id
stringlengths 30
36
| source
stringclasses 1
value | format
stringclasses 1
value | text
stringlengths 5
878k
|
|---|---|---|---|
no-problem/9812/astro-ph9812118.html
|
ar5iv
|
text
|
# Two–component galaxy models: phase–space constraints
## 1 Introduction
In the study of stellar dynamical models the fact that the Jeans equations have a physically acceptable solution is not a sufficient criterion for the validity of the model: the essential requirement to be met is the positivity of the DF of each distinct component. A model satisfying this minimal requirement is called a consistent model. In order to recover the DF of spherical models with anisotropy, the OM technique has been developed (Osipkov 1979; Merritt 1985), and numerically applied (see, e.g., Ciotti & Pellegrini 1992, CP92; Carollo, de Zeeuw, & van der Marel 1995; Ciotti & Lanzoni 1997, CL97). In the OM framework, a simple approach in order to check the consistency of spherically symmetric, multi–component models (avoiding the recovering of the DF itself), is described in CP92. It is now accepted that a fraction of the mass in galaxies is made of a dark component, whose density distribution – albeit not well constrained by observations – differs from that of the visible one (see, e.g., Bertin et al.1994; Carollo et al.1995; Buote & Canizares 1997; Gerhard et al.1998). Moreover, there is an increasing evidence of the presence of massive BHs at the center of most (if not all) elliptical galaxies (see, e.g., Harms et al.1994; van der Marel et al.1997; Richstone 1998). Unfortunately, only few examples of two–component systems in which both the spatial density and the DF are analytically known are at our disposition, namely the Binney–Evans model (Binney 1991; Evans 1993), and the two–component Hernquist model (HH model, Ciotti 1996, C96). It is therefore of interest the result proved in C98 that also the DF of H0 models with OM anisotropy is completely expressible in analytical way. This family of models is made by the superposition of a density distribution following the Hernquist profile (Hernquist 1990), and another density distribution following the $`\gamma =0`$ profile \[see eq. (3.5)\], with different total masses and core radii. OM orbital anisotropy is allowed in both components. Strictly related to the last point above, is the trend shown by the numerical investigations of CP92, i.e., the difficulty of consistently superimposing a centrally peaked distribution to a centrally flat one. More specifically, CP92 showed numerically that King (King 1972) or quasi–isothermal density profiles can not be coupled to a de Vaucouleurs (de Vaucouleurs 1948) model, because their DFs run into negative values near the model center. On the contrary, the DF of the de Vaucouleurs component is qualitatively unaffected by the presence of centrally flat halos. From this point of view, the C96 work on HH models is complementary to the investigation of CP92: in the HH models the two density components are both centrally peaked, and their DF is positive for all the possible choices of halo and galaxy masses and concentrations (in the isotropic case). The implications of these findings have been not sufficiently explored. For example, one could speculate that in presence of a centrally peaked dark matter halo, King–like elliptical galaxies should be relatively rare, or, viceversa, that a galaxy with a central power–law density profile cannot have a dark halo too flat in the center. In fact observational results on the central surface brightness profiles of elliptical galaxies (see, e.g., Jaffe et al.1994; Møller, Stiavelli, & Zeilinger 1995; Lauer et al.1995), and bulges of spirals (Carollo & Stiavelli 1998), as well as high–resolution numerical simulations of dark matter halos formation (Dubinsky & Carlberg 1991; Navarro, Frenk, & White 1997) seem to point in this direction. In C98, I explore further the trend emerged in CP92 and in C96, considering the analytical DFs of the H0 models and determining the structural and dynamical limitations imposed to them by dynamical consistency.
## 2 The consistency of multi–component systems
For a multi–component spherical system, where the orbital anisotropy of each component is modeled according to the OM parameterization, the DF of the density component $`\rho _\mathrm{k}`$ is given by:
$$f_\mathrm{k}(Q_\mathrm{k})=\frac{1}{\sqrt{8}\pi ^2}\frac{d}{dQ_\mathrm{k}}_0^{Q_\mathrm{k}}\frac{d\varrho _\mathrm{k}}{d\mathrm{\Psi }_\mathrm{T}}\frac{d\mathrm{\Psi }_\mathrm{T}}{\sqrt{Q_\mathrm{k}\mathrm{\Psi }_\mathrm{T}}},\varrho _\mathrm{k}(r)=\left(1+\frac{r^2}{r_{\mathrm{ak}}^2}\right)\rho _\mathrm{k}(r),$$
(2.1)
where $`\mathrm{\Psi }_\mathrm{T}(r)=_\mathrm{k}\mathrm{\Psi }_\mathrm{k}(r)`$ is the total relative potential, $`Q_\mathrm{k}=L^2/2r_{\mathrm{ak}}^2`$, and $`0Q_\mathrm{k}\mathrm{\Psi }_\mathrm{T}(0)`$. $``$ and $`L`$ are respectively the relative energy and the angular momentum modulus per unit mass, $`r_\mathrm{a}`$ is the anisotropy radius, and $`f_\mathrm{k}(Q_\mathrm{k})=0`$ for $`Q_\mathrm{k}0`$. If each $`f_\mathrm{k}`$ is non negative over all the accessible phase–space, the system is consistent. In C92 it was proved that
Theorem: A necessary condition (NC) for the non negativity of $`f_\mathrm{k}`$ given in eq. (2.1) is:
$$\frac{d\varrho _\mathrm{k}(r)}{dr}0,0r\mathrm{}.$$
(2.2)
If the NC is satisfied, a strong (SSC) and a weak sufficient condition (WSC) for the non negativity of $`f_\mathrm{k}`$ are respectively:
$$\frac{d}{dr}\left[\frac{d\varrho _\mathrm{k}(r)}{dr}\frac{r^2\sqrt{\mathrm{\Psi }_\mathrm{T}(r)}}{M_\mathrm{T}(r)}\right]0,\frac{d}{dr}\left[\frac{d\varrho _\mathrm{k}(r)}{dr}\frac{r^2}{M_\mathrm{T}(r)}\right]0,0r\mathrm{}.$$
(2.3)
Some considerations follow looking at the previous conditions. The first is that the violation of the NC is connected only with the radial behavior of $`\rho _\mathrm{k}`$ and the value of $`r_{\mathrm{ak}}`$, and so this condition applies independently of any other interacting component added to the model. Even when the NC is satisfied, $`f_\mathrm{k}`$ can be negative, due to the radial behavior of the integrand in eq. (2.1), which depends on the total potential, on the particular $`\rho _\mathrm{k}`$, and on $`r_{\mathrm{ak}}`$; so, a range of permitted values of $`r_{\mathrm{ak}}`$ satisfying the NC must be discarded. Naturally, the true critical anisotropy radius is always larger than or equal to that given by the NC, and smaller than or equal to that given by the SSC (WSC). To summarize: a model failing the NC is certainly inconsistent, and a model satisfying the SSC (WSC) is certainly consistent; the consistency of a model satisfying the NC and failing the SSC (WCS) can be proved only by direct inspection of the DF.
## 3 Results and conclusions
Both density distributions defining the H0 models belong to the family of the $`\gamma `$ models (Dehnen 1993):
$$\rho (r)=\frac{3\gamma }{4\pi }\frac{Mr_\mathrm{c}}{r^\gamma (r_\mathrm{c}+r)^{4\gamma }},0\gamma <3,$$
(3.4)
where $`M`$ is the total mass and $`r_\mathrm{c}`$ a characteristic scale–length. The main results obtained in C98 can be summarized as follows:
(1) The NC, WSC, and SSC that the model parameters must satisfy, in order to correspond to an H0 system for which the two physically distinct components have a positive DF, are analytically derived using the method introduced in CP92. Some conditions are obtained for the wider class of two–component $`\gamma _1+\gamma _2`$ models (of which the H0 models are a special case). In particular, it is shown that the DF of the $`\gamma _1`$ component in isotropic $`\gamma _1+\gamma _2`$ models is nowhere negative, independently of the mass and concentration of the $`\gamma _2`$ component, whenever $`1\gamma _1<3`$ and $`0\gamma _2\gamma _1`$. As an interesting application of this result, it follows that a BH of any mass can be consistently added at the center of any isotropic member of the $`\gamma `$ models family, when $`1\gamma <3`$. Two important consequences follow. The first is that the consistency of isotropic HH (or H+BH) models proved in C96 using an “ad hoc” technique is not exceptional, but a common property of a large class of two–component $`\gamma `$ models: for example, also isotropic two–component Jaffe \[Jaffe 1983, $`\gamma =2`$ in eq. (3.4)\] or Jaffe+BH models can be safely assembled. The second is that in two–component isotropic models, the component with the steeper central density distribution is usually the most robust against inconsistency.
(2) It is shown that an analytical estimate of a minimum value of $`r_\mathrm{a}/r_\mathrm{c}`$ for one–component $`\gamma `$ models with a massive (dominant) BH at their center can be explicitly found. As expected, this minimum value decreases for increasing $`\gamma `$.
(3) It is shown that the analytical expression for the DF of H0 models with general OM anisotropy can be found in terms of elliptic functions; the special cases in which each one of the two density components are embedded in a dominant halo are also discussed.
(4) The region of the parameter space in which H0 models are consistent is explored using the derived DFs: it is shown that, at variance with the H component, the $`\gamma =0`$ component becomes inconsistent when the halo is sufficiently concentrated, even in the isotropic case. This is an explicit example of the negative result found by CP92 described in the Introduction.
(5) The combined effect of halo concentration and orbital anisotropy is finally investigated. The trend of the minimum value for the anisotropy radius as a function of the halo concentration is qualitatively similar in both the components, and to that found for HH models in C96: a more diffuse halo allows a larger amount of anisotropy. A qualitatively new behavior is found and explained investigating the DF of the $`\gamma =0`$ component in the halo–dominated case for high halo concentrations. It is analytically shown the existence of a small region in the parameter space where a sufficient amount of anisotropy can compensate the inconsistency produced by the halo concentration on the structurally analogous – but isotropic – case.
(6) As a final remark, it can be useful to point out some general trends that emerge when comparing different one and two–component models with OM anisotropy, as those investigated numerically in CP92 and CL97, and analytically in C96 and C98. The first common trend is that OM anisotropy produces a negative DF outside the galaxy center, while the halo concentration affects mainly the DF at high (relative) energies. The second is that the possibility to sustain a strong degree of anisotropy is weakened by the presence of a very concentrated halo. The third is that in two–component models, in case of very different density profiles in the central regions, the component with the flatter density is the most “delicate” and can easily be inconsistent: particular attention should be paid in constructing such models.
|
no-problem/9812/astro-ph9812115.html
|
ar5iv
|
text
|
# The Gas Dynamics of Shell Galaxies
## 1 Introduction
It is widely accepted that shell galaxies may form as a result of a merger between an elliptical and a small disk galaxy. Simulations of the stellar component have shown that the shells or ripples are created either by “phase-wrapping” of debris on nearly radial orbits (Quinn 1984, ApJ, 279, 596), or by “spatial-wrapping” of matter in thin disks (Dupraz & Combes 1987, A&A, 185, L1; Hernquist & Quinn 1989, ApJ, 342, 1). Stars, often modeled as collisionless particles, which were originally bound to the merging companions are liberated and oscillate with different periods around the new common central potential. Before they relax and due to their different periods of oscillation, they accumulate near the apocenters of their orbits, to form shell-like features.
The fact that this scenario is indeed occuring has been questioned by the recent observations of neutral hydrogen in systems containing shells. Schiminovich et al (1994, ApJ, 432, L101; 1995, ApJ, 444, L77) have detected diffuse HI gas shells displaced just to the outside of the stellar ripples of Centaurus A and NGC 2865. These observational results were a priori surprising, since we believe that shells are phase-wrapped stellar structures and that the diffuse gaseous and stellar components do not have the same behavior when approaching the center of the potential well, in quasi-radial orbits (Weil & Hernquist 1993, ApJ, 405, 142). However, as it was pointed out (Kojima & Noguchi 1997, ApJ, 481, 132) the details of the gas dynamics depend strongly on the specific model adopted for the interstellar medium, and modeling of the gas using smooth particle hydrodynamics may not be very appropriate.
## 2 Numerical Simulations and Discussion
Our numerical simulations (Combes & Charmandaris, 1998, A&A, in preparation) indicate that these new HI observations can actually be accommodated within the standard picture for the formation of shell galaxies (see Fig. 1). We model the sinking disk galaxy using a realistic distribution of its stellar and gaseous component, taking into account dynamical friction and a proper treatment of the dissipation of the gas (using a cloud-cloud collision code). During the process of merging the tidal forces affect the gaseous component first since it is less bound than the stars, which in turn are liberated later. Due to the effect of dynamical friction the stars loose energy as the galaxy spirals inwards in the potential of the elliptical galaxy. By modifying the dissipation of the gas we can easily reproduce the observed spatial distribution of gaseous and stellar shells.
|
no-problem/9812/astro-ph9812426.html
|
ar5iv
|
text
|
# Temporal evolution of the pulse width in GRBs
## 1 Introduction
The cosmological origin of GRBs, established as a result of optical follow-up observations of fading X-ray counterparts to GRBs, requires an extraordinarily large amount of energy to flood the entire universe with gamma rays. The lack of apparent photon-photon attenuation of high energy photons implies substantial bulk relativistic motion. The bulk Lorentz factor, $`\mathrm{\Gamma }=(1\beta ^2)^{1/2}`$, must be on the order of $`10^2`$ to $`10^3`$. Two major scenarios involving relativistic shells have been developed. In the external shock models, a relativistic shell, that expands outward for a long period of time, is generated by a single release of energy during the merger. The shell coasts in a gamma-ray quiet phase for a certain period. Eventually, the shell becomes gamma-ray active due to the interactions with the external medium. If the shell has a velocity, $`v=\beta c`$, then the photons emitted $`on`$ $`axis`$ over a period $`t^{}`$ (“proper time” in the comoving frame of the shell) arrive at a detector over a much shorter period, $`T=\frac{t^{}}{2\mathrm{\Gamma }}`$. The duration of the event is set by the expansion of the shell and the complex temporal structure is due to inhomogeneities in the shell and/or the ambient material. The alternative theory is that a central site releases energy in the form of a wind or multiple shells over a period of time commensurate with the observed duration of GRB. Each subpeak in the GRB is the result of a separate explosive event in the central site. If the emission sites do indeed lie on a relativistically expanding shell, the pulse width in the time GRBs histories scales as $`\mathrm{\Delta }T=\mathrm{\Lambda }\mathrm{\Delta }t^{}`$, where $`\mathrm{\Lambda }`$ is the Doppler factor, $`\mathrm{\Lambda }=\mathrm{\Gamma }(1\beta \mathrm{cos}\theta )`$. Here $`\theta `$ is the angle of the motion of the emitting region with respect to the direction of the emission. In this paper, we proposed to determine the angular spread of the emitting region and the amount of deceleration from the time histories of many GRBs.
## 2 Pulse width evolution obtained from time histories
A visual inspection of the BATSE catalog of multiple-peaked time histories reveals that peaks usually have about the same duration at the beginning of the burst as near the end of the burst. Our aim is to characterize and measure the pulse shape as a function of arrival time. The aligned peak method measures the average pulse temporal structure, each burst contributes to the average by aligning the largest peak (Mitrofanov (1997)). We used all 53 bursts from the BATSE 4B Catalog that were longer than 20s and brighter than 5 photons s<sup>-1</sup> cm<sup>-2</sup>. Each burst must have at least one peak, as determined by a peak-finding algorithm (similar to Li & Fenimore (1996)), in each third of its duration. The largest peak in each third was normalized to unity and shifted in time, bringing the largest peaks of all bursts into common alignment. This method was applied in each third of the duration of the bursts. Thus, we obtained one curve of the averaged pulse shape for each different section of the bursts (as shown in Figure 1). The average profile is notably identical in each 1/3 of $`T_{90}`$ (we estimate the differential spread, $`S`$, to be $``$ 1%).
We have shown the lack of temporal evolution of the peak width in the context of an average of many bursts. Now, we expand our analysis to individual bursts. An excellent analysis has been provided by Norris et al. (1995), where they examined the temporal profiles of bright GRBs by fitting those profiles with pulses. From the set of bursts that they analysed, we used the 28 bursts with the characteristic of having seven or more fitted pulses within their duration. To obtain the temporal dependency of the pulse width, we selected the five largest peaks in each burst and fit their FWHM to a function of the arrival time, $`\mathrm{\Delta }T=k\left(\frac{TT_c}{T_{90}}\right)^\alpha `$. We chose this model because the expected dependency from the “external” shock model scales as $`\frac{T/T_{90}}{\mathrm{\Gamma }(1+\beta )}`$ (if a relativistic shell is responsible for the shape in the time histories, $`T_0`$ should be proportional to $`T_{90}`$; see Fenimore et al. (1996)). The purpose of the $`T_c`$ parameter is to correct the BATSE time to the time since the beginning of the explosion. Figure 2 shows the distribution of the power law indexes ($`\alpha `$) for the set of bursts analysed. The distribution shows that in individual bursts the pulse width does not increase throughout their duration as predicted by the external shock model, for which $`\alpha `$ is expected to be 1 (or larger if deceleration is ocurring). In fact, most bursts show a diminution in pulse width.
## 3 Discussion
We have uncovered that the width of the peaks remains remarkably constant throughout a burst. The width should scales as $`\mathrm{\Gamma }(1\beta \mathrm{cos}\theta )`$. In the external shock model, such lack of temporal evolution implies that $`\mathrm{\Gamma }`$ must be nearly constant. Peaks occur late in the burst because they are from regions off axis where the delay is caused by the curvature of the shell. The later peaks would be wider and delayed because off axis regions have larger $`\theta `$’s. A constant peak width indicates that all emitting entities must have similar $`\theta `$. Since the maximum angular size of the shell allowed by a differential spread of $`S`$ is $`S\mathrm{\Gamma }^1`$, the entire size of the shell must be a few percent of $`\mathrm{\Gamma }^1`$. Thus, the only external shock model that is consistent with the observations is one where the overall size of the shell is much smaller than $`\mathrm{\Gamma }^1`$ and there is no deceleration during the classic GRB phase. This adds to our previous arguments (Fenimore et al. (1996), Fenimore et al. (1998)) that a central engine (internal shocks) is the more likely explanation for the observed choatic time history. In the context of the internal shock model, we cannot place a limit on angular extent but the lack of evolution of the peak width indicates that $`\mathrm{\Gamma }`$ must be remarkably constant throught the internal shock phase.
|
no-problem/9812/nucl-th9812071.html
|
ar5iv
|
text
|
# References
LANL Report LA-UR-98-6000 (1998)
Talk given at the Fourth Workshop on Simulating Accelerator Radiation Environments (SARE4),
Knoxville, Tennessee, September 14-16, 1998
Production and Validation of Isotope Production Cross Section Libraries for Neutrons and Protons to 1.7 GeV
S. G. Mashnik, A. J. Sierk, K. A. Van Riper, and W. B. Wilson
T-2, Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545
White Rock Science, PO Box 4729, Los Alamos, NM 87545
## Abstract
For validation and development of codes and for modeling isotope production in high power accelerators and APT Materials studies, we have produced experimental, calculated, and evaluated activation libraries for interaction of nucleons with nuclides covering about a third of all natural elements. For targets considered here, our compilation of experimental data is the most complete we are aware of, since it contains all data available on the Web, in journal papers, laboratory reports, theses, and books, as well as all data included in the large compilation by Sobolevsky with co-authors (NUCLEX) published recently by Springer-Verlag in 4 volumes. Our evaluated library was produced using all available experimental cross sections together with calculations by the CEM95, LAHET, and HMS-ALICE codes and with the European Activation File EAF-97 and LANL Update II of the ECNAF Neutron Activation Cross-Section Library.
1. Introduction
Data on isotope production yields from various reactions are necessary to validate and develop models of nuclear reactions and are a decisive input for many applications, e.g., for accelerator transmutation of waste (ATW), accelerator-based conversion (ABC), accelerator-driven energy production (ADEP), accelerator production of tritium (APT), for the optimization of commercial production of radioisotopes used in medicine, mining, and industry, for solving problems of radiation protection of cosmonauts, aviators, workers at nuclear facilities, and for modeling radiation damage to computer chips, etc. (see details and references in ). Also, residual product nuclide yields in thin targets irradiated by medium- and high-energy projectiles are extensively used in cosmochemistry and cosmophysics to interpret the production of cosmogenic nuclides in meteorites by primary galactic particles.
Ideally, it would be desirable to have a universal library that includes data for all nuclides, projectiles, and incident energies. At present, neither the measurements nor the available codes permit one to create a reliable library covering all data needed at intermediate energies. First, experiments are costly and it is impossible to measure all data, in principle. Second, predictions by the best of available models, codes, and phenomenological systematics may differ for yields of certain isotopes at energies above 100 MeV by a factor of 100 or more, so that current models have to be further developed before they become reliable predictive tools (see, e.g., ). Construction of a universal comprehensive library would be very time consuming and costly, and is beyond the scope of the present work. Instead, we have collected experimental data, performed calculations with the most reliable codes, and evaluated cross sections for nucleon-nucleus interactions at intermediate energies for a number of targets of interest. Our progress in this activity is briefly described in this paper.
2. Experimental Data Library
We compile nucleon-induced isotope production experimental cross sections for two reasons. First, for validation and further development of the Cascade-Exciton Model (CEM) code CEM95 and to investigate the applicability of the recent ALICE code with the new Hybrid Monte Carlo Simulation model (HMS-ALICE) to produce activation libraries up to 150 MeV . Second, we need these data to evaluate reliable libraries for our medical isotope production and material studies at APT. For these tasks, we need data only up to several GeV. But realizing that such a compilation will be useful in the future for other problems, we do not limit ourselves to these energies, but compile all available data at any energy above several MeV. Since there are very few data on neutron-induced isotope production cross sections at energies above 100 MeV, we focus mainly on compiling proton-induced reaction data.
Many efforts have been previously made to compile experimental production yields from proton-induced reactions at intermediate energies. To the best of our knowledge, the most complete compilation was performed by Sobolevsky and co-authors and was published recently by Springer-Verlag in eight separate subvolumes . Sobolevsky and co-authors have performed a major work and compiled all data available to them for target elements from Helium to transuranics for the entire energy range from thresholds up to the highest energy measured. For proton-induced reactions, this compilation contains about 37,000 data points published in the first four Subvolumes I/13a-d (the following Subvolumes I/13e-h concern pion, deutron, triton, <sup>3</sup>He, and alpha induced reactions). This valuable compilation is also currently available in an electronic version as an IBM PC code named NUCLEX . Unfortunately, since this compilation is very expensive (Springer Verlag sells a single new subvolume for $1647.00 or $2020.00; interested buyers may find information at: http://www.springer-ny.com/catalog/np/nov96np/DATA/3-540-61045-6.html) it is not easily available to individual researchers or to small libraries. Of more immediate concern is the fact that NUCLEX does not contain a large volume of data obtained during recent years, especially for proton-induced reactions.
Due to the increasing interest in intermediate-energy data for ATW, ABC, ADEP, APT, astrophysics, and other applications, precise and voluminous measurements of proton-induced spallation cross sections have been performed recently, and are presently in progress, by R. Michel et al. from Hannover University , Yu. E. Titarenko et al. at ITEP, Moscow , Yu. V. Aleksandrov et al. at JINR, Dubna , B. N. Belyaev et al. at B. P. Konstantinov St. Petersburg Institute of Nuclear Physics , N. I. Venikov et al. at Kurchatov Institute, Moscow A. S. Danagulyan et al. at JINR, Dubna , H. Vonach et al. at LANL, Los Alamos , S. Sudar and S. M. Qaim at KFA, Jülich , D. W. Bardayan et al. at LBNL, Berkeley , J. M. Sisterson et al. at TRIUMF and other accelerators , etc. Finally, we note another, “new” type of nuclear reaction intensively studied in recent years, which provides irreplacable data for our needs. These are from reactions using reverse kinematics, when relativistic ions interact with hydrogen targets and they often provide the only way to obtain reliable data for interaction of intermediate energy protons with separate isotopes of an element with a complex natural isotopic composition. Much data from this type of reaction have been recently obtained, e.g., by W. R. Webber et al. at the LBL Bevalac and L. Tassan-Got et al. at GSI, Darmstadt . These new data, as well as a number of other new and old measurements have not been covered by NUCLEX. Therefore, we do not confine ourselves solely to NUCLEX as a source of experimental cross sections; instead, we compile all available data for the targets in which we are interested, searching first the World Wide Web, then any other sources available to us, including the compilation from NUCLEX.
Table 1 shows the whole list of isotope production cross sections we have so far included in our experimental library file. Actually, our experimental library consists at present of 32 files, a separate file for each element, stored in a simple and easy to read format, and a README file with the description of
Table 1
Experimental Data Library for Proton-Induced Isotope Production
| Target Z | Target # | Target | No. of | The same, | No. of data | The same, |
| --- | --- | --- | --- | --- | --- | --- |
| | | | reactions | in NUCLEX | points | in NUCLEX |
| 6 | | | | | 652 | 487 |
| | 1 | <sup>13</sup>C | 1 | 1 | | |
| | 2 | <sup>12</sup>C | 7 | 0 | | |
| | 3 | <sup>nat</sup>C | 13 | 8 | | |
| 7 | | | | | 493 | 419 |
| | 4 | <sup>14</sup>N | 11 | 0 | | |
| | 5 | <sup>13</sup>N | 5 | 5 | | |
| | 6 | <sup>nat</sup>N | 9 | 8 | | |
| 8 | | | | | 400 | 347 |
| | 7 | <sup>18</sup>O | 5 | 5 | | |
| | 8 | <sup>16</sup>O | 14 | 0 | | |
| | 9 | <sup>nat</sup>O | 16 | 11 | | |
| 9 | | | | | 153 | 136 |
| | 10 | <sup>19</sup>F | 9 | 9 | | |
| 10 | | | | | 28 | 7 |
| | 11 | <sup>22</sup>Ne | 1 | 1 | | |
| | 12 | <sup>20</sup>Ne | 21 | 0 | | |
| | 13 | <sup>nat</sup>Ne | 4 | 4 | | |
| 11 | | | | | 96 | 91 |
| | 14 | <sup>23</sup>Na | 8 | 8 | | |
| 12 | | | | | 596 | 497 |
| | 15 | <sup>26</sup>Mg | 3 | 3 | | |
| | 16 | <sup>25</sup>Mg | 3 | 3 | | |
| | 17 | <sup>24</sup>Mg | 26 | 3 | | |
| | 18 | <sup>nat</sup>Mg | 17 | 14 | | |
| 13 | | | | | 1170 | 816 |
| | 19 | <sup>27</sup>Al | 45 | 23 | | |
| 15 | | | | | 42 | 37 |
| | 20 | <sup>31</sup>P | 5 | 4 | | |
| 16 | | | | | 93 | 49 |
| | 21 | <sup>34</sup>S | 1 | 1 | | |
| | 22 | <sup>32</sup>S | 33 | 0 | | |
| | 23 | <sup>nat</sup>S | 10 | 9 | | |
| 17 | | | | | 25 | 25 |
| | 24 | <sup>nat</sup>Cl | 4 | 4 | | |
| 18 | | | | | 102 | 66 |
| | 25 | <sup>40</sup>Ar | 36 | 0 | | |
| | 26 | <sup>nat</sup>Ar | 17 | 17 | | |
| 19 | | | | | 20 | 20 |
| | 27 | <sup>nat</sup>K | 3 | 3 | | |
Table 1 (continued)
| Target Z | Target # | Target | No. of | The same, | No. of data | The same, |
| --- | --- | --- | --- | --- | --- | --- |
| | | | reactions | in NUCLEX | points | in NUCLEX |
| 20 | | | | | 660 | 258 |
| | 28 | <sup>48</sup>Ca | 1 | 1 | | |
| | 29 | <sup>44</sup>Ca | 4 | 4 | | |
| | 30 | <sup>43</sup>Ca | 2 | 2 | | |
| | 31 | <sup>42</sup>Ca | 1 | 1 | | |
| | 32 | <sup>40</sup>Ca | 57 | 0 | | |
| | 33 | <sup>nat</sup>Ca | 26 | 12 | | |
| 26 | | | | | 2638 | 1189 |
| | 34 | <sup>58</sup>Fe | 4 | 4 | | |
| | 35 | <sup>57</sup>Fe | 5 | 5 | | |
| | 36 | <sup>56</sup>Fe | 99 | 7 | | |
| | 37 | <sup>44</sup>Fe | 2 | 2 | | |
| | 38 | <sup>nat</sup>Fe | 92 | 58 | | |
| 27 | | | | | 2213 | 871 |
| | 39 | <sup>59</sup>Co | 54 | 48 | | |
| 30 | | | | | 997 | 996 |
| | 40 | <sup>70</sup>Zn | 2 | 2 | | |
| | 41 | <sup>68</sup>Zn | 6 | 6 | | |
| | 42 | <sup>67</sup>Zn | 3 | 3 | | |
| | 43 | <sup>66</sup>Zn | 3 | 3 | | |
| | 44 | <sup>64</sup>Zn | 6 | 6 | | |
| | 45 | <sup>nat</sup>Zn | 43 | 42 | | |
| 31 | | | | | 356 | 356 |
| | 46 | <sup>71</sup>Ga | 12 | 12 | | |
| | 47 | <sup>69</sup>Ga | 13 | 13 | | |
| | 48 | <sup>nat</sup>Ga | 6 | 6 | | |
| 32 | | | | | 651 | 651 |
| | 49 | <sup>76</sup>Ge | 31 | 31 | | |
| | 50 | <sup>74</sup>Ge | 3 | 3 | | |
| | 51 | <sup>73</sup>Ge | 2 | 2 | | |
| | 52 | <sup>72</sup>Ge | 6 | 6 | | |
| | 53 | <sup>70</sup>Ge | 29 | 29 | | |
| | 54 | <sup>nat</sup>Ge | 15 | 15 | | |
| 33 | | | | | 198 | 180 |
| | 55 | <sup>75</sup>As | 62 | 59 | | |
| 39 | | | | | 1509 | 614 |
| | 56 | <sup>89</sup>Y | 65 | 50 | | |
| 40 | | | | | 2351 | 812 |
| | 57 | <sup>96</sup>Zr | 14 | 14 | | |
| | 58 | <sup>94</sup>Zr | 30 | 30 | | |
| | 59 | <sup>92</sup>Zr | 4 | 4 | | |
Table 1 (continued)
| Target Z | Target # | Target | No. of | The same, | No. of data | The same, |
| --- | --- | --- | --- | --- | --- | --- |
| | | | reactions | in NUCLEX | points | in NUCLEX |
| 40 | | | | | 2351 | 812 |
| | 60 | <sup>91</sup>Zr | 44 | 42 | | |
| | 61 | <sup>90</sup>Zr | 40 | 40 | | |
| | 62 | <sup>nat</sup>Zr | 73 | 34 | | |
| 41 | | | | | 898 | 205 |
| | 63 | <sup>93</sup>Nb | 70 | 61 | | |
| 42 | | | | | 1432 | 1339 |
| | 64 | <sup>100</sup>Mo | 8 | 8 | | |
| | 65 | <sup>98</sup>Mo | 7 | 7 | | |
| | 66 | <sup>97</sup>Mo | 5 | 5 | | |
| | 67 | <sup>96</sup>Mo | 25 | 25 | | |
| | 68 | <sup>95</sup>Mo | 8 | 8 | | |
| | 69 | <sup>94</sup>Mo | 8 | 8 | | |
| | 70 | <sup>92</sup>Mo | 8 | 8 | | |
| | 71 | <sup>nat</sup>Mo | 113 | 38 | | |
| 54 | | | | | 145 | 94 |
| | 72 | <sup>126</sup>Xe | 2 | 0 | | |
| | 73 | <sup>124</sup>Xe | 6 | 2 | | |
| | 74 | <sup>nat</sup>Xe | 1 | 1 | | |
| 55 | | | | | 370 | 370 |
| | 75 | <sup>133</sup>Cs | 83 | 83 | | |
| 56 | | | | | 834 | 238 |
| | 76 | <sup>nat</sup>Ba | 79 | 30 | | |
| 57 | | | | | 122 | 122 |
| | 77 | <sup>nat</sup>La | 66 | 66 | | |
| 77 | | | | | 114 | 94 |
| | 78 | <sup>93</sup>Ir | 1 | 1 | | |
| | 79 | <sup>nat</sup>Ir | 46 | 28 | | |
| 79 | | | | | 2104 | 935 |
| | 80 | <sup>197</sup>Au | 278 | 271 | | |
| 80 | | | | | 73 | 73 |
| | 81 | <sup>202</sup>Hg | 10 | 10 | | |
| 83 | | | | | 1142 | 995 |
| | 82 | <sup>209</sup>Bi | 262 | 174 | | |
| Total: | 82 | | 2272 | 1574 | 22679 | 13389 |
the format and of references. Our library is still in progress and we hope to extend it, depending on our needs, and to make it available for users through the Web.
Presently, our experimental library contains 22,679 data points for 82 targets of 32 elements covering 2,272 proton-induced reactions. We also have begun to store in our library data for intermediate energy neutron-induced reactions, but so far we have only 95 data points for Bi and C targets covering 14 reactions induced by neutrons.
For comparison, we also show in Table 1 the statistics of available data for the same targets in NUCLEX . One can see, that for several targets like Cl, K, Ga, Ge, Cs, La, and Hg, no new measurements were performed in recent years, and we have not found more data than is in NUCLEX. On the other hand, for such targets like Ca, Fe, Co, Y, Zr, Ba, and Au, we have available 2-3 times more data points than in NUCLEX. Also, there are a number of targets like <sup>12</sup>C, <sup>14</sup>N, <sup>16</sup>O, <sup>20</sup>Ne, <sup>32</sup>S, <sup>40</sup>Ar, <sup>40</sup>Ca, and <sup>126</sup>Xe, for which there are no data at all in NUCLEX, while we presently have data for 181 reactions on these targets.
3. Calculated Cross Section Library
We needed to calculate a library of isotope production by nucleons at energies above 100 MeV to simulate the production of medical radioisotopes in a high-energy neutron and proton environment, e.g., at an APT Facility , to benchmark the CEM95 code , and to see how a similar library created by M. B. Chadwick using the HMS-ALICE code for energies below 150 MeV agrees with calculations at higher energies .
We use the CEM95 code to perform most of the calculations for our activation library above 100 MeV. A number of reactions are calculated also with the LAHET code system (version 2.83) , and just a few reactions, with the recently improved version of the CEM code . For energies below 100 MeV, we use the LA150 activation library by M. B. Chadwick .
We store calculated cross sections for the production of all possible isotopes from 78 targets (see the list in Table 2), and then use the results we need. Most of the reactions are calculated for energies from 100 MeV to 1.7 GeV, according to the needs of our medical isotope production study , while several reactions are calculated only up to 1.0 GeV for , while some others used extensively in our benchmark , were calculated from 10 MeV to 5 GeV. Having already a tested method for production of activation cross sections at these energies, our library can be extended easily for other targets, when necessary.
We have benchmarked our calculations against all available experimental data. Figures with comparison of more than 1000 excitation functions calculated with CEM95, LAHET, and HMS-ALICE codes with the available experimental data and predictions of other well-known models can be found in . From this comparison we see that, for the majority of nuclides the results agree quite well with each other and with the data. We can therefore conclude that the calculated activation cross sections are reasonably reliable and can be used together with available experimental data to produce an evaluated activation library, which we describe in the next section.
4. Evaluated Library
As we mentioned previously, neither available experimental data nor any of the current models or phenomenological systematics can be used alone to produce a reliable evaluated activation library covering a large area of target nuclides and incident energies. Therefore, we choose to create our evaluated library by approximating by hand smoothly, wherever possible, excitation functions using all available experimental data along with calculations using some more reliable codes, employing each of them in the corresponding regions of targets and incident energies where they work better. When we have reliable experimental data, they are taken as the highest priority for our approximation as compared to model results.
Table 2
List of targets covered by the CEM95 Activation Library: p(n) + A $``$ any isotope
( $`10/100`$ MeV $`T_01.7/5.0`$ GeV)
| <sup>12</sup>C | <sup>58</sup>Ni, <sup>60</sup>Ni, <sup>61</sup>Ni, <sup>62</sup>Ni, <sup>64</sup>Ni, <sup>nat</sup>Ni |
| --- | --- |
| <sup>14</sup>N | <sup>63</sup>Cu, <sup>65</sup>Cu, <sup>nat</sup>Cu |
| <sup>16</sup>O, <sup>18</sup>O | <sup>67</sup>Zn, <sup>68</sup>Zn, <sup>70</sup>Zn |
| <sup>19</sup>F | <sup>69</sup>Ga, <sup>71</sup>Ga |
| <sup>21</sup>Ne, <sup>22</sup>Ne | <sup>70</sup>Ge, <sup>73</sup>Ge, <sup>74</sup>Ge, <sup>76</sup>Ge |
| <sup>23</sup>Na | <sup>75</sup>As |
| <sup>24</sup>Mg, <sup>25</sup>Mg, <sup>26</sup>Mg , <sup>nat</sup>Mg | <sup>89</sup>Y |
| <sup>27</sup>Al | <sup>90</sup>Zr, <sup>91</sup>Zr, <sup>92</sup>Zr, <sup>94</sup>Zr, <sup>96</sup>Zr, <sup>nat</sup>Zr |
| <sup>32</sup>S, <sup>33</sup>S, <sup>36</sup>S , <sup>nat</sup>S | <sup>93</sup>Nb |
| <sup>35</sup>Cl, <sup>37</sup>Cl | <sup>92</sup>Mo, <sup>94</sup>Mo, <sup>95</sup>Mo, <sup>96</sup>Mo, <sup>97</sup>Mo, <sup>98</sup>Mo, <sup>100</sup>Mo |
| <sup>36</sup>Ar, <sup>38</sup>Ar, <sup>40</sup>Ar | <sup>132</sup>Xe, <sup>134</sup>Xe |
| <sup>39</sup>K, <sup>40</sup>K, <sup>41</sup>K | <sup>133</sup>Cs |
| <sup>40</sup>Ca | <sup>134</sup>Ba, <sup>135</sup>Ba, <sup>136</sup>Ba, <sup>137</sup>Ba, <sup>138</sup>Ba , <sup>nat</sup>Ba |
| <sup>54</sup>Fe, <sup>56</sup>Fe, <sup>57</sup>Fe, <sup>58</sup>Fe, <sup>nat</sup>Fe | <sup>138</sup>La, <sup>139</sup>La |
| <sup>59</sup>Co | <sup>197</sup>Au |
The recent International Code Comparisons for Intermediate Energy Nuclear Data organized by NEA/OECD at Paris , our own comprehensive benchmarks , and several studies by Titarenko et al. have shown that CEM95 and LAHET generally have the best predictive powers for spallation reactions at energies above 100 MeV as compared to other available models. Therefore, we choose them above 100 MeV to create our evaluated library. Actually, we employ the calculated library described in Sec. 3. The same benchmarks have shown that at lower energies, the HMS-ALICE code does one of the best jobs in comparison with other models. So, we use the activation library calculated by M. Chadwick with the HMS-ALICE code for protons below 100 MeV and neutrons between 20 and 100 MeV. In the overlapping region, between 100 and 150 MeV, we use both HMS-ALICE and CEM95 and/or LAHET results. For neutrons below 20 MeV, data of the European Activation File EAF-97, Rev. 1 with some recent improvements by M. Herman seem to be the most reliable, therefore we use them as the first priority in our evaluation.
Measured cross-section data from the compilation described in Sec. 2, when available, are included together with theoretical results and are used to evaluate cross sections for our medical isotope production study for 70 nuclides of 25 elements. We note that when we put together all these different theoretical results and experimental data, rarely do they agree perfectly with each other, providing a smooth continuity of evaluated excitation functions. Often, the resulting compilations show significant disagreement at energies where the available data progresses from one source to another. These sets are thinned to eliminate discrepant data, providing data sets of more-or-less reasonable continuity defining our evaluated cross sections used in calculations .
Examples with typical results of evaluated activation cross sections for several proton and neutron reactions are shown in Figs. 1 and 2 by broad gray lines. 51 similar color figures for proton-induced reactions and 56 figures for neutrons, can be found on the Web, in our detailed report .
Figure 1. Example of several evaluated proton-induced activation cross sections. Evaluated cross sections are shown by broad gray lines, other notation are given in the plots and described in the text and in .
Figure 2. The same as in Fig. 1, but for neutron-induced reactions.
5. Summary
We have produced experimental, calculated, and evaluated activation libraries for the interaction of nucleons with nuclides covering about a third of all natural elements. Our compilation of experimental proton-induced cross sections contains 22,679 data points for 82 targets of 32 elements covering 2,272 proton-induced reactions and is the most complete we are aware of for these targets.
The methods developed are applicable to an extended set of reactions. We plan to extend our libraries, depending on our needs, and to make them available for users through the Web.
Acknowledgements
We express our gratitude to R. E. MacFarlane and L. S. Waters for interest in and support of the present work. This study was supported by the U. S. Department of Energy.
|
no-problem/9812/cond-mat9812093.html
|
ar5iv
|
text
|
# Phonon Mediated Transresistivity in a Double Layer Composite Fermion System
## Abstract
We consider the fractional drag in a double layer system of two-dimensional electrons in the half-filled lowest Landau level. At sufficiently large inter-layer separations the drag is dominated by exchange of acoustic phonons and exhibits novel temperature and inter-layer distance dependences. At low temperatures the phonon mediated drag is strongly enhanced with respect to the case of zero magnetic field.
In the recent years, the problem of two-dimensional electrons in a double quantum well has attracted a lot of, both, experimental and theoretical attention. Apart from other interesting properties, this system offers a new opportunity for studying electron correlations by virtue of the phenomenon of the frictional drag. The latter manifests itself in the form of a finite transresistivity defined as a voltage induced in an open circuit (passive) layer while a current is flowing through the other (active) one (see and references therein).
In the absence of electron tunneling, the transresistivity which takes its origin in momentum transfer from the active to the passive layer can only result from inter-layer Coulomb coupling. Thus, being of the interaction origin, the transresistivity provides a sensitive probe for such important characteristics of the two-dimensional electron gas as the electron polarization operator and the dynamically screened inter-layer coupling.
Detailed experimental studies of the frictional drag in semiconductor heterostructures revealed features associated with inter-layer plasmons . However, it was predicted that, despite its dominance at small inter-layer distances $`d`$, the drag caused by the direct Coulomb coupling decays rapidly with $`d`$ . Since the experimentally measured drag showed a weaker distance dependence at large $`d`$, it was suggested that some other mechanism of momentum transfer ought to be at work.
As such, exchange of acoustic phonons was proposed as the most likely alternative. Improving on the earlier results obtained in , the authors of Ref. considered the effects of, both, deformation potential (DP) and piesoelectric (PE) electron-phonon couplings. This work was primarily concerned with the behavior at temperatures comparable to or higher than the Bloch-Gruneisen temperature $`T_{BG}=2uk_F`$ defined in terms of the Fermi momentum $`k_F`$ and the speed of sound $`u`$, where the DP coupling is known to be far more important than the PE one. At temperatures smaller than $`T_{BG}`$ their relative strength reverts though.
According to the general argument , the low-temperature transresistivity is controlled by the phase-space volume available for thermally excited bosonic modes of an appropriate kind. Thus, measuring an exponent $`n`$ in the low-temperature dependence $`\rho _{12}T^n`$ allows one to identify a mechanism responsible for the drag. Namely, the Coulomb drag gives $`n=2`$ while the phonon mediated drag tends to yield substantially higher $`n`$ values .
Since at high enough temperatures the frictional drag of any origin is expected to undergo a crossover to a universal linear $`T`$ dependence , an observation of a maximum of $`\rho _{12}(T)/T^2`$ at $`TT_{BG}`$ provides a strong evidence in favor of the phonon mechanism.
Recent drag measurements performed in the Quantum Hall regime revealed new features which are believed to be associated with a formation of compressible metal-like electron states at half-filling as well as other even-denominator filling fractions $`\nu 1/2p`$ of the lowest Landau level. In the framework of the theory by Halperin, Lee, and Read these states are described in terms of spinless fermionic quasiparticles dubbed composite fermions (CFs). In the mean field picture of the best studied $`\nu =1/2`$ Quantum Hall state, the CFs experience zero effective field and occupy all states inside some ostensible Fermi surface of the size $`k_F=(4\pi n_e)^{1/2}`$ given in terms of the electron density $`n_e`$.
Among other differences from ordinary electrons in zero field, the compressible CF states were predicted to feature a peculiar Coulomb drag behavior $`\rho _{12}T^{4/3}`$ characteristic of an anomalously slow relaxation of overdamped density fluctuations .
Further work in this direction focused on the experimentally observed tendency of $`\rho _{12}(T)`$ to saturate at low $`T`$ or even increase upon decreasing the driving current , the features that were attributed to a possible CF inter-layer pairing which might result in a formation of an incompressible (intrinsically double-layer) gapful electron state .
However, on the general grounds, one expects that neither the direct Coulomb coupling nor the proposed pairing mechanism remain relevant at large enough inter-layer separations. Instead, the drag measurements performed in this regime may reveal important details of coupling between the CFs and acoustic phonons.
Motivated by this expectation, in the present paper we consider the phonon mediated drag in a widely-separated double layer CF system.
Earlier insight into the problem of the CF-phonon coupling has been gained from the analysis of the available experimental data on phonon-limited mobility , phonon drag contribution to thermopower , and energy loss rate by hot electrons due to phonon emission at $`\nu =1/2`$.
The bare coupling between CFs and PE phonons is described by the standard electron-phonon vertex
$$M_\lambda (𝐐)=eh_{14}(A_\lambda /2\rho u_\lambda Q)^{1/2},A_l=\frac{9q_z^2q^4}{2Q^6},A_{tr}=\frac{8q_z^4q^2+q^6}{4Q^6}$$
$`(1)`$
where $`𝐐=(𝐪,q_z)`$ is the 3D phonon momentum, $`\rho `$ is the bulk density of $`GaAs`$, $`u_\lambda `$ is the speed of sound with polarization $`\lambda =l,tr`$, and $`h_{14}`$ is the non-zero component of the piesoelectric tensor which relates the local electrostatic potential to the lattice displacement.
As demonstrated in Ref., the bare PE vertex (1) undergoes full dynamical screening governed by the intra-layer $`V_{11}(q)=2\pi e^2/ϵ_0q`$ and inter-layer $`V_{12}(q)=V_{11}(q)e^{qd}`$ Coulomb potentials, $`ϵ_012.9`$ being the bulk dielectric constant of $`GaAs`$.
In the CF theory, the effects of screening are described in terms of the intra-layer CF density $`\mathrm{\Pi }_{00}(\omega ,𝐪)`$ and current $`\mathrm{\Pi }_{}(\omega ,𝐪)`$ polarization functions. However, when computing physical quantities, these polarizations always appear in a particular combination, which one can readily recognize as the irreducible electron density response function
$$\chi (\omega ,𝐪)=\frac{\mathrm{\Pi }_{00}}{1(4\pi )^2\mathrm{\Pi }_{00}\mathrm{\Pi }_{}}=\frac{q^2}{q^2(dn_e/d\mu )^1(4\pi )^2i\omega \sigma (q)}$$
$`(2)`$
This peculiar form of the density response which is characteristic of the compressible CF states is given in terms of the CF compressibility $`dn_e/d\mu `$ and the momentum-dependent conductivity $`\sigma (q)=\frac{1}{2}k_Fmin(l,2/q)`$, where $`l`$ stands for the CF mean free path . Throughout this paper, we refer to the momenta $`q`$ larger (smaller) than $`l^1`$ as ballistic (diffusive) regime, respectively.
Moreover, as a straightforward analysis shows, the possibility to express such quantities of physical interest as the transresistivity solely in terms of $`\chi (\omega ,𝐪)`$ also holds for the dynamically screened CF-phonon coupling included alongside with the direct inter-layer Coulomb interaction.
With both present, the two interactions constitute the effective inter-layer coupling
$$W_{12}=\frac{V_{12}+D_{12}}{(1+\chi (V_{11}+D_{11}))^2\chi ^2(V_{12}+D_{12})^2}$$
$`(3)`$
where the two-dimensional electron-phonon interaction function $`D_{ij}(\omega ,𝐪)`$ is computed with the use of the bulk phonon propagator ($`\omega _+=\omega +i(u/2l_{ph})sgn\omega `$)
$$D_\lambda (\omega ,𝐐)=\frac{2u_\lambda Q}{\omega _+^2u_\lambda ^2Q^2},$$
$`(4)`$
which depends on the three-dimensional momentum $`𝐐=(𝐪,q_z)`$ and accounts for a finite phonon mean free path $`l_{ph}`$ due to boundary scattering, and the wave function of the lowest occupied transverse electron subband $`\psi (z)ze^{z/w}`$ that determines the formfactor $`F_i(q_z)=𝑑z|\psi _i(z)|^2e^{iq_zz}`$.
For $`\omega `$ close to $`u_\lambda q`$ the relevant values of $`q_z`$ turn out to be small. Therefore one can only keep the contribution of transverse phonons $`(A_lA_{tr}1/4)`$ with velocity $`u3\times 10^5cm/s`$ which for $`qw1`$ results in the formula :
$$D_{ij}(\omega ,𝐪)=\underset{\lambda }{}\frac{dq_z}{2\pi }F_i(q_z)F_j(q_z)|M_\lambda (𝐐)|^2𝒟_{ij}(\omega ,𝐐)$$
$$\frac{(eh_{14})^2}{8u^2\rho \sqrt{q^2(\omega _+/u)^2}}(\delta _{ij}+(1\delta _{ij})e^{d\sqrt{q^2(\omega _+/u)^2}})$$
$`(5)`$
Since perpendicular magnetic field does not affect the $`z`$-axis confinement, Eq.5 is applicable at arbitrary magnetic fields.
To the (lowest nontrivial) second order in the effective inter-layer coupling (3) the transresistivity is given by the standard expression :
$$\rho _{12}=\frac{1}{8\pi ^2}\frac{h}{e^2}\frac{1}{Tn_e^2}\frac{d𝐪}{(2\pi )^2}_0^{\mathrm{}}𝑑\omega (\frac{qIm\chi (\omega ,𝐪)}{\mathrm{sinh}\frac{\omega }{2T}})^2|W_{12}(\omega ,𝐪)|^2$$
$`(6)`$
which can also incorporate, under some modifications, the effects of a magnetic field and/or disorder.
In order to facilitate a direct comparison with the calculation of $`\rho _{12}(T)`$ in zero field, we will refer to the following expression
$$ϵ(\omega ,𝐪)=\frac{dn_e}{d\mu }(\chi ^1+V_{11})=1+\frac{dn_e}{d\mu }(V_{11}(q)+\frac{i\omega \sigma (q)}{q^2})$$
$`(7)`$
as the effective dielectric function of a single layer CF system.
By making use of Eq.7 and putting $`D_{12}D_{11}=D`$ we arrive at the expression
$$\rho _{12}=\frac{1}{16\pi ^3}\frac{h}{e^2}\frac{1}{Tn_e^2}\frac{\sigma ^2(q)dq}{q}\frac{\omega ^2d\omega }{\mathrm{sinh}^2\frac{\omega }{2T}}\frac{|V_{12}(q)+D(\omega ,𝐪)|^2}{|(ϵ(\omega ,𝐪)V_{12}(q))(ϵ(\omega ,𝐪)+V_{12}(q)+2D(\omega ,𝐪))|^2}$$
$`(8)`$
where the upper limit in the integral over momentum is set by either the maximum span of the CF Fermi surface $`2k_F`$ or by the inverse width of the quantum well $`w^1`$, which, for the sake of simplicity, we choose to be equally restrictive ($`2k_Fw1`$).
The relative strength of the PE phonon coupling is controlled by a small parameter $`\eta =h_{14}^2ϵ_0/16\pi u^2\rho 10^3`$. Therefore it proves to be convenient to divide the $`q`$-integral onto the momenta smaller and greater than $`d^1`$. In the former, the small parameter $`\eta `$ allows one to neglect $`D`$ in Eq.8 altogether and, in this way, to recover the pure case of the Coulomb drag. In this range of momenta, the integral over frequencies receives its main contribution from $`\omega q^2V_{11}(q)/\sigma (q)`$ corresponding to the overdamped density mode whose (purely imaginary) dispersion is derived from the equation $`ϵ(\omega ,𝐪)=0`$.
Depending on the values of the two dimensionless parameters $`\xi =k_Fd`$ and $`\sigma =k_Fl`$, the contribution of the momenta $`qd^1`$ undergoes a number of crossovers. Provided that both $`\sigma `$ and $`\xi `$ are sufficiently large, Eq.8 yields a variety of regimes characterized by different dependences on the dimensionless temperature $`\tau =T/\mathrm{\Delta }`$ measured in units of the average Coulomb energy $`\mathrm{\Delta }=k_Fe^2/ϵ_0100K`$:
$$\rho _{12}^C(T)\left(\begin{array}{c}(\tau \sigma /\xi )^2\mathrm{ln}(\tau \sigma ^3/\xi ),\\ \tau <\xi /\sigma ^3\\ (\tau /\xi )^{4/3},\\ \xi /\sigma ^3<\tau <1/\xi ^2\\ \tau /\xi ^2,\\ 1/\xi ^2<\tau \end{array}\right)or\left(\begin{array}{c}(\tau \sigma /\xi )^2\mathrm{ln}(\tau /\sigma \xi ),\\ \tau <1/\sigma \xi \\ \tau \sigma /\xi ^3,\\ 1/\sigma \xi <\tau \end{array}\right)$$
$`(9)`$
in the situations $`l>d`$ and $`l<d`$, respectively.
In the case of a long CF mean free path, the asymptotical low-$`T`$ regimes $`\rho _{12}^CT^{4/3}`$ and $`T^2\mathrm{ln}T`$ were previously obtained under the assumptions of a ballistic and diffusive CF dynamics, respectively . According to Eq.9, upon increasing the inter-layer separation beyond the CF mean free path the former regime ceases to be accessible.
Since neither Eq.9 nor its zero field counterpart $`\rho _{12}^{C,0}\tau ^2\xi ^4`$ depend on the mass of the charge carriers, the only reason behind the observed three orders of magnitude enhancement of the $`\nu =1/2`$ drag at $`d300A`$ is because is was measured at low enough $`T`$.
As follows from Eq.9, when treated in the second order perturbation theory, the pure Coulomb drag at $`\nu =1/2`$ can only be responsible for a temperature dependence $`\rho _{12}(T)T^n`$ with $`n<2`$, including the regime of the logarithmic enhancement.
Since the inter-layer Coulomb correlations decay rapidly with $`d`$, and the perturbative result (9) becomes more and more accurate, one might conclude that the Coulomb drag alone could not explain a maximum in $`\rho _{11}(T)/T^2`$, should one be observed.
Nonetheless, the recent experiment on the double layer $`\nu =1/2`$ system with $`d5000A`$ did reveal such a maximum occurring at roughly the same $`T2K`$ as in the zero field case . By analogy with the situation in zero field, this observation may call for an alternate mechanism of inter-layer momentum transfer.
In our unified approach that led to Eq.8, an additional contribution due to exchange of PE phonons (both real and virtual ones corresponding to the contributions proportional to $`ImD`$ and $`ReD`$, respectively ) comes from the momenta $`qd^1`$, at which one can instead neglect $`V_{12}`$ in the integrand.
First, we consider the case of a short phonon mean free path due to boundary scattering (the exact conditions under which this approximation holds are to be specified later).
The phonon mediated contribution to the transresistivity can be cast in the form
$$\rho _{12}^{ph}\frac{1}{16\pi ^3}\frac{h}{e^2}\frac{\eta ^2}{Tn_e^2}_{1/d}^{2k_F}\frac{\sigma ^2(q)dq}{q}_0^{\mathrm{}}\frac{\omega ^2d\omega }{\mathrm{sinh}^2\frac{\omega }{2T}}\frac{e^{d\sqrt{(q^2(\omega _+/u)^2)}}}{|ϵ(\omega ,𝐪)|^4\sqrt{(q^2(\omega /u)^2)^2+(q/l_{ph})^2}}$$
$`(10)`$
By introducing the parameter $`\delta =uϵ_0/e^210^2`$ we present the results of the integrations in the form
$$\rho _{12}^{ph}(T)/\eta ^2\mathrm{ln}(l_{ph}/d)\left(\begin{array}{c}\tau ^4\sigma ^2\delta ^2(1+\sigma ^2\delta ^2)^2,\\ \delta \xi ^1<\tau <\delta \sigma ^1\\ \tau ^6\delta ^8,\\ max(\delta \xi ^1,\delta \sigma ^1)<\tau <\delta ^2\\ \tau ^2,\\ max(\delta \xi ^1,\delta \sigma ^1,\delta ^2)<\tau <\delta \\ \tau \delta ,\\ \delta <\tau \end{array}\right)$$
$`(11)`$
All of the above regimes can only be accessible if there exist sizable intervals of temperature within the bounds imposed by the conditions $`\xi ^1\sigma ^1\delta 1`$. With all the regimes present, the exponent $`n`$ in the approximate power-law temperature dependence of $`\rho _{12}^{ph}(T)`$ does not remain constant and changes non-monotonously as $`T`$ gets smaller.
The $`\mathrm{ln}(l_{ph}/d)`$ dependence of the transresistivity on the inter-layer separation holds as long as $`dl_{ph}`$, while at higher $`T`$ Eq.11 yields $`\rho _{12}^{ph}\mathrm{exp}(d/l_{ph})`$.
A different variety of regimes occurs in the case of a long phonon mean free path where, in order to arrive at a finite result, one has to keep $`D`$ in the denominator of Eq.8 that now reads as
$$\rho _{12}^{ph}\frac{1}{16\pi ^3}\frac{h}{e^2}\frac{\eta ^2}{Tn_e^2}_{1/d}^{2k_F}\frac{\sigma ^2(q)dq}{q}_0^{\mathrm{}}\frac{\omega ^2d\omega }{\mathrm{sinh}^2\frac{\omega }{2T}}\frac{e^{d\sqrt{q^2(\omega /u)^2}}}{|ϵ(\omega ,𝐪)(ϵ(\omega ,𝐪)\sqrt{q^2(\omega /u)^2}2\eta )|^2}$$
$`(12)`$
With $`D`$ included, the integrand in (12) develops a new weakly damped pole at $`\omega uq\sqrt{1\eta ^2}`$, which corresponds to a collective electron-phonon mode that forms in a double layer system (cf.). When integrating over frequencies, we restrict ourselves to the vicinity of this pole.
The coupled electron-phonon mode can only occur provided that neither the real nor the imaginary part of the denominator in (12) is affected by the intrinsic phonon lifetime at all $`qd^1`$. The corresponding criteria can be expressed as $`l_{ph}\eta ^2max(d,(k_F\delta )^1,d\sigma ^1\delta ^1)`$.
Expanding the denominator near the pole and carrying out the integrations we arrive at the following regimes:
$$\rho _{12}^{ph}(T)/\eta ^2\left(\begin{array}{c}\tau ^4\sigma \delta ^3(1+\sigma ^2\delta ^2)^1,\\ \delta \xi ^1<\tau <\delta \sigma ^1\\ \tau ^5\delta ^6,\\ max(\delta \xi ^1,\delta \sigma ^1)<\tau <\delta ^2\\ \tau ^3\delta ^2,\\ max(\delta \xi ^1,\delta \sigma ^1,\delta ^2)<\tau <\delta \\ \tau ,\\ \delta <\tau \end{array}\right)$$
$`(13)`$
Unlike the Coulomb drag (9) that originates from small momenta $`qd^1`$, the integral (12) receives only a negligible correction from the region where the dielectric function $`ϵ(\omega ,𝐪)`$ approaches zero.
According to Eq.13, the contribution of the electron-phonon mode remains nearly independent of $`d`$ as long as the condition $`d\sqrt{q^2(\omega /u)^2}qd\eta 1`$ is satisfied for all momenta less than $`2k_F`$, which implies $`\xi \eta <1`$.
Similar to the case of ordinary electrons in zero field Eq.12 includes, apart from the Lorenzian-like contribution associated with the collective mode that Eq.13 does account for, a substantial non-Lorenzian tail due to processes involving virtual phonons, which gives rise to a non-trivial overall dependence on $`d`$ (cf. Ref.). For this reason one should not expect the estimates (11) and (13) obtained under the assumptions of, respectively, short and long $`l_{ph}`$ to match smoothly at $`l_{ph}d\eta ^2`$.
In order to simplify a comparison between our results and the zero field analysis we make the following choice of the parameters: $`\xi \sigma \delta ^1`$, which greatly reduces the number of the relevant regimes. Incidentally, the above choice is not that far off the experimental situation of Ref.: $`d5000A`$ and $`l1\mu m`$ which corresponds to a single layer resistivity $`\rho _{11}2000\mathrm{\Omega }`$.
At these parameter values, in the range of temperatures $`0.01T/T_{BG}1`$ one can only expect to observe the asymptotical behavior $`\rho _{12}^{ph}(T)\eta ^2(T/T_{BG})^2\delta ^2\mathrm{ln}(l_{ph}/d)`$ or $`\eta ^2(T/T_{BG})^3\delta `$ for the phonon mean free path being shorter or longer than its crossover value estimated as $`d\eta ^2`$, which is well above a typical one $`l_{ph}0.1mm`$ .
In either case, the exponent is substantially reduced as compared to the case of ordinary electrons in zero field: $`\rho _{12}^{ph,0}\eta ^2(T/T_{BG}^0)^6\delta ^2\mathrm{ln}(l_{ph}/d)`$ and $`\eta ^2(T/T_{BG}^0)^5\delta `$, respectively.
Thus, compared to its zero field counterpart, the phonon mediated drag in the half-filled Landau level is strongly enhanced at all temperatures smaller than $`T_{BG}`$ (in Ref. the measured ratio of the transresistivities was $`\rho _{12}^0/\rho _{12}10^3`$ at $`T1K`$), while they become comparable at $`TT_{BG}`$.
For lower values of the exponent $`n`$ in the power-law temperature dependence of the transresistivity the maximum of the ratio $`\rho _{12}(T)/T^2`$ shifts towards lower $`T`$. When contrasted against its position in zero field, this downward shift could, at least partly, compensate for another, upward, shift due to the extra factor $`T_{BG}/T_{BG}^0=\sqrt{2}`$ resulting from the ratio between the CF and the ordinary electron’s Fermi momenta at a given density . Interestingly enough, in Ref. the maximum was found to occur in both cases at roughly the same position, which was, however, lower than the estimated value of $`T_{BG}^05K`$.
Compared to the direct Coulomb drag in the CF system (9), the relative strength of the phonon mediated drag is weaker than in the case of ordinary electrons in zero field. Indeed, on the basis of Eq.13 one would conclude that even at $`TT_{BG}`$ the two competing contributions to the transresistivity can only become comparable at $`\xi \eta ^1`$ which corresponds to inter-layer separations in excess of $`10\mu m`$, while the use of Eq.11 would result in even higher values of $`d`$.
Conversely, at zero field the phonon-mediated drag outpowers the direct Coulomb one already at $`\xi \delta ^{1/4}\eta ^{1/2}`$ which only requires separations $`d10^3A`$.
Thus, it does not seem to be obvious that under the experimental conditions of Ref. the phonon mediated drag indeed dominates over the direct Coulomb one in the CF regime, although it indeed appears to be the main contribution at zero field and $`TT_{BG}^0`$.
It is worthwhile mentioning that, by itself, an observation of a peak in $`\rho _{12}(T)/T^2`$ in a magnetic field may not yet provide an unambiguous evidence in favor of the phonon mechanism. In Ref. such a peak was predicted to occur in strong enough magnetic fields with no reference to phonons or the peculiar CF behavior in the lowest Landau level. Seemingly corroborating, is the observation of the peak’s evolution with magnetic field at fixed $`k_F`$ in a whole range of magnetic fields (see also ).
Finally, a comparison between the data from Refs. and shows that upon increasing $`d`$ by a factor of $`20`$ both $`\rho _{12}^0(1K)`$ and $`\rho _{12}(1K)`$ drop by the factors of $`50`$ and $`250`$, respectively.
In the case of $`\rho _{12}^0`$, this change is markedly less pronounced than the $`\xi ^4`$ behavior of the direct Coulomb drag , which could be naturally explained if a crossover to the weakly $`d`$-dependent phonon mediated regime occurred already at $`d10^3A`$.
On the contrary, the change in $`\rho _{12}`$ is somewhat more consistent with the roughly $`\xi ^2`$ behavior of the CF Coulomb drag than with the nearly $`d`$-independent phonon contribution that has taken over at some intermediate $`d`$. Thus, despite some qualitative agreement with the experimental findings of Ref., our results call for further experimental studies of the frictional drag in the CF regime, which have to be performed at even larger inter-layer separations $`d`$ in order to unequivocally point at the phonon-related origin of the phenomenon.
To summarize, we carried out a comprehensive analysis of the frictional drag in the double layer compressible CF state of electrons in the lowest Landau level. We considered the effect of the inter-layer electron-phonon coupling and contrasted it against the drag caused by the direct Coulomb interaction. We found that in the CF system the low temperature phonon mediated drag is strongly enhanced compared to the case of ordinary electrons in zero field. However, in order to single out the phonon drag in the CF system, the transresistivity measurements have to be done at larger inter-layer distances than those which are generally sufficient in the zero field case.
|
no-problem/9812/cond-mat9812331.html
|
ar5iv
|
text
|
# Comment on “Charged impurity scattering limited low temperature resistivity of low density silicon inversion layers” (Das Sarma and Hwang, cond-mat/9812216)
In a recent preprint, Das Sarma and Hwang propose an explanation for the sharp decrease in the $`B=0`$ resistivity at low temperatures which has been attributed to a transition to an unexpected conducting phase in dilute high-mobility two-dimensional systems (see Refs.\[1-4\] in ). The anomalous transport observed in these experiments is ascribed in Ref. to temperature-dependent screening and energy averaging of the scattering time. The model yields curves that are qualitatively similar to those observed experimentally: the resistivity has a maximum at a temperature $`E_F/k_B`$ and decreases at lower temperatures by a factor of 3 to 10. The anomalous response to a magnetic field (e.g., the increase in low-temperature resistivity by orders of magnitude ), is not considered in Ref. .
Two main assumptions are made in the proposed model: (1) the transport behavior is dominated by charged impurity scattering centers with a density $`N_i`$, and (2) the metal-insulator transition, which occurs when the electron density ($`n_s`$) equals a critical density ($`n_c`$), is due to the freeze-out of $`n_c`$ carriers so that the net free carrier density is given by $`nn_sn_c`$ at $`T=0`$. The authors do not specify a mechanism for this carrier freeze-out and simply accept it as an experimental fact. Although not included in their calculation, Das Sarma and Hwang also note that their model can be extended to include a thermally activated contribution to the density of “free” electrons.
In this Comment, we examine whether the available experimental data support the model of Das Sarma and Hwang.
(i) Comparison with the experimental data (see Fig. 1 of Ref. ) is made for an assumed density of charged impurities of $`3.5\times 10^9`$ cm<sup>-2</sup>, a value that is too small. In an earlier publication, Klapwijk and Das Sarma explicitly stated that the number of ionized impurities is “$`3\times 10^{10}`$ cm<sup>-2</sup> for high-mobility MOSFET’s used for the 2D MIT experiments. There is very little room to vary this number by a factor of two”. Without reference to this earlier statement, the authors now use a value for $`N_i`$ that is one order of magnitude smaller .
(ii) According to the proposed model, the number of “free” carriers at zero temperature is zero at the “critical” carrier density ($`n_s=n_c`$) and it is very small ($`n=n_sn_c<<n_c`$) near the transition. In this range, the transport must be dominated by thermally activated carriers, which decrease exponentially in number as the temperature is reduced. It is known from experiment that at low temperatures the resistance is independent of temperature at $`n_c`$ (the separatrix between the two phases) and depends weakly on temperature for nearby electron densities. In order to give rise to a finite conductivity $`e^2/h`$ at the separatrix, an exponentially small number of carriers must have an exponentially large mobility, a circumstance that is rather improbable.
(iii) Recent measurements of the Hall coefficient and Shubnikov-de Haas oscillations yield electron densities that are independent of temperature and equal to $`n_s`$ rather than a density $`n=n_sn_c`$ of “free” electrons . This implies that all the electrons contribute to the Hall conductance, including those that are frozen-out or localized. Although this is known to occur in quantum systems such as Hall insulators, it is not clear why it can hold within the simple classical model proposed by Das Sarma and Hwang.
|
no-problem/9812/astro-ph9812342.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
In the past, the analysis of giant H ii regions relied on single star photoionization models. Recent advances in stellar evolution and population synthesis made it possible to build models of H ii regions that are ionized by star clusters (, , , , , , ). Such models are helpful in the quest of the best diagnostics of giant H ii regions and of the populations of hot stars embedded in them .
After a short introduction to photoionization modelling, we will mainly address two points. One concerns the influence of the nebular density structure. The other concerns the strong line method for deriving elemental abundances. This method has already been widely discussed in the literature, but deserves further attention because it is the only possible way to derive abundances in low surface brightness systems as well as in distant galaxies.
## 2 Photoionization modelling in brief
The primary parameters defining a model are (i) the oxygen abundance (oxygen usually provides most of the cooling and can be considered as the element defining the metallicity Z), (ii) the mean effective temperature $`<T_{eff}>`$ of the ionizing radiation field (to first order, it can be characterized by the ratio $`Q_{He^o}`$/ $`Q_{H^o}`$ of the total number of photons above 24.6 and 13.6 eV, and (iii) the ionization parameter $`U_S`$ whose definition for a spherical nebula of constant density is $`U_S=Q_{H^o}/(4\pi R_S^2nc)`$ where $`R_S`$ is the Strömgren radius, $`n`$ is the gas density and $`c`$ the speed of light.
The intensities of the emission lines are proportional to the ionic populations of the elements producing them, but are strongly modulated by the electron temperature. The electron temperature results from a balance between the energy gains (due to ionization of hydrogen and helium, and essentially fixed by the effective temperature) and the energy losses (mainly due to line emission from the ”metals”, i.e. the elements heavier than H and He, except for the lowest metallicities where hydrogen losses become dominant). The ionization parameter sets the distribution of the various ions within the nebula, once the radiation field is specified.
Among the secondary parameters are: (i) the relative abundances of the elements in the gas with respect to oxygen, (ii) the possible presence of dust and its composition, (iii) the details of the energy distribution of the stellar ionizing photons, and (iv) the gas density distribution and the location of the ionizing stars with respect to the gas.
The species whose abundance variations affect the thermal balance most are nitrogen, carbon, and the refractory elements. The N/O and C/O ratios are expected to vary from one object to another since nitrogen and carbon are not produced in the same nucleosynthetic sites as oxygen. If abundant, nitrogen and carbon contribute significantly to the cooling of the gas. Refractory elements like Si, Mg, Fe may be depleted in the gas phase if dust is present, and deprive it from a cooling source which becomes important at high metallicities .
The direct effects of dust (apart from reddening) are mainly to compete with gas for the absorption of photons but, as shown by , dust also participates in the heating or the cooling of the gas. These effects strongly depend on the physical conditions inside the nebula and are difficult to model in detail, but they are of secondary importance in the thermal balance.
The energy distribution of the ionizing radiation field is dictated by the population of ionizing stars and the properties of their atmospheres. Even for a given mean effective temperature (or a mean $`Q_{He^o}`$/ $`Q_{H^o}`$), the stellar energy distribution influences the emission line properties by affecting the ionization structure of the various elements . A review of the impact of stellar evolution and atmospheres can be found in .
The effect of the distribution of the stars with respect to the gas has been considered in a simplified way by a few authors (, ) who modelled giant H ii regions as aggregates of single star Strömgen spheres rather than entities ionized by a dense star cluster located in their center, as is done in most studies. There is an increasing number of objects (e.g. 30 Doradus , NGC 1569 , I Zw 18 ) where the position of individual stars is known, and different situations exist.
So far, the influence of the density distribution has not been much studied. This partly stems from the fact that most photoionization codes assume spherical symmetry . However, even in such a case, there is room for exploration, as seen below.
## 3 The effect of the density structure on emission line ratios
Here, we will briefly explore several simple geometries, using models obtained with the photoionization code PHOTO and a synthetic evolving starburst as described in Stasińska & Leitherer . A more extended grid of models than the one presented in that paper has now been constructed. The details of all the models, including the intensities of about a hundred of lines from the far IR to the UV and other helpful parameters are available by anonymous ftp from ftp.obspm.fr/pub/obs/grazyna/cd-crete.
### 3.1 Filled sphere versus hollow shell
Two geometries have been mostly used to model giant H ii regions. One is a full sphere of constant density (e.g. ) or uniformly filled with gas clumps of constant density (e.g. ). It has been shown on various occasions that any combination of $`Q_{H^o}`$, $`n`$, and the filling factor $`ϵ`$ keeping the ionization parameter $`U_S`$ constant results in the same nebular ionization structure. A few authors have used another geometry, namely a hollow shell (e.g. ). This choice is motivated by the ring structures often seen in giant H ii regions in nearby galaxies (, , ) and by the idea that the interaction of stellar winds and supernovae explosions from the central star cluster with the ambient interstellar gas creates large bubbles , .
In order to illustrate the role played by geometry, we compare in Fig. 1 several line intensity ratios as a function of $`Q_{He^o}`$/$`Q_{H^o}`$ for three sequences of evolving models of giant H ii regions identical in every respect except the geometry. Filled circles represent full sphere models. Empty circles represent hollow sphere models of same $`U_S`$ . Empty squares represent hollow sphere models of same volume averaged ionization parameter $`\overline{U}`$ as the full sphere models (as mentioned by , for a full sphere of constant density and given filling factor, one has $`\overline{U}=3U_S`$). It is clearly seen that the emission line spectra are different in all three models. At a given $`Q_{He^o}`$/$`Q_{H^o}`$ , the \[Oiii\]/ \[Oii\] or \[Siii\]/ \[Sii\] ratios are not simply dependent of the ”ionization parameter” (whatever definition one takes), and in the example shown in Fig. 1, the dependence on the geometry is not the same for the two line ratios. As a consequence, the ”radiation softness parameter” (\[Oii\]/\[Oiii\]) / (\[Sii\]/\[Siii\]) used for ranking the mean effective temperature of H ii regions and believed to be relatively insensitive to ionization conditions , does depend on the geometry.
Unfortunately, there is no easy rule to predict the magnitude and even the direction of the changes when altering the geometry, even in such simple cases as considered here. Therefore, unless something is known about the geometry of the H ii regions one is studying, it is highly recommended to test the robustness of the inferences derived from photoionization mmodels by considering several exemplary geometries.
It is also important to realize that, in ab initio theoretical models of evolving starbursts, the prescription adopted for the geometry has an incidence on the evolution of the emission line intensity ratios.
García-Vargas et al. adopted a thin shell geometry defined by a distance to the central star cluster held constant during the whole evolution of the cluster. In such a circumstance, we have $`\overline{U}Q_{H^o}(t)`$. Stasińska & Leitherer adopted a full sphere geometry of uniform density, and in such a case it is easily shown that $`\overline{U}(Q_{H^o}(t))^{\frac{1}{3}}`$. That is, the ionization parameter decreases with $`Q_{H^o}`$(and therefore with time) at a much slower rate than in the models of García-Vargas et al. Clearly, any inference on starburst ages based on such models strongly depends on the adopted nebular properties. So, one cannot simply use the ionization parameter as a measure of the age of a given starburst, as suggested by García-Vargas et al. Actually, a starburst can possibly be unveiled using emission line intensities of the surrounding H ii region only if additional information is available.
Note that, if, adopting a thin shell model, the radius were taken to vary with time like in the theory of stellar wind bubbles (), one would have $`\overline{U}t^{\frac{6}{5}}Q_{H^o}(t)`$. On the other hand, adopting a full sphere geometry, if gas were supposed to expand at constant velocity, one would have $`\overline{U}t^1(Q_{H^o}(t))^{\frac{1}{3}}`$. Of course, probably none of these two descriptions is realistic, although it is likely that the the ionization parameter decreases with time not only on the account of a decrease in $`Q_{H^o}`$ but also because of ram pressure effects and gas dissipation.
### 3.2 Integrated spectra of core-halo H ii regions
There is ample documentation that many giant H ii regions exhibit a core-halo structure. There is also evidence, in galaxies, for the existence of a diffuse ionized medium (, , , ). This diffuse gas is probably ionized by photons leaking out from the core H ii regions, although other mechanisms may be at work in addition . In such a picture, the core is either density bounded, at least in some directions, or ionization bounded with a covering factor less than one. The diffuse medium has a much lower ionization parameter than the core, therefore it emits strongly in \[Sii\] and \[Oi\] and weakly in \[Oiii\]. When studying a nearby giant H ii region, the diffuse ionized medium will hardly contribute to the observed spectrum. On the other hand, when observing a galaxy at high redshift, the slit will encompass the bright core together with the diffuse halo, and the resulting spectrum will be very different from that of an H ii region in a nearby galaxy . This ”aperture effect” can be mistaken for a sign of ”activity” or shock excitation in galaxies at large redshifts. An example of such a situation is shown in Table 1 and in Fig. 2.
## 4 Abundances from the strong line method
While abundances in H ii regions are readily obtained from line intensity ratios using electron temperature based empirical methods (the problem of temperature fluctuations should be kept in mind however, see , , and references therein), these methods cannot be used for distant or low surface brightness galaxies. Indeed, they require a measurement of the faint \[Oiii\] $`\lambda `$4363 line. Even in close by H ii regions, they cannot be used if the metallicity is higher than about half solar, because the electron temperature is then too low for \[Oiii\] $`\lambda `$4363 to be emitted. An interesting alternative to overcome this problem has been proposed by Pagel et al. , using only the strongest lines. Initially, the method was intended to allow the determination of abundances in the metal rich giant H ii regions in the inner parts of spiral galaxies. This method has been further discussed in a number of papers (e.g. , ). Then, Skillman showed its potential use for low abundance systems. Mc Gaugh (, ) provided the first two-dimensional calibration taking explicitely into account the role of the ionization parameter.
The leading idea is that, with two line ratios (\[Oii\]/ H$`\beta `$ and \[Oiii\]/ H$`\beta `$ ) it is possible to estimate the three main defining parameters of an H ii region (O/H, $`<T_{eff}>`$ and $`\overline{U}`$) if there is a universal link between these parameters. And indeed, it seems that the initial mass function of the ionizing stars of giant H ii regions is universal (see Leitherer, this volume). This implies that, on average, one expects $`<T_{eff}>`$ and O/H to be related in such objects.
Mc Gaugh has built a grid of photoionization models of various metallicities and ionization parameters ionized by clusters of OB stars, and provided a convenient diagram to read out the values of $`\overline{U}`$ and O/H from \[Oiii\] / \[Oii\] versus (\[Oiii\]+ \[Oii\]) / H$`\beta `$ . One problem is that the relation between (\[Oiii\]+ \[Oii\]) / H$`\beta `$ and O/H is double valued: at low metallicities, the intensities of \[Oii\] and \[Oiii\] with respect to H$`\beta `$ increase with increasing O/H. But when the oxygen abundance gets high enough for this element to become the main coolant, the electron temperature drops and the cooling is gradually transferred from the optical forbidden lines to the infra red fine structure lines. Consequently, (\[Oiii\]+ \[Oii\]) / H$`\beta `$ becomes a decreasing function of O/H. If nothing is suspected a priori about the metallicity of an object, one can use \[Nii\] / \[Oii\] as a discriminant between the two regimes: this ratio is expected to be small ($`<`$ 0.1) at low metallicities .
Mc Gaugh analyzed various sources of errors in the method and estimated an intrinsic uncertainty of 0.05 dex in log (O/H) at the low abundance end, 0.1 dex at the high abundance end, and 0.2 dex in the turn over region. The expected overall uncertainties in the oxygen abundance are obtained by combining these formal error bars with those due to observational errors , giving an estimated accuracy of about 0.2 dex for reasonably high signal-to-noise data, and worse in the turn over region.
Note that, on the low abundance end, comparing the values of O/H derived from the strong line method and from the electron temperature based method provides a direct estimate of the error. Unfortunately, this cannot be done on the high abundance end. Even tailored photoionization models of metal rich H ii regions fitting all the important lines do not provide a reliable calibration: the observational constraints are generally far from sufficient to allow a precise determination of the oxygen abundance in metal rich objects where, as developed below, the emission line intensities are extremely sensitive to the physical conditions in the nebulae.
Mc Gaugh’s computations were done assuming a zero age starburst of given upper stellar mass limit, $`M_{up}`$. The evolution of a starburst was mimicked by considering different values of $`M_{up}`$. Mc Gaugh’s diagram is constructed for $`M_{up}=60M_{}`$, corresponding to an average age of $``$ 3 Myr for a burst having initially $`M_{up}=100M_{}`$.
Using evolving synthetic starburst models , it is possible to follow the distortion of Mc Gaugh’s diagram as the starburst ages. This is done in Figs. 3, 4 and 5, which show the diagram obtained by bicubic spline interpolation on the grid of photoionization models of at 1 Myr, 3 Myr and 5 Myr respectively. The continuous lines correspond to loci of same O/H, while the dotted lines correspond to loci of same total initial mass of the stars (and not same ionization parameter as in Mc Gaugh).
We see that, on the low abundance end, the main effect of ageing is a drop in \[Oiii\]/ \[Oii\], which is due to a decrease of $`\overline{U}`$ implied by the lowering of the number of ionizing photons. The isoabundance curves are not much displaced, especially at low excitation. Indeed, in this regime, cooling is mainly due to collisional excitation of H Ly $`\alpha `$ and is a steep function of the electron temperature. Therefore, the electron temperature is insensitive to changes in the heating rate, and so are the intensities of the \[Oiii\] $`\lambda `$5007 and \[Oii\] $`\lambda `$3727 . As a consequence, the oxygen abundance derived from such a diagram is not too dependent on age. A typical value of the corresponding uncertainty is $`\pm `$ 0.1 dex for an age of 3 $`\pm `$ 1 Myr.
On the high abundance end, the distortion of Mc Gaugh’s diagram is impressive. At high metallicities, cooling is provided by the infrared fine structure lines \[Oiii\] $`\lambda `$$`\lambda `$52, 88$`\mu `$, which are insensitive to the electron temperature. Therefore, the electron temperature is extremely affected by changes in the energy gains, and so is the intensity of \[Oiii\] $`\lambda `$5007 . To make things worse, at high metallicities, the mean effective temperature of the stellar radiation is subject to large variations during the evolution (see Table 3 of ). An important hardening occurs when stars reach the Wolf-Rayet phase, boosting the \[Oiii\] $`\lambda `$5007 line. As a result, oxygen abundances derived from the strong line method are quite uncertain at the high metallicity end. Typically, an H ii region having log (O/H)+12 = 9 and an age of 5 Myr would appear less metal rich by 0.3 dex if interpreted with a diagram corresponding to 3 Myr. Other sources of uncertainties contribute as well, owing to the fact that the electron temperature is so ill defined. For example, any density irregularity, any change in the geometry will greatly affect the emission in \[Oiii\] $`\lambda `$5007 . An example is shown in Fig. 6, which presents Mc Gaugh’s diagram at 3 Myr like Fig. 4, but for hollow spheres of same $`\overline{U}`$ as the full sphere models used in Fig. 4. Note that Figs. 3 – 6 were constructed by extrapolating the solar metallicity models towards higher abundances. We have also constructed models at twice solar metallicity (they are available on the ftp site mentioned above). Using these models produces very twisted isoabundance curves even around log (O/H)+12 = 9, because at metallicities twice solar, the strong cooling may completely quench the \[Oiii\] $`\lambda `$5007 line. The diagram is further complicated by the fact that, at high metallicities, recombination takes over collisional excitation in the production of the \[Oii\] line.
As regards the turnover region, warnings have already been expressed by several authors. In this regime, the main cooling agent is \[Oiii\] $`\lambda `$5007 , therefore the intensity of this line is independent of O/H, and is a function of the heating rate only. Nevertheless, one can say that an H ii regions showing (\[Oiii\]+ \[Oii\]) / H$`\beta `$ $`>`$ 0.7 must have a log (O/H)+12 within a few dex of 8.
In summary, the strong line method is rather robust at low abundances (log (O/H)+12 $`<`$ 7.7) but is rather uncertain at higher metallicities. The fact that in spiral galaxies, one does see a trend in (\[Oiii\]+ \[Oii\]) / H$`\beta `$ with galactocentric radius (e.g. ) is probably indeed related to an abundance effect. In this case, observing a large number of H ii regions at a given galactocentric radius will smooth out the effects due to non uniformity of the stellar populations. Still, the unknown geometry of the H ii regions affects any calibration of the method. Trustworthy abundance determinations in metal rich H ii regions have to await the observation of the infrared fine structure lines.
It has been suggested that the line ratio \[Aiii\] $`\lambda `$7136 / \[Siii\] $`\lambda `$9532 could be an indicator of metallicity in metal rich object, being an indicator of electron temperature. We do not share this optimistic view since, as commented above, the electron temperature is ill defined at high metallicities and strongly dependent on the properties of the ionizing clusters and on the density structure of the nebulae.
The determination of the N/O ratio has been specifically discussed by several authors , , with the help of single star photoionization models. Using the grid of models for evolving starburst presented in Sect. 3 as an interpolating device, one can obtain the N/O ratios from the observed \[Nii\]/ \[Oii\] with a typical uncertainty of $`\pm `$ 0.15 dex.
|
no-problem/9812/cond-mat9812146.html
|
ar5iv
|
text
|
# Monte Carlo Simulations of a Generalized n–spin facilitated kinetic Ising Model
## I introduction
There is a continuous effort in describing of supercooled liquids using different approaches . However the phenomenon is generally not complete understood. Supercooled fluids reveal normally a stretched exponential decay of typical (e.g. density–density) correlation functions and a non-Arrhenius behavior of the associated relaxation times. This slowing down in the dynamical behavior can be illustrated by a strongly curved trajectory in the Arrhenius plot (relaxation time $`\tau `$ versus the inverse temperature $`T^1`$), empirically described by the well known Williams–Landel–Ferry (WLF) relation . But in contrast to conventional phase transitions a long range order is not developed.
The characteristic slowing down of the dynamics is usually explained by an increasing cooperativity of local processes with decreasing temperature . This behavior is an universal phenomena of the glass transition.
Mode coupling theories (MCT) predict the existence of an ergodic behavior above a critical temperature $`T_c`$ and a nonergodic behavior below $`T_c`$. Note that $`T_c`$ is in the range between the melting temperature $`T_m`$ and the glass temperature $`T_g`$, i.e. $`T_m>T_c>T_g`$. At $`T_c`$ the system undergoes a sharp transition from an ergodic state to a state with partially frozen (density) fluctuations. The slow $`\alpha `$–process within the MCT is thought to correspond to the actual dynamic glass transition whereas the fast $`\beta `$–process is often identified with a cage rattling or the boson peak.
Actually, the nonergodic state obtained from the original MCT below $`T_c`$ are approximately stable only for a finite time interval. Strongly cooperative processes lead to a slow decay of apparently frozen structures. This slow decay shows the typical above mentioned properties corresponding to the dynamics of the main glass transition (WLF like behavior of the relaxation time, stretched exponential decay of the correlation function). This effect can be partially described in terms of an extended mode coupling theory introducing additional hopping processes.
There exist also various alternative descriptions which explain the cooperative motion of the particles inside a supercooled liquid below $`T_c`$. One of these possibilities is the spin facilitated Ising model , originally introduced by Fredrickson and Andersen. The basic idea of these models consists of a coarse graining of space and time scales and simultaneously a reduction of the degrees of freedom. In details that means:
1. Coarse graining of spatial scales: The supercooled liquid is divided into cells in such a way that each cell contains a sufficiently large number of particles which realize a representative number of molecular motions.
2. Reduction of the degrees of freedom: Each cell will be characterized by only one degree of freedom, i.e. the cell structure enables us to attach to each cell an observable $`s_j`$ (usually denoted as spin) which characterizes the actual dynamic state of particles inside the cell $`j`$. The usual realization is given by the local density $`\rho _j`$ (particles per cell) with $`s_j=1`$ if $`\rho _j>\overline{\rho }`$ and $`s_j=1`$ if $`\rho _j<\overline{\rho }`$ where $`\overline{\rho }`$ is the averaged density of the system. This mapping implies consequently different mobilities of the particles inside such a cell, i.e. $`s_j=1`$ corresponds to the immobile solid like state and $`s_j=1`$ to the mobile state of cell $`j`$. The set of all spin observables forms a configuration, the time expansion of the corresponding probability distribution obeys a master equation.
3. Coarse graining of the time scale: This step bases on the assumption that fast processes (e.g. the $`\beta `$–process) are well separated from the slow $`\alpha `$–process. Hence, the original Liouville equation of the supercooled liquid can be projected onto a simple master equation without any memory terms. Therefore, the spin facilitated kinetic Ising model is suitable for a description of a supercooled liquid well below $`T_c`$ within the MCT and for sufficiently large time scales.
To make the time evolution of the glass configurations more transparent we use the argumentation following the idea of Fredrickson and Andersen , i.e. we suppose that the basic dynamics is a simple process $`s_j=+1s_j=1`$ controlled by the thermodynamical Gibb’s measure and by self-induced topological restrictions. In particular, an elementary flip at a given cell is allowed only if the number of the nearest neighbored mobile cells ($`s_j=1`$) is equal or larger than a restriction number $`f`$ with $`0<f<z`$ ($`z`$: coordination number). So, elementary flip processes and geometrical restrictions lead to the cooperative rearrangement of the underlying system and therefore to a mesoscopic model describing a supercooled liquid below $`T_c`$. Such models are denoted as $`f`$–spin facilitated Ising model on a $`d`$–dimensional lattice, SFM$`[f,d]`$. The SFM$`[f,d]`$ can be classified as an Ising-like model the kinetics of which is confined by restrictions of the ordering of nearest neighbors to a given lattice cell. This self–adapting environments influence in particular the long time behavior of the spin-spin and therefore of the corresponding density-density correlation functions. These models were studied numerically (SFM$`[2,2]`$) and recently also analytically (SFM$`[f,d]`$).
For the present investigations, we generalize the usual SFM$`[f,d]`$ by introduction of additional short range interactions which favor (antiferromagnetic case) or prevent (ferromagnetic case) the formation of liquid–solid (mobile–immobile) interfaces. It is the aim to study a two–dimensional generalized SFM$`[2,2]`$ using Monte–Carlo simulations.
## II model
We consider a generalized spin facilitated Ising model with nearest-neighbor interactions in two dimensions. The Hamiltonian of the model is the same as that of the standard two-dimensional Ising model with an external field
$$H=h(J\underset{<ij>}{}s_is_j+\underset{i}{}s_i),s_i=\pm 1.$$
(1)
In our notation, the inverse temperature $`T`$ and the Boltzmann constant $`k`$ has been absorbed into the field $`h`$. Physically, the field corresponds to the difference of the energy per cell in the liquid and the solid state. In our later discussions, for convenience, we simply denote $`h=1/T`$. Here the coupling constant $`J`$ describes the nearest-neighbor interactions. In case of $`J=0`$, the original Fredrickson model is recovered. As above mentioned, the dynamic evolution of the generalized SFM$`[2,2]`$ is subjected to a topological constraint that a spin flip is only possible if
$$\frac{1}{2}\underset{i}{}(1+s_i)f.$$
(2)
In our simulations, the Metropolis algorithm is used and $`f`$ takes its typical value $`f=d`$ with $`d`$ being the space dimension, here we chose $`d=2`$. To assure that the system evolves into the physical section of the phase space, the initial configuration is always taken to be $`s_i1`$ which means that we start from the complete liquid–like state. After the system has reached its equilibrium, we measure the auto-correlation function
$$A(t)=\frac{1}{L^d}<\underset{i}{}s_i(t^{})s_i(t+t^{})>$$
(3)
with $`L`$ being the lattice size. Practically an average over $`t^{}`$ is made in the numerical measurements. The lattice sizes are taken to be $`L=50`$ or $`L=100`$ depending on $`h`$ and $`J`$. Up to the time regime of our simulations, no visible finite size effect has been observed. To achieve reliable results and estimated statistical errors, we have performed five runs of simulations. Total samples for average range from $`\mathrm{50\; 000}`$ to $`\mathrm{500\; 000}`$. More samples are for larger values of $`h`$ and/or $`J`$.
## III results and discussion
In the low temperature regime, the original Fredrickson Andersen model gives rise to a drastical enhancement of the relaxation time which is characterized inevitably with glassy materials. As demonstrated in there is no indication for a real glass transition or a critical temperature as predicted by mode–coupling theory . For large time $`t`$, empirical approaches suggest a stretched exponential decay of the auto-correlation function
$$A(t)exp[(t/\tau )^\gamma ].$$
(4)
where the exponent $`\gamma `$ is presumable not an universal exponent, i.e. weakly depending on $`h`$. As a function of the $`h`$ (inverse temperature), the relaxation time $`\tau `$ increases faster than according to an exponential law $`\mathrm{ln}\tau c+h`$ manifested as a non-Arrhenius behavior but there is no singularity $`\tau \mathrm{}`$ at finite temperatures as suggested by the Williams–Landel–Ferry relation . The exponent $`\gamma `$ offers a weak dependence on $`h`$ which is also confirmed by our simulations presented below. For the generalized SFM$`[2,2]`$ we are interested in the role of the extra coupling $`J`$. Physically, a positive exchange coupling $`J>0`$ and $`h>0`$ intend to support the creation of solid–solid pairs which are partially frozen in, e.g. such a coupling tends to enhance the relaxation time.
We observe that for fixed $`h`$, even if it is small, the auto-correlation decays in a stretched exponential form for large time $`t`$ and more interestingly, the relaxation time $`\tau `$ increases also rapidly following a non-Arrhenius law when the coupling $`J`$ increases. The exponent $`\gamma `$ exhibits also a weak dependence on the coupling $`J`$. In Fig. 1, the auto-correlation for $`h=0.40`$ and for different values of the coupling $`J`$ is displayed in semi-log scale with lines of circles. Obviously the decay is not an exponential one. The dotted lines are the stretched exponential fit to the curves. We see clearly the fit is rather good. The resulting relaxation time $`\tau `$ and the exponent $`\gamma `$ for different $`h`$ and $`J`$ are listed in Table I. For comparison, results for the original Fredrickson Andersen model ($`J=0`$) are also included. In Fig. 2, we have plotted the correlation time $`\tau `$ as a function of the coupling $`J`$ for different values of $`h`$ in semi-log scale. Clearly it is a non-Arrhenius behavior.
Our model has two parameters $`h`$ and $`J`$. For large $`h`$ and/or $`J`$, a strong freezing process manifested in a strong slowing down of the relaxation time is observed and the auto-correlation shows similar dependence on both $`h`$ and $`J`$, respectively. Alternatively to the autocorrelation function let us consider the magnetization $`M(h,J)`$. This quantity is an essential one also within our re-interpretation of the kinetic Ising model as an appropriate candidate to describe glasses. The magnetization corresponds to the density of the immobile particles. In Fig. 3, the dependence of the relaxation time $`\tau `$ and the exponent $`\gamma `$ on $`M(h,J)`$ are depicted for different couplings $`J`$ and different fields $`h`$. A nice collapse of the data is observed. A non-zero coupling $`J`$ practically induces a short range spatial correlation. However, our results show that such a short spatial correlation length does not change dramatically the properties of the glass system.
The interaction can be also anti-ferromagnetic, i.e. with a negative coupling constant $`J`$. In this case, the relaxation time $`\tau `$ decreases when the magnitude of $`J`$ becomes larger. This can be seen in the last block of Table I. However, we again find a nice collapse of th data with negative $`J`$ and positive $`J`$ when the magnetization $`M`$ is chosen to be the scaling variable. This is shown in Fig.4 (a). The situation is slightly different for the exponent $`\gamma `$. In Fig. 4 (b), for small $`|J|`$, it joins to the data points with positive $`J`$. But for bigger $`|J|`$, i.e. small $`M`$ and with short relaxation times $`\tau `$, the exponent $`\gamma `$ decreases rather than increases as in the case of positive $`J`$. This phenomenon is also understandable. A static antiferromagnetic coupling favors the coexistence of both liquid and solid like regions in the neighborhood. But the limit of $`J\mathrm{}`$ for fixed $`h`$ does not exactly correspond to a high temperature state.
## IV conclusions
We obtain as a main result that the dynamics of the generalized SFM$`[2,2]`$ is controlled only by the magnetization $`M(h,J)`$, i.e. the density of up (or down) spins. When we plot the correlation time $`\tau `$ as a function of $`M(h,J)`$, all data points for different values of $`h`$ and $`J`$ collapse to a single curve. The collapse of the data points for the exponent $`\gamma `$ is also observed except for the case with a strong antiferromagnetic coupling. These results show that the stretched exponential behavior is rather universal.
Our simulations have not yet covered the regime near the critical point of the standard Ising model ($`h0`$ but $`hJ`$ remains finite at its critical value). In this critical regime, both the glass transition and the second order phase transition take place. To study the mixed critical behavior of two phase transitions is an interesting extension of the present work.
Finally, it should be remarked that the empirical stretching exponent $`\gamma `$ is rather close to 0.5. This result is in agreement with recent analytical calculations .
|
no-problem/9812/hep-ph9812534.html
|
ar5iv
|
text
|
# Gamma photons from parametric resonance in neutron stars
## I Introduction
Neutron star collisions rank among the most energetic events expected to take place in the Universe, which makes them a natural candidate for the source of observed gamma-ray bursts (GRB) . The discovery of the afterglow associated with some of the GRB and the isotropy of the GRB both support the hypothesis that the GRB originate at cosmological distances with a redshift of order one. The total energy $`10^{53}`$ erg released when the two neutron stars merge is sufficient to explain the GRB, although it is not clear what fraction of that energy is converted to gamma rays. Another puzzling feature of the GRB is their non-thermal spectrum.
Understanding the physics of the explosion that follows the collision of two neutron stars is of great importance and is the subject of intense studies . The approach and the early stages of the interaction of the two neutron stars are accompanied by powerful acoustic shock waves that propagate through nuclear matter and eventually dissipate their energy into heat. During this period of time, the repetitive superconducting phase transitions can take place in part of the star’s volume due to the sharp dependence of the proton energy gap on density. The relaxation of the proton condensate to the potential minimum can be accompanied by a non-thermal resonant production of gamma rays in the MeV energy range. If the parametric resonance is efficient, the power transferred to the coherent gamma quanta may exceed $`10^{52}`$ erg/s. Eventually, the nuclear matter is heated above the critical temperature $`0.5`$ MeV and the production of the gamma rays comes to a halt.
The photons produced inside a neutron star cannot decay through pair production $`\gamma e^+e^{}`$ because the degenerate electrons have the chemical potential in excess of 100 MeV, far greater than the photon energy. The decay into the electron-positron pairs is, therefore, prohibited by the Pauli exclusion principle. The photons undergo a Compton scattering off the electrons near the Fermi surface which is sufficiently strong to keep the photons from escaping. While such scattering (or comptonization) may change the spectrum of the gamma-component, the number of photons remains the same. Since the extreme electron degeneracy is maintained even at the highest temperatures achieved in the fireball , the gamma-ray component is present in the nuclear matter at the time the latter is dispersed by the explosion.
The latter can have several consequences. In particular, $`\gamma `$-quanta, abundantly present in nuclear matter at the onset of the explosion, can be released when the fireball erupts. The detailed investigation of this signal lies outside the scope of this paper. However, it is plausible that the gamma-rays emitted at that point would have a non-thermal spectrum. At later times, when the fireball reaches a high temperature, other sources of gamma emission become dominant. We predict, therefore, a qualitative difference in the spectrum of gamma-rays emitted during the first milliseconds of the collision.
In this paper we will not attempt to understand the emission of gamma-rays from the fireball. Instead, we will concentrate on the phenomenon of resonant production of photons which transfers a fraction of the gravitational energy into the gamma-quanta inside the neutron star. This process is of fundamental interest on its own and it may have important consequences.
## II Basic idea
Nuclear matter in the interior of a neutron star may be superconducting. The existence of the superfluid proton condensate depends on several theoretical assumptions, some of which may be hard to justify. The proper usage of the many-body techniques and the choice of macroscopic degrees of freedom are by no means obvious. However, assuming that the theoretical framework of Refs. is valid, one can elucidate some generic features of proton superfluidity. The most important one for us is that the energy gap depends sharply on density, as shown in Fig. 1. An acoustic shock wave passing through the star can produce significant changes in the density and cause repetitive superconducting phase transitions occurring with periods of order the acoustical time scale $`\tau _a10^7`$ to $`10^3`$ s . The relaxation time of the proton condensate is of order MeV<sup>-1</sup>, or $`10^{20}\mathrm{s}\tau _a`$, and therefore the system quickly settles in the potential minimum after the passage of each shock wave.
In the presence of a magnetic field the superconducting phase transition is first-order and occurs in two stages. First, a bubble of the superconducting phase nucleates and expands; then the proton condensate may oscillate around the minimum. If the magnetic field is close to critical, $`BB_c`$, the two minima in the potential shown in Fig. 2 are nearly degenerate and no subsequent coherent motion of the condensate $`\varphi `$ takes place after the transition through bubble nucleation. However, for a smaller magnetic field, the so called “escape point” $`\varphi _e`$ is different from the global minimum $`\varphi _0`$ and the scalar condensate may oscillate around the minimum.
During the oscillations of the order parameter $`\varphi `$, a photon has effectively a time-dependent mass proportional to the value of $`\varphi `$. This may, in some cases, signal a copious production of gamma-quanta through parametric resonance . In this paper we study the parametric resonance both analytically and numerically for different for some sample values of nuclear density and magnetic field that ranges from $`10^{12}`$ to $`10^{15}`$ G.
## III Potential and equation of motion
A natural framework for describing the relaxation of the proton condensate after the phase transition is time-dependent Ginzburg-Landau theory . The applicability of such description is limited to cases where the Joule heat loses are small . Otherwise, the interactions become essentially non-local and a simple equation of motion for a scalar field seizes to be valid.
Here we will assume the validity of the time-dependent Ginzburg-Landau theory and will apply it to the superconducting proton condensate inside a neutron star. One should take this approach with a grain of salt given the lack of empirical knowledge with respect to the proton superconductivity in nuclear matter. However, we hope that – crude an approximation this may be – it may help identify the main features of the resonant production of gamma-rays.
We write the equation of motion for the order parameter as
$$\ddot{\varphi }+\frac{8\epsilon _F}{3c}\dot{\varphi }\frac{2\epsilon _F}{3cm_{}}^2\varphi +U^{}(\varphi )=0,$$
(1)
where $`\epsilon _F=p_F^2/2m_{}`$ is the Fermi energy of the proton condensate and $`c=(28\zeta (3)/3\pi ^3)\epsilon _F/T_c`$ is a constant characterizing the condensate. The friction term then has a magnitude $`8\epsilon /3c7.37T_c4.2\mathrm{\Delta }_0`$, of order a few MeV, which is comparable to the oscillating frequency $`\omega `$ of the condensate around its minimum $`\varphi _0`$, and therefore cannot be ignored. On the other hand, the gradient term in Eq. (1) is negligible in our case , and thus we are dealing with a (locally) homogeneous condensate.
The effective potential for the order parameter $`\varphi `$ can be determined from the physical properties of the condensate. After the acoustic wave has restored the symmetry, the energy density subsides and a new superconducting phase transition takes place. The difference in energy density associated with the proton condensate is
$$\mathrm{\Delta }U(\varphi _0)=\frac{m_{}p_F}{4\pi ^2}\mathrm{\Delta }_0^2,$$
(2)
where $`\mathrm{\Delta }_0`$ is the proton energy gap, and $`m_{}`$ is the effective proton mass in nuclear matter, which is somewhat lower than the bare proton mass $`m_p`$. The equilibrium value of the order parameter is given by $`\varphi _0^2=n_p/2m_{}`$, where $`n_p=Y_p\rho /m_p`$ is the number density of protons inside the neutron star, which make up a fraction $`Y_p0.03`$ of all baryons.
The way the superconducting phase transition occurs depends very strongly on the presence of a magnetic field $`B`$. Such a field creates a barrier in the effective potential between the symmetric and the superconducting phases. The height of the barrier is approximately given by the total energy density in the magnetic field,
$$\mathrm{\Delta }U(\varphi _1)=\frac{B^2}{4\pi }200\left(\frac{B}{10^{15}\mathrm{G}}\right)^2\mathrm{MeV}^4.$$
(3)
This barrier makes the phase transition first-order, with creation of bubbles of superconducting phase. After the bubble nucleation, the proton condensate oscillates and soon settles (due to friction) at its equilibrium value $`\varphi _0`$. We have shown in Fig. 2 the effective potential $`U(\varphi )`$ for a region inside the neutron star with matter density $`\rho =10^{14}`$ g/cm<sup>3</sup> and magnetic field $`B=10^{15}`$ G, which corresponds to $`\mathrm{\Delta }_00.35`$ MeV, $`\varphi _050.3`$ MeV, $`\mathrm{\Delta }U(\varphi _0)3000`$ MeV<sup>4</sup> and $`\mathrm{\Delta }U(\varphi _1)200`$ MeV<sup>4</sup>.
Depending on the relative size of the friction term, the condensate settles at $`\varphi _0`$ after a few oscillations, or without oscillations at all. We have shown in Fig. 3 the oscillations of the order parameter $`\varphi `$ in the case of the parameters of Fig. 2, which give a friction coefficient of order $`4.2\mathrm{\Delta }_00.31\omega `$ and $`\omega 4.7`$ MeV. This allows a few oscillations to take place, which is crucial for the resonant production of photons, as we will see in the next Section.
## IV Resonant photoproduction
The proton condensate comprises Cooper pairs of charge $`2e`$, which couple to the electromagnetic field through the term $`(2e)^2\varphi ^2A_\mu A^\mu `$ in the unitary gauge, in which the scalar field $`\varphi `$ is real. As the condensate oscillates around $`\varphi _0`$, it induces an effective periodic mass, which in some cases may stimulate parametric resonant production of photons . The mechanism is similar to that discussed in connection with axion clumps , as well as preheating after inflation , where explosive production of bosons may occur under special circumstances .
We will write the gauge field as $`A_\mu (x)=\chi (x)e_\mu (x)`$, where $`e_\mu (x)`$ is a polarization vector and $`\chi (x)`$ can be expanded in Fourier modes, $`\chi _k(t)`$, that satisfy the evolution equation
$$\ddot{\chi }_k+\left(k^2+2(2e)^2\varphi ^2(t)\right)\chi _k(t)=0,$$
(4)
with an effective mass proportional to the condensate
$$\varphi (t)\varphi _0\left(1+\mathrm{\Phi }\mathrm{exp}(ϵ\omega t/2)\mathrm{sin}\omega t\right),$$
(5)
where $`\mathrm{\Phi }=\varphi _e/\varphi _0`$ and $`ϵ4.2\mathrm{\Delta }_0/\omega `$ is the decay constant for the condensate oscillations. Equation (4) can be written as a Mathieu equation with coefficients ($`z=\omega t/2`$)
$`A_k`$ $`=`$ $`{\displaystyle \frac{4k^2}{\omega ^2}}+4q_0,`$ (6)
$`q(z)`$ $`=`$ $`4q_0\mathrm{\Phi }\mathrm{exp}(ϵz),`$ (7)
$`q_0`$ $`=`$ $`8e^2{\displaystyle \frac{\varphi _0^2}{\omega ^2}}=32\pi ^2\alpha _{\mathrm{em}}{\displaystyle \frac{\varphi _0^2}{\omega ^2}}.`$ (8)
Note that the effective parameter $`q(z)`$ that determines the strength of the resonance decreases exponentially with time. If $`ϵ`$ is too large, the parametric resonance is weak, and this mechanism is inefficient. On the other hand, if $`ϵ`$ is small, then the condensate oscillates several times and causes an explosive production of gamma rays with energy of a few MeV.
At density $`\rho 10^{14}`$ g/cm<sup>3</sup> and magnetic field $`B10^{15}`$ G, parametric amplification of the MeV photons can take place. In this case, $`ϵ=0.31`$ and $`4q_0340`$. We have plotted $`\mu _k`$ as a function of $`k`$, after one oscillation of the condensate, in Fig. 4, from which we can deduce the spectrum
$$n_k\frac{1}{2}\mathrm{exp}(\mu _k\omega t).$$
(9)
Nuclear matter is opaque for photons with energy of order the electron chemical potential, which limits the spectral band in which the modes are amplified. The spectrum $`n_k`$ has, therefore, an ultraviolet cut-off at the momentum associated with the electron Fermi energy inside the neutron star, beyond which photons are absorbed and the resonance shuts down. This occurs at wavelengths $`k_c100m_e=51`$ MeV. For the model discussed above, with frequency of oscillations $`\omega =4.7`$ MeV, this corresponds to $`k/\omega 11`$.
Furthermore, after a few oscillations, backreaction occurs, that is, the number of created photons is so large that it dominates the frequency of oscillations of the proton condensate, $`m^2=\omega ^2+8e^2\chi ^2`$. The backreaction sets in when
$$\chi ^2=\frac{1}{2\pi ^2}_0^{k_c}𝑑kk^2\frac{n_k(t)}{\omega _k}\frac{\omega ^2}{8e^2}\omega ^2,$$
(10)
where $`\omega _k^2=k^2+8e^2\varphi ^2(t)`$ is the effective frequency of the mode $`k`$, and the integration is up to the physical cut-off, $`k_c`$. For the model in hand, we find that backreaction takes place after about one oscillation, much before the friction term in Eq. (1) has significantly decreased the oscillation amplitude of the condensate.
The total energy density in photons produced during the resonance is, therefore,
$$\rho _\gamma (t)=\frac{1}{2\pi ^2}_0^{k_c}𝑑kk^2\omega _kn_k(t),$$
(11)
We have plotted this energy density in Fig. 5. $`\rho _\gamma (t)`$ reaches an asymptotic value after a few oscillations. Because of backreaction, the energy density produced is that after the first oscillation, of order $`10^5\mathrm{MeV}^4`$. It may seem like a large value, but actually the energy density produced via this mechanism is just a small fraction of the neutron star density, $`\rho _\gamma 8\times 10^4\mathrm{MeV}^410^4\rho `$.
We have analyzed the resonant photoproduction numerically for magnetic field from $`10^{12}`$ G to $`10^{16}`$ G and found no significant deviation from the sample values described above. The main effect of the magnetic field is to produce a potential barrier at $`\varphi =\varphi _1`$, as shown in Fig. 2. A large magnetic field may, however, have some other effects on nuclear matter. For example, it can modify the particle composition, in particular the distribution of protons and electrons . This, in turn, can affect the resonant photoproduction. Other parameters, such as the electron chemical potential, the proton fraction $`Y_p`$, etc., can differ significantly from the sample values we took. However, our numerical analyses show that a resonant production of gamma rays from the oscillating charged condensate is possible for a wide variety of parameters.
## V Implications for the observation of gamma-ray bursts
The MeV-energy gamma rays produced in each cycle of oscillations cannot decay the way they would decay in vacuum, through electron-positron pair production, because the electrons are highly degenerate inside the neutron star. Pauli exclusion principle prevents a production of electrons with energies less than the electron chemical potential, which is of order 100 MeV. Compton scattering off the electrons and protons near the Fermi surface is kinematicly suppressed but is not forbidden. The corresponding mean free path is $`\lambda 10^9`$ cm. Comptonization generally preserves the number of photons and tends to equalize the temperatures of electrons and photons . Given enough time, the photons would diffuse out of the star. However, since the merger of the two neutron stars is characterized by very short time scales, all the gamma rays non-thermally produced by the coherent oscillations of the proton condensate remain inside the star when the fireball erupts. At that point they can leak out and be observed as a short gamma-ray burst, or more likely, as a component of the GRB.
The energy stored in the non-thermal bath of gamma ray photons inside the neutron star is $`E_\gamma =\rho _\gamma V3\times 10^{50}`$ erg. If the collision of neutron stars releases these photons within a time scale of order $`1\mathrm{ms}\tau 0.1`$ s, then the power generated is of order $`10^{52\pm 1}`$ erg/s, which corresponds to the power emitted in the observed gamma ray bursts .
The resonant photoproduction is active in a spherical shell with density $`10^{14}10^{15}`$ g/cm<sup>3</sup> that contains most of the neutron star mass , and where the acoustically driven superconducting phase transitions can take place.
## VI Conclusion
Strong mechanical shock waves, such as those expected to be generated by a collision of two neutron stars, can cause repetitious superconducting phase transitions in nuclear matter. The relaxation of the proton condensate to its potential minimum can result in the non-thermal resonant production of $`\gamma `$-quanta in the MeV energy range.
There are some important consequences. The presence of energetic gamma quanta in nuclear matter, relatively transparent to photons thanks to electron degeneracy, can affect the equation of state. This, in turn, can affect the dynamics of the neutron star coalescence.
In addition, the gamma rays trapped until the onset of the fireball but then released by the explosion, can also contribute to gamma-ray bursts.
## Acknowledgements
J.G.B. is supported by a Research Fellowship of the Royal Society. A. K. is supported in part by the US Department of Energy grant DE-FG03-91ER40662.
|
no-problem/9812/astro-ph9812062.html
|
ar5iv
|
text
|
# Evolution of DA white dwarfs in the context of a new theory of convection
## 1 introduction
White dwarf stars with hydrogen lines (spectral type DA) constitute the most numerous group of the observed white dwarfs. Since the first DA white dwarf was reported to exhibit multiperiodic luminosity variations (see McGraw 1979), the interest in these objects has greatly increased. Rapid progress in this field has been possible thanks to the development of powerful theoretical tools as well as to a increasing degree of sophistication in observational techniques. Over the years, various studies have presented strong evidence that pulsating DA white dwarfs (DAV) or ZZ Ceti stars represent an evolutionary stage in the cooling history of the majority of, if not all, DA white dwarfs (Weidemann & Koester 1984 and references therein).
Studies carried out notably by Dolez & Vauclair (1981) and Winget et al. (1982) on the basis of detailed non - radial, non - adiabatic pulsation calculations, demonstrated that the $`\kappa `$-mechanism operating in the hydrogen partial ionization zone is responsible for the g(gravity)-mode instabilities in ZZ Ceti stars. From then on, pulsating white dwarfs have captured the interest of numerous investigators, who have employed the powerful technique of white dwarf seismology to derive fundamental parameters of these stars, such as the stellar mass, chemical composition and stratification of the outer layers (see, for example, Winget et al. 1994 and Bradley 1996 in the context of DB and DA, respectively). In particular, pulsation calculations have shown that the theoretical determination of the location of the blue (hot) edge of the instability strip (which marks the onset of pulsations) in the Hertzsprung - Russell diagram is mostly sensitive to the treatment of convection in envelope models. Accordingly, a comparison with the location of the hottest observed ZZ Ceti stars allows us to obtain valuable information about the efficiency of convection in such stars (Winget et al. 1982; Tassoul, Fontaine & Winget 1990; Bradley & Winget 1994; Bradley 1996 and references cited therein).
Unfortunately, the determination of atmospheric parameters of ZZ Ceti stars is strongly dependent on the assumed convection description in model atmosphere calculations, against which the observed spectrum is compared. The theoretical analysis carried out by Bergeron, Wesemael & Fontaine (1992a) is particularly noteworthy in this regard. In fact, these authors demonstrated that the predicted emergent fluxes, color indices and equivalent widths of DA white dwarfs, notably of the ZZ Ceti stars, are very sensitive to the details of the parametrization of the mixing - length theory (MLT) of convection. In particular, Bergeron et al. (1992a) concluded that the uncertainties in the observational definition of the blue edge of the ZZ Ceti instability strip brought about by different convective efficiencies may be appreciable, irrespective of the observational technique that is used.
Needless to say, the employment of MLT to deal with convection represents a serious shortcoming in theoretical studies of white dwarfs. In particular, MLT approximates the whole spectrum of eddies necessary to describe the stellar interiors by a single large eddy. This representation introduces in the description of the model certain free parameters (Böhm \- Vitense 1958) which, in most stellar studies, are reduced to a single one: the distance $`l`$ travelled by the eddy before thermalizing with the surrounding medium. $`l`$ is written as $`l=\alpha H_\mathrm{p}`$ where $`H_\mathrm{p}`$ is the pressure scaleheight and $`\alpha `$ is the free parameter. Obviously, the existence of such an adjustable parameter prevents stellar evolution calculations from being completely predictive. In white dwarf studies, several parametrizations of the MLT are extensively employed. Specifically, the ML1, ML2 and ML3 versions, which differ in the choice of the coefficients appearing in the expressions of the convective flux, average velocity and convective efficiency (see Tassoul et al. 1990 for details) are associated, respectively, with increasing convective efficiency (for the ML1 and ML2 versions $`\alpha =1`$, while ML3 is the same as ML2 but with $`\alpha =2`$).
Over the last two decades, an unparalleled effort has been devoted to the observational determination of the location of the blue edge of the instability strip of the ZZ Ceti stars (see Wesemael et al. 1991 for a review). Based on different observational techniques, the majority of the studies cited by Wesemael et al. point towards a temperature of the blue edge, defined by the ZZ Ceti star G117-B15A, of approximately 13000 K. New observational data and improved model atmospheres tend to indicate a significantly lower temperature for the blue edge. In fact, from ultraviolet observations performed with the Hubble Space Telescope, Koester, Allard & Vauclair (1994) have shown that model atmospheres calculated with the ML1 parametrization do not fit satisfactorily to the ultraviolet spectrum of G117-B15A. Instead, a more efficient convection, as given by the ML2 version, is required to explain the observed spectrum. On the basis of this result, Koester et al. derived for G117-B15A an effective temperature of 12250 $`\pm `$ 125 K. In view of the fact that the ultraviolet spectra of other candidates for the blue edge of the ZZ Ceti instability strip are similar to that of G117-B15A, Koester et al. concluded that the blue edge temperature is likely to be around 12250 K, a value which is considerably lower than found in previous studies.
In a still more recent analysis, Bergeron et al. (1995) inferred an even lower effective temperature for G117-B15A. More precisely, on the basis of new optical and ultraviolet analyses of ZZ Ceti stars, Bergeron et al. succeeded in constraining the convective efficiency used in model atmosphere calculations of these stars. Indeed, they demonstrated that a parametrization less efficient than ML2, particularly the ML2/$`\alpha `$=0.6 version (intermediate in efficiency between ML1 and ML2) yields ultraviolet temperatures that are completely consistent with the optical determinations. In this respect, they derived for G117-B15A an effective temperature of 11620 K. In the sample analysed by Bergeron et al., the blue edge of the instability strip is defined by the star G226-29, for which they obtained a temperature of 12460 K and a surface gravity of $`\mathrm{log}g=8.29`$. Finally, Fontaine et al. (1996) presented a reanalysis of the observed pulsation modes of G117-B15A that reinforces the conclusions of Bergeron et al. (1995) on the atmospheric parameters of this variable white dwarf. Fontaine et al. concluded that a model atmosphere of G117-B15A characterized by an effective temperature of $``$ 11500 K and a ML2/$`\alpha `$=0.6 convection is consistent with optical and ultraviolet spectroscopic observations as well as with the observed pulsation amplitudes in different wavelength bands.
Another conclusion drawn by Bergeron et al. (1995) (see also Wesemael et al. 1991) worthy of comment is related to the fact that convective efficiency in the upper atmosphere appears to be quite different from that in the deeper envelope. In fact, according to non - adiabatic pulsation calculations (Bradley & Winget 1994 and references cited therein), a more efficient convection than that provided by ML2/$`\alpha `$=0.6 is needed in the deeper envelope to explain the observed DA blue edge. This result, which reflects the inability of MLT to describe convection throughout the entire outer convection zone adequately, has indeed been borne out by recent detailed hydrodynamical simulations of convection in ZZ Ceti stars (see later in this section). Other objections have been raised against MLT. For instance, the value of $`\alpha `$, which is usually derived from solar radius adjustments, is larger than unity, in contrast to the basic postulates of MLT. In addition to this, the procedure of applying the solar $`\alpha `$ parameter to other stars is not entirely satisfactory, at least in the context of red giant stars for which a wide range of $`\alpha `$ values is needed (Stothers & Chin 1995, 1997). Finally, MLT does not fit laboratory convection data (Canuto 1996).
Fortunately, a considerable effort has been devoted in recent years to improving some of the basic postulates of MLT. In particular, Canuto & Mazzitelli (1991, 1992, hereafter CM) and more recently, Canuto, Goldman & Mazzitelli (1996, hereafter CGM) developed two convection models based on the full - spectrum turbulence (FST) theory, which represent a substantial progress in modelling stellar convection (we refer the reader to Canuto & Christensen - Dalsgaard 1997 for a recent review on local and non \- local convection models). Both of the treatments resolve essentially two main shortcomings of the MLT
(i) The MLT one - eddy approach is replaced by a full spectrum of turbulent eddies derived on the basis of amply tested, modern theories of turbulence. As discussed by CM, the one - eddy approximation is only justified for viscous fluids but becomes completely inadequate for almost - inviscid stellar interiors.
(ii) There are no adjustable parameters.
These features render the CM and CGM models substantially more solid and complete than MLT. The CM model in particular has successfully passed a wide variety of laboratory and astrophysical tests, performing in all cases much better than MLT. With regard to the astrophysical case, the CM model has been applied to the study of different kind of stars and evolutionary phases. Applications include the solar model (see CM), the effective temperature of which is fitted within $`0.5`$ per cent accuracy; red giants (D’Antona, Mazzitelli & Gratton 1992 and Stothers & Chin 1995, 1997), for which MLT may require various values of $`\alpha `$ for different stars, helioseismology (Paternò et al. 1993 and Monteiro, Christensen-Dalsgaard & Thompson 1996 amongst others), low - mass stars (D’Antona & Mazzitelli 1994) and stellar atmospheres (Kupka 1996). In the white dwarf context, the CM model has been shown to be a very valuable tool as well. In fact, Althaus & Benvenuto (1996) (see also Mazzitelli & D’Antona 1991) studied the evolution of carbon \- oxygen white dwarfs with helium - rich envelopes (spectral type DB) and found that the CM model predicts theoretical blue edges for the DB instability strip in good agreement with the observations of Thejll, Vennes & Shipman (1991). Mazzitelli (1995) has also applied the CM model to the study of convection in DA white dwarfs. More recently, Althaus & Benvenuto (1997a) used the CM model to study the evolution of helium white dwarfs of low and intermediate masses.
CGM have recently improved the CM model by developing a self-consistent model for stellar turbulent convection, which, like the CM model, includes the full spectrum of eddies but computes the rate of energy input self - consistently. More precisely, CGM resolve analytically a full turbulence model in the local limit, taking into account the fact that the energy input from the source (buoyancy) into the turbulence depends on both the source and the turbulence itself. At low and intermediate convective efficiencies, the CGM gives rise to higher convective fluxes than those given by the CM model. CGM find that the main sequence evolution of a solar model does not differ appreciably from the CM predictions but demands a smaller overshooting to fit the solar radius, in agreement with new observational data. D’Antona, Caloi & Mazzitelli (1997) have also applied the CGM model to study the problem of globular cluster ages. In the white dwarf domain, the CGM model also fares better than the CM model (Althaus & Benvenuto 1997b and Benvenuto & Althaus 1997). Finally, Benvenuto & Althaus (1997b) have applied the CGM model to study the evolution of low \- and intermediate - mass helium white dwarfs with hydrogen envelopes.
In order to assess to what extent the MLT is able to provide a reliable description of convection in the outer layers of DA white dwarfs, Ludwig, Jordan & Steffen (1994) have performed detailed two - dimensional hydrodynamical simulations of convection at the surface of a ZZ Ceti star. They demonstrated that the temperature profile in the outer layers of a $`T_{\mathrm{eff}}`$ = 12600 K, $`\mathrm{log}g`$ = 8.0 DA model cannot be reproduced with model atmospheres calculated with MLT and with a single value of $`\alpha `$. Instead, the value of $`\alpha `$ is to be varied from 1.5 in the upper atmosphere to 4 in much deeper layers in order for the hydrodynamical temperature profile to be reasonably well represented. Another result obtained by Ludwig et al. is that their hydrodynamical simulations lead to convective zones that are too shallow for the $`\kappa `$, $`\gamma `$ mechanism to give rise unstable g modes, implying a blue edge for the ZZ Ceti instability strip significantly cooler than 12600 K. This important finding has recently been borne out by Gautschy, Ludwig & Freytag (1996) on the basis of non - radial, non - adiabatic pulsation calculations. Gautschy et al. included in their analysis Ludwig et al.’s hydrodynamical simulations to derive the structure of the convective layers of their white dwarf models. This feature represents substantial progress compared with earlier investigations, which employ MLT to deal with convection. Gautschy et al. are unable to find a theoretical blue edge for the ZZ Ceti instability strip consistent with the observations of Bergeron et al. (1995) and conclude that, according to their calculations, it lies between 11400 K and 11800 K (for a model with $`\mathrm{log}g=8`$).
The present study is devoted to presenting new calculations of the evolution of carbon - oxygen DA white dwarfs considering the model for stellar turbulent convection of CGM as well as the various parametrizations of MLT commonly used in the study of these stars. Attention is focused mainly on those evolutionary stages corresponding to the ZZ Ceti effective temperature range. The calculations are carried out with a detailed white dwarf evolutionary code in which we include updated equations of state, opacities and neutrino emission rates. Furthermore, we take into account crystallization, convective mixing and hydrogen burning. To explore the sensitivity of the results to various input model parameters, we vary both the model mass from 0.50 - 1.0 $`M_{\mathrm{}}`$, which covers most of the observed mass distribution of DA white dwarfs (Bergeron, Saffer & Liebert 1992b, Marsh et al. 1997), and the hydrogen layer mass $`M_\mathrm{H}`$ in the interval $`10^{13}M_\mathrm{H}/M10^4`$ where $`M`$ is the model mass. Finally the hydrogen - helium transition zone is assumed to be almost discontinuos. Details of our evolutionary code and its main physical constituents are presented in Section 2. Section 3 is devoted to analysing our results and, finally, in Section 4 we summarize our findings.
## 2 computational details and input physics
The white dwarf evolutionary code we employed in this study was developed by us independently of other researchers and it has been used to study different problems connected to white dwarf evolution as well as the cooling of the so - called strange dwarf stars (Benvenuto & Althaus 1996). Details of the code and its constitutive physics can be found in Althaus & Benvenuto (1997a) and Benvenuto & Althaus (1997, 1998). In what follows we restrict ourselves to a few brief comments.
The equation of state for the low - density regime is that of Saumon, Chabrier & Van Horn (1995) for hydrogen and helium plasmas. The treatment for the high - density, completely ionized regime appropriate for the white dwarf interior includes ionic and photon contributions, coulomb interactions, partially degenerate electrons and electron exchange and Thomas - Fermi contributions at finite temperature. Radiative opacitites for the high - temperature regime ($`T8000`$ K) are those of OPAL (Iglesias & Rogers 1993) with metallicity $`Z=0.001`$, whilst for lower temperatures we use the Alexander & Ferguson (1994) molecular opacities. Molecular effects become relevant at temperatures as high as 5000 K and below 2500 K they are dominat (Alexander & Ferguson 1994). We shall see that convective mixing at low effective temperatures may in some cases change the outer layer composition from almost pure hydrogen to almost pure helium. This being the case, we had to rely, in the low - temperature regime, on the older tabulation of Cox & Stewart (1970) for helium composition. Conductive opacities for the low - density regime are taken from Hubbard & Lampe (1969) as fitted by Fontaine & Van Horn (1976). Conductive opacities for the liquid and crystalline phases, and the various mechanisms of neutrino emission relevant to white dwarf interiors are taken from the works of Itoh and collaborators (see Althaus & Benvenuto 1997a and references cited therein). It is worth mentioning that neutrino cooling is dominant during the hot phases of white dwarf evolution but they become completely negligible at the ZZ Ceti effective temperature range.
The procedure we followed to generate the initial models of different masses is described in Benvenuto & Althaus (1995). Such initial models were constructed starting from a $`0.55M_{\mathrm{}}`$ carbon - oxygen white dwarf model kindly provided to us by Professor Francesca D’Antona. We assumed the same core chemical composition (consisting of a mixture of carbon and oxygen, see Fig. 1 of Benvenuto & Althaus 1995) for all of them, notwithstanding the changes in the internal chemical composition that are expected to occur for models with different masses because of differences in the pre - white dwarf evolution of progenitors objects. The carbon - oxygen core of our models is surrounded by an almost pure helium envelope, the mass of which ($`M_{\mathrm{He}}`$) was varied in the range $`10^6M_{\mathrm{He}}/M10^2`$. In DA white dwarfs there is an almost pure hydrogen envelope on the top of the helium layers. Unfortunately, the mass of this hydrogen envelope is only weakly constrained by pre - white dwarf evolutionary calculations (D’Antona & Mazzitelli 1991) and by non - adiabatic pulsation studies (Fontaine et al. 1994). In recent years, however, strong evidence has been accumulated in favour of the idea that some ZZ Ceti stars appear to have thick hydrogen layers (Fontaine et al. 1994). In the present study, we decided to treat the mass of the hydrogen envelope as basically a free parameter within the range $`10^{13}M_\mathrm{H}/M10^4`$. We refer the reader to Benvenuto & Althaus (1998) for details about the procedure we follow to add a hydrogen envelope at the top of our models. We want to mention that we assumed the hydrogen/helium transition zone to be almost discontinuous.
We also included in our evolutionary calculations the release of latent heat during crystallization (see Benvenuto & Althaus 1997) and the effect of the convective mixing. We shall see in particular that the onset of crystallization in our more massive models occurs at effective temperatures close to the observed blue edge. Finally, we also considered the effects of hydrogen burning by including in our numerical code the complete network of thermonuclear reaction rates for hydrogen burning corresponding to the proton - proton chain and the CNO bi - cycle. Nuclear reaction rates are taken from Caughlan & Fowler (1988) and $`\beta `$ \- decay rates from Wagoner (1969), taking into account the corrections for their Q - values resulting from the effect of neutrino losses. Electron screening is from Wallace, Woosley & Weaver (1982). We consider the following chemical species: <sup>1</sup>H, <sup>2</sup>H, <sup>3</sup>He, <sup>4</sup>He, <sup>7</sup>Li, <sup>7</sup>Be, <sup>8</sup>B, <sup>12</sup>C, <sup>13</sup>C, <sup>13</sup>N, <sup>14</sup>N, <sup>15</sup>N, <sup>15</sup>O, <sup>16</sup>O, <sup>17</sup>O and <sup>17</sup>F.
The most relevant feature of the present study is the employment of the FST theory to deal with energy transport by convection. This theory represents a substantial improvement over MLT, particularly in the treatment of low - viscosity fluids like the ones present in stellar interiors, for which the MLT one - eddy assumption is completely inadequate. In particular, a self - consistent model without adjustable parameters based on the FST approach has been recently formulated by CGM. These authors fitted their theoretical results for the convective flux $`F_\mathrm{c}`$ with the expression
$$F_\mathrm{c}=KTH_\mathrm{p}^1(_{\mathrm{ad}})\mathrm{\Phi },$$
(1)
where $`K=4acT^3/3\kappa \rho `$ is the radiative conductivity, $`H_\mathrm{p}`$ is the pressure scaleheight, and $``$ and $`_{\mathrm{ad}}`$ are, respectively, the true and adiabatic temperature gradients. $`\mathrm{\Phi }`$ is given by
$$\mathrm{\Phi }=F_1(S)F_2(S),$$
(2)
where
$$F_1(S)=(\frac{Ko}{1.5})^3aS^k[(1+bS)^l1]^q,$$
(3)
and
$$F_2(S)=1+\frac{cS^{0.72}}{1+dS^{0.92}}+\frac{eS^{1.2}}{1+fS^{1.5}}.$$
(4)
Here, $`Ko`$ is the Kolmogorov constant (assumed to be 1.8), $`S=162A^2(_{\mathrm{ad}})`$, and the coefficients are given by $`a=10.8654`$, $`b=4.89073\times 10^3`$, $`k=0.149888`$, $`l=0.189238`$, $`q=1.85011`$, $`c=1.08071\times 10^2`$, $`d=3.01208\times 10^3`$, $`e=3.34441\times 10^4`$, and $`f=1.25\times 10^4`$. $`A`$ is given by
$$A=\frac{c_\mathrm{p}\rho ^2\kappa z^2}{12acT^3}(\frac{g\delta }{2H_\mathrm{p}})^{1/2}.$$
(5)
A important characteristic of the CGM model is the absence of free parameters. In particular, the mixing length is taken as $`l=z`$ where $`z`$ is the geometrical distance from the top of the convection zone. Stothers & Chin (1997) have recently shown that $`l=z`$ performs very well in different kind of stars. In order to avoid numerical difficulties, we allowed for a very small overshooting at the moment of evaluating $`l`$, that is, we considered $`l=z+\beta H_\mathrm{p}^{top}`$, where $`H_\mathrm{p}^{top}`$ is the pressure scaleheight at the top of the outer convection zone. We elected $`\beta <0.1`$, which does not affect the temperature profile of the convection zone of our evolving models.
For the sake of comparison, we also included in our calculations the ML1, ML2 and ML3 parametrizations of MLT. These versions, which are associated with different convective efficiencies, has been amply employed in white dwarf investigations (see Tassoul at al. 1990 for details).
## 3 evolutionary results
Here, we present the main results of our calculations. We computed the evolution of DA white dwarf models with masses ranging from $`M=0.5M_{\mathrm{}}`$ to $`M=1.0M_{\mathrm{}}`$ at intervals of $`0.1M_{\mathrm{}}`$ and with a metallicity of $`Z=0.001`$. We varied $`M_\mathrm{H}`$ within the range $`10^{13}M_\mathrm{H}/M2\times 10^4`$ and $`M_{\mathrm{He}}`$ in the range $`10^6M_{\mathrm{He}}/M10^2`$. We employ the FST approach given by CGM, and also the ML1, ML2 and ML3 versions of MLT. The models were evolved from the hot white dwarf stage down to $`\mathrm{log}(L/L_{\mathrm{}})=5`$. Unless stated otherwise, we shall refer in what follows to those models parametrized by $`M_\mathrm{H}=10^6M`$ and $`M_{\mathrm{He}}=10^2M`$.
We begin by examining Fig. 1 in which the surface gravity of our DA models with $`\mathrm{log}q(\mathrm{H})=10,6,4`$ and no hydrogen envelope is depicted in terms of the effective temperature ($`q`$ is the external mass fraction given by $`q=1M_r/M`$)<sup>1</sup><sup>1</sup>1For the sake of clarity, in this and the following figures, we do not depict the results corresponding to 1.0 $`M_{\mathrm{}}`$ models. The first observation we can make from this figure is that thick hydrogen envelopes appreciably modify the surface gravity of no - hydrogen models, especially in the case of low - mass configurations. The effect of adding an outer hydrogen envelope of $`M_H=10^{10}`$ M is barely noticeable. Note also that finite - temperature effects are relevant even in massive models (see also Koester & Schönberner 1986). At low effective temperatures, convective mixing between hydrogen and helium layers (see later in this section) change the surface gravity of models with thin hydrogen envelopes (see the discussion in Benvenuto & Althaus 1998 in the context of low - mass helium white dwarfs for more details about this topic).
It is worthwhile to mention that our more massive models begin to develop a crystalline core at effective temperatures close to the temperature range of the ZZ Ceti instability strip. More specifically, the onset of crystallization at the centre of 1.0, 0.9, 0.8 and 0.7 $`M_{\mathrm{}}`$ DA models with $`\mathrm{log}q=6`$ takes place at $`T_{\mathrm{eff}}=`$ 16200, 13100, 10870 and 8900 K, respectively. For details concerning the process of growth of the crystal phase of our models, see Benvenuto & Althaus (1995).
The role played by hydrogen burning in our evolving models is worthy of comment. Iben & Tutukov (1984) were the first in drawing the attention to the fact that hydrogen burning in cooling white dwarfs could be, at low luminosities, a major energy source. We find in particular that for models with $`\mathrm{log}q(\mathrm{H})>4`$, hydrogen burning may contribute non - negligibly to the surface luminosity. For more details, we depict in Fig. 2 the relative contribution of hydrogen burning as a function of effective temperature for various stellar masses having a hydrogen envelope of $`\mathrm{log}q(\mathrm{H})=4`$. Note that for intermediate stellar masses the hydrogen - burning contribution reaches a maximum near the ZZ Ceti instability strip. Another feature worthy of mention shown by Fig. 2 is that, even for more massive models, hydrogen burning never becomes the dominant source of stellar energy. We also include in Fig. 2 the neutrino luminosity of the models. For the effective temperature range shown in the figure, neutrino luminosity strongly decreases with cooling and it is negligible by the time models enter the instablitiy strip.
The importance of nuclear burning is strongly sensitive to the excat value of the hydrogen layer mass, as was recognised by Koester & Schönberner (1986) (see also D’Antona & Mazzitelli 1979). To show this, we have computed additional sequences by considering the cases with $`\mathrm{log}q(\mathrm{H})=`$ -3.824; -3.699. The results for 0.6 and 0.8 $`M_{\mathrm{}}`$ models are illustrated in Fig. 3. It is clear that a small difference in the hydrogen layer mass is responsible for the different role of hydrogen burning. In particular, for the 0.6 $`M_{\mathrm{}}`$ model with $`M_\mathrm{H}=9\times 10^5M_{\mathrm{}}`$, we find that the relative contribution of hydrogen burning at low luminosities remains always below 9 per cent, while for the same model but with $`M_\mathrm{H}=1.2\times 10^4M_{\mathrm{}}`$, the hydrogen - burning contribution rises up to $`18`$ per cent
The behaviour of the evolving outer convection zone is depicted in Figs. 4 - 7 for DA white dwarf models with $`M=`$ 0.5, 0.6, 0.8 and 1.0 $`M_{\mathrm{}}`$ and for the different theories of convection. In each figure, we plot the location of the top and the base of the outer convection zone in terms of the mass fraction $`q`$ as a function of effective temperature. The first observation we can make from these figures is that at both hot and cool extremes the convection zone profile is independent of the treatment of convection. In fact, at high effective temperatures only a negligible fraction of energy is carried by convection and, consequently, the temperature stratification remains essentially radiative, while at low temperatures convection is very efficient and most of the convection zone assumes an adiabatic stratification. At intermediate effective temperatures, however, where the ZZ Ceti stars are observed, convection treatment is decisive in fixing the structure of the outer zone of the models. Note that, in contrast to the situation encountered for DB objects (Benvenuto & Althaus 1997), the CGM predictions are roughly intermediate to those of the ML1 and ML2 convection. Another feature shown by these figures and worthy of
comment is that the thickness of the convection zone in the CGM model begins to increase at a given effective temperature much more steeply than in any of the MLT versions. A similar trend is also found in DB white dwarf models, though it is worth remarking that the global behaviour of the evolving convection zone is markedly different in both types of stars. In agreement with previous results (D’Antona & Mazzitelli 1979), the final extent reached by the base of the convection zone is smaller in the more massive models. We should mention that the small jump at the top of the convection zone at $`T_{\mathrm{eff}}9000K`$ is brought about, not surprisingly, by the change from OPAL to Alexander & Ferguson’s (1994) molecular opacities. This jump affects the $`z`$ values, thus giving rise to some irregularities in the CGM convection profile, which does not alter the conclusions of this work.
To clarify better the role played by convective efficiency in determining the thermal structure of the outer layers, we show first of all in Fig. 8 the dimensionless convective temperature gradient in terms of the mass fraction $`q`$ for the outer layers of a 0.6 $`M_{\mathrm{}}`$ DA model with $`T_{\mathrm{eff}}=10890K`$ according to the CGM model and the ML1 and ML2 versions of MLT. Note the presence of a strong peak in $``$ at $`\mathrm{log}q16.1`$ in the CGM case. This arises simply because $`z<H_\mathrm{p}`$ in the outermost layers, thus giving rise to very inefficient convection. In somewhat deeper layers, however, the CGM model turns out to be more efficient than ML1. This is a result of the fact that the larger number of eddies characterizing the CGM model, compared with MLT, begins to play a significant role. For $`\mathrm{log}q>15.6`$, both the CGM and ML2 models provide an adiabatic stratification and hence the behaviour of $``$ in the outermost layers will be critical for determining the temperaute profile in the envelope and thus the extent of the convection zone of the model. The resulting temperature profile of the model analysed in Fig. 8 is depicted in Fig. 9. It is clear that the CGM model predictions are very different from those of the ML1 and ML2 models. Indeed, the temperature profile cannot be reproduced with an MLT model with a single value of $`\alpha `$.
A final remark about the convection profiles is that at low effective temperatures they ultimately reach a maximum depth, which is determined by the location of the boundary of the degenerate core. For the 0.60 $`M_{\mathrm{}}`$ DA white dwarf model, in particular, such a depth occurs around $`\mathrm{log}q6`$. (see Tassoul et al. 1990 for a similar result). Accordingly, mixing episodes between hydrogen - rich and the underlying helium - rich zones are possible only for values of $`\mathrm{log}q(\mathrm{H})`$ lower than this limit. It is apparent from Figs. 4 - 7 that the mixing effective temperature for models with very thin hydrogen envelopes will be strongly dependent on the assumed convective efficiency and not so much on the stellar mass.
To put this assertion on a more quantitative basis, we list in Table 1 the effective temperatures at which the hydrogen surface abundance begins to decrease appreciably as a result of convective mixing, together with the final hydrogen abundances (for models with thick hydrogen layers, mixing process takes place gradually in a finite range of effective temperature; in that case we only give approximate values of the hydrogen abundance). Clearly, for models with very thin hydrogen envelopes, convective mixing drastically modifies the surface composition from a hydrogen - dominated to a helium - dominated one. This behaviour can be understood directly by examinig Fig. 10, which shows the evolving convection zones of the $`0.6M_{\mathrm{}}`$ model with $`\mathrm{log}q(\mathrm{H})=12`$ according to the CGM convection. At $`T_{\mathrm{eff}}10600K`$, the merging of the hydrogen convection zone with the helium convection zone (located just below the hydrogen/helium transition region) causes the base of the convection zone to reach very deep helium layers quickly, giving rise to an almost complete dilution of the hydrogen content. From then on, the subsequent evolution corresponds to that of a DB model (see Fig. 11 of Benvenuto & Althaus 1997).
In order to estimate approximate effective temperatures for the theoretical blue edge of DA instability strip we employ thermal time - scale arguments $`\tau _{\mathrm{th}}`$ for our evolving models. Several past studies (Cox 1980, Winget et al. 1982, Tassoul et al. 1990 and references cited therein) have shown that the behaviour of pulsational instabilities in white dwarf stars can be understood in terms of $`\tau _{\mathrm{th}}`$ of the driving regions defined by
$$\tau _{\mathrm{th}}=_0^{q_{\mathrm{bc}}}\frac{C_\mathrm{V}T}{L}M_{}𝑑q$$
(6)
In Eq. 6, $`C_\mathrm{V}`$ is the specific heat at constant volume, $`L`$ and $`M`$ are, respectively, the luminosity and the mass of the model, and the subscript “bc” corresponds
to the location of the bottom of the superficial convection zone. In particular, the blue edge of the instability strip (the onset of pulsations) corresponds approximately to the effective temperature at which $`\tau _{\mathrm{th}}`$ becomes comparable to 100 s, which for a white dwarf is of the order of the shortest observable g - mode periods. In this connection, we list in Table 2 (see also Fig. 11) the effective temperature of our theoretical blue edges of the ZZ Ceti instability strip according to the different stellar masses and convection treatments we have considered. Note, as found by other investigators (e.g. Tassoul et al. 1990 and Bradley & Winget 1994) the dependence of the blue edge temperature on the stellar mass and, more importantly, on the assumed convective efficiency. In particular, our MLT blue edges are consistent with those of Bradley & Winget (1994) obtained on the basis of detailed pulsation calculations. Clearly the CGM theory predictions are intermediate to the ML1 and ML2 results. In Fig. 11, we have also included the theoretical result obtained by Gautschy et al. (1996) for a $`0.6M_{\mathrm{}}`$ ZZ Ceti model and the observational data of Bergeron et al. (1995) as well (we picked out those ZZ Ceti stars analysed by Bergeron et al. which, for a given stellar mass, are characterized by the highest temperature).
With regard to the inclusion of the observational data in this figure, some comments seem to be appropriate at this point. As mentioned earlier, the effective temperature of the stars that define the observational blue edge is computed, as usual, by employing the emergent spectrum of a model atmosphere, which in the case of DA white dwarfs is
strongly dependent on the prescription adopted for the treatment of convection (see e.g. Bergeron et al. 1995). Thus, in order to perform a self - consistent comparison with observations, we should employ a model atmosphere computed considering the CGM theory. In view of the fact that such models are still not available, any comparison of our theoretical results with observation should be taken with caution.
In spite of the above warnings, it is remarkable that both the CGM theory and Gautschy et al.’s (1996) predictions tend to point towards a cooler blue edge than observations. In particular, on the basis of very detailed hydrodynamical simulations of convection in a $`0.6M_{\mathrm{}}`$ ZZ Ceti model, Gautschy et al. derived an effective temperature between 11400 K and 11800 K for the blue edge. As far as the blue edge dependence on the stellar mass is concerned, the MLT and CGM theory predictions are rather similar, spanning $`1000K`$ over the range of stellar masses we considered. Note that dependence of the blue edge on the stellar mass predicted by both theories of convection is rather similar to that shown by observations.
## 4 Summary
In this paper we compute the structure and evolution of carbon - oxygen white dwarf models with hydrogen envelopes (DA type). We consider stellar masses ranging from M=0.5 to $`M=1.0M_{\mathrm{}}`$ at intervals of $`M=0.1M_{\mathrm{}}`$, and we treat the mass of the hydrogen and helium envelopes as free parameters within the range $`10^{13}M_\mathrm{H}/M10^4`$ and $`10^6M_{\mathrm{He}}/M10^2`$, respectively. The models were evolved from the intermediate effective temperature stage down to $`\mathrm{log}(L/L_{\mathrm{}})=5`$.
The calculations were made with a white dwarf evolutionary code including updated radiative and conductive opacities and neutrino emission rates, and very detailed equations of state for hydrogen and helium plasmas. We also include the effects of crystallization, convective mixing and hydrogen burning (pp chain and CNO bi - cycle). The most important feature of our study, however, is that we treat the energy transport by convection within the formalism of the so - called full - spectrum turbulence theory, which constitutes an improvement over most previous studies of DA white dwarf evolution. In particular, we employ a new model based on such theory developed by Canuto, Goldman & Mazzitelli (1996) (CGM). This new model includes the full spectrum of eddies, has no free parameter and computes the rate of energy input self - consistently. In the white dwarf domain, the CGM models has been recently shown (Althaus & Benvenuto 1997b) to provide a good fit to new observational data of pulsating DB objects. Finally, for the sake of comparison, we also consider the most common parametrizations of the mixing length theory (MLT) usually employed in this kind of studies.
In agreement with previous studies, we find that very thick hydrogen layers substantially modify the surface gravity of the models and that the importance of nuclear burning is strongly sensitive to the excat value of the hydrogen layer mass. In particular, we find that for the 0.6 $`M_{\mathrm{}}`$ model with $`M_\mathrm{H}=9\times 10^5M_{\mathrm{}}`$ the relative contribution of hydrogen burning at low luminosities remains always below 9 per cent, while for the same model but with $`M_\mathrm{H}=1.2\times 10^4M_{\mathrm{}}`$, the hydrogen burning contribution rises up to $`18`$ per cent.
One of our main interest in this work has been to study the evolution of ZZ Ceti models with the aim of comparing the CGM and MLT predictions. In this connection, we find that the temperature profile given by the CGM model is markedly different from that of the ML1 and ML2 models, and it cannot be reproduced by any MLT model with a single value of $`\alpha `$. The evolving outer convection zone also behaves differently in both theories. In particular, the thickness of the convection zone in the CGM model begins to increase at a given effective temperature much more steeply than in any of the MLT versions and remains intermediate to those with ML1 and ML2 convection.
We have also computed approximate effective temperatures for the theoretical blue edge of DA instability strip by using thermal time - scale arguments for our evolving DA models. In this context, we find that the effective temperature of the theoretical blue edges of the ZZ Ceti instability strip depends strongly on the stellar mass and more importantly on the assumed convective efficiency, in agreement with previuos studies. In particular, our MLT blue edges are consistent with those of Bradley & Winget (1994) obtained on the basis of detailed pulsation calculations. We find that the CGM theory predicts blue edges cooler (by $`1000K`$) than the observed ones. Specifically, we find an effective temperature of $`11000K`$ for the blue edge of a 0.6 $`M_{\mathrm{}}`$ DA model. It is remarkable that recent non - adiabatic pulsation calculations based on numerical simulations of convection tend also to indicate a somewhat cooler blue edge. However, we remaind the reader that the physical ingredients we employed in the present paper are not consistent with those of the stellar atmosphere calculations employed by Bergeron et al. (1995), particularly in the treatment of convection. Although a difference of 1000 K between the effective temperature of the observed and theoretical blue edges for the ZZ Ceti instability strip seems to be large, we should wait for model atmospheres also computed in the frame of the full spectrum turbulence theory. Only after such models become available we shall be able to gauge the actual importance of such a discrepancy.
Detailed tabulations of the evolution of our DA models, which are not reproduced here, are available at http://www.fcaglp.unlp.edu.ar/ $``$althaus/
## acknowledgments
We thank Professor R. Stothers for providing us with material before its publication. We also appreciate e - mail communications with H.-G. Ludwig. This work has been partially supported by the Comisión de Investigaciones Científicas de la Provincia de Buenos Aires, the Consejo Nacional de Investigaciones Científicas y Técnicas (Argentina) through the Programa de Fotometría y Estructura Galáctica (PROFOEG) and the University of La Plata.
|
no-problem/9812/astro-ph9812365.html
|
ar5iv
|
text
|
# The Radio Emission of the Seyfert Galaxy NGC 7319 1footnote 11footnote 1Based on observations made with the Very Large Array operated by the National Radio Astronomy Observatory and on observations made with the NASA/ESA Hubble Space Telescope, obtained from the data archive at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555.
## 1 INTRODUCTION
A number of investigations have been made of the relationship between the radio emission and the narrow-line regions (NLR) of Seyfert galaxies. From ground-based observations, it is found that the NLR are aligned and cospatial with the radio emission (Haniff, Wilson, & Ward 1988). There is also evidence that the kinematics of the NLR are related to the radio-emitting plasma. Whittle et al. (1988) studied \[O III\] $`\lambda 5007`$ emission-line profiles in several Seyfert galaxies which have linear (i.e. double, triple or jet-like) radio sources. They found that substructures (subpeaks and shoulders) in the line profiles are more conspicuous close to the radio components. This result suggests that the radio-emitting plasma interacts with the optical line-emitting gas.
The Hubble Space Telescope (HST) has allowed much higher resolution imaging of the NLR and several studies have focussed on the relationship between the NLR and the radio emission (Bower et al. 1994, 1995; Capetti et al. 1996; Capetti, Macchetto & Lattanzi 1997; Falcke, Wilson & Simpson 1998). These observations have shown detailed correspondences between optical emission-line and radio continuum images, and indicate that interactions between radio plasma and the ambient gas determine the morphology of the NLR, with the radio ejecta sweeping up and compressing the interstellar medium. Detailed observations of such interactions are, therefore, necessary to understand the NLR within several hundreds pc of the nucleus.
NGC 7319, a member of Stephan’s Quintet, is a type 2 Seyfert galaxy with circumnuclear, outflowing, ionized gas which aligns with the radio emission (Aoki et al. 1996). Multiple components are found in the optical emission-line profiles on the SW side of the nucleus. The velocity of the outflow ranges up to 500 km s<sup>-1</sup> and its extent is 4 kpc. NGC 7319 thus exhibits one of the largest circumnuclear outflows known in Seyfert galaxies.
In this paper, we report sub-arcsecond resolution radio imaging of NGC 7319 with the Very Large Array (VLA). These radio images are compared to an HST archival broad-band WFPC2 image and the results of ground-based optical spectroscopy by Aoki et al. (1996) in order to study in detail the relationship between the radio emission and the outflowing gas. Our VLA observations and the HST image are described in Section 2. The results are presented in Section 3, while in Section 4 we discuss the properties of the radio-emitting plasma in NGC 7319 and interpret our results. We give concluding remarks in Section 5. The heliocentric systemic velocity of 6740 km s<sup>-1</sup> (Aoki et al. 1996) gives a distance of 86 Mpc for NGC 7319 assuming a Hubble constant $`H_0`$=75 km s<sup>-1</sup> Mpc<sup>-1</sup> and the solar motion relative to the CMB radiation field (Smoot et al. 1991). Thus 1″ corresponds to 420 pc.
## 2 OBSERVATIONS AND REDUCTION
### 2.1 VLA Data
The VLA observations of NGC 7319 were made in the ‘A-configuration’ at 20, 6 and 3.6 cm on 1996 November 4. The integration times on NGC 7319 were 1.4, 1.3, and 1.2 hours at 20, 6, and 3.6 cm, respectively. Each 6 and 3.6 cm band consisted of two contiguous channels with a total bandwidth of 100 MHz centered at 4860 and 8440 MHz, respectively. Two separate 50 MHz channels centered at 1365 and 1435 MHz were used for the 20 cm observations. The data were phase calibrated by means of observations of the source 2236+284, which is 6 away from NGC 7319. The calibrator and NGC 7319 were observed sequentially with a 15 min cycle. Flux calibration was achieved by observations of 3C48.
Data reduction was done with AIPS in Socorro. The observations of NGC 7319 were flux and phase calibrated, mapped, and CLEANed to give the final images. Due to an error, we observed 3C48 at 1385 and 1465 MHz, which are different from the 20 cm bands used for NGC 7319 and 2236+284. The fluxes of 2236+284 at its observed frequencies of 1365 and 1435 MHz were therefore determined by interpolation and extrapolation of the 3C 48 observations, and used to flux calibrate the observations of NGC 7319.
### 2.2 HST Data
There is a broad-band image of NGC 7319 in the HST archive. It was obtained with WFPC2 through the filter F606W (effective wavelength/width 5843 Å/1578.7 Å). The nuclear region of NGC 7319 was imaged with the Planetary Camera, which has a pixel size of 0$`\stackrel{}{\mathrm{.}}`$0455. Bias and dark subtraction, and flat fielding were done with pipeline processing at the Space Telescope Science Institute. We removed cosmic ray events from the data using IRAF <sup>2</sup><sup>2</sup>2IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc. under cooperative agreement with the National Science Foundation.. Note that the F606W filter contains strong emission-lines, such as \[O III\] $`\lambda 5007`$, H$`\alpha `$, and \[N II\] $`\lambda 6584`$, from NGC 7319. The system throughputs at \[O III\] $`\lambda 5007`$ and H$`\alpha `$ are 60 % and 90 % of the peak, respectively, so the structures in this image are influenced by emission-lines.
## 3 RESULTS
### 3.1 Radio Images
The full resolution 3.6, 6, and 20 cm maps are presented in Figure 1, 2, and 3, respectively. Three compact components, plus diffuse emission, are visible in the higher resolution maps. The positions, flux densities, and sizes (FWHM) of the compact components were measured using the AIPS task JMFIT, which is a two-dimensional elliptical gaussian fitting program. These parameters are summarized in Table 1. The sizes are given after deconvolution from the beams.
In the highest resolution ($`0\stackrel{}{\mathrm{.}}27\times 0\stackrel{}{\mathrm{.}}26`$ FWHM) map (Fig. 1), the three compact components, labeled A, B and C, are found to be aligned in P.A. $`24\pm 1.4`$°. The separations of A and B, B and C, A and C are $`0\stackrel{}{\mathrm{.}}97`$ (410 pc), $`3\stackrel{}{\mathrm{.}}36`$ (1.4 kpc), and $`4\stackrel{}{\mathrm{.}}32`$ (1.8 kpc), respectively. Component A has a size of $`0\stackrel{}{\mathrm{.}}3\times 0\stackrel{}{\mathrm{.}}2`$ (130 pc $`\times `$ 80 pc) and is elongated in P.A.$`0\mathrm{°}`$. Component C has a size of $`0\stackrel{}{\mathrm{.}}25\times 0\stackrel{}{\mathrm{.}}2`$ (110 pc $`\times `$ 80 pc) and is elongated in P.A.$`40\mathrm{°}`$. There is a diffuse component which extends $`2\mathrm{}`$ towards the north from C. There is also diffuse emission extending to the south of A.
The structure in the 6 cm naturally weighted map ($`0\stackrel{}{\mathrm{.}}47\times 0\stackrel{}{\mathrm{.}}45`$ FWHM) (Fig. 2) is similar to that in the 3.6 cm map. The three components (A, B, and C) are clearly seen, with diffuse emission extending between A and C. The overall appearance of two outer ‘hot spots’ (A and C) with diffuse emission extending back towards the nucleus is reminiscent of FRII-class radio galaxies, though with much lower radio power and smaller spatial extent in NGC 7319.
The A and B components which are seen in the 3.6 cm and 6 cm images are merged together in the 20 cm uniformly weighted image ($`1\stackrel{}{\mathrm{.}}31\times 1\stackrel{}{\mathrm{.}}29`$ FWHM) (Fig. 3). The position of this merged peak agrees with that given by van der Hulst & Rots (1981) to within 0$`\stackrel{}{\mathrm{.}}`$2. The resolution of our 20 cm image is higher than that of van der Hulst & Rots (1981), and the jet-like feature seen in their Figure 3 corresponds to our component C. Our 20 cm total flux density of $`28.5\pm 0.5`$ mJy agrees with that found by van der Hulst & Rots (1981).
The spectral indices ($`\alpha `$: $`f_\nu \nu ^\alpha `$) of the three compact components were calculated from the 3.6 and 6 cm flux densities and are given in Table 1. The spectral index between 20 and 6 cm is also derived for component C. The spectral indices of all three components are similar ($`\alpha 1.3`$). Our derived spectra for components with diffuse emission may be systematically too steep as the shorter wavelength maps may miss extended flux. However, the steep spectra are consistent with the spectral index of the total radio emission ($`\alpha =1.2`$, van der Hulst & Rots (1981); $`\alpha =1.1\pm 0.3`$, Kaftan-Kassim, M. A., Sulentic, J. W., & Sistal, G. (1975)).
### 3.2 Optical Image
The HST F606W image of NGC 7319 is shown in Figure 4. Prominent spiral-like structure is seen around the nucleus. The spiral features extend 1$`\stackrel{}{\mathrm{.}}`$4 (590 pc) to the north and 1″(420 pc) to the south of the nucleus. The northern arm is associated with a dust lane. There is also a V-shaped feature 3$`\stackrel{}{\mathrm{.}}`$4 south and 1$`\stackrel{}{\mathrm{.}}`$4 west of the nucleus. To enhance the contrast of the fine-scale structure, we smoothed the image with a gaussian function of 1$`\stackrel{}{\mathrm{.}}`$5 FWHM and subtracted it from the original (unsmoothed) one. The difference image is shown in Figure 5. The V-shaped feature, as well as the spiral-like structure mentioned above, are clearly seen.
Both the spiral-like structure and the V-shaped feature may, at least in part, be emission-line regions, since strong emission-lines (\[O III\] $`\lambda 5007`$, H$`\alpha `$, and \[N II\] $`\lambda 6584`$) are present in the filter passband. We have estimated the contribution of emission-lines to the spiral-like structure and the V-shaped feature in the F606W image using ground-based spectra of the corresponding region. Nuclear (aperture size 4$`\stackrel{}{\mathrm{.}}`$4 in NS $`\times `$ 1$`\stackrel{}{\mathrm{.}}`$8 in EW) and off-nuclear (aperture size 1$`\stackrel{}{\mathrm{.}}`$8 in NS $`\times `$ 4$`\stackrel{}{\mathrm{.}}`$4 in EW) spectra were extracted. The two extraction windows include the spiral-like structure and the V-shaped feature, respectively. These spectra were multiplied by the transmissions of HST, WFPC2 and F606W using the synthetic photometry package SYNPHOT in STSDAS. The result is that emission lines contribute 17 % and 3 % of the total flux of the nuclear region and off-nuclear region, respectively. We also derived the fluxes in the HST original (Fig. 4) and difference (Fig. 5) images within the same apertures as used to extract the ground-based spectra. The spiral-like structure (as measured from the difference image) comprises 14% of the total flux (from the original image) through the nuclear aperture. The corresponding number for the V-shaped feature through the off-nuclear aperture is 3%. These fractions are similar to the contributions of emission-lines in each aperture, so it is possible that both structures are emission-line features with the smoothly distributed background being stellar light.
## 4 DISCUSSION
### 4.1 The Physical Properties of Radio Plasma
Double or triple radio structures similar to NGC 7319 are found in many Seyfert galaxies (Ulvestad & Wilson 1984a, 1984b, 1989; Kukula et al. 1995). The steep spectra of the three compact components in NGC 7319 are consistent with synchrotron radiation. Adopting the minimum energy condition for cosmic rays plus magnetic field, we have estimated the magnetic field strength for components A and C. We have assumed power-law spectra ($`\alpha =1.2`$ for A, $`\alpha =1.3`$ for C) between 10 MHz and 100 GHz and a value of 100 for the ratio of the total cosmic ray energy to the relativistic electron energy.
The results for components A and C are $`6\times 10^4`$ G and $`5\times 10^4`$ G, respectively. The relativistic pressures derived for A and C are $`3\times 10^8`$, and $`2\times 10^8`$ dyn cm<sup>-2</sup>, respectively. The thermal gas pressure of the emission-line region of NGC 7319, for which the electron density is $`n_e`$=300-600 cm<sup>-3</sup> (Aoki et al. 1996), is $`48\times 10^{10}`$ dyn cm<sup>-2</sup>, assuming an electron temperature of T<sub>e</sub> = 10,000 K. The relativistic pressure is thus at least one or two orders of magnitude higher than the thermal gas pressure.
### 4.2 Comparison of Radio Data with HST and Ground-based Spectroscopic Data
The absolute astrometric uncertainties for HST data are of the order of 1″, so the HST image cannot be registered with the VLA images by relying on the internal HST astrometry. We therefore attempted to obtain more accurate astrometric coordinates for the HST image using the following procedure. The HST image was rotated to the cardinal orientation by means of the keyword ‘ORIENTAT’ in the data header. Next, the image was smoothed with a gaussian function of 1$`\stackrel{}{\mathrm{.}}`$5 FWHM to simulate ground-based seeing. Various reasonable sizes of the gaussian were tried, but in all cases the positions of the peak agreed to 0$`\stackrel{}{\mathrm{.}}`$1. The peaks in the smoothed and unsmoothed images agree with each other to 0$`\stackrel{}{\mathrm{.}}`$08 . We then assigned the absolute position of the nucleus determined by ground-based astrometry (Clements 1983) to the peak of the smoothed image.
However, this registration caused a systematic separation ($``$ 0$`\stackrel{}{\mathrm{.}}`$7) between features in the HST and radio images. The shift of 0$`\stackrel{}{\mathrm{.}}`$7 is larger than the internal errors of Clements’ (1983) and the VLA astrometry, but there can be uncertainties in Clements’ position due to structure in the nuclear region and the fact that the spectral sensitivity of Clements’ plates is different to the HST image. There are also systematic differences between the optical and radio astrometric frames. We finally decided to shift the optical peak onto the position of radio component B (Figs 1 and 2), while recognizing the somewhat arbitrary nature of this alignment and the consequent uncertainty in the interpretation which follows. The resulting shift of the original HST coordinates is 1$`\stackrel{}{\mathrm{.}}`$26. Using this registration, contours of the 3.6 cm radio image are overlaid on the HST image in Figure 6. Radio component A coincides with the northern arm of the spiral-like structure and component C is just inside of the V-shaped feature. Even with the alignment based on Clements’ nuclear position, radio component C is still to the NE of the V-shaped feature, so we regard this displacement as secure.
HST observations have revealed various associations between radio continuum and optical line-emission in Seyfert galaxies. In some Seyfert galaxies, the emission-line features are shell-like or arc-like and surround the radio emission; Mrk 573 is an excellent example (Capetti et al. 1996; Falcke, Wilson & Simpson 1998). This type of structure is interpreted as evidence that the ionized gas is compressed by the shocks created by the outward motion of the radio plasma. Such compression increases the gas density and enhances the surface brightness in line emission. The relationship between the V-shaped feature and radio component C is similar to that between the emission-line arcs and the radio hotspots in Mrk 573.
The emission-lines show a pronounced blueshift and the line profiles are flat-topped or double-peaked near the V-shaped feature (Fig. 7). There is also a blueward sloping asymmetry to the \[O III\] line profiles at the nucleus and northeast of the nucleus (Fig. 7).
We interpret these observational results as follows. The V-shaped feature represents compressed gas behind a bow shock driven into the photoionized gas by the outwardly moving radio component C. This compressed gas is given a motion outward by component C, and thus shows a large blueshift although there is not any independent evidence that the SW is the near side of the outflow yet. It is interesting that to the NE of component C, \[N II\] 6584 $`>`$ H$`\alpha `$, while to the SW of component C, H$`\alpha `$ $`>`$ \[N II\] 6584 (Fig. 7). There thus appears to be a change of excitation possibly associated with the bow shock. H$`\alpha `$/\[O III\], however, does not significantly change across the position of component C when we reanalyzed the spectra of Aoki et al. (1996). This may be due to that each emission line was observed on different nights under different seeing conditions. More accurate observations are necessary to study change of excitation across component C. The blue wing in \[O III\] at the nucleus and on the northeastern side of the nucleus may be related to the spiral-like structure in the HST image and/or components A and B in the radio images. This spiral-like structure could be a curved jet similar to that seen in ESO 428-G14 (Falcke et al. 1996) and NGC 4258 (Cecil, Wilson, & Tully 1992).
## 5 CONCLUSIONS
We have observed the Seyfert galaxy NGC 7319 with the VLA and found a triple radio source straddling the nucleus. We have also found a spiral-like structure and a V-shaped feature in an HST archival WFPC2 image of this galaxy. These optical features are closely related to the radio emissions. Combining these results with ground-based optical spectroscopy, we interpret the V-shaped feature as gas compressed by a bow shock driven into the ambient medium by the outwardly moving radio plasmoid. To confirm this interpretation, high resolution emission-line imaging and spectroscopy are needed.
We thank the staff of the Array Operation Center in Socorro, especially G. Taylor, for their kind help during observations and data reduction. We also thank Masaru Watanabe and Hiroshi Ohtani for their valuable suggestions and encouragements. Part of the data analysis was done at the Astronomical Data Analysis Center, National Astronomical Observatory of Japan, which is an inter-university research institute operated by Ministry of Education, Science, Culture and Sports. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. This research was supported in part by NASA through grant NAG 81027.
|
no-problem/9812/cond-mat9812167.html
|
ar5iv
|
text
|
# REFERENCES
Electric field and potential around localized scatterers in thin metal films studied by scanning tunneling potentiometry
Geetha Ramaswamy<sup>1</sup><sup>1</sup>1email: geetha@physics.iisc.ernet.in
Department of Physics, Indian Institute of Science, Bangalore 560 012, India
A. K. Raychaudhuri<sup>2</sup><sup>2</sup>2email: arup@csnpl.ren.nic.in
National Physical Laboratory, New Delhi 110 012, India.
## Abstract
Direct observation of electric potential and field variation near local scatterers like grain boundaries, triple points and voids in thin platinum films studied by scanning tunneling potentiometry is presented. The field is highest at a void, followed by a triple point and a grain boundary. The local field near a void can even be four orders of magnitude higher than the macroscopic field. This indicates that the void is the most likely place for an electromigration induced failure. The field build up near a scatterer strongly depends on the grain connectivity which is quantified by the average grain boundary reflection coefficient, estimated from the resistivity.
Electromigraion (EM) failure is an important reliability issue for metallic interconnects in integrated circuits. Increased device density has caused a reduction in interconnect size, resulting in very high current densities which in turn increase the probability of EM failure. EM failures occur either by voids developing across the interconnects producing an open circuit or by extrusions which cause a short circuit with neighbouring lines. Due to its technological importance, the mechanism of field induced void formation has been studied by various techniques including microscopy based techniques . Bamboo lines have a nanostructure consisting of a number of grain boundaries (GB) spanning the length of the interconnect. Near bamboo lines contain triple points (TP) and voids (V) along with a network of GBs. The electric field across such a thin film is not uniform at the nanometric scale and inhomogeneities in the field and current density occur near these localized scatterers . The technique of scanning tunneling potentiometry (STP) invented by Muralt and Pohl , which is built around a scanning tunneling microscope (STM), offers a novel method of studying the potential variations and hence the electric field at these scattering centers with a nanometric resolution, by providing a simultaneous map of the topograpy and the potential distribution in a current carrying film. Thus it provides information regarding the nanostructure and the field distribution, both of which are of vital importance in EM studies.
In this letter we present our results on the study of electronic transport in polycrystalline thin platinum films by STM/STP. By simultaneously mapping out the topography and the potential distribution in the current carrying thin film, we investigate the variation in the potential and the build up of an electric field near various types of localized scatterers, in the scale of the spatial extent of the scatterers. We show that the spatial distribution and the magnitude of the local field depends on the type of the scatterer as well as the connectivity of the grains. Specifically, our experiments show that of the different types of scatterers present in a polycrystalline thin film, the maximum build up of an electric field occurs near a void, which then is the most likely spot for an EM induced failure. Further our study also shows that for a given type of scatterer, the magnitude of the local field depends on the grain connectivity and it is low for well connected grains.
Previous STP studies on metallic films have focused on charging effects in granular gold films and on local variations in potential in Au-Pd , Au and Bi thin films . However, local variations in electric field around various kinds of extended defects occurring in a thin film has, so far, not been studied. Ideally, the STP experiments should be performed on conventional interconnect materials like Al and Cu. However, the formation of surface oxides in these materials on exposure to ambient atmosphere makes the interpretation of STP results difficult. In an attempt to understand the microscopic origin of the field inhomogeneities and the relative importance of the different kinds of defects in inducing an EM failure, we decided to study the underlying physical process in platinum, which is resistant to oxidation.
Platinum films (thickness$``$10 nm, deposition rate $``$2.8 nm/min) were deposited using shadow masks on cleaned glass substrates by e-beam in a turbo pumped chamber at a base pressure of 5 X 10<sup>-8</sup> torr, using material of 6N purity. Simultaneous STM/STP images were obtained by the double feedback technique, imaging the films under ambient conditions in an STM built in-house, using Pt-Rh(13%) tips. The details of the STM and the double feedback operation have been described elsewhere . The topographic images were obtained with a tunnel current of $`I_{pp}`$ = 0.8 nA and an AC bias of $`V_{pp}`$ = 0.05 V at a frequency of 2 kHz. The potentiometic images were obtained with a macroscopic field of 5.2 V/cm and a current density j $``$ $`10^5`$ A/cm<sup>2</sup>. To avoid any artifacts that might arise if the current direction and the scan direction are the same, j was kept at an angle ($`70^o`$) to the fast scan direction (X axis).
Figure 1 (a) and (b) show the simultaneous STM and STP images respectively, in a 62 nm X 44 nm region of the film. The topographic image shows a number of grains, GB, TP and V. Representative ones are marked in the figure. The nanostructure consists mostly of circular grains having an average grain diameter $`<D>`$ $``$ 14.7 nm and an r.m.s surface roughness $``$ 1.6 nm, as obtained from several topographic scans performed in different regions of the film. Fig. 1(c) is a line profile across the topographic image (marked across the image) showing the z-height corrugations of the GBs and TPs. From the potentiometric image and the line profile shown in Fig. 1(d), it is immediately apparent that the potential does not drop uniformly across the film surface and that the local potential distribution is severely affected in the vicinity of the scatterers. The line profile in Fig. 1(d) is obtained across the potential image in the same region as the profile in Fig. 1(c). From the two line profiles we discern a one to one correspondence between the scattering centers in the topographic image and the voltage variations in the potential image.
We analysed the STP images obtained from various regions of the film to ascertain the magnitude of the potential inhomogeneities caused by different types of extended defects and their relative importance in contributing to a possible EM failure. Our analysis shows that the typical potential variations $`\mathrm{\Delta }`$$`\varphi `$ at the GB are $``$ 1 - 3 mV occurring over a range of 0.5 - 3.5 nm. At the TP and voids $`\mathrm{\Delta }`$$`\varphi `$ is $``$ 1 - 4 mV and 3 - 9 mV respectively, occurring over a distance of 1 - 3 nm. It is to be noted that the potential image is flat and featureless in the absence of a DC current through the sample, except for a few 0.3 mV quantization noise wiggles of the A/D converter, as seen from Fig. 1(e). This proves that the potential variations are the result of an actual build up of a field near the scatterer and are not due to tip related artifacts.
In order to further understand the nature of the grain connectivity and the extent and magnitude of the scattering at different scattering centers such as GB, TP and V, we investigated the electric field values in their vicinity. The local transport field in the surface plane is calculated from the gradient of the local potential, $`\varphi (x,y)`$, as: $`𝐄_{}(x,y)=\varphi (x,y)`$. $`\varphi (x,y)`$ is what is measured in an STP scan. We computed the electric fields along the X, Y directions numerically from the 128 X 128 potentiometric data array and used quiver plots of the field data to visually show the distribution of field lines near the scatterer. The results of the field calculation are shown in Fig. 2. The length of the arrow indicates the magnitude and the head points along the direction of the field.
Figure. 2 shows the electric field and a line scan across the potential image around a GB, a TP and a V. These regions have been labeled and marked by a rectangle in Fig. 1(a), with a short line across the rectangle depicting the profile. A comparison of the figures brings out the distinct nature of the electric field distribution around the three kinds of scatterers. The field lines in a GB (Fig. 2(a)) are concentrated along the periphery of the grain and very low in its interior. At a TP the field lines are stronger (Fig. 2(c)). The field radiates outward from the GB and concentrates at the TP. At a V, the build up of the field is most prominent (Fig. 2(e)). This is also brought out by the STP line profiles (Figs. 2(b),(d),(f)), which show that the potential inhomogeneity is a maximum at a V and minimum at a GB. Typical field values obtained from above are $``$ 1.6 X $`10^4`$ V/cm, 3.2 X $`10^4`$ V/cm and 7.2 X $`10^4`$ V/cm respectively for a GB, TP and V.
Our study also indicates that the field build up around a scatterer will depend on the grain connectivity for a given kind of scatterer. The grain connectivity of the film can be characterized by an average GB reflection coefficient ¡R<sub>g</sub>¿, which is obtained from an analysis of the temperature dependence of resistivity of the film . From the analysis of the resistivity data ($`\rho _{4.2K}`$$``$47$`\mu \mathrm{\Omega }`$cm and $`\rho _{300K}`$/$`\rho _{4.2K}`$=1.05), we obtain an ¡R<sub>g</sub>¿$``$0.9 for this film. In order to better understand the dependence of ¡$`R_g`$¿ on the grain connectivity, which is reflected in the field build up, we repeated the STM/STP experiments on a platinum film which was grown on a rougher surface but with similar thickness and average nanostructural parameters ($`<D>`$$``$12.35 nm and an r.m.s roughness of 1.7nm). The film had a $`\rho _{4.2K}`$$``$160$`\mu \mathrm{\Omega }`$cm and $`\rho _{300K}`$/$`\rho _{4.2K}`$=1.22, resulting in a ¡$`R_g`$¿$``$0.97. This shows that the grains are not as well connected as in the previous film. In this film, although the STP and the field patterns are qualitatively similar to the previous film, for the same macroscopic field, the magnitude of the field at the defects (GB, TP and V) are an order of magnitude larger. This clearly indicates that the field build up, for a particular kind of defect, strongly depends on the grain connectivity.
This study has brought out a number of important observations. The quiver plots conclusively prove that the field across the film is not uniform and it is severely altered at the scatterers. Further, the field tends to concentrate at the scatterer and it is very low in the interior of the grain. The magnitude of the field, for a given scatterer, depends on the grain connectivity. Among the different kinds of scatterers, the field is highest at a void. The mapping out of the potential and field at the different scattering sites is an important result as this has significant implications on our understanding of EM phenomena in an interconnect. It identifies regions of excessive field build up, which are the likely ”weak spots” where an EM failure can originate. Previous studies on thin films have shown that during EM, vacancies migrate predominantly along GBs and accumulate at GBTPs due to vacancy flux divergence . These vacancies coalesce into a void, which then becomes the most likely spot for an EM failure to occur, when it reaches a critical size. Our STP study supports this observation. If the grain connectivity is poor, the local field at a triple point may even be three orders of magnitude higher than the average macroscopic value. Such a high field can result in a void formation, which has a higher field build up and hence has the greatest probability for EM induced film failure.
In conclusion, we have established a procedure for identifying the hot spots for a field induced EM failure. Our study shows that wide line metal interconnects should have well connected grains with a minimum number of triple points and voids for reliable long term performance.
We wish to thank Prof. N. Chandrasekhar for allowing us the use of his e-beam facility and K. Das Gupta for depositing the films.
FIGURE CAPTIONS
(1) FIG.1. Simultaneous STM/STP scans in a 62 nm X 44 nm area. (a) topographic scan (b) potentiometric scan. The arrow indicates the direction of the macroscopic j. The profiles across (a) and (b) are shown in (c) and (d) respectively. (e) STP profile in the absence of a field.
(2) FIG. 2. The field distributions and line profiles at various types of scatterers, calculated from the STP image, obtained in the regions marked by rectangles and short lines respectively in Fig. 1(a). (a) and (b) Field and profile at a GB, (c) and (d) at a TP and (e) and (f) at a V. It is seen that the field is maximum at a void, followed by a TP and a GB and it is very low in the interior of the grain.
|
no-problem/9812/cond-mat9812143.html
|
ar5iv
|
text
|
# Nonlinear dynamic susceptibilities of interacting and noninteracting nanoparticle systems
## I Introduction
The magnetic response of ensembles of single-domain magnetic particles has been in focus since the pioneering work of Néel and Brown. In the case of noninteracting particle systems, much theoretical work have been devoted to the linear response. Experimental results from magnetic relaxation and ac susceptibility measurements reasonably well support the now existing models.
The linear response of interacting magnetic particle systems is more involved. It has recently been shown that dipole-dipole interactions introduce collective behavior, as evidenced by the appearance of magnetic aging and a significantly broadened magnetic relaxation. Moreover, the dynamics of a magnetic particle system of monodispersive nature is indicative of critical slowing down at a finite temperature, implying the existence of a low temperature spin-glass-like phase.
Much less work have focused on the nonlinear response of magnetic particle systems. Bitoh and co-workers studied experimentally the third harmonic of the ac magnetization to obtain information related to the cubic dynamic susceptibility. However, the ac field used in this study was comparably large, which makes it possible that the results for the third harmonic of the magnetization are contaminated with higher order susceptibility terms. In order to obtain theoretical expressions for the dynamic nonlinear susceptibility for systems without magnetic anisotropy, one can directly translate the expressions obtained for the dielectric relaxation. The inclusion of magnetic anisotropy is more involved and it is only recently that this has been done for systems with the simplest uniaxial anisotropy. Of special interest is the study of the cubic dynamic susceptibility by Raikher and Stepanov, who used the Fokker-Planck equation to obtain numerically exact results on the linear and cubic dynamic susceptibilities. They also suggested approximate analytical expressions for these quantities.
In the case of interacting magnetic particle systems, experimental work related to the nonlinear magnetic response are even more scarce. However, in a recent study it was shown that the equilibrium nonlinear response of an interacting magnetic particle system of monodispersive nature indicates that the cubic equilibrium susceptibility diverges at a finite temperature, thus providing further evidence for a low temperature spin-glass-like phase. To the best of our knowledge, no experimental work has been reported for the dynamic higher order susceptibility terms of interacting particle systems.
In this paper, we report a study of the linear and cubic dynamic susceptibilities of a magnetic particle system consisting of nanosized maghemite particles. The effects of interparticle interactions are investigated by studying three samples with different volume concentrations of magnetic particles. The results for the most dilute sample are compared with the expressions for the susceptibility proposed by Raikher and Stepanov. The dynamic response of the most concentrated sample is compared with the dynamic response of a Ag(11at.%Mn) spin glass.
## II Experimental
The experiments were performed on three samples consisting of nanosized maghemite ($`\gamma `$-Fe<sub>2</sub>O<sub>3</sub>) particles, with a typical particle diameter of 6 nm and almost spherical particle shape (observed in TEM analysis). The particles were suspended in a hydrocarbon oil and coated with a surfactant layer preventing the particles from agglomerating. Since the measurements were performed at low temperatures, the oil was frozen and the particles fixed randomly in the sample. Three samples with different volume concentration of particles, $``$0.3%, 3%, and 17%, were used in order to study the effects of magnetic dipole-dipole interactions. The lowest concentration should be sufficiently low to make the dipole-dipole interactions between the particles negligible. Experiments were also performed on a Ag(11at.%)Mn spin glass sample exhibiting long range spin-spin interactions of RKKY type. . The spin glass sample was prepared by melting pure Ag and Mn together at 1000 C in an evacuated and sealed silica tube. After annealing the sample at 850 C for 72h it was quenched in water to room temperature.
Two different experimental equipments have been used: (i) A commercial ac-susceptometer was used to measure the 1st and 3rd harmonics of the magnetization for different ac-field amplitudes in the range $`100`$$`2000`$ A/m. (ii) A non-commercial low-field superconducting quantum interference device (SQUID) magnetometer was used to perform sensitive studies of the frequency dependence of the linear and cubic dynamic susceptibilities. Frequencies in range of 2 Hz to 200 Hz were used.
The magnetization, $`M`$, can be expanded with respect to an applied field $`H`$ as
$$M=\chi H+\chi _3H^3+\chi _5H^5+\mathrm{},$$
(1)
where $`\chi `$ is the linear susceptibility and $`\chi _3`$ the cubic susceptibility. The dynamic susceptibility can be probed by applying an ac field, $`H=h_0\mathrm{cos}(\omega t)`$. The linear susceptibility is then obtained from the magnetization measured at the fundamental frequency as $`\chi (\omega )=M_\omega /h_0`$ and the cubic dynamic susceptibility is obtained from the third harmonic of the magnetization as $`\frac{1}{4}\chi _3(\omega )=M_{3\omega }/h_0^3`$. These equations are only valid if the applied ac field is sufficiently small, so that contributions from higher order susceptibility terms in the expansion of $`M_\omega `$ and $`M_{3\omega }`$ are negligible. For the measurements of the linear and cubic dynamic susceptibilities we used different ac-fields to ascertain that no mixing with higher order susceptibility terms occured.
## III Theoretical background
### A Equilibrium susceptibilities
The potential energy of a single-domain particle in an external magnetic field is the sum of the Zeeman energy and the anisotropy energy. For a particle with uniaxial anisotropy the total magnetic potential is given by
$$U=D(\widehat{e}\widehat{n})^2\mu _0M_sV(\widehat{e}\stackrel{}{H}),$$
(2)
where $`D`$ is the anisotropy energy barrier and $`\stackrel{}{H}`$ is the external filed, while $`\widehat{n}`$ and $`\widehat{e}`$ are the unit vectors along the anisotropy axis and magnetic moment, respectively. In the case of volume anisotropy $`D=KV`$, where $`V`$ is the particle volume and $`K`$ is the volume anisotropy constant, and for surface anisotropy $`D=K_SS`$, where $`S`$ is the particle surface and $`K_S`$ is the surface anisotropy constant.
The linear and cubic equilibrium susceptibilities can be derived from the partition function associated with the magnetic potential. For a mono-dispersive system with randomly distributed anisotropy axes, the linear and cubic equilibrium susceptibilities are given by
$$\chi ^{\mathrm{eq}}(T)=\frac{\mu _0M_s^2V}{3k_BT},\chi _3^{\mathrm{eq}}(T)=\frac{\mu _0^3M_s^4V^3}{(k_BT)^3}\frac{(1+2S_2^2)}{45},$$
(3)
respectively, where
$$S_2(\sigma )=Z_0^1(\sigma )_1^1P_2(z)\mathrm{exp}(\sigma z^2)𝑑z.$$
(4)
Here $`P_2(z)=\frac{1}{2}(3z^21)`$ is the second Legendre polynomial, $`z=(\widehat{e}\widehat{n})`$, $`\sigma =D/k_BT`$, and $`Z_0=_1^1\mathrm{exp}(\sigma z^2)𝑑z`$ is the partition function in the absence of an applied magnetic field.
### B Dynamic susceptibilities
Recently, Raikher and Stepanov studied theoretically the linear and cubic dynamic susceptibilities of noninteracting ensembles of single-domain particles with uniaxial anisotropy, solving numerically the Fokker-Planck equation in the overdamped case. It is difficult to derive analytical expressions for these quantities, but for a particle system with randomly distributed anisotropy axes, they suggested the following simple formulae for the linear dynamic susceptibility,
$$\chi (T,\omega )=\chi ^{\mathrm{eq}}\left[\frac{1+2S_2(\sigma )}{3(1+i\omega \tau )}+\frac{2}{3}(1S_2(\sigma ))\right].$$
(5)
This expression has been shown to be a good approximation to the exact linear dynamic susceptibility for frequencies below the ferromagnetic resonance frequency regime (see, for instance, Ref.). In the above expression $`\tau `$ is the relaxation time, for which various analytical expressions have been suggested (see, for instance, Ref. ). In the moderate to high-energy barrier case, the relaxation time is approximately given by an Arrhenius law as $`\tau =\tau _0\mathrm{exp}(D/k_BT)`$, where $`\tau _0`$ is a constant. Raikher and Stepanov also proposed an expression for the cubic dynamic susceptibility
$$\chi _3(T,\omega )=\chi _3^{\mathrm{eq}}\frac{(1i\omega \tau )}{(1+i\omega \tau )(1+3i\omega \tau )},$$
(6)
which they assert to be a good approximation of the numerically exact result in the low frequency regime. From the above equations we can see that for $`\omega \tau 1`$, e.g., for high temperatures, the dynamic susceptibilities reduce to the equilibrium susceptibilities Eq. (3).
## IV Results and Discussion
The linear susceptibilities for the three maghemite samples measured at the frequency $`\omega /2\pi =125`$ Hz and with an ac-field amplitude of 100 A/m are shown in Fig. 1. Dipole-dipole interactions shift the susceptibility peaks to higher temperatures, lower the magnitude of the peaks, but produce a higher equilibrium susceptibility. Figure 2 shows the corresponding cubic susceptibilities measured with an ac-field amplitude of 400 A/m. Similar to the case of the linear susceptibility, dipole-dipole interactions shift the susceptibility peaks to higher temperatures and reduce their magnitudes. The sample with the highest concentration of particles has a second positive peak at low temperatures in both the real and imaginary components of the cubic susceptibility.
The results for the most dilute sample and the most concentrated sample will be discussed separately below. For the most dilute sample, the linear and cubic susceptibilities are compared to the theoretical expressions for noninteracting particle systems discussed in Sec III. The sample with the highest particle concentration has been reported to show collective behavior at low temperatures. Its cubic dynamic susceptibility is therefore compared with $`\chi _3`$ obtained for the spin glass sample.
### A Noninteracting particles
To compare the measured linear and cubic susceptibilities with theoretical expressions, the polydispersivity of the particle system needs to be taken into account. The conventional approach is to choose a trial volume distribution and determine its parameters by fitting theoretical curves to experimental data. The log-normal distribution and the gamma distribution have been shown to work well for a variety of nanoparticle systems. The log-normal distribution is given by
$$g(V)=\frac{1}{\sqrt{2\pi }Vs}\mathrm{exp}\left(\frac{\mathrm{ln}(V/V_0)}{2s^2}\right),$$
(7)
where $`V_0`$ is the median particle volume and $`s`$ is the logarithmic standard deviation. The gamma distribution is given by
$$g(V)=\frac{1}{\mathrm{\Gamma }(1+\beta )V_0}\left(\frac{V}{V_0}\right)^\beta \mathrm{exp}(V/V_0),$$
(8)
where $`\mathrm{\Gamma }()`$ is the gamma function and $`V_0`$ and $`\beta `$ are parameters related to the mean particle volume and standard deviation.
The susceptibility of a polydispersive system is calculated as $`\chi _{\mathrm{poly}}=_V\chi _{\mathrm{mono}}g_V𝑑V`$, where $`g_V`$ is the volume weighted volume distribution and $`\chi _{\mathrm{mono}}`$ is either the linear or the cubic susceptibility given by Eq. (5) and Eq. (6), respectively. In the theoretical model the anisotropy is assumed to have uniaxial symmetry. In other work on $`\gamma `$-Fe<sub>2</sub>O<sub>3</sub> nanoparticles, it has been shown that the dominating magnetic anisotropy is uniaxial and of surface type. In these studies particles exhibiting almost spherical particle shape and with diameters in the range 4.8 - 10 nm were investigated, which correspond well to the particles investigated in the present study.
The temperature dependence of the saturation magnetization at low temperatures has earlier been modeled as $`M_s(T)=M_s(0)(11.8\times 10^5T^{3/2})`$, with $`M_s(0)=4.2\times 10^5`$ A/m. The Nelder-Mead simplex method was used to perform nonlinear fitting to the measured linear and cubic susceptibilities for frequencies in the 2 Hz to 200 Hz range. The fitting parameters were: the anisotropy constant, the preexponential factor in the expression for the relaxation time ($`\tau _0`$), and the two parameters from the volume distribution. The parameters obtained from simultaneous fits to $`\chi `$ and $`\chi _3`$ are presented in Table I. (We do not present the combination surface anisotropy and volume distribution modeled by a gamma distribution, since it gave unphysical fitting parameters.) The anisotropy constant in the case of surface anisotropy $`K_S=2.3\times 10^5`$ Jm<sup>-2</sup> is similar to the value $`K_S=2.7\times 10^5`$ Jm<sup>-2</sup> obtained in Ref. . The different volume distributions give a typical particle diameter of $`6\pm 1`$ nm in good agreement with the typical particle size of 6 nm observed in TEM studies.. The results of the simultaneous fits to $`\chi `$ and $`\chi _3`$ are shown in Fig. 3 for the linear susceptibility and in Fig. 4 for the cubic susceptibility. We can see that the best fit is obtained for a volume distribution modeled by a gamma distribution. However, the differences between the three cases are not sufficiently large to make any conclusions about the origin or the uniaxial anisotropy or the functional form of the volume distribution in this sample. We have also performed the fitting with the expression for the relaxation time proposed by Cregg et al. in Ref. , but the change of the quality of the fit was negligible.
The discrepancies between the calculated and the measured dynamic cubic susceptibilities may have several possible origins. One obvious origin is that the models of the polydispersivity of the system used here are too simple and that the real situation is more involved. It may be that the symmetry of the magnetic anisotropy in the experimental system is different from uniaxial. Such a difference need not originate from the magnetocrystalline anisotropy but can also be due to the geometrical shape of the individual particles. The lack of theoretical work where a magnetic anisotropy different from the uniaxial case is assumed makes it difficult to predict what effect the symmetry of the anisotropy will have on the nonlinear dynamic susceptibilities. Another possible cause for the observed differences is linked to the details of the Fokker-Planck equation used in the theoretical work of Raikher and Stepanov, who studied the overdamped case (the damping constant, in the dynamical Landau Lifshitz equation from which the Fokker-Planck equation is derived, is assumed to fulfill $`\alpha 1`$). However, a more realistic value of the damping constant for magnetic nano-particles is $`\alpha 1`$. The form of the dynamic linear susceptibility curves is not very sensitive to the value of the damping constant (it essentially changes the value of $`\tau _0`$), but the same does not necessarily hold for the nonlinear response and a relatively more pronounced dependence of $`\chi _3(\omega )`$ on the damping parameter cannot, in principle, be ruled out. Taking all these considerations into account, we can conclude that measurements of both the linear and cubic susceptibilities potentially allow for extracting more detailed information concerning the intrinsic properties of real particle samples.
### B Interacting particles
Figures 1 and 2 clearly show that dipole-dipole interactions have a strong influence on the magnetic relaxation of the particle system. Some recent experimental studies show that strongly interacting particle systems approach a spin glass phase at low temperatures. For example, spin glass characteristics such as: aging effects, critical slowing down, and a divergence of the cubic equilibrium susceptibility have been reported in such systems.
In a model often used to analyze the dynamics of interacting particle systems, the effect of interparticle interactions is accounted for by shifting the energy barrier distribution to higher energies and thereby increasing the relaxation times. Some features shown in Figs. 1 and 2, such as a shift of the maxima in $`\chi (T,\omega )`$ to higher temperatures and the appearance of weak positive peaks in $`\chi _3(T,\omega )`$ at low temperatures, can in principle be explained by a significant increase of the individual particle relaxation times. However, for the here investigated particle system, it has previously been shown that the low temperature magnetic relaxation for the most interacting sample resembles that observed for archetypal spin glasses. At temperatures below 45 K, the relaxation time spectrum broadens significantly and the magnetic relaxation is significantly different from that of the noninteracting sample. This together with the observed aging effect give strong indications that the low temperature magnetic relaxation is dominated by collective particle dynamics. With increasing temperature, however, the time scale of the collective dynamics is gradually shifted to shorter time scales and the slow magnetic relaxation remaining is due to single-particle relaxation of a comparably small number of large particles. Still, the relaxation of these large particles is influenced by small particles surrounding them - relaxing particles will experience a magnetic field from neighboring polarized superparamagnetic particles, increasing the energy barriers and thereby the relaxation times. This gradual change of the magnetic relaxation with increasing temperature, from collective to single particle dynamics, obstructs the observation of critical slowing down in this particle system. For the same reason, it will not be possible to experimentally observe a divergent cubic equilibrium susceptibility. Nevertheless, the observation of collective behavior of the magnetic relaxation at low temperature makes it interesting to compare the behavior of the most interacting sample to that of a typical spin glass with long range interactions.
Figure 5 shows the susceptibility of a Ag(11 at.% Mn) spin glass measured at the frequency $`\omega /2\pi =125`$ Hz and with an ac field amplitude of 1600 A/m. The real part of the linear susceptibility has a sharp cusp at about the same temperature at which there is a sharp rise from zero of the imaginary part. The two components of the cubic susceptibility have sharp negative peaks at high temperatures followed by broad positive peaks at low temperatures. Before comparing this behavior with that of the most concentrated particle system, we will point out some fundamental differences between an interacting particle system and a spin glass.
In a particle system the single particle relaxation time has an exponential temperature dependence and in addition depends upon the particle volume and the anisotropy constant, while in a spin glass the individual spins have a relaxation time in the order of $`10^{12}`$ s to $`10^{14}`$ s. In a spin glass the slow dynamics is only due to collective phenomena and thus the imaginary component of the susceptibility disappears when the maximum relaxation time of the spin system becomes shorter than the observation time $`1/\omega `$. The dynamics of the investigated particle system is, as discussed above, dominated by collective particle behavior at low temperatures, while at higher temperatures the slow dynamics are dominated by single particle relaxation. When comparing the two systems, we are thus restricted to the low temperature regime where both systems display collective dynamics. For the investigated interacting particle system, this implies temperatures $`T<45`$ K.
The linear and cubic susceptibilities of the most concentrated particle sample (see Figs. 1 and 2) display features much broader in temperature as compared to the corresponding features observed for the spin glass sample. This can be explained as a combination of two effects: the polydispersivity of the particle system and that the maxima in the linear susceptibility as well as the negative peaks in the cubic susceptibility are partly due to single particle relaxation. It is tempting though to attribute the low temperature positive peaks in the interacting particle sample to collective dynamics, since also the spin glass shows broad positive peaks in the cubic dynamic susceptibility at low temperatures. This feature is found at temperatures $`T<40`$ K, i.e., in the temperature range where the aforementioned study of the magnetic relaxation of the most interacting sample revealed evidence for collective particle behavior.
## V Conclusions
We have studied the linear and cubic dynamic susceptibilities of solid dispersions of maghemite nanoparticles with various strengths of the inter-particle interactions. We have found that the expressions for the dynamical susceptibilities proposed by Raikher and Stepanov, describe the experimental results for the most dilute sample with a reasonable degree of accuracy. Nevertheless, further developments of the theoretical modeling (including other symmetries of the anisotropy energy, effects of the damping parameter, etc.) would be desirable to reduce the gap between theory and experiments.
Concerning the features observed in the most concentrated sample an explanation in terms of single particle dynamics with equilibrium susceptibilities and energy barriers modified by the interactions cannot be excluded only on the basis of the present study. However, previous evidence of collective magnetic behavior in the same system strongly suggests this as a more consistent interpretation of the obtained results. This is also supported by the comparison of the results obtained for the most concentrated particle system with the nonlinear response of an archetypal spin glass.
###### Acknowledgements.
This work was financially supported by The Swedish Natural Science Research Council (NFR).
|
no-problem/9812/astro-ph9812343.html
|
ar5iv
|
text
|
# Cygnus X-2, super-Eddington mass transfer, and pulsar binaries
## 1 INTRODUCTION
Cygnus X-2 is a persistent X-ray binary with a long orbital period ($`P=9.84`$ d, Cowley, Crampton & Hutchings 1979). The observation of unambiguous Type I X-ray bursts (Smale, 1998) shows that the accreting component is a neutron star rather than a black hole. The precise spectroscopic information found by Casares, Charles & Kuulkers (1998), and the parameters which can be derived from it, is summarized in Table 1. The mass ratio $`q=M_2/M_10.34`$ implies that mass transfer widens the system, and is therefore probably driven by expansion of the secondary star. Normally in long-period low-mass X-ray binaries (LMXBs) this occurs because of the nuclear evolution of a subgiant secondary along the Hayashi line, with typical effective temperatures $`T_{\mathrm{eff},2}30004000`$ K. However Casares et al.’s observations show that this cannot be the case for Cygnus X-2. The secondary is in the Hertzsprung gap (spectral type A9 III): use of Roche geometry and the Stefan–Boltzmann law gives $`L_2150\mathrm{L}_{}`$ with $`T_{\mathrm{eff},2}7330`$ K (see Table 1). Moreover the mass ratio $`q0.34`$, and the assumption that the primary is a neutron star and thus obeys $`M_12\mathrm{M}_{}`$, implies that the secondary has a low mass ($`M_2=qM_10.68\mathrm{M}_{}`$). In contrast, an isolated A9 III star would have a mass of about $`4\mathrm{M}_{}`$. More recently Orosz & Kuulkers (1998) have modelled the ellipsoidal variations of the secondary and thereby derived a model-dependent inclination of $`i=62.5^{}\pm 4^{}`$ which translates into component masses $`(M_1=1.78\pm 0.23)\mathrm{M}_{}`$ and $`(M_2=0.60\pm 0.13)\mathrm{M}_{}`$.
In this paper we consider explanations for the unusual nature of the secondary in Cygnus X-2. We find only one viable possibility, namely that this star is currently close to the end of early massive Case B mass transfer, and thus that the neutron star has somehow managed to reject most of the mass ($`3\mathrm{M}_{}`$) transferred to it in the past. In support of this idea, we show that this type of evolution naturally explains the surprisingly large companion masses in several millisecond pulsar binaries.
## 2 MODELS FOR CYGNUS X-2
In this Section we consider four possible explanations for the unusual nature of the secondary in Cygnus X-2. We shall find that three of them are untenable, and thus concentrate on the fourth possibility.
### 2.1 A normal star at the onset of Case B mass transfer?
The simplest explanation is that the position of the secondary in the HR diagram is just that of a normal star crossing the Hertzsprung gap. Since such a star no longer burns hydrogen in the core, this is a massive Case B mass transfer as defined by Kippenhahn & Weigert (1967), hereafter KW. Provided that the initial mass ratio $`q_\mathrm{i}1`$ the binary always expands on mass transfer, which occurs on a thermal time-scale. Kolb (1998) investigated this type of evolution systematically and found that the secondary’s position on the HR diagram is always close to that of a single star of the same instantaneous mass. For Cygnus X-2 the A9 III spectral type would require a current secondary mass $`M_24\mathrm{M}_{}`$, and thus a primary mass $`M_1=M_2/q12\mathrm{M}_{}`$. This is far above the maximum mass for a neutron star and would require the primary to be a black hole, in complete contradiction to the observation of Type I X-ray bursts from Cygnus X-2 (Smale, 1998).
We conclude that the secondary of Cygnus X-2 cannot be a normal star. Accordingly we must consider explanations in which it is far bigger and more luminous than expected from its estimated mass $`(0.490.68)\mathrm{M}_{}`$.
### 2.2 A stripped subgiant?
The type of Case B evolution described in subsection (2.1) above is known as ‘early’, in that mass transfer starts when the donor’s envelope is still largely radiative rather than convective (as it would become as the star approached the Hayashi line), and ‘massive’, meaning that the helium core has a large enough mass that upon core contraction it does not become degenerate but instead ignites central helium burning. The corresponding minimum core mass is about $`0.35\mathrm{M}_{}`$, corresponding to a total ZAMS mass of $`22.5\mathrm{M}_{}`$ (depending on the assumed degree of convective overshooting during central hydrogen burning). For lower initial masses we have ‘low-mass’ Case B (Kippenhahn, Kohl & Weigert 1967). Here the donor’s helium core becomes degenerate and the envelope is convective. After a possible early rapid mass transfer phase in which the binary mass ratio $`q=M_2/M_1`$ is reduced to $`1`$, (Bhattacharya & van den Heuvel, 1991; Kalogera & Webbink, 1996) the donor reaches the Hayashi line and mass transfer is driven by its nuclear expansion under hydrogen shell burning. The star remains on the Hayashi line, increasing its radius and the binary period as its core mass grows. Webbink, Rappaport & Savonije (1983) describe this type of ‘stripped giant/subgiant’ evolution in detail, and indeed fit Cygnus X-2 in this way. However there is a clear discrepancy between the observed effective temperature and that required for a Hayashi-line donor, which should be about 4100 K in this case. Webbink et al. (1983) appeal to X-ray heating by the primary to raise the temperature to the observed 7330 K, but remark that since the heating only operates on the side of the secondary facing the primary, one would expect a much larger orbital modulation of the optical flux than actually observed ($`\mathrm{\Delta }V_{\mathrm{obs}}0.3`$ mag) unless the orbital inclination $`i`$ is low. Simple estimates (see the Appendix) show that such a small modulation would require $`i<13.4^{}`$. However the mass function and mass ratio for Cygnus X-2 can be combined to show that $`M_1\mathrm{sin}^3i=(1.25\pm 0.09)\mathrm{M}_{}`$, so such small inclinations would imply $`M_1>100\mathrm{M}_{}`$, again clearly incompatible with the very strong observational evidence for a neutron star primary. A still stronger argument can be constructed on the basis that the spectral type of the secondary is not observed to vary during the orbital cycle.
### 2.3 A helium white dwarf undergoing a hydrogen shell flash?
It is known that newly-born helium white dwarfs can undergo one or more hydrogen shell flashes during their evolution from the giant branch to the white dwarf cooling sequence. During these flashes the star has a much larger photosphere. Calculations by Driebe et al. (1998) show that only low-mass He white dwarfs in the interval $`0.21\mathrm{M}_{}M_{\mathrm{WD}}0.30\mathrm{M}_{}`$ can undergo such a flash which in turn can put the star in the same position on the HR diagram as the secondary of Cygnus X-2 (e.g. the first shell flash of the sequence with $`M_{\mathrm{WD}}=0.259\mathrm{M}_{}`$). However, the required low secondary mass has a price: the observed mass ratio implies $`M_1=qM_2(0.76\pm 0.09)\mathrm{M}_{}`$, much smaller than required by the mass function ($`M_1\mathrm{sin}^3i=(1.25\pm 0.09)\mathrm{M}_{}`$). To makes matters worse, the evolutionary track crosses the relevant region of the HR diagram in an extremely short time: the star’s radius expands on a time-scale $`\tau =\mathrm{d}t/\mathrm{d}\mathrm{ln}R_233`$ yr. Not only does this give the present system an implausibly short lifetime, the radius expansion would drive mass transfer at a rate $`\mathrm{\Delta }M_\mathrm{H}/\tau `$ few $`\times 10^4\mathrm{M}_{}`$ yr<sup>-1</sup>, again totally inconsistent with observations. We conclude that the secondary of Cygnus X-2 cannot be a low-mass He white dwarf undergoing a hydrogen shell flash.
### 2.4 A star near the end of early massive Case B mass transfer?
We saw in subsection (2.1) above that a secondary near the onset of early massive Case B mass transfer (i.e. with $`q1`$ throughout) is ruled out for Cygnus X-2, as the required stellar masses conflict with observation. However, a more promising assignment is a secondary near the end of an early massive Case B evolution which began with $`q_\mathrm{i}1`$.
KW have investigated this process in detail. In contrast to the case $`q_\mathrm{i}1`$ discussed by Kolb (1998), and considered in (2.1) above, the ratio $`q_\mathrm{i}1`$ means that the binary and Roche lobe initially shrink on mass transfer. Adiabatic stability is nevertheless ensured because the secondary’s deep radiative envelope (‘early’ Case B) contracts on rapid mass loss. Mass transfer is therefore driven by the thermal-time-scale expansion of the envelope, but is more rapid than for $`q_\mathrm{i}1`$ because of the orbital shrinkage. Once $`M_2`$ is reduced to the point that $`q1`$, the Roche lobe begins to expand. This slows the mass transfer, and shuts it off entirely when the lobe reaches the thermal-equilibrium radius of the secondary, since the latter then has no tendency to expand further (except possibly on a much longer nuclear time-scale). Calculations by KW and Giannone, Kohl & Weigert (1968), hereafter GKW, show that in some cases the orbit can shrink so much that the process ends in the complete exhaustion of the donor’s hydrogen envelope, ultimately leaving the core of the secondary in a detached binary. Alternatively, the rapid Case B mass transfer may end with the donor on the Hayashi line, still retaining a large fraction of its original hydrogen envelope. However, there will be no long-lasting phase of mass transfer with the donor on the Hayashi line because the star starts shrinking with ignition of central helium burning (KW). Neither of these two cases describes Cygnus X-2. However, there is an intermediate possibility: the initial mass ratio $`q_\mathrm{i}`$ may be such that the donor retains a small but non-negligible hydrogen envelope as mass transfer slows. The current effective temperature of 7330 K shows that the companion’s envelope is mainly radiative, with only a very thin surface convection zone. (A paper in preparation by Kolb et al. shows this in detail.) In the example computed by KW the donor, at the end of mass transfer, is not on the Hayashi line, but at almost the same point in the HR diagram as it occupied immediately before mass transfer began (this is confirmed by the detailed numerical calculations of Kolb et al.). Just before the process ends we then have an expanding low-mass donor, driving a modest mass transfer rate in a long-period expanding binary, but at the HR diagram position of a much more massive normal star. As we shall show below, for an initial donor mass of about $`3.6\mathrm{M}_{}`$ the end point of such an evolution can be made to match closely that observed for the secondary of Cygnus X-2. (Note that we have not performed detailed numerical calculations for this paper, but rather used the results of KW and GKW. The forthcoming paper by Kolb et al. reports detailed calculations.)
Clearly this idea offers a promising explanation of the secondary in Cygnus X-2. However there is an obvious difficulty in accepting it immediately. KW’s calculations assumed that the total binary mass and angular momentum were conserved, and in particular that the primary retained all the mass transferred to it. But the primary of Cygnus X-2 is known to be a neutron star, with a mass presumably $`2\mathrm{M}_{}`$, so we must require instead that it accretes relatively little during mass transfer. This agrees with the idea that a neutron star cannot accrete at rates greatly in excess of its Eddington limit ($`10^8\mathrm{M}_{}`$ yr<sup>-1</sup>), and the fact that almost all of the mass is transferred at much higher rates ($`10^6\mathrm{M}_{}`$ yr<sup>-1</sup>). We thus follow earlier authors (Bhattacharya & van den Heuvel, 1991; Kalogera & Webbink, 1996) in postulating that the neutron star is extremely efficient in ejecting most of the super–Eddington mass transfer, rather than allowing the excess mass to build up into a common envelope. Clearly common-envelope evolution could not produce Cygnus X-2: the current binary period of 9.8 d shows that far too little orbital energy could have been released to remove the envelope of any plausible progenitor for the secondary (see e.g. Section 4 below). By contrast, it is at least energetically possible to expel most of a super–Eddington mass transfer rate, provided that this is done at large enough distance $`R_{\mathrm{ej}}`$ from the neutron star. If the matter is given just the escape velocity at $`R_{\mathrm{ej}}`$ the ratio of ejection to accretion rate is
$$\frac{\dot{M}_{\mathrm{ej}}}{\dot{M}_{\mathrm{acc}}}=\frac{R_{\mathrm{ej}}}{R_{}},$$
(1)
where $`R_{}`$ is the radius of the neutron star. Ejecting all but about 1% of the transferred matter thus requires
$$R_{\mathrm{ej}}100R_{}10^8\mathrm{cm}$$
(2)
which is far smaller than the size of the accretion disc around the neutron star for example. (This point is discussed further by King and Begelman, 1999.)
In the next section we will consider an early massive Case B evolution for Cygnus X-2. We will find that the hypothesis of efficient mass ejection by the neutron star allows excellent agreement with the current state of the system, as well as plausible explanations for the observed states of several detached pulsar binaries.
## 3 EARLY MASSIVE CASE B EVOLUTION FOR NEUTRON-STAR BINARIES
The main features of early massive Case B evolution can be understood by considering the relative expansion or contraction of the donor’s Roche lobe $`R_\mathrm{L}`$ and the thermal equilibrium radius $`R_{2,\mathrm{e}}`$ which the donor attains at the end of mass transfer (see e.g. GKW). As discussed above, we assume that the neutron star ejects any super–Eddington mass inflow. Since the mass transfer rate exceeds the Eddington limit by factors $`100`$ (see above), almost all of the transferred mass must be ejected, and to an excellent approximation we can assume that the neutron star mass $`M_1`$ remains fixed during the mass transfer (the equation for $`R_\mathrm{L}`$ can actually be integrated exactly even without this assumption, but at the cost of some algebraic complexity). We assume further that the ejected mass carries the specific angular momentum of the neutron star’s orbit. This is very reasonable, since the ejection region is much smaller than the size of the disc (see eq. 2). Then we can use the result quoted by Kalogera & Webbink (1996) to write
$$\frac{R_\mathrm{L}}{R_{\mathrm{L},\mathrm{i}}}=\left(\frac{M_{2\mathrm{i}}}{M_2}\right)^{5/3}\left(\frac{M_\mathrm{i}}{M}\right)^{4/3}e^{2(M_2M_{2\mathrm{i}})/M_1},$$
(3)
where $`M_2`$ and $`M=M_1+M_2`$ are the donor and total binary mass at any instant, and $`M_{2\mathrm{i}},M_\mathrm{i}`$ their values at the onset of mass transfer. In writing (3) we have used the simple approximation $`R_\mathrm{L}/a(M_2/M)^{1/3}`$, where $`a`$ is the binary separation. Using this and Kepler’s law we get the change of binary period $`P`$ as
$$\frac{P}{P_i}=\left(\frac{R_\mathrm{L}}{R_{\mathrm{L},\mathrm{i}}}\right)^{3/2}\left(\frac{M_{2\mathrm{i}}}{M_2}\right)^{1/2}$$
(4)
so that
$$\frac{P}{P_i}=\left(\frac{M_{2\mathrm{i}}}{M_2}\right)^3\left(\frac{M_\mathrm{i}}{M}\right)^2e^{3(M_2M_{2\mathrm{i}})/M_1}.$$
(5)
Conventional massive Case B evolution always begins with a mass ratio $`q_\mathrm{i}=M_{2\mathrm{i}}/M_1`$ sufficiently large that $`R_\mathrm{L}`$ initially shrinks. From (3) it is easy to show that this requires $`q_\mathrm{i}>1.2`$. If during the evolution $`M_2`$ decreases enough that $`q<1.2`$, $`R_\mathrm{L}`$ begins to expand again. The curves of $`\mathrm{log}R_\mathrm{L}`$ thus have the generic U-shaped forms shown in Figs. 1 – 3.
The thermal equilibrium radius $`R_{2,\mathrm{e}}`$ depends on the relative mass $`M_{\mathrm{He}}/M_2`$ of the donor’s helium core. Figs. 1 – 3 show the so-called ‘generalized main sequences’ of GKW, the first schematically, and the latter two for $`M_{2\mathrm{i}}=3\mathrm{M}_{},5\mathrm{M}_{}`$. For a core-envelope structure to be applicable, the star must have at least finished central hydrogen burning. Since the mass transfer takes place on the thermal time-scale of the donor there is little nuclear evolution, and we can regard $`M_{\mathrm{He}}`$ as fixed. The quantity $`R_{2,\mathrm{e}}`$ shown in Figs. 1 – 3 therefore gives the thermal equilibrium radius attained by the star after transferring varying amounts of its hydrogen envelope. The evolution of the system is now specified by the initial mass ratio $`q_\mathrm{i}`$ and the radius $`R_{2\mathrm{i}}`$ of the donor at the onset of mass transfer. This can lie between the maximum radius $`R_\mathrm{B}`$ reached during central hydrogen burning and one almost as large as the value $`R_{\mathrm{HL}}`$ at the Hayashi line (mass transfer is adiabatically unstable if the donor develops a deep convective envelope). The allowed initial radius range is about a factor 2 for a donor with $`M_{2\mathrm{i}}=2.5\mathrm{M}_{}`$, increasing to a factor $`6`$ for $`M_{2\mathrm{i}}=5\mathrm{M}_{}`$ (Bressan et al. 1993).
We see from the schematic Fig. 1 that three qualitatively different outcomes of early massive Case B mass transfer are possible:
1. ‘small’ $`q_\mathrm{i}`$, i.e. $`M_{2\mathrm{i}}`$ only slightly larger than $`M_1`$. Here the Roche lobe soon begins to expand, so the curves of $`\mathrm{log}R_\mathrm{L}`$ and $`\mathrm{log}R_{2,\mathrm{e}}`$ cross before much mass is transferred. Mass transfer ends with ignition of central helium burning as this makes the star shrink somewhat. At this point the secondary still has a thick hydrogen envelope and lies on the Hayashi line, with the binary having a long orbital period. During central helium burning the secondary stays on the Hayashi line but remains detached. Mass transfer resumes only after central helium burning when the star approaches the asymptotic giant branch (AGB). Mass transfer is again driven by nuclear evolution (double shell-burning). Since core helium-burning has ceased this is Case C. Here (unusually) the mass transfer is adiabatically stable, despite the secondary’s deep convective envelope, since the mass ratio is already $`1`$. The average mass transfer rate is again of order $`10^6\mathrm{M}_{}`$ yr<sup>-1</sup>, but may become up to an order of magnitude higher during thermal pulses (Pastetter & Ritter 1989). The donor may also lose large amounts of envelope mass in a wind. If the binary again manages to avoid common-envelope evolution by ejecting most of the transferred mass, mass transfer will finally end once the secondary’s envelope has been lost, leaving a very wide (period $`10`$ yr) binary containing a neutron star and a CO white dwarf.
2. ‘critical’ $`q_\mathrm{i}`$. We define this as the case where the $`R_\mathrm{L},R_{2,\mathrm{e}}`$ curves cross at the ‘knee’ in the mass–radius curve, i.e. with $`M_2`$ only slightly larger than $`M_{\mathrm{He}}`$, so the mass transfer depletes almost the entire envelope. The remnant donor retains a thin hydrogen envelope, and lies between the main sequence and the Hayashi line on the HR diagram. Thus the envelope mass is low enough to prevent the star lying on the Hayashi line, but not so low that the remnant is small and hot, i.e. to the left of the main sequence. The initial separation must be small enough that mass transfer starts before central helium burning, but large enough that it starts only after the donor has reached the Schönberg–Chandrasekhar limit. This limit is defined as the point where the isothermal helium core has reached the maximum mass which is able to support the overlying layers of the star, i.e. the point at which core collapse begins. The orbital period is shorter than in case 1., but longer than in case 3. below. Cygnus X-2 is an example of this evolution, viewed at the point where the donor has almost attained its thermal-equilibrium radius, and mass transfer is well below the maximum thermal-time-scale rate. This evolution ends with nuclear evolution of the donor to smaller radii as the mass of the hydrogen-rich envelope is further reduced by shell burning. The system detaches, leaving a helium-star remnant which subsequently ignites central helium burning and finally becomes a CO white dwarf.
3. ‘large’ $`q_\mathrm{i}`$. The curves cross only when the hydrogen envelope is effectively exhausted. The remnant is a helium star and the orbital period is short. If $`M_{\mathrm{He}}0.9\mathrm{M}_{}`$ the helium star evolves directly into a CO white dwarf. If $`M_{\mathrm{He}}1\mathrm{M}_{}`$ this star re-expands during helium shell-burning. This in turn can give rise to a further phase of (so-called Case BB) mass transfer, e.g. Delgado & Thomas, 1981; Law & Ritter, 1983; Habets, 1985, 1986).
Table 2 summarizes the possible Case B evolutions with a neutron-star primary. The various outcomes all reflect the general tendency of larger $`q_\mathrm{i}`$ (i.e. larger $`M_{2\mathrm{i}}`$) to produce greater orbital contraction, and thus smaller and relatively less massive remnants (their mass is a smaller fraction of a larger $`M_{2i}`$) and short orbital periods. Conversely, larger $`R_{2\mathrm{i}}`$ in the range $`R_\mathrm{B}R_{\mathrm{HL}}`$, where $`R_\mathrm{B}`$ is the maximum radius reached during central hydrogen burning, produces exactly the opposite trends. To model Cygnus X-2 one thus needs $`q_\mathrm{i}`$ close to the critical value.
As can be seen from the following arguments the range of possible solutions is strongly constrained by equation (3): with the assumption that $`M_1`$ remains essentially constant during the evolution, $`M_{2\mathrm{f}}`$ is fixed by the observed mass ratio $`q_\mathrm{f}=0.34\pm 0.04`$. On the other hand, $`q_\mathrm{i}`$ is also essentially fixed by the model assumption that the current donor is close to the end of Case B mass transfer. This in turn means that the donor is now close to thermal equilibrium, with its luminosity therefore coming mainly from hydrogen shell burning.
Use of the generalized main sequence for $`3\mathrm{M}_{}`$ stars given in GKW (their fig. 7, shown here in Fig. 2) demonstrates that one can account for the current state of the donor in Cygnus X-2 if this star had $`M_{2\mathrm{i}}3\mathrm{M}_{}`$ (i.e. $`q_\mathrm{i}2.1`$) and $`0.55R_{\mathrm{HL}}R_{2\mathrm{i}}0.75R_{\mathrm{HL}}`$ when mass transfer began, corresponding roughly to $`0.5\mathrm{M}_{}M_{2\mathrm{f}}0.7\mathrm{M}_{}`$. The numerical results of GKW (their Table 2 and Fig. 4) suggest that the current state of the donor is then given approximately by $`M_2=0.79\mathrm{M}_{},M_{\mathrm{He}}=0.67\mathrm{M}_{}`$ and thus $`q_0=M_{\mathrm{He}}/M_2=0.85`$; with $`T_{\mathrm{eff},2}=7060`$ K, $`L_2=126\mathrm{L}_{}`$ and $`R_2=7.5\mathrm{R}_{}`$. These values correspond to an orbital period $`P8.4`$ d. From (5) we infer that mass transfer started at an orbital period $`P_\mathrm{i}3.5`$ d. These quantities are close to those given in Table 1, although the predicted $`M_2`$ is slightly larger than the observational estimate. This may either result from the fact that the chemical composition of the models defining the generalized main sequences is characterized by a step function at $`M_\mathrm{r}=M_\mathrm{c}`$ and thus differs from that of evolutionary models, or from the fact that the initial chemical composition ($`X=0.602,Z=0.044`$) and the opacities used by GKW are rather outdated. In addition the core mass $`M_{\mathrm{He}}=0.67\mathrm{M}_{}`$ inferred above points to an initial mass higher than the $`3\mathrm{M}_{}`$ suggested earlier. In fact to fit the observed value of $`0.5\mathrm{M}_{}M_{2\mathrm{f}}0.7\mathrm{M}_{}`$, more modern calculations than those of GKW (e.g. Bressan et al. 1993) yield the required core mass if the initial mass was $`3.2M_{2\mathrm{i}}4.1\mathrm{M}_{}`$.
We conclude that the observational data for Cygnus X-2 are well reproduced if we assume it is a remnant of early massive Case B evolution with $`q_\mathrm{i}`$ close to the critical value $`2.32.9`$, if $`M_11.4\mathrm{M}_{}`$, or $`1.82.3`$, if we adopt the value $`M_11.8\mathrm{M}_{}`$ derived by Orosz & Kuulkers (1998). For the remainder of this paper we shall adopt $`M_1=1.4\mathrm{M}_{}`$.
## 4 END PRODUCTS
The discussion above shows that for both low-mass Case B, and for early massive Case B with $`q_\mathrm{i}`$ below a critical value ($`2.6`$ for $`M_{2i}3.6\mathrm{M}_{}`$), the evolution leads to a long-period binary with the donor on the Hayashi line. As is well known, the luminosity and radius of such a star are fixed by its degenerate core mass $`M_\mathrm{c}`$ rather than its total mass $`M_2`$. Even though the degenerate core is different in nature in the two cases, it is possible to give a single formula for the radius, i.e
$$r_2=\frac{3.7\times 10^3m_\mathrm{c}^4}{1+m_\mathrm{c}^3+1.75m_\mathrm{c}^4},$$
(6)
where $`r_2=R_2/\mathrm{R}_{},m_\mathrm{c}=M_\mathrm{c}/\mathrm{M}_{}`$ (Joss, Rappaport & Lewis, 1987). Then using the well-known relation
$$P=0.38\frac{r_2^{3/2}}{m_2^{1/2}}\mathrm{d}$$
(7)
($`m_2=M_2/\mathrm{M}_{}`$) which follows from Roche geometry, we get a relation between $`M_\mathrm{c},M_2`$ and the orbital period $`P`$ (e.g. King, 1988). Once all of the envelope mass has been transferred we are left with a wide binary containing a millisecond pulsar (the spun-up neutron star) in a circular orbit with the white-dwarf core of the donor. Since at the end of mass transfer we obviously have $`m_2=m_\mathrm{c}`$, such systems should obey the relation
$$P8.5\times 10^4\left(\frac{m_2^{11/2}}{[1+m_2^3+1.75m_2^4]^{3/2}}\right)\mathrm{d}.$$
(8)
The timing orbit of the millisecond pulsar allows constraints on the companion mass, so this relation can be tested by observation. Lorimer et al. (1995), Rappaport et al. (1995) and Burderi, King & Wynn (1996) show that while the relation is consistent with the data for a majority of the $`25`$ relevant systems, there are several systems (currently 3 or 4; see Fig. 4 and Table 3) for which the white dwarf mass is probably too large to fit. While one might possibly exclude B0655+64 because of its long spin period $`P_\mathrm{s}`$ (but see below), the other three systems are clearly genuine millisecond pulsars.
Our considerations here offer a simple explanation for this discrepancy. If the initial mass ratio $`q_\mathrm{i}`$ lies above the critical value ($`2.6`$ for $`M_{2i}3.6\mathrm{M}_{}`$), the donor radius will be less than $`R_{\mathrm{HL}}`$ at the end of early massive Case B mass transfer, and the orbital period relatively shorter. When such systems finally detach from the Roche lobe, the WD companion is considerably more massive than expected for the orbital period on the basis of the Hayashi-line relation (8). We see from (5) that systems with large initial companion masses $`M_{2\mathrm{i}}`$ ($`4\mathrm{M}_{}`$) can end as short-period systems. Table 4 and Fig. 4 show the expected minimum final periods $`P_\mathrm{f}`$ and companion masses $`M_2`$ for various $`M_{2\mathrm{i}}`$, i.e. assuming that mass transfer began with the donor at the Schönberg–Chandrasekhar limit.
Table 4 and Fig. 4 show that early massive Case B evolution with $`q_\mathrm{i}`$ larger than the critical value can explain the discrepant pulsar systems of Table 3. Indeed it appears that the process can end with very short orbital periods, offering an alternative to the usual assumption of a common-envelope phase (Bhattacharya, 1996; Tauris, 1996) The limiting factor for this kind of evolution may be the so–called ‘delayed dynamical instability’ (Webbink, 1977; Hjellming, 1989). For a sufficiently massive initial donor, mass transfer eventually becomes dynamically unstable because the adiabatic mass radius exponent of a strongly stripped radiative star becomes negative and the donor begins to expand adiabatically in response to mass loss. For typical neutron star masses $`M_11.4\mathrm{M}_{}`$ this would limit $`M_{2\mathrm{i}}`$ to values $`(44.5)\mathrm{M}_{}`$ (Hjellming 1989; Kalogera & Webbink, 1996). Given the uncertainties in our estimates, this is probably consistent with the initial mass $`M_{2\mathrm{i}}5\mathrm{M}_{}`$ needed to explain the current state of the most extreme discrepant system (B0655+64). The other three systems can all be fitted with $`M_{2\mathrm{i}}4\mathrm{M}_{}`$, so there is no necessary conflict with the mass limits for the delayed dynamical instability.
We may now ask about the end states of such an evolution if the neutron star were unable to eject the mass transferred at the very high rates expected once the delayed dynamical instability set in and the system went through a common envelope phase instead. Using the standard prescription for estimating the parameters of a post common envelope system (e.g. Webbink 1984) with the values for $`M_{2\mathrm{i}}`$ and $`M_{2\mathrm{f}}=M_2`$ given in Table 4, $`M_1=1.4\mathrm{M}_{}`$ and $`\alpha _{\mathrm{CE}}\lambda =0.5`$, where $`\lambda 0.5`$ is a structural parameter and $`\alpha _{\mathrm{CE}}`$ the common envelope efficiency parameter defined by Webbink (1984), we find that the final orbital period is $`0.04P_\mathrm{f}(d)0.2`$ if mass transfer set in when the donor was already near the Hayashi line, and smaller still if mass transfer set in earlier or if $`\alpha _{\mathrm{CE}}`$ is smaller than unity. On the other hand, unless $`\alpha _{\mathrm{CE}}`$ is significantly smaller than unity, common envelope evolution starting from a system with the donor on the asymptotic giant branch would end with periods much longer than those of the systems listed in Table 3 and shown in Fig. 4. Table 5 shows the outcome of common–envelope evolution in the two extreme cases where mass transfer starts at the Schönberg–Chandrasekhar limit, and on the Hayashi line. Although common–envelope efficiencies exceeding unity are discussed in the literature (i.e. energy sources other than the orbit are used to expel the envelope), the values found for $`\beta _{\mathrm{CE}}=\lambda \alpha _{\mathrm{CE}}`$ in Table 5 show that to produce Cyg X–2–like systems requires absurdly large efficiencies, even starting from the most distended donor possible. Thus common envelope evolution does not offer a promising explanation for these systems. Their very existence may thus indicate that an accreting neutron star can eject mass efficiently even at the very high mass transfer rates encountered in the delayed dynamical instability. We conclude therefore that even very rapid mass transfer on to a neutron star does not necessarily result in a common envelope (cf King & Begelman, 1999).
We note finally that all of the pulsars of Table 3 have spin periods much longer than their likely equilibrium periods (i.e. they lie far from the ‘spinup line’, cf Bhattacharya & van den Heuvel, 1991), suggesting that they have accreted very little mass ($`<<0.1\mathrm{M}_{}`$) during their evolution. This agrees with our proposal that these systems are the direct outcome of a super–Eddington mass transfer phase in which almost all the transferred mass is ejected.
## 5 SPACE VELOCITY AND POSITION OF CYGNUS X-2 IN THE GALAXY
The distance to Cygnus X-2 derived from the observations of Type I X-ray bursts (Smale 1989) is $`d=(11.6\pm 0.3)`$ kpc. From the galactic coordinates $`l=87.33^{}`$ and $`b=11.32^{}`$ and the solar galactocentric distance $`R_0=(8.7\pm 0.6)`$ kpc one derives a galactocentric distance for Cygnus X-2 of $`d_{\mathrm{GC}}=(14.2\pm 0.4)`$ kpc and a distance from the galactic plane of $`z=(2.28\pm 0.06)`$ kpc. Thus Cygnus X-2 has a very peculiar position indeed, being not only in the halo but also in the very outskirts of our galaxy. But not only is its position peculiar, its space velocity with respect to the galactic centre is also surprising. It can be shown that the observed heliocentric radial velocity of $`\gamma =(208.6\pm 0.8)`$ km s<sup>-1</sup> (Casares et al. 1998) is totally incompatible with prograde rotation on a circular, even inclined orbit around the galactic centre (Kolb et al., in preparation). The orbit is either highly eccentric and/or retrograde. In either case Cygnus X-2 must have undergone a major kick in the past, presumably when the neutron star was formed in a Type II supernova. Since prior to the supernova explosion the primary was much more massive ($`M_{1\mathrm{i}}10\mathrm{M}_{}`$) than the secondary ($`M_{2\mathrm{i}}3.6\mathrm{M}_{}`$), the latter was still on the main sequence when the supernova exploded. Thus the age of Cygnus X-2 (and the time elapsed since the supernova) is well approximated by the nuclear time-scale of the secondary, which is $`410^8`$ yr for a $`3.6\mathrm{M}_{}`$ star. This means that Cygnus X-2 must have gone around the galactic center a few times since its birth or supernova explosion and that, therefore, its birthplace in the galaxy cannot be inferred from its current position and velocity.
## 6 DISCUSSION
We have shown that the unusual nature of the secondary star in Cygnus X-2 can be understood if the system is near the end of a phase of early massive Case B evolution in which almost all of the transferred material is ejected. The system is unusual in having had an initial mass ratio $`q_\mathrm{i}=M_2/M_1`$ in a narrow critical range near $`q_\mathrm{i}2.6`$; smaller ratios lead to detached systems with the secondary near the Hayashi line, and larger ratios produce binary pulsars with fairly short orbital periods and relatively massive white dwarf companions. During this evolution, much of the companion’s original mass ($`3\mathrm{M}_{}`$ for Cygnus X-2) is transferred and consequently lost on the thermal time-scale $`10^6`$ yr of this star. Evidently the huge mass loss rate and the short duration of this phase make it difficult to detect any systems in this state; they would probably resemble Wolf–Rayet stars of the WNe type (i.e. showing hydrogen).
Cygnus X-2 is currently near the end of the thermal time scale mass transfer phase, so that its mass transfer rate is now well below the thermal time-scale value, and probably given by the accretion rate. At $`\dot{M}_{\mathrm{acc}}2\times 10^8\mathrm{M}_{}`$ yr<sup>-1</sup> (Smale, 1998), this is nevertheless one of the highest in any LMXB, making it easily detectable. Only a full calculation of the evolution, with in particular a detailed model for the secondary, can predict the duration of the current phase; this is not any easy task, as this star deviates strongly from thermal equilibrium during most of the evolution. But it is clear that the mass transfer rate will decline as the remaining few tenths of a solar mass in the hydrogen envelope are transferred. Cygnus X-2’s long orbital period and large accretion disc mean that even its current mass transfer rate only slightly exceeds the critical value required for a persistent rather than a transient LMXB (cf. King, Kolb & Burderi 1996), so the system will eventually become transient. Once the envelope has been transferred, mass transfer will stop, and the system will become a pulsar binary with about the current orbital period $`P=10`$ d, and a white dwarf companion with a mass which is slightly higher than that of the companion’s present helium core. Clearly since the present core mass is at least $`0.35\mathrm{M}_{}`$ this $`Pm_2`$ combination will not obey the Hayashi-line relation (8), so Cygnus X-2 will become another ‘discrepant’ system like those in Table 3.
The reasoning of the last paragraph shows that Cygnus X-2 will cease to be a persistent X-ray binary within the current mass transfer time-scale $`t_\mathrm{M}=(M_2M_\mathrm{c})/\dot{M}_{\mathrm{acc}}10^7`$ yr. Its past lifetime as a persistent source before the current epoch, and its future one as a detectable transient after it, are both likely to be of a similar order, although full evolutionary calculations are required to check this. The fact that we nevertheless observe even one system like Cygnus X-2 strongly suggests that the birthrate of such systems must be relatively high, i.e. $`10^7`$ yr<sup>-1</sup> in the Galaxy. Since the binary pulsar end-products of these systems have enormously long lifetimes, this may suggest that systems like Cygnus X-2 play a very important role in providing the Galactic population of millisecond pulsars.
Cygnus X-2 thus fits naturally into a unified description of long-period LMXBs in which super–Eddington Case B mass transfer is efficiently ejected by the neutron star. While the ejection process can already be inferred for the formation history of Hayashi-line LMXBs resulting from low-mass Case B evolution (Bhattacharya & van den Heuvel, 1991; Kalogera & Webbink, 1996), Cygnus X-2 supplies the most powerful evidence that this process must occur. The work of Section 2 shows that it is very hard otherwise to reconcile the rather low current mass ($`M_20.50.7\mathrm{M}_{}`$) of the secondary with its large radius ($`R_27\mathrm{R}_{}`$) and high luminosity ($`L_2150\mathrm{L}_{}`$). From Section 4 we see that the orbital period $`P=9.84`$ d is far too long for the system to be the product of common-envelope evolution, leaving no realistic alternative for driving the required mass ejection.
## 7 Acknowledgments
ARK thanks the Max–Planck–Institut für Astrophysik for its hospitality during 1998 August, and the U.K. Particle Physics and Astronomy Research Council for a Senior Fellowship. HR thanks the Leicester University Astronomy Group for its hospitality during 1998 November/December, and support from its PPARC Short-Term Visitor grant.
## 8 Appendix: X–ray Heating in Cygnus X–2
In Section 2.2 we considered a stripped subgiant model for Cyg X–2, and asserted that the observed orbital modulation of the optical flux ($`\mathrm{\Delta }V_{\mathrm{obs}}0.3`$ mag) would require an extremely low inclination if one appeals to X–ray heating of the companion to raise its observed effective temperature to 7330 K. Here we justify this claim.
We consider a simple picture in which the hemisphere of the (spherical) companion facing the neutron star has effective temperature 7330 K, while the other hemisphere has the Hayashi–line effective temperature 4100 K. We consider the effect of relaxing these assumptions below. Then viewing the heated face at the most favourable phase the observer sees hot and cool areas $`2\pi R_2^2(1/2+i/\pi ),2\pi R_2^2(1/2i/\pi )`$, where $`i`$ is the inclination in radians, with the two expressions reversing at the least favourable phase. Neglecting limb–darkening, the ratio of maximum to minimum flux is
$$\frac{F_{\mathrm{max}}}{F_{\mathrm{min}}}=\frac{(1/2+i/\pi )B_{\mathrm{hot}}+(1/2i/\pi )B_{\mathrm{cool}}}{(1/2+i/\pi )B_{\mathrm{cool}}+(1/2i/\pi )B_{\mathrm{hot}}},$$
(9)
where $`B_{\mathrm{hot}},B_{\mathrm{cool}}`$ are the optical surface brightnesses of the hot and cool regions respectively. Approximating these by Planck functions at $`5500\AA `$, we find $`B_{\mathrm{hot}}/B_{\mathrm{cool}}15`$. Requiring $`F_{\mathrm{max}}/F_{\mathrm{min}}1.3`$ ($`\mathrm{\Delta }V_{\mathrm{obs}}0.3`$ mag) in (9) shows that $`i0.0745\pi `$, or $`i13.4^{}`$, as used in Section 2.2
In reality the heated region would be smaller than a hemisphere, and its temperature higher than 7330 K in order to produce an average observed temperature of this value. However relaxing these limits clearly requires even smaller inclinations than the estimate above, because the contrast in optical surface brightness between the hot and cool regions would be even larger than the ratio $`15`$ we found above.
|
no-problem/9812/cond-mat9812294.html
|
ar5iv
|
text
|
# Strongly Correlated Electrons and Neutron Scattering
## 1 Introduction
Interactions between electrons in solids result in various states of matter and phase transitions. Coulomb repulsive forces are known to be the cause of magnetism. Especially in the presence of strong Coulomb interaction, i.e. in strongly correlated systems, Mott insulators are realized in the half-filled non-degenerate band, or in the case with integer number of electrons per site in general, where the low lying excitations are exclusively due to spin degrees of freedom. This situation is properly described by the Heisenberg spin Hamiltonian. Here spins and charges of electrons are completely separated. Although various types of magnetic ground states have been explored in a transparent way by this Heisenberg spin Hamiltonian, much of the recent interest are in the nonmagnetic singlet ground states, which are usually termed as spin-gapped systems. These singlet ground states are purely quantum-mechanical in their origin in comparison to magnetic states which are basically classical. Such studies on singlet ground states may be motivated by the fact that the high temperature superconductivity in cuprates, which has singlet $`d_{x^2y^2}`$ symmetry, are realized next to AF Mott insulators, i.e. in the doped Mott insulators where small amount of holes are introduced into the Mott insulating state . Here it is to be noted that AF superexchange interaction, $`J`$, has two distinct aspects; it leads quite generally to AF Néel ground states in lattices, while it results in the singlet states obviously for two spins and even in lattices in the presence of strong quantum fluctuations or frustrations. Actually it turned out that there exists a close relationship between antiferromagnetism and singlet $`d_{x^2y^2}`$ superconductivity . In this paper some of the recent development of the studies on the remarkable features of this interrelationship between antiferromagnetism and singlet ground states are introduced where neutron scattering experiments have played decisive roles.
## 2 Disordered Spin-Peierls Systems
The spin-Peierls transition has been known for some time but the discovery of this transition in inorganic compound, CuGeO<sub>3</sub>, by Hase et al is rather recent. Surprizing and puzzling was the report of the neutron scattering experiment by Regnault et al on CuGeO<sub>3</sub>, with a small amount of replacement of Ge by Si. This experiment has revealed the coexistence of resolution-limited sharp Bragg spots, one corresponding to the lattice dimerization stabilized below around the critical temperature of spin-Peierls transition of the parent clean systems and the other to AF ordering stabilized at much lower temperatures. Since these two states, dimerized and AF states, had been considered to be exclusive , some kind of inhomogeneity was suspected. At the same time, the sharpness of these two Bragg spots together with the apparent transfer of the spectral intensity between these two spots as the temperature is varied had suggested that the experiment had in fact disclosed the coexistence of the true long range order of these two conflicting order parameters. A theoretical proposal has been made, which indicates that such coexistence is possible once the spacial variations of the competing order parameters are taken into account. In this study the lattice distortions are treated completely classical because of the three-dimensionality of the actual crystals and the quantum mechanical features of one-dimensional spins are expressed in terms of the phase Hamiltonian derived by the bosonization , which has been treated in a mean field approximation in view of the existence of the interchain exchange interaction. Basic physics behind turns out to be that the realization of the singlet ground state by the dimerization is purely quantum mechanical and the complete coherency of the wave functions of spins in whole crystal (of the order of 10<sup>23</sup> !!) is needed and that, once the quantum phases of wave functions are perturbed, magnetizations are created locally, which naturally order in the absence of frustrations. The excitations above this new type of ground state with the coexistence turned out to have unique features : i.e. there are two distinct excitations, one with a gap reflecting dimerization which is present even in the clean systems and the other at very low energy and at around AF wavevector only with the very little total spectral weight in proportion to the degree of disorder,i.e. impurity concentration. In the coexisting ground state this small spectral weight at low energy forms well-defined spin wave mode reflecting the long range AF order. This is schematically shown in Fig.1.
Such a feature of two-mode structure has actually been observed in neutron scattering experiment and the existence of the well-defined spin waves has also been proven by the ESR experiments as well. At present the onset of the Néel ordering has been observed even down to 28.5mK at $`x=5\times 10^3`$ of Cu<sub>1-x</sub>Zn<sub>x</sub>GeO<sub>3</sub> by Manabe et al. . Note that there is little weight, i.e. “transparent”, in the energy region between the gap and spin wave mode. This is very unusual and unexpected since the effects of perturbations had been expected to be treated “perturbatively” in the presence of the gap in the excitation spectrum and then only the modifications of the gapped modes could be expected. At the same time this finding that the disorder introduces spectral weight only at very low energy (as low as the elastic Bragg spots) even in the presence of large gap will be a warning to correlate the experimental data of neutron scattering and NMR without serious considerations on the sample quality, since only the NMR will be affected by such low energy excitations. The present problem of disorder-induced antiferromagnetism in spin-Peierls systems can be considered to be a typical example showing how the spectral weight of the excitations emerges in the process of nucleation. Here the first order transition is expected in the clean systems between AF ordering and spin-Peierls dimerized state as a result of the competition between the interchain exchange interaction and the spin-lattice coupling along the chain . Once the disorder is induced, however, the antiferromagnetism is nucleated around the impurities in otherwise singlet ground state.
## 3 Heavy Electrons, Ce<sub>x</sub>Cu<sub>2</sub>Si<sub>2</sub>
Among strongly correlated electron sysytems, heavy electrons in lantanides and actinides have been studied for a long time and it is now known that magnetism and superconductivity are next to each other. Very recently Ishida et al found very mysterious properties in Cu-NQR spectra of series of Ce<sub>x</sub>Cu<sub>2</sub>Si<sub>2</sub>. At $`x`$=0.975 this system is AF ordered, while at $`x`$=1.025 the heavy electron superconductivity is observed. However at intermediate $`x`$=0.99 they found that there exist magnetic fluctuations with very low frequencies comparable to the NQR frequency ($`\omega `$=10<sup>6∼7</sup>)Hz even though no signatures of magnetism are seen in neutron scattering whose characteristic frequency is $`\omega `$=10<sup>11∼12</sup>Hz. Since there should be intrinsic disorder due to the spacial distribution of Ce atoms, a similar consideration as in the preceding disordered spin-Peierls systems will apply and then a similar feature as in Fig.1 is expected though the gapped mode here is due to superconductivity instead of dimerization in disordered spin-Peierls systems.
## 4 High $`T_\mathrm{c}`$ Cuprates
Disclosures of the true nature of magnetic exciatations have played crucial roles in the understanding of high $`T_\mathrm{c}`$ cuprates, for which both neutron scaterring and NMR have contributed very much. Findings by the former will be fully discussed in this Symposium. On the other hand NMR experiment on the undredoped Y(123) by Yasuoka et al. has first indicated the existence of the spin gap: one of the most remarkable features of the cuprates. All these experimental studies on magnetic excitations have revealed that the disorder affects the excitations in essential ways as in the case of disordered spin-Peierls systems. Few of the examples will be discussed in the following. First the complete destructions of the spin-gap in NMR by a very small amount of replacement of Cu by Zn in Y(248) will be understood again by Fig.1 if the spin-gap is associated with the formation of the short range order of spin-singlet as described by the RVB theory based on the slave boson mean field approximation of the $`t`$-$`J`$ model as shown in Fig.2 .
Here only the general trends of the spin-charge separation and their confinements are stressed. This spin gap can be identified with the pseudogap observed later by ARPES experiments since electrons are convolutions of spinons and holons , as has been explicitly demonstrated . In this context the underdoped region of La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> is very interesting. Early neutron scattering data by Keimer et al indicated the existence of spin glass phase, which at the same time has a feature close to the long-range ordered antiferromagnetgism of parent compounds. This finding has led the present author to conjecture that the true long range AF ordering might actually been realized at lower temperatures . However more recent neutron studies unambiguously demonstrated that the ground state is a canonical spin glass phase. A possible cause for this stability of spin-glass over the long range AF ordered state may be due to the extended nature of the wave function of the holes, which will lead to the frustration among spins . This difference between extended and local disorder is obviously dependent on the mobility of holes, i.e. insulating or (weakly) conducting, and may be related with the experimental findings of shift of the spectral weight toward the low energy in accordance with the resistivity increase as the temperature is lowered in underdoped YBCO .
## 5 NaV<sub>2</sub>O<sub>5</sub>
This system, which is insulating at all temperature, exhibits a magnetic transition at $`T_\mathrm{c}=35K`$ below which the spin susceptibility drops sharply but continuously , a feature typical of spin-gap systems, and has been believed to be an example of spin-Peierls transition.In this case the valences of V atoms are considered to be arranged as in Fig.3(b).
However recent NMR experiment on V atoms by Ohama et al indicated that the charge disproportionation sets in below $`T_\mathrm{c}`$ even though the the charge distribution on V atoms is uniform above $`T_\mathrm{c}`$. This finding stimulated many studies both theoretical and experimental. A theoretical study based on the selfconsistent Hartree approximation for both onsite and intersite Coulomb interaction has disclosed that the ground state will have a charge ordering of zigzag type as seen in Fig.3(c), which has a spin dimer structure. Above $`T_\mathrm{c}`$ the charge distributions are expected to be as in Fig.3(a). Basically all experimental data appear to be consistent so far with this state, including the spin excitation spectra probed by the neutron scattering by Yosihama et al . It will be very interesting to probe the transient region of this charge ordering which leads to the formation of the singlet ground state, i.e. the quenching of the quantum spins.
## Acknowledgements
The present research has been financially supported by the Grant-in Aid of Monbushou on the Priority Area “Anomalous Metallic State near the Mott Transition” and is based on the collaborations with M. Saito, T. Tanimoto, N. Nagaosa, M. Sigrist, A. Furusaki for subject 2 H. Kohno and T. Tanamoto for subject 4 and and H. Seo and H. Kino for subject 5 whom I thank deeply. Discussions on experiments with J.Akimitsu, Y. Endoh, Y. Fujii, K. Hirota, K. Kakurai, K. Katsumata, B. Keimer, N. Mori, M. Motokawa, M. Nishi, H. Nojiri, M. Sato, M. Sera, Z.X. Shen, G. Shirane, K. Yamada, H. Yasuoka K. Uchinokura and Y. Ueda are fully appreciated. Special thanks are to Y. Kitaoka for stimulating discussions on subject 3.
|
no-problem/9812/math9812038.html
|
ar5iv
|
text
|
# On locally 𝐿𝐶-spaces1991 Math. Subject Classification — Primary: 54D20, 54A05, Secondary: 54D45, 54G99. Key words and phrases — 𝐿𝐶-space, locally 𝐿𝐶-space, Lindelöf set.
## 1 Introduction
Classical generalizations of Lindelöf spaces such as hereditarily Lindelöf and maximal Lindelöf spaces have had their major impact in the development of General Topology. A certain class of spaces, relatively new as a concept but extensively studied in recent years, is the class of $`LC`$-spaces. A topological space $`(X,\tau )`$ whose Lindelöf subsets are closed is called an $`LC`$-space by Gauld, Mršević, Reilly and Vamanamurthy and by Mukherji and Sarkar . This concept emerged from the study of maximal Lindelöf spaces as being a notion having a close relationship to P-spaces.
$`LC`$-spaces generalize Wilansky’s KC-spaces and Hausdorff P-spaces . On the other hand every $`LC`$-space is a cid-space, i.e., all countable sets are closed and discrete, and hence $`T_1`$ and anti-compact (= pseudo-finite) . An extensive list of references on $`LC`$-spaces as well as some generalizations of the concept can be found in .
Recently, Ganster, Kanibir and Reilly introduced the class of locally $`LC`$-spaces. By definition, a topological space $`(X,\tau )`$ is called a locally $`LC`$-space if each point of $`X`$ has a neighborhood which is an $`LC`$-subspace. In , the authors proved that a space $`(X,\tau )`$ is an $`LC`$-space if and only if each point of $`X`$ has a closed neighborhood that is an $`LC`$-subspace. Thus every regular locally $`LC`$-space is an $`LC`$-space, a result first proved by Hdeib and Pareek in . The following example shows that we cannot replace ‘regular’ by ‘Hausdorff’.
###### Example 1.1
There exists a Hausdorff, locally $`LC`$-space $`(X,\tau )`$, which is not an $`LC`$-space. Let $`Z`$ be a set of cardinality $`\mathrm{}_1`$ with a distinguished point $`z_0`$. The topology on $`Z`$ is defined as follows: each $`zz_0`$ is isolated while the basic neighborhoods of $`z_0`$ are the cocountable subsets of $`Z`$ containing $`z_0`$. Note that $`Z`$ is a Lindelöf $`LC`$-space. The space $`(X,\tau )`$ will be constructed from copies of $`Z`$. For each $`n\omega `$, let $`X_n`$ be a copy of $`Z`$, where $`x_n`$ denotes the non-isolated point of $`X_n`$. Let $`X^{}=_{n\omega }X_n`$ denote the topological sum of the spaces $`X_n`$ and let $`X=X^{}\{p\}`$ with $`pX^{}`$. A topology $`\tau `$ on $`X`$ can be defined if, in addition, we specify the basic open neighborhoods of $`p`$. They are the union of $`\{p\}`$ and a cocountable subset of $`\{X_n\{x_n\}:nk\}`$ for some $`k\omega `$. $`(X,\tau )`$ is a Hausdorff space that fails to be an $`LC`$-space . However, as shown in , $`(X,\tau )`$ is a locally $`LC`$-space.
## 2 Locally $`LC`$-spaces
###### Proposition 2.1
For a topological space $`(X,\tau )`$ the following conditions are equivalent:
(1) $`X`$ is a locally $`LC`$-spaces.
(2) Every point of $`X`$ has an open neighborhood, which is an $`LC`$-subspace of $`X`$.
Proof. Follows from the fact that every subspace of an $`LC`$-space is an $`LC`$-space. $`\mathrm{}`$
###### Proposition 2.2
Every subspace of a locally $`LC`$-space is a locally $`LC`$-space.
Proof. Let $`(X,\tau )`$ be a locally $`LC`$-space and let $`AX`$. By assumption, for each $`xA`$, there exists $`U\tau `$ such that $`(U,\tau |U)`$ is an $`LC`$-space. Note that $`V=UA`$ is an open neighborhood of $`x`$ in $`(A,\tau |A)`$ and that $`(V,\tau |V)`$ is an $`LC`$-subspace. By Proposition 2.1, $`(A,\tau |A)`$ is a locally $`LC`$-space. $`\mathrm{}`$
###### Proposition 2.3
If $`(X,\tau )`$ has an open cover by locally LC-subspaces, then $`X`$ is a locally $`LC`$-space.
Proof. Let $`X=_{iI}G_i`$ be on open cover of $`X`$ where each $`G_i`$ is a locally $`LC`$-space, and let $`xX`$. Choose $`jI`$ such that $`xG_j`$. If $`U_j`$ is an open neighborhood of $`x`$ in $`G_j`$ such that $`U_j`$ is an $`LC`$-subspace of $`G_j`$, then $`U_j`$ is also open in $`(X,\tau )`$. By Proposition 2.1, $`(X,\tau )`$ is a locally $`LC`$-space. $`\mathrm{}`$
###### Corollary 2.4
Let $`(X_\alpha ,\tau _\alpha )_{\alpha \mathrm{\Omega }}`$ be a family of topological spaces. For the topological sum $`X=_{\alpha \mathrm{\Omega }}X_\alpha `$ the following conditions are equivalent:
(1) $`X`$ is a locally $`LC`$-space.
(2) Each $`X_\alpha `$ is a locally $`LC`$-space.
Proof. Follows from Proposition 2.2 and Proposition 2.3. $`\mathrm{}`$
###### Proposition 2.5
Let $`(X_i,\tau _i)_{iF}`$ be a finite family of Hausdorff spaces. If each $`X_i`$ is a locally $`LC`$-space, then the product space $`X=_{iF}X_i`$ is also a locally $`LC`$-space.
Proof. Let $`(X,\tau )`$ and $`(Y,\sigma )`$ be Hausdorff locally $`LC`$-spaces. Let $`(x,y)X\times Y`$. By assumption, there exists $`U\tau `$ and $`V\sigma `$ such that $`xU`$, $`yV`$ and both $`U`$ and $`V`$ are $`LC`$-subspaces of $`X`$ and $`Y`$, respectively. By \[3, Theorem 2\], $`U\times V`$ is an $`LC`$-subspace of the product space $`X\times Y`$. By Proposition 2.1, $`X\times Y`$ is a locally $`LC`$-space. $`\mathrm{}`$
###### Remark 2.6
We note that the Hausdorff condition can be reduced to the weaker separation property $`R_1`$. Recall that a space $`(X,\tau )`$ is called an $`R_1`$-space if $`x`$ and $`y`$ have disjoint neighborhoods whenever $`\mathrm{cl}\{x\}\mathrm{cl}\{y\}`$. Clearly, a space is Hausdorff if and only if it is $`T_1`$ and $`R_1`$.
Question. Does Proposition 2.5 remain true if we drop the requirement that the spaces in question have to be Hausdorff (or $`R_1`$)?
###### Proposition 2.7
Every locally hereditarily Lindelöf, locally $`LC`$-space is discrete.
Proof. Let $`(X,\tau )`$ be a locally hereditarily Lindelöf and a locally $`LC`$-space. We may assume that every point $`xX`$ has an open neighbourhood $`W`$ that is both hereditarily Lindelöf and an $`LC`$-space. But this means that $`W`$ is an open discrete subspace of $`(X,\tau )`$. Thus $`(X,\tau )`$ is discrete as well. $`\mathrm{}`$
###### Corollary 2.8
A locally $`LC`$-space is discrete if and only if it is locally finite. $`\mathrm{}`$
###### Proposition 2.9
Every locally $`LC`$-space is a $`T_1`$-space.
Proof. Assume that $`(X,\tau )`$ is a locally $`LC`$-space and let $`xX`$. For every $`yx`$, there exists $`U\tau `$ such that $`yU`$ and $`(U,\tau |U)`$ is an $`LC`$-space and hence a $`T_1`$-space. Clearly, $`yU\{x\}\tau `$ and $`xU`$. This shows that $`X`$ is $`T_1`$. $`\mathrm{}`$
###### Proposition 2.10
If every $`LC`$-subspace of every Lindelöf subset of a topological space $`(X,\tau )`$ is Lindelöf, then $`X`$ is a locally $`LC`$-space if and only if $`X`$ is an $`LC`$-space.
Proof. Assume that $`X`$ is a locally $`LC`$-space. Let $`AX`$ be Lindelöf and let $`xA`$. Since $`X`$ is an locally $`LC`$-space, there exists $`U\tau `$ such that $`xU`$ and $`(U,\tau |U)`$ is an $`LC`$-space. Since every subspace of an $`LC`$-space is an $`LC`$-space, $`UA`$ is an $`LC`$-space. By assumption, $`UA`$ is Lindelöf and hence closed in $`(U,\tau |U)`$. Thus, $`UA`$ is open in $`(X,\tau )`$, contains $`x`$ and is disjoint from $`A`$. This shows that $`A`$ is closed and consequently $`X`$ is an $`LC`$-space. $`\mathrm{}`$
###### Proposition 2.11
Open (and hence also closed) bijective images of locally $`LC`$-spaces are locally $`LC`$-spaces.
Proof. Let $`f:(X,\tau )(Y,\sigma )`$ be open and bijective and let $`(X,\tau )`$ be a locally $`LC`$-space. Let $`yY`$. Choose $`xX`$ such that $`f(x)=y`$. Since $`(X,\tau )`$ is a locally $`LC`$-space, then there exists $`U\tau `$ such that $`xU`$ and $`(U,\tau |U)`$ is an $`LC`$-space. Since $`f`$ is open, then $`f(U)`$ is an open neighborhood of $`y`$ in $`(Y,\sigma )`$. Since open, bijective images of $`LC`$-spaces are also $`LC`$-spaces, then $`(f(U),\sigma |f(U))`$ is an $`LC`$-subspace of $`(Y,\sigma )`$. By Proposition 2.1, $`(Y,\sigma )`$ is a locally $`LC`$-space. $`\mathrm{}`$
###### Corollary 2.12
The property ‘locally $`LC`$-space’ is a topological property. $`\mathrm{}`$
Department of Mathematics
PL 4, Yliopistonkatu 5
University of Helsinki
00014 Helsinki
Finland
e-mail: dontchev@cc.helsinki.fi
Department of Mathematics
Graz University of Technology
Steyrergasse 30
A-8010 Graz
Austria
e-mail: ganster@weyl.math.tu-graz.ac.at
Department of Mathematics
Faculty of Science
Haceteppe University
06532 Beytepe/Ankara
Turkey
e-mail: kanibir@eti.cc.hun.edu.tr
|
no-problem/9812/hep-lat9812017.html
|
ar5iv
|
text
|
# Hybrid Quarkonia on Asymmetric Lattices
## Abstract
We report on a study of heavy quark bound states containing an additional excitation of the gluonic degrees of freedom. To this end we employ the NRQCD approach on coarse and asymmetric lattices, where we discard vacuum polarisation effects and neglect all spin-correction terms. We find a clear hybrid signal on all our lattices ($`a_s=0.15\mathrm{}0.47`$ fm). We have studied in detail the lattice spacing artefacts, finite volume effects and mass dependence. Within the above approximations we predict the lowest lying hybrid excitation in Charmonium to be 1.323(13) GeV above the ground state, where we use the 1P-1S splitting to set the scale. The bottomonium hybrid was found to be 1.542(8) GeV above its ground state.
Gluonic excitations are ideal objects for investigating the nonperturbative nature of the gluonic degrees of freedom in QCD. Hybrid mesons can be thought of as hadronic bound states with an additional excitation of the gluon flux. This can give rise to states with non-conventional quantum numbers and has triggered an intense experimental and theoretical search for such particles. Previous predictions for the energies of hybrid states come from phenomenological models , static potential models and lattice simulations with propagating quarks . So far, lattice QCD is the only approach in which hybrids can be treated from first principles. However, the errors from such calculations on isotropic lattices are still much larger than for conventional states. This is because the correlation functions decay too rapidly when the excitation energy is very large with respect to the inverse lattice spacing. To obtain a similar signal-to-noise ratio one needs a much finer resolution in the temporal direction.
While in recent years there has been much progress in obtaining more reliable results from improved actions on spatially coarse lattices, it has also been demonstrated that anisotropic lattices can be employed to accommodate different physical scales on the same lattice. In particular the study of glueball states on coarse and anisotropic lattices has prompted us to study heavy hybrid states on such lattices in order to increase both the scope and the precision of a previous calculation significantly. There a non-relativistic approach (NRQCD) was used for the heavy $`b`$ quarks on an isotropic lattice with $`a0.08`$ fm. NRQCD has frequently been employed to allow high precision measurements for the $`b\overline{b}`$-system . Also the combination of improved gluon actions and the NRQCD approach for heavy quarks has already been used to determine the spectrum of conventional quarkonia . Previous attempts to measure heavy hybrid states on the lattice were reviewed in . An initial preliminary study using asymmetric lattices for Bottomonium hybrids was presented in .
In this paper we implemented such an efficient approach to study in detail the lattice artefacts and finite size effects for both the $`b\overline{b}g`$ and $`c\overline{c}g`$ hybrid. For the latter state we obtain excellent agreement with a relativistic simulation on isotropic lattices . As with this other study we have neglected dynamical sea quark effects, but we have succeeded in lowering the statistical error to about $`1\%`$.
In our study we generated the gauge field configurations using a tadpole-improved action which has been employed by different groups :
$`S=\beta \xi ^1{\displaystyle \underset{x,\mathrm{i}>\mathrm{j}}{}}\left\{{\displaystyle \frac{5}{3}}P_{\mathrm{ij}}{\displaystyle \frac{1}{12}}\left(R_{\mathrm{ij}}+R_{\mathrm{ij}}\right)\right\}\beta \xi {\displaystyle \underset{x,\mathrm{i}}{}}\left\{{\displaystyle \frac{4}{3}}P_{\mathrm{it}}{\displaystyle \frac{1}{12}}R_{\mathrm{it}}\right\},U_iU_i/u_s.`$ (1)
Here $`P_{\mathrm{ij}}`$ and $`R_{\mathrm{ij}}`$ denote the trace of the standard spatial plaquette and rectangle, respectively. Where the index $`t`$ appears the plaquette/rectangle extends only one link into the temporal direction. This theory has two parameters, $`\beta `$ and $`\xi `$, the second of which determines the asymmetry of our lattices. At tree level the “aspect ratio” is $`\xi =a_s/a_t`$, where $`a_s`$ and $`a_t`$ are the spatial and temporal lattice spacings, respectively. From we note that the radiative corrections to this relation are small when tadpole improvement is implemented, as described below.
The action in Equation 1 is designed to be accurate up to $`𝒪(a_s^4,a_t^2)`$ classically. To account for radiative corrections all spatial gauge links, $`U_i`$, are self-consistently tadpole-improved with $`u_s=0|(1/3)P_{\mathrm{ij}}|0^{1/4}`$, as suggested in . Such a mean-field treatment was demonstrated to reduce significantly the leading corrections, which are due to unphysical tadpoles in lattice perturbation theory. Our scaling analysis shows that errors $`𝒪(\alpha a_s^2)`$ are indeed negligible if the lattice spacing is sufficiently small. Since we will use this action only for small temporal lattice spacings we expect $`𝒪(\alpha a_t^2)`$ errors to be very small. Therefore we have not employed the tadpole improvement description for the temporal gauge links. Our results for two different aspect ratios $`\xi `$ justify such an assumption.
To propagate the heavy quarks through the lattice we expand the NRQCD Hamiltonian correct to $`𝒪(mv^2)`$, where all spin-dependent terms are absent. This accuracy was already employed in . The only additional improvement is to correct for temporal and spatial lattice spacing errors by adding two extra terms ($`c_7,c_8`$) to the evolution equation in :
$`H_0={\displaystyle \frac{\mathrm{\Delta }^2}{2m_b}},\delta H=c_7{\displaystyle \frac{a_t\mathrm{\Delta }^4}{16nm_b^2}}+c_8{\displaystyle \frac{a_s^2\mathrm{\Delta }^{(4)}}{24m_b}}.`$ (2)
In those operators all spatial links are also tadpole-improved using the same $`u_s`$ as for generating the configurations. After this modification we take the tree-level values for all the coefficients in the Hamiltonian. In this case, $`c_7=c_8=1`$.
For the non-relativistic meson operators we only used the simplest possible choices and have not tuned the overlap to optimise the signal. The gauge-invariant construction of S-state and P-state operators is described in . For the magnetic hybrid signal studied here, we have inserted the lattice version of the magnetic field into the $`Q\overline{Q}`$-state ($`B_i=ϵ_{ijk}\mathrm{\Delta }_j\mathrm{\Delta }_k`$):
$${}_{}{}^{1}H_{i}^{}(x)=\psi ^{}(x)B_i\chi (x),{}_{}{}^{3}H_{jk}^{}(x)=\psi ^{}(x)\sigma _jB_k\chi (x).$$
(3)
For the leading order in the NRQCD Hamiltonian, the operators in Equation 3 create a whole set of degenerate states: $`1^{},0^+,1^+,2^+`$. These are the spin-singlet and spin-triplet states with zero orbital angular momentum, including the exotic combination $`1^+`$. This degeneracy will be lifted when higher order relativistic corrections are re-introduced into the NRQCD Hamiltonian. We expect this to be a small effect, in as much as the heavy quarks are very slow in the shallow hybrid potential . By the same argument, we expect hybrid states with additional orbital angular momentum to be almost degenerate as it was observed in . The definition of the operators in Equation 3 has been augmented by a combination of fuzzing for the links and Jacobi-smearing for the quark fields . No effort has been made to optimise the signal further, but one could do so if even higher precision is needed, or if higher excited states are to be determined. To extract the hadron masses we simply fit the meson correlators to a single exponential;
$$C_\alpha (t)=H_\alpha ^{}(t)H_\alpha (0)=A_\alpha e^{m_\alpha t}.$$
(4)
We measured correlators every 10 trajectories in the Monte Carlo update, and for the error analysis we binned 50 such measurements into one. After this binning we still have an ensemble of 100–1000 configurations depending on the lattice and the state of interest. As it can be seen from the representative example of an effective massplots in Figure 1, the data is very good and the goodness of the single exponential fits is always bigger than $`Q=0.1`$, which we called acceptable. In Tables I and II we present our results and the simulation parameters.
To determine the lattice spacing $`a_t^1`$, we used the 1P-1S splitting in Charmonium and Bottomonium. As expected, the values from Charmonium are smaller than those from Bottomonium, because in the quenched approximation the coupling does not run as in full QCD. At $`(\beta ,\xi )=(2.4,5)`$ we observe a $`16\%`$ effect. In order to give another estimate of quenching errors we also determined the radial excitations, $`nS`$, and calculated the ratio $`R_{\mathrm{SP}}`$= 2S-1S/1P-1S, which can be compared with the experimental value of 1.28 for Bottomonium. At $`(\beta ,\xi )=(2.7,5)`$ we find $`R_{\mathrm{SP}}=1.424(89)`$. From these findings we quote quenching errors of $`1020\%`$. Several suggestions have been made how to measure the spatial lattice spacing , but to convert our results into dimensionful numbers we can use $`a_t^1`$ throughout.
We have also tested the velocity expansion and included relativistic corrections up to $`𝒪(mv^6)`$ for some of our lattices. At this level of accuracy we could not resolve any significant change in our results. A more detailed analysis of the spin structure is subject of a future project.
In this study we were mainly interested in the gluonic excitations of heavy quarkonia. For this purpose it was irrelevant to adjust the quark masses to their exact values. For some lattices we have changed the quark masses by 25% and did not find any noticeable change in the ratio $`R_\mathrm{H}`$ = (1H-1S)/(1P-1S).
Finite volume effects were a source of immediate concern for us. This is because hybrid states are expected to reside in a very flat potential . The bag model also suggests a very large bound state as the result of the gluonic excitation . As shown in Figure 2, we have found that for spatial extents of 1.2 fm or larger the masses of all Bottomonium and Charmonium states remain constant within small statistical errors. For even smaller volumes we can resolve a slight increase in the mass of the $`b\overline{b}g`$ hybrid. In other words, the spatial extent of the hybrid excitation seems to be almost independent of the heavy quark mass. In this sense the $`b\overline{b}g`$ hybrid is more difficult to calculate, as we need similar volumes but finer lattices than for Charmonium.
Finally, we carried out a scaling analysis to demonstrate that discretisation errors are under control. This is of utmost importance for the NRQCD approach since one cannot extrapolate to zero lattice spacing in this effective field theory and one has to model continuum behaviour already for finite lattice spacings. In Figure 3 we show the scaling of the hybrid excitation above the ground state. From this, one can see that we have found convincing scaling windows for both Bottomonium and Charmonium. Scaling violations can only be seen on the coarsest lattices ($`a_s>0.36`$ fm for $`c\overline{c}g`$ and $`a_s>0.19`$ fm for $`b\overline{b}g`$), but this is not totally unexpected - it is questionable how well our simple minded implementation of the tadpole prescription works to remove the $`𝒪(\alpha p^2a^2)`$ errors for heavy quark systems on such coarse lattices. In the case of $`b\overline{b}g`$ we also plot the results for two different aspect ratios. Both results are consistent and confirm the initial assumption of small temporal lattice spacing errors. To quote our final result for the lowest lying hybrid excitations we take the averaged value of all the results in the scaling region and employ the experimental values for the 1P-1S splitting to set the scale. We find 1.323(13) GeV for the case of Charmonium and 1.542(8) GeV for the first gluonic excitation in Bottomonium, in good agreement with a previous estimate of 1.68(10) GeV .
In conclusion, we have demonstrated the usefulness of coarse and anisotropic lattices for the nonperturbative study of gluonic excitations in heavy quark systems. Furthermore, this should also be considered a success of the NRQCD approach, which allowed us to predict the lowest lying Charmonium hybrid state at the same mass as from a relativistic calculation . Apart from the very accurate predictions for hybrid quarkonia, it is also interesting to notice that all of the above results could be obtained in a comparatively short period of time. Whereas our present calculation confirms that the $`b\overline{b}g`$ hybrid will lie above the S+S threshold for decay into B-mesons, the issue whether it may be found below the S+P threshold has to be decided in a simulation where dynamical sea quarks are included in order to control this last remaining systematic error.
TM would like to thank R.R. Horgan and I.T. Drummond for many useful and constructive discussions. The calculations were done using workstations and the CP-PACS facilities at at the Center for Computational Physics at the University of Tsukuba. This work is supported in part by the Grants-in-Aid of Ministry of Education (No. 09304029). TM, HPS and AAK are supported by the JSPS Research for Future Program, and SE and KN are JSPS Research Fellows.
|
no-problem/9812/cond-mat9812171.html
|
ar5iv
|
text
|
# Effect of Ferromagnetic Spin Correlations on Superconductivity in Ferromagnetic Metals
\[
## Abstract
We study the renormalization of the quasiparticle properties in weak ferromagnetic metals, due to spin fluctuations, away from the quantum critical point for small magnetic moment. We explain the origin of the $`s`$-wave superconducting instability in the ferromagnetic phase and find that the vertex corrections are small and Migdal’s theorem is satisfied away from the quantum critical point.
\] More than thirty years ago Doniach and Berk and Schrieffer showed that, in the paramagnetic phase, the phonon-induced, $`s`$-wave superconductivity in exchange-enhanced transition metals is suppressed by ferromagnetic spin fluctuations, in the neighborhood of the Curie temperature. At the same time a theory of superconductivity coexisting with long-range ferromagnetic order was developed by Larkin and Ovchinnikov , and by Fulde and Ferrell for magnetic-impurity-induced ferromagnetism in metals. Without experimental evidence for the coexistence of superconductivity and ferromagnetism, this theory has been only of academic interest. It is generally accepted that ferromagnetism suppresses superconductivity and the apparent contradiction between the above two pictures has not been clarified.
On the experimental side, recent advances has allowed for the investigation of the quantum critical region in correlation-induced, weak ferromagnetic metals as well as in some heavy-fermion compounds. When hydrostatic pressure is applied on a transition metal compound such as MnSi or ZrZn<sub>2</sub>, the Curie temperature can be driven down to zero at a critical pressure. In the neighborhood of this critical pressure the paramagnetic-ferromagnetic phase transition is driven by quantum critical fluctuations. So far, experiments have failed to find superconductivity in the paramagnetic phase of these compounds and as we argue below, the physics close to the phase transition is not well understood.
These experiments have motivated us to study the ferromagnetic regime relatively close to the critical point which is described as a highly correlated but weakly ferromagnetic metal. In this investigation we extend the Doniach, Berk, Schrieffer (DBS) theory into the ferromagnetic phase of the transition metal compounds. We develop a microscopic theory of the ferromagnetic state, based on the interactions mediated by spin-fluctuations between the fermions, and explain the microscopic origin of the unexpected $`s`$-wave superconductivity recently predicted by us on the basis of phenomenological considerations. In doing so, we find that, in contrast to the paramagnetic case, ferromagnetic fluctuations enhance pairing correlations – resolving the longstanding dilemma referred to in the opening paragraph.
Our starting point is the Stoner state which is a Hartree-Fock solution of some Hamiltonian, below the mean-field ferromagnetic instability. This state is a product of two Slater determinants with an electron mass possibly renormalized by the band structure. At this level of approximation the correlations do not renormalize the mass. Although the Stoner state has non-zero magnetization, it is known that the Hartree-Fock approximation overestimates the exchange. However, we will assume that fluctuations about this mean-field saddle point do not completely destroy the ferromagnetic order. We cannot over-emphasize that this starting point is not perturbatively connected to the paramagnetic Fermi-Liquid state. The next step is to include the correlations which produce weakly interacting quasiparticles with renormalized mass $`m^{}`$ near the Fermi surface. The renormalization in the neighborhood of the two Fermi surfaces is described by the single particle Green’s function
$$G_\sigma (\stackrel{}{p},\omega )=\frac{z}{\omega v_F(|\stackrel{}{p}|p_\sigma )+i\delta _\sigma [\stackrel{}{p}]}+G_{inc}$$
(1)
which is diagonal with the quantization axis parallel to the $`z`$-axis. Here $`\stackrel{}{p}`$ is the three-dimensional momentum of the particle, $`p_\sigma `$ ($`\sigma =`$,$``$) is the Fermi momentum of the spin-$`\sigma `$ electrons, $`v_F`$ is the Fermi velocity, and $`\delta _\sigma [\stackrel{}{p}]\delta \times \text{sign}(|\stackrel{}{p}|p_\sigma ),`$ with $`\delta `$ an infinitesimal real number. The quasiparticle properties are hidden in the quasiparticle residue $`z<1`$ and the effective mass. In principle the Fermi velocity and the quasiparticle residue also depend on the spin index, but in the neighborhood of the phase transition, $`p_{}p_{}p_{}`$,$`p_{}`$ and they are equal. In the case of weak ferromagnetic metals the incoherent part of the Green’s function describing the physics away from the Fermi surface is a smooth function of $`\stackrel{}{p}`$ and $`\omega `$ which renormalizes the properties at the Fermi surface and introduces no new physics. The Green’s function describes a system of quasiparticles with spontaneous magnetization given by Dzyaloshinskii’s theorem
$$m_0=\frac{1}{12\pi ^2}(p_{}^3p_{}^3)=\frac{n_{}n_{}}{2}\text{,}$$
(2)
where $`m_0`$ is the uniform, static magnetization and $`n_\sigma `$ is the density of spin-$`\sigma `$ particles. For weak ferromagnets the magnetization $`m_0`$ is proportional to the exchange splitting $`\mathrm{\Delta }=p_{}p_{}`$. Here we have assumed that all particles are in an eigen-state of the $`z`$ component of the spin operator and for definiteness we will assume that $`p_{}>p_{}`$. The low-energy excitations of the system described by this Green’s function are quasiparticle excitations in the neighborhood of the two Fermi surfaces as well as collective spin excitations. The spontaneously broken $`SU(2)`$ symmetry guarantees the existence of a massless Goldstone mode described by the propagator
$$D_G(\stackrel{}{q},\omega )=\frac{\mathrm{\Delta }N(0)v_F}{2}\frac{\omega _s(\stackrel{}{q})}{(\omega +i\delta )^2\omega _s^2(\stackrel{}{q})}\text{,}$$
(3)
$`N(0)`$ is the density of states at the Fermi surface. In the case of a ferromagnetic metal the magnetization is a conserved quantity and the spin-wave dispersion is $`\omega _s(\stackrel{}{q})=D\left|\stackrel{}{q}\right|^2`$ where $`D=v_F\mathrm{\Delta }/p_F^2`$ is the spin stiffness. The longitudinal response of the system is described by the propagator
$$D_l(\stackrel{}{q},\omega )=\frac{N(0)p_F^2}{2}\frac{1}{\xi ^2+\left|\stackrel{}{q}\right|^2i\pi p_F^2\omega /2v_F\left|\stackrel{}{q}\right|},$$
(4)
where $`\xi m_0^1`$ is the correlation length. The interaction of the quasiparticles with these collective spin excitations can be described by the interaction
$$H_{\text{sf}}=g_0\underset{\stackrel{}{k}\stackrel{}{q}\alpha \beta }{}c_{\stackrel{}{k}\alpha }^{}\stackrel{}{\sigma }_{\alpha \beta }c_{\stackrel{}{k}+\stackrel{}{q}\beta }\stackrel{}{S}_\stackrel{}{q}\text{,}$$
(5)
where $`g_0`$ is the bare momentum-independent coupling constant, $`c_{\stackrel{}{k}\alpha }^{}`$ and $`c_{\stackrel{}{k}\alpha }`$ are the anticommuting, quasiparticle creation and annihilation operators respectively, $`\stackrel{}{\sigma }_{\alpha \beta }`$ are the Pauli matrices and $`\stackrel{}{S}_\stackrel{}{q}=<_{\stackrel{}{p}\gamma \delta }c_{\stackrel{}{p}\gamma }^{}\stackrel{}{\sigma }_{\gamma \delta }c_{\stackrel{}{p}\stackrel{}{q}\delta }>`$ is the three component spin fluctuation field. The vector field $`\stackrel{}{S}_\stackrel{}{q}`$ is the average magnetization at a particular wave vector. In the ferromagnetic phase this average is different from zero, while in the paramagnetic phase it is strictly zero. Nevertheless it has been used to describe magnetically enhanced paramagnetic metals, although it can be mathematically justified only in the ferromagnetic phase.
Recently we have shown, ignoring the vertex corrections, that the self energy leading to the exact Green’s function, Eq. (2), is local and leads to logarithmic dependence of the quasiparticle residue on the magnetization. When the magnetization approaches the quantum critical point the quasiparticle residue vanishes and the Fermi liquid theory breaks down. At finite temperatures in the neighborhood of the Curie temperature the spin fluctuations lead to a non-Fermi liquid specific heat $`C/T\mathrm{ln}T`$ consistent with recent experiments on MnSi and ZrZn<sub>2</sub> as well as on some of the heavy-fermion compounds .
Weak ferromagnetic metals are very interesting because the gapless Goldstone mode coexists with the longitudinal excitations which are gaped. The longitudinal spin-fluctuation propagator Eq. (4) is similar to the model susceptibility (peaked at the $`\stackrel{}{Q}=(\pi ,\pi )`$ nesting vector) used in the theory of antiferromagnetic metals. However, in our case the expression for the susceptibility is rigorous, following from the poles of the 4-point vertex at small momentum transfer.
The first vertex corrections to the 3-point vertex in the weak ferromagnetic metal, Fig. (1), are
$`\mathrm{\Lambda }_{}^{(1)}(p,p+k)=\mathrm{\Lambda }_l^{(1)}+\mathrm{\Lambda }_G^{(1)}\text{,}`$ (6)
$`\mathrm{\Lambda }_{}^{(1)}(p,p+k)=\mathrm{\Lambda }_l^{(1)}+\mathrm{\Lambda }_G^{(1)}\text{,}`$ (7)
where
$`\mathrm{\Lambda }_l^{(1)}`$ $`=`$ $`ig_0^2{\displaystyle 𝑑qG_{}(q)D_l(qp)G_{}(q+k)\text{,}}`$ (8)
$`\mathrm{\Lambda }_G^{(1)}`$ $`=`$ $`ig_0^2{\displaystyle 𝑑qG_{}(q)D_G(qp)G_{}(q+k)\text{,}}`$ (9)
$`\mathrm{\Lambda }_l^{(1)}`$ $`=`$ $`ig_0^2{\displaystyle 𝑑qG_{}(q)D_l(qp)G_{}(q+k)\text{,}}`$ (10)
$`\mathrm{\Lambda }_G^{(1)}`$ $`=`$ $`ig_0^2{\displaystyle 𝑑qG_{}(q)D_G(qp)G_{}(q+k)\text{,}}`$ (11)
$`dq`$ $``$ $`{\displaystyle \frac{d^4q}{(2\pi )^4}}\text{.}`$ (12)
Here we have assumed an expansion of the full vertex
$$z\mathrm{\Lambda }_{\alpha \beta }(p,p+k)=1+\mathrm{\Lambda }_{\alpha \beta }^{(1)}+\mathrm{}\text{ ,}$$
(13)
and $`p`$, $`q`$, and $`k`$ are 4-vectors.
It is important to distinguish the order of the small momentum-transfer and energy-transfer limits. In the limit which defines the Fermi liquid parameters through the 4-point vertex, the Ward identity
$$\underset{\omega 0}{lim}\underset{\stackrel{}{q}0}{lim}\mathrm{\Lambda }_{\alpha \beta }(p,p+q)=\left(1\frac{\mathrm{\Sigma }}{\omega }\right)\delta _{\alpha \beta }=\frac{1}{z}\delta _{\alpha \beta }\text{,}$$
(14)
shows that the vertex is proportional to the inverse quasiparticle residue. The effective pairing potential in principle can be constructed from the 3-point vertex with the requirement that the triplet scattering amplitude is zero. In second order perturbation theory, however, the momentum independence of the self-energy and the vanishing of the triplet scattering amplitude are incompatible and so far we have not been able to construct a pairing potential with the above properties. Nevertheless, one can see that the singlet scattering amplitude is attractive leading to a pairing instability in the singlet channel. Physically, the Pauli exclusion principle keeps quasiparticles with the same spin apart, leading to a negative charge depletion between them. This charge distribution attracts another quasiparticle with the opposite spin leading to the singlet pairing.
The Ward identity which we mentioned earlier shows that the effective pairing is enhanced for small magnetizations since $`z^1\mathrm{ln}m_0`$ and this enhancement is due to the longitudinal collective mode.
In the physical limit where energy is conserved, the corresponding Ward identity is
$$\underset{\stackrel{}{q}0}{lim}\underset{\omega 0}{lim}\mathrm{\Lambda }_{\alpha \beta }(p,p+q)=\frac{v_F}{z}\frac{dp_\alpha }{d\mu }\sigma _{\alpha \beta }^z$$
(15)
where, $`\sigma _{\alpha \beta }^z`$ is the Pauli matrix, $`v_F^0`$ is the Fermi velocity of the noninteracting Fermi gas, and there is no summation over repeated indexes.
In calculating the vertex corrections we first set the frequency to zero and then take the limit for the momentum. Because, we are working in the broken symmetry phase a distinction must be made for vertex corrections involving particles on one of the two Fermi surfaces and vertex corrections involving particles on different Fermi surfaces. In the former case the limit
$$\mathrm{\Lambda }_{\sigma \sigma }(|\stackrel{}{p}|p_\sigma ,|\stackrel{}{p}|p_\sigma )$$
(16)
while in the latter the limit
$$\mathrm{\Lambda }_{\sigma \sigma ^{}}(|\stackrel{}{p}|p_\sigma ,|\stackrel{}{p}|p_\sigma ^{}+\mathrm{\Delta })$$
(17)
must be taken. In both cases we use the spectral representation for the propagators $`D_{l/G}`$
$$D_{l/G}(\stackrel{}{q},\omega )=\frac{2}{\pi }_0^{\mathrm{}}\frac{zImD_{l/G}(\stackrel{}{q},z)}{z^2\omega ^2i\delta }$$
(18)
Using that
$`G_\sigma (\stackrel{}{p}+\stackrel{}{q},ϵ+\omega )G_\sigma ^{}(\stackrel{}{p}+\stackrel{}{q}+\stackrel{}{k},ϵ+\omega )=`$ (19)
$`{\displaystyle \frac{z}{v_F}}{\displaystyle \frac{G_\sigma (\stackrel{}{p}+\stackrel{}{q},ϵ+\omega )G_\sigma ^{}(\stackrel{}{p}+\stackrel{}{q}+\stackrel{}{k},ϵ+\omega )}{|\stackrel{}{p}+\stackrel{}{q}||\stackrel{}{p}+\stackrel{}{q}+\stackrel{}{k}|(p_\sigma p_\sigma ^{})}}`$ (20)
it is not difficult to obtain the expansion
$`z\mathrm{\Lambda }_{\sigma \sigma ;l}(p_\sigma ,p_\sigma )=`$ (21)
$`1+{\displaystyle \frac{g_0^2N^2(0)z^2}{16p_\sigma }}\mathrm{ln}{\displaystyle \frac{\pi ^2p_\sigma ^4+4\mathrm{\Delta }^4}{\pi ^2p_\sigma ^4+4(\mathrm{\Delta }^2+p_c^2)^2}}+\mathrm{}`$ (22)
and
$$z\mathrm{\Lambda }_{\sigma \sigma ;G}(p_\sigma ,p_\sigma )=1+\frac{g_0^2N^2(0)z^2}{4}ln(1+\frac{\mathrm{\Delta }p_c}{p_F^2})+\mathrm{}.$$
(23)
We have used a momentum cutoff $`p_c`$ reflecting the different physics at very small distances. Very similar logarithmic behavior can be seen in the vertex expansion of $`\mathrm{\Lambda }_{\sigma \sigma ^{};l/G}(p_\sigma ,p_\sigma ^{}+\mathrm{\Delta })`$. This implies that the self-energy is weakly momentum dependent close to the phase transition. Therefore a local ferromagnetic Fermi liquid theory can be used to describe weak ferromagnetic metals in a regime where the magnetization is sufficiently small, but away from criticality (since the fluctuations in the critical regime are beyond the scope of Fermi liquid theory). This confirms the $`s`$-wave pairing instability in the ferromagnetic phase.
The above adiabaticity is a consequence of the smallness of the exchange splitting $`\mathrm{\Delta }`$ compared to the Fermi momentum $`p_F(p_{}+p_{})/2`$ and the smallness of the maximum spin wave velocity $`\omega _G`$ compared to the Fermi energy $`ϵ_F`$ and leads to the validity of Migdal’s theorem for weak ferromagnetic metals.
In the DBS theory of spin-fluctuation-enhanced paramagnetic metals it is argued that the sharply peaked static spin susceptibility
$$\chi (0,0)=\frac{\chi _p}{1N(0)V_c}$$
(24)
close to the Curie point suppresses the $`s`$-wave pairing, because ferromagnetic spin fluctuations act as an effective repulsive force between electrons with opposite spins. Here the $`\chi _p`$ is the Pauli susceptibility and $`V_c`$ is a pseudopotential. To see why the spin-fluctuations have opposite effect in the ferromagnetic phase it is convenient to write the factor $`1N(0)V_c`$ in the denominator of the spin susceptibility in terms of the Landau Fermi liquid parameter $`F_0^a`$. Then in both paramagnetic and ferromagnetic phases the static spin susceptibility is positive and in the ferromagnetic phase is
$$\chi (0,0)=\frac{N(0)/2}{\left|1+F_0^a\right|},$$
(25)
while in the paramagnetic phase is
$$\chi (0,0)=\frac{N(0)}{1+F_0^a},$$
(26)
In approaching the quantum critical point from the paramagnetic side $`F_0^a`$ approaches the value $`1`$ from above, while approaching from the ferromagnetic side $`F_0^a`$ approaches $`1`$ from below. The different sign of $`1N(0)V_c1+F_0^a`$ has dramatic effect on the sign of the spin-fluctuation mediated quasiparticle interaction on the singlet channel which can be seen in the $`t`$-matrix
$$t(0,0)=V_c+\frac{N^1(0)(F_0^a)^2}{1+F_0^a}\text{.}$$
(27)
In the paramagnetic phase the second term is positive, while in the ferromagnetic it is negative leading to $`s`$-wave pairing.
What is the physics in the neighborhood of the quantum critical point in weak ferromagnetic metals is still an open question. Another interesting point is that the BCS theory of superconductivity can not give a quantum critical point, because as the critical temperature approaches zero so does the pairing interaction. Whether a different type of superconductivity exists or a different phase exists in the neighborhood of the quantum critical point on the paramagnetic side of the phase diagram is still an open question.
In Fig.(1) we represent a schematic phase diagram of a “typical” weak ferromagnetic metal. Because, the energy scale $`T^{}=T_c^2/ϵ_F`$ below which the superconducting instability occurs vanishes the $`s`$-wave superconducting state must also vanish at the quantum phase transition. On the paramagnetic side the $`p`$-wave supercondicting state is expected as predicted by the DBS theory. At finite temperatures close to the ferromagnetic phase transition on the ferromagnetic side the spin fluctuations renormalize the physical quantities leading to a non-Fermi liquid specific heat $`CTlnT`$ and in Fig.(1) we have shown the crossover between the Fermi liquid and the non-Fermi liquid state. We also expect the $`p`$-wave superconducting state in the paramagnetic phase to vanish at the quantum critical point. Another possibility is that it remains finite as we go through the phase transition at finite Curie temperature. However, the superconducting transition temperature must be less than the Fermi liquid scale set by $`T^{}`$ which vanish at the quantum phase transition and this implies the vanishing of the $`p`$-wave paired state.
The $`s`$-wave superconducting state in the ferromagnetic phase is unusual and is a generalization of the Larkin-Ovchinnikov-Fulde-Ferrell (LOFF) state studied in the 60’s in metals with magnetic impurities. The difference between this generalized LOFF state and the one originally studied is that the magnetic moment in the former is caused by the quasiparticles which also participate in the pairing, while in the latter the magnetic field is external to the quasiparticle system. Therefore the response of the two systems to an external magnetic field must be quite different. The understanding of how the spin fluctuations are modified by the superfluid density is an interesting question which can shed light on the nature of this state. Another interesting possibility is that this state has an odd-gap close to half filling induced by the presence of magnon excitations. The details of this state are beyond the scope of the current paper and will be investigated in a future publication.
In conclusion, in this paper we described the physics of a weak ferromagnetic metal from microscopic principles. We have shown that the vertex corrections in the physical limit are small and that the self-energy is local. In the limit of small momentum transfer the vertex function enhances the effective coupling between the quasiparticles in the neighborhood of the quantum phase transition leading to an $`s`$-wave superconducting instability.
We would like to thank J.L. Smith, A. Balatsky, P. Coleman, G. Kotliar, A.E. Ruckenstein, L.B. Ioffe, G.G. Lonzarich, R.B. Laughlin, and W. Mao for the stimulating discussions on this subject. This work was sponsored by the DOE Grant DEFG0297ER45636.
|
no-problem/9812/quant-ph9812072.html
|
ar5iv
|
text
|
# EPR Correlations as an Angular Hanbury-Brown—Twiss Effect
\[
## Abstract
It is shown that EPR correlations are the angular analogue to the Hanbury-Brown—Twiss effect. As insight provided by this model, it is seen that, the analysis of the EPR experiment requires conditional probabilities which do not admit the derivation of Bell inequalities.
> 3.65.Bz, 01.70.+w
\]
Bell’s Theorems purport to prove that an objective, local hidden-variable extension of Quantum Mechanics (QM) is impossible. This result has been called “beautiful” and the century’s most significant discovery. For some, however, this result is a symptom of error or misunderstanding.
Of course, a theorem does not establish a universal, unrestricted truth; it only tests symbolic manipulations, that is mathematics, for consistency, given an hypothesis. A search for error in Bell’s analysis, therefore, is nothing but a critical review of its hypothesis. The obdurate realist, who wishes to challenge Bell’s conclusion, has only two options: QM must be wrong (perhaps incomplete or otherwise defective on the margins) or, the hypothesis contains error.
Within QM, all that is needed to obtain the expressions relevant to the Einstein, Rosen and Podolsky (EPR) experiment, as modified by Bohm (EPRB), at the heart of Bell’s analysis, is a superposition state; the rest follows from simple geometric transformations. Indeed, exactly this feature of QM has been questioned, starting with Furry. He suggested that for macroscopic distances, a superposition state converts to a mixed state.
The second option, seeking error in the hypothesis of what should constitute an objective local extension of QM, is likewise lean on possibilities. The hypothesis Bell used was scarcely more than the assertion that the coincidence intensity for the EPRB experiment is to be given by:
$`P(a,b)={\displaystyle I_𝒜(a,\lambda )I_{}(b,\lambda )\rho (\lambda )𝑑\lambda },`$ (1)
where notation and content are taken from Bell with the modification that $`I_A`$ stands for the count rate, or photoemission probability, at measuring station $`A`$, etc. For ideal photodetectors, this count rate is proportional to the impinging field intensity; i.e., to the square of the field strength.
It is the purpose here to analyze just these assumptions, in particular the second, and to show that in fact application of the principles underlying the Hanbury-Brown—Twiss Effect, permits an objective, local interpretation of the EPR correlations.
The application of Furry’s proposal to the EPRB experiment, proceeds as follows. It is assumed that the source emits classical electromagnetic radiation polarized in a particular but random direction. It is taken that in each arm of the setup, this radiation is to be directed through a polarizer and then detected using a photodetector which obeys the square law; i.e., it emits photoelectrons in proportion to the square of the intensity of the absorbed radiation. That is, the probability of emission of a photoelectron in each arm of an EPR experiment is $`(\mathrm{cos}(\theta ))^2`$ where $`\theta `$ is the angle between the polarization direction of the signal and the axis of the polarizer used in the detector. A coincidence detection is then taken to be proportional to the product of detection probabilities in each channel, $`\mathrm{cos}^2(\theta )\mathrm{cos}^2(\theta \varphi )`$, where $`\varphi `$ is the angle between the axes of the measurement polarizers if the coordinate system is aligned with one of them. That this product gives a coincidence probability is based on the proposition that the probability of coincidence of local and therefore statistically independent events is the product of the individual probabilities. In the notation of Equation 1, $`I_𝒜=\mathrm{cos}^2(\theta ),I_{}=\mathrm{cos}^2(\theta \varphi )`$and$`\rho (\lambda )=d\lambda /2\pi `$.
Finally, the total coincidence rate is obtained by averaging over many pairs of signals, each with its own randomly given polarization angle $`\theta `$, that is
$`{\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}[\mathrm{cos}(\theta )\mathrm{cos}(\theta \varphi )]^2𝑑\theta =1/4+1/8cos(2\varphi ).`$ (2)
To convert this intensity to a probability it must be recast as the ratio of a coincidence rate divided by the total count rate. For ideal detectors, the total number of detections is linearly proportional to the sum of the field intensities at both detectors, that is: $`2(_0^{2\pi }I(\theta )𝑑\theta )=1`$. This model was examined as a semiclassical EPR variant in Ref. .
This expression seems perfectly rational and, as the resulting correlation is $`cos(2\varphi )/2`$, it does not violate a Bell Inequality. It would resolve the conundrums evoked by Bell’s Theorems were it to agree with experiment. However, this result has a nonzero minimum, whereas the QM equivalent, $`\mathrm{cos}^2(\varphi )/2`$, does go to zero and this difference has been observed and reported in Ref. . Eq. (2) does not conform to Nature.
If Furry’s Ansatz is to benefit a realist program, it must therefore be modified in some essential. The search need not be carried far. The radiation in an EPR experiment emanates from a single source (which may comprise many microunits, atoms say) and is by assumption such that the twin emissions pairwise are not to carry off angular momentum. In the case of radiation, this effectively means that if one is polarized so as to pass a polarizer in any particular direction, the other must be blocked by a polarizer in this direction. Obviously, such emissions are not statistically independent in each arm. In turn, without statistical independence, the assumption of factorizability of the joint probability as employed in writing Eq. (1), is not admissible. Factorization is overly restrictive and not valid. Quite reasonably so, as the question is: given that a particular result is obtained in one arm, – \- absolute simultaneity is impossible — what are the probabilities of outcomes in the second arm? This calculation demands conditional probabilities and they are not factorizable.
This in no way, however, implies nonlocality; rather, it implies just statistical dependence; i.e., the probabilities of ‘Bertlmann’s socks.’ Nonlocality, taken as a violation of Einstein’s principle that all influences effective at a particular event (point) in Minkowski space, must originate at points in the past light cone of that event, is not violated. The correlation resulting from most forms (a realist holds: all forms) of statistical dependence is simply derived from a common cause. Of course, as factorizability always holds for the coincidence probability of statistically independent events, it inevitably implies no violation of locality. Indeed, such events have no cause-effect relationship. In summary, factorizable coincidences map onto but are not one-to-one on the set of all coincidence functions for events respecting locality.
Thus, as an alternate to the Furry inspired model described above, consider the following:
The source is assumed to emit in the $`\pm \widehat{z}`$ direction circularly polarized signals; clockwise in one direction and counterclockwise in the other. Thus, the signal impinging on photodetector A, say, is:
$`E_A(\theta )=\widehat{x}cos(\theta )+e^{i\pi /4}\widehat{y}sin(\theta ),`$ (3)
where factors of the form $`exp(kx\omega t)`$ are supressed, $`\widehat{x},\widehat{y}`$ are orthogonal unit vectors, the factors $`cos(\theta ),sin(\theta )`$ project the individual components of the circularly polarized signal onto the axis of the polarizer and the factor, $`exp(i\pi /4),`$ represents the fixed phase difference between the orthogonal components which give circular polarization. Likewise, the signal impinging on the photodetector B, oriented at angle $`\varphi `$ with respect to A, as expressed in $`A`$’s coordinate system, is:
$`E_B(\theta ,\varphi )=\widehat{y}cos(\theta \varphi )e^{i3\pi /4}\widehat{x}sin(\theta \varphi ).`$ (4)
Use is made now of a generalized coincidence probability inspired by second order coherence theory:
$`P(a,b)={\displaystyle \frac{E_AE_BE_BE_A}{|E_A|^2+|E_B|^2}},`$ (5)
where the angle brackets indicate an ensemble average over all values of $`\theta `$, the angle of attack of each separate signal, or, on an ergotic principle, over the random phases of the individual atomic sources. The dot product is with respect to the orthogonal set $`\{\widehat{x},\widehat{y}\}`$.
The numerator in Equation 5 is the probability of a coincidence count; as usual for ideal detectors, it is the product of the intensities of the separate signals, but in the form taught by coherence theory. Traditionally, intensity correlation calculations were based on the direct product of intensities, $`I_1I_2`$, whereas coherence theory teaches that the correct form for this calculation is $`E_1E_2E_2E_1`$. The effective difference is that the later form allows the phase to contribute to calculation. It is the information in the phase that is required to explain the Hanbury-Brown—Twiss effect as well as other coherence phenomena.
Note that all the information used in the calculation of the numerator of Equation 5 is propagated to stations $`A`$ and $`B`$ from events in the past light cones of these events. The signals arriving at the measurement stations are, in this case, just classical electromagnetic signals for which there is no question of a violation of locality. Here it is seen clearly that factorizability is not a valid encodification of Einsteinian locality.
The denominator in Equation 5 is equal to the total intensity of both signals in both detectors and is, therefore, proportional to the total photoelectron count, again, for ideal detectors. The ratio of the numerator to the denominator then is by definition the probability of coincidence counts.
Taking all the above into account, provides the following expression for the coincidence count rate:
$`P(\widehat{𝐚},\widehat{𝐛})=`$ (6)
$`{\displaystyle \frac{_0^{2\pi }(cos(\theta )sin(\theta \varphi )sin(\theta )cos(\theta \varphi ))^2𝑑\theta }{2_0^{2\pi }(cos^2(\theta )+sin^2(\theta ))𝑑\theta }}.`$ (7)
Evaluated, this integral equals the QM result, Eq. (3):
$`P(\widehat{𝐚},\widehat{𝐛})={\displaystyle \frac{1}{2}}\mathrm{sin}^2(\varphi ).`$ (8)
This model, comprising non quantum components, is fully local in the Einsteinian sense; and, as it agrees with QM, it is in accord with those laboratory observations verifying QM. In essence it is, given the vector character of electromagnetic radiation, just the angular analogue of the Hanbury-Brown—Twiss Effect. It stands as a counterexample to Bell’s conclusion.
Like QM, however, it violates Bell Inequalities. Such inequalities, however, are derived under the assumption that the relevant coincidence probabilities factor into two terms, which a second order coherence function does not in general allow. That is: Bell inequalities are not valid for all forms of fully local coincidences.
This can also be shown as follows. As a matter of fact, the most general form for the coincidence count between stations $`A`$ and $`B`$ in the EPRB experiment is a function of three sets of variables: $`P(𝐚,𝐛,\lambda )`$, where $`𝐚,𝐛`$ are those variables that specify conditions at the measuring stations $`A`$ and $`B`$, and $`\lambda `$ specifies all common causes pertaining to the generation of the two signals. The $`\lambda `$, not being explicit in QM, have been denoted “hidden variables.” The coincidence count considered in QM is the marginal probability derived from the full coincidence probability by integration over $`\lambda `$:
$`P(𝐚,𝐛)={\displaystyle P(𝐚,𝐛,\lambda )𝑑\lambda }.`$ (9)
Now, the identity from probability theory:
$`P(𝐚,𝐛,\lambda )=P(\lambda )P(𝐚|\lambda )P(𝐛|𝐚,\lambda ),`$ (10)
where $`P(x|y)`$ is the conditional probability of $`x`$ contingent on $`y`$, exposes the intrinsic structure of such a coincidence.
This form reduces to that of the integrand of Equation 1 when $`P(𝐛|𝐚,\lambda )=P(𝐛|\lambda )`$; i.e., when the events at $`A`$ and $`B`$ are statistically independent; in other words, when there is no relationship between them. This is fundamentally contrary to the structure of the EPRB experiment in which it is taken that the emissions are correlated. It is easy to verify that no derivation of a Bell inequality goes through using Equation 10. Thus, such inequalities do not pertain to correlated events.
Equation 10 does not imply that information is telegraphed from station $`A`$ to station $`B`$. It means only that the counts registered at both stations will exhibit correlations that will become evident when the data is brought together at a later time for comparison. Such a comparison can be made, naturally, only at a point in Minkowski space for which the the past light cone includes the measuring stations $`A`$ and $`B`$. Likewise, the correlations did not arise with the help of superluminal, or any other, communication. The structure yielding the correlations when the measuring stations are specified by $`𝐚,𝐛`$, is built into these signals at their source which is in the past light cones of both stations. $`P(𝐛|𝐚,\lambda )`$ being contingent on a is a realization not of communication between stations $`A`$ and $`B`$, but of correlations invested in the signals at the common source. For the EPRB experiment, clearly, there can be no coincident count when the polarizers are parallel regardless of the orientation of the signals (so long as they are orthogonal, as assumed in the first place). Thus, the dependence of the conditional probability is the consequence of the necessity of the detectors to be set so as to admit detection of the correlated characteristics of the signals, here orthogonal polarization.
Of additional interest is the fact that the new model moves the nonfactorizable structure from a superposition wave function to the form of the coincidence probability. This affords considerable simplification of discussions on the interpretation of QM. Superposition wave functions have been the source of much confusion, requiring as they do, “collapse” for ontological meaning.
In conclusion, using the correct classical-physics method to calculate a coincidence count in the EPRB experiment, yields the QM result and exposes an inappropriate assumption in the derivation of Bell inequalities. No error has been found in QM, rather, just the argument against an objective local extension of QM has been put aside. This is at no cost to any established theoretical or empirical result from QM. The fact that experiments to test Bell inequalities have virtually beyond all argument supported QM, do not by themselves imply that QM is nonlocal. They prove no more than that inequalities that should obtain for objective local extensions of QM, but were derived under a false premise in any case, are not valid. Indeed, we see, they can not be.
|
no-problem/9812/astro-ph9812454.html
|
ar5iv
|
text
|
# THE SUBMILLIMETER FRONTIER: A SPACE SCIENCE IMPERATIVE
## Abstract
A major goal of modern astrophysics is to understand the processes by which the universe evolved from its initial simplicity, as seen in measurements of the Cosmic Microwave Background (CMB), to the universe we see today, with complexity on all scales. The initial collapse of the subtle seeds seen in the CMB results in the formation of galaxies and stars. The formation of the first stars marks the beginning of heavy element nucleosynthesis in the universe, which has a profound effect on the formation of subsequent generations of stars. The density fluctuations in the CMB from which all this structure grows are primordial. The development of these fluctuations into progressively more complex systems is the history of the universe; galaxies from seed structures, clouds from uniform interstellar gas, stars from clouds, elements from the initial hydrogen and helium, molecules from elements, dust from molecules, and planets from dust. These processes have resulted, at least on Earth, in the remarkable range of physical, chemical, and biological systems we see around us.
While the diffuse background measurements of COBE reveal the importance of the far infrared and submillimeter in early galaxy and star formation, the understanding of the development of complex structure requires high resolution imaging and spectroscopy. It is clear from COBE observations of the far infrared background that much of the luminosity emitted in the critical initial phases of structure formation is emitted at far infrared wavelengths. Dust, which is responsible for the far infrared emission, hides much of the activity in the early universe from optical and near infrared study.
We present a concept for an instrument that will enable us to observe the dominant far infrared and submillimeter emission from the epoch of structure formation. The instrument, called the Submillimeter Probe of the Evolution of Cosmic Structure (SPECS), will produce high resolution images and spectroscopic data, allowing us our first clear view into the hidden environments where the structures in the universe developed. These observations will become powerful new tools to understand this crucial phase in the development of complex structures in the universe. SPECS will also open up new realms of discovery in the local universe. For example, individual star forming regions in a wide variety of nearby galaxies can be studied in detail, showing how the star formation process is affected by variable conditions such as metallicity and interstellar radiation field intensity. Such information will further improve our understanding of galaxies in the early universe.
SPECS is not only scientifically exciting, it is technically feasible. If a concerted effort is made to advance and test the required technologies during the next decade, it will be possible to build the SPECS observatory in the succeeding years.
## 1 Introduction
One of the most remarkable understandings developed by modern astrophysics is that the very complex local universe evolved from an earlier hot phase of almost perfect uniformity. Observations of the CMB reveal density fluctuations of only a few parts in $`10^5`$ at $`z1000`$. By the time we see the distant universe of galaxies in the Hubble Deep Field $`(z13)`$, these density fluctuations have developed into isolated galaxies and clusters in mostly empty space, with a whole array of shapes and sizes. The primary goal of SPECS is to provide a definitive observational basis for understanding the history of and the processes that drive the development of complex structure from the homogeneous early universe. While simple observations are sufficient to characterize simple systems, the rich complexity of this era of galaxy formation requires observations with sufficient spatial and spectral resolution to characterize the total luminosities, physical conditions, and morphological characteristics of these developing systems.
Observations with HST, ISO, Keck, JCMT and other large Earthbound telescopes have resulted in the identification of a large number of galaxies at redshifts out to $`z>3`$, and have begun to produce a consistent picture of cosmic evolution. The optically selected samples of galaxies show a rapid increase in star formation rate, hence luminosity, as a function of redshift, peaking at about 10 times that in the local universe at $`z1.5`$ (Madau, Pozzetti & Dickinson 1998). The far-infrared/submillimeter extragalactic background measured by the DIRBE and FIRAS instruments on COBE (Hauser et al. 1998; Fixsen et al. 1998) requires a similarly high rate of star formation at $`z1.5`$ (Dwek et al. 1998).
Conclusions regarding the earlier history of star formation are more tentative; the star formation rate may remain near the elevated rate seen at $`z1.5`$, or it may decline significantly. The background measurements provide a weak constraint on the star formation at $`z\genfrac{}{}{0pt}{}{_>}{^{}}1.5`$. Ground based submillimeter observations from the SCUBA camera on the JCMT and measurements with ISOCAM on ISO have recently revealed high z galaxies with very high star formation rates. These galaxies are completely undistinguished in an optical survey. Observations of the Hubble Deep Field at 450 and 850 $`\mu `$m suggest that optically faint galaxies undergoing massive starbursts may be responsible for 50% of the cosmic infrared background seen by FIRAS and DIRBE, and that as much as 80% of the luminosity of the early universe may be emitted in the far infrared (Hughes et al. 1998). Apparently most of the nucleosynthesis and its associated energy release, the motive force behind galactic evolution, occurs in environments that in optical studies are shrouded in dust. Indeed, extinction by dust is an important and ill-quantified source of systematic error in many cosmological surveys.
Thus we should strive to obtain a detailed view of the optically obscured star forming systems in the early universe; we need the ability to measure the luminosities, redshifts, metal abundances and morphologies of galaxies back to the epoch of their formation. SPECS achieves this goal with sensitive submillimeter interferometry and spectroscopy. As outlined in Table 2, SPECS provides high sensitivity and HST-like angular resolution in the far infrared, a wide field of view, and spectral resolution $`10^4`$. Since submillimeter radiation from the early universe is faint, cryogenic telescopes with background-limited direct detectors are required. The angular scales of the relevant structures are very small, so interferometric baselines ranging up to $`1`$ km are required.
Here we present a program of development that can, by late in the next decade, open the era of galaxy formation to detailed study. The emission from stars from this era is redshifted into the near infrared spectral region, where it can be studied by NGST. When combined with NGST observations, SPECS observations will provide direct measures of the total luminosities of forming and young galaxies by measuring their fluxes in the two spectral regions where the bulk of their energy is emitted.
SPECS will help answer several key questions about the basic characteristics of the universe:
1. When was “first light”? Did the first generation of stars form in early galaxies or before such systems existed?
2. What is the history of energy release and nucleosynthesis in the universe? How did carbon, oxygen, other heavy elements, and dust build up over time? What mechanisms were responsible for dispersing the metals?
3. Did the process or rate of star formation change over the course of cosmic history? How might any change in the star formation process be attributed to the gradual enrichment of the interstellar medium with heavy elements, or other factors still unknown?
4. What are the processes of structure formation in the universe? Were these processes hierarchical? What is the role of collisions between clouds and galaxy fragments? When and how did the first bulges, spheroids and disks form? How, ultimately, did the galaxies in today’s universe form?
In addition to these primary objectives, SPECS will provide unprecedented observations needed to understand the formation of stars and planets and the interaction of stars, at birth and death, with the interstellar medium.
In this paper, we argue for the necessity of a sensitive submillimeter interferometer for spectroscopic and imaging studies of the development of structure in the universe and the enrichment of the universe with heavy elements over cosmic time. Section 2 provides a theoretical and observational context in which gaps in our understanding are evident and the niches to be filled by SPECS are identified. Section 3 relates the physical processes that yield far-IR and submillimeter emission to the astrophysical systems that produce the emission, and illustrates how the SPECS observations can be used to address the scientific questions mentioned above. A technical concept for the SPECS observatory is presented in §5. We show that many of the technologies required for its realization exist, or are being developed as a part of the NASA program (§4). We identify the new technologies required (§5.3), and outline an incremental scientific program in which they can be developed during the next decade (§5.4).
## 2 Cosmic Evolution - Current Understandings and <br>Observational Context
The canonical picture of the evolution of the universe given below provides a framework for discussion of the need for an instrument like SPECS. As they are currently understood, the major developments were as follows:
* $`z>>10^7`$ – The expanding universe begins in a hot, dense Big Bang, including a period of cosmic inflation that produced a smooth distribution of matter over the scale of our horizon and the density fluctuations, a part in 10<sup>5</sup>, seen in the cosmic microwave background radiation. A few minutes later, nucleosynthesis results in the production of H, D, He, Li, Be, and B from the initial protons and neutrons. Observed abundances are consistent with this picture. Dark matter, perhaps both cold and hot, begins to move under its own gravity as soon as the distant universe came within the causal horizon.
* $`z1000`$ – The decoupling of radiation from matter allows baryonic matter to cluster around the dark matter. The universe becomes transparent, leaving behind the microwave background radiation observed by COBE. We know the statistics of the density field at this time, and believe that a linear theory describes the growth of density fluctuations. This is the basis of the claim that the cosmic microwave background fluctuations can measure the main cosmic parameters. The MAP (2000) and Planck (2007) missions, and many ground based measurements, will do this.
* $`z20`$ – The first luminous objects form in fluctuations of greatest density enhancement ($`>3\sigma `$). They must cool off as they collapse, emitting spectral lines from H<sub>2</sub> and H. Such lines would be extremely weak (of the order of pico-Janskys) and are not expected to be observable. However, the first star-forming systems might be detectable; whereas the binding energy, only $`1`$ – 10 eV/nucleon, is released during cloud collapse, nuclear fusion releases a few MeV per nucleon. In this epoch massive stars (Population III?) form and begin to ionize the intergalactic medium, and some expel heavy elements in supernova explosions and stellar winds. These objects are entirely hypothetical, but we know that something ionized the intergalactic medium by a redshift of 5. With SPECS we will be able to find and characterize the first stars even if our view is obscured by dust, learn when these objects produced heavy elements, and determine when the first dust formed. At a redshift of 20, Ly$`\alpha `$ is observable by NGST. The redshifted Brackett lines could be observed by SPECS, as could the near IR emission from cool stars, continuum emission from dust in H II regions, and several important diagnostic fine structure lines from heavy elements (see §3). The \[C II\] 158 $`\mu \mathrm{m}`$ line is redshifted to 3.3 mm and could be seen with the Millimeter Array (MMA), along with thermal emission from dust associated with the neutral phases of the ISM.
* $`z320`$ – Secondary structure formation. Cloud cooling is enhanced by the inclusion of newly synthesized heavy elements. Galaxies grow by collisions and absorption of smaller fragments, with a rate governed by the statistics of the primordial density fluctuations and their growth. Many are very dusty, with star formation obscured by very local dust from young hot stars and supernovae. Interstellar shocks reprocess the dust. Some heavy elements enrich the newly ionized intergalactic medium, driven by high pressures and outflows from small galaxies with insufficient gravity to retain the debris. Gas liberated from the galaxies by collisions is heated and radiates X-rays. Deep potential wells form in clusters and gravitational lenses allow far IR measurements of even more distant objects. SIRTF will count the galaxies and measure their luminosity functions. NGST will observe the stars in the galaxies, and the obscuration by dust. SPECS will observe the dust luminosity directly, allowing the inference of the hidden stellar luminosity.
* $`z13`$ – Star formation peaks with frequent collisions of galaxies and fragments. Cooling flows allow hot gas to fall to the centers of galaxy clusters and disappear from view, possibly forming stars (O’Connell and McNamara 1991). In many galaxies Active Galactic Nuclei (AGN) form in dense cores with accretion disks of infalling material and produce jets observable from radio to X-ray. These AGN have strong effects on their host galaxies. SIRTF establishes the relationship between AGN and the ultraluminous infrared galaxies first seen by IRAS. HST reveals the morphologies and colors of young galaxies and tells us that collisions are common. NGST sees the stars and, depending on its long wavelength coverage, part way into the dense cores of AGN. SPECS sees into the dense cores, penetrating the opaque dust and seeing its emission. A number of H<sub>2</sub>O lines, which are important diagnostics of star forming molecular clouds (Harwit et al. 1998), will be visible to SPECS. The combination of NGST and SPECS allows us to distinguish the emission from individual star forming regions and learn whether and how much star formation is driven by collisions of galaxies, spiral density waves in existing galaxies, feedback from other star formation, or other causes.
* $`z0,`$ the local universe – Galaxy mergers have nearly ceased. Nearby galaxies allow us the maximum visibility into star formation processes, which now occur presumably under very different densities and radiation fields than at higher redshifts. NGST and SPECS are complementary, and both are necessary for a complete picture. In the Milky Way and more than a hundred nearby galaxies, SPECS allows us to see inside dusty clouds where stars and planets are forming. The high angular resolution provides a new dimension of information. Studies of the carbon, water, nitrogen, and oxygen lines show the cooling of the warm phase of the interstellar gas at high spatial resolution. Protostellar disks in local star forming complexes can be resolved in these lines, enabling the development of more detailed astrophysical models of protostars.
In cosmology, the far IR and submillimeter region is centrally important. A plot of the luminosity of a typical spiral galaxy (the Milky Way, Figure 1) has just two large bumps, one from 0.2 to 2 $`\mu \mathrm{m}`$ from the luminosity of stars, and one nearly as large from 50 - 400 $`\mu \mathrm{m}`$, from the energy absorbed and reradiated by dust. These two terms dwarf all the others, but only one has been widely observed; the far IR emission from most galaxies is barely detectable with present techniques, and very few galaxies can be resolved into parts. Early galaxies tend to be even more profuse infrared emitters than typical galaxies in the local universe, as illustrated here by the spectrum of a representative starburst galaxy (Figure 2). It seems likely from direct measurements of high redshift galaxies (Blain et al. 1998; Barger et al. 1998) and from measurements of the cosmic far infrared background (Hauser et al. 1998; Fixsen et al. 1998) that much of the luminosity of the early universe has been hidden by dust absorption and is reemitted in the far infrared. The history and even the list of constituents of the universe is presently quite incomplete. There is a huge uncharted far IR universe awaiting the first discoverers with adequately sensitive instruments.
To summarize, an instrument like SPECS is needed to fill a gap in observational capability at far infrared and submillimeter wavelengths that translates directly into an inability to solve some of the greatest mysteries currently recognized in astronomy. The SPECS observations would be complementary to those of the NGST and MMA. A triad of observatories consisting of these three would allow us to take the next leap forward in our understanding of the universe.
## 3 Physical Processes in Submillimeter Emission
The purpose of this section is to relate the observational capabilities envisioned for SPECS (Table 2) to the primary scientific objectives (Table 1). To make this connection one must consider the physical processes that give rise to the observable emission.
The main processes known to produce far infrared radiation are thermal emission from interstellar dust, fine structure line emission from interstellar atoms and ions, rotational line emission from molecules, and synchrotron and bremsstrahlung radiation from hot electrons in dense regions like active galactic nuclei. Atomic and molecular features that play key roles in the collapse of star-forming clouds and in the energy balance of the interstellar medium are unique to the far IR. The line emission allows detailed investigation of physical conditions such as temperature, density, chemical composition, and ionization state. Very small dust grains emit “PAH features” and nonthermal radiation as they spin at many GHz rates. Condensed objects (stars, planets, asteroids, and comets) all emit at far infrared wavelengths, and if they are cold enough ($`<40`$ K) they have no other emission. In addition, rest frame near- and mid-IR emission from high-z galaxies appears in the far infrared, making this the relevant spectral window for observations of hydrogen Brackett line and cool star emission from such galaxies. A wide range of phenomena are uniquely well observed in the far IR.
### 3.1 Structure in the Early, Metal-free Universe
Since the heavy elements responsible for the cooling and collapse of gas clouds in the present universe are not available for the formation of the first generation of stars and galaxies, the detailed physical mechanisms of initial formation must be very different. SPECS will provide us with a unique view of the first galaxies and protogalactic systems. For example, assuming that molecular hydrogen clouds were the progenitors of the first stars, the redshifted H<sub>2</sub> 17 and 28 $`\mu \mathrm{m}`$ rotational lines could be observable. SPECS will resolve a hypothetical $`z=10`$ object at the 100 pc scale in the 17 $`\mu \mathrm{m}`$ line. If stars formed as early as $`z=20`$, then SPECS would detect the Br$`\alpha `$ and Br$`\gamma `$ lines at the redshifted wavelengths 85 $`\mu \mathrm{m}`$ and 45 $`\mu \mathrm{m}`$, respectively. From SPECS observations it will be possible to learn whether the first stars formed in pre-existing galaxies, or if stars formed earlier in pre-galactic clouds.
### 3.2 The History of Energy Release and Nucleosynthesis
What produces the cosmic far infrared background? It is usually said that the far IR emission comes from starburst activity (e.g., Haarsma & Partridge 1998) and (in combination with the UV-optical emission) reveals the nucleosynthesis history of the universe. However, it would be more accurate to say that the far IR background reveals the luminosity history, since we do not know whether the luminosity comes from starbursts or from AGN. Some have argued, combining ISO, SCUBA, and X-ray data, that much of the X-ray and far IR background comes from AGN sources associated with and partially obscured by intense nuclear starbursts (Almaini, Lawrence, and Boyle 1998; Almaini et al. 1998; Fabian et al. 1998; Genzel et al. 1997; Lutz 1998; Lilly et al. 1998). The actual nature of the far IR sources detected with the SCUBA array (Hughes et al. 1998; Barger et al. 1998) is unknown.
There are several ways to investigate such far IR objects. Many have far IR luminosities that exceed those at other wavelengths by an order of magnitude, so it is clearly important to observe this dominant luminosity as well as possible. Low spectral resolution photometry over the entire wavelength range from X-ray to far IR is required to get the luminosity; extrapolations from wavelengths that are conveniently observed are not really sufficient. Those objects that are not too heavily obscured can be seen with HST, and high resolution UV images might reveal star clusters or compact disks. Hard X rays ($`>`$ 3 keV) can penetrate high column densities of obscuring material, directly revealing a non-thermal source. Mid IR (5 - 30 $`\mu `$m) spectroscopy (preferably spatially resolved) can observe the 5 - 12 $`\mu `$m AGN/starburst diagnostic lines and dust features found by ISO (Genzel et al. 1997), with the observable spectra depending on the depth of obscuration and on the redshift. Long wavelength high angular resolution imaging (e.g., with the MMA or a space interferometer) is also required, to locate the sources precisely and determine their spatial structure. Very high spatial resolution is required to measure the diameter of a compact source at the core, if there is one, calling for a millimeter array or a far IR space interferometer with very long baselines.
SPECS would fill a very important gap. The objects seen with SCUBA are at the milliJansky level at 850 $`\mu \mathrm{m}`$, and are considerably brighter (though less easily seen from the ground) at shorter wavelengths. The SPECS sensitivity is of the order of 10 $`\mu `$Jy (see §5), two or three orders of magnitude below the SCUBA source brightness, and it would have the combined spatial resolution and sensitivity to resolve the SCUBA sources into hundreds of pixels. In the end, we would know how and where most of the cosmic energy in the whole X-ray to far IR range is liberated, something we can not know today.
In normal spiral galaxies, the \[C II\] line at 158 $`\mu \mathrm{m}`$ is often the brightest emission line at any wavelength, dominating the cooling of the Warm Neutral Medium. In typical spirals, the line can have a total luminosity about 1/3% of the total system luminosity (Stacey et al. 1991). Although a deficit is seen in the 158 $`\mu \mathrm{m}`$ line emission from ultraluminous infrared galaxies, suggesting that perhaps high redshift galaxies show weak \[C II\] emission (Luhman et al. 1998), the line still may prove to be a useful marker for such galaxies. SPECS observations of galaxies over a wide range of redshifts in the \[C II\] and other fine structure lines from oxygen and nitrogen will enable us to trace the buildup of heavy elements in the universe over time. Moderate spectral resolution ($`10^3`$ \- $`10^4`$) will be needed to detect the line emission. High angular resolution is required to examine the mechanisms responsible for dispersing the metals in a large number of galaxies.
### 3.3 The Evolution of Galaxies and Structures
High spectral and spatial resolution observations will yield both kinematic and morphological information about the structures in the universe. We would like to know how these structures interacted and developed over time into the $`z=0`$ galaxies. What were the progenitors of halos, bulges and disks, and galaxy clusters and superclusters, and when did these structures form? Did galaxies form from the “top down,” or were they assembled from bits and pieces? How did the different galaxy morphological types develop? The broad and continuous spectral coverage provided by NGST, SPECS and the MMA, will enable us to follow the development of structure from $`z=10`$ to $`z=0`$ in a fixed set of spectral lines and features.
### 3.4 Galaxies as Astrophysical Systems
Following the initial phase of nucleosynthesis in the universe, the dominant cooling and diagnostic lines are in the far IR region, and are not accessible to detailed study from the ground (except at very high redshift). This spectral region includes the fine structure lines of \[C II\], \[O I\] and \[N II\], which dominate the cooling of and reveal the physical conditions in neutral and ionized gas clouds. These elements are also the building blocks of life, and it is important to know when and where they are produced and turned into planets.
Molecular lines in the far IR from CO and H<sub>2</sub>O dominate the cooling of dense molecular clouds and are thus critical players in the formation of stars. Additional diagnostic lines, such as those of the hydrides (OH, CH, FeH, etc.) are also found in this spectral region. Thus spectral line measurements will allow us to probe dense, cold material in dark clouds, providing insight into the details of cloud collapse in star formation. Longer wavelength studies with ground based telescopes of other less energetically important lines have been remarkably successful, but we are seriously limited at present by our inability to observe the lines that dominate the energy balance.
In addition to the far IR fine structure lines mentioned above, many galaxies are luminous sources of mid IR fine structure emissions from Ne II (12.8 $`\mu `$m), O III (52, 88 $`\mu `$m), Ne III (15.6, 36.0 $`\mu `$m), Ne V (15.6, 36.0 $`\mu `$m) and several other ions that result from photoionization by radiation shortward of the Lyman limit (e.g., Moorwood et al. 1996). Even at moderate redshift, all of these important diagnostic lines are shifted longward of even the longest wavelengths contemplated for NGST. The mid IR lines provide unique probes of the metallicity and gas density in ionized regions, as well as the spectral shape of the ionizing radiation field (e.g., Voit 1992). These transitions have several important advantages over the optical wavelength lines traditionally used to probe H II regions: they are not heavily extinguished by interstellar dust; their luminosities are only weakly dependent on temperature and therefore provide model-independent estimates of metallicity; and they provide line ratios that are useful diagnostics of density over a wide dynamic range. The availability of noble gas elements (e.g., Ne, Ar) allows the metallicity to be determined without the complicating effects of interstellar depletion, and the availability of a wide range of ionization states (e.g., NeII and NeV) provides an excellent discriminant between regions that are ionized by hot stars and those that are ionized by a harder source of radiation such as an AGN.
The spectral bump characteristic of thermal dust emission (see Figures 1 and 2) will be evident in galaxies dating back to the first epoch of dust formation. Recent observations of an individual galaxy indicate that dust existed at z = 5.34 (Armus et al. 1998). Extinction effects become important as soon as there is dust.
SPECS will spatially resolve individual galaxies and allow us to study them in the spectral lines that dominate the energy balance of the interstellar medium, cloud collapse, and star formation, and in the spectral features that signify the presence of dust. We can expect to observe changes in the physical properties of the interstellar medium, and the star formation rate or process, as a function of redshift. It might be possible to infer, for example, the effects of increasing metallicity on molecular cloud cooling and the rate of star formation. In any case, clearly, such a rich data set would provide the basis for new and improved astrophysical models.
In summary, observations of the early universe in the far infrared and submillimeter will allow us to:
1. Obtain a complete census of energy release in the universe as a function of cosmic time;
2. Probe physical conditions associated with star formation in the early universe, including metallicity and depletion in the interstellar media of high redshift galaxies;
3. Investigate the kinematics of the newly formed galaxies to help understand the role of collisions in the initiation of star formation and the growth of complex structure; and
4. Measure the relative importance of ionization by starlight and by active galactic nuclei in the high redshift universe.
Observations of star formation in the local universe are important in providing a detailed picture of this process in environments enriched with heavy elements.
If high sensitivity and angular resolution can be extended to the far IR/submillimeter, then new kinds of objects may be detectable. Old planets far from their parent stars, either in distant orbits or already escaped, could be numerous and detectable only in the far IR.
## 4 Submillimeter Astronomy Today, Opportunities for Tomorrow
To put our current submillimeter observational capability in perspective, we can compare the present facilities with the human eye. At a wavelength of 1 mm, even a large telescope like the 15 meter JCMT (James Clerk Maxwell Telescope) at Mauna Kea has an angular resolution only as good as a 7.5 mm diameter telescope at 0.5 $`\mu `$m; in other words, far IR astronomy now has the angular resolution of an ideal human eye. Had the Hubble Deep Field been observed at comparable resolution, most of the interesting information would have been missed. The far IR telescope and atmosphere are also glowing brightly, compared to the faint objects being observed. In one sense this is tremendously disappointing, but in another it is a tremendous opportunity.
To meet this opportunity, several major projects are planned or being built. The SIRTF (Space Infrared Telescope Facility), with an 0.8 meter telescope at 4 K, is now under construction. The Far IR Submillimeter Telescope (FIRST), an ESA Cornerstone mission with a US-contributed 3.5 meter telescope operating at a temperature of 70 K, will fly around 2007. The Japanese H2L2 mission is being planned for a launch in 2012 with a 3.5 m telescope at 4 K (Matsumoto 1998). These new telescopes represent a continuing series of more sensitive instruments, but they will be limited by angular resolution only a little better than the human eye. The images of distant objects obtained with these telescopes will be fuzzy and overlapping. We need a far-infrared telescope that can distinguish individual distant galaxies. On the ground, the MMA is to be built in Chile, an extraordinary facility with coherent receivers, upward of 40 dishes of 10 m aperture, and spacings up to kilometers apart. This equipment will provide excellent angular resolution, comparable to the HST. It will be limited in sensitivity by antenna spillover and atmospheric emission at ambient temperature, and in dynamic range by atmospheric fluctuations, but with long exposures will be able to find immense numbers of galaxies because of its huge collecting areas.
The science drivers for SPECS are derived from the considerations of the previous sections. We have already shown how several fundamental questions pertaining to the evolution of structure in the universe can be addressed observationally. None of the existing or planned facilities will provide the combination of high sensitivity, high angular resolution, and large solid angle at far IR wavelengths that we believe is necessary to meet the scientific challenge. The obvious answer is a far IR space interferometer with cold, photon-counting array detectors.
The development of space interferometry was recommended by the HST and Beyond committee (Dressler et al. 1996) and by the Bahcall decade survey committee (Bahcall et al. 1991), although the emphasis in those reports was on shorter wavelengths. One reason for this emphasis is the strong desire to find and learn about planets around other stars, and the mid IR band (5 - 20 $`\mu \mathrm{m}`$) contains the key spectral signatures of photosynthetic life (ozone, water, and carbon dioxide). Also, far IR astronomy is so far behind other areas that its potential can be hard to recognize. Acting on the Bahcall and Dressler Committee recommendations, NASA and ESA are both preparing for space interferometers at visible and near IR wavelengths, and the technological basis is under development.
The key new development that enables sensitive far infrared space interferometers is the detector system. The current generation of bolometric detectors is already good enough to be limited by the photon fluctuation noise of the cosmic infrared background light in wide bandwidths. Semiconductor bolometers can be made in arrays by two different techniques, and improvements using superconducting thermometers and amplifiers promise to reach the extreme sensitivities required for this mission (Irwin 1995; Lee et al. 1996, 1997; Lee et al. 1998). Another promising detector technology is direct detection with superconducting tunnel junctions. These devices convert far IR light directly into photocurrents that can be amplified and measured. The necessary superconducting electron counting amplifier design has just been developed (Schoelkopf et al. 1998), and there is every reason to be optimistic about continued improvements. For many applications, current direct detectors are already far more sensitive than even ideal (“quantum limited”) amplifiers or heterodyne receivers, which suffer additional quantum fluctuation noise equivalent to receiving one photon per second per unit bandwidth and polarization, and can be orders of magnitude more sensitive for the SPECS application.
Photon detectors are preferred whenever the background photon rates are small compared with the detector bandwidths of interest. This is true for broadband systems at wavelengths shorter than about 1 mm, although for very high spectral resolution the coherent receivers are competitive down to about 300 $`\mu \mathrm{m}`$ wavelength. The sensitivity of an ideal photon detector and amplifier are shown in Figure 3, using the COBE DIRBE measurements of the brightness of the darkest part of the sky. Note that the sensitivity worsens as the wavelength increases, but longer observing times and slightly larger collecting areas can compensate.
An interferometer would have the sensitivity of a single dish with the same collecting area, but the angular resolution of a telescope with the full aperture. Considering the sensitivity improvements that are possible, it seems that the individual apertures could be relatively small (a few meters) while the maximum useful separation could easily be hundreds to thousands of meters. Even accounting for all the penalties of interferometry, such a device could map thousands of $`10^{10}\mathrm{L}_{}`$ galaxies at $`z=1`$ in a very reasonable time.
## 5 The Submillimeter Probe of the Evolution of Cosmic Structure (SPECS)
The SPECS that we envision has the characteristics outlined in Table 2. In particular, the science drivers (§3) are satisfied with several to 10 $`\mu `$Jy sensitivity, several tens of milliarcsecond angular resolution, $`10^4`$ spectral resolution, and a 15$`^{}`$ field of view.
Below we describe the basic concepts and configurations to be considered.
### 5.1 Basic Concepts
The principle of operation of a spatial interferometer is to determine the two-point correlation function of the incoming waves, and from that to compute a map and spectrum. A monochromatic plane wave can be described as $`\psi (x,t)=\psi _0\mathrm{exp}i(\stackrel{}{𝐤}\stackrel{}{𝐱}+\omega t)`$, where $`\stackrel{}{𝐱}`$ and $`t`$ are position and time, and $`\stackrel{}{𝐤}`$ and $`\omega `$ are the wavevector $`(|\stackrel{}{𝐤}|=2\pi /\lambda )`$ and frequency $`(\omega =2\pi f)`$. Its correlation function is
$$C(\stackrel{}{𝐱},\stackrel{}{𝐱}^{},t,t^{})=E(\psi (\stackrel{}{𝐱},t)\psi ^{}(\stackrel{}{𝐱}^{},t^{}))=|\psi _0^2|\mathrm{exp}i(\stackrel{}{𝐤}(\stackrel{}{𝐱}\stackrel{}{𝐱}^{})+\omega (tt^{})),$$
(1)
where $`E`$ is the expectation value function. For a steady sky signal, the correlation function depends only on the spatial separation $`\stackrel{}{𝐱}\stackrel{}{𝐱}^{}`$ and the time delay $`tt^{}`$ between the two points of observation. By custom the component of $`\stackrel{}{𝐱}\stackrel{}{𝐱}^{}`$ perpendicular to the central line of sight is labeled by $`(u,v)\lambda `$. The component of $`\stackrel{}{𝐱}\stackrel{}{𝐱}^{}`$ parallel to the line of sight is absorbed in the time delay term.
If there are only two telescopes at $`\stackrel{}{𝐱}`$ and $`\stackrel{}{𝐱}^{}`$, and the time delay $`tt^{}`$ is fixed, then the angular response function of the correlation function is a simple cosine pattern with angular scale of 1 cycle per $`\delta \theta =1/|(u,v)|`$. If the two telescopes are moved around to measure more data points, the angular resolution of the deduced map is approximately $`\delta \theta =1/\mathrm{max}|(u,v)|`$. The number of observations at different values of $`(u,v)`$ gives the maximum number of parameters of the map that can be deduced from the data. If the values of $`(u,v)`$ are uniformly spaced in a square grid pattern, the algorithm to recover the sky map is a simple Fourier transform. Indeed, the Van Cittert-Zernike theorem says that the correlation function for a given $`\stackrel{}{𝐱},\stackrel{}{𝐱}^{}`$ and an intensity distribution on the sky is just a Fourier component
$$C(0)I(\stackrel{}{𝐬})exp(2\pi i\stackrel{}{𝐬}(\stackrel{}{𝐱}\stackrel{}{𝐱}^{})/\lambda )𝑑\mathrm{\Omega },$$
(2)
where $`I(\stackrel{}{𝐬})`$ is the brightness in the direction $`\stackrel{}{𝐬}`$, $`\stackrel{}{𝐬}`$ is a unit vector, and $`\mathrm{\Omega }`$ is solid angle. Here the $`C(0)`$ notation refers to setting the time delay equal to zero, so that the waves being correlated are in phase.
Spectral resolution can be obtained either by choosing a narrow band filter, or by performing measurements of the dependence of the correlation function on the time delay. Some spectral resolution is required to support the spatial interferometry, because the angular resolution and the whole scale of the reconstructed Fourier transform image depends on wavelength. In general the reciprocal spectral resolution $`R^1\delta \lambda /\lambda `$ should be significantly less than the ratio of the size of the synthesized beam $`\delta \theta `$ to that of the primary beam ($`\lambda /D_t`$), or $`R>\mathrm{max}|(u,v)|\lambda /D_t`$; otherwise the reconstruction is ambiguous.
We can now compare coherent and optical technologies. With coherent receivers, the wavefunction $`\psi `$ is measured directly, amplified, and relayed to a central computer, and all the correlation properties can be computed electronically in a giant digital correlator. With optical methods, the correlation function is determined using square law detectors (photon or energy detectors) and beamsplitters. Assume a beamsplitter with transmission coefficient $`t`$ and reflection coefficient $`r`$ is used to combine two beams, one reflected and one transmitted. The output amplitude on one side is then $`\psi =r\psi _1+t\psi _2`$, and the intensity is $`I=|r\psi _1|^2+|t\psi _2|^2+2\mathrm{}(rt^{}\psi _1\psi _2^{})`$, where $`\mathrm{}`$ is the operator that finds the real part of a complex number. This last term contains the needed information about the correlation function of the two input waves.
With coherent receivers, it is convenient to measure the correlation function at many different time delays (lags) and to deduce the spectrum of the incoming radiation from the Fourier transform of this distribution. Since the waves have already been amplified, this can all be done simultaneously in a digital correlator. In the optical case, the equivalent is a grating or prism spectrometer, which performs the coherent transformations instantaneously and disperses the output photons across a detector array. This has the advantage that the photon fluctuation noise is also distributed with the signal, and is the ideal when it is possible to provide enough detectors. Optical technology offers another method as well, the Michelson spectrometer. In this method, the time delay is varied by a moving mirror, and an interferogram is measured. The Fourier transform of this interferogram is the desired spectrum, just as in the coherent case. The disadvantage is that the photon fluctuation noise aflicts all the computed frequency bins regardless of the input spectrum. On the other hand, the Michelson spectrometer can easily cover a wide field of view simultaneously. This offers many potential advantages for the SPECS mission.
The distribution of points in the $`(u,v)`$ plane governs the quality of the image that can be reconstructed, and there is a large literature on the question of ideal arrangements. With ground based interferometers the optimization favors a large number $`m`$ (10-50) of separate antennas in a Y shaped or circular pattern. Then the rotation of the Earth changes the orientation of the array relative to the sky, and each of the $`m(m1)/2`$ pairs of antennas produces a track in the $`(u,v)`$ plane. Adjustable delay lines (e.g., coax cables) compensate for the fact that the antennas are not located in a plane perpendicular to the line of sight. With enough pairs and enough Earth rotation, a large number of points can be measured in a day. To get more coverage, the antennas can be picked up and moved, but this is time consuming.
With a space optical interferometer, complexity grows rapidly with the number of telescopes, so the smallest possible number is favored. With three telescopes, there are phase closure relations among the observations that enable self-calibration of certain instrument properties (Pearson and Readhead, 1984). With four antennas, there are amplitude closure relations as well, and with five antennas there are enough combinations to make redundant self-calibrations. We have not analyzed this question fully, but are hopeful that the number of telescopes can be kept down to three by careful use of the information from multiple pixels within each field of view, by careful calibration of the equipment response functions, and by depending on good long term stability. This is not generally done with ground based instruments because the complexity of the receivers favors the use of more antennas instead of more detectors. Also, on the ground, atmospheric fluctuations are very rapid, demanding the use of the self-calibration relations on a second-by-second basis. We hope this is not necessary in space.
With a space interferometer, there is a choice of methods of moving the telescopes in the $`(u,v)`$ plane. The large scale of the desired SPECS precludes the use of rigid physical structures, so there must be separated spacecraft flying in formation. These can be completely independent, or they can be tethered together to reduce fuel consumption. If they are tethered together, the entire collection of telescopes can easily spin at relatively high speed around the line of sight, producing observations of a circular set of points in the $`(u,v)`$ plane. If the tethers are played in and out together, the radius of the circle in the $`(u,v)`$ plane changes too, so it may be easy to produce a spiral pattern. It would be possible to sample the $`(u,v)`$ plane completely in 2 days with 1 km maximum baselines if the scanning mirror could be made to stroke at about 1 Hz. If the spacecraft are not tethered together, then a more natural scan pattern may be a simple raster, in which velocity changes are abrupt at the end of each sweep. Whether this is feasible depends on the details of the propulsion system.
An advantage of a space interferometer is that in principle all spacings are available. There is no need to move heavy, fixed antennas from one attachment fitting to another, an approach often used on the ground. Elaborate algorithms such as CLEAN and Maximum Entropy have been developed to handle the irregularly spaced data collected with ground-based interferometers. Although perfectly uniformly sampled data could be Fourier transformed directly to produce an image, adaptations of CLEAN and Maximum Entropy likely will be useful in the construction of SPECS images, and might allow us to produce high-quality images even if the $`(u,v)`$ plane is undersampled. Unlimited by atmospheric fluctuations, and with the freedom to position the light-collecting telescopes where they are needed, a space-based interferometer can achieve much greater dynamic range than a ground-based system.
There are important choices to be made about the spatial and spectral resolution to be used for each target. For unresolved sources, the sensitivity is determined primarily by the collecting area and the observing time, and is about the same as for a filled aperture telescope of the same area. Excess resolution beyond what is needed for the interesting features is a waste of observing time. For a space interferometer, the spatial resolution is controlled by the scan pattern of the moving remote mirrors. The spectral resolution should be chosen to match either the spatial resolution, or the intrinsically interesting spectral structure of the sources.
### 5.2 Interferometer Configuration
A typical far IR imaging interferometer concept would include the following features. At least three telescopes are required, to take advantage of the possibilities of self-calibration of the telescope positions with phase closure algorithms. They would be cooled to a low temperature, so that their thermal emission is less than the sky brightness and they do not dominate the noise, as shown in Figure 4. These temperatures are now achievable with radiation baffles and active coolers, in deep space.
One way to arrange the telescopes in a far-infrared interferometer is illustrated in Figures 5 and 6. There are three telescopes at the central beam combining station, each looking out sideways at a diagonal flat. The diagonal flats move in formation to change the location of the observing stations, and all are located equally far from their respective telescopes. This arrangement produces the least possible beam divergence between the telescopes and the diagonal flats, since the telescopes at the central station have the largest possible apertures. It also allows the diagonal flats to be packed close together, providing low spatial frequency information. The telescopes are conceived as off-axis Cassegrain afocal systems with a magnification factor of about 10. Their output beams reflect from small diagonal flats to become parallel to the original line of sight from the sky, as illustrated in Figure 6.
The telescope beams would be combined to measure the correlation functions, using beamsplitters (as in the Michelson spectrometer) or geometrically (as in the Michelson stellar interferometer). We suggest the beamsplitter approach, as that is the most similar to the successful imaging interferometers done with microwave receivers, and the algorithms are fully developed. In this case, the path difference between the input beams is modulated to measure the spectrum of the coherence function. This is like the digital correlator for the MMA, which determines spectra by correlating one input signal with another as a function of time delay. It has been shown that in the photon noise limited case, all forms of beam combination produce approximately the same sensitivity (Prasad and Kulkarni, 1989). With $`m`$ telescopes, there are $`m(m1)/2`$ beam combiners, each with two output beams, so there are $`m(m1)`$ focal planes. One such concept is shown in Figure 7. The numbers of reflections are equal in each path, so that the images from different telescopes can be superposed exactly.
With the beamsplitter approach each focal plane can be a direct image of the sky, feeding an array of detectors just as though it were at the focal point of a single telescope. The field of view can be as large as the telescope aberrations and detector technology allow; we think a field of 1/4 degree may be feasible. This is totally different from the standard microwave array, in which each antenna feeds a single pixel receiver. As a result, the far IR version can have a huge advantage in observing speed, proportional to the number of detectors. Effectively there are now $`N_{\mathrm{pix}}\times N_{\mathrm{pix}}`$ separate interferometers operating simultaneously, each with its own field of view of approximately $`\mathrm{\Delta }\theta =\lambda /D_t`$, where $`D_t`$ is the diameter of each telescope. Since this $`\mathrm{\Delta }\theta `$ can be much smaller than the geometrical field of view of the telescope, it is possible to provide a large number of pixels. This is in addition to the sensitivity advantage of each detector over a coherent receiver. With the beamsplitter approach, all frequencies are modulated and observed simultaneously, much as they are in the digital correlators. However, the digital correlators use amplified copies of the input signals and therefore have different noise characteristics. Digital correlators have the advantage that they can measure many different time delays simultaneously, partially compensating for the lower detector sensitivity of coherent systems in low background conditions, but they can not be used in photon background limited receivers.
The maximum spectral resolution achievable is governed by the beam divergence at the beam combiner, because the path difference is multiplied by the cosine of the angle of each ray from the central ray, and a range of angles corresponds to an apparent range of wavelengths. With a large diameter system like that described here, there is no problem achieving resolutions of the order of $`R=10^4`$, although the path difference required becomes large, approximately $`R\lambda `$. If this much spectral resolution is required, a more compact spectrometer such as the Fabry-Perot filter would be favored, and would have the advantage of reducing the photon noise from signals outside the bandpass of interest.
The spatial resolution obtained is governed by the maximum spacing of the mirrors. At a wavelength of 0.5 $`\mu \mathrm{m}`$, the HST mirror is 4.8 million wavelengths across. For the same angular resolution at 200 $`\mu \mathrm{m}`$, the mirrors need to be 960 meters apart. This is clearly impossible with a rigid physical structure in space, but it could be achieved with formation flying. The technology to do this will be demonstrated with the DS-3, a New Millennium program mission scheduled for launch in 2002. The DS-3 is intended to achieve much higher positional accuracy and stability than are required for a far IR interferometer. Because the remote mirrors must be propelled in a pattern, it is desirable that their masses should be as low as possible. For this level of accuracy, it may be that a stretched membrane on a hoop could be made flat enough and controlled well enough with electrostatic forces. In any case the spacecraft engineering challenge is significant.
The minimum spectral resolution needed is governed by the number of pixels into which each point source may be resolved (i.e., the ratio of the primary beam diameter to that of the synthesized beam). Since the synthesized beam diameter $`\delta \theta `$ is wavelength dependent, the wavelength must be known to a certain tolerance. For a general image, the spectral resolution should be greater than the ratio of the mirror separation to the mirror diameter.
The stroke of the Michelson interferometer retroreflectors produces interferograms that are Fourier transformed to obtain spectra of the correlation function between two apertures. The spectral resolution $`R`$ obtained is governed by the path difference range, which must be approximately $`R\lambda `$ long. In the case of the wide field interferometer described here, each individual detector pixel has its zero path difference point located at a different part of the mirror stroke, and the total path difference range needed to cover the whole field of view is $`\theta b_{max}`$, where $`\theta `$ is the width of the field and $`b_{max}`$ is the spacing between the remote reflectors. Although this is much longer than the stroke needed for a single detector pixel, by a factor of the number of rows or columns of detectors in each array, the array is faster than a single pixel detector in proportion to the number of rows of elements. At $`R=10^4`$, $`R\lambda \theta b_{max}`$.
The necessary orbit for such an interferometer must certainly be in deep space, far from the Earth. Otherwise, the viewing geometry relative to the Sun and Earth is so variable that good radiative cooling is impossible, and the long observing times required are not achievable.
The telescopes would each be shielded from the Sun by a radiative baffle like that developed for NGST. Additional cooling would come from an active cooler, and at least four types are currently popular: the reverse Turbo Brayton cycle; the Stirling cycle; the pulse tube; and the sorption pump cooler. These coolers could reach temperatures of the order of 4 K with some effort. The detectors will certainly need to be colder, of the order of 0.05 to 0.5 K depending on the technology chosen. The additional cooling may come from adiabatic demagnetization, as developed for ASTRO-E, or from helium-3 dilution, as developed for FIRST.
The sensitivity achievable by such a system is shown in Figure 8. The sensitivity is $`\delta (\nu S_\nu )=\mathrm{NEP}R/(Aϵt^{1/2})`$, where $`\delta (\nu S_\nu )`$ is the $`1\sigma `$ uncertainty in the brightness, $`R`$ is the spectral resolution $`\nu /\delta \nu =\lambda /\delta \lambda `$, $`\mathrm{NEP}`$ is the noise equivalent power from photon noise, $`ϵ`$ is the system efficiency including a sine wave modulation factor of $`1/8^{1/2}`$, and $`t`$ is the observing time. The $`\mathrm{NEP}`$ is for a single detector and must be calculated allowing for the loss of photons on the way to the detectors. The collecting area $`A`$ is for the entire system, including all $`m`$ apertures. The subtleties about the details of beam combination are embedded in the efficiency factor. For Figure 8 we have chosen three apertures of 3 m each, an efficiency of 10%, a background limited photon noise for a 100% bandwidth and that efficiency, and an observing time of 10<sup>5</sup> seconds, about 1 day. We also assumed a spatial resolution of 0.05$`^{\prime \prime }`$, which implies a maximum mirror separation and a required spectral resolution. The plot shows two curves, one with the full spectral resolution, and one with the spectra smoothed to give a wide band image and gain a factor of R<sup>1/2</sup> in sensitivity. The sensitivity is easily adequate to reach the brightness of a high-z galaxy, and to resolve such a galaxy into much fainter sub-units.
At 450 $`\mu \mathrm{m}`$, where Noise Equivalent Flux Density data are readily available (Hughes and Dunlop 1997), the sensitivity of SPECS would exceed that of the other major submillimeter facilities by factors of 50 (FIRST and the South Pole 10 m telescope), 150 (MMA), 200 (SOFIA), 700 (JCMT), or more.
We note that a smaller version of the SPECS could still have high enough sensitivity to match the Hubble Space Telescope if the detector $`\mathrm{NEP}`$ can be improved. This could be feasible with new technology detectors combined with dispersive spectrometers to limit the photon noise on each one.
### 5.3 Technology Requirements
Many items need to be developed to bring SPECS to fruition. Clearly an engineering study with a complete performance simulation is needed to define the detailed configuration. The major technology developments needed are: high sensitivity, large format arrays of detectors; cold, lightweight mirrors for the telescopes and remote reflectors; formation flying with tethers; position measurement for the multiple spacecraft; active coolers for telescopes and detectors; and a smooth, long-stroke mirror scanning mechanism. There are many design choices to consider as well. Narrow band filters can reduce the spectral bandwidth and the detector noise. Dispersive spectrometers with large detector arrays could be combined with the imaging Michelson spectrometer to obtain improved sensitivity. There is no guarantee that the suggested optical configuration is close to optimal.
There is also little practical experience with image reconstruction from far IR detectors and spectral-spatial interferometers like SPECS. Simple laboratory tests should be done to wring out the problems in the analysis algorithms; those in use at the NRAO took decades of development, and although our requirements are similar, they are not identical.
We anticipate that the new generation of detectors will be some kind of superconducting device, either a superconducting transition edge (TES) bolometer with a multiplexed SQUID readout, or a superconducting junction, with photons converted to quasiparticles that can be collected and moved to the input of a single electron transistor. For the junction detectors, the following problems need to be addressed: finding efficient devices for collecting and converting photons to quasiparticles; transporting the quasiparticles to the amplifiers; running the amplifiers (current designs require RF excitation); multiplexing the detectors and electronics to enable large arrays; and building on-chip integrated circuitry. It is possible that an improved TES bolometer array will prove the best choice. In that case, developments to anticipate include: large format arrays, on-chip readout electronics using superconducting thermometers and SQUID readouts, and on-chip multiplexing electronics.
### 5.4 The Road Leading to SPECS
If a concerted effort is made to advance and test the required technologies (Table 3) during the next decade, it will be possible to build the SPECS observatory in about 15 years. A number of relevant efforts already underway (e.g., the DS-3 mission to test formation flying) require sustained support. Funding for efforts that require long lead times, such as advanced array detector and lightweight cryogenic optical system development, should be augmented as soon as possible. The investment needed to make the detectors a convenient and affordable reality is vital to the future of this subject. Early support will also be required for SPECS tradeoff and design studies.
A likely outcome of an evaluation of cost and risk is the recommendation to fly one or more “precursor” missions. The primary purpose of such missions would be to advance the technical frontiers in sensible increments toward the ultimate goal, SPECS. However, these missions would have potential scientific benefits as well. For example, a mission to test the detectors and cryo-coolers might conduct an unprecedented all-sky, confusion limited far IR/submillimeter survey. A dispersive spectrometer with a detector array could provide simultaneous spectra. A smaller version of the SPECS, say with 1 m mirrors, would still be remarkably sensitive, particularly if the photon counting detector arrays work out well and dispersive spectrometers are used to reduce the photon noise on each detector.
Also, the US should pursue relevant opportunities to partner with other nations. Japanese IR astronomers (Matsumoto and Okuda) have expressed interest in contributions to their H2L2 mission (see §4). The US could contribute, for example, improved detectors and a scanning Michelson interferometer, which could be used to conduct a deep spectroscopic survey at R $`10^3`$ (i.e., enough to reach the linewidths of typical galaxies). While this survey would still be confusion limited, objects at different redshifts could be distinguished by their spectral lines.
Any opportunity that arises to test technologies applicable to SPECS on a precursor mission required for NGST should also be exploited if it is deemed cost-effective.
## 6 Summary and Recommendations
A space-based far infrared/submillimeter imaging interferometer is needed to answer some of the most fundamental cosmological questions, those concerning the development of structure in the universe (§1). Such an instrument would be complementary to the already-planned Next Generation Space Telescope and the ground-based Millimeter Array (§2); it would allow access to a large number of important cooling and diagnostic spectral lines from ions, atoms, and molecules, and to the bulk of the thermal emission from dust clouds (§3). In light of recent technology developments, especially the possibility of background limited photon counting detectors in the far infrared, it is now practical to consider building the interferometer (§§4 and 5.2). The authors recommend that we set our sights on this goal and, over the course of the next decade, design, build and test the technology (§5.3) that will be needed to deploy a Submillimeter Probe of the Evolution of Cosmic Structure.
The authors thank Chuck Bennett, Alan Kogut, Simon Radford, Ramesh Sinha and Mark Swain for helpful comments and suggestions. We are grateful for and encouraged by the warm reception our SPECS concept has thus far received. MH and DS acknowledge funding support from NASA through grants NAG5-3347 and NAG5-7154, respectively.
## 7 References
> Acosta-Pulido, J. A., et al. 1998, A&A, in press (see http://www.ipac.caltech.edu/iso/AandA/I0026.html).
> Almaini, O., et al. 1998, Astronomische Nachrichten, 319, 55.
> Almaini, O., Lawrence, & Boyle, B. J. 1998, COSPAR talk, Nagoya, Japan.
> Armus, L., Matthews, K., Neugebauer, G., & Soifer, B. T. 1998, ApJ, 506, L89, astro-ph/9806243.
> Bahcall, J., et al. 1991, The Decade of Discovery in Astronomy and Astrophysics, Astronomy and Astrophysics Survey Committee, National Academy Press (http://www.nap.edu/bookstore/).
> Barger, A. J., et al. 1998, Nature, 394, 248, astro-ph/9806317.
> Blain, A. W., Ivison, R. J., & Smail, I. 1998, MNRAS, 296, L29, astro-ph/9710003.
> Dressler, A., et al. 1996, Exploration and the Search for Origins: A Vision for Ultraviolet-Optical-Infrared Space Astronomy, Report of the HST & Beyond Committee, (Washington, DC: AURA).
> Dwek, E. et al. 1998, ApJ, 508, 106, astro-ph/9806129.
> Fabian, A., et al. 1998, MNRAS, 297, L11.
> Fixsen, D., et al. 1998, ApJ, 508, 123, astro-ph/9803021.
> Genzel, R., et al. 1998, ApJ, 498, 579.
> Haarsma, D. B. & Partridge, R. B. 1998, ApJ, 503, L5, astro-ph/9806093.
> Harwit, M., Neufeld, D. A., Melnick, G. J. & Kaufman, M. J. 1998, ApJ, 497, L105, astro-ph/9802346.
> Hauser, M. G., et al. 1998, ApJ, 508, 25, astro-ph/9806167.
> Hughes, D., et al. 1998, Nature, 394, 241, astro-ph/9806297.
> Hughes, D. and Dunlop, J. 1997 “Using new submillimetre surveys to identify the evolutionary status of high-z galaxies,” in Observational Cosmology with New Radio Surveys, astro-ph/9707255.
> Irwin, K. 1995, Ph. D. thesis, Stanford University, p. 116f.
> Lee, A. T., et al. 1996, Appl. Phys. Lett. 69, 1801.
> Lee, A. T., Lee, S.-F., Gildemeister, J. M. & Richards, P. L. 1997, LTD-7 Conf. Proc., p. 123.
> Lee, S.-F., et al. 1998, Appl. Optics, 37, 3391.
> Lilly, S., et al. 1998, for ESA Conf. Proc. of 34th Liege International Astrophysics Colloquium “NGST: Science and Technological Challenges”, astro-ph/9807261.
> Luhman, M., et al. 1998, ApJ, 504, L11.
> Lutz, D., et al. 1998, ApJ, 505, L103.
> Madau, P., Pozzetti, L. & Dickinson, M. 1998, ApJ, 498, 106.
> Matsumoto, T. 1998, in Science with the NGST, E. P. Smith & A. Koratkar, eds., ASP Conf. Ser., v. 133, p. 63.
> Moorwood, A. F. M., et al. 1996, A&A, 315, L109.
> O’Connell, R. W. & McNamara, B. R. 1991, ApJ 393,597.
> Pearson, T. J., and Readhead, A. C. S. 1984, Ann. Rev. Astron. Astrophys. 22, 97.
> Prasad, S. & Kulkarni, S. R. 1989, Opt. Soc. Am. J. A, 6, 1702.
> Schoelkopf, R., et al. 1998, Science, 280, 1238.
> Soifer, B. T., Boehmer, L., Neugebauer, G., & Sanders, D. B. 1989, AJ, 98, 766.
> Stacey, G. J., et al. 1991, ApJ, 393, 423.
> Voit, M. 1992, ApJ, 339, 495.
> Web site for SPECS:
> http://www.gsfc.nasa.gov/astro/specs
|
no-problem/9812/hep-th9812115.html
|
ar5iv
|
text
|
# A Note on ADE-Spectra in Conformal Field Theory
## I Introduction
footnotetext: e-mail addresses: bytsko@pdmi.ras.ru, fring@physik.fu-berlin.de.
It is well known, that a large class of off-critical integrable models is related to affine Toda field theories or RSOS-statistical models , which posses a rich underlying Lie algebraic structure. Since these models can be regarded as perturbed conformal field theories, it is suggestive to recover the underlying Lie algebraic structure also in the conformal limit. Of primary interest is to identify the conformal counterparts of the off-critical particle spectrum. One way to achieve this is to analyze the quasi-particle spectrum, which results from certain expressions of the Virasoro characters $`\chi (q)`$ or their linear combinations. Hitherto this analysis was mainly performed for formulae of the form
$$\chi (q)=q^{\mathrm{const}}\underset{\stackrel{}{l}}{}\frac{q^{\stackrel{}{l}^tA\stackrel{}{l}+\stackrel{}{B}\stackrel{}{l}}}{(q)_{l_1\mathrm{}}(q)_{l_r}}.$$
(1)
Here $`r`$ is the rank of the related Lie algebra g, the matrix $`A`$ coincides with the inverse of the Cartan matrix, $`\stackrel{}{B}`$ characterizes the super-selection sector, $`(q)_l:=_{k=1}^l(1q^k)`$, and there may be certain restrictions on the summation over $`\stackrel{}{l}`$. Following the prescription of one can always obtain a quasi-particle spectrum once a character admits a representation in the form of equation (1). It should be noted that such spectra can not be obtained form the standard form of the Virasoro characters (10).
In the following we will demonstrate that one also recovers Lie algebraic structures in certain Virasoro characters or their linear combinations which admit the factorized form
$$\frac{q^{\mathrm{const}}}{\left\{1\right\}_1^{}}\{x_1;\mathrm{};x_N\}_y^{}\{x_1^{};\mathrm{};x_M^{}\}_y^+,$$
(2)
where we adopt the notations of
$`\{x\}_y^\pm :={\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}\left(1\pm q^{x+ky}\right),\{x_1;\mathrm{};x_n\}_y^\pm :={\displaystyle \underset{a=1}{\overset{n}{}}}\{x_a\}_y^\pm .`$
In many cases (see for details) expressions of the type (2) can be rewritten in the form (1), but now $`A`$ is entirely absent or, at most, is a diagonal matrix. There are no restrictions on the summation over $`\stackrel{}{l}`$, and we allow terms of the form $`(q^y)_l`$ in the denominator (which may be regarded as an anionic feature ).
Unlike the conventional form for the Virasoro characters (10), formulae (1) and (2) allow to extract the leading order behaviour in the limit $`q1^{}`$ by means of a saddle point analysis, see e.g. . For a slightly generalized version of (1), in the sense that all $`(q)_l`$ are replaced by $`(q^y)_l`$, this analysis leads to
$$z_i^y=\underset{j=1}{\overset{r}{}}(1z_j)^{(A_{ij}+A_{ji})},c_{\mathrm{eff}}=\frac{6}{y\pi ^2}\underset{i=1}{\overset{r}{}}L(z_i).$$
(3)
This means solving the former set of equations for the unknown quantities $`z_i`$, we may compute the effective central charge thereafter by means of the latter equation in terms of Rogers dilogarithm $`L(x)`$. Recall that the effective central charge is defined as $`c_{\mathrm{eff}}=c24h^{}`$, where $`h^{}`$ is the lowest conformal weight occurring in the model. There exist inequivalent solutions to equations (3) leading to the same effective central charge corresponding either to the form (1) or (2). When treating these equations as formal series, such computations give a first hint on possible candidates for characters.
Alternatively, with regard to factorization, we can exploit the essential fact that the blocks $`\{x\}_y^\pm `$ are closely related to the so-called quantum dilogarithm and we can easily compute their contributions to the effective central charge. As explained in , each block $`(\{x\}_y^{})^{\pm 1}`$ and $`(\left\{x\right\}_y^+)^{\pm 1}`$ in expressions of type (2) contributes
$$\mathrm{\Delta }c_{\mathrm{eff}}=\begin{array}{c}\frac{1}{y}\end{array},\mathrm{and}\mathrm{\Delta }\mathrm{c}_{\mathrm{eff}}=\pm \begin{array}{c}\frac{1}{2\mathrm{y}}\end{array},$$
(4)
respectively. In the course of our argument, i.e. when we consider the difference of the Virasoro characters, we will also need the notion of the secondary effective central charge
$$\stackrel{~}{c}=124h^{\prime \prime },$$
(5)
where $`h^{\prime \prime }`$ is the next to lowest conformal weight occurring in the model.
## II ADE-Structure
Let g be a Lie algebra of rank $`r`$ and $`h`$ be its Coxeter number. We define the following function related to g
$$\mathrm{\Xi }^𝐠(\stackrel{}{x},q)=\frac{q^{\mathrm{const}}}{\{x_1;\mathrm{};x_r\}_{\frac{h}{2}+1}^{}},$$
(6)
with $`\stackrel{}{x}`$ obeying the condition
$$x_a+x_{r+1a}=h/2+1,a=1,\mathrm{},r,$$
(7)
and for odd $`r`$ we put $`x_{\frac{r+1}{2}}=\frac{h}{4}+\frac{1}{2}`$. Our aim is to find conformal models such that their characters or possibly linear combinations coincide with (6) for appropriately chosen $`q^{\mathrm{const}}`$ and $`\stackrel{}{x}`$. Such conformal models have quasi-particle spectra, generated by (6) for the related sectors, with the number of different particle species equal to the rank $`r`$.
The question arises for which conformal models can we expect (6) to be a character? Exploiting (4), we readily find the corresponding effective central charge
$$c_{\mathrm{eff}}(𝐠)=\frac{2r}{h+2}=\frac{2r^2}{\mathrm{dim}𝐠+1}.$$
(8)
On the other hand, the analysis of the ultra-violet limit of the thermodynamic Bethe ansatz for the ADE related minimal scattering matrices of affine Toda field theory leads to the following effective central charges:
| g | $`A_n^{(1)}`$ | $`D_n^{(1)}`$ | $`E_6^{(1)}`$ | $`E_7^{(1)}`$ | $`E_8^{(1)}`$ | $`A_{2n}^{(2)}`$ |
| --- | --- | --- | --- | --- | --- | --- |
| c<sub>eff</sub> | $`\frac{2n}{n+3}`$ | $`1`$ | $`\frac{6}{7}`$ | $`\frac{7}{10}`$ | $`\frac{1}{2}`$ | $`\frac{2n}{2n+3}`$ |
Table 1. Effective central charges for affine Toda field theories.
Thus, we see that, upon substitution of the related Lie algebraic quantities<sup>*</sup><sup>*</sup>*For a twisted affine Lie algebra of type $`X_N^{(k)}`$ one introduces $`h`$ – the Coxeter number and $`h^{(k)}=kh`$. We should use $`h`$ in (8) for $`A_{2n}^{(2)}`$. of the simply laced algebras (see e.g. ), eq. (8) recovers all the effective central charges in Table 1. Furthermore it turns out that for g from this table corresponding to minimal models or $`c=1`$ models we are always able to identify several $`\mathrm{\Xi }^𝐠(\stackrel{}{x},q)`$ with single Virasoro characters or specific linear combinations of them.
In addition, there exist characters which exhibit even stronger Lie algebraic features. They are given by (6) with the values of $`x_a`$ chosen as follows (which is a particular case of (7))
$$2x_a1=e_a,a=1,\mathrm{},r,$$
(9)
where $`\{e_a\}`$ stands for the set of the exponents of the Lie algebra g. We denote this particular character as $`\mathrm{\Xi }^𝐠(q)`$.
### A Minimal Models
The minimal models are parameterized by a pair $`(s,t)`$ of co-prime positive integers and the corresponding central charge is $`c(s,t)=1\frac{6(st)^2}{st}`$. Labeling the highest weights as $`h_{n,m}^{s,t}=\frac{(ntms)^2(st)^2}{4st}`$, with the restrictions $`1ns1`$ and $`1mt1`$, the usual form of the characters of irreducible highest weight representations reads
$$\chi _{n,m}^{s,t}(q)=\eta _{n,m}^{s,t}\underset{k=\mathrm{}}{\overset{\mathrm{}}{}}q^{stk^2}\left(q^{k(ntms)}q^{k(nt+ms)+nm}\right).$$
(10)
Here we abbreviated the ubiquitous factor $`\eta _{n,m}^{s,t}:=q^{h_{n,m}^{s,t}\frac{c(s,t)}{24}}/\left\{1\right\}_1^{}`$ by an analogy with the eta-function. The (secondary) effective central charge is easy to find (see e.g. )
$$c_{\mathrm{eff}}(s,t)=1\frac{6}{st},\stackrel{~}{c}(s,t)=1\frac{24}{st}.$$
(11)
Thus, the values of $`c_{\mathrm{eff}}`$ which are less than 1 can be matched as follows
$`c_{\mathrm{eff}}(A_1^{(1)})=c_{\mathrm{eff}}(E_8^{(1)})=c_{\mathrm{eff}}(3,4),`$ (12)
$`c_{\mathrm{eff}}(A_2^{(1)})=c_{\mathrm{eff}}(5,6)=c_{\mathrm{eff}}(3,10)=c_{\mathrm{eff}}(2,15),`$ (13)
$`c_{\mathrm{eff}}(E_6^{(1)})=c_{\mathrm{eff}}(6,7)=c_{\mathrm{eff}}(3,14)=c_{\mathrm{eff}}(2,21),`$ (14)
$`c_{\mathrm{eff}}(E_7^{(1)})=c_{\mathrm{eff}}(4,5),`$ (15)
$`c_{\mathrm{eff}}(A_{2n}^{(2)})=c_{\mathrm{eff}}(2,2n+3).`$ (16)
We see, that the matching is, in general, not unique. For (16) it depends on $`n`$ – the first non-unique representations occur for $`n=6`$ and $`n=9`$ and coincide with (13) and (14), respectively. Therefore we might have for instance relations between $`A_2^{(1)}A_{12}^{(2)}`$ and $`E_6^{(1)}A_{18}^{(2)}`$. Some of these apparent ambiguities are easily explained as the consequence of a symmetry property of the characters. For instance we observe that eq. (10) possesses the symmetry: $`\chi _{\alpha n,m}^{\alpha s,t}(q)=\chi _{n,\alpha m}^{s,\alpha t}(q)`$, for instance $`\chi _{2,m}^{6,5}(q)=\chi _{1,2m}^{3,10}(q)`$.
Of course eqs. (12)-(16) are only to be understood as a first hint on a possibility for characters in the corresponding models to be of the form $`\mathrm{\Xi }^𝐠(\stackrel{}{x},q)`$. In order to make the identifications more precise, we have to resort to more stringent properties of the characters. We shall be using previously obtained results on representation of characters of minimal models in the form (2). In it was proven, that for $`M=0`$ and $`x_ix_j`$ for $`ij`$ in (2) the only possible factorizable single characters are
$`\chi _{n,m}^{2n,t}(q)`$ $`=`$ $`\eta _{n,m}^{2n,t}\{nm;nt;ntnm\}_{nt}^{},`$ (17)
$`\chi _{n,m}^{3n,t}(q)`$ $`=`$ $`\eta _{n,m}^{3n,t}\{2nt;nm;2ntnm\}_{2nt}^{}`$ (19)
$`\times \{2nt2nm;2nt+2nm\}_{4nt}^{}.`$
Combining now characters in a linear way, the property to be factorizable remains still exceptional. It was argued in that
$$\chi _{n,m}^{s,t}(q)\pm \chi _{n,tm}^{s,t}(q)$$
(20)
are the only combinations of characters in the same model which have a chance to acquire the form (2) with reasonably small $`N`$ and $`M`$. The limit $`q1^{}`$ of the upper and lower signs in (20) is governed by $`c_{\mathrm{eff}}`$ and $`\stackrel{~}{c}`$, respectively, which can be seen form their properties with respect to the S-modular transformation .
The following factorizable combinations of type (20) where found in
$`\chi _{n,m}^{3n,t}(q)`$ $`\pm `$ $`\chi _{2n,m}^{3n,t}(q)=\eta _{n,m}^{3n,t}\{nm;ntnm\}_{nt}^{}`$ (23)
$`\times \begin{array}{c}\left\{\frac{nt}{2}\right\}_{\frac{nt}{2}}^{}\{\frac{nt2nm}{4};\frac{nt+2nm}{4}\}_{\frac{nt}{2}}^\pm \end{array},`$
$`\chi _{n,m}^{4n,t}(q)`$ $``$ $`\chi _{3n,m}^{4n,t}(q)=\eta _{n,m}^{4n,t}\begin{array}{c}\{\frac{nt}{2};\frac{nt}{2}nm;nm\}_{\frac{nt}{2}}^{}\end{array},`$ (25)
$`\chi _{n,m}^{4n,t}(q)`$ $`+`$ $`\chi _{3n,m}^{4n,t}(q)=\eta _{n,m}^{4n,t}\{nm;nt;ntnm\}_{nt}^{}`$ (28)
$`\times \begin{array}{c}\{\frac{nt}{2}nm;\frac{nt}{2}+nm;\frac{nt}{2}\}_{nt}^+\end{array},`$
$`\chi _{n,m}^{6n,t}(q)`$ $``$ $`\chi _{5n,m}^{6n,t}(q)=\eta _{n,m}^{6n,t}\{nt;nm;ntnm\}_{nt}^{}`$ (30)
$`\times \{nt2nm;nt+2nm\}_{2nt}^{}.`$
For the differences of type (20) it was proven that besides the presented formulae no other combinations are possible to factorize in the form (2) with $`M=0`$ and $`x_ix_j`$ for $`ij`$.
Remarkably, as we demonstrate in Table 2, all identifications presented in (12)-(16) can be realized in terms of characters with the help of (17)-(30). In particular, we identify $`\mathrm{\Xi }^𝐠(q)`$ with the following characters
$`\mathrm{\Xi }^{A_1^{(1)}}(q)=\chi _{1,2}^{3,4}(q),\mathrm{\Xi }^{A_{2n}^{(2)}}(q)=\chi _{1,n+1}^{2,2n+3}(q),`$ (31)
$`\mathrm{\Xi }^{A_2^{(1)}}(q)=\chi _{1,2}^{5,6}(q)+\chi _{1,4}^{5,6}(q)=\chi _{1,2}^{3,10}(q)+\chi _{1,8}^{3,10}(q),`$ (32)
$`\mathrm{\Xi }^{E_6^{(1)}}(q)=\chi _{2,1}^{6,7}(q)+\chi _{2,6}^{6,7}(q)=\chi _{1,2}^{3,14}(q)+\chi _{1,12}^{3,14}(q),`$ (33)
$`\mathrm{\Xi }^{E_7^{(1)}}(q)=\chi _{2,1}^{4,5}(q),\mathrm{\Xi }^{E_8^{(1)}}(q)=\chi _{1,3}^{3,4}(q).`$ (34)
Since these identifications hint on the connection with massive models, i.e. affine Toda field theories, it is somewhat surprising that also the non-simply laced algebras $`G_2`$ and $`F_4`$ occur in Table 2.We thank W. Eholzer for pointing out to us that our formulae in may include $`\mathrm{\Xi }^{G_2}(q)=\chi _{1,2}^{3,4}(q)`$ and $`\mathrm{\Xi }^{F_4}(q)=\chi _{1,2}^{2,7}(q)`$ as well. No connection is known between $`G_2`$ and $`F_4`$affine Toda models and $`(3,4)`$ and $`(2,7)`$ minimal models, respectively.Notice also that a substitution of the inverse Cartan matrix for $`G_2`$ into (1) yields in the standard $`q1^{}`$ analysis $`c_{\mathrm{eff}}0.453`$. At present it seems to be just an intriguing coincidence. Replacing in (6) the Coxeter number by the dual Coxeter number would exclude the non-simply laced cases without altering the formulae for the simply laced ones.
It is interesting to notice that, as seen from Table 2, the differences of type (30) can also be of the form
$$\frac{q^{\mathrm{const}}}{\{x_1;\mathrm{};x_r\}_b^{}}.$$
(35)
For instance, $`\chi _{1,1}^{5,6}(q)\chi _{1,5}^{5,6}(q)`$ corresponds to $`b=2h+4`$, $`x_a=2(e_a+1)`$, and $`\chi _{1,2}^{6,7}(q)\chi _{1,5}^{6,7}(q)`$ corresponds to $`b=h+2`$, $`x_a=e_a+1`$, where $`h`$ and $`\{e_a\}`$ are the Coxeter number and exponents of $`A_2`$ and $`E_6`$, respectively.
| g | $`\mathrm{sectors}`$ | $`\{x_1;\mathrm{};x_r\}_y^{}`$ |
| --- | --- | --- |
| $`E_8`$ | $`\chi _{1,1}^{3,4}`$ | $`\{2;3;4;5;11;12;13;14\}_{16}^{}`$ |
| | $`\chi _{1,2}^{3,4}`$ | $`\{1;3;5;7;9;11;13;15\}_{16}^{}`$ |
| | $`\chi _{1,3}^{3,4}`$ | $`\{1;4;6;7;9;10;12;15\}_{16}^{}`$ |
| $`A_1`$ | $`\chi _{1,2}^{3,4}`$ | $`\{1\}_2^{}`$ |
| $`E_7`$ | $`\chi _{2,1}^{4,5}`$ | $`\{1;3;4;5;6;7;9\}_{10}^{}`$ |
| | $`\chi _{2,2}^{4,5}`$ | $`\{1;2;3;5;7;8;9\}_{10}^{}`$ |
| $`A_2`$ | $`\chi _{1,2}^{5,6}+\chi _{1,4}^{5,6}`$ | $`\{1;\frac{3}{2}\}_{5/2}^{}`$ |
| | $`\chi _{2,2}^{5,6}+\chi _{2,4}^{5,6}`$ | $`\{\frac{1}{2};2\}_{5/2}^{}`$ |
| | $`\chi _{1,1}^{5,6}\chi _{1,5}^{5,6}`$ | $`\{2;8\}_{10}^{}`$ |
| | $`\chi _{2,1}^{5,6}\chi _{2,5}^{5,6}`$ | $`\{4;6\}_{10}^{}`$ |
| $`E_6`$ | $`\chi _{2,1}^{6,7}+\chi _{4,1}^{6,7}`$ | $`\{1;\frac{\mathrm{𝟓}}{\mathrm{𝟐}};3;4;\frac{\mathrm{𝟗}}{\mathrm{𝟐}};6\}_7^{}`$ |
| | $`\chi _{2,2}^{6,7}+\chi _{4,2}^{6,7}`$ | $`\{1;\frac{\mathrm{𝟑}}{\mathrm{𝟐}};2;5;\frac{\mathrm{𝟏𝟏}}{\mathrm{𝟐}};6\}_7^{}`$ |
| | $`\chi _{2,3}^{6,7}+\chi _{4,3}^{6,7}`$ | $`\{\frac{\mathrm{𝟏}}{\mathrm{𝟐}};2;3;4;5;\frac{\mathrm{𝟏𝟑}}{\mathrm{𝟐}}\}_7^{}`$ |
| | $`\chi _{1,1}^{6,7}\chi _{1,6}^{6,7}`$ | $`\{2;3;4;10;11;12\}_{14}^{}`$ |
| | $`\chi _{1,2}^{6,7}\chi _{1,5}^{6,7}`$ | $`\{1;4;6;8;10;13\}_{14}^{}`$ |
| | $`\chi _{1,3}^{6,7}\chi _{1,4}^{6,7}`$ | $`\{2;5;6;8;9;12\}_{14}^{}`$ |
| $`G_2`$ | $`\chi _{1,2}^{3,4}`$ | $`\{1;3\}_4^{}`$ |
| $`F_4`$ | $`\chi _{1,1}^{2,7}`$ | $`\{2;3;4;5\}_7^{}`$ |
| | $`\chi _{1,2}^{2,7}`$ | $`\{1;3;4;6\}_7^{}`$ |
| | $`\chi _{1,3}^{2,7}`$ | $`\{1;2;5;6\}_7^{}`$ |
Table 2. Representation of characters in the form of (6) and for differences of the type (30) in the form (35). The replacement of the blocks $`\{x\}^{}`$ typed in bold by $`\{x\}^+`$ yields the corresponding differences of characters.
Eq. (6) does not exhaust all manifest Lie algebraic functions in which combinations of characters of the type (20) can be represented. For instance, the following representation
$$\overline{\mathrm{\Xi }}^𝐠(\stackrel{}{w},q)=q^{\mathrm{const}}\frac{\left\{\frac{h+2}{8}\right\}_{\frac{h+2}{4}}^+}{\{w_1;\mathrm{};w_{r1}\}_{\frac{h}{2}+1}^{}}$$
(36)
with $`\stackrel{}{w}`$ obeying the condition $`w_a+w_{ra}=h/2+1`$ also occurs (see Table 3).
| g | $`\mathrm{sectors}`$ | $`\{w_1;\mathrm{};w_{r1}\}_y^{}`$ |
| --- | --- | --- |
| $`A_1`$ | $`\chi _{1,1}^{3,4}+\chi _{1,3}^{3,4}`$ | 1 |
| $`E_7`$ | $`\chi _{1,1}^{4,5}+\chi _{1,4}^{4,5}`$ | $`\{\frac{\mathrm{𝟑}}{\mathrm{𝟐}};2;\frac{\mathrm{𝟕}}{\mathrm{𝟐}};\frac{\mathrm{𝟏𝟑}}{\mathrm{𝟐}};8;\frac{\mathrm{𝟏𝟕}}{\mathrm{𝟐}}\}_{10}^{}`$ |
| | $`\chi _{1,2}^{4,5}+\chi _{1,3}^{4,5}`$ | $`\{\frac{\mathrm{𝟏}}{\mathrm{𝟐}};4;\frac{\mathrm{𝟗}}{\mathrm{𝟐}};\frac{\mathrm{𝟏𝟏}}{\mathrm{𝟐}};6;\frac{\mathrm{𝟏𝟗}}{\mathrm{𝟐}}\}_{10}^{}`$ |
Table 3. Representation of characters in the form of (36). For the arguments in bold the same convention applies as in table 2, including the numerator.
### B Compactified free Boson
In general the Fock space of a free boson may simply be constructed from a Heisenberg algebra and the corresponding Virasoro central charge equals $`c=1`$. The character of the Heisenberg module ($`\widehat{U}(1)`$-Kac-Moody) is simply the inverse of the $`\eta `$-function, $`q^{1/24}/\left\{1\right\}_1^{}`$. When compactifying the boson on a circle of rational square radius $`R=\sqrt{2s/t}`$ one can associate highest weight representations of a $`\widehat{U}(1)_k`$-Kac-Moody algebra to this theory. The $`\widehat{U}(1)_k`$-algebra has an integer level, which is $`k=st`$. The corresponding characters read
$$\widehat{\chi }_m^k(q)=\widehat{\eta }_m^k\underset{l=\mathrm{}}{\overset{\mathrm{}}{}}q^{kl^2+ml}=\widehat{\eta }_m^k\{2k\}_{2k}^{}\{km;k+m\}_{2k}^+.$$
(37)
where we denoted $`\widehat{\eta }_m^k:=q^{h_m^k\frac{1}{24}}/\left\{1\right\}_1^{}`$. The highest weight may take on the values $`h_m^k=\frac{m^2}{4k}`$ with $`m=0,1,\mathrm{},k1`$.
As Table 1 indicates, the $`D_n`$-affine Toda models are related to compactified bosons. In order to recover the $`D_n`$-structure at the conformal level, we shall find realizations of $`\mathrm{\Xi }^{D_n}(q)`$ in terms of (37). Similar to the case of minimal models, it will be helpful to study factorization of (combinations of) the Kac-Moody characters.
First, choosing the constant in (6) as $`h_{n/2}^n\begin{array}{c}\frac{1}{24}\end{array}`$ we identify for even $`n`$
$`\mathrm{\Xi }^{D_n}(q)`$ $`=`$ $`\widehat{\eta }_{n/2}^n{\displaystyle \frac{\{n\}_n^{}}{\left\{\frac{n}{2}\right\}_n^{}}}=\widehat{\eta }_{n/2}^n\{n\}_n^{}\begin{array}{c}\left\{\frac{n}{2}\right\}_{\frac{n}{2}}^+\end{array}`$ (39)
$`=`$ $`\widehat{\eta }_{n/2}^n\{2n\}_{2n}^{}\begin{array}{c}\left\{\frac{n}{2}\right\}_n^+\end{array}=\widehat{\chi }_{n/2}^n(q).`$ (41)
Formally this expression also holds for odd $`n`$, albeit in this case the right hand side may not be interpreted as the character related to a compactified boson. Therefore we need another way to construct $`\mathrm{\Xi }^{D_n}(q)`$ in case $`n`$ is odd. For this purpose we consider the combinations $`\widehat{\chi }_m^n(q)\pm \widehat{\chi }_{nm}^n(q)`$ which are analogues of (20). In particular, the $`q1^{}`$ limit of these sums and differences is governed by $`c_{\mathrm{eff}}=c=1`$ and $`\stackrel{~}{c}`$. According to (5) and the possible values for the highest weights we have $`\stackrel{~}{c}=16/n`$.
Exploiting the identity (2.31) obtained in , we find (for $`0<m<n/2`$)
$`\widehat{\chi }_m^n(q)\pm \widehat{\chi }_{nm}^n(q)=\widehat{\eta }_m^n\begin{array}{c}\left\{\frac{n}{2}\right\}_{\frac{n}{2}}^{}\{\frac{n}{4}\frac{m}{2};\frac{n}{4}+\frac{m}{2}\}_{\frac{n}{2}}^\pm \end{array}.`$ (43)
The counting, based on (4), gives the expected values of $`c`$ and $`\stackrel{~}{c}`$. Notice that for the upper sign the r.h.s. can be identified as the product side of (37):
$$\widehat{\chi }_m^n(q)+\widehat{\chi }_{nm}^n(q)=\widehat{\chi }_{m/2}^{n/4}(q).$$
(44)
Comparison with (41) then yields a formula for $`\mathrm{\Xi }^{D_n}(q)`$ valid for both odd and even $`n`$
$$\mathrm{\Xi }^{D_n}(q)=\widehat{\chi }_n^{4n}(q)+\widehat{\chi }_{3n}^{4n}(q).$$
(45)
Finally, it is interesting to observe that, employing (17)-(23), we may express (37) and (43) entirely in terms of the minimal Virasoro characters:
$`\widehat{\chi }_m^n(q)`$ $`=`$ $`\chi _{1,m}^{3,n}(q){\displaystyle \frac{\chi _{1,n}^{2,3n}(q)}{\chi _{1,m}^{2,n}(q)}},`$ (46)
$`\widehat{\chi }_m^n(q)\pm \widehat{\chi }_{nm}^n(q)`$ $`=`$ $`\left(\chi _{1,m}^{3,n}(q)\pm \chi _{2,m}^{3,n}(q)\right){\displaystyle \frac{\chi _{1,n}^{2,3n}(q)}{\chi _{1,m}^{2,n}(q)}}.`$ (47)
Here $`n=6l\pm 1`$, $`lN`$ if we really regard all components on the r.h.s. as characters of irreducible Virasoro representations. This restriction can be omitted if we regard (17)-(23) just as formal series. With regard to the central charge eqs. (46)-(47) imply
$$c_{\mathrm{eff}}(D_n^{(1)})=c_{\mathrm{eff}}(3,n)+c_{\mathrm{eff}}(2,3n)c_{\mathrm{eff}}(2,n).$$
(48)
Thus, the connection with minimal models is more subtle than one would expect at first sight from a simple matching of the central charges, e.g. $`c_{\mathrm{eff}}(D_n^{(1)})=2c_{\mathrm{eff}}(3,4)`$.
### C Parafermions
The $`A_n^{(1)}`$-series of affine Toda theories is known to be related in the ultra-violet limit (see e.g. ) to the $`Z_{n+1}`$-parafermions . The corresponding central charge, $`c(k)=2(k1)/(k+2)`$ and characters may be obtained from the $`\widehat{SU}(2)_{k1}/\widehat{U}(1)`$-coset, where $`k=n+1`$. Introducing the quantity $`\mathrm{\Delta }_{j,m}^k=j(j+1)/(k+2)m^2/k`$ the characters of the highest weight representation, which appear as branching functions in the coset, acquire the form
$`\stackrel{~}{\chi }_{j,m}^k(q)={\displaystyle \frac{\stackrel{~}{\eta }_{j,m}^k}{\{1\}_1^{}}}{\displaystyle \underset{r,s=0}{\overset{\mathrm{}}{}}}(1)^{r+s}q^{rs(k+1)+\frac{r(r+1)}{2}+\frac{s(s+1)}{2}}`$ (49)
$`\times \left(q^{r(j+m)+s(jm)}q^{r(k+1jm)+s(k+1j+m)+k+12j}\right),`$ (50)
where $`\stackrel{~}{\eta }_{j,m}^k:=q^{\mathrm{\Delta }_{j,m}^k\frac{c(k)}{24}}/\{1\}_1^{}`$. The labels are restricted as $`jmkj`$, $`0jk/2`$ and $`(jm)Z`$. In particular the $`\stackrel{~}{\chi }_{0,m}^k(q)`$ are the characters of the parafermionic currents $`\psi _m^k`$. The characters possess the symmetries
$$\stackrel{~}{\chi }_{j,m}^k(q)=\stackrel{~}{\chi }_{j,m}^k(q)=\stackrel{~}{\chi }_{k/2j,k/2m}^k(q).$$
(51)
From our observations made above for characters of the ADE-related conformal models one may expect that expressions (49) exhibit $`A_n`$-type structures (e.g. posses $`n`$ quasi particles and moreover acquire the form of the type $`\mathrm{\Xi }^{A_n}(q)`$) in some of the cases when they admit a factorized form. We shall now discuss this issue in detail for several of the lowest ranks.
As we have seen above, factorization of linear combinations of characters occurs usually only for the specific type of combinations. Now the analogue of (20) is $`\stackrel{~}{\chi }_{j,m}^k(q)\pm \stackrel{~}{\chi }_{j,2km}^k(q)`$. One expects that the $`q1^{}`$ limit of these sums and differences is governed by $`c_{\mathrm{eff}}=c(k)`$ and $`\stackrel{~}{c}`$, respectively. Since $`h^{\prime \prime }=\mathrm{\Delta }_{1,1}^k`$, eq. (5) yields $`\stackrel{~}{c}=c(k)(k6)/k`$. This is confirmed by all the examples given below.
For the parafermionic formulae (49) we do not have such powerful analytical tools (analogues to the factorization formulae (17)-(30) ) at hand as in the case of the minimal models. Therefore, as a first step, we resort to an analysis with Mathematica. Typically we expand the characters up to $`q^{100}`$.
$`\underset{¯}{A_1}:`$ In this case there are only three distinct (up to the symmetries (51)) characters and they can be matched with those of the (3,4) minimal model:
$`\stackrel{~}{\chi }_{0,0}^2(q)=\chi _{1,1}^{3,4}(q),\stackrel{~}{\chi }_{0,1}^2(q)=\chi _{1,3}^{3,4}(q),`$ (52)
$`\stackrel{~}{\chi }_{\frac{1}{2},\frac{1}{2}}^2(q)=\chi _{1,2}^{3,4}(q)=\mathrm{\Xi }^{A_1}(q).`$ (53)
Thus, all the characters in this case are factorizable and moreover $`\mathrm{\Xi }^{A_1}(q)`$ is present among them.
$`\underset{¯}{A_2}:`$ There are four distinct characters in this case and they can be matched with those of the (5,6) minimal model:
$`\stackrel{~}{\chi }_{0,0}^3(q)=\chi _{1,1}^{5,6}(q)+\chi _{1,5}^{5,6}(q),\stackrel{~}{\chi }_{0,1}^3(q)=\chi _{1,3}^{5,6}(q),`$ (54)
$`\stackrel{~}{\chi }_{\frac{1}{2},\frac{1}{2}}^3(q)=\chi _{2,3}^{5,6}(q),\stackrel{~}{\chi }_{\frac{1}{2},\frac{3}{2}}^3(q)=\chi _{2,1}^{5,6}(q)+\chi _{2,5}^{5,6}(q).`$ (55)
Only $`\stackrel{~}{\chi }_{0,1}^3(q)`$ and $`\stackrel{~}{\chi }_{\frac{1}{2},\frac{1}{2}}^3(q)`$ are factorizable (see subsection II.A).
$`\underset{¯}{A_3}:`$ Since $`c=1`$, it is suggestive to try to relate the characters to those of the compactified bosons. This turns out to be possible for all the characters (thus factorizability is guaranteed):
$`\stackrel{~}{\chi }_{0,1}^4(q)=\widehat{\chi }_6^{12}(q),`$ $`\stackrel{~}{\chi }_{0,0}^4(q)+\stackrel{~}{\chi }_{0,2}^4(q)=\widehat{\chi }_0^3(q),`$ (56)
$`\stackrel{~}{\chi }_{1,0}^4(q)=\widehat{\chi }_2^3(q),`$ $`\stackrel{~}{\chi }_{1,1}^4(q)=\widehat{\chi }_1^3(q),`$ (57)
$`\stackrel{~}{\chi }_{\frac{1}{2},\frac{1}{2}}^4(q)=\widehat{\chi }_1^4(q),`$ $`\stackrel{~}{\chi }_{\frac{1}{2},\frac{3}{2}}^4(q)=\widehat{\chi }_3^4(q).`$ (58)
Furthermore, some of linear combinations can be expressed in terms of the characters of the (3,4) minimal model (notice that $`c=1`$, $`\stackrel{~}{c}=1/2`$ for $`A_3`$ and $`c=1/2`$, $`\stackrel{~}{c}=1`$ for the (3,4) minimal model):
$`\stackrel{~}{\chi }_{0,0}^4(q)\stackrel{~}{\chi }_{0,2}^4(q)=\left(\chi _{1,2}^{3,4}(q)\right)^1,`$ (59)
$`\stackrel{~}{\chi }_{\frac{1}{2},\frac{1}{2}}^4(q)\pm \stackrel{~}{\chi }_{\frac{1}{2},\frac{3}{2}}^4(q)=\left(\chi _{1,1}^{3,4}(q)\chi _{1,3}^{3,4}(q)\right)^1.`$ (60)
$`\underset{¯}{A_4}:`$ No characters or linear combinations factorize.
$`\underset{¯}{A_5}:`$ Several characters and combinations are factorizable and can be expressed via those of the (3,4) minimal model and $`D_n`$, for instance
$`\stackrel{~}{\chi }_{\frac{3}{2},\frac{1}{2}}^6(q)`$ $`=`$ $`\chi _{1,2}^{3,4}(q)\left(\widehat{\chi }_6^{24}(q)\widehat{\chi }_{18}^{24}(q)\right),`$
$`\stackrel{~}{\chi }_{\frac{1}{2},\frac{3}{2}}^6(q)`$ $`=`$ $`\chi _{1,2}^{3,4}(q)\left(\widehat{\chi }_8^{24}(q)\widehat{\chi }_{16}^{24}(q)\right),`$
$`\stackrel{~}{\chi }_{0,1}^6(q)\pm \stackrel{~}{\chi }_{0,2}^6(q)`$ $`=`$ $`(\chi _{1,1}^{3,4}(q)\pm \chi _{1,3}^{3,4}(q))(\widehat{\chi }_9^{24}(q)\widehat{\chi }_{15}^{24}(q)),`$
$`\stackrel{~}{\chi }_{1,1}^6(q)\pm \stackrel{~}{\chi }_{1,2}^6(q)`$ $`=`$ $`(\chi _{1,1}^{3,4}(q)\pm \chi _{1,3}^{3,4}(q))(\widehat{\chi }_3^{24}(q)\widehat{\chi }_{21}^{24}(q)).`$
Also we notice that $`\stackrel{~}{\chi }_{\frac{1}{2},\frac{1}{2}}^6(q)\stackrel{~}{\chi }_{\frac{1}{2},\frac{5}{2}}^6(q)=q^{5/96}`$. Such an identity can occur only in this parafermionic model since it requires $`\stackrel{~}{c}=0`$.
$`\underset{¯}{A_6}:`$ – no combinations factorize and the only factorizable single characters are
$`\stackrel{~}{\chi }_{1,m}^7(q)`$ $`=`$ $`\stackrel{~}{\eta }_{1,m}^7{\displaystyle \frac{\{3\}_3^{}\{m;7m;7\}_7^{}}{\{1\}_1^{}\{3m;213m\}_{21}^{}}},m=1,2,3.`$ (61)
Summarizing these data, we see that, apart from the $`A_1`$ case, none of the factorizable (combinations of) characters provided by eq. (49) for $`A_n`$ can be identified as $`\mathrm{\Xi }^{A_n}(\stackrel{}{x},q)`$. However, it is plausible to speculate that in general the $`\mathrm{\Xi }^{A_n}(\stackrel{}{x},q)`$ might be identifiable as characters of other conformal models having the central charge $`2n/(n+3)`$. This conjecture is supported by the $`A_2`$ case (see Table 2) and $`A_3`$ case, in which we can identify
$$\mathrm{\Xi }^{A_3}(q)=\widehat{\chi }_3^{12}(q)+\widehat{\chi }_9^{12}(q).$$
(62)
To conclude the discussion on factorizable parafermionic characters, we notice an intriguing fact – some characters in the $`A_7`$ case exhibit an $`E_7`$ structure (cf. Tables 2 and 3):
$`\stackrel{~}{\chi }_{0,1}^8(q)+\stackrel{~}{\chi }_{0,3}^8(q)=\left(\chi _{2,1}^{4,5}(q)\right)^2,`$
$`\stackrel{~}{\chi }_{1,1}^8(q)+\stackrel{~}{\chi }_{1,3}^8(q)=\left(\chi _{2,2}^{4,5}(q)\right)^2,`$
$`\stackrel{~}{\chi }_{0,0}^8(q)\pm 2\stackrel{~}{\chi }_{0,2}^8(q)+\stackrel{~}{\chi }_{0,4}^8(q)=\left(\chi _{1,1}^{4,5}(q)\pm \chi _{1,4}^{4,5}(q)\right)^2,`$
$`\stackrel{~}{\chi }_{1,0}^8(q)\pm 2\stackrel{~}{\chi }_{1,2}^8(q)+\stackrel{~}{\chi }_{1,4}^8(q)=\left(\chi _{1,2}^{4,5}(q)\pm \chi _{1,3}^{4,5}(q)\right)^2.`$
This is the first case in which we have to combine three characters in order to obtain a factorized form. A more detailed account on the factorization of $`A_n`$-related characters will be presented elsewhere.
## III One Particle States
The functions $`\{x\}_y^+`$ and $`\frac{1}{\{x\}_y^{}}`$ can be written as double series in $`q`$ with coefficients being $`𝒫(n,m)`$ (or $`𝒬(n,m)`$) – the number of partitions of an integer $`n0`$ into $`m`$ distinct (or smaller than $`m+1`$) non-negative integers (see e.g. ).
Applying this fact to a character of the type (2) with $`x_ix_j`$, we obtain it in the form of a series $`\chi (q)=_{k=0}^{\mathrm{}}\mu _kq^k`$, where the level $`k`$ admits the partitioning, $`k=_a_{i_a}p_a^{i_a}`$, into parts of a specific form (e.g. (63) and (64) below). The interpretation of the $`p_a^{i_a}`$ as momenta of massless particles gives rise to a quasi-particle picture (developed originally for characters of the form (1) in ), where a character is regarded as a partition function, $`\chi (q)=_k\mu _ke^{\beta E_k}`$. Here $`q=e^{2\pi \beta v/L}`$, with $`v`$ being the speed of sound, and $`L`$ – the size of the system. A quasi-particle spectrum constructed in this way is in one-to-one correspondence to the Verma module of the corresponding irreducible representations of the Virasoro algebra or some modules related to linear combinations. It is crucial to stress that this procedure is not applicable to the standard representation of the characters (i.e. of the type (10)) and is a very specific feature of the representations (1) and (2). Note that the modules which are of the form (1) do in general (if they do, they give rise to Rogers-Ramanujan type identities ) not factorize, such that the spectra related to (2) do not only differ in nature from the ones obtainable from (1), but are also related to different sectors.
As just explained, a quasi-particle representation can be constructed for any factorizable character of the type (2) provided that $`x_ix_j`$. For instance, the characters (41) related to a compactified boson admit a representation with $`(2k+1)`$ particles. However, since we are particularly interested in spectra with Lie algebraic features, it is most natural to perform the quasi-particle analysis for the characters which admit the form $`\mathrm{\Xi }^𝐠(\stackrel{}{x},q)`$. In this way we obtain the following fermionic spectrum (if we employ the series involving $`𝒫(n,m)`$) in the units of $`2\pi /L`$
$$p_a^i(\stackrel{}{m})=x_a+\begin{array}{c}\left(\frac{h}{4}+\frac{1}{2}\right)\left(1m_a\right)+\left(\frac{h}{2}+1\right)N_a^i\end{array},$$
(63)
or bosonic spectrum (if we use the series with $`𝒬(n,m)`$)
$$p_a^i=x_a+\begin{array}{c}\left(\frac{h}{2}+1\right)N_a^i\end{array}.$$
(64)
Here $`\stackrel{}{x},\stackrel{}{m}`$ and $`\stackrel{}{N}`$ parameterize the possible states. In eq. (63) the numbers $`N_a^i`$ are distinct positive integers such that $`_{i=1}^{m_a}N_a^i=N_a`$, whereas in eq. (64) they are arbitrary non-negative integers. Notice that for the combination of characters the levels may be half integer graded, such that also the momenta take on half integer values in this case. A sample spectrum is presented in Table 4 which illustrates how the available momenta of the form (64) are to be assembled in order to represent a state at a particular level.
| $`k`$ | $`\mu _k`$ | $`p_1^i=1+\frac{5i}{2}`$, $`p_2^i=\frac{3}{2}+\frac{5i}{2}`$ |
| --- | --- | --- |
| $`\frac{1}{2}`$ | 0 | |
| 1 | 1 | $`|p_1^0`$ |
| $`\frac{3}{2}`$ | 1 | $`|p_2^0`$ |
| 2 | 1 | $`|p_1^0,p_1^0`$ |
| $`\frac{5}{2}`$ | 1 | $`|p_1^0,p_2^0`$ |
| 3 | 2 | $`|p_1^0,p_1^0,p_1^0`$, $`|p_2^0,p_2^0`$ |
| $`\frac{7}{2}`$ | 2 | $`|p_1^0,p_1^0,p_2^0`$, $`|p_1^1`$ |
| 4 | 3 | $`|p_1^0,p_1^0,p_1^0,p_1^0`$, $`|p_1^0,p_2^0,p_2^0`$, $`|p_2^1`$ |
Table 4. Bosonic spectrum for $`\chi _{1,2}^{5,6}(q)+\chi _{1,4}^{5,6}(q)`$. $`k`$ denotes the level and $`\mu _k`$ its degeneracy.
Naturally the questions arise if we can interpret these spectra more deeply and if we can possibly find alternative representations for the related modules. First of all we should give a meaning to the particular combinations which occur in our analysis. In we provided several possibilities. In particular the combination $`\chi _{1,2}^{5,6}(q)+\chi _{1,4}^{5,6}(q)`$ is of interest in the context of boundary conformal field theories, since this combination of characters coincides with the partition function $`Z_{A,F}`$ for the critical 3-state Potts model with boundaries ($`F`$ denotes the free boundary condition). It is intriguing that this combination possesses a manifestly Lie algebraic quasi-particle spectrum. The combination $`\chi _{2,2}^{5,6}(q)+\chi _{2,4}^{5,6}(q)`$, which coincides with $`Z_{BC,F}`$ in the same model possesses a slightly weaker relation to $`A_2`$.
To answer the question concerning possible representations, we recall the fact that the fields corresponding to the highest weight states satisfy the quantum equation of motion of Toda field theory . It is therefore very suggestive to try to identify the presented spectra in terms of the $`W`$-algebras . For $`\mathrm{\Xi }^𝐠(q)`$ we can make this more manifest. Changing the units of the momenta to $`\pi /L`$, we obtain from (48)
$$p_a^i=e_a+1+(h+2)N_a^i.$$
(65)
Here $`e_a`$ belongs to the exponents of the Lie algebra. Since the generators of the W-algebras $`W_{s+1}`$ are graded by the exponents plus one , we may associate the following generators to this quasi-particle spectrum
$$p_a^iW_a\left(W_aW_{ra}\right)^{N_a^i}.$$
(66)
In particular, the critical 3-state Potts model with boundaries would be related to the $`W_3`$-algebra. We leave it for the future to investigate this conjecture in more detail.
Acknowledgment: We would like to thank W. Eholzer for useful discussions. A.B. is grateful to the members of the Institut für Theoretische Physik, FU-Berlin for hospitality. A.F. is grateful to the Deutsche Forschungsgemeinschaft (Sfb288) for partial support.
|
no-problem/9812/cond-mat9812078.html
|
ar5iv
|
text
|
# References
COMPUTER SIMULATIONS OF DEFECTS IN PEROVSKITE KNbO<sub>3</sub> CRYSTALS
R. I. EGLITIS<sup>a,b</sup>, E. A. KOTOMIN<sup>b,c</sup>, A. V. POSTNIKOV<sup>c</sup>, N. E. CHRISTENSEN<sup>d</sup>, M. A. KOROTIN<sup>e</sup>, and G. BORSTEL<sup>c</sup>
<sup>a</sup>Institute of Materials Research & Engineering, National University of Singapore, Singapore 119260; <sup>b</sup>Institute of Solid State Physics, University of Latvia, 8 Kengaraga, Riga LV-1063, Latvia; <sup>c</sup>Universität Osnabrück – Fachbereich Physik, Osnabrück D-49069, Germany; <sup>d</sup>Institute of Physics and Astronomy, University of Aarhus, Aarhus C, DK-8000, Denmark; <sup>e</sup>Institute of Metal Physics, Yekaterinburg GSP-170, Russia
An ab initio LMTO approach and semi-empirical quantum chemical INDO method have been used for supercell calculations of basic point defects – $`F`$-type centers and hole polarons bound to cation vacancy – in partly covalent perovskite KNbO<sub>3</sub>. We predict the existence of both one-site and two-site (molecular) polarons with close absorption energies ($``$ 1 eV). The relevant experimental data are discussed and interpreted.
Keywords: ferroelectrics, atomic and electronic structure, vacancies, polarons, ab initio and semi-empirical methods
INTRODUCTION
Perovskite KNbO<sub>3</sub> crystals are widely used in non-linear optics and holography. Their properties are influenced by point defects, primarily by vacancies. Relatively little is known about such intrinsic point defects in KNbO<sub>3</sub>. A broad absorption band around 2.7 eV has been observed in electron-irradiated crystals and ascribed to the $`F`$-type centers (O vacancy which trapped one or two electrons – $`F^+`$ and $`F`$ centers, respectively).$`^{\text{[1]}}`$ Transient optical absorption at 1.2 eV has been associated recently,$`^{\text{[2]}}`$ in analogy with other perovskites, with a hole polaron (a hole bound to some defect). The ESR study of KNbO<sub>3</sub> doped with Ti<sup>4+</sup> gives a proof that holes could be trapped by such negatively charged defects.$`^{\text{[3]}}`$ For example, in BaTiO<sub>3</sub>, hole polarons were also found which are bound to Na and K alkali ions replacing Ba and thus forming a negatively charged site attracting a hole.$`^{\text{[4]}}`$ Primary candidates for such defects are cation vacancies. In irradiated MgO they are known to trap one or two holes giving rise to the V<sup>-</sup> and V<sup>0</sup> centers$`^{\text{[5, 6]}}`$ which are nothing but bound hole polaron and bipolaron, respectively. Some preliminary theoretical study has been already done by us on $`F`$ centers in KNbO<sub>3</sub>$`^{\text{[7]}}`$ and on hole polarons in this material.$`^{\text{[8]}}`$ In the present paper, we report the results of additional computer simulations using the same two different approaches as e.g. in Ref. . We restrict ourselves to the cubic phase of KNbO<sub>3</sub>, with the lattice constant $`a_0`$=4.016 Å.
METHODS
For the study of the ground-state atomic and electronic structure we used the ab initio linearized muffin-tin orbital (LMTO) method based on the local density approximation (LDA). For structure optimizations, the full-potential version of LMTO by van Schilfgaarde and Methfessel has been used.$`^{\text{[9]}}`$ As an extension of the analysis previously done in Ref. for the $`F`$ center, we applied the tight-binding LMTO method$`^{\text{[10]}}`$ in the modification incorporating the LDA+$`U`$ formalism.$`^{\text{[11]}}`$ The latter allows to maintain the orbital dependency of the potential and, to some extent, to treat Coulomb correlation effects within localized shells beyond the LDA. We used this approach for introducing an ad hoc upward shift of the conduction band (mostly consisting of Nb $`4d`$ states) and splitting-off of the $`F`$ center level from it in a different way than it was done in Ref. . Moreover, this allowed us to analyze the symmetry of the $`F`$ center wavefunction. We used $`U`$=8 eV and $`J`$=0 (for the Nb $`4d`$ shell) as parameters of the LDA+$`U`$ method.
In parallel with the LMTO, the semi-empirical method of the Intermediate Neglect of the Differential Overlap (INDO) modified for ionic and partly ionic solids$`^{\text{[12]}}`$ has been used (see Ref. for details of its application to KNbO<sub>3</sub>). Differently from the LMTO, the INDO method is based on the Hartree-Fock formalism and allows self-consistent calculations of the excited states of defects and thus the relevant absorption energies using the so-called $`\mathrm{\Delta }`$SCF method. The simulation of all defects has been done within a supercell approach, with one (neutral) O atom removed to model the $`F`$ center and two atoms, O and K, removed to simulate the $`F^+`$ center. For the modeling of the hole centers, a K atom has been removed, and the actual type of the hole polaron (one-site and two-site) was set by the symmetry of the local lattice relaxation. In LMTO calculations, the $`2\times 2\times 2`$ supercells containing 40 atoms were used in all cases, i.e. with repeated point defects separated by $``$8 Å. Only the positions of nearest neighbors to the defect were relaxed. In the INDO calculations we used much larger, $`4\times 4\times 4`$ supercells (320 atoms), and allowed for the relaxation of more distant neighbors. Besides decreasing the residual interaction between impurities in adjacent supercells, this effectively takes into account the dispersion of energy bands over the Brillouin zone of KNbO<sub>3</sub> up to a better extent than it was possible in previous defect calculations$`^{\text{[7, 8]}}`$ with smaller supercells.
RESULTS AND DISCUSSIONS
$`F`$-type centers
In the cubic phase all O atoms are equivalent and have the local symmetry C<sub>4v</sub> (due to which the excited state of the $`F`$-type centers could be split into a nondegenerate and a doubly-degenerate levels). The optimized atomic relaxation around the $`F`$ center as done by the LMTO indicates the outward shift of the Nb neighbors to the O vacancy by 3.5% a<sub>0</sub>. The associated lattice relaxation energy is shown in Table 1.
The optimized Nb relaxation found in the INDO simulations was 3.9%, i.e. very close to the ab initio calculations. The outward relaxation of nearest K atoms and inward displacements of O atoms are much smaller. They give $``$20% of the net relaxation energy of 1.35 eV. The $`F`$ center local energy level lies $``$0.6 eV above the top of the valence band. Its molecular orbital contains primarily the contribution from the atomic orbitals of the two nearest Nb atoms. Only $``$0.6 $`e`$ resides at the orbitals centered at the vacancy site; hence the electron localization at the defect is much smaller than is known for $`F`$ centers in ionic oxides (see Ref. for comparison). The symmetry analysis of the ground-state wave function associated with the $`F`$ center, done by the TB-LMTO method with the use of the LDA+$`U`$ formalism and by INDO, revealed the same result, namely that the major contribution comes from the $`e_g`$ states centered at Nb neighbors (more specifically, it is essentially the $`3z^2r^2`$ component, with $`z`$ in the direction towards the $`F`$ center). The partial densities of states from the LMTO calculation are shown in Fig. 1.
FIGURE 1 Local density of states of the $`F`$-center (left panel) and of the Nb atom nearest to it as calculated by the LMTO method.
For the $`F^+`$ center the relaxation energy of 2.23 eV and the Nb displacements of 5.1% are larger than those for the $`F`$ center due to a stronger Coulomb repulsion between unscreened O vacancy and Nb atoms: a share of the electron density inside the O vacancy decreases to 0.3 $`e`$.
The optical absorption energies calculated by means of the $`\mathrm{\Delta }`$SCF method for the $`F^+`$ and $`F`$ centers are given in Table 1. The two absorption bands predicted for the former center are shifted to the low-energy side, which is in agreement with similar defects in ionic oxides.$`^{\text{[14]}}`$ However, both defects are predicted to have one of the bands around 2.6–2.7 eV, in agreement with the experimental observation.$`^{\text{[1]}}`$
Hole Polarons
Both ab initio and semi–empirical calculations agree that there are $`two`$ energetically favorable atomic configurations in which a hole is well localized: one-site and two-site (molecular) polarons. In the former case, a single O<sup>-</sup> ion is displaced towards the K vacancy by 1.5 % (LMTO) or 3% (INDO). The INDO calculations show that simultaneously, 11 other nearest oxygens surrounding the vacancy are slightly displaced outwards the vacancy. In the two-site (molecular) configuration, a hole is shared by the two O atoms which approach each other – by 0.5% (LMTO) or 3.5% (INDO) – and both shift towards a vacancy – by 1.1% (LMTO) or 2.5% (INDO). The lattice relaxation energies (which could be associated with the experimentally measurable hole thermal ionization energies) are presented in Table 1. In both methods the two-site configuration of a polaron is lower in energy.
In spite of general observation of a considerable degree of covalency in KNbO<sub>3</sub> (see, e.g., Ref. for a discussion) and contrary to a delocalized character of the $`F`$ center state, the one-site polaron state remains well localized at the displaced O atom, with only a small contribution from atomic orbitals of other O ions but not K or Nb ions. In agreement with Schirmer’s theory for the small-radius polarons in ionic solids,$`^{\text{[15]}}`$ the optical absorption corresponds to a hole transfer to the state delocalized over nearest oxygens. The calculated absorption energies for one-site and two-site polarons are close (Table 1) and twice smaller than the experimental value for a hole polaron trapped near Ti.$`^{\text{[3]}}`$ This shows that the optical absorption energy of small bound polarons could be strongly dependent on the defect involved.
CONCLUSIONS
(i) The two different methods used for defect calculations reveal a qualitative agreement, despite the fact that the INDO (as is generally typical for the Hartree-Fock-based schemes) systematically gives larger atomic displacements and relaxation energies. Both two–electron $`F`$-center calculations demonstrate a strong electron delocalization from the O vacancy over the two nearest Nb atoms; very likely due to a considerable covalency of the chemical bonding in KNbO<sub>3</sub>, and predict the e<sub>g</sub> symmetry of the wave function which could be checked experimentally. Both methods agree also that both one–site and two–site hole polarons bound to the cation vacancy are energetically favorable, with a preference to the latter.
(ii) The INDO calculations of the optical absorption energies strongly support the interpretation of the experimentally observed band at 2.7 eV as due to the $`F`$-type centers. To distinguish between the $`F`$ and $`F^+`$ centers, the ESR measurements should be used.
(iii) The calculated hole polaron absorption ($``$1 eV) is close to the observed short-lived absorption band energy$`^{\text{[2]}}`$ and thus could arise due to a hole polaron bound at cation vacancy. Further detailed study is needed to clarify whether such hole polarons are responsible for the effect of the blue-light-induced-infrared-absorption (BLIIRA) reducing the second-harmonic generation efficiency in KNbO<sub>3</sub>.$`^{\text{[2]}}`$
Acknowledgments
This study was partly supported by the DFG (a grant to E. K.; the participation of A. P. and G. B. in the SFB 225), Volkswagen Foundation (grant to R. E.), and the Latvian National Program on New Materials for Micro- and Optoelectronics (E. K.). M. K. appreciates the hospitality of the University of Osnabrück during his stay there. Authors are greatly indebted to Prof. M. R. Philpott for valuable discussions.
|
no-problem/9812/astro-ph9812113.html
|
ar5iv
|
text
|
# Shading and Smothering of Gamma Ray Bursts
## Abstract
The gamma ray burst (GRB) 980425 is distinctive in that it seems to be associated with supernova (SN) 1998bw, has no X-ray afterglow, and has a single peak light curve and a soft spectrum. The supernova is itself unusual in that its expansion velocity exceeds c/6. We suggest that many of these features can be accounted for with the hypothesis that we observe the GRB along a penumbral line of sight that contains mainly photons that have scattered off ejected baryons. The hypothesis suggests a baryon poor jet (BPJ) existing within a baryon rich outflow. The sharp distinction can be attributed to whether or not the magnetic field lines thread an event horizon. Such a configuration suggests that there will be some non-thermal acceleration of pick-up ex-neutrons within the BPJ. This scenario might produce observable spallation products and neutrinos.
The problem of baryon contamination of gamma ray bursts is well known. In order to be detectable at cosmological distances, a GRB must put out enough power $`10^{50}`$ erg s<sup>-1</sup> to drive away at least $`10^3M_{}`$ of baryonic matter over the duration of the burst, and this would be enough to obscure gamma rays originating within $`10^{14}`$ cm of the point of energy release. Because the maximum Lorentz factor $`\mathrm{\Gamma }`$ of the bulk expansion could not be large compared to unity, however, the gammasphere would have to be placed well within this radius if the observed burst duration is to be only seconds, leaving the non-thermal nature of the spectra and short timescale a puzzle.
Suggested solutions for this problem include hypotheses that a) the GRB is powered by a merger of strange stars (Haensel, Paczynski, and Amsterdamski, 1992; Usov, 1998) which hold their baryons using strong interactions and b) the burst originates on field lines that thread an event horizon (Levinson and Eichler, 1993; Iwamoto, 1998; Paczynski 1998). Other proposals (e.g. Eichler et. al. 1989; Mezaros and Rees, 1992; Mochkovitch, R. et. al. 1993) involve some combination of geometry and centrifugal force that keep the axis of an accretion disk free of baryons, but they would need to be quantified and are harder to analyze.
If the baryon purity is enforced by an event horizon that forms during the same process that powers the GRB, then a baryonic outflow would be driven from the matter that had not yet fallen into the event horizon. It would most likely surround the BPJ that emerges from the event horizon and its photosphere would be considerably larger than that of the BPJ. The external observer could, therefore, see the primary emission from the BPJ only if looking down the hole it makes in the baryonic outflow. If too far off to the side, the observer might not see anything. A third possibility, however, is that the observer is offset from the BPJ axis enough that the central engine of the GRB is obscured, but the walls of the baryonic outflow that interface the BPJ are partly visible to the observer (see figure 1). An observer viewing from such a ”penumbral” line of sight could see photons that were scattered from the walls that envelop the BPJ but not photons in its primary beam.
GRB 80425 has a high probability of being associated with the unusual supernova (SN) 1998bw (Galama et. al., 1998). Kulkarni, et. al. (1998) have reported radio emission beginning several days after the supernova, and argued that the GRB, being intrinsically weak if indeed associated with the SN, is then a member of a separate class of GRB’s. Iwamoto (1998) has proposed a model in which the GRB is generated by an accreting black hole that was made during the same collapse that generated the SN. In his model, the GRB energy and associated afterglows, are intrinsically weak as argued by Kulkarni et. al.
We propose that GRB 80425 is in fact being viewed along a ”penumbral line of sight” as defined above, that it represents a class of viewing angles along which GRB appear much weaker, and that it is perhaps the tip of an even larger class of buried GRB. There are several motivations for this given below, apart from the possible association of the GRB with a supernova. We envision that most of the gamma-ray emission in the BPJ is accompanied by a baryonic outflow. The latter may be a pre-existing star or it may be a wind driven from a neutron star as it merges. If it is a wind, we assume that the outflow velocity exceeds the BPJ photosphere divided by the burst duration, so that during the rise of the primary flux the baryonic outflow extends out to a radius much larger than the radius of the primary emission zone. Viewing the source from outside the primary beam we only see the scattered component. Failure to see X-ray afterglow could be attributed to our being outside the beaming cone of afterglow emission.
The variation of the scattered emission is anticipated to occur on a timescale of order the light travel time across the system (30 light seconds in the case of GRB 980425), provided the primary burst persists for a shorter time. Temporal substructure in the primary emission would be smeared out in the scattered component, rendering the time profile singly peaked. Having been scattered at least once, the spectrum would cut off sharply above several hundred KeV. The scattered component, though extending over a broader angular region, could be considerably weaker, depending on the detailed assumptions regarding the geometry and kinematics: The inner wall of the baryonic outflow may catch only a small fraction of the BPJ’s emission, and the angle of incidence between photon and wall may be sharper than the angle between the wall and the observer, leading to a net loss of energy even in the Thomson scattering limit. Compton recoil in any case sets in at several hundred KeV.
The explanation of a single peak light curve is of course presented after the observation. Situations that may lead to different temporal characteristics (e.g. flash-in-the-pan effects) could be envisaged, but a prolonged, smeared light curve seems like a reasonable signature of scattering off baryonic matter that accompanies the burst. Softening of the spectrum where Klein Nishina effects set in would seem to be an unavoidable consequence of scattering.
As an illustrative example we consider a cylindrical wall illuminated by a beamed, point source located on the axis of the cylinder a distance $`z_{ph}`$ from the photosphere of the baryonic outflow. The cone of emission is taken to be coaxial with the cylinder with an opening angle $`\eta _{max}`$. The wall is assumed to move at a constant velocity, $`\beta `$, and to have an infinitely large Thomson depth. To simplify the analysis, the scattering cross section is assumed to be isotropic in the frame of the wall, and KN effects are ignored. Let us denote by $`_{GRB}(ϵ)`$ the spectral energy distribution of the illuminating source. The power per unit energy per unit solid angle emitted in a direction $`\widehat{\mathrm{\Omega }}`$ by the section of the wall exposed to the observer is then given by
$$_{em}(ϵ_s,\widehat{\mathrm{\Omega }})=\frac{\mathrm{sin}\theta }{4\pi ^2(1\beta \mathrm{cos}\theta )^3}_{\eta _\theta }^{\eta _{ph}}\left\{F(\eta )\mathrm{sin}\eta (1\beta \mathrm{cos}\eta )^3_{GRB}\left[\frac{(1\beta \mathrm{cos}\theta )}{(1\beta \mathrm{cos}\eta )}ϵ_s\right]\right\}𝑑\eta $$
(1)
where $`\theta `$ is the angle between $`\widehat{\mathrm{\Omega }}`$ and the wall axis, $`\eta _{ph}=\mathrm{cot}^1(z_{ph/R})`$, with $`R`$ being the cross-sectional radius of the wall, $`\mathrm{cot}\eta _\theta =`$ max $`\{\mathrm{cot}\eta _{ph}2\mathrm{cot}\theta ;\mathrm{cot}\eta _{max}\}`$, and $`F(\eta )=[4\mathrm{tan}^2\theta (\mathrm{cot}\eta _{ph}\mathrm{cot}\eta )^2]^{1/2}`$. Note that when $`\theta =\pi /2`$ the lower limit $`\eta _\theta =\eta _{ph}`$ and the integral in eq. (1) vanishes, as it should. For a power law spectrum, viz., $`_{GRB}(ϵ)=_oϵ^p`$; $`ϵ_{min}<ϵ<ϵ_{max}`$, eq. (1) reduces to
$$_{em}(ϵ_s,\widehat{\mathrm{\Omega }})=\frac{_oϵ_s^p}{4\pi ^2}G(\beta ,\theta ,ϵ_s)$$
(2)
with
$$G(\beta ,\theta ,ϵ_s)=\frac{\mathrm{sin}\theta }{(1\beta \mathrm{cos}\theta )^{3+p}}_{\eta _\theta }^{\eta _{ϵ_s}}\left\{F(\eta )\mathrm{sin}\eta (1\beta \mathrm{cos}\eta )^{3+p}\right\}𝑑\eta ,$$
(3)
where $`\mathrm{cos}\eta _{ϵ_s}=`$ min $`\{\mathrm{cos}\eta _{ph};[1(ϵ_s/ϵ_{max})(1\beta \mathrm{cos}\theta )]/\beta \}`$. As seen, the spectrum emitted by the wall is also a power law with an upper cutoff at energy $`ϵ_b=ϵ_{max}\frac{1\beta \mathrm{cos}\eta _{ph}}{1\beta \mathrm{cos}\theta }`$. Plots of $`G`$ versus $`\beta `$ for energies below the cutoff, and for different viewing angles, are exhibited in fig. 2.
The radio afterglow appears to come from an emission region moving with a bulk Lorentz factor of about 2 (Kulkarni et. al. 1998), (but see Loeb and Waxman, 1998). This could be because the jet has slowed down from higher values of $`\mathrm{\Gamma }`$ by the time the radio afterglow is produced. But it could also be that the radio emission is generated by the forward shock of the ejected baryons. The required Lorentz factor of 2 or so implies a velocity only 3 or 4 times faster than the ejecta and could be accounted for by the acceleration of a shock front as it moves down the density gradient in the outer part of the presupernova star. The enormous ejecta velocity, five or six times that of a typical type 1C supernova, could be correlated with the appearance of gamma rays. One could argue that a low baryon density jet is formed in a much larger, less peculiar class of supernova, and that the gamma rays are usually smothered. Although not confident about the following point, we do not rule out the possibility that the baryonic output of SN 1998bw is low because it came directly off a preexisting neutron star that underwent a merger with some other object. This might account for the unusually high level of radioactivity in the ejecta as well as the high velocity.
The above hypothesis may also have the following observational consequences:
i) Shaded bursts, being effectively weaker in the direction of the observer, would have a different V/V<sub>max</sub> distribution from the usual ones, possibly even approaching an average value of 1/2. This is a retrodiction, since it has already been noted by Tavani (1998) that long duration, soft GRB’s have a different distribution from the majority, which show a cosmological imprint in their distribution.
ii) The scattered hard X-/soft gamma rays should be linearly polarized, and it may be technologically possible to measure their polarization in the not too distant future.
iii) The coexistence of a medium that is low enough density to admit shock acceleration with dense baryonic material that envelops it could be the site of Be and B production in the Galaxy. Here we have assumed that buried gamma ray bursts are not uncommon. (In this regard , it is worth distinguishing between allowing gamma rays to escape, which requires extreme baryon purity on the scales of GRB photospheres, and allowing shock acceleration, which may be less demanding.)
iv) Particle acceleration within a very dense environment also implies neutrino production via pions if the accelerated particles are baryons. The possibility of high energy neutrino production in GRB has been considered by a number of authors.
With a nominal GRB rate of 10<sup>-6</sup>/galaxy-yr, process (iii) is irrelevant and and process (iv) is probably not observable with neutrino detectors now under construction (the largest being AMANDA) unless the total GRB output greatly exceed $`10^4`$ erg/cm<sup>2</sup> (Eichler, 1994). However, if we are to consider the possiblity that any stellar collapse that yields a black hole also (much of the time, at least) yields a smothered GRB, then one could postulate that they are far more frequent than GRB, and perhaps almost as frequent as supernovae.
Consider now a BPJ emanating from a black hole that has formed in the midst of a supernova (or hypernova). Some neutron rich material should be expelled behind the rest of the ejecta by the same arguments that exist for GRB. Neutrons can cross field lines and drift into the BPJ. If they decay there, and the jet is already relativistic, then they immediately become cosmic rays in the frame of the jet. Moreoever, they become the principal injection component to any shock acceleration that occurs in the jet. Estimating the infusion velocity v of neutron into the BPJ by equating the transverse momentum imparted to a proton at the wall to a significant fraction of the protons outward momentum $`m_pu`$ we obtain, neglecting the proton - neutron mass difference,
$$v^2\sigma n=u/\delta t$$
(4)
where n is the density of neutrons at the wall of the BPJ, $`\delta t`$ is the hydrodynamical timescale, and $`\sigma `$ is the neutron proton scattering cross section, whence
$$v/uR/(N\sigma )^{1/2}$$
(5)
where we have left aside dimensionless geometric factors. Here $`NnuR^2\delta t`$ is the total number of neutrons, maybe $`10^{54}`$, and $`\sigma `$ is of order 5 barns, The quantity R is the radius at which the crossing takes place, which we take to be $`10^{12}`$ cm, so that the neutrons have not quite decayed out of the BPJ, but soon will decay inside it given that v is small compared to u. With these numbers, $`10^{50.5}`$ neutrons can be picked up by the BPJ. This would be enough to give of order 4 $`\times 10^{48}`$ Beryllium nuclei (the needed quantity as estimated by several authors, e.g., Drury and Parizot, and references therein) if a fair fraction of these ex-neutrons make their way into CNO-rich ejecta. The detectability of the neutrinos, which would be made in comparable numbers to the number of pick-up neutrons but at a factor of 10 or so less energy, depends on the assumed bulk Lorentz factor in BPJ at the point of pick-up. Assuming the bulk Lorentz factor of $`10^3`$ and, say $`10^{50}`$ pick-up neutrons, gives of the order of $`10^{50}`$ 100 GeV neutrinos. This would be detectable by AMANDA at distances of an Mpc or so. Shock acceleration of the ex-neutrons could raise the average energy (and hence the detectability) of the neutrinos.
The various scenarios outlined above, if ever confirmed, would represent a qualitative manifestation of the hypothesized event horizon, which is responsible for the sharp transverse density and velocity gradients within the explosion. One should note, however, that any baryon-retaining mechanism, such as magnetic field lines threading strange matter, could also maintain sharp gradients provided that it is immersed in a baryonic outflow.
AL acknowledges support by Alon Fellowship and a grant from the Israeli Science Foundation. DE acknowledges a grant from the Israeli Science foundation. We are grateful to L. Drury, E. Waxman, and V. Usov for helpful conversations.
|
no-problem/9812/gr-qc9812030.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Charlie Misner was the first to recognize that the subject of harmonic mappings of Riemannian manifolds finds an important application in general relativity. In a pioneering paper with Richard Matzner he found that stationary, axially symmetric Einstein field equations can be formulated as a harmonic map. Eells and Sampson’s theory of harmonic mappings of Riemannian manifolds provides a geometrical framework for thinking of a set of pde’s, in the same spirit as “mini-superspace” that Charlie was to introduce for ode Einstein equations a little later. The subsequent developement of the subject of space-times admitting two Killing vectors, that eventually led to its recognition as a completely integrable system \- has employed another formulation of the stationary, axi-symmetric field equations due to Ernst which is equivalent to that of Matzner and Misner.
Charlie’s later work on harmonic maps encompasses a scope much broader than this specific problem and its power and elegance is bound to make a major impact on theoretical physics.
I was privileged to be in contact with Charlie’s ideas at that time and worked on the two Killing vector problem , . It was the hyperbolic version of gravitational fields admitting two Killing vectors that attracted my attention. This is the problem of colliding impulsive plane gravitational waves for which Khan and Penrose had presented a famous solution . My work finally led to the exact solution for colliding impulsive plane gravitational waves with non-collinear polarizations which is physically the most general solution of this type. It turned out that the Matzner-Misner formalism was the one more readily amenable to the hyperbolic problem, even though it was originally intended for the elliptic case, whereas the Ernst formulation fitted the elliptic problem, ie the exterior field of rotating stars, best. The relationship between these two formalisms is given by a Neugebauer-Kramer involution .
A few years after the solution appeared, there was a remarkable avalanche of papers on colliding plane gravitational waves. There were important papers examining the singularity structure of spacetimes resulting from the collision of gravitational waves \- . However, there was also a mass of new exact solutions which are all essentially devoid of any physical interest because their authors had chosen not to solve the Cauchy problem with the initial data appropriate to generic plane waves, but rather they started with a “solution” and derived (!) the initial data. This type of derived initial data for the collision problem describes some very peculiar plane waves indeed. An inordinately large number of such references can be found in .
Nevertheless, physically interesting colliding wave problems are still open and waiting for an exact solution! Remarkably enough, the interaction of plane impulsive gravitational and electromagnetic shock waves is in this category. We have the Khan-Penrose and Bell-Szekeres solutions describing the interaction of either two impulsive gravitational, or two electromagnetic shock waves alone and also the solution of Griffiths for the interaction of an impulsive gravitational wave with an electromagnetic shock wave. But the generic case where we must consider the collision of both type of waves is missing even in the case of collinear polarization. The important open problem here is the construction of an exact solution of the Einstein-Maxwell equations that reduces to all, the Khan-Penrose, Bell-Szekeres and Griffiths solutions. There are various unsatisfactory treatments of this problem in the literature , . I shall give its harmonic map formulation.
## 2 Initial Data
The problem of colliding plane gravitational waves was proposed and in essence solved by Penrose in 1965 even though most people writing on this subject do not seem to be familiar with it. We shall use Penrose’s formulation of the Cauchy problem , to discuss the interaction of two plane waves, every one of which will consist of a superposition of an impulsive gravitational wave and an electromagnetic shock wave. The interaction will be determined by an integration of the Einstein-Maxwell equations with initial data defined on a pair of intersecting null characteristics. The initial values of the fields will be those appropriate to a plane wave which is given by the Brinkman metric ,
$$ds^2=2du^{}dv^{}dx^{\mathrm{\hspace{0.33em}2}}dy^{\mathrm{\hspace{0.33em}2}}+2H(v^{},x^{},y^{})dv^{\mathrm{\hspace{0.33em}2}}$$
(1)
and the superposition of an impulsive gravitational wave and an electromagnetic shock wave, with amplitudes proportional to $`a,b`$ respectively, is obtained for
$$H=\frac{a}{2}\left(y^{\mathrm{\hspace{0.33em}2}}x^{\mathrm{\hspace{0.33em}2}}\right)\delta (v^{})\frac{b^2}{2}\left(x^{\mathrm{\hspace{0.33em}2}}+y^{\mathrm{\hspace{0.33em}2}}\right)\theta (v^{})$$
(2)
where $`\delta `$ is the Dirac delta-function and $`\theta `$ is the Heaviside unit step-function.
The Brinkman coordinate system employed in eq.(1) is useful because the superposition of waves travelling in the same direction is obtained simply by addition. However, the Brinkman coordinates are not suitable for the collision problem because of the explicit dependence of the metric on $`x^{},y^{}`$. For this purpose we must transform to the Rosen form where the metric coefficients will depend on $`v`$ alone. This is accomplished by the Khan-Penrose transformation
$$\begin{array}{ccc}v^{}\hfill & =& v,\hfill \\ u^{}\hfill & =& u+\frac{1}{2}x^2FF_v+\frac{1}{2}y^2GG_v,\hfill \\ x^{}\hfill & =& xF,\hfill \\ y^{}\hfill & =& yG,\hfill \end{array}$$
(3)
which results in
$$ds^2=2dudvF^2dx^2G^2dy^2,$$
(4)
provided
$$\begin{array}{ccc}F_{vv}\hfill & =& \left[a\delta (v)b\theta (v)\right]F,\hfill \\ G_{vv}\hfill & =& \left[a\delta (v)+b\theta (v)\right]G.\hfill \end{array}$$
(5)
These are linear, distribution-valued ordinary differential equations which can be solved using the Laplace transform
$$(s)=_0^{\mathrm{}}e^{sv}F(v)𝑑v$$
(6)
and from eq.(5) we find
$$(s)=\frac{1}{s^2+b^2}\left[(sa)F(0)+F_v(0)\right]$$
(7)
where of $`F(0),F_v(0)`$ are initial values. They are obtained from the continuity of the metric and its first derivatives across $`v=0`$ which requires $`F(0)=1`$, $`F_v(0)=0`$. In this case inverting the Laplace transform we get
$$F=cos(bv\theta (v))\frac{a}{b}sin(bv\theta (v))$$
(8)
and the result for $`G`$ is obtained by letting $`aa`$ in eq.(8) as indicated by eqs.(5).
In Rosen coordinates the general form of the metric that admits two hypersurface orthogonal Killing vectors is given by
$$ds^2=2e^Mdudve^U\left(e^Vdx^2+e^Vdy^2\right)$$
(9)
where $`U,V,M`$ depend on only $`u,v`$ and comparison with eqs.(4) and (8) shows that for the initial value problem the data is given by
$$\begin{array}{ccc}e^U\hfill & =& cos^2(bv\theta (v))\frac{a^2}{b^2}sin^2(bv\theta (v))\hfill \\ e^V\hfill & =& \frac{b+a\mathrm{tan}\left(bv\theta \left(v\right)\right)}{ba\mathrm{tan}\left(bv\theta \left(v\right)\right)}\hfill \\ e^M\hfill & =& 1.\hfill \end{array}$$
(10)
The limiting values of this result are familiar. If we have just an impulsive gravitational wave, we must pass to the limit $`b0`$ which yields
$$\begin{array}{ccc}e^U\hfill & =& 1av^2\theta (v)\hfill \\ e^V\hfill & =& \frac{1+av\theta \left(v\right)}{1av\theta \left(v\right)}\hfill \\ e^M\hfill & =& 1\hfill \end{array}$$
(11)
as in the case of Khan and Penrose. Furthermore, in the limit $`a0`$ we have only an electromagnetic shock wave
$$\begin{array}{ccc}e^U\hfill & =& 1sin^2(bv\theta (v))\hfill \\ e^V\hfill & =& 1\hfill \\ e^M\hfill & =& 1\hfill \end{array}$$
(12)
which is the result for the Bell-Szekeres case.
Spacetimes describing colliding plane waves are divided into four regions:
Region I : $`u<0,v<0`$ empty space before the collision
Region II: $`u>0,v<0`$ a plane wave
Region III: $`u<0,v>0`$ another wave travelling in the opposite direction
Region IV: $`u>0,v>0`$ the interaction region
The initial values given above are on $`v=0`$, the boundary between Regions III, IV and similar results hold on $`u=0`$, the boundary between Regions II, IV, determining the $`u`$-dependence. In the latter case the amplitudes of these waves will be given by different constants, say $`ap`$ and $`bq`$, cf eq.(28) in sequel. The case of Griffiths’ solution is a mixture where we have eqs.(11) between Regions III, IV and eqs.(12) with $`u`$ replacing $`v`$ on the boundary between Regions II, IV.
## 3 Einstein-Maxwell Equations
The Einstein-Maxwell field equations governing the interaction of two plane waves is well-known . Starting with the metric (9) and the Maxwell potential 1-form $`𝒜`$
$$𝒜=Adx$$
(13)
where $`A`$ depends only on $`u,v`$, we find a set of Einstein field equations which can be grouped into two categories. First we have the initial value equations
$$\begin{array}{ccc}2U_{vv}U_v^{\mathrm{\hspace{0.33em}2}}+2M_vU_vV_v^{\mathrm{\hspace{0.33em}2}}\hfill & =& 2\kappa e^{UV}A_v^{\mathrm{\hspace{0.33em}2}}\hfill \\ 2U_{uu}U_u^{\mathrm{\hspace{0.33em}2}}+2M_uU_uV_u^{\mathrm{\hspace{0.33em}2}}\hfill & =& 2\kappa e^{UV}A_u^{\mathrm{\hspace{0.33em}2}}\hfill \end{array}$$
(14)
and their integrability conditions
$$\begin{array}{c}U_{uv}U_vU_u=0\\ 2A_{uv}V_vA_uV_uA_v=0\\ 2V_{uv}U_vV_uU_uV_v+2\kappa e^{UV}A_uA_v=0\\ 2M_{uv}+U_vU_uV_vV_u+2\kappa e^{UV}A_uA_v=0\end{array}$$
(15)
where $`\kappa `$ is Newton’s constant in geometrical units.
In eqs.(15) we have the wave equation for $`e^U`$ and its solution is immediate from the initial values. The following two equations are the main equations and the last equations is irrelevant as $`M`$ can be obtained from quadratures once the main equations are solved.
The problem consists of finding a solution to eqs.(15) satisfying the initial data (10). Finally, we shall remark that eqs.(11), or (12) can be regarded as the solution of an initial value problem themselves, namely one between Region I and either a gravitational impulsive wave, or an electromagnetic shock wave across the null plane $`v=0`$ in Region III. In this mini-problem the first one of eqs.(14) serves as the field equation.
## 4 Harmonic Maps
We refer to and for a review and survey of the principal results on harmonic mappings of Riemannian manifolds. Here we shall briefly recall the most basic definitions in order to fix the notation. We shall consider two Riemannian manifolds endowed with metrics
$$\begin{array}{cc}ds^2=g_{ik}dx^idx^k,& i=1,\mathrm{},n\\ ds^{\mathrm{\hspace{0.33em}2}}=g_{\alpha \beta }^{}dy^\alpha dy^\beta ,& \alpha =1,\mathrm{},n^{}\end{array}$$
(16)
and a map
$$f:^{}$$
(17)
between them. This map is called harmonic if it extremizes the energy fuctional of Eells and Sampson, $`\delta =0`$,
$$(f)=g_{\alpha \beta }^{}\frac{f^\alpha }{x^i}\frac{f^\beta }{x^k}g^{ik}\sqrt{\text{ }g\text{ }}d^nx$$
(18)
where the Lagrangian consists of the trace with respect to the metric $`g`$ of the induced metric $`fg`$ on $``$. When the target space $`^{}`$ is 1-dimensional, harmonic maps satisfy Laplace’s equation on the background of $``$ and on the other hand if $``$ is 1-dimensional, then harmonic maps coincide with the geodesics on $`^{}`$. The nonlinear sigma model corresponds to the harmonic map $`f:R^2S^2`$.
There are at least three different ways in which the Einstein-Maxwell equations (15) can be formulated as a harmonic map.
1. We can take a 4-dimensional target space $`^{}`$ with the metric
$$ds^{\mathrm{\hspace{0.33em}2}}=e^U\left(2dMdU+dU^2dV^2\right)\mathrm{\hspace{0.17em}2}\kappa e^VdA^2$$
(19)
which has sections of constant curvature and the flat metric
$$ds^2=2dudv$$
(20)
on $``$. This is the electrovac analogue of the formulation given in . It is not the most economical approach because $`M`$, which can be obtained by quadratures, appears explicitly in the metric (19).
2. We can get rid of $`M`$ and consider a 2-dimensional target space with constant curvature at the expense of regarding $`U`$ as a given function on $``$ which does not enter into the variational problem as one of the local components of the harmonic map. Thus assuming that $`e^U`$ satisfies the wave equation, as in eqs.(15), we can take the metric on the target space as
$$ds^{\mathrm{\hspace{0.33em}2}}=e^U\frac{dd\overline{}}{\text{ }Re\text{ }^2}$$
(21)
where
$$=e^{(VU)/2}+i\sqrt{\frac{\kappa }{2}}A$$
(22)
is an Ernst potential type of complex coordinate. The metric (21) is that of a space of constant negative curvature and $``$ is the complex coordinate for the Poincaré upper half plane. There exists another representation, namely Klein’s unit disk for the space of constant negative curvature. This is obtained by the transformation
$$=\frac{\xi +1}{\xi 1}$$
(23)
and
$$ds^{\mathrm{\hspace{0.33em}2}}=e^U\frac{d\xi d\overline{\xi }}{(1\xi \overline{\xi })^2}$$
(24)
is the resulting form of the metric. Frequently this is the most convenient representation. The original Matzner-Misner as well as the Neugebauer-Kramer formulations are of this type.
The metric on $``$ is the same as the one in eq.(20).
This approach is by far the most common procedure followed in the literature.
3. It is possible to reformulate the reduced problem by avoiding the ad hoc introduction of $`U`$ on $`^{}`$ at the expense of adding a new Killing direction to $``$ where the magnitude of the Killing vector is $`e^U`$. In this case we have
$$ds^{\mathrm{\hspace{0.33em}2}}=\frac{d\xi d\overline{\xi }}{(1\xi \overline{\xi })^2}$$
(25)
where the definition of $`\xi `$ is the same as before and
$$ds^2=2dudve^{2U}dz^2$$
(26)
where we consider the mapping to be independent of $`z`$ and once again $`U`$ is specified a priori. It appears that this possibility has not been discussed before in the literature.
The first formulation where it is not necessary to introduce information from outside the variational principle is attractive. However, the close resemblance of the latter formulations to non-linear $`\sigma `$-models is an advantage.
## 5 Solutions
The metric (20) on M is not in a form most suitable for the construction of harmonic maps. We need to rewite it using new coordinates in such a way that some information about the initial data (10) is already incorporated into the system with the result that we can look for simple harmonic maps automatically satisfying the initial data. For this purpose we start with the wave equation for $`e^U`$ and following Szekeres write its solution as
$$e^U=f(u)+g(v)$$
(27)
where from eq.(8) we know that
$$\begin{array}{ccc}f\hfill & =& \frac{1}{2}\left(1+\frac{a^2}{b^2}\right)sin^2(bv\theta (v)),\hfill \\ g\hfill & =& \frac{1}{2}\left(1+\frac{p^2}{q^2}\right)sin^2(qv\theta (v)).\hfill \end{array}$$
(28)
We can now consider a trivial coordinate transformation
$$uf(u)vg(v)$$
which amouts to a replacement of $`u,v`$ by $`f,g`$ in the Einstein-Maxwell equations (15). We shall further introduce the following definitions which are formally the same as those given by Khan and Penrose
$$\begin{array}{ccccccc}p\hfill & =\hfill & \sqrt{\frac{1}{2}f}\hfill & & q\hfill & =\hfill & \sqrt{\frac{1}{2}g}\hfill \\ r\hfill & =\hfill & \sqrt{\frac{1}{2}+f}\hfill & & w\hfill & =\hfill & \sqrt{\frac{1}{2}+g}\hfill \end{array}$$
(29)
in terms of which it will be convenient to introduce new coordinates on M.
In the vacuum case with non-collinear polarization we had found that
$$\sigma =pwqr\tau =pw+qr$$
(30)
were useful new coordinates because of two reasons. First of all, $`\sigma `$ and $`\tau `$ can be recognized as prolate spheroidal coordinates and we have the Kerr-Tomimatsu-Sato solutions of the main equations. Furthermore, the simplest and the most familiar solution of this type
$$\xi =cos(\alpha )\tau +isin(\alpha )\sigma $$
(31)
satisfies the initial data.
This situation changes for the electrovac case. It turns out that a useful definition of new coordinates is
$$\sigma =\frac{pq}{r+w}\tau =\frac{p+q}{r+w}$$
(32)
in terms of which the metric on M is given by
$$ds^{\mathrm{\hspace{0.33em}2}}=\frac{d\tau ^2}{(1+\tau ^2)^2}\frac{d\sigma ^2}{(1+\sigma ^2)^2}$$
(33)
and from eqs.(27) and (28) we have
$$\begin{array}{cc}e^U\hfill & =\frac{\left(1\sigma ^2\right)\left(1\tau ^2\right)}{\left(1+\sigma ^2\right)\left(1+\tau ^2\right)}\hfill \\ & =1p^2q^2.\hfill \end{array}$$
(34)
Then the main equations become
$$\begin{array}{cc}\frac{\left(1+\sigma ^2\right)^2}{\left(1\sigma ^2\right)}\frac{}{\sigma }\left((1\sigma ^2)\frac{\xi }{\sigma }\right)+\frac{\left(1+\tau ^2\right)^2}{\left(1\tau ^2\right)}\frac{}{\tau }\left((1\tau ^2)\frac{\xi }{\tau }\right)& \\ =\frac{2\overline{\xi }}{\xi \overline{\xi }1}\left[(1+\sigma ^2)^2\left(\frac{\xi }{\sigma }\right)^2+(1+\tau ^2)^2\left(\frac{\xi }{\tau }\right)^2\right]& \end{array}$$
(35)
which is similar to the prolate spheroidal case but differs from it in some important respects. Its advantage lies in the fact that
$$\xi =ϵ\sigma ,ϵ^2=\pm 1$$
(36)
is a solution of eq.(35) that leads to the Bell-Szekeres solution. Furthermore it can be readily verified that $`\xi =ϵ\tau `$ is also a solution as in eq.(31) which is again the Bell-Szekeres solution. In terms of these coordinates the Bell-Szekeres solution is given by
$$ds^{\mathrm{\hspace{0.33em}2}}=\frac{d\tau ^2}{(1+\tau ^2)^2}\frac{d\sigma ^2}{(1+\sigma ^2)^2}\left(\frac{1\tau ^2}{1+\tau ^2}\right)^2dx^2\left(\frac{1\sigma ^2}{1+\sigma ^2}\right)^2dy^2$$
(37)
which may help to clarify the meaning of $`\sigma `$ and $`\tau `$.
So the remaining problem is to a find a one-parameter complex solution of eq.(35) that reduces to eq.(36). Such a solution will be of physical interest as it will automatically satisfy the proper initial data.
## 6 Conclusion
The only proper conclusion of a paper such as this, namely the exact solution of eqs.(35) satisfying the initial data in eqs.(10) is missing. In this paper I have given a list of the essential properties that we must require from a physically acceptable solution describing the interaction of plane impulsive gravitational and electromagnetic shock waves and presented some preparatory material towards such a solution. In my case this solution has been missing since 1978 and that is why I felt it inappropriate to publish the work reported here earlier. However, I now feel that the abundance of so many irrelevant “solutions” of this problem in the literature has made the presentation of the real problem imperative.
## 7 Acknowledgement
This work was in part supported by The Turkish Scientific Research Council TÜBITAK under tbag-cg-1.
|
no-problem/9812/cond-mat9812288.html
|
ar5iv
|
text
|
# Parafermion Statistics and Quasihole Excitations for the Generalizations of the Paired Quantum Hall States
## I Introduction
The quantum Hall plateaus on the first excited Landau levels appear to be different from those of the lowest Landau level. The most striking difference is the $`5/2`$ plateau for which there is no analog in the lowest Landau level.
To explain the $`5/2`$ plateau, several trial wave functions were proposed. One of them, the Haldane-Rezayi state , assumes that the electrons in it are spin unpolarized. Yet another trial wave function, the Pfaffian , assumes that electrons are polarized, just as for the ordinary Laughlin trial wave functions . Recent numerical evidence , in the absence of the Landau level mixing, supports the Pfaffian as a trial wave function for the $`5/2`$ plateau.
In the paper by N. Read and one of us a further generalization of the Pfaffian state was proposed. Moreover, it was conjectured, and supported by the numerical evidence, that these generalizations may be better candidates for other plateaus on the first excited Landau levels than the Laughlin state and its hierarchical or composite fermion generalizations.
An unusual feature of the generalizations of the Pfaffian state given in was that the trial wave functions were found explicitly for nonsimple fractions. That was done with the help of conformal field theory.
Conformal field theory discovered in is essentially a method to solve various scale invariant 1+1 dimensional quantum field theories exactly. It was since proved extremely useful for understanding various 1 dimensional quantum and 2 dimensional classical statistical mechanics systems. Its relevance for the quantum Hall systems was first discussed in .
The relevant conformal field theory for the generalization of the Pfaffian is the so-called parafermion conformal field theory . The parafermions, the direct generalizations of the fermions, play an important role in describing the critical points of the $`Z_k`$ invariant statistical mechanics systems. A particularly well known example of the $`Z_2`$ invariant system is the Ising model. The Ising model can be described by a Majorana fermion. A $`Z_k`$ invariant system must be described, on the other hand, by $`Z_k`$-parafermions.
A Majorana fermion $`\psi (z)`$ can be used to generate the Pfaffian trial wave function, as was discussed in .
$$\mathrm{\Psi }_{\mathrm{Pfaffian}}=\psi (z_1)\psi (z_2)\mathrm{}\psi (z_N)\underset{i<jN}{}(z_iz_j)^2$$
(1)
The filling factor of this trial wave function is $`1/2`$. Moreover, the so-called spin fields of the fermion conformal field theory (the order parameter of the Ising model) $`\sigma `$ can be used to generate the excitations of such a system,
$$\mathrm{\Psi }_{\mathrm{Pfaffian}}^{\mathrm{excited}}(\eta _1,\eta _2,\mathrm{},\eta _{2L})=\sigma (\eta _1),\sigma (\eta _2),\mathrm{},\sigma (\eta _{2L}),\psi (z_1),\psi (z_2),\mathrm{},\psi (z_N)\underset{i,j}{}(\eta _iz_j)^{\frac{1}{2}}\underset{i<j}{}(z_iz_j)^2.$$
(2)
On the other hand, as was discussed in , for each $`k`$ there are $`k1`$ parafermions, denoted $`\psi _l`$, with dimensions $`h_k=l(kl)/k`$. In it was proposed to use the first of them to generate a parafermion wave function,
$$\mathrm{\Psi }_{\mathrm{para}}^M=\psi _1(z_1)\psi _1(z_2)\mathrm{}\psi _1(z_N)\underset{i<jN}{}(z_iz_j)^{M+\frac{2}{k}}$$
(3)
where $`M`$ has to be taken odd or even integer depending on whether the particles are fermions or bosons. The filling factor of this wave function is given by $`\nu =k/(kM+2)`$.
It was further shown in that this wave function should be an exact ground state of a system of $`N`$ bosons in a magnetic field in the presence of the $`k`$-body interaction Hamiltonian
$$H=V\underset{i_1<i_2<\mathrm{}<i_{k+1}}{}\delta ^2(z_{i_1}z_{i_2})\delta ^2(z_{i_2}z_{i_3})\mathrm{}\delta ^2(z_{i_k}z_{i_{k+1}}),$$
(4)
thereby generalizing the Pfaffian case for which the interaction Hamiltonian was (4) at $`k=2`$. We do not need to consider fermions because their wave function can be obtained from that of bosons by a simple Jastrow factor.
To find the wave function directly from (3) is a difficult task. However, it was shown in that the following explicit construction has the correct properties (it is symmetric and vanishes whenever $`k+1`$ particles coincide), which we reproduce for future reference.
To write down $`\mathrm{\Psi }_{\mathrm{para}}^{(0)}`$ we need to break the coordinates of $`N`$ electrons into clusters of $`k`$ ($`N`$ should be divisible by $`k`$). For each pair of distinct clusters, say $`z_1,\mathrm{},z_k`$ and $`z_{k+1},\mathrm{},z_{2k}`$, we define factors $`\chi `$ by
$$\chi _{1,2}(z_1,\mathrm{},z_k;z_{k+1},\mathrm{},z_{2k})=(z_1z_{k+1})(z_1z_{k+2})(z_2z_{k+2})(z_2z_{k+3})\mathrm{}(z_kz_{2k})(z_{2k}z_{k+1}).$$
(5)
The subscript $`1,2`$ of $`\chi _{1,2}`$ labels the clusters (the first, starting with $`z_1`$ and the second, starting with $`z_k`$).
The wave function $`\mathrm{\Psi }_{\mathrm{para}}^{(0)}`$ is defined in terms of these as
$$\mathrm{\Psi }_{\mathrm{para}}^{(0)}=\underset{P}{}\underset{0r<s<N/k}{}\chi _{r+1,s+1}(z_{P(kr+1)},\mathrm{},z_{P(k(r+1))};z_{P(ks+1)},\mathrm{},z_{P(k(s+1))}).$$
(6)
It is not hard to see that for $`k=2`$ this wave function reproduces the Pfaffian for bosons at $`\nu =1`$.
It was also observed in that the quasihole excitations above (6) can be described by the insertion of the spin fields $`\sigma _1`$ of the parafermion conformal field theory. While some of the excitations for the Pfaffian state were found explicitly in , no explicit wave functions for the excitations of the parafermion states have yet been found.
The purpose of this paper is to continue the work begun in . In particular, we explain the numerically observed degeneracies of the excitations of the parafermion states. It turns out that to reproduce the numerically observed degeneracies we need to assume that these excitations obey the parafermion statistics.
Here we would like to emphasize that although the parafermions have been shown in to obey the Haldane exclusion principle in the statistical sense which is a valid description for large quantities of parafermions, for small number of parafermions we need to know the concrete rules which govern the combinatorics of parafermions. These rules asymptotically approach the exclusion statistics rules as the number of parafermions becomes large. These rules are described below and by themselves do not have much to do with Haldane exclusion statistics.
Our finding that the excitations of the parafermion states behave as a combination of bosons and parafermions is a direct generalization of the well known fact that the excitations of the Pfaffian state behave as a combination of bosons and fermions.
We also show that the wave function (6) is indeed given by the parafermion correlation function (3), thus confirming the conjecture of .
## II Excitations of the $`Z_3`$-Parafermion State on a Sphere
In this section we are going to explain how to calculate the degeneracy of the excitations of the parafermion states. To do that, we will have to construct what could be called parafermion quantum mechanics and learn how to add angular momenta of the quantum mechanical particles obeying parafermion statistics.
It was shown in that if you put $`N`$ particles on a sphere with a magnetic monopole in the center, turn on $`k+1`$ particle interaction (4) and then adjust the total flux of the magnetic field through the surface of the sphere to be equal to $`N_\varphi =\nu ^1N(M+2)`$ with $`\nu `$ being the filling factor, $`\nu =k/(Mk+2)`$ and $`M`$ being either odd (for fermions) or even (for bosons) nonnegative integer, you discover that the electrons settle into one unique zero energy ground state with the wave function
$$\mathrm{\Psi }_{\mathrm{GS}}=\underset{i<j}{}(z_iz_j)^M\mathrm{\Psi }(z_1,\mathrm{},z_N),$$
(7)
with $`\mathrm{\Psi }(z_1,\mathrm{},z_N)`$ given by (6). The total angular momentum of that state is equal to $`0`$.
Now if you start increasing the flux by units of one flux quanta, you discover that the zero energy eigenstates of (4) are degenerate. These states can all be grouped into the angular momentum multiplets.
The reason for this is more or less clear. By increasing the flux by 1, we create quasiholes. There is more than one way of creating those quasiholes which explains the degeneracy of states at a higher flux.
As was explained in the elementary quasiholes at arbitrary $`k`$ are not Laughlin quasiholes. Rather, they are quasiholes which carry flux $`1/k`$ and we create $`k`$ of them at once. The wave functions with quasiholes can be found with the help of conformal field theory, by inserting the fields $`\sigma _1`$ of the parafermion conformal field theory . However, we are not going to be interested in the explicit wave functions.
What we want to explain in this section is the numerically observed degeneracy of the states with quasiholes on a sphere and how exactly one could generate all the degeneracies. We will show how the correct angular momentum multiplets can be obtained by putting parafermions into orbitals on a sphere and combining their angular momenta in a way consistent with their fractional statistics. That generalizes the prior treatment of the Pfaffian state in .
Now we would like to present, first without explanation, the rules for finding the degeneracies of the excitations above the $`k=3`$ state, that is, for $`Z_3`$-parafermions.
It was found in , by matching the numerical data, that, as we increase the flux, the degeneracy of states we get for the $`Z_3`$ parafermions is given by the following formulas. The number of excitations at the excess flux 1 (3 quasiholes, each carrying $`1/3`$ of the flux quantum) is given by a binomial coefficient
$$\left(\genfrac{}{}{0pt}{}{N/3+3}{3}\right),$$
(8)
with $`N`$ being the total number of the electrons. $`N`$ should be divisible by three. This coefficient has a simple interpretation as the number of ways you can put 3 bosons into $`N/3+1`$ orbitals. The reason why the excitations at the excess flux $`1`$ behave as bosons was explained in and is essentially the same for the parafermion states as for the Pfaffian state .
Moreover, following , we can assign the angular momentum quantum numbers to these states. For this purpose we interpret the $`N/3+1`$ states as the orbitals on a sphere, in the multiplet of the angular momentum $`L=N/6`$. When we put the bosons on this sphere, we generate the angular momentum multiplets by combining their angular momenta while keeping their wave function totally symmetric.
While we could refer to any standard quantum mechanics textbooks for the rules on how to combine the angular momenta of bosons, we can recast these rules into the following simple form. Let us visualize $`N/3+1`$ orbitals as boxes. Each box has a number $`L_z`$ assigned to it which varies from $`N/6`$ to $`N/6`$ by steps of $`1`$. We are allowed to put as many bosons as we wish into each particular box.
For example, for $`N=6`$ we have $`3`$ boxes with $`L_z=1`$, $`L_z=0`$, and $`L_z=1`$. Fig. 1 shows one possible way to put 3 bosons into 3 boxes, by putting 2 of them into $`L_z=1`$ box and one into $`L_z=1`$ box, thus giving the total $`L_z=1`$. Of course, there are $`\left(\genfrac{}{}{0pt}{}{5}{3}\right)=10`$ ways to put 3 bosons into 5 boxes. Out of these 10 ways, the angular momentum projections $`L_z=\pm 3`$ or $`L_z=\pm 2`$ can be obtained in 1 way each, while the angular momentum projections $`L_z=\pm 1`$ or $`L_z=0`$ can be obtained in 2 different ways each. We immediately recognize the sum of the $`L=3`$ and $`L=1`$ angular momentum multiplets. And indeed, this is confirmed by the numerical data.
As the flux is increased we observe many more excited states. It was found in that the number of states at the excess flux $`2`$ is given by
$$\left(\genfrac{}{}{0pt}{}{N/3+6}{6}\right)+3\left(\genfrac{}{}{0pt}{}{N/3+5}{6}\right)+\left(\genfrac{}{}{0pt}{}{N/3+4}{6}\right),$$
(9)
and at the three excess flux quanta
$$\left(\genfrac{}{}{0pt}{}{N/3+9}{9}\right)+10\left(\genfrac{}{}{0pt}{}{N/3+8}{9}\right)+10\left(\genfrac{}{}{0pt}{}{N/3+7}{9}\right).$$
(10)
The higher fluxes turned out to be harder to analyze as there is less numerical data available for them.
We observe that the number of states at these flux quanta is reminiscent of the corresponding formula for the Pfaffian which was found in to be
$$\underset{F,(1)^F=(1)^N,FN}{}\left(\genfrac{}{}{0pt}{}{n}{F}\right)\left(\genfrac{}{}{0pt}{}{(NF)/2+2n}{2n}\right)$$
(11)
with $`n`$ being the number of the excess flux quanta. The extra $`\left(\genfrac{}{}{0pt}{}{n}{F}\right)`$ was interpreted as the number of ways one can put $`F`$ fermions into $`n`$ boxes. By analogy, we can write down a formula which generalizes (11) to the case of the $`k=3`$ parafermions,
$$\underset{F0,\mathrm{exp}\left(\frac{2\pi Fi}{3}\right)=1,FN}{}\left\{\genfrac{}{}{0.0pt}{}{n}{F}\right\}\left(\genfrac{}{}{0pt}{}{(NF)/3+3n}{3n}\right)$$
(12)
where $`\left\{\genfrac{}{}{0.0pt}{}{n}{F}\right\}`$ is the number of ways you can put $`F`$ $`Z_3`$-parafermions into $`n`$ orbitals.
What does it exactly mean, to put parafermions into orbitals? Inspired by the parafermion mode counting derived in , we present the following rules. While the rules are relatively complicated, they are the only way known to us which allows to obtain $`\left\{\genfrac{}{}{0.0pt}{}{n}{F}\right\}`$ and, at the same time, to break the configurations into angular momentum multiplets.
We replace the boxes by the ‘positions’. As we move from right to left, each next position carries $`1/3`$ more of the $`z`$-component of the angular momentum, as in Fig. 2, ranging from $`(3n2)/6`$ to $`(3n2)/6`$. Thus, we have $`3n1`$ positions. Fig. 2 depicts the positions at $`n=2`$.
There are two types of the $`Z_3`$-parafermions we can put into these positions. We depict them by either the light shaded or dark shaded circles. The following rules should be used to put the parafermions into the positions:
1. The light shaded circle carries $`L_z`$ equal to the $`L_z`$ of its position.
2. The dark shaded circle carries $`L_z`$ equal to twice the $`L_z`$ of its position. In addition, the dark shaded circle is counted as two parafermions.
3. The dark shaded parafermion can occupy any position, while the light shaded parafermion is not allowed to occupy the rightmost or the leftmost position. The number of empty positions to the left of the leftmost parafermion has to be $`3l`$ if it is dark shaded, or $`3l+1`$ if it is light shaded, where $`l`$ is any nonnegative integer, including $`0`$. For example, if the leftmost parafermion is the light shaded one, it can occupy the position number $`2`$, or $`5`$, or in general $`2+3l`$ counting from the left, and if it is a dark shaded one, it can occupy the position number $`1`$, or $`4`$, $`\mathrm{}`$.
4. The adjacent light shaded parafermions are allowed to have $`3l`$ empty positions between them, while the adjacent dark shaded parafermions have to have $`3l+1`$ empty positions between them. If a light shaded parafermions is adjacent to a dark shaded parafermion, they have to be separated by $`3l+2`$ empty positions, where $`l`$ is any nonnegative integer including $`0`$.
Examples of the allowed configurations are shown on Fig. 3. This figure depicts the $`3`$ ways one can put $`3`$ parafermions into $`5`$ positions at $`n=2`$, consistent with (9) and (12). Moreover, we immediately observe that these configurations form an $`L=1`$ multiplet.
It is not very convenient to draw positions all the time, so we introduce the following compact notations. Each light circle will be represented by $`\psi _s`$ with $`s`$ equal to $`L_z`$ of the circle. Each dark circle will be represented by $`\varphi _s`$. For example, the three configurations of the Fig. 3 can be represented as
$$\psi _{\frac{1}{3}}\psi _0\psi _{\frac{1}{3}},\psi _{\frac{1}{3}}\varphi _{\frac{2}{3}},\varphi _{\frac{2}{3}}\psi _{\frac{1}{3}}$$
(13)
(we remember that in accordance with the rule 2, $`\varphi `$ is counted as two parafermions).
There is only one way to put $`6`$ parafermions into $`5`$ positions at $`n=2`$,
$$\varphi _{\frac{2}{3}}\varphi _0\varphi _{\frac{2}{3}},$$
(14)
consistent with $`\left\{\genfrac{}{}{0.0pt}{}{2}{6}\right\}=1`$.
Let us demonstrate the power of the technique by breaking the $`n=3`$ configurations into multiplets. There should be $`8`$ positions, carrying $`L_z`$ from $`7/6`$ to $`7/6`$, that is, $`7/6`$, $`5/6`$, $`3/6`$, $`1/6`$, $`1/6`$, $`3/6`$, $`5/6`$, and $`7/6`$.
It is obvious there is only $`1`$ way to put zero parafermions into these positions.
According to (10) and (12) $`\left\{\genfrac{}{}{0.0pt}{}{3}{3}\right\}=10`$. Indeed, we can count the configurations directly
$`L_z\left(\psi _{\frac{5}{6}}\psi _{\frac{3}{6}}\psi _{\frac{1}{6}}\right)={\displaystyle \frac{3}{2}},L_z\left(\psi _{\frac{5}{6}}\psi _{\frac{3}{6}}\psi _{\frac{5}{6}}\right)={\displaystyle \frac{1}{2}},L_z\left(\psi _{\frac{5}{6}}\psi _{\frac{3}{6}}\psi _{\frac{5}{6}}\right)={\displaystyle \frac{1}{2}},L_z\left(\psi _{\frac{1}{6}}\psi _{\frac{3}{6}}\psi _{\frac{5}{6}}\right)={\displaystyle \frac{3}{2}},`$ (15)
$`L_z\left(\varphi _{\frac{7}{6}}\psi _{\frac{1}{6}}\right)={\displaystyle \frac{5}{2}},L_z\left(\varphi _{\frac{7}{6}}\psi _{\frac{5}{6}}\right)={\displaystyle \frac{3}{2}},L_z\left(\varphi _{\frac{1}{6}}\psi _{\frac{5}{6}}\right)={\displaystyle \frac{1}{2}},`$ (16)
$`L_z\left(\psi _{\frac{5}{6}}\varphi _{\frac{1}{6}}\right)={\displaystyle \frac{1}{2}},L_z\left(\psi _{\frac{5}{6}}\varphi _{\frac{7}{6}}\right)={\displaystyle \frac{3}{2}},L_z\left(\psi _{\frac{1}{6}}\varphi _{\frac{7}{6}}\right)={\displaystyle \frac{5}{2}}.`$ (17)
We observe $`10`$ different configurations. Moreover, we observe that $`L_z=\pm 5/2`$ occurs only once each, and $`L_z=\pm 3/2`$ or $`L_z=\pm 1/2`$ occurs twice each. We immediately recognize the sum of $`L=5/2`$ and $`L=3/2`$ angular momentum multiplets.
Quite analogously, for $`6`$ parafermions at $`n=3`$ we find
$`L_z\left(\psi _{\frac{5}{6}}\psi _{\frac{3}{6}}\psi _{\frac{1}{6}}\psi _{\frac{1}{6}}\psi _{\frac{3}{6}}\psi _{\frac{5}{6}}\right)=0,`$ (18)
$`L_z\left(\psi _{\frac{5}{6}}\psi _{\frac{3}{6}}\psi _{\frac{1}{6}}\psi _{\frac{1}{6}}\varphi _{\frac{7}{6}}\right)=1,L_z\left(\varphi _{\frac{7}{6}}\psi _{\frac{1}{6}}\psi _{\frac{1}{6}}\psi _{\frac{3}{6}}\psi _{\frac{5}{6}}\right)=1,`$ (19)
$`L_z\left(\varphi _{\frac{7}{6}}\varphi _{\frac{3}{6}}\psi _{\frac{3}{6}}\psi _{\frac{5}{6}}\right)=2,L_z\left(\psi _{\frac{5}{6}}\psi _{\frac{3}{6}}\varphi _{\frac{3}{6}}\varphi _{\frac{7}{6}}\right)=2,L_z\left(\varphi _{\frac{7}{6}}\psi _{\frac{1}{6}}\psi _{\frac{1}{6}}\varphi _{\frac{7}{6}}\right)=0,`$ (20)
$`L_z\left(\varphi _{\frac{7}{6}}\varphi _{\frac{3}{6}}\varphi _{\frac{1}{6}}\right)=3,L_z\left(\varphi _{\frac{7}{6}}\varphi _{\frac{3}{6}}\varphi _{\frac{7}{6}}\right)=1,L_z\left(\varphi _{\frac{7}{6}}\varphi _{\frac{3}{6}}\varphi _{\frac{7}{6}}\right)=1,L_z\left(\varphi _{\frac{1}{6}}\varphi _{\frac{3}{6}}\varphi _{\frac{7}{6}}\right)=3.`$ (21)
The total multiplicity is $`10`$, in agreement with (10). Counting $`L_z`$ we observe that $`L_z=\pm 3`$ and $`L_z=\pm 2`$ occurs in one configuration each, while $`L_z=\pm 1`$ and $`L_z=0`$ occurs twice each. We recognize the direct sum of $`L=3`$ and $`L=1`$ multiplets.
One can further check that it is impossible to fit $`9`$ or more parafermions into these positions.
The Tables I, II, and III in the appendix below contain the numerical data available for the Hamiltonian (4) at $`k=3`$ and at the number of electrons fixed at $`N=6`$, $`N=9`$ and $`N=12`$. The data is in the form $`L`$ versus $`n`$. For each excess flux quanta $`n`$ the intersection of $`n`$th column and $`L`$th row contains the multiplicity of the angular momentum $`L`$ at given $`n`$.
To reproduce this data using the parafermion counting rules, we need to combine the angular momenta of the parafermions with those of bosons, by analogy with (12). This is now trivial to do, since parafermions and bosons are distinguishable particles, and their angular momenta combine in accordance with standard rules.
For example, at $`N=6`$ we can obtain the multiplets in the following way. First we need to put $`9`$ bosons into $`3`$ orbitals as represented by the term of (12) with $`F=0`$. Their angular momenta are calculated in exactly the same way as we calculated the angular momenta for $`3`$ bosons in $`3`$ orbitals in the text directly preceding Fig. 1.
Then we have to put $`9`$ bosons into $`2`$ orbitals and combine their angular momenta with the $`L=5/2`$ and $`L=3/2`$ angular momenta of the parafermions, as represented by the term of (12) with $`F=3`$.
And finally, we put $`9`$ bosons into $`1`$ orbital. There is only one such state and its angular momentum is $`0`$. Therefore the total angular momentum is just $`L=3`$ and $`L=1`$, coming from the parafermion contribution.
We are not going to go through explicit counting since it is rather standard. The only nontrivial step was the parafermion angular momentum contribution, and that was worked out above. We have checked, however, that the result reproduces the numerical data given in the Table I. In fact, we have verified all the data in all three Tables, expect the results for $`n=5`$ at $`N=6`$. In particular, we found that the parafermion multiplets at $`n=4`$ are the following. At $`F=3`$, there is one each of $`L=0,2,3,4`$ states giving the multiplicity of $`22`$ (compare with (25)). At $`F=6`$, there are two each of $`L=0,2,4`$ states, and one each of $`L=3,6`$ states giving the overall multiplicity of $`50`$. For $`F=9`$ it is one of each $`L=0,2,4`$, with a total multiplicity of $`15`$. And for $`F=12`$ there is one $`L=0`$ state.
It would be nice if we could obtain the parafermion multiplicities $`\left\{\genfrac{}{}{0.0pt}{}{n}{F}\right\}`$ in a more closed form rather than by having to count configurations all the time. The most closed form that we are aware of can be obtained with the help of the equation for the partial partition functions derived in .
We define the polynomials $`Y_l(x)`$ in the following way.
$$Y_{l+1}(x)=xY_{l+\frac{2}{3}}(x)+x^2Y_{l+\frac{1}{3}}(x)+(1x^3)Y_l(x),$$
(22)
with the initial conditions $`Y_{\frac{1}{3}}=0`$, $`Y_0=1`$, and $`Y_{\frac{1}{3}}=0`$. Then we claim that the following expansion is valid
$$Y_n(x)=\underset{k=0,1,\mathrm{}}{}\left\{\genfrac{}{}{0.0pt}{}{n}{3k}\right\}x^{3k}.$$
(23)
It is in fact possible to derive (22) directly from the counting rules presented above, see .
It is not hard to check that
$`Y_1=1,Y_2=1+3x^3+x^6,Y_3=1+10x^3+10x^6`$ (24)
in agreement with (9) and (10). We can continue further,
$$Y_4=1+22x^3+50x^6+15x^9+x^{12},Y_5=1+40x^3+168x^6+140x^9+28x^{12}.$$
(25)
These multiplicities are also in agreement with the available numerical data for the degeneracies at various flux quanta $`n`$.
We note that according to the sum of the multiplicities at fixed $`n`$ should give us Fibonacci numbers $`F_{3n2}`$,
$$\underset{k=0,1,2,\mathrm{}}{}\left\{\genfrac{}{}{0.0pt}{}{n}{3k}\right\}=F_{3n2}.$$
(26)
That is indeed true, since this sum is equal to $`Y_n(1)`$, and by substituting $`x=1`$ into (22) we obtain
$$Y_{l+1}(1)=Y_{l+\frac{2}{3}}(1)+Y_{l+\frac{1}{3}}(1),$$
(27)
which is the defining relation for the Fibonacci numbers.
## III Conformal Field Theory and the Counting Rules
The rules we have just presented were in fact deduced from the parafermion conformal field theory. Here we would like to explain how conformal field theory should be employed to derive these rules.
While we were interested in parafermion quantum mechanics, conformal field theory, as a field theory, gives us the counting rules in the second quantized formalism. It is not so hard to read statistics off the second quantized formalism.
Let us concentrate on the Majorana fermions as an example. We know, of course, that there are $`\left(\genfrac{}{}{0pt}{}{n}{F}\right)`$ ways to put $`F`$ fermions into $`n`$ boxes, in accordance with Pauli exclusion principle. This can be obtained also from the conformal field theory of the Majorana fermions.
Expanding the Majorana fermion $`\psi _n`$ in terms of modes we get
$$\psi (z)=\underset{n}{}\psi _nz^{n\frac{1}{2}},$$
(28)
with $`n`$ going over half integers $`1/2+ZZ`$. The modes $`\psi _n`$ obey the anticommutation relations,
$$\psi _n\psi _m+\psi _m\psi _n=\delta _{n+m,0}.$$
(29)
All these are well known facts of conformal field theory.
Since the Majorana fermion defined at every point in space and time cannot be infinite when acting on the vacuum, $`\psi (z)|0`$ for any $`z`$ including $`z=0`$, it follows that $`\psi _n|0=0`$ for all $`n=1/2,3/2,\mathrm{}`$. We say that the modes of $`\psi (z)`$ with positive index act as annihilation operators. The modes $`\psi _n`$ with $`n=1/2,3/2,\mathrm{}`$ act as creation operators. We can use them to create new states. A generic state will look like
$$\psi _{n_F}\psi _{n_{F1}}\mathrm{}\psi _{n_1}|0,$$
(30)
with all the $`n_i`$ being negative half integers. Note that some of the states in (30) are not linearly independent and can actually be transformed into each other by repeatedly using the anticommutation relations (29). We can select a linear independent subset of (30) by ordering all the $`n`$, say making them increase from left to right, and making sure neither of them are equal to each other, thereby making the state zero according to $`\psi _n^2=0`$.
Now observe that there is the following correspondence between the states (30) and the angular momentum wave functions. Consider the states (30) together with the relevant powers of the coordinates, as in (28).
$$\psi _{n_F}\psi _{n_{F1}}\mathrm{}\psi _{n_1}|0z_1^{n_F\frac{1}{2}}z_2^{n_{F1}\frac{1}{2}}\mathrm{}z_F^{n_1\frac{1}{2}}.$$
(31)
Let us now sum (31) over all the states accessible via the anticommutation relations (29). In other words, we want to sum over all the permutations $`\sigma `$ of the numbers $`n_1`$, $`n_2`$, $`\mathrm{}`$, $`n_F`$. Obviously we obtain
$`{\displaystyle \underset{\sigma }{}}\psi _{\sigma (n_F)}\psi _{\sigma (n_{F1})}\mathrm{}\psi _{\sigma (n_1)}|0z_1^{\sigma (n_F)\frac{1}{2}}z_2^{\sigma (n_{F1})\frac{1}{2}}\mathrm{}z_F^{\sigma (n_1)\frac{1}{2}}=`$ (32)
$`\psi _{n_F}\psi _{n_{F1}}\mathrm{}\psi _{n_1}|0{\displaystyle \underset{\sigma }{}}\mathrm{sign}(\sigma )z_1^{\sigma (n_F)\frac{1}{2}}z_2^{\sigma (n_{F1})\frac{1}{2}}\mathrm{}z_F^{\sigma (n_1)\frac{1}{2}}.`$ (33)
One immediately recognizes the totally antisymmetric polynomials of the sort discussed in . The angular momentum $`L_z`$ computed on these polynomials can be defined, up to an additive constant, as generated by the operator $`z\frac{}{z}`$, to give $`_in_iF/2`$. These totally antisymmetric polynomials are in fact the wave functions of the fermions in the first quantized formalism!
If we consider all possible states (32), with the restrictions $`n_N<n`$, $`n`$ being some integer, there is going to be exactly $`\left(\genfrac{}{}{0pt}{}{N}{F}\right)`$ of them which are the multiplicities in (11). Now we see how these states break naturally into angular momentum multiplets. One can check that the multiplets we obtain in this way matches the numerical data for the Pfaffian state .
What we just did so far looks like a complicated and not very intuitive way of rederiving the results of . However, as we move on to the parafermion states, conformal field theory becomes the only consistent and dependable way to construct the angular momenta multiplets.
$`k=3`$ parafermion conformal field theory contains two fields of dimension $`2/3`$, $`\psi ^{(1)}(z)`$ and $`\psi ^{(2)}(z)`$ . Expanding them in terms of modes we obtain
$$\psi ^{(1)}(z)=\underset{n}{}\psi _n^{(1)}z^{n\frac{2}{3}},\psi ^{(2)}(z)=\underset{n}{}\psi _n^{(2)}z^{n\frac{2}{3}}.$$
(34)
It is obvious that the modes $`\psi _n^{(1)}`$ and $`\psi _n^{(2)}`$ have to annihilate the vacuum if $`n>2/3`$. Applying the modes with $`n2/3`$ to the vacuum we can in fact create parafermionic states. Not all the states we create in this way are linearly independent. To check which ones are independent we have to employ the generalized commutation relations of the sort derived in . These relations are very complicated. Fortunately for $`k=3`$ the set of independent states have already been derived in . Here for completeness we are going to quote the answer.
The full set of linearly independent states is created by applying the modes of the parafermions of the first kind $`\varphi _s^{(1)}\psi _s^{(1)}`$ or a certain combination of the modes $`\varphi _s^{(2)}\psi _s^{(1)}\psi _{s\frac{2}{3}}^{(1)}`$. We do not need to use the modes of the parafermion of the second kind $`\psi ^{(2)}`$ as its modes create states which are linearly dependent on the states created by $`\varphi _s^{(1)}`$ and $`\varphi _s^{(2)}`$.
The allowed states have the form
$$\varphi _{s_N}^{\left(i_N\right)}\mathrm{}\varphi _{s_2}^{\left(i_2\right)}\varphi _{s_1}^{\left(i_1\right)}|0,$$
(35)
with the spacing specified as
$`\mathrm{if}i_{l+1}=1,i_l=1\mathrm{then}s_{l+1}s_l=m+{\displaystyle \frac{1}{3}},`$ (36)
$`\mathrm{if}i_{l+1}=2,i_l=1\mathrm{then}s_{l+1}s_l=m+{\displaystyle \frac{2}{3}},`$ (37)
$`\mathrm{if}i_{l+1}=1,i_l=2\mathrm{then}s_{l+1}s_l=m+{\displaystyle \frac{4}{3}},`$ (38)
$`\mathrm{if}i_{l+1}=2,i_l=2\mathrm{then}s_{l+1}s_l=m+{\displaystyle \frac{2}{3}},`$ (39)
where $`m`$ is any nonnegative integer number including $`0`$.
We note that these states are in one to one correspondence with the counting rules of the previous section. In fact, this is how the counting rules should be derived.
We can go slightly further and conjecture a way to derive the parafermion wave function. We need to consider the sum over all the states which can be obtained from one of the states (35) by the generalized commutation relations
$$\underset{\mathrm{dependent}\mathrm{states}}{}\psi _{s_N}^{\left(j_N\right)}\mathrm{}\psi _{s_1}^{\left(j_1\right)}|0z_N^{s_N\frac{2}{3}}\mathrm{}z_1^{s_1\frac{2}{3}},$$
(40)
with different $`j`$ being either $`1`$ or $`2`$. By using those generalized commutation relations we can in principle bring the sum to the form (compare with (32))
$$\varphi _{s_N}^{\left(i_N\right)}\mathrm{}\varphi _{s_2}^{\left(i_2\right)}\varphi _{s_1}^{\left(i_1\right)}|0f(z_1,\mathrm{},z_N).$$
(41)
We interpret $`f(z_1,\mathrm{},z_N)`$ to be the wave function of the parafermions!
While finding $`f`$ explicitly is beyond the scope of this paper, we note that it indeed combines the single particle states of the form $`z^s`$ into the multiparticle wave function obeying the right particle exchange properties.
One should also in principle be able to derive $`f`$ from the quasihole wave functions in direct analogy with the corresponding derivation for the Pfaffian state . This also lies beyond the scope of this paper.
At the end we would like to note that for $`k>3`$ the mode counting rules can also be derived by using the generalized commutation relations of . Some partial results at $`k=4`$ and $`k=5`$ are presented in .
## IV The Correlation Functions of the Parafermion CFTs
All the preceding discussion of this paper was based on the assumption that the wave function (6) can be generated as a correlation function of the parafermions, in the sense of .
While it was conjectured in that (6) is such a correlation function, no proof was found. It is important to prove this relationship, otherwise, our manipulations with parafermions lose their relevance.
In this section we would like to present a proof that the wave function found in , which is the zero energy state of the $`k+1`$ particle interaction Hamiltonian (4), is indeed the correlation function of a parafermion conformal field theory. For this purpose we recall that a parafermion conformal field theory consists of $`k1`$ fields $`\psi _l`$, $`l=1,\mathrm{},k1`$ with the operator product expansion
$$\psi _l(z)\psi _l^{}(z^{})c_{l,l^{}}(zz^{})^{2ll^{}/k}(\psi _{l+l^{}}+\beta _{l,l^{}}(zz^{})\psi _{l+l^{}}+\mathrm{}),$$
(42)
where $`c`$ and $`\beta `$ are certain numbers – structure constants. The index $`l`$ of the fields $`\psi _l`$ in (42) has to be understood in the mod $`k`$ sense, $`\psi _{k+l}\psi _l`$. Additionally, $`\psi _0`$ is identified with the unit operator. In that case, $`\psi _0`$ vanishes and therefore, when $`l+l^{}=0`$ mod $`k`$, the linear term in (42) proportional to $`\beta `$ vanishes.
According to the conjecture of the correlation function of $`N`$ fields $`\psi _1`$ is equal to
$$\psi _1(z_1)\mathrm{}\psi _N(z_N)=\frac{\mathrm{\Psi }(z_1,z_2,\mathrm{},z_N)}{_{i<j}(z_iz_j)^{\frac{2}{k}}}$$
(43)
with $`\mathrm{\Psi }`$ given by (6).
To show that the right hand side of (43) is indeed consistent with (42), we will glue $`k1`$ parafermions $`\psi _1`$ together to obtain the correlation function with $`\psi _{k1}`$. Then we will glue $`\psi _{k1}`$ and $`\psi _1`$ together and show that the correlation function is consistent with the fact that
$$\psi _{k1}(z)\psi _1(z^{})(zz^{})^{2\frac{k1}{k}}\psi _0+\mathrm{}$$
(44)
with a vanishing first derivative $`\psi _0`$.
The proportionality sign in (44) and throughout this section means that an unimportant numerical constant has been dropped.
We start with taking the limit $`z_1z_2`$. In that limit the right hand side of (43) is proportional to $`(z_1z_2)^{2/k}`$. This is indeed consistent with the operator product expansions
$$\psi _1(z)\psi _1(z^{})(zz^{})^{\frac{2}{k}}\psi _2(z^{})+\mathrm{}$$
(45)
By multiplying (43) by $`(z_1z_2)^{2/k}`$ and taking $`z_1z_2`$ we arrive at the following correlation function
$$\psi _2(z_2)\psi _1(z_3)\mathrm{}\psi _1(z_N)\frac{\mathrm{\Psi }(z_2,z_2,z_3,\mathrm{},z_N)}{_{i>2}(z_2z_i)^{\frac{4}{k}}_{2<i<jN}(z_iz_j)^{\frac{2}{k}}}.$$
(46)
As $`z_2`$ approaches $`z_3`$ the right hand side of (46) has just the right singularity $`(z_2z_3)^{4/k}`$, matching the singularity of (42) as $`\psi _2`$ approaches $`\psi _1`$. Therefore, we continue this process further and take $`z_2z_3`$. By doing so we indeed recover the correlation function with the field $`\psi _3`$,
$$\psi _3(z_3)\psi _1(z_4)\mathrm{}\psi _1(z_N)\frac{\mathrm{\Psi }(z_3,z_3,z_3,z_4,\mathrm{},z_N)}{_{i>3}(z_3z_i)^{\frac{6}{k}}_{3<i<jN}(z_iz_j)^{\frac{2}{k}}}.$$
(47)
It is clear at this point that as we continue ‘gluing’ parafermion fields together at some point we will arrive at
$$\psi _{k1}(z_{k1})\psi _1(z_k)\mathrm{}\psi _1(z_N)\frac{\mathrm{\Psi }(z_{k1},z_{k1},\mathrm{},z_{k1},z_k,z_{k+1},\mathrm{},z_N)}{_{i>k1}(z_{k1}z_i)^{\frac{2(k1)}{k}}_{k1<i<jN}(z_iz_j)^{\frac{2}{k}}}.$$
(48)
At this stage we should be careful. By taking $`z_{k1}`$ to $`z_k`$ we should be able to recover the identity operator $`\psi _k\psi _0`$. Let us check that this is indeed the case. The main singularity of (48) as $`z_{k1}z_k`$ is $`(z_{k1}z_k)^{\frac{k1}{k}}`$ which indeed matches (42) and therefore,
$$\psi _0(z_k)\psi _1(z_{k+1})\mathrm{}\psi _1(z_N)\frac{\mathrm{\Psi }(z_k,z_k,\mathrm{},z_k,z_{k+1},z_{k+2},\mathrm{},z_N)}{_{i>k}(z_kz_i)^2_{k<i<jN}(z_iz_j)^{\frac{2}{k}}}.$$
(49)
By its definition, the identity operator does not depend on its position, and therefore the right hand side of (49) should not depend on $`z_k`$. To see that, let us recall that in the following theorem was proved. The wave function $`\mathrm{\Psi }`$ for $`N`$ particles whose first $`k`$ particles live at the same point $`z_k`$ can be expressed in terms of the wave function for $`Nk`$ particles in the following way,
$$\mathrm{\Psi }(z_k,z_k,\mathrm{},z_k,z_{k+1},\mathrm{},z_N)\underset{k<iN}{}(z_kz_i)^2\mathrm{\Psi }(z_{k+1},z_{k+2},\mathrm{},z_N).$$
(50)
Substituting (50) to (49) we see that
$$\psi _0(z_k)\psi _1(z_{k+1})\mathrm{}\psi _1(z_N)\frac{\mathrm{\Psi }(z_{k+1},z_{k+2},\mathrm{},z_N)}{_{k<i<jN}(z_iz_j)^{\frac{2}{k}}},$$
(51)
that is, $`\psi _0`$ is indeed an identity operator. The correlation function is insensitive to its insertion. That completes the first part of our proof.
If we continue to expand (48) in powers of $`z_{k1}z_k`$ we observe that we should reproduce the operator product expansion (42) with $`l+l^{}=0`$ mod $`k`$. In particular, we must see that the linear term of that expansion vanishes, $`\beta _{k1,1}=0`$. If this term didn’t vanish, that would mean that in addition to the parafermion fields, the conformal field theory whose correlation function is given by (43) has a dimension 1 operator $`J`$ which generates a U$`(1)`$ symmetry (see ). Such an operator should be absent in the parafermion theory. Let us check that it is indeed absent.
Continuing the expansion of (48) in powers of $`z_{k1}z_k`$ we find that the linear term is obviously proportional to
$$\frac{}{z_{k1}}\left[\frac{\mathrm{\Psi }(z_{k1},z_{k1},\mathrm{},z_{k1},z_k,z_{k+1},\mathrm{},z_N)}{_{i>k}(z_{k1}z_i)^{\frac{2(k1)}{k}}}\right]|_{z_{k1}=z_k}.$$
(52)
To show that this term is zero it is sufficient to show that
$$\frac{}{z_{k1}}\mathrm{\Psi }(z_{k1},z_{k1},\mathrm{},z_{k1},z_k,z_{k+1},\mathrm{},z_N)|_{z_{k1}=z_k}=\frac{2(k1)}{k}\left(\underset{i>k}{}\frac{1}{z_kz_i}\right)\mathrm{\Psi }(z_k,z_k,\mathrm{},z_k,z_{k+1},\mathrm{},z_N).$$
(53)
We can further simplify the equality (53) by noting that since $`\mathrm{\Psi }`$ is a symmetric function of its arguments, it is enough to differentiate it with respect to just one variable,
$$\frac{}{z_k}\mathrm{\Psi }(z_{k1},z_{k1},\mathrm{},z_{k1},z_k,z_{k+1},\mathrm{},z_N)|_{z_{k1}=z_k}=\frac{2}{k}\left(\underset{i>k}{}\frac{1}{z_kz_i}\right)\mathrm{\Psi }(z_k,z_k,\mathrm{},z_k,z_{k+1},\mathrm{},z_N).$$
(54)
To prove (54) we again recall the definition of $`\mathrm{\Psi }(z_1,\mathrm{},z_N)`$ as given in (6). To construct it, we break all its coordinates into $`N/k`$ clusters of $`k`$ coordinates and then sum a certain expression, written in terms of these clusters, over all the permutations of $`N`$ particles. Let us first look at the right hand side of (54). In accordance with the theorem proved in , the only terms which contribute to $`\mathrm{\Psi }(z_k,\mathrm{},z_k,z_{k+1},\mathrm{},z_N)`$ in the sum over permutations (6) are those where all the $`z_k`$ belong to the same cluster. This is how (50) could be derived. Let us now make an assumption, which we will justify later, that the only terms which contribute to the left hand side of (54) when we substitute (6) for $`\mathrm{\Psi }`$ are also those where all the $`z_{k1}`$ and $`z_k`$ belong to the same cluster.
It is possible to convince oneself that after some algebra the sum of these terms can be reduced to
$$\mathrm{\Phi }=\underset{r=1}{\overset{k}{}}\underset{P}{}\underset{0<j<N/k}{}\frac{(z_kz_{P\left(r+kj\right)})(z_kz_{P\left(r+kj+s\left(r\right)\right)})}{(z_{k1}z_{P\left(r+kj\right)})(z_{k1}z_{P\left(r+kj+s\left(r\right)\right)})}\underset{k<iN}{}(z_{k1}z_i)^2\underset{0<j<l<N/k}{}\chi _{j+1,l+1},$$
(55)
where $`P(n)`$ gives a permutation of the integer numbers $`k+1`$, $`k+2`$, $`\mathrm{}`$, $`N`$. And $`s(n)=n+1`$ for $`0<n<k`$ with $`s(k)=1`$. $`\chi `$ are the expressions (5). Note that $`\chi `$ do not depend on $`z_{k1}`$ and $`z_k`$. Applying the derivatives as in (54) we get
$`{\displaystyle \frac{}{z_k}}\mathrm{\Phi }|_{z_{k1}=z_k}=`$ (56)
$`{\displaystyle \underset{r=1}{\overset{k}{}}}{\displaystyle \underset{P}{}}{\displaystyle \underset{0<j<N/k}{}}\left({\displaystyle \frac{1}{z_kz_{P\left(r+kj\right)}}}+{\displaystyle \frac{1}{z_kz_{P\left(r+kj+s\left(r\right)\right)}}}\right){\displaystyle \underset{k<iN}{}}(z_kz_i)^2{\displaystyle \underset{0<j<l<N/k}{}}\chi _{j+1,l+1}=`$ (57)
$`{\displaystyle \frac{2}{k}}\left({\displaystyle \underset{k<iN}{}}{\displaystyle \frac{1}{z_kz_i}}\right)\mathrm{\Phi }|_{z_{k1}=z_k}.`$ (58)
The last line in (56) follows from the fact that the summation over $`r`$ completely symmetrizes the sum in the second line of (56) over all the permutations $`P`$.
Therefore we have proved (54) on the assumptions that the only terms that contribute are those where the first $`k`$ coordinates of $`\mathrm{\Psi }`$ belong to the same cluster. To see that it is indeed so, let us recall again that according to the theorem proved in all such term should vanish as $`z_{k1}`$ approach $`z_k`$. Since they are polynomials they vanish at least as $`z_kz_{k1}`$, or perhaps even faster. If they vanish faster, their derivative with respect to $`z_k`$ with the setting $`z_{k1}=z_k`$ after differentiation is definitely zero. If, on the other hand, they vanish linearly, then we can argue that for each term vanishing as $`z_kz_{k1}`$ there is another permutation of the coordinates different from the first one by exchanging exactly the two coordinates in that difference. This term will be proportional to $`z_{k1}z_k`$. After differentiating and setting $`z_k=z_{k1}`$, these terms will cancel each other.
This concludes the proof that the wave function conjectured in to be the correlation function of the parafermions is indeed the correlation function of the parafermions. This in fact allows us to use the parafermionic conformal field theory to describe the ground states of the hamiltonian (4), as we did throughout this paper.
## V Acknowledgements
The authors are grateful to N. Read for important discussions and to K. Schoutens for explaining the results of his paper . This work was initiated and completed during the ITP program “Disorder and Interaction in Quantum Hall and Mesoscopic Systems” and was supported by the NSF grants PHY-94-07194 and DMR-9420560 (ER). ER is also grateful to ITP for an ITP Scholar award.
## A Numerical Data
In this appendix we present the numerical data for the degeneracy of the angular momenta multiplets at various excess flux quanta $`n`$. Columns of the tables are labeled by $`n`$. Rows are labeled by the angular momentum $`L`$. The intersection of the $`n`$th column and $`L`$th row gives us the degeneracy of the angular momentum $`L`$ multiplets at a given $`n`$. The three tables represent the angular momentum multiplets at the total particle number $`N=6`$, $`N=9`$, and $`n=12`$. All the data is at $`k=3`$.
|
no-problem/9812/astro-ph9812160.html
|
ar5iv
|
text
|
# Multiwavelength Observations of GX 339–4 in 1996. I. Daily Light Curves and X-ray and Gamma-Ray Spectroscopy
## 1 Introduction and Overview
Most Galactic black hole candidates exhibit at least two distinct spectral states (see Liang & Narayan 1997, Liang 1998, Poutanen 1998 for reviews). In the hard state (= soft X-ray low state) the spectrum from $``$ keV to a few hundred keV is a hard power law (photon index $`1.5\pm 0.5`$) with an exponential cutoff. This can be interpreted as inverse Comptonization of soft photons. In the soft state (often, but not always, accompanied by the soft X-ray high state), the spectrum above $`10`$ keV is a steep power law (photon index $`>2.2`$) with no detectable cutoff out to $``$ MeV. This multi-state behavior is seen in both persistent sources (e.g. Cygnus X-1) and transient black hole X-ray novae (e.g. GRS 1009–45). GX 339–4 is unusual in that it is a persistent source, being detected by X-ray telescopes most of the time, but it also has nova-like flaring states.
In 1996, we performed a series of multiwavelength observations of GX 339–4 when it was in a hard state (= soft X-ray low state). This paper is one of a series that describes the results of that campaign. In Paper II (Smith & Liang (1999)) we discuss the rapid X-ray timing variability in an observation made by the Rossi X-Ray Timing Explorer (RXTE) 1996 July 26 UT. In Paper III (Smith, Filippenko, & Leonard (1999)) we discuss our Keck spectroscopy performed on 1996 May 12 UT. These papers expand significantly on our preliminary analyses (Smith et al. 1997a ; Smith et al. 1997b ).
We start this paper by presenting in §2 the radio spectra from 1996 July taken when there was a possible radio jet (Fender et al. (1997)). We show that the radio spectrum is flat and significantly variable, and that the spectral shape and amplitude were not unusual for the source at that time.
In §3 we discuss the observing and data analysis details for our pointed observation 1996 July 9-23 using the Oriented Scintillation Spectrometer Experiment (OSSE) on the Compton Gamma-Ray Observatory (CGRO).
In §4 we show the daily X-ray and gamma-ray light curves from 1996 July. In addition to our OSSE results, we show the data obtained by the All Sky Monitor (ASM) on RXTE and by the Burst and Transient Source Experiment (BATSE) on CGRO using the Earth occultation technique. The soft X-ray flux during our observations was very low, and the spectrum was very hard. However, unlike at the time of our Keck observation on 1996 May 12, the source was significantly detected during most of the month. Most importantly, there was no significant change in the higher energy emission during the time the radio jet-like feature was seen.
In §5, we present the OSSE spectrum and fit it with simple power law times exponential (PLE) and Sunyaev-Titarchuk (ST; Sunyaev & Titarchuk (1980)) functions. They both give equally good fits to the OSSE data alone.
In §6, we discuss the observing and data analysis details for the 1996 July 26 RXTE observation.
In §7, we present the X-ray spectroscopy from the RXTE observation. Both a PLE and a ST model can explain the RXTE data alone. An additional soft component is required, as well as a broad emission feature centered on $`6.4`$ keV. This may be an iron line that is broadened by orbital Doppler motions and/or scattering off a hot medium. Its equivalent width is $`600`$ eV.
Combining the RXTE and CGRO spectra, we show in §8 that the PLE model easily fits the joint spectrum, but that the ST model does not. In Böttcher, Liang, & Smith (1998) we use the GX 339–4 spectral data to test our detailed self-consistent accretion disk corona models.
## 2 Radio Observations
The radio counterpart to GX 339–4 was discovered in 1994 by the Molonglo Radio Observatory, Australia (MOST) at 843 MHz (Sood & Campbell-Wilson (1994)). Monitoring over the following years has found that the radio emission is variable, with a flux density $`10`$ mJy (Sood et al. (1997); Corbel et al. (1997); Hannikainen et al. (1998)).
High resolution 3.5 cm radio observations with the Australia Telescope Compact Array (ATCA) detected a possible jet-like feature 1996 July 11–13 (MJD 50275–7; Fender et al. (1997)). An ATCA observation 1997 Feb 3 may have a small extension in the direction opposite this jet, but no strongly significant confirmation of the jets has been found (Corbel et al. (1997)).
Figure 1 shows the daily radio spectra in 1996 July. All the data points are from ATCA, except the one at 843 MHz from MOST (Hannikainen et al. (1998)). The ATCA observations are the ones made at the time that the possible jet-like feature was detected: see Fender et al. (1997) for the observing details.
The radio spectrum is approximately flat, and shows a significant variability. The spectral shape and amplitude were not anomalous for this source during the time of the possible radio jet.
The radio emission suggests the continual presence of an outflow in this state. The approximately flat spectrum is not consistent with optically thin emission (unless the electron distribution is exceptionally hard) and is indicative of some absorption in the radio-emitting regions. One possible geometry would be a conical jet such as discussed in Hjellming & Johnston (1988).
## 3 OSSE Observations and Reductions
OSSE observed GX 339–4 as a ToO during an outburst for 1 week beginning 1991 September 5. The OSSE flux was $`300`$ mCrab between 50 and 400 keV (Grabelsky et al. (1995)). A second 1 week observation was carried out beginning 1991 November 7, when the source was $`40`$ times weaker (Grabelsky et al. (1995)). Our observation was made 1996 July 9-23 (MJD 50273–287). The flux is a factor of $`2`$ lower than that found in the (brighter) 1991 September observation.
The OSSE instrument consists of four separate, nearly identical NaI-CsI phoswich detectors with a field of view $`3.8\times 11.4`$ degrees FWHM (see Johnson et al. (1993) and Grabelsky et al. (1995) for instrumental details and observing techniques). Our observations consist of a series of alternating on- and off-source pointings with 2.048 minutes spent on each pointing. The on-source pointings (at $`l=339.0\mathrm{°}`$, $`b=3.7\mathrm{°}`$) were centered close to GX 339–4 ($`l=338.9\mathrm{°}`$, $`b=4.3\mathrm{°}`$), with the long dimension of the collimator oriented perpendicular to the Galactic plane. The total on-source observing time (per detector) was $`1.59\times 10^5`$ seconds. The off-source pointings were located along the galactic plane at $`\pm 12\mathrm{°}`$ from the on-source pointing (they were centered at $`l=351.0\mathrm{°}`$, $`b=3.2\mathrm{°}`$ and $`l=327.0\mathrm{°}`$, $`b=4.1\mathrm{°}`$). Both backgrounds gave the same results, indicating there were no bright sources in either one.
Version 7.4 of the IGORE OSSE data analysis package was used to subtract the background and to generate the daily light curves and spectra in the standard way (Johnson et al. (1993)). To perform joint fits with the RXTE spectrum, we used IGORE to generate a spectrum and response matrix that could be read by XSPEC.
Grabelsky et al. (1995) estimated the contribution from the diffuse component from the Galactic plane. They found that this is a negligible effect when the source is bright, as in our observation. The spectrum of the diffuse component also has a similar shape to that of GX 339–4. We have therefore not subtracted an estimate for the diffuse component.
We checked all of our results for the individual detectors separately, and found no discrepancies. We have therefore combined all four detectors in all the results presented here.
## 4 Daily Light Curves
Figure 2 shows the daily light curves for our OSSE data in the 50–70 and 70–270 keV bands, as well as the hardness ratio of these two bands. These are combined with the data from the RXTE ASM and BATSE Earth occultation. The BATSE light curve assumes an optically thin thermal bremsstrahlung (OTTB) model with a fixed $`kT=60`$ keV (Rubin et al. (1998)).
The soft X-ray flux in 1996 July was low, and the spectrum was very hard. However, unlike at the time of our Keck observation on 1996 May 12, the source was significantly detected during most of the month. During the month, the fluxes at all energies were generally rising. Figure 1 of Smith et al. (1997a) and Figure 1 of Rubin et al. (1998) show that the X-ray and gamma-ray fluxes peaked on $``$ TJD 50290.
A constant does not fit the OSSE 50–70 keV light curve. The reduced $`\chi _\nu ^2`$= 3.86 for $`\nu =14`$ degrees of freedom: the probability that a random set of data points would give a value of $`\chi _\nu ^2`$ as large or larger than this is $`Q=1.3\times 10^6`$. Similarly, a constant does not fit the OSSE 70–270 keV light curve with $`\chi _\nu ^2`$= 2.88, $`\nu =14`$, $`Q=2.2\times 10^4`$. For both bands, the linear relationships shown in Figure 2 give good fits ($`Q=0.35,0.44`$ respectively).
While there is an indication from the OSSE (70–270 keV)/(50–70 keV) hardness ratio that the spectrum may be softening during the two week observation, this is not statistically significant. A constant fit to the hardness ratio gives acceptable results, $`\chi _\nu ^2`$= 1.11, $`\nu =14`$, $`Q=0.34`$. This fact and the linear rise in the flux are important in §8, when we combine the non-simultaneous RXTE and OSSE spectra.
Most importantly, there was no significant change in the higher energy emission during the time the possible radio jet-like feature was seen. One might have expected that the physics behind the jet formation could have led to a significant change in the higher energy emission. But since the radio flux was not unusually bright during this time, it is possible that the energy release in this case was relatively small. Further multiwavelength observations during radio jet or radio flaring events will be very important to understanding whether the synchrotron emitting electrons are involved in the Compton scattering that produces the hard X-rays. In other black hole candidates, violent changes in the high energy emission may or may not be associated with large radio flares, e.g. GRO J1655–40 (Harmon et al. (1995); Tavani et al. (1996)).
## 5 OSSE Spectra
The OSSE spectrum was extracted by averaging over the whole two week observation. Since the hardness ratio did not change significantly over this time, this gives a reliable measure of the spectral shape during these two weeks, though Figure 2 shows that the normalization on any given day will be different from the average presented here.
The OSSE spectrum was fit over the 0.05–10 MeV energy range using a forward folding technique. Concerns about the OSSE instrumental response and the precision of the cross calibration of the detectors (Grabelsky et al. (1995)) are mitigated by looking at the results for the individual detectors separately. We found that the results were consistent in all four detectors, and thus we present the results for the four detectors combined.
Throughout this paper we will only consider simple phenomenological models for the spectral fitting: see Böttcher, Liang, & Smith (1998) for detailed self-consistent accretion disk corona model fits to all the GX 339–4 data. In the OSSE range, the spectrum has a power law shape with a cut-off. We found that it is only necessary to use one model component to explain the OSSE spectrum, and we consider two simple functional forms:
1) Power law times exponential (PLE). In this model, the flux has the form $`F(E)E^\alpha \mathrm{exp}(E/kT)`$. An optically thin Compton spectrum can be roughly described using this model (Haardt et al. (1993)).
(2) Sunyaev-Titarchuk function (ST; Sunyaev & Titarchuk (1980)). In this thermal Comptonization model with a spherical geometry, for energies below $`kT`$ of the scattering medium, the spectrum is a power law with photon index $`\alpha =(1/2)+[(9/4)+\gamma ]^{1/2}`$, where $`\gamma =\pi ^2mc^2/3kT(\tau +2/3)^2`$, and $`\tau `$ is the optical depth.
Figure 3 shows the best PLE and ST model fits to the OSSE data. Both functional forms give equally good fits to the OSSE data alone. See §8 for the effect of doing joint fits with the RXTE spectrum.
The best fit PLE model has $`\alpha =1.15\pm 0.07`$, $`kT=97\pm 6`$ keV, and flux $`0.094\pm 0.001\mathrm{photons}\mathrm{cm}^2\mathrm{s}^1\mathrm{MeV}^1`$ at 100 keV. (The errors are $`1\sigma `$). This has $`\chi _\nu ^2`$= 1.01, $`Q=0.37`$. The flux normalization at 100 keV is a factor of 2.0 lower than that found by Grabelsky et al. (1995) in the 1991 September observation, while our power law index is slightly steeper (theirs was $`0.88\pm 0.05`$), and our $`kT`$ is higher (they had $`68\pm 2`$).
The best fit ST model has $`\tau =2.5\pm 0.1`$, $`kT=47\pm 1`$ keV, and flux $`0.093\pm 0.001\mathrm{photons}\mathrm{cm}^2\mathrm{s}^1\mathrm{MeV}^1`$ at 100 keV. This has $`\chi _\nu ^2`$= 1.00, $`Q=0.45`$. The flux normalization at 100 keV is a factor of 2.0 lower than that found by Grabelsky et al. (1995) in the 1991 September observation, while our $`\tau `$ is smaller (theirs was $`3.0\pm 0.1`$), and our $`kT`$ is higher (they had $`37\pm 1`$). However, the value we derive for $`\alpha =1.9`$ is the same as theirs.
Grabelsky et al. (1995) noted that their ST fit was marginal at best. They found that the model dropped off much more rapidly than the data at higher energies. As we have shown, this disagrees with our results where we get good fits to the OSSE data alone. However, in §8 we will show that this is no longer the case when the OSSE data is combined with the RXTE spectrum.
## 6 RXTE Observations and Reductions
The RXTE pointed observation was made as a Target of Opportunity on 1996 July 26 (MJD 50290), just after the OSSE run ended. We generated the RXTE spectrum using two of its instruments, the Proportional Counter Array (PCA), and the High Energy X-ray Timing Experiment (HEXTE).
The PCA consists of five Xe proportional counters with a total effective area of about 6500 cm<sup>2</sup> (Jahoda et al. (1996)). These cover a scientifically validated energy range of 2 – 60 keV with an energy resolution $`<18`$% at 6 keV.
The HEXTE consists of two independent clusters each containing four phoswich scintillation detectors (Rothschild et al. (1998)). These have a total effective area of about 1600 cm<sup>2</sup>, and cover an energy range of 15 – 250 keV with an energy resolution of 15% at 60 keV. Each cluster can “rock” (beam switch) along mutually orthogonal directions to provide background measurements. In our observation, the background offset was $`1.5\mathrm{°}`$ with a switching time of 16 seconds.
Both instruments have a $`1\mathrm{°}`$ field of view. No other X-ray sources were in the GX 339–4 field of view, or in the HEXTE background regions. For both instruments, the background dominates at higher energies. For GX 339–4 this is at $`70`$ keV for the PCA and $`130`$ keV for HEXTE. The PCA does not make separate background measurements. Instead we used version 1.5 of the background estimator program. The HEXTE background was determined from the off source pointings in both chopping directions. For both instruments, we found that beyond the energies where the background dominated, the model (PCA) and observed (HEXTE) background measurements were the same as what was observed in the on-source observation. This gives us confidence that the background measurements are valid. The backgrounds have been subtracted from all the spectra shown here.
For both instruments, data were only used when the spacecraft was not passing through the South Atlantic Anomaly and when the source was observed at elevations $`>10\mathrm{°}`$ above the Earth’s limb and the pointing was offset $`<0.02\mathrm{°}`$ from the source. The observation was short, with good data starting 18:20:32.625 and ending 20:14:48.497 UT. Because of the above constraints, valid data was only available in four segments: see Smith & Liang (1999) for the light curve. We combined all this data to make one spectrum, although it should be cautioned that the source was extremely variable during this observation (Smith & Liang (1999)). For the PCA, the total on-source exposure time was 3424 sec. After correcting for dead time, this gave an effective exposure of 3337 sec. The dead time correction was more important for the HEXTE data, where the time taken to rock also has to be accounted for in the background exposures. For HEXTE, the adjusted on-source exposure time was 1085 sec for cluster A, and 1070 sec for cluster B.
For the PCA, we used the Standard 2 production data set to generate the spectrum with 129 spectral channels. We used the RXTE tasks in FTOOLS version 4.0 to extract the data. We performed the extraction twice, once using all five Proportional Counter Units (PCU), and once using just the top PCU layer. The top layer is responsible for $`80`$% of the detected counts, so has the best S/N. Using all the layers gives a higher count rate, and thus a better sensitivity at higher energies, but with a lower S/N. The background contributes $`13`$% of the on-source count rate when using all the layers, and $`8`$% using just the top layer. We found that the spectral results made little difference which of these two methods was used, and we chose to retain both of them for our fitting. For both cases, we used version 2.1.2 of the PCA response matrices. We ignored the PCA data below 2 keV that are not scientifically valid. We also ignored the PCA data above 60 keV, well below where the background dominated: we found that making this upper limit to the energy range smaller had no effect on the spectral fitting. We simply dropped the data around 4.78 keV where there is a problem with the response matrix due to the xenon L edge.
For the HEXTE, we used the E\_8us\_256\_DX1F event list data. These record each individual event with $`8\mu `$sec timing and 256 energy channel resolution. As with the PCA, the RXTE tasks in FTOOLS version 4.0 were used to extract the two HEXTE spectra, one for each cluster. For both clusters, we used the 1997 March 20 versions of the HEXTE response matrices. We ignored the HEXTE data above 120 keV, well below where the background dominated. We found that changing this upper limit to the energy range had no effect on the spectral fitting. Given the short exposure time, the HEXTE error bars are large, and the spectra do not extend far enough beyond the break energy to be constraining. We also ignored the HEXTE data below 20 keV. Again, this choice of energy cut-off had no effect on the fitting results.
We used the two PCA and two HEXTE spectra jointly in our spectral fitting. This was performed using XSPEC version 10.0. The relative overall normalization of the PCA and HEXTE spectra was left as a free parameter in the fitting, while all the other model parameters were the same for both instruments. We found that the relative normalization was $`0.71\pm 0.01`$ in all of the results presented here. We did not find it necessary to include separate overall normalizations between the two PCA spectra or between the two HEXTE spectra. The fitting was performed using the complete spectral resolution of both instruments, though the data have been rebinned in the figures so that each bin is at least $`5\sigma `$.
## 7 RXTE Spectra
As in §5, in this paper we only consider two simplistic phenomenological models for the higher energy emission. A detailed fitting of this data using a more self-consistent model is presented elsewhere (Böttcher, Liang, & Smith (1998)).
### 7.1 Power Law Times Exponential Fit
A PLE model fits the RXTE data (alone) above 15 keV. The cut-off energy is very poorly determined by the RXTE data because the HEXTE error bars are large in this short observation, and the RXTE data does not sample far beyond the cut-off energy. We therefore chose to fix $`kT=96.9`$ keV at the value found for the best OSSE fit.
Figure 4 shows the results of fitting the PLE model to the data above 15 keV. This uses $`\alpha =1.26\pm 0.01`$, and flux $`0.130\pm 0.005\mathrm{photons}\mathrm{cm}^2\mathrm{s}^1\mathrm{keV}^1`$ at 1 keV. (The errors are 90% confidence regions for varying one parameter). This has $`\chi _\nu ^2`$= 0.93, $`Q=0.79`$ (for fitting above 15 keV).
It is apparent from Figure 4 that two extra components are required to fit the spectrum at lower energies:
(1) A soft component that peaks at $`2.5`$ keV. Such a soft component has been seen in previous GX 339–4 observations, eg. by Tenma (Makishima et al. (1986)), EXOSAT (Ilovaisky et al. (1986); Méndez & Van der Klis (1997)), and Ginga (Miyamoto et al. (1991); Ueda et al. (1994)). It is present in all states, and dominates during the soft state.
(2) A broad emission centered on $`6.4`$ keV. A similar feature was seen in Ginga observations of GX 339–4 in the low (=hard) state, and was modeled using a reflection model (Ueda et al. (1994)). Using a broad iron line improved the fits in the EXOSAT observations of GX 339–4 made in the high (=soft) state (Ilovaisky et al. (1986)). A broad line was seen in RXTE-ASCA observations of Cygnus X-1 in the soft (=high) state (Cui et al. (1998)). A very similar effect was seen in RXTE observations of Cygnus X-1 in the low (=hard) state by Dove et al. (1998), though they chose to force the Fe K$`\alpha `$ line to be narrow (with a line width of 100 eV giving an equivalent width of 60 eV) which resulted in obvious residuals. RXTE observations of the Seyfert galaxy MCG –5-23-16 also found this feature, confirming previous ASCA observations (Weaver, Krolik, & Pier (1998)): they interpreted it as being a broad iron line.
As we show in the next sub-section, the exact form of the soft components is poorly determined by the RXTE data. Although we showed a data point below 2 keV in Figure 4, the PCA response matrix is currently not reliable below 2 keV, and this data point was dropped from our fitting. We therefore do not get a good measurement of the roll-over, and hence cannot determine $`N_H`$ reliably. Thus for all our spectra it has been fixed at $`N_H=5\times 10^{21}\mathrm{cm}^2`$, as was found in the EXOSAT observations that more reliably measured the lower energy spectrum (Ilovaisky et al. (1986); Méndez & Van der Klis (1997)).
An extra broad emission feature around $`6.4`$ keV is always required to get a good fit. Here we assume it is an iron line feature, that may have been broadened by orbital Doppler motions and/or scattering off a hot medium. Detailed Compton scattering line profiles have been generated using our Monte Carlo codes and are presented elsewhere (Böttcher, Liang, & Smith (1998)). Here we simply use a broad Gaussian and an iron edge. We caution that it is possible that the broad RXTE feature seen in the sources listed above is due to a currently unidentified systematic flaw in the PCA response matrices, although the observations of MCG –5-23-16 make this unlikely.
Unlike for the Ginga–OSSE observation of GX 339–4 in 1991 (Zdziarski et al. (1998)) our simplistic continuum fitting of the 1996 data does not require a significant reflection component. Dove et al. (1998) found that no reflection component was needed to explain their RXTE observations of Cygnus X-1. They noted that the broad spectral range covered by RXTE is an important improvement over previous observations for accurately modeling the continuum. However, we caution that we are only using simple phenomenological models in this paper; for a more self-consistent modeling of this data, see Böttcher et al. (1998), which also concludes that there is no need for a strong reflection component because most of the incident flux from the corona goes into heating of the disk surface layer and is not reflected.
Figure 5 shows an example of a complete fit to the RXTE data where we have used a simple power law for the soft component. The PLE component has $`\alpha =1.22\pm 0.01`$, consistent with the OSSE best fit value. The iron emission line is very broad. Its equivalent width (EW) is 640 eV.
### 7.2 Sunyaev-Titarchuk Function Fit
A ST model can also be used to give a good fit to the RXTE data alone. (In the next section we show that the ST model does not give a good fit to the combined RXTE-OSSE data.)
Figures 6–8 show three sample fits. We show three different models for the soft component (power law, black body, and thermal bremsstrahlung) to illustrate how this affects the relative contributions of the soft components. In all cases, a broad iron emission line is required. Its EW = 475, 700, and 570 eV in Figures 6–8 respectively. Note that in the unfolded spectra shown, we have not normalized the PCA and HEXTE spectra: this is discussed in the next section.
It is important to note that the spectra are very hard. The ST component dominates the entire RXTE spectrum, even down to 2 keV. (The PLE and soft power law components cross at 4 keV in Figure 5.) This is particularly important in our variability study using this data (Smith & Liang (1999)) where we divide the PCA data into three bands 2–5, 5–10, and 10–40 keV to try to highlight the separate components. There we show that it is the soft 2–5 keV band that is the most variable.
## 8 Joint RXTE-OSSE Spectra
We now combine the RXTE and OSSE data to generate the joint spectrum. The results presented in this section should be treated with care, because the two data sets are not quite simultaneous. The fact that the OSSE hardness ratio did not change significantly over our two week observation makes it reasonable to assume that the average gamma-ray spectral shape was the same at the time of the RXTE observation, but this remains an assumption.
In our rapid variability study of this RXTE data we show that the hardness ratios do not change dramatically with time or brightness (Smith & Liang (1999)). Thus while the source is extremely variable, with the 2–5 keV band showing the greatest variability, the average spectra are representative of the spectral shape throughout the RXTE observation. However, it should be remembered that the spectrum presented here is a representative average, and the actual spectrum at any given instant will have an overall normalization that can differ by a factor $`3`$.
### 8.1 Power Law Times Exponential Fit
The PLE best fits to the separate OSSE (Figure 3(a)) and RXTE (Figure 5) data used the same $`kT=96.9`$ keV and had consistent values of $`\alpha `$. We therefore will get a good joint RXTE–OSSE fit provided the relative normalizations between the different instruments can be matched. This depends on the absolute calibrations of the OSSE, PCA, and HEXTE instruments. We could simply shift the unfolded OSSE, PCA, and HEXTE spectra arbitrarily to make them match. But instead we have used the RXTE ASM and BATSE data to try to be more rigorous.
The uncorrected PCA data gives a 2–10 keV flux of 77 mCrab. However, the ASM gave a 2–10 keV flux of only 60 mCrab for this day. This suggests that the unfolded PCA spectrum needs to be normalized by $`\times 0.78`$ to get the correct value.
The uncorrected HEXTE data gives a 20–100 keV flux of $`3.6\times 10^2\mathrm{photons}\mathrm{cm}^2\mathrm{s}^1`$. This is close to the BATSE measured flux of $`4.0\times 10^2\mathrm{photons}\mathrm{cm}^2\mathrm{s}^1`$. This suggests that the unfolded HEXTE spectrum needs to be normalized by $`\times 1.1`$ to get the correct value.
We therefore infer that the relative normalization of the HEXTE and PCA spectra should be 0.71, which is exactly what was found independently in the joint fitting.
Based on the 50–70 keV OSSE daily light curve in Figure 2, we would expect to have to multiply our average OSSE spectrum by a factor of $`(110/94)=1.17`$ to get the correct normalization on 1996 July 26. This agrees with doing a joint RXTE-OSSE fit in XSPEC, where we get a relative normalization between OSSE and the PCA of 1.5 (which is 1.17/0.78).
Figure 9 shows the effect of shifting the unfolded spectra from Figures 3(a) and 5 by these amounts. As expected, the data sets join smoothly. The model curve is the same as in Figure 5, but with an overall normalizing factor of $`\times 0.78`$ that from before (the RXTE fits assumed the PCA was perfectly calibrated).
Figure 9 highlights that it is extremely important that future observations of GX 339–4 should make simultaneous X-ray and gamma-ray observations to accurately measure the spectrum both above and below the cut-off energy.
### 8.2 Sunyaev-Titarchuk Function Fit
The ST best fits to the separate OSSE (Figure 3(b)) and RXTE (Figures 6–8) data do not give consistent ST fit parameters. We therefore do not expect to get a good joint RXTE–OSSE fit.
Figure 10 shows the results of adding the OSSE data to Figure 6. We have fixed all the model parameters to those given in Figure 6, except for the relative normalization of the OSSE and RXTE instruments. It is clear that the model fit to the RXTE data does not explain the shape of the joint RXTE-OSSE spectrum: the model drops off too rapidly with increasing energies to give an acceptable joint fit. This now agrees with Grabelsky et al. (1995).
In §5 we showed that an ST model could fit the OSSE data alone. However, we find that these model parameters do not give a good fit to the RXTE data. Even if we ignore all the soft components and just try to use a ST model alone to simultaneously fit the OSSE data and the RXTE data above 15 keV, we are unable to get an acceptable fit (the best $`\chi _\nu ^2`$= 1.5, $`Q=4\times 10^9`$, for $`kT=35`$ keV, and $`\tau =3.8`$).
It is clear that the correct shape cannot be generated by the ST model when a larger energy range is available. We cannot rule out that the gamma-ray spectrum changed dramatically between the end of our OSSE observation and our RXTE observation, though the prior evolution of the source makes this unlikely. Again this highlights the need for future observations of GX 339–4 to make simultaneous X-ray and gamma-ray observations.
More realistic and self-consistent Compton scattering models can explain the harder spectrum presented here. A full study is beyond the scope of this paper, but in Böttcher, Liang, & Smith (1998) we develop such a detailed simulation, and illustrate that it can fit the joint GX 339–4 spectrum.
## 9 Summary
As part of our multiwavelength campaign of observations of GX 339–4 in 1996 we presented our radio, X-ray, and gamma-ray observations made in July, when the source was in a hard state (= soft X-ray low state).
The radio observations were made at the time when there was a possible radio jet. We showed that the radio spectrum was flat and significantly variable, and that the radio spectral shape and amplitude were not anomalous for the source at this time. Daily light curves from our pointed OSSE observation July 9–23, from BATSE, and from the ASM on RXTE also showed that there was no significant change in the X- and gamma-ray flux or hardness during the time the radio jet-like feature was seen.
The higher energy portion of our pointed RXTE observation made July 26 is equally well fit using simple PLE and ST functions. An additional soft component is required, as well as a broad emission feature centered on $`6.4`$ keV. This may be an iron line that is broadened by orbital Doppler motions and/or scattering off a hot medium. Its equivalent width is $`600`$ eV. Both a PLE and a ST model also fit our OSSE spectrum on its own. Although the observations are not quite simultaneous, combining the RXTE and CGRO spectra we find that the PLE model easily fits the joint spectrum. However, the ST model drops off too rapidly with increasing energies to give an acceptable joint fit.
Our results show that it is extremely important that future studies of GX 339–4 should make truly simultaneous multiwavelength observations, particularly given the variability of the source. It is essential to accurately measure the spectrum both above and below the gamma-ray cut-off energy. It will also be extremely interesting to combine future unusual radio activity with the behavior at high energies.
We thank the referee for carefully reading the manuscript and providing useful suggestions and clarifications. This work was supported by NASA grants NAG 5-1547 and 5-3824 at Rice University. This work made use of the RXTE ASM data products provided by the ASM/RXTE teams at MIT and at the RXTE SOF and GOF at NASA’s Goddard Space Flight Center. The BATSE daily light curves were provided by the Compton Observatory Science Support Center at NASA’s Goddard Space Flight Center.
|
no-problem/9812/physics9812034.html
|
ar5iv
|
text
|
# Studies of 100 𝝁m-thick silicon strip detector with analog VLSI readout
## 1 Introduction
A silicon strip detector (SSD) has the highest position resolution among the electric tracking devices in particle physics experiments. However, an error in measuring the track angle is dominated by the multiple scattering effect for particles with a low velocity. If the effect is reduced with a very thin SSD, new experiments which are impossible by the present technology will be realized. One example is a search for the $`T`$ violation in the decay of B mesons , in which the $`T`$-violating transverse $`\tau ^+`$ polarization in the decay $`\mathrm{B}\mathrm{D}\tau ^+\nu `$ will be measured to a precision of 10<sup>-2</sup>. In order to obtain the $`\tau `$ polarization the decay vertices of B and $`\tau `$ must be measured separately. A simulation shows that the experiment will be feasible only with very thin SSD’s at asymmetric-energy B factories.
In general, a thin SSD has a small signal-to-noise (S/N) ratio because the energy deposit in the detector is proportional to the thickness and its large capacitance results in a large noise. Thus careful treatment of a noise in the off-line analysis is important.
In this paper, we evaluate the performances of a 100 $`\mu `$m-thick silicon strip detector. The performances are compared with those of the 300 $`\mu `$m and 500 $`\mu `$m-thick silicon strip detectors.
## 2 Detector
Single-sided silicon detectors with the dimensions of 1 cm $`\times `$ 1.3 cm have been fabricated by Hamamatsu Photonics. The strip pitch is 100 $`\mu `$m. The widths of implantation strips and aluminum electrodes are 42 $`\mu `$m and 34 $`\mu `$m, respectively. Three detectors with different thicknesses (100 $`\mu `$m, 300 $`\mu `$m, and 500 $`\mu `$m) were tested. The 100 $`\mu `$m-thick SSD was made by etching a 300 $`\mu `$m-thick wafer. The analog VLSI chips (VA2<sup>1</sup><sup>1</sup>1Produced by Integrated Detector and Electronics AS (IDEAS), Oslo, Norway. are used as a readout circuit of the detectors. An SSD and a VLSI chip were mounted on a printed circuit board called “hybrid board”.
## 3 Experiment
Two different particles were used for evaluation of the detector performances. A proton beam was used to measure the response of detectors for baryons or heavy particles. To see the response for light and high velocity particles which satisfy the minimum ionizing particle (MIP) condition ($`E/m>3`$), electrons from a $`{}_{}{}^{S}90`$ $`\beta `$-ray source was used.
The experiment was carried out with a proton beam at the Research Center for Nuclear Physics, Osaka University. Scattered protons from a $`{}_{}{}^{C}12`$ target were momentum analyzed by a magnetic spectrometer. A detector system that consists of an SSD and two trigger plastic scintillation counters was placed at the focal plane of the spectrometer. The momentum of detected protons was 800 MeV/$`c`$ with the momentum spread of $`<0.05\%`$. The energy loss for a proton with 800 MeV/$`c`$ is 68 keV for the 100 $`\mu `$m-thick SSD, which is about 1.7 times larger than that for the minimum ionizing protons.
The readout system is schematically shown in Fig. 1.
The hybrid board consisting of a silicon strip and a VA2 chip was connected to a “repeater card”, which contained level converters for logic signals, buffer amplifier for analog output signal, and adjustable bias supply for the VA2 chip. The VA2 chip was controlled by a VME based timing board which received a trigger signal and generated clock pulses for VA2 and a VME based flash ADC board. Analog multiplexed output from VA2 was sent to a flash ADC through the repeater card. Two layers of trigger counters were placed in front of the SSD. The repeater card was connected to the hybrid board with a ribbon cable for both the analog and logic signals. The length of the ribbon cable was about 15 cm.
In order to compare the characteristics of silicon strip detectors, the operation parameters of the VA2 readout chips were fixed to standard values without optimization for each measurement. Signal shaping time was about 700 ns. Signals were read out in 4 $`\mu `$sec clock repetition. Typical trigger rate was about 30 Hz.
In addition to a proton beam test, measurements with a $`{}_{}{}^{S}90`$ $`\beta `$-ray source were also performed. The $`{}_{}{}^{S}90`$ $`\beta `$-ray source was placed at 15 mm from the SSD. A collimator with a size of 2 mm in diameter and 10 mm in thickness was used to irradiate electrons perpendicularly to the SSD. In order to realize the minimum ionizing condition, a high energy component of $`\beta `$-rays was selected by a trigger scintillation counter placed behind the SSD. Operation parameters of the VA2 chip was the same as those at the proton beam test. Readout clock was 400 ns. The trigger rate at the $`\beta `$-ray source test was about 7 Hz.
## 4 Analysis and Results
An output from each strip has a different offset level. These differences have been trimmed at the first step of the off-line analysis. Solid lines in Fig. 2 show the maximum pulse height distributions after the pedestal trimming for 100 $`\mu `$m, 300 $`\mu `$m, and 500 $`\mu `$m-thick SSD’s at the proton beam test.
Note that we have neglected the effect of charge division among adjacent strip. Dotted lines show the same distributions under the condition that no charged particle hit the detector. The noise peak and proton signal peak have overlapped for the 100 $`\mu `$m-thick SSD, while the proton signals are clearly distinguished from noises for the 500 $`\mu `$m and 300 $`\mu `$m-thick SSD’s.
For 100 $`\mu `$m-thick SSD, a strong noise level correlation between non-adjacent channels has been observed. This indicates that the main component of the noise has a common phase and amplitude among the strips. This component called common-mode noise (CMN) has been calculated as an averaged pulse height over all strips. In the calculation, channels with significantly large pulse heights; larger than 3 standard deviation $`(\sigma )`$ of the noise distribution, have been excluded. Fig. 3 shows the maximum pulse height distribution after the CMN subtraction for the 100 $`\mu `$m SSD.
Proton events are clearly separated from the noise.
We have investigated the characteristics of noise more carefully. Fig. 4(a) shows the strip dependence of the noise width after the CMN subtraction. The width depends on the strip number, whereas pulse height differences between adjacent two strips shown in Fig. 4(b) have a constant value of about 6.
This indicates that the intrinsic $`\sigma `$ of the noise is expected to be about 4.2 ($`=6/\sqrt{2}`$) for all strips. Thus, we conclude that the CMN has a non-constant component. Instead of simply averaging the pulse heights, we fit them with a linear curve to get CMN as a function of a channel number. Fig. 4(c) shows the noise widths after this method is applied. The widths are about 4.2 for all strips as expected.
If the CMN is not removed correctly by assuming a constant CMN, a noise width depends on a strip number (Fig. 4(a)). This may cause a strip dependent S/N separation which are not desirable for any experiments. Fitting the CMN with a linear curve is particularly important for the detection of MIP’s with a thin SSD where the S/N ratio is small. The maximum pulse height distribution for electrons with 100 $`\mu `$m-thick SSD after subtracting the CMN by linear-fitting is shown in Fig. 5(b) compared with that with a constant CMN subtraction (Fig. 5(a)). Although electron events are not separated from the noise for both cases, the separation of signals from noises is improved by the linear-fitting method<sup>2</sup><sup>2</sup>2$`\beta `$-rays were irradiated at the central strips by using a collimator. It is expected that this improvement is clearly seen for the strips near the edge of the detector..
Fig. 5 indicates that there is a finite probability of misidentifying a noise as a particle track by selecting the maximum pulse height. The detection efficiency and signal misidentification probability for electrons with 100 $`\mu `$m-thick SSD are plotted as a function of threshold energies in Fig. 6.
When a threshold level is set to detect the electron with an efficiency more than 99% the probability of misidentification obtained by linear-fitting of CMN is 27% smaller compared to that by the constant CMN-subtraction method.
The S/N ratio obtained from $`\beta `$-ray source tests for the 100 $`\mu `$m and 300 $`\mu `$m SSD’s are summarized in Table 1.
Better S/N ratio is obtained by fitting CMN with a linear curve. The S/N ratios obtained with the assumption of a constant CMN for both the 100 $`\mu `$m and 300 $`\mu `$m SSD’s are slightly worse. For a 300 $`\mu `$m SSD, the difference of two methods in subtracting the CMN is not very important in an actual application because the S/N ratio is sufficiently large. Noise width obtained in the $`{}_{}{}^{S}90`$ $`\beta `$-ray source test and the proton beam test are summarized in Table 2 in energy unit (keV).
The width of CMN at the $`\beta `$-ray source test is different from that at the proton beam test. But the noise after the CMN subtraction is almost the same.
There remains a possibility to improve the S/N ratio by considering the charge division among adjacent strips during finding a particle trajectory. Performances of the prototype detector might be improved by optimizing its operating conditions.
## 5 Conclusion
An SSD with a thickness of 100 $`\mu `$m was tested with 800 MeV/$`c`$ protons and $`\beta `$-rays from $`{}_{}{}^{S}90`$ source. By using an analog VLSI chip for readout, we remove the CMN. Assuming that CMN is constant among all strips, proton signals are separated from noises for the 100 $`\mu `$m, 300 $`\mu `$m and 500 $`\mu `$m-thick SSD’s after the CMN subtraction. We found that a non-constant component in CMN makes the energy resolution worse. For the 100 $`\mu `$m SSD, the signal and noise separation was improved by fitting CMN with a linear curve at a $`\beta `$-ray source test. We conclude that a 100 $`\mu `$m SSD with analog VLSI readout can be used as a very thin tracking device in a future experiment. However, careful treatment of CMN is important for detecting MIP’s.
This work has been supported by the Grant-in-Aid for General Science Research (No. 07454057 and No. 09640358) by the Ministry of Education, Science and Culture.
|
no-problem/9812/astro-ph9812040.html
|
ar5iv
|
text
|
# On the origin of Damped Lyman–𝛼 systems: a case for LSB galaxies?
## 1 Introduction
Since bright spiral disk galaxies contain most of the H I mass in the nearby universe (Rao & Briggs (1993)), it has always seemed likely that these types of galaxies — or at least their progenitors — would be the origin of the high-redshift DLy$`\alpha `$ systems seen in background QSO spectra. The detection of $`z<\mathrm{\hspace{0.25em}1}`$ DLy$`\alpha `$ systems (Boisse et al. (1998); Turnshek (1997)) has now made it possible to search for the galaxies which cause the absorption. Although redshifts are largely unavailable for the objects imaged, recent observations suggest that a wide range of morphological types are responsible for the absorption (Lanzetta et al. (1997); Le Brun et al. (1997)), and that bright spiral galaxies do not dominate the DLy$`\alpha `$ population. In particular, several low surface brightness (LSB) galaxies have been found close to QSO sightlines (Steidel et al. (1994); Le Brun et al. (1997)), while in one case, only an LSB or dwarf galaxy can account for the complete non-detection of a suitable galaxy responsible for a $`z=\mathrm{\hspace{0.25em}0.0912}`$ DLy$`\alpha `$ line (Rao & Turnshek (1998)).
Further evidence that DLy$`\alpha `$ systems may not originate in bright spiral galaxies, even at higher redshift, comes from measurements of the metallicity of the absorbing gas. The abundances of alpha elements in DLy$`\alpha `$ systems do not appear to be enhanced relative to iron group elements but similar to the solar value (Molaro, Centurion & Vladillo 1998; Centurion et al. 1998), while the overall metallicity seems to remain low even at $`z<1.5`$ (Pettini et al. 1998). These results suggest that DLy$`\alpha `$ lines may actually arise in systems which are less common than the main star-forming disks we see around us today. In this Letter we use our model of disk formation and evolution to explore differences in metallicity, dust content, and neutral-gas mass density between high surface brightness (HSB) and LSB galaxies. We find that we are able to better reproduce the observational data with models of LSB galaxies, and predict how future abundance measurements may be able to further differentiate between the two.
## 2 Disk Models
Several authors have shown how differences in surface brightness between galaxies can be readily understood if LSB galaxies are hosted in dark halos with values of the spin parameter, $`\lambda `$, of the dark halo ($`\lambda =J|E|^{1/2}G^1M^{5/2}`$, where $`J`$ is the angular momentum, $`|E|`$ is the total energy and $`M`$ is the total mass of the halo) larger than those of HSB galaxies (Fall & Efstathiou (1980); Kashlinsky (1982); Dalcanton, Spergel & Summers (1997); Mo, Mao & White (1997); Jimenez et al. (1997)). Jimenez et al. (1998) used a detailed chemo-spectro-photometric disk model to show (abandoning the assumption of a constant $`M/L`$ ratio) that not only surface brightness, but also color, color gradients and metallicity of LSB disks can be explained if the spin parameter is larger for LSB than for HSB galaxies.
In order to investigate the properties of both HSB and LSB disks, and to determine whether their characteristics are akin to those identified in DLy$`\alpha `$ systems, we have computed models of galactic disks with $`\lambda =\mathrm{\hspace{0.25em}0.03}`$ (HSBs) and $`\lambda =\mathrm{\hspace{0.25em}0.06}`$ (LSBs; see Jimenez et al. (1998)). The model is described in detail in Jimenez et al. (1997) and we only summarize it here. We assume that the specific angular momentum of baryonic matter is the same as the dark matter and that gas settles into a given dark halo until centrifugally supported . We also assume that the dark matter profile is that found in numerical simulations (Navarro, Frenk & White (1997)), although using the isothermal sphere profile does not significantly change the computed initial surface density for the settling disk. Once the initial surface density of the disks has been computed, we used the Schmidt law (Kennicutt (1998)) to compute the star formation rate. In particular, the SFR law adopted is $`\psi (r,t)=\nu \mathrm{\Sigma }_{gas}(r,t)^{k_1}\mathrm{\Sigma }(r,t)^{k_2}`$, with $`k_1=1.5`$ and $`k_2=0.5`$, and where $`\mathrm{\Sigma }_{gas}(r,t)`$ is the gas surface density and $`\mathrm{\Sigma }(r,t)`$ is the total mass surface density. The HSB disks have a initial burst of star formation which then declines with time, while the LSB galaxies have constant, but less active, star-formation. The primordial gas infall rate in the disk is assumed to be the same function of total surface density as used in models of the Milky Way (see Jimenez et al. (1998)). The infall rate is higher in the center than in the outermost regions of the disk, with the infall law expressed as $`\dot{\mathrm{\Sigma }}_{inf}(r,t)=A(r)X_{inf}^ie^{t/\tau (r)}`$, where $`\tau (r)`$ is the time-scale for the formation of the disk at a radius $`r`$. The values of $`\tau (r)`$ are chosen to fit the present time radial distribution of the gas surface density in the disk. In analogy with what is required for the disk of the Milky Way, we assumed $`\tau (r)`$ is increasing towards larger radii (see Table 2 of Jimenez et al. (1998)). $`X_{inf}^i`$ is the abundance of element $`i`$ in the infalling gas and the chemical composition is assumed to be primordial. The parameter $`A(r)`$ is obtained by requiring the surface density now to be $`\mathrm{\Sigma }(r,t_{now})`$, so that $`A(r)=\frac{\mathrm{\Sigma }(r,t_{now})}{\tau (1e^{t_{now}/\tau (r)})}`$. The evolution of several chemical species (H, D, He, C, N, O, Ne, Mg, Si, S, Ca, Fe and Zn), as well as the global metallicity $`Z`$, is followed by taking into account detailed nucleosynthesis prescriptions (Jimenez et al. 1998). The IMF is taken from Scalo (1986). For simplicity, we have adopted an Einstein-deSitter Universe ($`\mathrm{\Omega }=1`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0.0`$, $`H_0=50`$ km s<sup>-1</sup> Mpc<sup>-1</sup>). Non-zero values of $`\mathrm{\Lambda }`$ actually strengthen the results discussed below.
## 3 Results
The initial goal of our modelling was to explore any differences in gas metallicity between HSB and LSB disks, to determine whether the measurements of \[Zn/H\] now available for DLy$`\alpha `$ systems could discriminate between the two types of galaxies. In Figure 1 we plot the values of \[Zn/H\] found by Pettini et al. (1997 and refs. therein; 1998) against the redshift of the absorption systems. Our models enable us to plot the metallicity of gas disks as a function of redshift for both a given galactic radius, $`r`$, and for a formation redshift. Hence, in Figure 1a, we plot the evolution in the metallicity of an LSB disk, forming at $`z=4`$, at $`r=0`$, 1.0, 2.5, and $`610`$ kpc. We take $`610`$ kpc to be the outer limit at which an LSB would still give rise to a DLy$`\alpha `$ line; our models predict $`N`$(H I)$`>\mathrm{\hspace{0.25em}5}\times 10^{20}`$ cm<sup>-2</sup> at this radius for an LSB baryonic mass of $`>\mathrm{\hspace{0.25em}5}\times 10^9M_{}`$. This agrees well with observed H I surface densities toward nearby LSB galaxies at similar radii (de Blok, McGaugh, & van der Hulst 1996).
In Figure 1b, we show the same variation in metallicity for $`r`$ 6 kpc but with the galaxy forming at $`z=`$3.2, 2.2 and 1.4. Figure 1a clearly shows that the predicted metallicities of LSB galaxies fit the observed data very well, particularly at outer radii, where the absorption cross-section will be largest and hence where the majority of QSO sightlines will intercept. Figure 1b demonstrates, however, that such LSB disks must form at $`z\mathrm{\hspace{0.25em}4}`$, and evolve slowly for such a model to be consistent with the data.
In contrast, Figure 1c shows the variation of metallicity with redshift for an HSB galaxy, again forming at $`z=\mathrm{\hspace{0.25em}4.0}`$, for $`r=`$0, 4.0, 6.0 and $`1020`$ kpc. The figure shows that HSB galaxies have enriched their ISMs much faster than LSB galaxies, leading to gas metallicities higher than observed. However, this does not rule out HSB disks as potential absorbers. Figure 1d shows that if HSB galaxy halos form over the range of redshifts measured for DLy$`\alpha `$ systems, as is indeed the case in CDM cosmogonies (e.g., Peacock & Heavens (1985)), then it is possible to reproduce the \[Zn/H\] measurements. In this case, most of the DLy$`\alpha `$ systems would be proto-galaxies in the very early stages of their evolution—it takes only 0.3 Gyr for HSB disks to reach \[Zn/H\] $``$ $`1`$, close to the average measured value. We note that in our models, however, these proto-galaxies have not yet formed rotating disks, hence we would not expect the velocity profiles of metals absorption lines in DLy$`\alpha `$ systems to be indicative of rotating disks. The fact that the profiles of metal lines do appear to support rotating thick-disk models (Wolfe & Prochaska (1998); Prochaska & Wolfe (1997)) tends to favor LSB disks. We note, however, that it remains possible that the initial merging of galaxy halos in a CDM-like scenario could still give rise to the observed line profiles (Haehnelt, Steinmetz, & Rauch 1998).
One possible way to distinguish between slowly evolving LSB disks and a continuously forming population of HSB galaxies is to consider the evolution of the dust content of each galaxy type. Figure 2 shows the dust formation rate for HSB and LSB disks at different radii (each radius normalized to $`10^9`$ M). The amount of dust produced by each generation of stars was accounted for by using the model of Draine & Lee (1984). As expected, HSB disks form about twice as much dust as their LSB counterparts due to their higher metallicity. More importantly, however, HSB galaxies produce most of their dust during the first Gyr, reaching a maximum at an age of 0.3 Gyr. On the other hand, LSB galaxies produce virtually no dust at radii larger than $`\mathrm{\hspace{0.25em}2}`$ kpc. Therefore, a population of continually forming proto-disks should quickly begin to obscure background QSOs, removing HSB disks from DLy$`\alpha `$ samples, as suggested by other researchers (Fall & Pei (1993)). LSB galaxies, of course, remain largely unobscured, and would show up readily against background QSOs. If, in reality, it is a mixture of LSB and HSB galaxies which actually give rise to DLy$`\alpha `$ systems, such an effect would at least explain why LSB galaxies have so readily been found responsible for low redshift DLy$`\alpha `$ absorbers.
Our model predicts one more difference between the observable properties of DLy$`\alpha `$ systems depending on whether HSB or LSB disks are responsible. DLy$`\alpha `$ systems have been used to infer the evolution in the cosmological mass density of neutral gas in the universe, $`\mathrm{\Omega }_g`$, from $`z\mathrm{\hspace{0.25em}0}4`$. In Figure 3 we reproduce the values of $`\mathrm{\Omega }_g`$ as a function of redshift (Storrie-Lombardi, McMahon & Irwin (1996)). We also show how H I gas is consumed by stars in our disk models, both for HSB (dashed line) and LSB (continuous line) galaxies. The curves are normalized such that that $`\mathrm{\Omega }_g`$ for LSB galaxies is zero at $`z=0`$. The true value of $`\mathrm{\Omega }_g`$ for LSB galaxies clearly lies somewhere between 0 (an extreme limit which would assume that LSB disks have used up all their gas — at odds with the observations) and that observed for galaxies today, a value dominated by the contribution from spirals and irregular galaxies (Rao & Briggs 1993). As can be seen in Figure 3, however, this range is extremely small, and it matters little whether we set $`\mathrm{\Omega }_g`$ for LSB disks to zero or the value observed for all galaxy types. Surprisingly, the HSB galaxies do not reproduce the data particularly well—the amount of gas needed at high-redshift to reproduce the neutral gas density measured in the local universe is much higher than observed. On the other hand, LSB disks fit the data much better, since they have less efficient star formation and thus transform less H I into stars. It would seem, therefore, that DLy$`\alpha `$ systems are very inefficient star producers.
The figure is also important in demonstrating that there need not be many LSB galaxies in the nearby universe to account for the number of DLy$`\alpha `$ systems found at high redshift. Although plausible arguments have been made that the size and number density of LSB galaxies in the local universe are sufficient to explain the frequency of DLy$`\alpha `$ systems at high redshift (Impey & Bothun (1997)), our models show that even if no LSB disks remained today, their gradual slow consumption of H I over time best fits the observed values of $`\mathrm{\Omega }_g`$, and that HSB disks simply convert H I into stars too efficiently to account for those measurements.
## 4 Future work
Since the star formation histories in LSB and HSB galaxies are so different, our models make it possible to predict differences in the $`\alpha `$nucleid to iron-peak element ratios for the two types of galaxies. In particular, we would expect that at a fixed epoch HSB disks have \[$`\alpha `$/Fe\] ratios around zero or less and lower than \[$`\alpha `$/Fe\] ratios in LSB disks. This because at the same epoch LSB disks have lower Fe abundances than $`\alpha `$-element abundances relative to HSB disks. This difference is mostly due to the fact that $`\alpha `$-elements are produced on very short timescales by Type II SNe whereas Fe is produced on long timescales (from several tenths of million years to several Gyr) by Type Ia SNe. Therefore, the different star formation history affects iron more than $`\alpha `$-elements. This is contrary to what happens if one looks at the \[$`\alpha `$/Fe\] vs. \[Fe/H\] relation at a fixed \[Fe/H\] instead of at a fixed epoch. In this case, galaxies with lower star formation rate shows lower \[$`\alpha `$/Fe\] ratios (Matteucci (1991)) than systems with higher star formation rates.
This effect is clearly seen in Figure 4 where we plot the redshift evolution of \[$`\alpha `$/Zn\] for LSB (continuous line) and HSB (dashed line) galaxies formed at $`z=\mathrm{\hspace{0.25em}1.8}`$ and 4.0, which encompasses a reasonable range in redshift for disk formation. As expected, the LSB disks have super-solar \[$`\alpha `$/Zn\] ratios during most of their evolution. The HSB galaxies exhibit super-solar ratios only during their first Gyr, but have mainly sub-solar ratios for the rest of their lives. It can be seen that any measured \[$`\alpha `$/Zn\] ratio at a given redshift is unique in determining whether the absorption originates in an LSB or HSB disk — although we only plot the curves after the first 0.4 Gyr of the disks’ formation, before which the \[$`\alpha `$/Zn\] ratio move vertically to higher values. Unfortunately, only two \[$`\alpha `$/Zn\] ratios have been measured in DLy$`\alpha `$ systems (for $`z_{\mathrm{abs}}<z_{\mathrm{em}}`$ systems), both of which are from observations of S II lines (see, e.g., Centurion et al. (1998) and refs. therein), plotted as diamonds in Figure 4. Both measurements have large errors, and it is clear that many, more accurate observations are needed in the future to determine the nature of the DLy$`\alpha `$ population in the Universe.
Figure Captions:
Figure 1: Redshift evolution of \[Zn/H\] for HSB (bottom panels) and LSB (upper panels) galaxies, as measured at different radii in a disk. a) The evolution in metallicity of LSB galaxies formed at $`z=\mathrm{\hspace{0.25em}4}`$ well fits the values measured by Pettini et al. (1997; 1998). Also shown as a dashed box at $`z=\mathrm{\hspace{0.25em}0}`$ is the range in metallicity of our own Galaxy (Roth & Blades (1995)); b) LSB disks that formed late, however, do not fit the data. c) Conversely, HSB disks that formed at $`z=\mathrm{\hspace{0.25em}4}`$ become metal-rich too quickly to explain the observed values; d) only if there is a continuously forming population of HSB disks between $`z\mathrm{\hspace{0.25em}1}4`$ can they account for the measured metallicities.
Figure 2: The dust formation rates for LSB and HSB galaxies, for several galactic radii. The dust production in LSB disks at $`r>2`$ kpc is negligible, while HSB disks produce a significant amount of dust at all radii in the early stages of their evolution. Such dust production may well obscure background QSOs and remove evolved HSB galaxies from DLy$`\alpha `$ samples.
Figure 3: The evolution of neutral H I with redshift (Storrie-Lombardi, McMahon & Irwin (1996)) derived from DLy$`\alpha `$ systems (squares). Also plotted are the predictions from our model for HSB (dashed line) and LSB (continuous line) galaxies, both normalized to the present day neutral H I abundances (Rao & Briggs (1993)). Both curves have been computed using an average of the radii larger than 6 kpc (HSB disks) and 2.5 kpc (LSB disks). We find that HSB galaxies require the existence of much more neutral gas at high redshift than is observed, because of their high star formation rates. LSB galaxies, however, provide an excellent fit to the data because their gas consumption is lower. This supports the idea that LSB galaxies may be responsible for most of the observed DLy$`\alpha `$ lines even at high redshift.
Figure 4: Predictions from our models for \[$`\alpha `$/Zn\] ratios in DLy$`\alpha `$ systems as a function of redshift. The diamonds correspond to measurements of \[S/Zn\] from the literature. LSB galaxies have super-solar values of \[$`\alpha `$/Zn\] since their low star formation rate means fewer Type Ia SNae produce smaller yields of Zn, even though the initial burst of star formation produced the same amount of $`\alpha `$ elements from Type II SNae.
|
no-problem/9812/astro-ph9812415.html
|
ar5iv
|
text
|
# The differential magnification of high-redshift ultraluminous infrared galaxies
## 1 Introduction
At wavelengths between about 10 and 1000 $`\mu `$m the spectral energy distribution (SED) of a galaxy is dominated by the thermal emission from interstellar dust, which typically peaks at a wavelength of about 100 $`\mu `$m (Sanders & Mirabel 1996). The dust is heated by absorbing the blue and ultraviolet light from young stars and active galactic nuclei (AGN). 100-$`\mu `$m radiation from galaxies at redshifts less than about unity has been detected directly by the IRAS and ISO space-borne telescopes. The redshifted dust emission from more distant galaxies can be detected very efficiently at longer submillimetre wavelengths (Blain & Longair 1993a), as demonstrated by the discovery of SMM J02399$``$0136 at redshift $`z=2.8`$ (Ivison et al. 1998; Frayer et al. 1998) in a 850-$`\mu `$m survey (Smail, Ivison & Blain 1997). Distant galaxies were also detected by IRAS, but only those with flux densities enhanced by gravitational lensing galaxies. Three such galaxies are known – IRAS F10214+4724 at $`z=2.3`$ (Rowan-Robinson et al. 1991), H1413+117 at $`z=2.6`$ (Barvainis et al. 1995) and APM 08279+5255 at $`z=3.9`$ (Irwin et al. 1998; Lewis et al. 1998b; Downes et al. 1999). The SEDs of these high-redshift galaxies, and two well studied low-redshift ultraluminous infrared galaxies (ULIRGs) – Arp 220 and Markarian 231 (Klaas et al. 1997; Rigopoulou, Lawrence & Rowan-Robinson 1996; Soifer et al. 1999) – are shown in Fig. 1.
## 2 Spectral energy distributions
The difference in the slope of the mid-infrared SED of Arp 220 and Markarian 231 is interpreted as evidence for an active galactic nucleus (AGN) in the core of Markarian 231. Powerful ionizing radiation from an AGN would heat a very small fraction of the dust grains in the galaxy to high temperatures near the nucleus, increasing the flux density of the galaxy at short wavelengths and thus producing a shallower spectrum. There have been various attempts to interpret the mid-/far-infrared SEDs of galaxies using models of radiative transfer (Granato, Danese & Franceschini 1996; Green & Rowan-Robinson 1996). However, given that a power-law spectrum convincingly accounts for the limited mid-infrared data available, it is reasonable to avoid such modeling and to assume that the mass of emitting dust in the source at temperatures between $`T`$ and $`T+\mathrm{d}T`$ is $`m_\mathrm{T}(T)\mathrm{d}T`$. Dust at each temperature can be associated with a $`\delta `$-function spectrum $`\delta (\nu \nu _0)`$, in which $`\nu _0(3+\beta )kT/h`$ (Blain & Longair 1993b). $`k`$ and $`h`$ are the Boltzmann and Planck constants respectively and $`\beta `$ is the index in the function that describes the spectral emissivity of dust $`ϵ_\nu \nu ^\beta `$; $`\beta 1`$–1.5. The mid-infrared SED of a source at redshift $`z`$ is
$$f_\nu _{T_{\mathrm{min}}}^{T_{\mathrm{max}}}m_\mathrm{T}(T)T^{4+\beta }\delta [\nu (1+z)\nu _0]dT.$$
(1)
If $`m_\mathrm{T}T^\alpha `$, then the result is a power-law SED $`f_\nu \nu ^a`$, with a spectral index $`a=4+\alpha +\beta `$. For dust temperatures spanning a range from $`T_{\mathrm{min}}=40`$ K to $`T_{\mathrm{max}}=2000`$ K, this emission covers a range of wavelengths from about 90 to 1.8 $`\mu `$m. The values of $`\alpha `$ associated with the galaxies above are listed in Table 1. $`\alpha 6`$ to $`9`$, and so only a very small fraction of the dust in these galaxies is heated to high temperatures.
In the case that a significant fraction of dust heating in a galaxy is caused by an AGN, it would be reasonable to assume that the dust temperature would depend on the distance from the nucleus of the galaxy $`r`$ as $`T_\mathrm{r}(r)r^\eta `$. A negative value of $`\eta `$ would be expected, reflecting a higher dust temperature in the more intense radiation field closer to the AGN. If the mass of dust enclosed in the spherical shell between radii $`r`$ and $`r+\mathrm{d}r`$ is $`m_\mathrm{r}(r)\mathrm{d}r`$, where $`m_\mathrm{r}r^\gamma `$, then the mid-infrared SED,
$$f_\nu _{r_{\mathrm{min}}}^{r_{\mathrm{max}}}m_\mathrm{r}(r)T_\mathrm{r}(r)^{4+\beta }\delta [\nu (1+z)\nu _0]dr,$$
(2)
again evaluates to a power law, with a spectral index $`a=3+\beta +(\gamma +1)/\eta `$. Note that for a blackbody in equilibrium with an unobscured point source $`\eta =0.5`$. In this case $`\gamma 2`$ is required in order to represent the SEDs of Markarian 231 and SMM J02399$``$0136 in Fig. 1 – the distribution expected if the radial density of dust is uniform. A more reasonable value of $`\eta `$ would be less than $`0.5`$, reflecting the effects of obscuration. This would imply that $`\gamma >2`$, and thus that the density of emitting dust increases with increasing radius.
Two of the three high-redshift galaxies with IRAS detections – APM 08279+5255 and H1413+117 – have a much flatter mid-infrared SED as compared with a typical low-redshift IRAS galaxy or SMM J02399$``$0136: see Fig. 1. SMM J02399$``$0136 lacks a detection by IRAS, but its mid-infrared SED is constrained by a 15-$`\mu `$m ISO measurement of Metcalfe: see Ivison et al. (1998) for further details. Given the limited mid-infrared data for SMM J02399$``$0136, it is certainly possible that its mid-infrared spectral index could be more negative than the value of $`1.7`$ listed in Table 1. An additional high-redshift source, the brightest detected in a 850-$`\mu `$m survey of the Hubble Deep Field (HDF; Hughes et al. 1998) with a flux density of 7 mJy, has a reported 15-$`\mu `$m flux density limit of less than 23 $`\mu `$Jy, indicating that its mid-infrared SED is steeper than that of SMM J02399$``$0136.
The slopes of these SEDs are not likely to be affected by contaminating sources that are picked up in the differently sized observing beams at each wavelength. The beam size of the 0.6-m telescopes used to determine the mid-infrared SEDs of these galaxies are about 5, 10, 25 and 40 arcsec at 12, 25, 60 and 100 $`\mu `$m respectively, as compared with about 7 and 14 arcsec for the ground-based submillimetre-wave telescopes used at 450 and 850 $`\mu `$m respectively. Any contamination from other sources nearby on the sky would thus be most significant at wavelengths of 60 and 100 $`\mu `$m, and lead to a artificial steepening of the mid-infrared SED.
The shallow SEDs of APM 08279+5255 and H1413 +117 could be caused by an unusually large fraction of hot dust in these galaxies; however, from optical observations of their QSO emission both are known to be gravitationally lensed by magnification factors of at least several tens. A systematically greater magnification for hotter dust components would increase the flux density at shorter wavelengths and so flatten the spectrum (Eisenhardt et al. 1996; Lewis et al. 1998b). This situation would arise very naturally if the hotter dust clouds were smaller and more central, as in the AGN model described in equation (2). A similar effect can be produced by the microlensing effect of individual stars within the lensing galaxy (Lewis et al. 1998a).
For a large magnification to occur, a distant source must lie very close to a caustic curve of a gravitational lens. The magnification is formally infinite on such a curve, but an upper limit $`A_{\mathrm{max}}`$ is imposed to the magnification if the source has a finite size $`d`$ (Peacock 1982). If the lensing galaxy can be modeled as a singular isothermal sphere (SIS), then $`A_{\mathrm{max}}d^1`$. Assuming an SIS lens, the image geometry and magnifications expected in such a situation are illustrated in Figs 2 and 3. Eisenhardt et al. (1996) present more sophisticated lens models that account for the geometry of high-resolution images of IRAS F10214+4724. Ellipticity in the lens modifies the magnification–size relation to $`A_{\mathrm{max}}d^{0.63}`$ on scales between 0.001 and 1 arcsec. In general, a similar magnification–size relationship holds regardless of both the geometry of the source and whether one or more objects is responsible for producing the lensing effect (Kneib et al. 1998).
The diagnostic feature of such a situation is the production of multiple images of comparable brightness. This is clearly the case for both APM 08279+5255 (Lewis et al. 1998b) and H1413+117 (Kneib et al. 1998), but less so for IRAS F10214+4724, which has a more asymmetric arc–counterimage geometry. However, in the lens models of both Broadhurst & Léhar (1995) and Eisenhardt et al. (1996), IRAS F10214+4724 lies very close to the tip of an astroid caustic, and so differential lensing would still be expected to flatten its mid-infrared SED. These references should be consulted for detailed lens models of these sources. The lens model if the most recently discovered source APM 08279+5255 is poorly constrained because of the lack of high-resolution optical and near-infrared images at present.
If the smallest, hottest dust clouds that are closest to the AGN lie on the caustic that is responsible for the multiple images of the AGN (equation 2), then differential magnification of the hot regions compared with the cooler regions is likely to modify the SED in the mid-infrared waveband. The effect on the SED can be calculated by including a magnification factor of $`A_{\mathrm{max}}(r)r^\mu `$ in the integrand of equation (2). For a SIS lens $`\mu =1`$. By evaluating equation (2) in this case, a power-law SED with a spectral index $`a=3+\beta +(\gamma +1\mu )/\eta `$ is obtained. The spectral index is reduced by $`\mathrm{\Delta }a=\mu /\eta `$ compared with the value obtained in the absence of lensing. Note that the modification is independent of the value of $`\gamma `$ and the form of $`m(r)`$. Following the same approach, but assuming an exponential decrease of $`T(r)`$, $`\mathrm{\Delta }a1`$.
## 3 Discussion
The mid-infrared SEDs of the high- and low-redshift AGNs SMM J02399$``$0136 and Markarian 231 have spectral indices $`a1.8`$. Despite being lensed by a cluster of galaxies SMM J02399$``$0136 does not lie close to a caustic, and so is not magnified differentially. Markarian 231 is not lensed. The strongly lensed distant galaxies APM 08279+5255 and H1413+117 have a spectral index $`a1.1`$. This would indicate that $`\mathrm{\Delta }a=\mu /\eta 0.7`$, and thus $`\eta 1.4`$ or $`0.9`$, depending on whether the lens is a SIS, with $`\mu =1`$, or matches the model of Eisenhardt et al. (1996), with $`\mu =0.6`$. Both spectral indices are considerably steeper than the unscreened blackbody value of $`0.5`$. Given the very great opacity of interstellar dust to ionizing radiation, this is entirely reasonable. To match the observed SEDs, the corresponding values of $`\gamma `$ of 7.3 and 4.0 are required, again indicating that most of the emitting dust is relatively cool/distant from the core.
IRAS F10214+4724 has a steeper mid-infrared SED, and a much smaller optical flux density than APM 08279+5255 and H1413+117 (Lewis et al. 1998b). Its mid-infrared SED is more similar to those of Markarian 231 and SMM J02399$``$0136. These features could both be explained if the optical depth of dust extinction into the central regions of IRAS F10214+4724 were sufficiently large to obscure not only the AGN in the optical waveband, but also the hottest dust clouds at longer mid-infrared wavelengths. Alternatively, the intrinsic unmagnified SED of IRAS F10214+4724 could be steeper than those of APM 08279+5255 and H1413+117, just as these two galaxies could have intrinsically flat mid-infrared SEDs.
Does the calculated value of $`\eta `$ correspond to a plausible size for the emitting objects? The temperature of the inner face of the dust cloud exposed to the AGN cannot exceed about 2000 K, or else the dust would sublime. A 2000-K blackbody would be in equilibrium with a $`10^{13}`$-L point source at a distance of 0.6 parsec. If $`\eta =1.4`$, then the dust temperature falls to 100, 50 and 30 K at distances of 5, 8 and 14 parsec respectively. If $`\eta =0.9`$, then the corresponding distances are 9, 17 and 26 parsec. These values are all in agreement with observations of low-redshift ULIRGs, in which the emitting region is less than several hundreds of parsecs in extent (Downes & Solomon 1998). Most of the luminosity of the galaxy is emitted by cool dust on larger scales. The equivalent radius of a blackbody sphere emitting 10<sup>13</sup> L at 50 K is 950 pc, several times larger than the observed sizes of nearby ULIRGs (Downes & Solomon 1998; Sakamoto et al. 1999).
The caustic curves predicted by the lens models that describe IRAS F10214+4724 (Broadhurst & Léhar 1995) and H1413+117 (Kneib et al. 1998) are about 0.8 and 0.2 arcsec in size in the plane of the source, larger than the scale of a 200-pc high-redshift source. Hence, the whole emitting region of the source should be subject to differential magnification; a difference in magnification by a factor of about 100 would be obtained between a 200-pc outer radius and a 2-pc inner radius for an SIS lens.
If the far-infrared SED of a galaxy is modelled by a single-temperature dust spectrum, then differential magnification would be expected to increase the temperature for which the best fit was obtained. An increase in temperature by a factor of about 20 per cent would typically be expected for discrete data points at the wavelengths given in Fig. 1. This effect may account for at least part of the difference between the very high rest-frame dust temperature of about 107 K inferred for APM 08279+5255 and the cooler dust temperatures inferred for nearby ULIRGs and SMM J02399$``$0136. If the redshift of a distant differentially magnified ULIRG were to be estimated from an observed SED, by fitting to a standard template, then the inferred redshift would be underestimated by a similar amount, up to 20 per cent.
The mid-infrared SEDs of distant galaxies that are known to be either unlensed or not to be subject to differential magnification are very uncertain. This is likely to remain true for some years, although the Wide-Field Infrared Explorer (WIRE) and Space InfraRed Telescope Facility (SIRTF) satellites will provide some valuable information. Operating at 12 and 25 $`\mu `$m, WIRE will specifically probe the mid-infrared SED of distant galaxies. The effect of differential magnification discussed here could increase the number of lensed AGN detected in these bands by a factor of about 10; however, this number is still expected to be less than about 1 per cent of the size of the full WIRE catalogue.
AGN-powered ULIRGs are expected to have flatter mid-infrared SEDs than starburst-powered ULIRGs, even in the absence of differential magnification. Hence, any further flattening due to differential magnification should make it easier to distinguish AGN from starbursts in the subsample of lensed galaxies in the catalogue.
The flux density from the inner regions of an AGN can experience strong variation on short time-scales. On scales of several pc, the response of the hottest, smallest components of the mid-infrared dust emission spectrum should be comparable to the light-crossing time – about 1 yr. Over the lifetime of SIRTF, any such variations, amplified by differential magnification, should be detectable. In order to probe the spatial structure and temperature distribution of dust in the inner regions of AGN directly, high-resolution mid-infrared imaging, and thus a space-borne mid-infrared interferometer, will be required (Mather et al. 1998a,b). Baselines of order 500 m at 20 $`\mu `$m will be required to probe 10-pc scales in high-redshift galaxies.
The conditions in the central regions of AGN can already be probed using a variety of methods in other wavebands. Reverberation mapping – observing the transient response of line emission to a variable continuum source – is possible in the optical waveband (Bahcall, Kozlovsky & Salpeter 1972; Blandford & McKee 1982; Wandel 1997). Observations of water masers using very long baseline radio interferometry (Miyoshi et al. 1995) have revealed the structure of the accretion disk in NGC 4258. Modifications to the profiles of X-ray fluorescence lines due to strong gravity in the innermost regions of accretion disks have also been observed (Reynolds & Fabian 1997).
## 4 Conclusions
1. Distant lensed ULIRGs are likely to have their mid-infrared SEDs made more shallow by the effects of differential magnification. While observations of these sources must be exploited in order to investigate the nature of this class of galaxies, careful account must be taken of the uncertainties introduced by differential magnification. Recent submillimetre-selected samples are expected to be largely immune to this problem; their mid-infrared SEDs will be probed by the forthcoming SIRTF mission.
2. The SED of the longest known high-redshift ULIRG IRAS F10214+4724 is probably affected by differential magnification in the same way as those of APM 08279+5255 and H1413+117. However, it has a steeper mid-infrared SED. This could be due to either a steeper intrinsic SED, which is still flattened by differential magnification, or to a greater optical depth to dust extinction, which obscures the most central regions of the galaxy, even in the mid-infrared waveband.
3. The SEDs of a sample of lensed galaxies can be used to probe the conditions in the cores of distant dusty galaxies. A larger sample of these objects will be compiled by the Planck Surveyor satellite (Blain 1998). A direct test of the properties of dust in the central regions of high-redshift ULIRGs will require a space-borne interferometer such as SPECS (Mather et al. 1998a,b).
## Acknowledgements
I thank Jean-Paul Kneib, Malcolm Longair, Priya Natarajan and an anonymous referee for their helpful comments on the manuscript, and PPARC and MENRT for support. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.
|
no-problem/9812/cond-mat9812188.html
|
ar5iv
|
text
|
# Spin-Pseudospin Coherence and CP3 Skyrmions in Bilayer Quantum Hall Ferromagnets
## Abstract
We analyze bilayer quantum Hall ferromagnets, whose underlying symmetry group is SU(4). Spin-pseudospin coherence develops spontaneously when the total electron density is low enough. Quasiparticles are CP<sup>3</sup> skyrmions. One skyrmion induces charge modulations on both of the two layers. At the filling factor$`\nu =2/m`$ one elementary excitation consists of a pair of skyrmions and its charge is $`2e/m`$. Recent experimental data due to Sawada et al. \[Phys. Rev. Lett. 80, 4534 (1998)\] support this conclusion.
The quantum Hall (QH) effect is a remarkable macroscopic quantum phenomenon in the two-dimensional electron system . Attention has recently been paid to quantum coherence in QH systems. The kinetic and Coulomb Hamiltonians have the spin SU(2) symmetry. Its spontaneous breakdown leads to a spin coherence and turns the system into a QH ferromagnet. The effective Hamiltonian is the SU(2) nonlinear sigma (NL$`\sigma `$) model . Quasiparticles are CP<sup>1</sup> skyrmions .
In this paper we analyze skyrmion excitations in bilayer QH (BLQH) ferromagnets. The lowest Landau level (LLL) contains four energy levels corresponding to the spin and layer (pseudospin) degrees of freedom. The SU(4) symmetry underlies the BLQH system. Its spontaneous breakdown leads to a spin-pseudospin coherence. The effective Hamiltonian is the SU(4) NL$`\sigma `$ model . Quasiparticles are CP<sup>3</sup> skyrmions . They have two characteristic features: (A) One skyrmion induces charge modulations on both of the two layers. The main part of the activation energy is the capacitive charging energy. (B) One elementary excitation consists of a pair of skyrmions at $`\nu =2/m`$ with $`m`$ odd. It carries the charge $`2e/m`$. Our theoretical analysis accounts for recent experimental data due to Sawada et al.. Throughout the paper we use the natural units $`\mathrm{}=c=1`$.
QH ferromagnets: We analyze skyrmion excitations at the filling factor $`\nu 2\pi \rho _0/eB_{}=1/m`$ with $`m`$ odd. We use an improved composite-boson (CB) theory , which is proposed based on a suggestion due to Girvin et al. . The advantage of the scheme is a direct connection between the semiclassical property of an excitation and its microscopic wave function. We start with a review of monolayer QH ferromagnets . The analysis of BLQH ferromagnets is its straightforward generalization with a replacement of SU(2) by SU(4).
The kinetic Hamiltonian for planar electrons with mass $`M`$ in the perpendicular magnetic field $`B_{}`$ is
$$H_K=\frac{1}{2M}d^2x\mathrm{\Psi }^{}(𝒙)(P_xiP_y)(P_x+iP_y)\mathrm{\Psi }(𝒙),$$
(1)
where $`P_j=i_j+eA_j`$ is the covariant momentum with $`A_j=\frac{1}{2}\epsilon _{jk}x_kB_{}`$; $`\epsilon _{12}=\epsilon _{21}=1`$ and $`\epsilon _{11}=\epsilon _{22}=0`$. The electron field $`\mathrm{\Psi }(𝒙)`$ is a two-component spinor made of the spin-up ($`\psi ^{}`$) and spin-down ($`\psi ^{}`$) field. The symmetry group of this Hamiltonian is U(2)=U(1)$``$SU(2).
When the Zeeman energy is small, a spin coherence develops spontaneously. This is described by introducing the two-component CB field by the formula ,
$$\mathrm{\Phi }(𝒙)=e^{𝒜(𝒙)}e^{ie\mathrm{\Theta }(𝒙)}\mathrm{\Psi }(𝒙),$$
(2)
where $`𝒜(𝒙)`$ is the auxiliary field obeying $`\mathbf{}^2𝒜(𝒙)=2\pi m[\rho (𝒙)\rho _0]`$; $`\rho (𝒙)\mathrm{\Psi }^{}(𝒙)\mathrm{\Psi }(𝒙)`$ is the electron density. The phase field $`\mathrm{\Theta }(𝒙)`$ attaches $`m`$ units of flux quantum $`2\pi /e`$ to each electron via the relation, $`\epsilon _{ij}_i_j\mathrm{\Theta }(𝒙)=(2\pi /e)m\rho (𝒙)`$. The effective magnetic field is
$$_{\text{eff}}(𝒙)=B_{}\epsilon _{ij}_i_j\mathrm{\Theta }(𝒙)=B_{}(2\pi /e)m\rho (𝒙).$$
(3)
It vanishes, $`_{\text{eff}}_\text{g}=0`$, on the ground state at $`\nu =1/m`$. Substituting (2) into (1), the kinetic Hamiltonian reads
$$H_K=\frac{1}{2M}d^2x\mathrm{\Phi }^{}(𝒙)(𝒫_xi𝒫_y)(𝒫_x+i𝒫_y)\mathrm{\Phi }(𝒙),$$
(4)
where $`𝒫_j=i_j(\epsilon _{jk}+i\delta _{jk})_k𝒜(𝒙)`$ is the covariant momentum. We have defined $`\mathrm{\Phi }^{}(𝒙)\mathrm{\Phi }^{}(𝒙)e^{2𝒜(𝒙)}`$, with which $`\rho (𝒙)=\mathrm{\Psi }^{}(𝒙)\mathrm{\Psi }(𝒙)=\mathrm{\Phi }^{}(𝒙)\mathrm{\Phi }(𝒙)`$. An analysis of the Lagrangian shows that the canonical conjugate of $`\phi ^\alpha (𝒙)`$ is not $`i\phi ^\alpha (𝒙)`$ but $`i\phi ^\alpha (𝒙)`$.
We decompose the CB field as
$$\mathrm{\Phi }(𝒙)=e^{𝒜(𝒙)}\varphi (𝒙)𝒏(𝒙),$$
(5)
with the U(1) component $`\varphi (𝒙)=e^{i\chi (𝒙)}\sqrt{\rho (𝒙)}`$ and the SU(2) component $`𝒏(𝒙)`$: It is the CP<sup>1</sup> field subject to the constraint, $`𝒏^{}(𝒙)𝒏(𝒙)=1`$. The spin density is
$$S^a(𝒙)=\frac{1}{2}\rho (𝒙)s^a(𝒙),s^a(𝒙)=𝒏^{}(𝒙)\lambda ^a𝒏(𝒙),$$
(6)
where $`\lambda ^a`$ are the Pauli matrices.
At sufficiently low temperature it is reasonable to focus our attention to physics taking place within the LLL. The Hilbert space $`_{\text{LLL}}`$ is constructed by imposing the LLL condition so that the kinetic term (4) is quenched. It has a simple expression in terms of the CB field,
$$(𝒫_x+i𝒫_y)\mathrm{\Phi }(𝒙)|𝔖=\frac{i}{\mathrm{}_B}\frac{}{z^{}}\mathrm{\Phi }(𝒙)|𝔖=0.$$
(7)
The complex number is $`z=(x+iy)/2\mathrm{}_B`$ with $`\mathrm{}_B`$ the magnetic length. Hence, the wave function for composite bosons is analytic and symmetric in all $`N`$ coordinates,
$$\mathrm{\Omega }[z]=0|\mathrm{\Phi }(𝒙_1)\mathrm{}\mathrm{\Phi }(𝒙_N)|𝔖.$$
(8)
The wave function for electrons is $`𝔖[𝒙]=\mathrm{\Omega }[z]𝔖_{\text{LN}}[𝒙]`$, where $`𝔖_{\text{LN}}[𝒙]`$ is the familiar Laughlin wave function. Here, $`[𝒙]=(𝒙_1,𝒙_2,\mathrm{},𝒙_N)`$ and $`[z]=(z_1,z_2,\mathrm{},z_N)`$. Any excitation confined to the LLL is described by a choice of the analytic spinor factor $`\mathrm{\Omega }[z]`$.
We analyze the CB theory semiclassically, where the bosonic field operator is approximated by a c-number field. It follows from (8) that the wave function is
$$𝔖[𝒙]=\underset{r}{}\mathrm{\Phi }(𝒙_r)𝔖_{\text{LN}}[𝒙],$$
(9)
where $`\mathrm{\Phi }(𝒙)`$ is analytic. If the Zeeman energy is neglected, the ground state is determined by minimizing the Coulomb energy. The one-point function $`\mathrm{\Phi }(𝒙)`$ is a constant vector pointing to an arbitrary direction in the SU(2) space, as implies a spontaneous breakdown of the SU(2) symmetry. Actually, a small Zeeman interaction fixes this direction so that $`s^a(𝒙)=\delta ^{az}`$, or
$$\phi ^{}(𝒙)_\text{g}=\sqrt{\rho _0},\phi ^{}(𝒙)_\text{g}=0.$$
(10)
The ground-state wave function is given by (9) with (10).
The semiclassical LLL condition follows from (5),
$$\phi ^\alpha (𝒙)=e^{𝒜(𝒙)}e^{i\chi (𝒙)}\sqrt{\rho (𝒙)}n^\alpha (𝒙)\omega ^\alpha (z),$$
(11)
where various fields are classical ones. This is solved by $`\chi (𝒙)=0`$ and $`n^\alpha (𝒙)=\omega ^\alpha (z)/\sqrt{_\alpha |\omega ^\alpha |^2}`$. The soliton equation follows trivially from (11),
$$\frac{\nu }{4\pi }\mathbf{}^2\mathrm{ln}\rho (𝒙)\rho (𝒙)+\rho _0=\nu Q(𝒙),$$
(12)
where $`Q(𝒙)`$ is the topological charge density,
$$Q(𝒙)=\frac{1}{4\pi }\mathbf{}^2\mathrm{ln}\left(\underset{\alpha }{}|\omega ^\alpha (z)|^2\right).$$
(13)
The lightest skyrmion on the ground state (10) is
$$\phi ^{}(𝒙)_{\text{sky}}=z\sqrt{\rho _0},\phi ^{}(𝒙)_{\text{sky}}=\kappa \sqrt{\rho _0},$$
(14)
with which the wave function is given by (9). For a large skyrmion ($`\kappa 1`$), the soliton equation (12) is solved iteratively and agrees with the familiar expression ,
$$\varrho (𝒙)\rho (𝒙)\rho _0\nu Q(𝒙)=\frac{\nu }{\pi }\frac{4(\kappa \mathrm{}_B)^2}{[r^2+4(\kappa \mathrm{}_B)^2]^2}.$$
(15)
The skyrmion is reduced to the vortex in the limit $`\kappa 0`$.
The excitation energy of one skyrmion consists of the exchange energy $`E_{\text{ex}}`$, the electrostatic energy $`E_C`$ and the Zeeman energy $`E_Z`$. Minimizing their sum, we determine the skyrmion size $`\kappa `$, the skyrmion energy $`E_{\text{sky}}`$ and the flipped-spin number $`N_{\text{spin}}`$ as follows :
$`\kappa `$ $`{\displaystyle \frac{1}{2}}(\beta \nu )^{1/3}\left\{\stackrel{~}{g}\mathrm{ln}\left({\displaystyle \frac{\sqrt{2\pi }}{32\stackrel{~}{g}}}+1\right)\right\}^{1/3},`$ (16)
$`E_{\text{sky}}`$ $`\left(\nu \sqrt{{\displaystyle \frac{\pi }{32}}}+{\displaystyle \frac{3\beta \nu ^2}{4\kappa }}\right)E_C^0,`$ (17)
$`N_{\text{spin}}`$ $`2\nu \kappa ^2\mathrm{ln}\left({\displaystyle \frac{\sqrt{2\pi }}{32\stackrel{~}{g}}}+1\right).`$ (18)
Here, $`\rho _s=\nu e^2/(16\sqrt{2\pi }\epsilon \mathrm{}_B)`$, $`E_C^0=e^2/\epsilon \mathrm{}_B`$ and $`\stackrel{~}{g}=g^{}\mu _BB/E_C^0`$. The parameter $`\beta `$ represents the strength of the Coulomb energy; it is calculated as $`\beta =3\pi ^2/64`$ for a sufficiently large skyrmion in an ideal planar system. However, an actual skyrmion is not large and there will also be a correction from a finiteness of the layer width. We use $`\beta `$ as a phenomenological parameter. We fix it as $`\beta =0.24`$ by requiring the skyrmion spin $`N_{\text{spin}}`$ to agree with the experimental data due to Barrett et al. at $`\nu =1`$: $`N_{\text{spin}}3.7`$ at $`B=7.05`$ Tesla, where $`\kappa 1.0`$ and $`\stackrel{~}{g}0.015`$. Note that $`N_{\text{spin}}5.3`$ ($`\kappa 1.10`$) at $`B=3`$ Tesla and $`N_{\text{spin}}2.7`$ ($`\kappa 0.96`$) at $`B=15`$ Tesla. Our results are consistent with the previous ones .
The excitation energy of a skyrmion-antiskyrmion pair will be given by $`\mathrm{\Delta }=2E_{\text{sky}}\mathrm{\Gamma }_{\text{offset}}`$ with a sample dependent offset $`\mathrm{\Gamma }_{\text{offset}}`$. In Fig.1 we have fitted the experimental data due to Schmeller et al. based on formula (17) with $`\beta =0.24`$, where an appropriate offset $`\mathrm{\Gamma }_{\text{offset}}`$ is used for each curve. The theoretical curves reproduce all data remarkably well.
BLQH systems: We proceed to analyze BLQH states at $`\nu =1/m`$ and $`2/m`$ with $`m`$ odd. There are three experimental techniques to elucidate various states. The first one is to change the total electron density $`\rho _0`$. By increasing $`\rho _0`$ the interlayer separation $`d`$ is effectively increased compared with the magnetic length, $`d/\mathrm{}_B\sqrt{\rho _0}`$. Hence, as $`\rho _0\mathrm{}`$ the BLQH system at $`\nu =2/m`$ will be decomposed into two independent $`\nu =1/m`$ monolayer systems; as $`\rho _00`$ an interlayer coherence will develop spontaneously, which has been argued at $`\nu =1/m`$ and will be argued at $`\nu =2/m`$ in this paper. Consequently, we expect a phase transition at $`\nu =2/m`$ between these two phases but not at $`\nu =1/m`$. These two phases are clearly distinguishable by using the second technique, i.e. by applying gate bias voltage. We can control the density difference $`\sigma _0`$ between the two quantum wells, where $`\sigma _0=(\rho _0^1\rho _0^2)/\rho _0`$ with $`\rho _0^\alpha `$ the density in the layer $`\alpha `$. Only the coherent state is stable against an arbitrary change of $`\sigma _0`$. All these features have been experimentally confirmed in recent works due to Sawada et al. at $`\nu =1`$ and $`2`$. The coherent state at $`\nu =2/3`$ has not been observed by them presumably because of a poor sample quality. The third technique is to tilt the sample with the perpendicular magnetic field $`B_{}`$ fixed. In the high-density data the activation energy is found to increase at $`\nu =2`$ and $`2/3`$ as normally as in the monolayer system: Indeed, we can fit the data by the monolayer skyrmion formula (17). We conclude that elementary excitations are monolayer skyrmions. In the low-density data it is found to decrease anomalously at $`\nu =1`$ and $`\nu =2`$, as is the phenomenon discovered by Murphy et al. at $`\nu =1`$: It is a behavior intrinsic to the interlayer coherent state in the BLQH system.
Spin-pseudospin coherence: We analyze elementary excitations in the coherent state of the BLQH system. The electron field $`\mathrm{\Psi }(𝒙)`$ has four components $`\psi ^1`$, $`\psi ^1`$, $`\psi ^2`$ and $`\psi ^2`$, where the layer is indexed by 1 and 2. The kinetic Hamiltonian is given by (1), whose symmetry group is U(4)=U(1)$``$SU(4). When the interlayer and intralayer Coulomb energies are nearly equal and dominate the system, we expect a spin-pseudospin coherence to emerge.
Such a new phase is described in terms of the CB field defined by (2). The CB field is decomposed into the U(1) and SU(4) components by (5). Here, $`𝒏(𝒙)`$ is the CP<sup>3</sup> field. The group SU(4) is generated by the hermitian, traceless, $`4\times 4`$ matrices $`\lambda ^a`$, $`a=1,2,\mathrm{},15`$, normalized as $`\text{Tr}(\lambda ^a\lambda ^b)=2\delta ^{ab}`$. They are the generalization of the Pauli matrices in case of SU(2). The SU(4) spin density is given by (6) with such $`\lambda ^a`$.
The Coulomb energy is decomposed into two terms,
$$E_C^\pm =\frac{1}{2}d^2xd^2yV^+(𝒙𝒚)\varrho _\pm (𝒙)\varrho _\pm (𝒚),$$
(19)
where $`V^\pm (𝒙)=(e^2/2\epsilon )\left(|𝒙|^1\pm (|𝒙|^2+d^2)^{1/2}\right)`$ with the interlayer separation $`d`$; $`\varrho _+\varrho (𝒙)=\rho (𝒙)\rho _0`$; $`\varrho _{}(𝒙)=\rho ^1(𝒙)\rho ^2(𝒙)\rho _0^1+\rho _0^2`$ with $`\rho ^1(𝒙)=\rho ^1(𝒙)+\rho ^1(𝒙)`$ and $`\rho ^2(𝒙)=\rho ^2(𝒙)+\rho ^2(𝒙)`$. The Coulomb energy $`E_C^+`$, possessing the SU(4) symmetry, is the driving force to realize the QH system. The term $`E_C^{}`$ describes the capacitive charging energy between the two layers. It vanishes in the limit $`d0`$.
The Hilbert space $`_{\text{LLL}}`$ is defined by the LLL condition (7). In the semiclassical approximation the wave function is given by (9) at $`\nu 1`$. The SU(4) spin-pseudospin coherence is shown to develop spontaneously, precisely as the SU(2) spin coherence is. The ground state at $`\nu 1`$ is given by
$`\phi ^1_\text{g}`$ $`=\sqrt{\rho _0/2}\sqrt{1+\sigma _0},\phi ^1_\text{g}=0,`$ (20)
$`\phi ^2_\text{g}`$ $`=\sqrt{\rho _0/2}\sqrt{1\sigma _0},\phi ^2_\text{g}=0,`$ (21)
where $`\rho _0^1=\frac{1}{2}\rho _0(1+\sigma _0)`$ and $`\rho _0^2=\frac{1}{2}\rho _0(1\sigma _0)`$. It is convenient to use a new set of CB fields,
$`\phi ^S`$ $`=\sqrt{(1+\sigma _0)/2}\phi ^1+\sqrt{(1\sigma _0)/2}\phi ^2,`$ (22)
$`\phi ^A`$ $`=\sqrt{(1\sigma _0)/2}\phi ^1\sqrt{(1+\sigma _0)/2}\phi ^2,`$ (23)
and a similar set for the spin-down fields. They are reduced to the symmetric and antisymmetric fields at the balanced point ($`\sigma _0=0`$). We call them the “bond” and “antibond” fields. The ground-state value (21) is transformed into $`\stackrel{~}{\mathrm{\Phi }}_\text{g}=\sqrt{\rho _0}(1,0,0,0)`$ in terms of the new fields, $`\stackrel{~}{\mathrm{\Phi }}(\phi ^S,\phi ^S,\phi ^A,\phi ^A)`$.
The tunneling energy is
$$E_T=\frac{1}{2}\mathrm{\Delta }_{\text{SAS}}\sqrt{1\sigma _0^2}d^2x[\varrho _{}^S(𝒙)\varrho _{}^A(𝒙)],$$
(24)
where $`\varrho _{}^S(𝒙)=\rho ^S(𝒙)\rho ^S(𝒙)_\text{g}`$ with $`\rho ^S=\phi ^S\phi ^S+\phi ^S\phi ^S`$ and similar equations for $`\varrho _{}^A(𝒙)`$. The tunneling gap is $`(1\sigma _0^2)^{1/2}\mathrm{\Delta }_{\text{SAS}}`$ on the state (21).
A skyrmion excitation flips in general spins and pseudospins. It is a CP<sup>3</sup> skyrmion described by
$$\stackrel{~}{\mathrm{\Phi }}_{\text{sky}}=\sqrt{\rho _0}(z,\kappa _1,\kappa _2,\kappa _3),$$
(25)
with constant parameters $`\kappa _i`$. Its classical configuration is determined by (11) $``$ (13), and (15) with $`\kappa ^2=_\alpha \kappa _\alpha ^2`$. For definiteness we assume hereafter that the tunneling energy is much larger than the Zeeman energy. (For instance, $`\mathrm{\Delta }_{\text{SAS}}/g^{}\mu _BB4`$ at $`B=`$ 5 Tesla in the sample of Ref..) Then, we have $`\kappa =\kappa _10`$, $`\kappa _2=\kappa _3=0`$. It is identical to the CP<sup>1</sup> skyrmion (14) in the spin space. At the balanced point the skyrmion size, the skyrmion energy and the flipped-spin number are given by the same formulas as (16) $``$ (18), where $`\beta `$ now depends on the layer separation $`d`$. Here, we concentrate our attention to its dependence on the imbalance parameter $`\sigma _0`$. The term involving $`\sigma _0`$ is only the charging term (19) for $`\nu 1`$. We give a numerical estimation of the activation energy at $`\nu =1`$ in Fig.2 by using the sample parameters ($`d=231`$Å, $`\mathrm{}_B=120`$Å) of Ref.. The theoretical curve explains the experimental data quite well with a reasonable skyrmion size $`\kappa 1`$.
BLQH ferromagnet at $`\nu =2`$: A caution is needed to analyze the BLQH system at $`\nu =2`$ since one Landau state contains two electrons. We attach one unit of flux to each electron and transform it into the CB field by formula (2) with $`m=1`$. The effective magnetic field is not given by (3) but by
$$2_{\text{eff}}=2B_{}\epsilon _{ij}_i_j\mathrm{\Theta }(𝒙)=2B_{}(2\pi /e)m\rho (𝒙).$$
(26)
It vanishes on the homogeneous ground state at $`\nu =2/m`$. Due to the fermi statistics the wave function is the antisymmetric product of two wave functions (9),
$`𝔖[𝒙]=`$ $`{\displaystyle \underset{r}{}}[\mathrm{\Phi }_1(𝒙_r)\mathrm{\Phi }_2(𝒙_r)\mathrm{\Phi }_2(𝒙_r)\mathrm{\Phi }_1(𝒙_r)]`$ (28)
$`\times 𝔖_{\text{LN}}[𝒙]^2,`$
where $`\mathrm{\Phi }_1(𝒙)`$ and $`\mathrm{\Phi }_2(𝒙)`$ are analytic and satisfy
$$\underset{\alpha }{}|\phi _1^\alpha (𝒙)|^2=\underset{\alpha }{}|\phi _2^\alpha (𝒙)|^2,$$
(29)
as follows from the semiclassical LLL condition (11).
When $`\mathrm{\Delta }_{\text{SAS}}g^{}\mu _BB`$, the spin-up and spin-down “bond” states are filled. Hence, the ground state is given by (28) with a set of two constant CB fields,
$$\stackrel{~}{\mathrm{\Phi }}_1_\text{g}=\sqrt{\frac{\rho _0}{2}}(1,0,0,0),\stackrel{~}{\mathrm{\Phi }}_2_\text{g}=\sqrt{\frac{\rho _0}{2}}(0,1,0,0),$$
(30)
in terms of the “bond” and “antibond” fields. This might be identified with the canted state for $`\mathrm{\Delta }_{\text{SAS}}g^{}\mu _BB`$. A skyrmion excitation flips pseudospins, or induces a coherent tunneling excitation. It is described by (28) with a set of two CB fields,
$`\stackrel{~}{\mathrm{\Phi }}_1_{\text{sky}}`$ $`=\sqrt{\rho _0/2}(z,\kappa _1,\kappa _2,\kappa _3),`$ (31)
$`\stackrel{~}{\mathrm{\Phi }}_2_{\text{sky}}`$ $`=\sqrt{\rho _0/2}(\kappa _1^{},z,\kappa _2^{},\kappa _3^{}),`$ (32)
with $`\kappa ^2_\alpha \kappa _\alpha ^2=_\alpha \kappa _{}^{}{}_{\alpha }{}^{2}`$. It consists of two CP<sup>3</sup> skyrmions (25), and the skyrmion charge is $`2e`$. We emphasize that there exists no skyrmion composed of one CP<sup>3</sup> skyrmion at $`\nu =2`$ because of the constraint (29).
An estimation of the excitation energy of the skyrmion (32) is straightforward. We concentrate our attention to its dependence on the imbalance parameter $`\sigma _0`$. The terms involving $`\sigma _0`$ are the charging energy (19) and the tunneling energy (24). The charging energy increases while the tunneling energy decreases as $`\sigma _0`$ increases. We give a numerical estimation in Fig.2 by using the sample parameters ($`d=231`$Å, $`\mathrm{}_B=228`$Å, $`\mathrm{\Delta }_{\text{SAS}}=`$6.8K) of Ref.. The vortex limit ($`\kappa 0`$) gives a best fitting of the experimental data because of a large tunneling gap ($`\mathrm{\Delta }_{\text{SAS}}=`$6.8K). We have also given a theoretical curve for a would-be skyrmion carrying charge $`e`$ at $`\nu =2`$ by using the same parameters, where the charging energy and the tunneling energy are found to cancel each other almost completely (Fig.2).
The driving force of the SU(4) spin-pseudospin coherence is the Coulomb exchange energy arising from the SU(4)-invariant Coulomb term $`E_C^+`$ in (19). Provided the exchange energy is dominant, it is obvious that the SU(4) coherence develops also at $`\nu =2/m`$ with the skyrmion charge $`2e/m`$, and at $`\nu =6,10,14,\mathrm{}`$ with charge $`2e`$.
I am very grateful to Kenichi Sasaki, Anju Sawada, Sankar Das Sarma, Eugene Demler, Allan MacDonald and Jim Eisenstein for fruitful conversations on the subject. Part of this work was done at ITP, Santa Barbara. This research was supported in part by the National Science Foundation under Grant No. PHY9407194.
|
no-problem/9812/astro-ph9812079.html
|
ar5iv
|
text
|
# Quasi–thermal Comptonization and gamma–ray bursts
## 1 Introduction
After the observational breakthrough of $`Beppo`$SAX (Costa et al. 1997; Van Paradijs et al. 1997) we are now starting to disclose the physics of gamma–ray bursts (GRB). The huge energy and power release required by their cosmological distances supports the fireball scenario (Cavallo & Rees 1978; Rees & Mészáros 1992; Mészáros & Rees 1993), whose evolution and behavior is unfortunately largely independent of its origin.
We do not know yet in detail how the GRB event is related to the afterglow emission, but in the most accepted scheme of formation of and emission from internal/external shocks (Rees & Mészáros 1992; Rees & Mészáros 1994; Sari & Piran 1997), the former is due to collisions of pairs of relativistic shells (internal shocks), while the latter is generated by the collisionless shocks produced by shells interacting with the interstellar medium (external shocks). The short spikes ($`t_{\mathrm{𝑣𝑎𝑟}}`$10 ms) observed in the high energy light curves suggest that shell–shell collisions occur at distances $`R10^{12}`$$`10^{13}`$ cm from the central source within a plasma moving with a bulk Lorenz factor $`\mathrm{\Gamma }100`$. The fireball starts to be decelerated by the interstellar medium further out, at a distance which depends on the assumed density of this material.
Up to now, the main radiation mechanism assumed to give rise to both the burst event and the afterglow is synchrotron emission (Rees & Mészáros 1994; Sari, Narayan & Piran 1996; Sari & Piran 1997; Panaitescu & Mészáros 1998). In fact if magnetic field, electrons and protons share the available energy $`E=10^{50}E_{50}`$ erg of the shell, the electrons reach typical random Lorentz factors $`\gamma m_p/m_e`$, while the assumption of a Poynting flux $`L_B=R^2\mathrm{\Gamma }^2B^2c/2=10^{50}L_{B,50}`$ erg s<sup>-1</sup> implies a comoving magnetic field of the order $`BL_B^{1/2}/(R\mathrm{\Gamma })10^5L_{B,50}^{1/2}/(R_{13}\mathrm{\Gamma }_2)`$ G. <sup>1</sup><sup>1</sup>1Here and in the following we parametrise a quantity $`Q`$ as $`Q=10^xQ_x`$ and adopt cgs units. For these values of $`\gamma `$ and $`B`$, the typical observed synchrotron frequency is $`\nu _s0.5L_{B,50}^{1/2}/[R_{13}(1+z)]`$ MeV, independently of the bulk Lorentz factor $`\mathrm{\Gamma }`$, and in excellent agreement with the observed values of the peak of GRB spectra in a $`\nu F(\nu )`$ representation.
In this model therefore the assumption of energy equipartition plays a key role. And, at a closer look, this implicitly requires a number of constraints to be satisfied in the emission regions. The main one concerns the acceleration of the electrons, which must be impulsive (i.e. on timescales much shorter than the cooling ones). Further requirements will be discussed in §2.
One can envisage an alternative scenario, which we describe in §3, where the key role is instead played by the balance between the cooling and heating processes. This would be favoured if the emitting region occupies the entire shell volume rather than the narrow region associated with a planar shock, as in the ‘equipartition’ model.
An immediate prediction following this hypothesis is that the typical energy of the emitting electrons is mildly relativistic, and the main radiation process is quasi–thermal Comptonization. Also this model implies the existence of a characteristic observed frequency of a few MeV, controlled by the feedback introduced by the effect of electron–positron pair production.
Earlier attempts to explain burst radiation with multiple Compton scatterings have been done by Liang et al. (1997), who considered emission by a very dense population of non–thermal relativistic electrons in a weak magnetic field, in the context of bursts located in an extended galactic halo. Liang (1997) later extended this model for bursts at Gpc distances, assuming an emitting region of the order of $`10^{15}`$ cm, a magnetic field value of the order of 0.1 G and thermal electrons at a temperature of a few keV.
We instead first discuss the constraints that the ‘synchrotron’ scenario must satisfy and then, using the very same parameters (except for the electron energy), propose that quasi–thermal Comptonization can quite naturally dominate the cooling.
Some consequences of our scenario are presented in §4, while our findings are discussed in §5.
For simplicity we will consider spherical shells of comoving width <sup>2</sup><sup>2</sup>2Primed quantities are measured in the comoving frame. The magnetic field, the random Lorentz factor $`\gamma `$, the optical depth $`\tau `$ and the Comptonization parameter $`y`$, not primed for clarity, are also measured in the comoving frame. $`\mathrm{\Delta }R^{}`$, moving with a bulk Lorentz factor $`\mathrm{\Gamma }`$, and a monoenergetic electron distribution peaked at a (random) energy $`\gamma m_ec^2`$. Generalizations to jet–like outflows and more complex particle distributions are straightforward.
## 2 Constraints on the ‘equipartition’ scenario
Shell–shell collision produces a shock that has to accelerate particles to an energy $`\gamma m_\mathrm{e}c^2`$ in a timescales $`t_{\mathrm{acc}}^{}`$ much shorter than the cooling time $`t_{\mathrm{cool}}^{}(\gamma )`$, since otherwise the final random energy of the particle would be controlled by its cooling rate and not by the equipartition condition.
Once accelerated to the equipartition Lorentz factor $`\gamma m_p/m_e`$ and radiatively cooled down, a particle will not be accelerated again during the shell–shell interaction, in order for the total random energy not to exceed the available energy in the relative bulk motion of the two shells.
The inverse Compton power must be at most of the same order of the synchrotron one, as otherwise the luminosity we observe in the hard X–rays would be largely dominated by the luminosity emitted beyond the GeV band, with obvious problems for the total energetics. This requires that $`\tau _e\gamma ^21`$, where $`\tau _e`$ is the optical depth of the emitting relativistic electrons (not to be confused with the total scattering optical depth $`\tau _T`$ of the shell which is of order unity at $`R10^{13}`$ cm). This condition is valid if the Compton scattering process is entirely in the Thomson regime, which is appropriate as long as $`\gamma 760B_5^{1/3}`$. Higher energy electrons would scatter in the Thomson limit synchrotron photons of lower frequencies: for $`\gamma 10^3`$ the reduction in the Compton luminosity due to Klein–Nishina effects is of the order of $`3B_5^{4/3}\gamma _3^4`$.
The condition $`\tau _e\gamma ^21`$ translates in demanding that the width $`dR^{}`$ of the emitting region of the shock (assumed to be plane–parallel) satisfies $`dR^{}/\mathrm{\Delta }R^{}\tau _T^1\gamma ^210^6`$. This in turn controls the cooling timescale, which must be of the order of $`dR^{}/c`$, and we obtain:
$$t_{cool}^{}\frac{\mathrm{\Delta }R^{}}{c\tau _T\gamma ^2}\gamma \frac{\mathrm{\Delta }R^{}}{\tau _T}\frac{\sigma _TB^2(1+\tau _e\gamma ^2)}{6\pi m_ec^2}86\frac{\mathrm{\Delta }R_{11}^{}B_5^2}{\tau _T}$$
(1)
indicating that either the magnetic field is stronger than $`10^5`$ G or the total shell optical depth $`\tau _T0.1`$, to allow the electrons to reach $`\gamma 10^3`$.
Consider also that if the density of electrons is increased by electron–positron pair production, the mean energy per lepton is less than $`m_p/m_e`$ by the factor equal to the ratio of leptons to proton densities $`n_e^{}/n_p^{}`$.
We consider these constraints – i.e. impulsive acceleration, limited Compton power and absence of copious pair production – quite demanding (at least among the limits imposed by considerations on radiation processes). The emitting region cannot be very compact and cannot have a width larger than $`dR^{}`$. Note that in some models the shocked region is instead considered to be complex and extended, as in the shock structure resulting from Rayleigh-Taylor instability, which can occupy the entire shell volume (Pilla & Loeb 1998). In this case there are many more electrons emitting at a given time, with a consequent building up of the radiation energy density and an increased Compton luminosity. In addition, the larger cooling rate can limit the typical electron energies to values much below the equipartition one.
In the next section we will therefore examine the possibility that the above conditions are not satisfied.
## 3 Quasi–thermal Comptonization
Let us assume that the heating process for a typical electron lasts for the duration time of the shell–shell interaction, $`\mathrm{\Delta }R^{}/c`$. The maximum amount of energy given to a single lepton is of the order of $`m_pc^2n_p^{}/n_e^{}`$ (not to violate the total energetics), which, when released over the above timescale corresponds to a total (average) heating rate $`\dot{E}_{heat}^{}=n_p^{}m_pc^3/\mathrm{\Delta }R^{}`$. The typical electron energy is given by balancing $`\dot{E}_{heat}^{}`$ and $`\dot{E}_{cool}^{}=(4/3)n_e^{}\sigma _TcU_r^{}\gamma ^2\beta ^2(1+U_B^{}/U_r^{})`$, where $`U_B^{}/U_r^{}`$ is the magnetic to radiation energy density ratio and where $`n_e^{}`$ includes a possible contribution from $`e^\pm `$ pairs:
$$\gamma ^2\beta ^2\frac{3\pi R^2n_p^{}m_pc^3}{\mathrm{\Delta }R^{}\sigma _Tn_e^{}L^{}(1+U_B^{}/U_r^{})}\frac{n_p^{}m_p}{n_e^{}m_e}\frac{1}{1+U_B^{}/U_r^{}}\frac{3\pi }{\mathrm{}^{}}$$
(2)
where $`\mathrm{}^{}[L^{}\sigma _\mathrm{T}/(Rm_\mathrm{e}c^3)](\mathrm{\Delta }R^{}/R)`$ is the compactness parameter of the region emitting a (comoving) luminosity $`L^{}`$. Electron–positron pairs can be important for values of the compactness greater than unity (Svensson 1982, 1984, 1987), and can even dominate the particle number density. A typical value in this situation is $`\mathrm{}^{}=270(L_{46}^{}/R_{13})(\mathrm{\Delta }R_{11}^{}/R_{13})`$, and therefore equation (2) yields typical electron (and positron) energies at most moderately relativistic. As detailed below, the small energy of the emitting particles implies:
1) the synchrotron emission is self–absorbed.
2) the main radiation mechanism is multiple Compton scattering and the self–absorbed synchrotron emission is the source of soft seed photons.
Even though the particle distribution may not have time to thermalize, it will be characterized by a mean energy, and possibly be peaked at this value. It is then convenient to introduce an ’effective temperature’ $`\mathrm{\Theta }^{}kT^{}/(m_ec^2)`$.
The synchrotron luminosity can then be estimated assuming that the spectrum is described by the Rayleigh-Jeans part of a blackbody spectrum, up to the self–absorption frequency $`\nu _T^{}`$:
$$L_s^{}\frac{8\pi }{3}m_eR^2\mathrm{\Theta }^{}(\nu _T^{})^37.6\times 10^{41}\mathrm{\Theta }^{}R_{13}^2(\nu _{T,14}^{})^3\mathrm{erg}\mathrm{s}^1$$
(3)
where $`\nu _T^{}`$ can be derived again approximating the particle distribution as a Maxwellian of temperature $`\mathrm{\Theta }^{}`$. An approximate prescription, which holds for $`0.1<\mathrm{\Theta }^{}<3`$, has been derived by interpolating the analytic approximations to the cyclo–synchrotron emission reported by Mahadevan et al. (1996). This gives, for $`B_51`$ and $`\tau _T1`$, $`\nu _T^{}2.75\times 10^{14}(\mathrm{\Theta }^{})^{1.191}\mathrm{Hz}.`$
A generalized Comptonization parameter $`y`$, which is approximately valid also in the trans–relativistic and quasi–transparent conditions, can be defined as $`y\mathrm{\hspace{0.17em}4}\tau _T\mathrm{\Theta }^{}(1+\tau _T)(1+4\mathrm{\Theta }^{}).`$
The ratio of the Compton to the synchrotron powers can then be approximated by $`e^y`$, and thus in order to emit a Compton comoving luminosity $`L_c^{}=10^{46}L_{c,46}^{}`$ erg s<sup>-1</sup>, the $`y`$ parameter must be of the order of $`y=\mathrm{ln}(L_c^{}/L_s^{})11.5\mathrm{ln}(L_{c,46}^{}/L_{s,41}^{})`$. With this value of $`y`$ and $`\tau _T`$ of order unity, the Comptonized high energy spectrum has a $`F(\nu )\nu ^0`$ shape, while the relatively modest optical depth prevents a strong Wien peak to form (Pozdnyakov, Sobol & Sunyaev 1983). Therefore, very schematically, the resulting observed spectrum would extend between the energies $`h\nu _T^{}\mathrm{\Gamma }/(1+z)`$ and $`2kT^{}\mathrm{\Gamma }/(1+z)`$, with $`F(\nu )`$ const.
The spectrum emitted by a single shell will evolve very rapidly: after the observed acceleration time $`\mathrm{\Delta }R^{}(1+z)/(\mathrm{\Gamma }c)`$, particles cool on a similar timescale, while the Comptonization spectrum steepens and the emitted power decreases. (Eventually, the same particles can be re–heated by a collision with another shell.) The time integrated emission (even for a short exposure time of – say – a second) will result from all the cooling and re–heating histories of many shell–shell interactions. Any Wien hump and/or feature in the spectrum of individual shells will be smoothed out. The hard power–law continuum, if typical of all shells, would instead be preserved even when integrating over the exposure time.
### 3.1 The role of electron–positron pairs
As anticipated, $`e^\pm `$ pair production can play a crucial role: this process would surely be efficient for intrinsic compactnesses $`\mathrm{}^{}>1`$, and would on one hand increase the optical depth, and on the other act as a thermostat, by maintaining the temperature in a narrow range. For the temperatures of interest here, photon–photon collisions are the main pair production process. Note that our definition of $`\mathrm{}^{}`$ corresponds to the optical depth for $`\gamma `$$`\gamma e^\pm `$ within the shell width. Additional pairs will be produced outside the shell region, increasing the lepton content of the surrounding medium. Detailed time dependent studies of the optical depth and temperature evolution for a rapidly varying source have not yet been pursued. Results concerning a steady source in pair equilibrium indicate that for $`\mathrm{}^{}`$ between 10 and $`10^3`$ the maximum equilibrium temperature is of the order of 30–300 keV (Svensson 1982, 1984), if the source is pair dominated (i.e. the density of pairs outnumbers the density of protons). Indeed we expect in this situation to be close to pair equilibrium, as this would be reached in about a dynamical timescale. Note that the quoted numbers refer to a perfect Maxwellian particle distribution. However, pairs can be created even if the temperature is sub–relativistic by the photons and particles in the high energy tails of the real distribution: for a particle density decreasing slower than a Maxwellian one, more photons are created above the threshold for photon–photon pair production, and thus pairs become important for values of $`\mathrm{\Theta }^{}`$ lower than in the completely thermal case.
We conclude that an ‘effective’ temperature of $`kT^{}`$ 50 keV ($`\mathrm{\Theta }^{}0.1`$) and $`\tau _T4`$ dominated by pairs, can be a consistent solution giving $`y11`$. We stress here that the assumption of a soft seed photon distribution of luminosity $`L^{}10^{41}`$ erg s<sup>-1</sup> implies that $`any`$ self–consistent solution $`must`$ give $`y10`$, in order to produce a Compton luminosity matching the observed one.
## 4 Some consequences
If the high energy spectrum is due to quasi–thermal Comptonization, it will be sensitive to the amount of available soft seed photons. Assume that the complex light curve of GRB can be explained by the internal shock scenario, in which the emission is produced by the collisions of many shells. The first colliding shells will give rise to a certain synchrotron self–absorbed radiation, while subsequent shells, besides producing synchrotron radiation, will be illuminated by photons coming from the previous shell–shell collisions. The amount of seed photons is then bound to increase, in turn increasing the cooling rate of the electrons and positrons, which will therefore reach a lower temperature and produce a softer spectrum. This can qualitatively explain the hard–to–soft behavior of GRB emission and the (weak) correlation between duration and hardness (shorter bursts have harder spectra, Fishman & Meegan, 1995 and references therein). More quantitative details demand the knowledge of the exact time dependent feedback introduced by pair emission, which is difficult to asses.
If the intrinsic effective temperature is of the order of 50 keV, then the observed Comptonized spectrum extends to $`10\mathrm{\Theta }_1^{}\mathrm{\Gamma }_2/(1+z)`$ MeV. Note that there are no severe constraints on such high values of the energy at which the $`\nu F(\nu )`$ spectrum of the burst peaks (Cohen, Piran & Narayan 1998), and these values maybe reached during the very first parts of the burst light curve (e.g. first second). Time integrated spectra may instead be well fitted as emission from particles of lower temperatures.
One would also expect that (again especially during the initial first phase) a Wien peak at the electron temperature would be formed for bursts with $`\tau _T`$ larger than 3–5. This values of the optical depth would be easily reached for particularly strong bursts, where more copious pair production might occur.
In Comptonization models, photons of higher energies undergo more scatterings: if variability is caused by a change in the seed photon population, this may cause the high energy flux to lag that at softer frequencies. In this case the relevant timescale is the average time between two scatterings, which in the observer frame is of the order of $`\mathrm{\Delta }R^{}(1+z)/(\mathrm{\Gamma }\tau _Tc)0.03\mathrm{\Delta }R_{11}^{}(1+z)/(\tau _T\mathrm{\Gamma }_2)`$ s, which is within the possibility of current detectors. However there can be no lag if variability is caused by a sudden increase (decrease) of the number of emitting electrons or of their effective temperature. Constraints on the Comptonization scenario may come from detailed studies on how to reproduce very fast variability, since multiple scatterings will tend to smooth out any very short change.
If the progenitors of GRB are hypernovae (Paczyński 1998) the density in the vicinity of the central source is dominated by the pre–hypernova wind. This can lead to optical depths around unity at distances $`R10^{12}`$$`10^{13}`$, just where shell–shell collisions are assumed to take place. There is then the possibility that the GRB events are due to shocks with this material, rather than shocks between the shells. The implications of this scenario for the emission models are very interesting and will be discussed elsewhere (Ghisellini et al., in prep.). Here we would like to stress that:
i) in the case of shocks between shells and the pre–hypernova wind the large densities involved suggest that inverse Compton emission is favored with respect to the synchrotron process;
ii) if the (still unshocked) interstellar material has total optical depth $`\tau _T`$ around or greater than unity, photons will be down–scattered, introducing a break in the emergent spectrum at the observed energy $`511/[\tau _T^2(1+z)]`$ keV (Guilbert, Fabian & Rees 1983; Pozdnyakov, Sobol & Sunyaev 1983). Furthermore, the interstellar matter will act as a ‘mirror’, sending back the scattered photons, thus increasing the amount of Compton cooling in the emitting region (see Ghisellini & Madau 1996 for an application of this ‘mirror’ model to blazars).
We then conclude that the high density environment of the hypernova poses problems to the non–thermal ‘equipartition’ scenario and favors quasi–thermal Comptonization as the main radiation process.
Afterglow emission would start at the deceleration radius $`R_d10^{16}[E_{50}/(n\mathrm{\Gamma }_2^2)]^{1/3}`$ cm, after $``$40 $`(E_{50}/n)^{1/3}/\mathrm{\Gamma }_2^{8/3}`$ s from the start of the burst (e.g. Wijers, Rees & Mèszàros 1997), where $`n`$ is the number density of the interstellar material. At $`R_d`$, the reduced densities of particles and photons diminish the importance of the Compton emission even for ultrarelativistic electron energies, but the requirement $`\tau _e\gamma ^2<1`$ can still limit $`\gamma `$, especially in the case of dense, star forming, environments.
## 5 Discussion
In the equipartition scenario, the rough equality between the energy density of protons, leptons and magnetic field leads to a remarkable good agreement with the observed characteristics of the spectra. In order for this to be achieved the acceleration of electrons has to be impulsive, take place in a very limited volume of the interacting shell, and e<sup>±</sup> pair density has to be small enough not to significantly lower the mean lepton energy.
At the other extreme, when the particle acceleration occupies the entire shell volume and/or lasts for a shell light crossing time, the mean lepton energy is controlled by the balance between the heating and the cooling rate. This leads to mildly or sub relativistic lepton energies. Therefore the synchrotron emission is inhibited by self–absorption, and provides soft photons for the dominant inverse Compton scattering process. The ratio of the observed multiple Compton scattering and the self–absorbed synchrotron luminosities is of the order of $`10^5`$, and determines the required Comptonization $`y`$ parameter, i.e. the product of the particle optical depth and temperature. As long as the compactness of the emission region is greater than unity, relativistic temperatures cannot be achieved, because in this case electron–positrons are copiously produced, increasing the optical depth and decreasing the temperature. If, on the other hand, the temperature is low and the optical depth is large, photons are trapped inside the shell, and part of the radiation luminosity is used to expand it. As a result, there is a narrow range of optical depths and temperatures which accounts for the observed spectra. This may be why GRB preferentially emit at $``$1 MeV.
Detailed analysis are needed to determine the exact shape and the time behavior of the predicted spectra: particularly relevant, in this respect, will be to study the time evolution of hot compact sources, relaxing the assumption of pair balance.
We conclude that there are at least two possible regimes yielding the observed spectrum and peak frequency of GRB, depending on the nature of the dissipation/acceleration mechanism. There is even the possibility that they coexist: one can in fact imagine that an initially planar (and ’narrow’) shocked region can be soon subjected to instabilities thus starting dissipating energy over a larger volume. In this situation both the non–thermal ‘equipartition’ and the quasi–thermal ‘heating/cooling balance’ regimes would be at work.
We thank Martin Rees for invaluable discussions and comments on the manuscript. AC acknowledges the Italian MURST for financial support.
|
no-problem/9812/cond-mat9812216.html
|
ar5iv
|
text
|
# Charged impurity scattering limited low temperature resistivity of low density silicon inversion layers
\[
## Abstract
We calculate within the Boltzmann equation approach the charged impurity scattering limited low temperature electronic resistivity of low density $`n`$-type inversion layers in Si MOSFET structures. We find a rather sharp quantum to classical crossover in the transport behavior in the $`05`$K temperature range, with the low density, low temperature mobility showing a strikingly strong non-monotonic temperature dependence, which may qualitatively explain the recently observed anomalously strong temperature dependent resistivity in low-density, high-mobility MOSFETs.
PACS Number : 73.40.-c, 73.40.Qv, 73.50.Bk, 71.30.+h
\] Several recent publications on low temperature resistivity measurements in various low density two dimensional (2D) systems report the observation of an anomalously strong temperature dependence as a function of carrier density, which has been interpreted as evidence for a zero temperature two dimensional metal-insulator transition (2D M-I-T), which is considered to be forbidden in two dimensions (at least for a non-interacting 2D system) by the one parameter scaling theory of localization . A number of theoretical papers have appeared in the literature providing many possible resolutions of this seemingly unanticipated (but apparently ubiquitous) phenomenon. In this Letter we propose a possible theoretical explanation for (at least a part of ) the observed phenomena. Our explanation is quantitative, microscopic, and physically motivated. Although our theory is quite general and generic (and thus applicable to all the systems exhibiting the so-called 2D M-I-T), we specifically consider here the electron inversion layer in Si MOSFETs, which is both the original system in which the 2D M-I-T was first reported and is also the most exhaustively experimentally studied system in this context. It is important to emphasize that, in contrast to much of the existing theoretical work on the subject, our theory does not address the existence (or not) of a zero temperature 2D M-I-T, but specifically addresses the issue of quantitatively understanding the strikingly unusual finite temperature experimental results on the effective “metallic” side of the transition.
We first summarize the key experimental features of the 2D M-I-T phenomenon (focusing on Si MOSFETs), emphasizing the specific aspects addressed in our theory. Experimentally one finds a “critical density” ($`n_c`$) separating an effective “metallic” behavior (for density $`n_s>n_c`$) from an effective “insulating” behavior ($`n_s<n_c`$). We concentrate entirely on the effective “metallic” behavior in this Letter since a 2D metal is “unusual” according to the conventional theory and a 2D insulator is not. The experimental insulating behavior (for $`n_s<n_c`$) is quite conventional for a strongly localized semiconductor and can be understood using standard transport models . The effective “metallic” behavior is characterized by a strong drop in the temperature dependent resistivity, $`\rho (T)`$, at low temperatures ($`0.1KT13K`$) and at low densities ($`n_sn_c`$). This novel and dramatically strong temperature dependence of $`\rho (T)`$, where $`\rho (T)`$ may drop by a factor of $`210`$ at low electron densities as temperature decreases from 2K to 100 mK, is one of the most significant experimental observations we qualitatively explain in this Letter. In addition the experimental resistivity, $`\rho (T,n_s)`$, as a function of temperature and electron density shows an approximate “scaling” behavior $`\rho (T,n_s)\rho (T/T_0)`$ with $`T_0T_0(n_s)`$ indicating consistency with quantum criticality. Our theoretical results show the same “scaling” behavior with our calculated $`T_0(n_s)`$ having very similar density dependence as the experimental observation. There are interesting aspects of the magnetic field and the electric field dependence of the observed resistivity, which we do not address here, concentrating entirely on the behavior of $`\rho (T,n_s)`$ in the $`n_sn_c`$ “metallic” regime. It is this “anomalous metallic” behavior (in the sense of a very strong metallic temperature dependence of the resistivity in a narrow density range above $`n_c`$) which has created the recent interest in the 2D M-I-T phenomena since in general, the temperature dependent resistivity of a metal should saturate as it enters the low temperature Bloch-Grüneisen regime without manifesting any strong temperature dependence.
Our theory, which provides good qualitative agreement with the existing experimental data on the metallic ($`n_s>n_c`$) side of the transition, is based on two essential assumptions: (1) transport is dominated by charged impurity scattering centers (with a density of $`N_i`$ per unit area) which are randomly distributed at the interface; (2) the M-I-T at $`n_s=n_c`$ is characterized by a “freeze-out” of free carriers due to impurity binding — the free carrier density responsible for “metallic” transport is thus ($`n_sn_c`$) for $`n_s>n_c`$, and on the insulating side, $`n_s<n_c`$, the free carrier density (at $`T=0`$) is by definition zero. Some justifications for these assumptions have been provided in ref. although our current model transcends the specific scenario envisioned in ref. and is more general. In contrast to ref. , we do not specify any particular mechanism for the carrier freeze-out and accept it as an experimental fact. We note that we could extend our model and go beyond the above two assumptions, for example, by making the effective free carrier density $`n=(n_sn_c)\theta (n_sn_c)+n_a(T)`$, where $`n_a(T)`$ is a thermally activated contribution to the carrier density (this relaxes the second assumption), and/or by introducing additional scattering mechanisms such as the short-range surface roughness scattering (this relaxes the first assumption). These extensions beyond our two essential approximations will undoubtedly produce better quantitative agreement between our theory and experiment (at the price of having unknown adjustable parameters). We, however, refrain from such a generalized theory, because we believe that the minimal theory, constrained by our two stringent assumptions and thus allowing for only one unknown parameter (the charged impurity density $`N_i`$) which sets the overall scale of resistivity in the system, already catches much of the essential physics in the problem.
We use the finite temperature Drude-Boltzmann theory to calculate the ohmic resistivity of the inversion layer electrons taking only into account long range Coulombic scattering by the random static charged impurity centers with the electron-impurity Coulomb interaction being screened by the 2D electron gas in the random phase approximation (RPA). The resistivity is given by $`\rho =\sigma ^1`$, where the conductivity $`\sigma =ne^2<\tau >/m`$ with $`m`$ as the carrier effective mass, and $`<\tau >`$ is the energy averaged finite temperature scattering time:
$$<\tau >=\frac{𝑑EE\tau (E)\left(\frac{f}{E}\right)}{𝑑EE\left(\frac{f}{E}\right)},$$
(1)
where $`f(E)`$ is the Fermi distribution function, $`f(E)=\{1+\mathrm{exp}[(E\mu )]/k_BT\}^1`$ with $`\mu (T,n)`$ as the finite temperature chemical potential of the free carrier system determined self-consistently. The energy dependent scattering time $`\tau (E)`$ for our model of randomly distributed interfacial impurity charge centers is given by
$$\frac{1}{\tau (E)}=\frac{2\pi N_i}{\mathrm{}}\frac{d^2k^{}}{(2\pi )^2}\left|\frac{v(q)}{\epsilon (q)}\right|^2(1\mathrm{cos}\theta )\delta \left(ϵ_𝐤ϵ_𝐤^{}\right),$$
(2)
with $`q=|𝐤𝐤^{}|`$, $`\theta \theta _{\mathrm{𝐤𝐤}^{}}`$ is the scattering angle between $`𝐤`$ and $`𝐤^{}`$, $`E=ϵ_𝐤=\mathrm{}^2k^2/2m`$, $`ϵ_𝐤^{}=\mathrm{}^2k^2/2m`$, $`v(q)`$ is the 2D Coulomb interaction between an electron and an impurity, and $`\epsilon (q)\epsilon (q;\mu ,T)`$ is the 2D finite temperature static RPA dielectric (screening) function . In calculating the Coulomb interaction and the RPA dielectric function in Eq. (2) we take into account subband quantization effects in the inversion layer through the lowest subband variational wavefunction . We note that there are two independent sources of temperature dependence in our calculated resistivity — one source is the energy averaging defined in Eq. (1) and the other is the explicit temperature dependence of the finite temperature dielectric function $`\epsilon (q)`$ which produces a direct temperature dependence through screening in Eq. (2). At very high temperatures, when $`TT_F`$ with $`T_F\mu (T=0)/k_B`$ as the free carrier Fermi temperature, the system is classical and it is easy to show that Eq. (1) leads to a decreasing resistivity with increasing temperature: $`\rho (T)A(T/T_F)^1`$ for $`TT_F`$. In the quantum regime, however, energy averaging by itself produces a weak quadratic (negative) temperature dependence according to Eq. (1): $`\rho \rho (T=0)B(T/T_F)^2`$, for $`TT_F`$. For Si inversion layer, however, this low temperature negative temperature dependence is overwhelmed by the temperature dependence of the screening function in Eq. (2) which dominates $`2k_F`$ scattering — this phenomenon arises from the specific form of the 2D screening function which is a constant upto $`q=2k_F`$, and has a cusp at $`2k_F`$ at $`T=0`$. This strong temperature dependence arising from the low temperature screening function produces a linear rise in the low temperature ($`TT_F`$) resistivity with increasing temperature according to Eq. (2): $`\rho (T)\rho (T=0)+C(T/T_F)`$, for $`TT_F`$. This linear temperature dependence is, however, cut off at very low temperatures due to the rounding of the sharp corner in the 2D screening function by impurity scattering effects — at very low temperature $`TT_D`$ where $`T_D`$ ($`=\mathrm{\Gamma }/\pi k_B`$ with $`\mathrm{\Gamma }`$ as the collisional broadening) is the collisional broadening induced Dingle temperature, the explicit temperature dependence of $`\epsilon (q,T)`$ is suppressed. At the densities and temperatures of interest in the 2D M-I-T phenomena all of these distinct physical effects are operational, and the actual behavior of $`\rho (T,n)`$ could be quite complicated because the four different asymptotic mechanisms discussed above compete with each other as the system crosses over from a non-degenerate classical ($`T>T_F`$) to a strongly screened degenerate quantum ($`T<<T_F`$) regime. We note that in general the temperature dependence is non-monotonic (particularly at lower densities where the energy averaging effects are significant), as has been experimentally observed , because the temperature dependence of Eq. (1) by itself produces a negative temperature coefficient whereas screening through Eq. (2) produces a positive temperature coefficient.
In Fig. 1 we show our numerically calculated resistivity for the Si-15 sample of ref. . We show calculated $`\rho (T)`$ as a function of $`T`$ in Fig. 1 for several values of $`n_s>n_c`$. We use several different Dingle temperatures to incorporate the impurity scattering induced
collisional broadening corrections in the screening function, including the pure RPA ($`T_D=0`$) case which completely neglects collisional broadening effects on screening. In Fig. 1(b) we show the calculated Si-15 results where the Dingle temperature varies as a function of electron density. For each density the appropriate $`T_D`$ (going into the screening calculation) is determined from the resistivity for that particular density. In Fig. 1(b) the temperature dependence of $`\rho (T)`$ at low temperatures is strongest at intermediate densities somewhat away from $`n_c`$ whereas in Fig. 1(a) the temperature dependence of $`\rho (T)`$ becomes stronger as one approaches $`n_c`$, and is the strongest at the lowest density. This arises from the competition in screening among $`T`$, $`T_F`$, and $`T_D`$ — at the lowest densities the temperature dependence is moderated by having relatively high values of $`T_D`$ whereas at high densities the temperature dependence is suppressed by the large value of $`T_F`$, implying that the strongest temperature dependence of $`\rho (T)`$ occurs at intermediate densities where neither $`T_D`$ nor $`T_F`$ is too high. Putting $`T_D=0`$ leads to stronger temperature dependence because the temperature dependence of screening is not cut off at low temperatures as it is in the $`T<T_D<T_F`$ regime for the $`T_D0`$ results.
The impurity density $`N_i`$ has been fixed by demanding agreement between experiment and theory at high temperatures ($`T=5K`$) and the highest densities. The impurity density $`N_i`$ thus sets the scale of the overall resistivity ($`\rho N_i`$), and does not affect the calculated $`T`$ and $`n_s`$ dependence of $`\rho (T,n_s)`$. It is important to emphasize that $`N_i`$ values needed in our calculation to obtain quantitative agreement with experiment are in the reasonable range of $`N_i10^{10}cm^2`$, which is known to be the typical effective random charged impurity scattering center density in high mobility Si MOSFETs. Since $`N_i`$ is the only “free” parameter of our theory, it is significant that we obtain a reasonable value for $`N_i`$ in order to achieve agreement between theory and experiment. We emphasize that our theory is valid even if the metallic ($`n_s>n_c`$) system is actually weakly localized as long as the effective localization length is larger than the system size or the phase coherence length. In general, our calculated resistivity is higher (by $`2540\%`$) than the experimental values at low densities ($`n_sn_c`$), i.e., our theory predicts a somewhat stronger $`n_s`$ dependence of $`\rho `$ than that observed experimentally. This discrepancy can be corrected by adding an activated carrier density $`n_a(T)`$ to our effective carrier density $`n=(n_sn_c)+n_a`$, which produces the strongest effect at the lowest densities (and essentially no effect at higher densities), and would reduce $`\rho (T)`$ at lower densities. One can also use a variable impurity density $`N_i(n_s)`$ which varies with the gate voltage (following the spirit of ref. ), and is lower at lower values of $`n_s`$, again producing quantitative agreement between theory and experiment. Given the overall excellent qualitative agreement between our results and the experimental data of ref. , we think that these refinements of our model are not particularly essential or meaningful.
We show the experimental data points for Si-15 taken from ref. in Fig. 1(a) to give an idea about the level of agreement between our calculation and the experimental results. We do not attach particularly great significance to the quantitative agreement achieved in Fig. 1(a) because of the various approximations in our theory. We do emphasize, however, that our calculations catch all the essential qualitative features of the low temperature experimental data . We obtain the observed non-monotonicity in $`\rho (T)`$ at low densities and also the strong drop in $`\rho (T)`$ at low densities in the $`0.12K`$ temperature range. Consistent with the experimental observations our calculated low density $`\rho (T)`$ could drop by an order of magnitude for $`12K`$ change in the temperature. Our high density results show weak monotonic increasing $`\rho (T)`$ with increasing $`T`$ similar to experimental observations . We have carried out calculations for all the reported Si samples (as well as GaAs samples) in the literature , and our level of qualitative agreement with experiment in uniformly good (typically as good as it is in Fig. 1) for all the existing experiments. For lower mobility samples, our calculated temperature dependence is
weaker (consistent with experimental findings) because screening is suppressed by stronger impurity scattering effects (through a higher value of level broadening or Dingle temperature).
In Fig. 2 we show our calculated “scaling” properties of $`\rho (T,n_s)\rho (T/T_0)`$ with $`T_0T_0(n_s)`$ for the Si-15 results shown in Fig. 1(b). The scaling is obtained entirely numerically by obtaining the $`T_0`$ which gives the best scaling fit to the calculated $`\rho (T,n_s)`$. Comparing Fig.2 with the corresponding experimental scaling plots of resistivity we conclude that our theoretical scaling behavior of $`\rho (T,n_s)`$ is about as good as the corresponding experimental scaling. In particular, our $`T_0(n_s)`$, shown as an inset in Fig. 2, agrees reasonably well with the experimental results . We obtain very similar “scaling” results for the other Si samples in refs. . The “scaling” we obtain in Fig. 2 underscores the important point that the experimentally observed scaling behavior in a narrow ($`T,n_s`$) range does not necessarily imply quantum criticality.
Before concluding we point out the approximations made in our calculations. We have assumed uncritically that the Drude-Boltzmann transport theory, which is extensively and successfully employed in the device simulation of Si MOSFETs, applies to the problem being studied. Our main justification for applying the standard transport theory to the current problem is our belief that such a “zeroth order”, “one-parameter” ($`N_i`$ being the only parameter in our model) theory must be applied to the problem and compared with the experimental data before one can discuss more speculative (and calculationally difficult) approaches. The fact that such a zeroth order theory already obtains good qualitative agreement with the experimental results indicates that charged impurity scattering, carrier binding and freeze-out, temperature and density dependence of 2D screening, and classical to quantum crossover (in the $`T=05K`$ range) are playing significant roles in the experiments and cannot be neglected in any theoretical analysis of the “2D M-I-T” phenomenon. Our other approximations of using the RPA screening (we actually incorporate a 2D Hubbard local field correction in our screening, which has no qualitative effect on our results) and the Dingle temperature approximation to incorporate collisional broadening effects on screening are quite reasonable (at least qualitatively) within our model and approximation scheme, and may be systematically improved (with a great deal of work) if future experiments warrant such a quantitative improvement of the theory. It is important to emphasize that quantum corrections, including localization effects, are left out of our semi-classical Drude-Boltzmann theory. In providing some justification we mention that the effective dimensionless parameter $`k_Fl`$ (where $`l`$ is the transport mean free path) is typically 2 or larger in our results (at $`T=0`$, which we believe to be the appropriate limit to consider), and therefore a Boltzmann theory may have reasonable validity. We estimate weak localization effects to be substantially weaker than the temperature dependence shown in Fig. 1 in the experimental temperature range ($`T>50mK`$) in refs. . The fact that the observed temperature dependence, particularly at lower temperatures, is somewhat stronger than our calculated results may very well be the manifestation of quantum fluctuation or interaction effects neglected in our theory. An important approximation of our theory (consistent with the Drude-Boltzmann approach) is the neglect of inelastic electron-electron interaction, which may well be significant in the low density 2D systems of experimental relevance. For example, it is possible that the insulating system ($`n_s<n_c`$) is an electron glass (arising from the competition/frustration between interaction and disorder). While a quantitative theory including disorder and interaction effects is extremely difficult, we speculate that our Boltzmann theory (in particular the quantum-classical crossover which leads to the strong temperature dependence) is sufficiently robust so that our qualitative conclusions will remain unaffected.
We conclude by emphasizing the specific salient features of the qualitative agreement between our theory and experimental data on the metallic side: (1) strong temperature dependence at low and intermediate densities ($`n_sn_c`$); (2) non-monotonicity in $`\rho (T)`$, arising from quantum-classical crossover, at low values of $`n_sn_c`$ where $`\rho (T)`$ increases weakly with decreasing $`T`$ at higher temperatures and decreases strongly with $`T`$ at lower temperatures; (3) scaling of $`\rho (T,n_s)\rho (T/T_0)`$ with the theoretical $`T_0(n_s)`$ agreeing with the experimental results; (4) our calculated zero temperature conductivity, $`\sigma (T=0,n_s)=1/\rho (T0,n_s)`$, shows an approximately (within $`25\%`$) linear density dependence, $`\sigma (T=0)n=(n_sn_c)`$, which is consistent with the $`T0`$ extrapolation of the experimental resistivity and also with several other experimental findings \[this dependence, $`\sigma (T=0)(n_sn_c)`$, also supports our basic freeze-out or binding model\]; (5) for increasing disorder (i.e., for lower mobility samples) we predict an increasing $`n_c`$ and weaker temperature dependence, as observed experimentally.
This work is supported by the U.S.-ARO and the U.S.-ONR.
|
no-problem/9812/hep-ph9812243.html
|
ar5iv
|
text
|
# JLAB-THY-98-48 Interpreting the Neutron’s Electric Form Factor: Rest Frame Charge Distribution or Foldy Term?
## I Introduction
In 1962, Sachs showed that the combinations of elastic nucleon form factors ($`N=p`$ or $`n`$)
$$G_E^N=F_1^N\frac{Q^2}{4m_N^2}F_2^N$$
(1)
$$G_M^N=F_1^N+F_2^N$$
(2)
have simple interpretations as the spatial Fourier transforms of the nucleons’ charge and magnetization distributions in the Breit frame (where momentum $`\stackrel{}{p}=\frac{\stackrel{}{Q}}{2}`$ is scattered to momentum $`\stackrel{}{p}^{}=+\frac{\stackrel{}{Q}}{2}`$). Here $`F_1^N`$ and $`F_2^N`$ are the Dirac and Pauli form factors, respectively, defined by
$$N(\stackrel{}{p}^{},s^{})|j_{em}^\mu (0)|N(\stackrel{}{p},s)=\overline{u}(\stackrel{}{p}^{},s^{})[F_1^N\gamma ^\mu +i\frac{\sigma ^{\mu \nu }q_\nu }{2m_N}F_2^N]u(\stackrel{}{p},s)$$
(3)
where $`q_\nu =p_\nu ^{}p_\nu `$ and $`F_1^N`$ and $`F_2^N`$ are functions of $`Q^2=q^2`$.
These form factors obviously contain vital information on the internal composition of the nucleons. Although it has proven elusive experimentally, the electric form factor of the neutron $`G_E^n`$ is particularly fascinating in this respect. In pion-nucleon theory, $`G_E^n`$ would arise from a $`\pi ^{}`$ cloud with convection currents producing the anomalous magnetic moments $`F_2^p=1.79\mu _p1`$ and $`F_2^n=1.91\mu _n`$. In contrast, in a valence quark model, the nucleon magnetic moments arise from the underlying charged spin-$`\frac{1}{2}`$ constituents with the famous $`SU(6)`$ relation
$$\frac{\mu _p}{\mu _n}=\frac{3}{2}$$
(4)
and with a scale set by
$$\mu _p=\frac{m_N}{m_d}3$$
(5)
where $`m_dm_u\frac{1}{3}m_N`$ is a valence quark effective mass. Within this model it has been argued that in the $`SU(6)`$ limit $`G_E^n(Q^2)`$ would be identically zero, but that the spin-spin forces which produce the $`SU(6)`$-breaking $`\mathrm{\Delta }N`$ splitting create a charge segregation inside the neutron and lead to a nonzero $`G_E^n`$ . The effect arises because the spin-spin forces push $`d`$ quarks to the periphery of the neutron and pull the $`u`$ quark to the center. Thus both the $`\pi ^{}`$ cloud picture and the hyperfine-perturbed quark model predict a negative neutron charge radius, as observed.
Nonrelativistically, the squared charge radius is simply the charge-weighted mean square position of the constituents. More generally
$$G_E^p(Q^2)1\frac{1}{6}r_{Ep}^2Q^2+\mathrm{}$$
(6)
and
$$G_E^n(Q^2)\frac{1}{6}r_{En}^2Q^2+\mathrm{}$$
(7)
define the proton and neutron charge radii, with
$$\frac{G_M^N(Q^2)}{\mu _N}=1\frac{1}{6}r_{MN}^2Q^2+\mathrm{}$$
(8)
defining the corresponding magnetic radii. Experimentally, the three form factors $`G_E^p`$, $`G_M^p`$, and $`G_M^n`$ are reasonably well known, although new measurements should soon determine them with much greater precision . From these measurements, which cover $`Q^2`$ in the multi-GeV<sup>2</sup> range, we have learned that all three form factors have roughly similar shapes:
$$G_E^p(Q^2)\frac{G_M^p(Q^2)}{\mu _p}\frac{G_M^n(Q^2)}{\mu _n}G_D(Q^2)$$
(9)
where the dipole form factor
$$G_D(Q^2)\frac{1}{[1+Q^2/M_{dipole}^2]^2}$$
(10)
with $`M_{dipole}^2=0.71`$ GeV<sup>2</sup>. I note in passing that this observation makes little sense in a picture where the nucleons have a point-like core surrounded by a pion cloud (since in such a picture $`G_M^n`$ is purely pionic while $`G_E^p`$ and $`G_M^p`$ are mixtures of a “bare” proton and a pion cloud). In contrast, Eq. (9) is very natural in a valence quark model where, to leading order in a relativistic expansion, $`G_D(Q^2)`$ would simply be the Fourier transform of the square of the ground state spatial wavefunction.
Recent advances in experimental technique should lead to a clear measurement of the neutron’s electric form factor $`G_E^n(Q^2)`$ for $`Q^2`$ in the GeV<sup>2</sup> range in the next few years . Its charge radius $`r_{En}^2`$ is known from low energy neutron-electron elastic scattering to be $`0.113\pm 0.005`$ fm<sup>2</sup> , but the electron-neutron scattering measurements needed to determine $`G_E^n`$ in the GeV<sup>2</sup> range (in order to roughly map out the neutron’s electric structure with a resolution of 10% of the proton’s size) have been plagued by the lack of a free neutron target . Fortunately, the method of Arnold, Carlson, and Gross , which uses spin observables sensitive to $`G_E^nG_M^n`$ interference, opens up new methods for measuring $`G_E^n`$, and recent advances in accelerator, target, and detector technology are beginning to exploit these new methods.
In anticipation of these measurements, there has been renewed discussion about their interpretation. I focus here on the belief that the measured $`r_{En}^2`$ is explained by the “Foldy term” . I.e., using Eqs. (1) and (7),
$$r_{En}^2=r_{1n}^2+\frac{3\mu _n}{2m_N^2}r_{1n}^2+r_{Foldy,n}^2,$$
(11)
where $`r_{1n}^2`$ is the “charge radius” associated with $`F_1^n\frac{1}{6}r_{1n}^2Q^2+\mathrm{}`$. The second term in Eq. (11), called the Foldy term, appears to arise as a relativistic correction associated with the neutron’s magnetic moment and so to have nothing to do with the neutron’s rest frame charge distribution. It has the value $`0.126`$ fm<sup>2</sup>, nearly coinciding with the measured value. On this basis it has been argued that any “true” charge distribution effect must be very small. In this paper I will show that while the Foldy term closely resembles $`r_{En}^2`$ numerically, it does not “explain it”. Indeed, I will demonstrate that, in the relativistic approximation to the constituent quark model in which the Foldy term first appears, it is cancelled exactly by a contribution to the Dirac form factor $`F_1`$ leaving $`r_{En}^2`$ correctly interpreted as arising entirely from the rest frame internal charge distribution of the neutron.
## II The Interpretation of the Neutron Charge Radius in a Constituent Quark Model
The relationship (1) between the Sachs form factor $`G_E`$ and the Dirac and Pauli form factors $`F_1`$ and $`F_2`$ is relativistic in origin. Unfortunately, relativistic constituent models of the nucleon are notoriously difficult: rest frame models are difficult to boost and infinite-momentum-frame (or light-cone) quark models have trouble constructing states of definite $`J^P`$. This could be the reason that the interpretation of $`G_E^n`$ has not been clarified in the context of such models.
While an accurate constituent quark model of nucleon structure must certainly be fully relativistic, the issue at hand can be resolved by using a relativistic expansion around the nonrelativistic limit. This is possible because the Foldy term $`r_{Foldy,n}^2`$ arises at order $`Q^2/m^2`$ and so its character may be exposed by an expansion of $`G_E^n`$ to order $`1/m^2`$. I will also exploit symmetries of the problem available in certain limits which will make the discussion independent of the details of models.
I begin with a simple “toy model” in which a “toy neutron” $`\stackrel{~}{n}_{\overline{S}D}`$ is composed of a scalar antiquark $`\overline{S}`$ of mass $`m_S`$ and charge $`e_D`$ and a spin-$`\frac{1}{2}`$ Dirac particle $`D`$ of mass $`m_D`$ and charge $`e_D`$ bound by flavor and momentum independent forces into a rest frame nonrelativistic $`S`$-wave. The calculation begins by noting that, from their definitions,
$$G_E^{\stackrel{~}{n}_{\overline{S}D}}(Q^2)=\stackrel{~}{n}_{\overline{S}D}(+\frac{Q\widehat{z}}{2},+)|\rho _{em}|\stackrel{~}{n}_{\overline{S}D}(\frac{Q\widehat{z}}{2},+)$$
(12)
and
$$G_M^{\stackrel{~}{n}_{\overline{S}D}}(Q^2)=\frac{m_{\stackrel{~}{n}_{\overline{S}D}}}{Q}\stackrel{~}{n}_{\overline{S}D}(+\frac{Q\widehat{z}}{2},+)|j_{em}^{1+i2}|\stackrel{~}{n}_{\overline{S}D}(\frac{Q\widehat{z}}{2},).$$
(13)
It is immediately clear that the calculation of these form factors requires boosting the rest frame $`S`$-wave bound state to momenta $`\pm \frac{Q\widehat{z}}{2}`$. Doing so can introduce a host of $`1/m^2`$ effects in the boosted counterpart of the $`S`$-wave state and it can also produce new $`P`$-wave-like components by Wigner-rotation of the $`D`$-quark spinors . I will show that the latter effect is subleading, and will deal with the former effect by exploiting an effective charge-conjugation symmetry of the system for $`m_D=m_Sm`$.
Since $`\mu _{\stackrel{~}{n}_{\overline{S}D}}`$ involves the limit of Eq. (13) as $`Q0`$, to the required order in $`1/m`$ it simply takes on its nonrelativistic value
$$\mu _{\stackrel{~}{n}_{\overline{S}D}}=\frac{e_Dm_{\stackrel{~}{n}_{\overline{S}D}}}{m_D}$$
(14)
where of course $`m_{\stackrel{~}{n}_{\overline{S}D}}=m_S+m_D`$ in this limit. The Foldy term is thus well-defined:
$$r_{Foldy,\stackrel{~}{n}_{\overline{S}D}}^2=\frac{e_D}{2m_Dm_{\stackrel{~}{n}_{\overline{S}D}}}.$$
(15)
We next compute $`G_E^{\stackrel{~}{n}_{\overline{S}D}}(Q^2)`$ directly from Eq. (12). To leading order in $`1/m^2`$, $`\rho _{em}`$ remains a one-body current, and the impulse approximation is valid. Within this approximation, we make use of the relation
$$D(\stackrel{}{p}+Q\widehat{z},s^{})|\rho _{em}|D(\stackrel{}{p},s)=e_D(1\frac{Q^2}{8m_D^2})\stackrel{~}{D}(\stackrel{}{p}+Q\widehat{z})|\rho _{em}|\stackrel{~}{D}(\stackrel{}{p})\delta _{ss^{}}+\rho _{spinflip}$$
(16)
where
$$\rho _{spinflip}\frac{e_DQ}{4m_D^2}(p_{}\delta _{s^{}+}\delta _sp_+\delta _s^{}\delta _{s+})$$
(17)
and where $`\stackrel{~}{D}`$ is a fictitious scalar quark with the mass and charge of $`D`$. This expression is easily obtained by making a nonrelativistic expansion of both the $`D`$ and $`\stackrel{~}{D}`$ charge density matrix elements.
The spin-flip term $`\rho _{spinflip}`$ can only contribute via transitions to and from the Wigner-rotated components of the wavefunction. However, the amplitudes of such components are proportional to $`Qk/m_D^2`$, where $`k`$ is an internal momentum. Since $`\rho _{spinflip}`$ already carries a factor $`1/m_D^2`$, such effects may be discarded. Note that non-flip Wigner-rotated contributions are of the same order and may also be neglected.
We conclude that $`r_{E\stackrel{~}{n}_{\overline{S}D}}^2`$ may be computed by replacing $`D`$ by $`\stackrel{~}{D}`$ provided the additional contribution $`e_DQ^2/8m_D^2`$ is added to $`G_E^{\stackrel{~}{n}_{\overline{S}D}}`$. (More precisely, this factor multiplies the $`\stackrel{~}{D}`$ contribution to $`G_E^{\stackrel{~}{n}_{\overline{S}D}}`$, but at $`Q^2=0`$ this is just unity.) I will denote the associated “zwitterbewegung” charge radius $`3e_D/4m_D^2`$ by $`r_{D,zwitter}^2`$. The effect of $`r_{D,zwitter}^2`$ is well-known in a variety of contexts, including atomic physics , nuclear physics , hadronic physics , and heavy quark physics ; at the most elementary and concrete level it appears as the additional factor of $`(1\frac{Q^2}{8m_D^2})^2`$ in the ratio of the Mott cross section to the Rutherford cross section. The problem of computing the remaining contributions to $`G_E^{\stackrel{~}{n}_{\overline{S}D}}`$ to this order from the fictitious $`\overline{S}\stackrel{~}{D}`$ scalar-scalar bound state would in general be highly nontrivial. However, in the limit $`m_D=m_Sm`$ these contributions vanish, since this system has in this limit a pseudo-charge-conjugation invariance under $`(\overline{S},\stackrel{~}{D})(S,\overline{\stackrel{~}{D}})`$. However, we note that in this limit $`m_{\stackrel{~}{n}_{\overline{S}D}}=2m`$ and so from Eq. (15) we have
$$r_{E\stackrel{~}{n}_{\overline{S}D}}^2=r_{D,zwitter}^2=\frac{3e_D}{4m^2}=r_{Foldy,\stackrel{~}{n}_{\overline{S}D}}^2,$$
(18)
i.e., in this model the “scalar charge distribution” is zero and the Foldy term would indeed account for the full charge radius of $`\stackrel{~}{n}`$. This conclusion is simply interpreted: the two scalar particles $`\overline{S}`$ and $`\stackrel{~}{D}`$ have perfectly overlapping and cancelling charge distributions, but the expansion of the $`\stackrel{~}{D}`$ distribution by $`r_{D,zwitter}^2`$ creates a slight excess of $`\stackrel{~}{D}`$ at large radii. In terms of its experimental significance, we have concluded that in an $`\overline{S}D`$ model of the neutron, the observation of an equality of $`r_{En}^2`$ and $`r_{Foldy,n}^2`$ would indeed indicate the absence of an intrinsic “scalar” charge distribution.
We shall soon be drawing quite another conclusion for the situation in the constituent quark model. However, before leaving the $`\overline{S}D`$ model, it is useful to consider another limit: the “hydrogenic limit” where $`m_S\mathrm{}`$. In this case (see Eq. (15)), $`r_{Foldy,\stackrel{~}{n}_{\overline{S}D}}^2=0`$ but
$$r_{E,\stackrel{~}{n}_{\overline{S}D}}^2=\frac{3e_D}{4m_D^2}+e_Dr_{wf}^2,$$
(19)
where $`r_{wf}^2`$ is the charge radius associated with the bound state $`\stackrel{~}{D}`$ problem. It may be that $`r_{wf}^2`$ contains other $`1/m_D^2`$ effects, but we note that the physics of the $`r_{D,zwitter}^2`$ effect previously associated with $`F_2`$ via $`r_{Foldy,\stackrel{~}{n}_{\overline{S}D}}^2`$ now must be asociated with $`F_1`$ via $`r_{1\stackrel{~}{n}_{\overline{S}D}}^2`$. The $`\overline{S}D`$ toy model thus simultaneously supplies us with a simple interpretation of the Foldy term and a warning about associating $`F_1`$ with the neutron’s “intrinsic charge distribution”.
While the $`\overline{S}D`$ toy model has some of the characteristics of diquark models for the nucleon ($`\overline{S}`$ has the quantum numbers and color of a scalar diquark), our main use for it was to introduce the basic elements of our discussion in a simple context. Indeed, it is now relatively trivial to extend our considerations to the realistic case of the valence quark model in which the neutron is in the leading approximation made of three mass $`m_q`$ spin-$`\frac{1}{2}`$ quarks $`ddu`$ bound by flavor and momentum independent forces into flavor-independent nonrelativistic relative $`S`$-waves. In this case
$$\mu _{\stackrel{~}{n}_{ddu}}=\frac{2m_{\stackrel{~}{n}_{ddu}}}{3m_q}2$$
(20)
so that
$$r_{Foldy,\stackrel{~}{n}_{ddu}}^2=\frac{1}{m_qm_{\stackrel{~}{n}_{ddu}}}.$$
(21)
In calculating $`r_{E\stackrel{~}{n}_{ddu}}^2`$ via Eq. (12), the transformation of the calculation of the charge radius of $`ddu`$ to that of three scalar quarks $`\stackrel{~}{d}\stackrel{~}{d}\stackrel{~}{u}`$ and residual $`r_{q,zwitter}^2=e_qQ^2/8m_q^2`$ terms proceeds as before, as does the neglect of Wigner-rotated components of the boosted state vectors. However, in this case, since the $`r_{q,zwitter}^2`$ terms are spin and flavor independent, and since the sum of the three charges is zero, they lead to no net $`Q^2/m^2`$ term! The reason for this is clear: the exactly overlapping and cancelling quark distributions remain exactly overlapping and cancelling after they are all equally smeared by $`r_{q,zwitter}^2`$ . This picture also anticipates the next stage of the argument: the analog of the pseudo-charge-conjugation invariance that we used for the scalar part of the $`\overline{S}D`$ matrix element is that the three quark wavefunction belongs to the symmetric representation of the permutation group $`S_3`$ so that the scalar part $`r_{E,\stackrel{~}{n}_{\stackrel{~}{d}\stackrel{~}{d}\stackrel{~}{u}}}^2`$ of the charge radius vanishes. Thus in the usual valence quark model
$$r_{E,\stackrel{~}{n}_{ddu}}^2=r_{E,\stackrel{~}{n}_{\stackrel{~}{d}\stackrel{~}{d}\stackrel{~}{u}}}^2+\mathrm{\Sigma }_ie_ir_{i,zwitter}^2=0,$$
(22)
which requires that
$$r_{1\stackrel{~}{n}_{ddu}}^2=r_{Foldy,\stackrel{~}{n}_{ddu}}^2$$
(23)
so it is appropriate to interpret the observed $`r_{En}^2`$ as due to an intrinsic internal charge distribution. Stated in another way, the coincidence of the predicted rest frame charge distribution in such models with the experimental value of $`r_{En}^2`$ may be claimed as a success, while the numerical coincidence of $`r_{En}^2`$ with the Foldy term may consistently be viewed as a potentially misleading accident. Such an accident is possible because while in the nonrelativistic limit $`r_{Foldy,\stackrel{~}{n}}^2<<r_{E\stackrel{~}{n}}^2`$, in QCD both constituent masses and hadronic radii are determined by $`\mathrm{\Lambda }_{QCD}`$ so they are expected to be of comparable magnitude.
## III Conclusions and Final Remarks
The principal conclusion of this paper is that, within the context of the valence quark model, the apparent contribution $`3\mu _n/2m_N^2`$ of the Foldy term to the charge radius of the neutron is illusory: in the leading approximation it is exactly cancelled by a “nonintuitive” contribution to the radius $`r_{1n}^2`$ of the Dirac form factor $`F_1^n`$. It is therefore totally appropriate to compare the measured $`r_{En}^2`$ and in general $`G_E^n(Q^2)`$ against quark model predictions (see, e.g., Fig. 1 of the second of Refs. ) for the rest frame internal charge distribution of the neutron.
Before too much is made of this successful prediction of the valence quark model, some other very fundamental questions must still be answered. Perhaps the most fundamental is the possible effect of nonvalence components in the neutron wavefunction. After all, the classic explanation for $`r_{En}^2`$ is that the neutron has a $`p\pi ^{}`$ component in its wavefunction (for a discussion in the more modern context of heavy baryon chiral perturbation theory, see Ref. ). Since both hyperfine interactions and $`q\overline{q}`$ pairs are $`1/N_c`$ effects, I know of no simple argument for why one should dominate.
Fortunately, there is both theoretical and experimental progress in resolving this old question. Recent theoretical work on “unquenching the quark model” indicates that there are strong cancellations between the hadronic components of the $`q\overline{q}`$ sea which tend to make it transparent to photons. These studies provide a natural way of understanding the successes of the valence quark model even though the $`q\overline{q}`$ sea is very strong, and in particular suggest that the precision of the OZI rule is the result of both a factor of $`1/N_c`$ and strong cancellations within this $`1/N_c`$-suppressed meson cloud. New data on the contributions of $`s\overline{s}`$ pairs to the charge and magnetization distribution of the nucleons is also beginning to constrain the importance of such effects and, by broken $`SU(3)`$, their $`u\overline{u}`$ and $`d\overline{d}`$ counterparts, and future experiments will either see $`s\overline{s}`$ effects or very tightly limit them (at the level of contributions of a few percent to $`r_E^2`$ and $`\mu _N`$). The resolution of the old question of the origin of $`\mu _n`$ and $`r_{En}^2`$ is thus within sight.
REFERENCES
|
no-problem/9812/cond-mat9812324.html
|
ar5iv
|
text
|
# 1 𝑞^(𝐿) in the MKA (lines without points) and from simulations of the 3𝑑 EA spin glass.
Comment on “Evidence for the Droplet/Scaling Picture of Spin Glasses”
In a recent letter Moore et al. claim to exhibit evidence for a non-mean-field behavior of the $`3d`$ Ising spin glass. We show here that their claim is insubstantial, and by analyzing in detail the behavior of the Migdal-Kadanoff approximation (MKA) as compared to the behavior of the Edwards-Anderson (EA) spin glass we find further evidence of a mean-field like behavior of the $`3d`$ spin glass.
The main point of does not concern the validity of the MKA in describing spin glasses, since it is well known, after the work of , that already at the mean field level the MKA describes a trivial droplet structure, completely missing the structure of the phase space of the model in any dimension.
Reference shows instead that the probability distribution of the order parameter $`P_{MK}(q)`$ computed in the MKA at $`T=0.7`$, close to the temperature where most of the numerical simulations have been run , has a spurious small $`q`$ “plateau”, very similar to the non-trivial $`P(q)`$ one finds numerically for the EA model. In these conditions, for values of the lattice size comparable to the ones used in numerical simulations, $`L16`$, the small $`q`$ region of $`P_{MK}(q)`$ does not seem to depend on $`L`$, even if one knows that eventually, for very large values of $`L`$, it will have to become trivial. The authors of explain this coincidence as a hint of the fact that asymptotically the EA model will also behave as a droplet model.
Here we show that this similarity in the behavior of the MKA and the true EA $`3d`$ spin glass does not concern observables that are crucial for determining replica symmetry breaking (RSB). We look at the link overlap (on a system of linear size $`L`$ and volume $`V=L^3`$) $`q^{(L)}1/(3V)^{}\sigma _i\sigma _{i+\widehat{\mu }}\tau _i\tau _{i+\widehat{\mu }}`$, where the sum runs over first-neighbor site pairs. $`q^{(L)}`$ is more sensitive than the usual overlap $`q`$ to the difference between a droplet and a mean field like behavior. The link overlap is, as discussed for example in , of crucial importance, since a non-trivial $`P(q)`$ could be simply due to the presence of interfaces, while a non-trivial $`P(q^{(L)})`$ is a non-ambiguous signature of RSB.
We show that one can see a clear difference, already at $`T=0.7`$ on medium-size lattices, among the MKA and the EA model. So, not only our observation makes the point of obsolete, but it also shows that simulations on reasonable-sized lattices are useful, when studying either disordered systems or normal statistical mechanical models (from the point of view of the advocates of in the case of disordered systems only simulations on systems of huge size could make the true nature of the system manifest).
We have analyzed the MKA of the $`3d`$ spin glass (averaging over $`1000`$ disorder samples), and the $`3d`$ EA model by numerical simulations (using a tempering algorithm and an annealing scheme, checking convergence and averaging over $`64`$ or more samples). In all cases we have considered binary couplings and a Hamiltonian $`H_ϵ[\sigma ,\tau ]=H_0[\sigma ]+H_0[\tau ]ϵ^{}\sigma _i\sigma _{i+\widehat{\mu }}\tau _i\tau _{i+\widehat{\mu }}`$, where $`H_0`$ is the usual EA $`3d`$ Hamiltonian.
In fig. (1) we show our results for $`q^{(L)}(ϵ)`$ versus $`ϵ^{\frac{1}{2}}`$. The MKA gives a smooth behavior: for small $`ϵ`$, $`q^{(L)}(ϵ)`$ behaves like $`ϵ^\lambda `$, with $`\lambda 1`$. Finite size effects look very small for these sizes (from $`4`$ to $`16`$). The EA model behaves in a completely different way. Here finite size effects are large, and the behavior for small $`ϵ`$ becomes more singular for larger sizes. The $`L=4`$ lattice is reminiscent of the MKA behavior, but already at $`L=8`$ the difference is clear. From our data we are not able to definitely establish the existence of a discontinuity, but the numerical evidence is strongly suggestive of that. The data are suggestive of the building up of a discontinuity as $`L\mathrm{}`$, i.e. $`q=q_++A_+ϵ^\lambda `$ for $`ϵ>0`$ and $`q=q_{}+A_{}|ϵ|^\lambda `$ for $`ϵ<0`$, with $`q_+q_{}`$ and an exponent $`\lambda `$ close to $`\frac{1}{2}`$: a continuous behavior (i.e. $`q_+=q_{}`$) cannot be excluded from these data, but in this case we find an upper limit $`\lambda <0.25`$, totally different from the behavior of MKA, $`\lambda 1`$. This is what is needed to show that when looking at observables that are very sensitive to RSB the difference among the trivial behavior of the MKA and true spin glasses is already clear at $`T0.6T_c`$ on lattices of size $`L16`$, as opposed to the claims of .
E. Marinari, G. Parisi, J. J. Ruiz-Lorenzo and F. Zuliani
|
no-problem/9812/astro-ph9812096.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The gamma-ray mission AGILE (Astro-rivelatore Gamma a Immagini LEggero) is currently in a study phase for the Italian Space Agency (ASI) program of Small Scientific Missions. AGILE ideally conforms to the faster, cheaper, better philosophy adopted by space agencies for scientific missions. Gamma-ray detection by AGILE is based on silicon tracking detectors developed for space missions by INFN and Italian University laboratories during the past years (Barbiellini et al., 1995; Morselli et al., 1995). AGILE is both very light ($`60`$ kg) and highly efficient in detecting and monitoring gamma-ray sources in the energy range 30 MeV–50 GeV. The accessible field of view is unprecedently large ($`\stackrel{>}{}1/5`$ of the whole sky) because of state-of-the-art readout electronics and segmented anticoincidence system. AGILE was selected by ASI (1997 December) for a phase A study to be completed within the end of 1998. The goal is to achieve an on-axis sensitivity comparable to that of EGRET on board of CGRO (a smaller background resulting from an improved angular resolution more than compensates the loss due to a smaller effective area) and a better sensitivity for large off-axis angles (up to $`60^{}`$). Planned to be operational during the year 2001 for a 3-year mission, AGILE will ideally ‘fill the vacuum’ between the end of EGRET operations and GLAST. AGILE’s data will provide crucial support for ground-based observations and several space missions including AXAF, INTEGRAL, XMM, ASTRO-E, SPECTRUM-X.
——————————————————————————-
($``$) Adapted from a paper presented at the Conference Dal nano- al Tera-eV: tutti i colori degli AGN, Rome 18-21 May 1998, to be published by the Memorie della Societa’ Astronomica Italiana.
Fig. 1 (left panel) shows the baseline instrument layout. We refer to a companion paper for more details on the instrument (Morselli et al., 1998). Spectral information ($`\mathrm{\Delta }E/E1`$) will be obtained by multiple scattering of created pairs in tungsten-silicon planes (for energies less than $`500`$ MeV) and by the use of a mini-calorimeter. We are also studying the possibility of adding an ultra-light coded mask imaging system sensitive in the energy band $``$10–40 keV on top of AGILE. Super-AGILE is an innovative concept<sup>1</sup><sup>1</sup>1Developed in collaboration with E. Costa, M. Feroci, L. Piro and P. Soffitta, combining silicon technology to simultaneously detect gamma-rays and hard X-rays with accurate imaging.
## 2 Scientific Objectives
Table 1 summarizes the expected performance of AGILE vs. that of EGRET. Because of the large field of view ($`0.8\pi `$ sr) AGILE will discover a large number of gamma-ray transients, monitor known sources, and allow rapid multiwavelength follow-up observations because of a dedicated data analysis and alert program. Fig. 1 (right panel) shows the off-axis response to gamma-ray detection of AGILE and EGRET. We summarize here AGILE’s scientific objectives.
$``$ Active Galactic Nuclei. For the first time, simultaneous monitoring of a large number of AGNs per pointing will be possible. Several outstanding issues concerning the mechanism of AGN gamma-ray production and activity can be addressed by AGILE including: (1) the study of transient vs. low-level gamma-ray emission and duty-cycles; (2) the relationship between the gamma-ray variability and the radio-optical-X-ray-TeV emission; (3) the correlation between relativistic radio plasmoid ejections and gamma-ray flares. A program for joint AGILE and ground-based monitoring observations is being planned. On the average, AGILE will achieve deep exposures of AGNs and substantially improve our knowledge on the low-level emission as well as detecting flares. We conservatively estimate that for a 3-year program AGILE will detect a number of AGNs 2–3 times larger than that of EGRET. A companion paper presents the impact of AGILE on the study of AGNs (Mereghetti et al., 1998). Super-AGILE will monitor, for the first time, simultaneous AGN emission in the gamma-ray and hard X-ray ranges.
$``$ Diffuse Galactic and extragalactic emission. The AGILE good angular resolution and large average exposure will further improve our knowledge of cosmic ray origin, propagation, interaction and emission processes. We also note that a joint study of gamma-ray emission from MeV to TeV energies is possible by special programs involving AGILE and new-generation TeV observatories of improved angular resolution.
$``$ Gamma-ray pulsars. AGILE will contribute to the study of gamma-ray pulsars in several ways: (1) improving photon statistics for gamma-ray period searches by dedicated observing programs with long observation times of 1-2 months per source; (2) detecting possible secular fluctuations of the gamma-ray emission from neutron star magnetospheres; (3) studying unpulsed gamma-ray emission from plerions in supernova remnants and searching for time variability of pulsar wind/nebula interactions, e.g., as in the Crab nebula (de Jager et al., 1996).
$``$ Galactic sources, new transients. A large number of gamma-ray sources near the Galactic plane are unidentified, and sources such as 2CG 135+1 or transients (e.g., GRO J1838-04) can be monitored on timescales of months/years. Also Galactic X-ray jet sources (such as Cyg X-3, GRS 1915+10, GRO J1655-40 and others) can produce detectable gamma-ray emission for favorable jet geometries, and a TOO program is planned to follow-up new discoveries of micro-quasars.
$``$ Gamma-ray bursts. About ten GRBs have been detected by EGRET’s spark chamber during $`7`$ years of operations (Schneid et al., 1996a). This number appears to be limited by the EGRET FOV and sensitivity and not by the GRB emission mechanism. GRB detection rate by AGILE is expected to be a factor of $`5`$ larger than that of EGRET, i.e., $`\stackrel{>}{}`$5–10 events/year). The small AGILE deadtime ($`\stackrel{>}{}100`$ times smaller than that of EGRET) allows a better study of the initial phase of GRB pulses (for which EGRET response was in many cases inadequate). The remarkable discovery by EGRET of ‘delayed’ gamma-ray emission up to $`20`$ GeV from GRB 940217 (Hurley et al., 1994) is of great importance to model burst acceleration processes. AGILE is expected to be highly efficient in detecting photons above 10 GeV because of limited backscattering. Super-AGILE will be able to locate GRBs within a few arcminutes, and will systematically study the interplay between hard X-ray and gamma-ray emissions.
$``$ Solar flares. During the last solar maximum, solar flares were discovered to produce prolonged high-intensity gamma-ray outbursts (e.g., Schneid et al., 1996b). AGILE will be operational during part of the next solar maximum and several solar flares may be detected. Particularly important for analysis will be the flares simultaneously detected by AGILE and HESSI (sensitive in the band 20 keV–20 MeV).
## 3 Mission
AGILE is planned to be integrated with a spacecraft of the MITA class currently being developed by Gavazzi Space with the support of ASI. AGILE’s pointing is obtained by a three-axis stabilization system with an accuracy near 0.5–1 degree. Pointing reconstruction reaching an accuracy of 1–2 arcmin is obtained by star trackers. The downlink telemetry rate is planned to be $`500\mathrm{kbit}\mathrm{s}^1`$, and is adequate for AGILE and Super-AGILE for a single contact per orbit. The ideal orbit is equatorial (550 – 650 km).
The AGILE mission is being planned as an Observatory open to the international scientific community. Planning of pointed observations, quicklook and standard data analysis results will be available to the community through a Guest Observer Program. The AGILE mission emphasizes a rapid response to the detection of gamma-ray transients. The AGILE Science Support Group will help coordinating multiwavelength observations of gamma-ray sources, and will stimulate investigations of observational and theoretical nature on gamma-ray sources detected by AGILE.
|
no-problem/9812/cond-mat9812420.html
|
ar5iv
|
text
|
# First Order Transitions and Multicritical Points in Weak Itinerant Ferromagnets
\[
## Abstract
It is shown that the phase transition in low-$`T_c`$ clean itinerant ferromagnets is generically of first order, due to correlation effects that lead to a nonanalytic term in the free energy. A tricritical point separates the line of first order transitions from Heisenberg critical behavior at higher temperatures. Sufficiently strong quenched disorder suppresses the first order transition via the appearance of a critical endpoint. A semi-quantitative discussion is given in terms of recent experiments on MnSi, and predictions for other experiments are made.
\] The thermal paramagnet-to-ferromagnet transition at the Curie temperature $`T_C`$ is usually regarded as a prime example of a second order phase transition. For materials with high $`T_C`$ this is well established both experimentally and theoretically. Recently there has been a considerable interest in the corresponding quantum phase transition of itinerant electrons at zero temperature ($`T=0`$), and in the related finite $`T`$ properties of weak itinerant ferromagnets, i.e. systems with a very low $`T_C`$. Experimentally, the transition in the weak ferromagnet MnSi has been tuned to different $`T_C`$ by applying hydrostatic pressure. Interestingly, the transition at low $`T`$ was found to be of first order, while at higher transition temperatures it is of second order. The tricritical temperature that separates the two types of transitions was found to roughly coincide with the location of a maximum in the magnetic susceptibility in the paramagnetic phase. Theoretically, it has been shown that in a $`T=0`$ itinerant electron system, soft modes that are unrelated to the critical order parameter (OP) or magnetization fluctuations couple to the latter. This leads to an effective long-range interaction between the OP fluctuations. In disordered systems, the additional soft modes are the same ‘diffusons’ that cause the so-called weak-localization effects in paramagnetic metals. In clean systems there are analogous, albeit weaker, effects that manifest themselves as corrections to Fermi liquid theory. A Gaussian theory is sufficient to obtain the exact quantum critical behavior in the most interesting dimension, $`d=3`$, for clean as well as for disordered systems (apart from logarithmic corrections in the clean case).
In this Letter we show that at sufficiently low temperatures the phase transition in itinerant ferromagnets is generically of first order. This surprising result is shown to be rooted in fundamental and universal many-body physics underlying the transition, viz. long-wavelength correlation effects, and hence to be independent of the band structure. This suggests that the behavior observed in MnSi is generic, and should also be present in other weak itinerant ferromagnets. We also make detailed predictions about how quenched disorder suppresses the first order transition, which allows for decisive experimental checks of our theory.
Let us start by deriving the functional form of the free energy of a bulk itinerant ferromagnet at finite $`T`$, and in the presence of quenched disorder that we parametrize by $`G=1/ϵ_\mathrm{F}\tau `$, with $`ϵ_\mathrm{F}`$ the Fermi energy, and $`\tau `$ the elastic mean-free time. The general Landau expansion of the free energy $`F`$ as a function of the magnetic moment $`m`$ in an approximation that neglects OP fluctuations is
$$F=tm^2+u_4m^4+u_6m^6+\mathrm{}.$$
(2)
The coefficients $`t`$, $`u_4`$, $`u_6`$, etc. in this expansion can have nontrivial properties and contain important physics. A derivation from a microscopic theory shows that they are given as frequency-momentum integrals over correlation functions in a ‘reference system’ that depends on the nature of the underlying microscopic model. If the critical magnetization fluctuations are the only soft modes in the system, then they are simply numbers. However, if in the process of deriving the Landau functional some other soft modes have been integrated out, then the coefficients will in general not exist, since they are represented as diverging integrals over the soft modes. In Refs. and it was shown that in an itinerant electron system at $`T=0`$ there are indeed such soft modes. In the disordered case, these are the ‘diffusons’ mentioned above, with a dispersion relation $`\omega k^2`$, and they lead to coefficients whose divergent parts have the form
$$u_{2m}_0^\mathrm{\Lambda }𝑑kk^2𝑑\omega \frac{1}{(\omega +k^2)^{2m}}.$$
(3)
Here $`\mathrm{\Lambda }`$ is a momentum cutoff, and all prefactors in the integrals have been omitted. In the clean case, the relevant soft modes are particle-hole excitations in the spin-triplet channel with a ballistic dispersion relation, $`\omega k`$. The resulting integrals are still divergent, although not as strongly as in the disordered case. It was shown in Refs. , that these divergent terms in the Landau expansion can be understood as an illegal expansion of a nonanalytic term in the free energy of the form
$$f(m)=m^4_0^\mathrm{\Lambda }𝑑kk^2_0^{\mathrm{}}𝑑\omega \frac{(1)^x}{\left[(\omega +k^x)^2+m^2\right]^2}.$$
(4)
In the disordered case, where $`x=2`$, this follows explicitly from Eq. (3.6’) of Ref. . In the clean case, an analogous treatment yields the same expression with $`x=1`$. Notice the different sign of the dirty case compared to the clean one, which we will come back to below. Equation (4) yields $`f(m)m^{5/2}`$ and $`f(m)m^4\mathrm{ln}m`$ in the disordered and clean cases, respectively. In either case the resulting singularity is protected by the magnetization, which gives the soft modes a mass. The leading effect of $`T0`$ is adequately represented by replacing $`\omega \omega +T`$. In addition, in the presence of disorder the ballistic modes in the clean case obtain a mass proportional to $`1/\tau `$, so the appropriate generalization of Eq. (4) for the clean case ($`x=1`$) to finite temperature and disorder is obtained by the replacement $`\omega \omega +T+1/\tau `$. Doing the integrals, and adding the usual terms of order $`m^2`$ and $`m^4`$, we obtain a free energy of the form
$`F`$ $`=`$ $`tm^2+G(N_\mathrm{F}\mathrm{\Gamma }_t)m^4\left[m^2+(\alpha T)^2\right]^{3/4}`$ (6)
$`+vm^4\mathrm{ln}\left(m^2+(T+\beta G)^2\right)+um^4+O(m^6),`$
where $`\mathrm{\Gamma }_t`$ is an effective spin-triplet interaction amplitude made dimensionless by means of a density of states at the Fermi level, $`N_\mathrm{F}`$. If we measure $`F`$, $`m`$, and $`T`$ in terms of a microscopic energy, e.g. $`ϵ_\mathrm{F}`$, then $`t`$, $`v`$, and $`u`$ are all dimensionless. $`v`$ is quadratic in $`\mathrm{\Gamma }_t`$. $`t=1N_\mathrm{F}U`$ is the dimensionless distance from the critical point. It depends on the physical spin-triplet interaction amplitude $`U`$, with $`N_\mathrm{F}U1`$ in a ferromagnetic or nearly ferromagnetic system, while $`\mathrm{\Gamma }_t`$ above is an effective interaction amplitude with $`N_\mathrm{F}\mathrm{\Gamma }_t<1`$. $`\mathrm{\Gamma }_t`$ is expected to be relatively larger in strongly correlated systems. Finally, $`\alpha `$ and $`\beta `$ are parameters that measure the relative strengths of the temperature and the disorder dependence, respectively, in the two nonanalytic terms. They are numbers of order unity, and like $`u`$ and $`v`$ they are non-universal. Equation (6) provides a functional form of the free energy that correctly describes the leading nonanalytic $`m`$-dependence for both clean and disordered systems, as well as the leading temperature cutoff for either term and the leading disorder cutoff for the clean nonanalyticity.
The sign of $`v`$ merits some attention. Perturbation theory to second order in $`\mathrm{\Gamma }_t`$ yields $`v>0`$. Further, $`v>0`$ indicates a decrease of the effective Stoner coupling constant $`I`$ due to correlation effects: $`I`$ is a homogeneous spin susceptibility, $`v>0`$ means that this susceptibility increases as the wavenumber increases from zero, and correlation effects decrease with increasing wavenumber. It is well known that correlation effects in general decrease $`I`$, and $`v>0`$ is consistent with that. Ref. has given some possible mechanisms for $`v`$ to be negative at least in some materials, and showed that in this case the ferromagnetic transition is always of second order. However, the generic case is $`v>0`$, which we will now discuss.
We first consider the case $`T=0`$. The transition in the clean system, $`G=0`$, is then of first order, since $`m^4\mathrm{ln}m<0`$ for small $`m`$. Upon disordering the system, $`G>0`$, the negative term is no longer the leading one at $`t=0`$. For small values of $`G`$, the transition remains first order. However, for $`G`$ exceeding a value $`G_{ce}`$ the first order transition occurs only at $`t<0`$, and it is pre-empted by a second order transition. Since the negative term is only the third term in an $`m`$-expansion of $`F`$, the multicritical point where the nature of the transition changes is a critical endpoint (CEP). The phase diagram in the $`G`$-$`t`$ plane is shown in Fig. 1. For $`G_{ce}<G<G_c`$, the second order transition at $`t=0`$ is followed by a second transition, the second one being of first order, to a state with a larger magnetization. The line of first order transitions ends in a critical point (CP) at a disorder value $`G_c`$, where the two minima in the free energy merge.
Before we consider $`T>0`$, let us discuss this result and the validity of our conclusions. To facilitate an analytic discussion, we put $`\beta =0`$. We then have $`F=tm^2+G(N_\mathrm{F}\mathrm{\Gamma }_t)m^{5/2}+2vm^4\mathrm{ln}m+um^4`$. At $`G=0`$ there is a first order transition at $`t=v\mathrm{exp}[(1+u/v)]`$, and the magnetization at the transition has a value $`m=\mathrm{exp}[(1+u/v)/2]`$. Notice that the nonanalytic term is the leading one in $`F`$ after the $`tm^2`$ term, and that we know the functional form of $`F`$ exactly up to $`O(m^4)`$. As long as $`u/v>>1`$, $`m`$ is exponentially small at the transition. For small $`v`$, our Landau expansion is therefore controlled in the sense that terms of $`O(m^6)`$ and higher would have to have exponentially large coefficients in order to change our results. For $`G>G_{ce}=(4v/3N_\mathrm{F}\mathrm{\Gamma }_t)\mathrm{exp}[(1+3u/4v)]`$, the first order transition is pre-empted by a second order one. At the CEP, the magnetic moment has the value $`m=\mathrm{exp}[(2/3+u/2v)]=e^{1/6}m(G=0)`$. Allowing for $`\beta =O(1)0`$, and repeating the calculation numerically, leads only to minor quantitative changes of these results.
At $`T>0`$, the free energy is an analytic function of $`m`$, but for small $`T`$ the coefficients in an $`m`$-expansion become very large. Our remarks about the validity of our truncated Landau expansion therefore still apply, i.e., at $`0<k_\mathrm{B}T<<ϵ_\mathrm{F}`$, our theory contains the most important terms to every order in an expansion in powers of $`m^2`$. Let us first consider the clean system, $`G=0`$. There is a tricritical point (TCP) at $`T_{\mathrm{tc}}=\mathrm{exp}(u/2v)`$, with a first order transition for $`T<T_{\mathrm{tc}}`$, and a line of Heisenberg critical points for $`T>T_{\mathrm{tc}}`$. To describe the (conventional) tricritical behavior in $`d=3`$ our mean-field theory is sufficient (apart from logarithmic corrections), for the critical behavior at $`T>T_{\mathrm{tc}}`$ it is of course not.
For the suppression of the first order transition by disorder at $`T>0`$ we find two different possibilities, depending on the value of the parameter $`\alpha `$. For small $`\alpha `$ ($`\alpha 1.5`$ with our choice of the remaining parameters, Fig. 2), the TCP is replaced by a CEP for $`G`$ larger than some $`G_{tce}<G_{ce}`$. At $`G=G_{ce}`$, the CEP reaches $`T=0`$, and for larger values of $`G`$ the transition is of second order for all $`T`$. At small $`T`$, it is followed by a first order transition. The line of first order transitions ends in a critical point, and disappears only for $`G=G_c`$. For larger values of $`\alpha `$ (Fig. 3), the TCP persists for a range of disorder larger than $`G_{ce}`$. The first order transition first gets pre-empted in a temperature window between two CEPs. At $`G=G_{ce}`$, the lower CEP reaches $`T=0`$, while the TCP at higher temperature survives. With further increasing disorder, two CPs appear in the ordered phase, and the remaining CEP gets replaced by a TCP. Finally, the two TCPs merge, and the remaining CP reaches $`T=0`$, eliminating the last temperature regions with first order transitions.
Notice that the interesting features of these phase diagrams do not depend on the logarithm in Eq. (6); similar features are obtained in standard phenomenological Landau expansions with a negative coefficient of the third term. We stress again, however, that in our case the expansion is controlled, and that we have a definite physical mechanism for the appearance of a negative term, in contrast to purely phenomenological theories.
We now turn to a discussion of the available experimental information on this subject. MnSi has a low $`T_C`$ ($`30\mathrm{K}`$) under ambient pressure, and $`T_c`$ can be driven to zero by a hydrostatic pressure $`p_c15\mathrm{k}\mathrm{b}\mathrm{a}\mathrm{r}`$. $`k_\mathrm{B}T/ϵ_\mathrm{F}<<1`$ always, and $`T`$ is low enough to suppress phase breaking processes, so the quantum critical behavior is easily accessible experimentally. This system has been studied in detail by Pfleiderer et al. These authors found from susceptibility measurements that the transition turns first order at a $`T_c`$ of about 12 K. The line of second order transitions was found to scale with pressure like $`T_c(p_cp)^{3/4}`$, while in the first order regime the transition temperature varies like $`T_1(pp_c)^{1/2}`$. The scaling of $`T_c`$ with pressure was explained by a scaling analysis based on the self-consistently renormalized (SCR) theory of Moriya and Kawabata, assuming a dynamical exponent $`z=3`$. The first order transition at low $`T`$ was attributed in Ref. to a sharp structure in the density of states at the Fermi level.
Let us look at the experiment in the light of the above discussion. In Ref. it was shown that the quantum phase transition in $`d=3`$ is indeed correctly described by SCR theory, apart from logarithmic corrections that would be very difficult to detect experimentally, and that the dynamical critical exponent in $`d=3`$ is $`z=3`$. The analysis of Ref. was therefore adequate, and in particular the quantum-to-classical crossover exponent $`\varphi `$, which determines the behavior of the critical temperature as a function of $`t`$ through the relation $`T_ct^\varphi `$, has a value $`\varphi =3/4`$. If one makes the plausible assumption that $`t`$ depends linearly on the hydrostatic pressure, at least for small $`t`$, then this is in agreement with both the experimental finding and the analysis in Ref. . As for the pressure dependence of $`T_1`$, one of the temperature scales in the problem is the Fermi liquid temperature scale, which arises from a quadratic $`T`$-dependence of $`t`$. Since the first order transition is determined by the condition $`t(T_1)=\mathrm{const}.`$, we immediately get $`T_1\sqrt{p_cp}`$, where we again assume a linear relation between $`p`$ and $`t`$.
We finally discuss the observation that the tricritical temperature roughly coincides with a minimum of the inverse magnetic susceptibility $`\chi ^1`$ in the paramagnetic phase. In $`d`$ dimensions, the leading $`T`$-dependence of the paramagnetic susceptibility is of the form
$$\chi /2N_\mathrm{F}=1+2\stackrel{~}{v}_dT^2T^{d3}\stackrel{~}{u}_dT^2,$$
(7)
In $`d=3`$, the nonanalyticity is of the form $`T^2\mathrm{ln}T`$. A calculation of $`\stackrel{~}{v}_3`$ to second order in $`\mathrm{\Gamma }_t`$ revealed that to that order, $`\stackrel{~}{v}_3=0`$, in agreement with prior results from Fermi liquid theory. Ref. also discussed that there are reasons to believe that the exact value of $`\stackrel{~}{v}_3=0`$ may be nonzero. If we assume that this is the case, then we obtain a minimum in $`\chi ^1`$ at a temperature $`T_{\mathrm{min}}=\mathrm{exp}(\stackrel{~}{u}_3/2\stackrel{~}{v}_31/2)`$. Since the nonanalyticities in $`F`$ and $`\chi `$ are manifestations of the same singularity, one expects $`\stackrel{~}{u}_3u`$ and $`\stackrel{~}{v}_3v`$, so that $`T_{\mathrm{min}}T_{\mathrm{tc}}`$. While this provides a possible explanation for the observation, we stress the speculative nature of the above considerations due to the theoretical uncertainty about a nonanalytic $`T`$ dependence of $`\chi `$ in $`d=3`$.
Our theory thus provides us with a complete explanation for the nature of the transitions observed in MnSi, and in particular for the existence of a first order transition at low $`T`$, which in Ref. was attributed to a band structure feature characteristic of MnSi. While this feature may well be sufficient to make the transition in MnSi of first order, the present theory leads to the surprising prediction that the first order transition is generic, and thus should be present in other weak itinerant ferromagnets as well. Our theory further predicts in detail how the first order transition will be suppressed by quenched disorder. Observations of such a suppression, or lack thereof, would be very interesting for corroborating or refuting the theory. Semi-quantitatively, the theory predicts that the $`T`$ region that shows a first order transition will be largest for strongly correlated systems. Conversely, since the dependence of the tricritical temperature on the system parameters is exponential, in some, or even many, systems the first order transition may take place only at very low temperatures. This may explain why in ZrZn<sub>2</sub> no first order transition has been observed, although the experiment does not seem to rule out a weakly first order transition.
We gratefully acknowledge helpful conversations with G. Lonzarich and C. Pfleiderer. This work was supported in part by the NSF under grant Nos. PHY94–07194, DMR–98–70597, and DMR–96–32978, and by the DFG under grant No. SFB 393/C2.
|
no-problem/9812/gr-qc9812066.html
|
ar5iv
|
text
|
# Acknowledgements
## Acknowledgements
I would like to thank Universidade do Estado do Rio de Janeiro for financial Support.
|
no-problem/9812/astro-ph9812202.html
|
ar5iv
|
text
|
# Extreme blazars
## 1 Introduction
Although united in a single class because of their similar properties, blazars come in different flavors, according to the strength of their emission lines, the level of their optical polarization and the ratio between their X–ray and radio fluxes. The most important recent realization has been the discovery that blazars are powerful $`\gamma `$–ray emitters: at least during flares, they can emit in this band up to 90% of their bolometric output (von Montigny et al. 1995; Thompson et al. 1995, 1996; Weekes et al. 1996; Petry et al. 1996). Thanks to EGRET, onboard CGRO, and to the impressive improvements of ground based Cherenkov telescopes, we at last know where most of the blazar power is emitted. Rapid variability indicates that the $`\gamma `$–ray and X–ray emitting regions are compact, and this is yet another proof that the bulk of blazar emission is beamed, since otherwise the source would be opaque to $`\gamma `$–rays (see e.g. Dondi & Ghisellini 1995).
On the theoretical side, this discovery led to major improvements in our understandings of blazars physics in particular and how jet must work in general. One important outcome of these studies is the realization that blazars probably form a sequence in their properties. As will be explained in somewhat more detailed below, there is a link between the maximum energy of the emitting electrons and the intrinsic luminosity of the source. Therefore there is a link between the overall spectral energy distribution (SED) and the bolometric luminosity of blazars. Very high energy electrons are present only in low luminosity sources, and these are the ones emitting significantly in the GeV–TeV band. Can even more extreme objects exist?
## 2 The SED of blazars
We now know that the SED of blazars is characterized by two broad peaks, the first in the IR–EUV and sometimes X–ray band, and the second in the MeV–GeV bands. There is almost unanimity about the interpretation of the first peak as due to incoherent synchrotron emission. The only remaining doubt concerns the interpretation of the fast (intraday) radio variability observed in a large fraction (about 1/4) of blazars (Wagner & Witzel 1995). If it is indeed an intrinsic property, and it is not due to interstellar scintillation, then this phenomenon will deeply change the present theoretical “paradigm”, requiring some coherent process to be at work (Benford & Lesh 1998). The origin of the second peak is still uncertain. It can be due to synchrotron Self Compton (SSC: Maraschi, Ghisellini & Celotti 1992; Bloom & Marscher 1996) or a mixture of SSC plus a contribution by inverse Compton scattering off photons produced externally to the jet (EC: Dermer & Schlickeiser 1993; Sikora, Begelman & Rees 1994; Blandford & Levinson 1995, Ghisellini & Madau 1996), or another more energetic synchrotron component (as in the ‘proton blazar’ model by Mannheim 1993).
In this framework, objects presenting extreme properties are the most interesting, since they best constrain our models. In this respect the discovery made by the existing Cherenkov telescopes of TeV emission from blazars is extremely important, and stimulates new and interesting ideas on the physics of relativistic jets. If, as the correlated variability suggests, the TeV flux is produced by the same electrons dominating by synchrotron emission in the X–ray band, then we have a powerful tool to study in detail the acceleration process at its limit.
Even before the advent of EGRET blazars were divided in subclasses according to their (low energy) SED. For instance, BL Lacs discovered through radio and X–ray surveys were recognized to have different radio to X–ray spectra, even if sharing other properties such as the absence of strong emission lines, the rapid and large amplitude variability and the same average X–ray luminosity. This led Maraschi et al. (1986) and Ghisellini & Maraschi (1989) to try to unify these two blazar subclasses assuming that they simply corresponded to a different viewing angle under which we see an accelerating, inhomogeneous jet.
But Giommi & Padovani (1994) later noticed that the SED of radio and X–ray selected BL Lacs showed peaks at different energies, and proposed that this difference was intrinsic, and not due to orientation effects. They then divided BL Lac objects into HBL (high energy peak BL Lac) and LBL (low energy peak BL Lac), the former being sources preferentially selected through X–ray surveys, and the latter through radio surveys. Now we can extend the Giommi & Padovani’s idea of a moving peak also to the high energy part of the SED, since we now know that the two peak energies correlate.
## 3 The blazar sequence
We attempted to find regularities in the SED of blazars by first considering the observational properties of complete samples of sources, and then by modeling the SED of those blazars detected in the $`\gamma `$–ray band, to find their intrinsic physical quantities.
The first attempt was performed by Fossati et al. (1998), considering the SLEW survey of BL Lacs, the 1 Jy sample of BL Lacs and the 2 Jy sample of flat spectrum radio sources, for a total of 126 objects. By collecting data from the literature in selected bands we could construct the SED of these blazars, to look for trends with their bolometric luminosities and/or their classification. To this end we divided all sources in radio luminosity bins (it is thought that the radio power traces the bolometric one, see the discussion in Fossati et al. 1998), averaging the data of the sources belonging to the same luminosity bin. The result is shown in Fig. 1: less powerful objects have the synchrotron peak in the soft-medium X–ray range, while their high energy peak is at the highest $`\gamma `$–ray energies. As the total power increases, both peaks shift to lower frequencies, and at the same time the $`\gamma `$–ray luminosity increases its relative importance, arriving to dominate the bolometric output.
The approach to fit all blazars with some spectral information in the $`\gamma `$–ray band was undertaken by Ghisellini et al. (1998), using an homogeneous EC model (but including the SSC contribution). The main result of this study is the finding of a tight correlation between the Lorentz factor of the electrons emitting at the peaks, $`\gamma _{\mathrm{peak}}`$, and the amount of energy density (both magnetic and radiative) present in the emitting region (see Fig. 2). This in turn correlates with the observed (beamed) luminosity, and the ratio between the power of the Compton and the synchrotron components. What is also remarkable is that different subclasses of blazars occupy different regions of Fig. 2, indicating a well defined blazar sequence:
* Low luminosity lineless BL Lacs (HBL) have large values of $`\gamma _{\mathrm{peak}}`$, synchrotron peak energy in the EUV–soft X–rays and a roughly equally powerful Compton component peaking in the GeV–TeV band.
* LBL are characterized by a greater overall luminosity, a smaller $`\gamma _{\mathrm{peak}}`$ and peak energies in the optical and GeV band.
* More powerful sources, such as HPQ and LPQ, have the smallest values of $`\gamma _{\mathrm{peak}}`$, peak energies in the mm–far IR and MeV band, a dominating Compton component in which photons produced externally to the jet are more important than the locally produced synchrotron photons.
* There is a correlation between the intrinsic power and the amount of external photons needed to fit the spectra. As shown in Fig. 2, HBL do not need this external photon component, while some LBL do (as BL Lac itself, see also Sambruna et al. 1999).
One can ask what is the physical reason of the found correlations between $`\gamma _{\mathrm{peak}}`$ and the energy density in the comoving frame of the source, the intrinsic power of the source, and the Compton dominance (i.e. the ratio between the $`\gamma `$–ray and the IR–optical luminosity). One possibility is that there is a competition between the acceleration and the cooling mechanisms in these sources, balancing at $`\gamma _{\mathrm{peak}}`$. Since we find $`\gamma _{\mathrm{peak}}U^{0.6}`$, where $`U`$ is the sum of the magnetic and radiation energy densities, we have that the synchrotron and inverse Compton cooling rate ($`\gamma ^2U`$) is, at $`\gamma _{\mathrm{peak}}`$, nearly the same for all sources. This then suggests the presence of some universal acceleration mechanism, independent of $`\gamma `$ and $`U`$: in powerful sources having a large radiation energy density the balance between gain and losses happens at a small value of $`\gamma _{\mathrm{peak}}`$, while in weaker sources this balance is reached at a larger $`\gamma _{\mathrm{peak}}`$. A problem with this interpretation is the behavior of individual sources, which is sometimes just the opposite: in the 1997 flare of Mkn 501 the bolometric power increased by a factor 20 with respect to the quiescent state, and $`\gamma _{\mathrm{peak}}`$ increased also, by at least a factor $``$10. If the mentioned interpretation is correct, one must assume that during the flare the acceleration rate was much faster than during quiescence, and this contrasts with the supposed “universality” of such mechanism. Further work to solve this discrepancy is needed.
## 4 TeV BL Lacs
At this meeting it was announced that 1ES 1959+658 was detected in the TeV band, bringing the total number of TeV BL Lacs to 5 (the other four are Mkn 421, Punch et al. 1992; Mkn 501, Quinn et al. 1996; 1ES 2344+514, Catanese et al. 1997 and PKS 2155–304, Chadwick et al. 1998). The SED of these 5 sources are shown in Fig. 4. As can be seen their SED is very similar, but note that Mkn 421 and PKS 2155–304 have always shown a $`steep`$ ($`\alpha _x1.5`$) X–ray spectrum, while the X–ray slopes of Mkn 501 and 1ES 2344+514, during flares, are flat ($`\alpha _x<1`$), making the synchrotron emission to peak in the very hard X–ray range (Pian et al. 1998; Giommi, Padovani & Perlman, 1998).
## 5 More extreme BL Lacs
It takes a little leap of imagination to suggest that the blazar sequence extends somewhat more than what already discovered by the recent high energy observations. Here I propose that a new subclass of BL Lac objects can exist, with lower luminosity and larger $`\gamma _{peak}`$ than those HBL already detected at high energies. To explore this possibility, we must first ask:
* Why have we not yet detected these extreme objects? Alternatively: is it possible that we already detected them in some bands, but we do not yet know their extreme nature because of lack of information, such as the X–ray slope and/or the $`\gamma `$–ray flux?
* What limits the maximum electron energies?
### 5.1 Limits on the maximum electron energy: shocks
In strong shocks, electrons can be accelerated efficiently up to the energy where their gains equals their energy losses, which in our case are mainly due to radiation losses. Assuming, as Guilbert, Fabian & Rees (1983), that at each gyroradius the electrons increase their energy by a factor $`\gamma `$, and that the radiative losses are dominated by the synchrotron process, it can be derived that the maximum obtainable $`\gamma B^{1/2}`$, where $`B`$ is the magnetic field. This in turn translates in a maximum synchrotron frequency of $``$70 MeV, independent of the magnetic field. This frequency is then blueshifted by the Doppler factor $`\delta `$, yielding an observable maximum frequency of $`700(\delta /10)`$ MeV. These values are within the EGRET band, but EGRET did not detect any extraordinary blazar whose synchrotron spectrum extends to such large energies. It is therefore unlikely that BL Lacs have synchrotron spectra reaching energies greater than $``$100 MeV. There must then be a more severe limit to the maximum observable synchrotron frequency.
### 5.2 Global energetics
Indeed, another limit can be obtained from the global energetics of jets. The argument is as follows: From the observed extended emission of radio sources we know that at least $`10^{45}`$$`10^{47}`$ erg s<sup>-1</sup> must be transported from the black hole to hundreds of kpc, in the form of Poynting flux and/or bulk kinetic energy of the particles (e.g. Rawlings & Saunders 1991). A fraction of this power can be dissipated into radiation along the way, and one of the main result of EGRET is to have demonstrated that the larger dissipation occurs in a well defined part of the jet (see Ghisellini and Madau 1996 for a more detailed discussion). In steady state, the power dissipated into radiation cannot exceed the power transported by the jet. Since the synchrotron and inverse Compton losses are $`\gamma ^2`$, there is an upper limit to $`<\gamma ^2>`$, corresponding to complete dissipation.
To be more quantitative, let us define $`L_\mathrm{k}`$ and $`L_\mathrm{B}`$ as the power of bulk kinetic motion of the emitting plasma and of Poynting flux, respectively (see e.g. Celotti & Fabian 1993, Ghisellini & Celotti 1998):
$$L_\mathrm{k}=\pi R^2\mathrm{\Gamma }^2\beta cn^{}m_\mathrm{e}c^2(<\gamma >+m_\mathrm{p}/m_\mathrm{e})$$
(1)
$$L_\mathrm{B}=\frac{1}{8}R^2\mathrm{\Gamma }^2\beta cB^2,$$
(2)
where $`R`$ is the cross sectional radius of the jet, $`n^{}=n/\mathrm{\Gamma }`$ is the comoving particle density of average energy $`<\gamma >m_\mathrm{e}c^2`$, and $`m_\mathrm{p}`$, $`m_\mathrm{e}`$ are the proton and electron rest masses, respectively. An electron proton plasma is assumed. The synchrotron intrinsic power is
$$L_\mathrm{s}^{}=Volumen^{}(\gamma )\dot{\gamma }_\mathrm{s}m_\mathrm{e}c^2𝑑\gamma =\frac{2}{9}R^3\sigma _\mathrm{T}cn^{}B^2<\gamma ^2>$$
(3)
where $`\dot{\gamma }_\mathrm{s}`$ is the synchrotron cooling rate, and $`n^{}(\gamma )\gamma ^p`$ between $`\gamma _{\mathrm{min}}`$ and $`\gamma _{\mathrm{peak}}`$. We assume that all electrons partecipating in the bulk flow are accelerated to random energies $`\gamma m_ec^2`$ (i.e. the density $`n^{}`$ in eq. 1 and 3 is the same). For a viewing angle $`1/\mathrm{\Gamma }`$, the luminosity calculated assuming isotropy is related to $`L_\mathrm{s}^{}`$ by $`L_{\mathrm{s},\mathrm{obs}}=\mathrm{\Gamma }^4L_\mathrm{s}^{}`$. The intrinsic power emitted over the entire solid angle equals $`\mathrm{\Gamma }^2L_\mathrm{s}^{}`$. We can then relate the synchrotron power $`\mathrm{\Gamma }^2L_\mathrm{s}^{}`$ to $`L_\mathrm{k}`$ (which is proportional to $`n^{}`$) and $`L_\mathrm{B}`$ (which is proportional to the magnetic energy density $`U_\mathrm{B}`$), obtaining
$$\mathrm{\Gamma }^2L_\mathrm{s}^{}=\frac{16\sigma _\mathrm{T}L_\mathrm{k}L_\mathrm{B}}{9\pi Rm_\mathrm{e}c^3\mathrm{\Gamma }^2}\frac{<\gamma ^2>}{<\gamma >+m_\mathrm{p}/m_\mathrm{e}}=\frac{2\sigma _\mathrm{T}L_\mathrm{k}B^2R}{9\pi m_\mathrm{e}c^2}\frac{<\gamma ^2>}{<\gamma >+m_\mathrm{p}/m_\mathrm{e}}$$
(4)
Requiring $`\mathrm{\Gamma }^2L_\mathrm{s}^{}<L_\mathrm{k}`$ implies:
$$\frac{<\gamma ^2>}{<\gamma >+m_\mathrm{p}/m_\mathrm{e}}<\frac{9\pi }{16}\frac{Rm_\mathrm{e}c^3\mathrm{\Gamma }^2}{\sigma _\mathrm{T}L_\mathrm{B}}=\frac{2\pi }{16}\frac{m_\mathrm{e}c^2}{\sigma _\mathrm{T}RB^2}$$
(5)
Assuming a given particle energy distribution $`n^{}(\gamma )`$ allows to calculate the left hand side of eq. (5) and therefore to derive a limit for $`\gamma _{\mathrm{peak}}`$. For instance, $`n^{}(\gamma )\gamma ^2`$ between $`\gamma _{\mathrm{min}}<m_\mathrm{p}/m_\mathrm{e}`$ and $`\gamma _{\mathrm{peak}}`$ yields
$$\gamma _{\mathrm{peak}}<\frac{9\pi }{2\gamma _{\mathrm{min}}}\frac{m_\mathrm{p}c^2}{\sigma _\mathrm{T}B^2}=\mathrm{\hspace{0.17em}3.2}\times 10^6\frac{1}{\gamma _{\mathrm{min}}R_{16}B^2}$$
(6)
Here the notation $`Q=10^xQ_x`$ is used, with cgs units. The observed synchrotron peak frequency $`\nu _{\mathrm{peak}}=3.7\times 10^6\gamma _{\mathrm{peak}}^2B\mathrm{\Gamma }`$ Hz corresponds to
$$\nu _{\mathrm{peak}}=\mathrm{\hspace{0.17em}1.6}\frac{\mathrm{\Gamma }_1}{\gamma _{\mathrm{min}}^2R_{16}^2B^3}\mathrm{MeV}$$
(7)
If the synchrotron power $`\mathrm{\Gamma }^2L_s^{}`$ is of the same order of $`L_\mathrm{B}`$ and $`L_\mathrm{k}`$, we obtain the scaling of $`\gamma _{\mathrm{peak}}`$ and $`\nu _{\mathrm{peak}}`$ with the intrinsic synchrotron power: $`\gamma _{\mathrm{peak}}(L_s^{})^1`$ and $`\nu _{\mathrm{peak}}(L_s^{})^{3/2}`$. Only the less powerful sources can have their high energy peak above the TeV band.
## 6 The predicted TeV flux of extreme BL Lacs
Given the scaling just obtained, we can easily calculate the predicted spectrum of an homogeneous source, in which the particle distribution is the result of continuous injection and radiative cooling. The main uncertainty in this is the amount of soft photons used as seeds for the Compton scattering process. Since BL Lacs, especially HBL, have very weak lines and no signs of thermal emission (i.e. the blue bump), it is natural to assume, for these sources, that a pure SSC model applies. However, it not easy with pure SSC to fit the spectrum of Mkn 501 during the 1997 flare and at the same time to account for the similar amplitude variability of the synchrotron X–rays and the Compton $`\gamma `$–rays. This is because the observed flattening of the X–ray spectrum for brighter states, if extrapolated towards lower frequencies, constrains the pure SSC model to underpredict the IR flux, and more severely so for brighter states. This implies that although the number of high energy electrons increases during high states, the number of IR photons, which are the seeds for the Compton process, decreases. As a result the TeV flux is predicted to vary much less than observed. This can be “cured” either by:
* assuming that the particle distribution is not the result of continuous injection and cooling. For instance, the injection can be impulsive, with the high energy synchrotron flux strictly following the injection phases, while at lower energies, where the cooling time is longer, the particle can accumulate and be considered the result of an “average” (over time) injection.
* assuming that the active region containing the most energetic electrons is embedded in a larger region (i.e. a contiguous part of the jet) which is much more steady, and that contributes substantially to the radiation energy density in the IR. Also this radiation is beamed with the same Doppler factor, since it is produced by the jet.
In both cases the result is that in the active region there is a population of more steady IR photons contributing to the scattering. In these conditions the hard X–rays and the TeV flux vary together with the same amplitude, as observed. Since this case is somewhat different both from the pure SSC scenario and from the EC case, I will called it “ambient” photon model. The level of the Compton emission is larger than in the pure SSC case, since we have an extra contribution to the seed photons. This can be seen in Fig. 5, where I made the comparison between pure SSC and the “ambient” photon case, assuming that the extra IR photons account for the observed IR flux (see also Ghisellini 1998).
Fig. 6 shows some models along these lines. These are SSC plus “ambient” photon models in which the maximum electron energies increases $`1/L_s^{}`$. Note that in the radio band these models severely underestimate the observed flux, since the radiaiton is self absorbed at high radio frequencies in these compact regions. Additional components must be present in the source, to account for the radio emission. We can, very roughly, estimate the level of the flux in the radio band in the following way: i) extrapolate the thin synchrotron power law of the model fitting Mkn 501 towards low frequencies, obtaining a factor 10 of discrepancy between the extrapolation and the data; ii) do the same extrapolation with the other curves, and multiply the extrapolated flux by the same factor 10.
## 7 Where are very extreme BL Lacs?
Fig. 6 reports for illustration the level of radio flux of 10 mJy at 5 GHz, appropriate for the large area deep radio survey, and the flux of $`5\times 10^{12}`$ erg cm<sup>-2</sup> s<sup>-1</sup> in X–rays, which is approximately the level of the SLEW survey (Perlman et al. 1996). It can be seen that objects like Mkn 501 could have been detected in existing large area X–ray survey (i.e. the ROSAT RASS), even if they lye at a redshift of 0.1 (i.e. 3 times more distant than Mkn 501). In this case the host galaxy contributes in the optical band as in Mkn 501, and these objects can be easily classified as BL Lacs. In the RASS (and possibly SLEW) survey there can therefore exist extreme BL Lacs. However, for the majority of these already detected and classified objects we only know the radio, optical and X–ray fluxes. It is therefore interesting to find a tool able to recognize extreme BL Lacs only on the basis of these fluxes and the corresponding broad band spectral indices, as done e.g. by Fossati (1998). Using this this tool we have selected a handful of BL Lac objects to be observed with BeppoSAX (0120+340, 0224+014, 0120+340, 0548–322, 1101–232, 1320+084, 1426+428, 2005–489, 2356-309): if their X–ray spectral index will turn out to be flat ($`\alpha _x<1`$), this would guarantee that their synchrotron spectrum peaks at high energies, and hence they would be good candidates for detection in the TeV band.
Decreasing the intrinsic power still further, we may find objects whose synchrotron peak is at energies even higher. In the soft X–ray range, the reduced flux could let these objects escape detection in large area X–ray surveys. In the radio, the flux can be below the $``$10 mJy level. Furthermore, even if detected in X–ray surveys, it may be difficult to classify these objects as BL Lacs, because in the optical the contrast between the galaxy and the non–thermal emission is reduced. It may be that the easiest way to find them is through MeV and TeV surveys. VERITAS can therefore play a crucial role in this respect, finding conspicuous TeV sources that look like normal elliptical galaxies in other bands.
How many of these sources do we expect? Since they are intrinsically weak BL Lacs, one expects to have a lot of them. The exact answer however depends on the lower end of the BL Lac luminosity function (which is still unknown), where these extreme sources are predicted to be. One can find a solid upper limit to the number of these sources by requiring that their MeV emission does not overproduce the MeV $`\gamma `$–ray background (Ghisellini et al., in preparation).
## 8 Conclusions
The blazar sequence discussed in this paper suggests a key role for future, more sensitive Cherenkov detectors, such as VERITAS. The most extreme (i.e. emitting at the higher energies) sources should in fact be the intrinsically weakest, therefore the most numerous. There is the possibility that a new subclass of BL Lacs exists, whose synchrotron spectrum peaks in the MeV band and whose Compton spectrum peaks at or even above one TeV, and whose existence may be discovered by instruments like VERITAS, in the TeV band, or like INTEGRAL, in the MeV band.
Very extreme BL Lacs may be the sources where the dissipation of the power carried by the jet is the most efficient, and where we can study the acceleration mechanism at its limit. If they exist, and if they can be detected, they will be very useful for the determination of the IR background: since their peak is above one TeV, it will be much less ambiguous to disentangle the effect of photon–photon absorption from the intrinsic curvature of the spectrum.
Acknowledgments
It is a pleasure to thank Annalisa Celotti and Laura Maraschi for constant help and for years of fruitful discussions
|
no-problem/9812/math9812090.html
|
ar5iv
|
text
|
# A note on one inverse spectral problem.
## Abstract
The note contains the proof of the uniqueness theorem for the inverse problem in the case of $`n`$-th order differential equation.
The inverse spectral problem is studied in papers of many authors (). Extensive bibliographies for the inverse spectral problem can be found in .
Let’s consider the spectral problem for the common differential differential equation:
$$F(x,y(x),y^{}(x),\mathrm{},y^{(n)},q_1(x),q_2(x),\mathrm{},q_m(x),\lambda )=0$$
(1)
with common boundary conditions
$$U_j(y(x),\lambda ,a_0,a_1,\mathrm{},a_s)=0,j=1,\mathrm{},n.$$
(2)
Here $`x[0,1],`$ $`\lambda `$ is eigenvalue parameter, $`q_i`$ ($`i=1,\mathrm{}m`$) are uncertain factors of the equation, $`a_i`$ ($`i=0,\mathrm{},s`$) are uncertain constants of boundary conditions, $`q_iC^1[0,1]`$ ($`i=1,\mathrm{}m`$), $`a_i\text{}`$ ($`i=0,\mathrm{},s`$).
The spectral problem defined by equalities (1)–(2) we shall name $`F.`$
Along with the problem $`F`$ we shall consider $`m`$ problems: $`A_i`$:
$`y^{\prime \prime }+q_i(x)y=\lambda y,`$
$`y^{}(0)=0,`$
$`y^{}(1)=0.`$
and one more problem $`A_{m+1}`$:
$`y^{\prime \prime }+3y^{}+2\lambda ^2y=0,`$ (3)
$`y(0)=0,`$ (4)
$`y^{}(1)+a(\lambda )y(1)=0,`$ (5)
Theorem.
If eigenvalues of the problems $`A_i`$ and $`\stackrel{~}{A}_i`$ ($`i=1,\mathrm{\hspace{0.17em}2},\mathrm{},m+1`$) coincide with their algebraic multiplicities, then the factors of the equations and the constant in the boundary conditions of the problems $`F`$ and $`\stackrel{~}{F}`$ coincide, that is $`q_i(x)\stackrel{~}{q}_i(x),`$ $`a_k=\stackrel{~}{a}_k`$ $`i=1,\mathrm{\hspace{0.17em}2},\mathrm{},m`$ $`k=1,\mathrm{\hspace{0.17em}2},\mathrm{},s.`$
Proof. It follows from Ambarzumijan’s theorem () that the equality $`q_i(x)=\stackrel{~}{q}_i(x)`$ is true.
Functions $`y_1(x,\lambda )=e^{2\lambda x}+2e^{\lambda x},`$ $`y_2(x,\lambda )=\frac{1}{\lambda }(e^{2\lambda x}e^{\lambda x})`$ are solutions of the differential equation (3), satisfying
$$y_1(0,\lambda )=1,y_1^{}(0,\lambda )=0,y_2(0,\lambda )=0,y_2^{}(0,\lambda )=1.$$
(6)
Let $`A(\lambda )`$ be a polynomial $`a_0+a_1\lambda +a_2\lambda ^2+\mathrm{}+a_s\lambda ^s.`$
The eigenvalues $`\lambda _i`$ of the problems (3)–(5) are the roots of a characteristic determinant, therefore its satisfy to the following equation:
$$\mathrm{\Delta }(\lambda )=\frac{1}{\lambda }\left(e^{2\lambda }e^\lambda \right)+a(\lambda )\left(e^{2\lambda }+2e^\lambda \right)=0.$$
(7)
This function has infinite number of the radicals. (For this reason the equation was selected by such: $`y^{\prime \prime }+3y^{}+2\lambda ^2y=0.`$ Generally speaking it was possible to select any equation having not less $`s`$ pairwise different nonzero eigenvalues.)
From (6) we have
$$1a_0+\lambda _ia_1+\lambda _i^{\mathrm{\hspace{0.17em}2}}a_2+\mathrm{}+\lambda _i^sa_s=\frac{e^{2\lambda _i}e^{\lambda _i}}{\lambda _ie^{2\lambda _i}+2\lambda _ie^{\lambda _i}}.$$
(8)
The equalities (8) is a system of $`(s+1)`$ linear equaions having $`(s+1)`$ unknown $`a_0,`$ $`a_1,`$ $`a_2,`$ $`\mathrm{},`$ $`a_s`$ of the boundary condition. The determinant of this system is the Vandermonde determinant. As the eigenvalues $`\lambda _i`$ are pairwise different and are not equal to zero, the Vandermonde determinant is not equal to zero. Therefore system (8) has a unique solution. From here Follows, that the constants $`a_0,`$ $`a_1,`$ $`a_2,`$ $`\mathrm{},`$ $`a_s`$ are determined univalently.
The theorem is proved.
|
no-problem/9812/astro-ph9812175.html
|
ar5iv
|
text
|
# Prompt and Delayed High-Energy Emission from Cosmological Gamma-Ray Bursts
## 1 Introduction
The blast-wave model has met with considerable success in explaining the X-ray, optical and radio afterglows of GRBs . The evolving blast wave accelerates particles to ultrarelativistic energies and has been suggested as the source of ultrahigh-energy cosmic rays (UHECRs) . GeV – TeV photon production, due primarily to proton synchrotron radiation from UHECR acceleration in GRBs during the afterglow phases, has recently been predicted . A large Compton luminosity when relativistic electrons scatter soft photons is expected as well. One source of seed photons is the electron synchrotron radiation, and the synchrotron self Compton (SSC) process has been calculated by . External soft photon sources could also be important in the Compton scattering process. Models for the GRB origin such as the hypernova/collapsar scenario suggest that the sources of GRBs are associated with star-forming regions and might thus be embedded in a starlight radiation field. Here we examine the Compton scattering of starlight by relativistic electrons in the blast wave.
## 2 Compton Emission
We compare high energy $`\gamma `$-ray emissions due to SSC and SSL (Scattered StarLight) processes, which can be estimated by comparing the comoving blast-wave frame energy densities $`u_{SL}^{}`$ and $`u_{sy}^{}`$ of starlight and synchrotron photons, respectively. Knowledge of the peak energy of the seed photon spectrum, which determines whether Compton scattering is suppressed due to Klein-Nishina effects, is important for this comparison. The relativistic electrons are assumed to be distributed according to a power-law with index $`\stackrel{>}{}\mathrm{\hspace{0.25em}3}`$ and a low-energy cutoff $`\gamma _{\mathrm{e},\mathrm{min}}=\xi _e(m_p/m_e)\mathrm{\Gamma }`$ , where $`\xi _e=0.1\xi _{e,1}`$ is an electron equipartition factor, and $`\mathrm{\Gamma }=300\mathrm{\Gamma }_{300}`$ is the bulk Lorentz factor of the blast wave. The magnetic field is $`H=\sqrt{8\pi rm_pc^2n_0\xi _H}\mathrm{\Gamma }`$, where $`r=10r_1`$ is the shock compression ratio, $`n_0`$ is the number density of external matter, assumed uniform, and $`\xi _H=10^6\xi _{H,6}`$ is a magnetic-field equipartition factor. A value of $`\xi _H10^6`$ is required to produce synchrotron spectra resembling observed spectra in the prompt phase of GRBs .
### 2.1 Synchrotron and SSC Processes
The peak photon energy of the synchrotron spectrum is $`ϵ_{sy}^{}10^4\mathrm{\Gamma }_{300}^3n_0^{1/2}q_4`$, where $`q=\sqrt{r_1\xi _H}\xi _e^2=10^4q_4`$ is a combined equipartition factor. The energy density of the synchrotron radiation field may be estimated using $`u_{sy}^{}\tau _T\gamma _{\mathrm{e},\mathrm{min}}^2u_H^{}`$, where $`\tau _T`$ is the radial Thomson depth of the blast wave. Specifying the width of the shell, $`\mathrm{\Delta }x^{}x^{}/\mathrm{\Gamma }`$, at the deceleration radius, we find
$$u_{sy}^{}[\mathrm{erg}\mathrm{cm}^3]7.310^6r_1^2n_0^{5/3}\xi _{H,6}E_{52}^{1/3}\xi _{e,1}^2\mathrm{\Gamma }_{300}^{10/3},$$
(1)
where $`E_0=10^{52}E_{52}`$ erg is the kinetic energy of the blast wave. SSC scattering can occur efficiently in the Thomson regime if $`\xi _{e,1}\mathrm{\Gamma }_{300}^4n_0^{1/2}q_4\stackrel{<}{}\mathrm{\hspace{0.25em}1}`$. In this case, the SSC spectrum in the observer’s frame peaks at
$$ϵ_{\mathrm{SSC}}=h\nu _{\mathrm{SSC}}/m_ec^210^8\mathrm{\Gamma }_{300}^6\xi _{e,1}^2n_0^{1/2}q_4/(1+z),$$
(2)
which cannot exceed $`ϵ_{\mathrm{IC},\mathrm{max}}1.610^7\xi _{e,1}\mathrm{\Gamma }_{300}^2/(1+z)`$.
### 2.2 Scattered Starlight
We assume that the blast wave approaches a star of luminosity $`L_{}=łL_{}`$ at a distance $`d=d_{15}`$ cm from a given point within the blast wave. Then the energy density of the stellar radiation field in the blast-wave frame is
$$u_{SL}^{}[\mathrm{erg}\mathrm{cm}^3]\frac{1.310^3ł\mathrm{\Gamma }_{300}^2}{d_{15}^2}$$
(3)
. By comparing eqs. (1) and (3), we see that $`u_{SL}^{}`$ dominates $`u_{sy}^{}`$ when $`d_{15}\stackrel{<}{}d_{SSL}13ł^{1/2}\mathrm{\Gamma }_{300}^{2/3}r_1^1n_0^{5/6}\xi _{H,6}^{1/2}\xi _{e,1}^1E_{52}^{1/6}`$.
If the starlight spectrum is approximated by a thermal blackbody of temperature $`T_0`$\[eV\], then this spectrum peaks in the comoving blast-wave frame at dimensionless photon energy $`ϵ_{SL}^{}1.610^3\mathrm{\Gamma }_{300}T_0`$. Compton scattering can occur in the Thomson regime if $`\mathrm{\Gamma }_{300}^2\xi _{e,1}T_0\stackrel{<}{}\mathrm{\hspace{0.25em}10}^2`$, indicating that the SSL process becomes efficient only if $`\mathrm{\Gamma }\stackrel{<}{}\mathrm{\hspace{0.25em}30}(\xi _{e,1}T_0)^{1/2}`$. This can be realized in dirty fireballs or in the later afterglow phase of blast-wave deceleration. When the above condition is satisfied, the SSL spectrum in the observer’s frame peaks at
$$ϵ_{SSL}1.510^5\mathrm{\Gamma }_{30}^4T_0\xi _{e,1}^2/(1+z),$$
(4)
where $`\mathrm{\Gamma }_{30}=\mathrm{\Gamma }/30`$. The SSL component can dominate once scattering occurs primarily in the Thomson regime, at which point it is likely to be the dominant electron radiation mechanism in the 10 – 100 GeV regime.
The minimum duration of flares from the SSL process is $`\mathrm{\Delta }t_{\mathrm{min}}d/(\mathrm{\Gamma }^2c)40d_{15}/\mathrm{\Gamma }_{30}^2`$ s. The fraction of energy in GeV – TeV photons produced by the SSL mechanism rather than through other processes is roughly given by $`[d_{SSL}/(x_{bw}/\mathrm{\Gamma })]^2`$, which represents the fraction of the observable blast wave area where $`u_{SL}^{}`$ dominates $`u_{sy}^{}`$. The term $`x_{bw}`$ is the distance of the blast wave from the explosion site. Čerenkov telescopes with threshold energies below $``$ 100 GeV and the planned GLAST satellite might be able to detect the predicted SSL radiation from nearby GRBs if they are indeed associated with star-forming regions.
## 3 Comparison with Gamma Rays from Hadronic Processes
The high energy $`\gamma `$-ray signatures of UHECR acceleration in GRB blast waves have been investigated in detail in . If protons are accelerated efficiently in GRB blast waves, the spectrum in the $`10`$$`100`$ GeV range is expected to be dominated by a hard power-law with $`\nu F_\nu \nu ^{+0.5}`$ due to proton synchrotron radiation. Since protons are expected to cool inefficiently, while electrons suffer strong radiative cooling, the temporal decay of the proton synchrotron radiation is slower than for the synchrotron component component. Let $`g`$ be the index parametrizing the deceleration and thereby the radiative regime of the blast wave, then both components decay as $`F_\nu (t)t^\chi `$, but with different temporal indices. For synchrotron radiation from cooling electrons, $`\chi _{\mathrm{sy}}=(4g2)/(1+2g)`$, while for uncooled syncrotron radiation from UHECR, $`\chi _{\mathrm{p},\mathrm{sy}}=(4g3)/(1+2g)`$. The temporal decay of the SSC radiation is more complicated, because scattering in the Klein-Nishina is important. Comparison of decay observations over a large range of photon energies with model calculations are necessary to accurately discriminate between the various processes. The SSL radiation can be distinguished from the SSC and the hadronic $`\gamma `$-ray emission by its rapid variability and by the flares it produces in the GeV range. Observations of flares of GeV - TeV radiation would support the hypothesis that GRBs occur within stellar associations and star-forming regions.
|
no-problem/9812/cond-mat9812232.html
|
ar5iv
|
text
|
# Coupling of Length Scales and Atomistic Simulation of MEMS Resonators
## 1 The Failures of Standard Finite Elements for Small Devices
The design of MEMS relies on a thorough understanding of the mechanics of the device itself. As system sizes shrink, MEMS are forced to operate in a regime where the assumptions of continuum mechanics are violated, and the usual finite element (FE) models fail. The behavior of materials begins to be atomistic rather than continuous, giving rise to anomalous and often non-linear effects:
* The devices become less stiff and more compliant than FE predicts.
* The roles of surfaces and defects become more pronounced.
* Anharmonic effects become more important.
* New mechanisms for dissipation become evident.
* Statistical Mechanics becomes a key issue, even to the point that thermal fluctuations cannot be neglected.
The inadequacy of FE for these phenomena will be an obstacle to further miniaturization of MEMS.
### 1.1 Micro-Resonator
The failings of continuum elastic theory are evident in many ways for micro-resonators. Gigahertz resonators (cf. Fig. 1) are roughly of the size $`0.2\times 0.02\times 0.01`$ microns. Devices of this size and smaller are so miniscule that materials defects and surface effects can have a large impact on their performance. Atomistic surface processes which would be negligible in large devices are a major source of dissipation in sub-micron devices, leading to a degradation in the Q-value of the resonator. Additionally, bond breaking at defects can produce plastic deformations. These effects vary with temperature, and in the smallest devices, the atomicity shows up through stochastic noise. Systems smaller than about 0.01 microns are too small to be in the thermodynamic limit, and anomalous statistical mechanical effects are important. These effects are beyond continuum elastic theory.
### 1.2 Micro-Gears
The anomalies are also evident in articulated devices. The effects of wear, lubrication and friction can be expected to have profound consequences on the performance of micron-sized machines, where areas of contact are a significant part of the system. An archetypical example is the gear train, something at the heart of many micro-machines of the future (See Fig. 2). The process of micro-gear teeth grinding against each other cannot be simulated accurately with FE. All of the failings listed above are evident. The teeth are predicted to be too rigid. Large amplitude high-frequency resonant modes may be missed. Bond breaking and formation at the point of contact can only be treated empirically.
## 2 Coupling of Length Scales
Sub-micron MEMS require atomistic modeling, but atomistic simulations are very computer intensive. The active region of a device whose dimensions approach a micron may be modeled with atomistics, but it is beyond the capabilities of even the largest supercomputers to model the entire region of interest of a micron-scale device. The coupling of the active region to the substrate is important, and this may involve volumes of many cubic microns and billions of atoms. Thus, many sub-micron MEMS are too small for finite elements and too large for atomistics.
We have developed a novel hybrid methodology which solves this problem. Surfaces and other regions of micro-gears and micro-resonators in which atomistic effects are critical can be modeled accurately with an atomistic simulation. Finite elements offers an adequate and efficient model of other regions such as the body and the axle of the micro-gears and the peripheral regions of the micro-resonator. Then the two, finite elements and atomistics, are melded together to run concurrently through consistent boundary conditions at the handshaking interface. Our multiscale algorithm combines atomistic, finite element and even electronic simulations into a seamless, self-consistent monolithic simulation. This coupling of length scales strikes a balance between computational accuracy and efficiency. This is part of our DOD HPC Grand Challenge project to model the dynamical behavior of MEMS. The project gives us vast computational resources at the Maui Supercomputer Center, where we are developing codes to simulate the behavior of the next generation of MEMS.
### 2.1 Hybrid of electronic structure, atomistics and continuum
Our codes employ electronic and atomistic simulation in the regions of bond-breaking and bond-deformation (large strain), so they are free from the assumptions of continuity that lead to the failure of FE. Instead, the codes are based on well-tested interatomic potentials and electronic parameterizations. The simulation tracks the motion of each individual atom as it vibrates or perhaps diffuses in thermal equilibrium. This means that the simulation automatically includes a wide range of phenomena. The increased compliance at small sizes arises naturally, and when the stresses are large enough to induce plastic deformation, this is simulated, too. No special phenomenology is required for effects due to surface relaxation, bond breaking and asperities.
The technique involves a state-of-the-art atomistic simulation (molecular dynamics, MD) augmented self-consistently with concurrent FE and electronic simulations (tight-binding, TB). This coupling of length scales is a novel finite temperature technique in materials simulation. It has never been attempted before, especially on parallel machines, and it allows the extension of an essentially atomistic simulation to much larger systems. Effects such as bond-breaking, defects, internal strain, surface relaxation, statistical mechanical noise, and dissipation due to internal friction are included. The trick is that the gear train is decomposed into different regions, FE, MD and TB, according to the scale of the physics within that region.
Our current codes will not run on existing desktop workstations, although they may in five years. Several factors will make this possible. First, the intrinsic speed and capacity of workstations will increase exponentially with time (Moore’s Law). Second, multiprocessor workstations will become more common, so parallel codes will offer the same advantage to workstations that they currently offer only to supercomputers. And third, our codes will become more efficient as we continue to make algorithmic advances.
The ultimate goal of this work is to understand the important atomistic effects both qualitatively and quantitatively. To quantify the effects, we have calculated new constitutive relations appropriate for sub-micron resonators. They include terms that describe atomistic surface effects, and other contributions to the energy of the structure. These constitutive relations provide the starting point for continuum and finite element models. Of course, the resulting models do not contain all of the information from the atomistic simulation, so they are only valid for a restricted range of device sizes and geometries, but the resulting finite element computations could be carried out on a workstation. This may be very useful for device design.
### 2.2 Coarse-grained molecular dynamics
The coupling between atomistics and finite elements described above works very well for many applications. For example, in simulations of crack propagation in silicon, the strain fields and elastic waves emanating from the crack tip pass fairly smoothly from the atomistic region into the finite element region. In particular, there is little coherent backscatter of elastic shock waves from the MD/FE interface, a problem that has plagued pure atomistic simulations of cracks. It is clear, however, that finite elements does not connect perfectly smoothly with atomistics in the limit that the element size becomes atomic scale. Finite element analysis assumes that the energy density is spread smoothly throughout each element, but at the atomic scale, the potential energy is localized to the covalent bonds of silicon and the kinetic energy is localized largely to the nuclei. This small atomic scale mismatch can cause problems in some types of simulation.
We have developed a substitute for finite elements which does connect seamlessly to molecular dynamics in the atomic limit. It also reproduces the results of finite elements (with slight improvements) in the limit of large element size. This new methodology is called Coarse-Grained Molecular Dynamics (CGMD).
CGMD has been constructed to provide a consistent treatment of the short wavelength modes which are present in the underlying atomistics but are missing from the coarse finite element mesh. These modes can participate in the dynamics and the thermodynamics of the device. In many situations, the short wavelength part of the spectrum is relatively unpopulated, and the missing modes are irrelevant to the behavior of the device. But this is not true for sufficiently small devices, or when there is a strong source of high frequency elastic waves (such as a propagating crack or when two micro-gears grind against each other). In addition to offering well-behaved thermodynamics, CGMD also models the elastic wave spectrum more accurately than conventional finite elements. Furthermore, CGMD includes non-linear effects that are compatible with the atomistics; i.e. it effectively provides non-linear constitutive equations that are derived from the atomistics without any free parameters.
## 3 Technical Approach
### 3.1 Micro-Resonator
The goal of a multiscale simulation is to balance accuracy with efficiency in an inhomogeneous system. An atomistic simulation is employed in regions of significantly anharmonic forces and large surface area to volume ratios or where internal friction due to defects is anticipated. As shown in Fig. 3, the atomistic simulation models the central region of the resonator. This corrects the expected, but previously unquantified, failure of continuum elastic theory in the smallest MEMS structures. Regions of the micro-resonator which are well-described by continuum elastic theory are simulated using finite elements. These peripheral regions include the coupling to the outside world, a substrate at a given temperature. Electronic structure calculations are used in regions of very large strain and bond breaking, such as at defects. The electronic, atomistic and continuous regions are joined seamlessly to form the complete simulation.
Coupling length scales for the micro-resonator is accomplished as follows. Any defects, the only regions of breaking bonds, are of necessity described by electronic structure methodology. These are coupled to the statistical mechanics of sub-micron system sizes (to provide necessary fluctuations) via conventional molecular dynamics. This region, in turn, is coupled to micron and larger scales via finite elements. In each case the coupling between the regions amounts to a set of consistent boundary conditions that enforce continuity.
This formulation of the coupling of length scales gives a natural domain decomposition to divide the computational load among parallel processors, as shown in Fig. 4. The MD region is partitioned lengthwise into subregions, each of which is assigned to a separate processor. The FE regions (one at each end of the resonator) at present comprise one processor each, since they are not excessively computationally intensive. Each electronic structure domain consisting of about 20 atoms is assigned to a separate processor. The MD region uses an empirical potential suitable for the material of interest; in this case, since the materials are silicon and quartz, the potentials are those due to Stillinger and Weber and Vashishta , respectively.
Our simulations have been limited to relatively small defect densities, so the multi-million atom atomistic region of the resonator requires the majority of the processors. Even though the electronic structure calculation is intrinsically a much more expensive computation, the total TB expense is less because there are relatively few TB atoms. Our codes are written in FORTRAN and MPI and run on IBM SP2s, allowing atomistic regions of millions of atoms coupled, of course, to large finite element regions, if necessary.
### 3.2 Micro-Gears
Fig. 2 shows an example of micro-gear technology. Such devices can presently be made on the 100 micron scale and rotate at speeds of 150,000 RPM. Materials may be either polysilicon or nickel, depending upon the method of manufacture; we concentrate on silicon. We can expect next-generation devices to reach the 1 micron level. The speed with which they could be made to rotate is a subject for our research.
Fig. 5 illustrates the multiscale decomposition for the micro-gear train. An inner region including the shaft is treated by finite elements. FE uses the energy density that comes from constitutive relations for the material of interest to produce a force on each nodal point which drives the displacement field at that point using an algorithm which looks just like molecular dynamics. We are also developing improved codes based on CGMD. The timestep used by the FE region has to be in lock-step with the MD region (and also therefore with the TB region). The handshaking between the FE and MD region is accomplished using a self-consistent overlap region.
Lastly, in regions at the gear-gear contact point in the non-lubricated case, a tight-binding description is used. When we study the lubricating properties of SAMs, the TB region is not required, since bond breaking is not an issue. Each TB region spreads across multiple processors. Tight binding is a fast electronic structure method—with careful parameterization, it can be very accurate. We use the parameterization due to Bernstein and Kaxiras . The effectiveness of this coupling of length scales has been demonstrated previously in simulations of a crack opening in silicon.
## 4 Results
We have simulated micro-resonators with various sizes, defect concentrations and temperatures, for comparison. The dimensions of the largest oscillator are shown in Fig. 6. All of the devices have the same aspect ratio, 25:2:1. The motion of the resonator is simulated as follows. The initial configuration of the atoms is taken to be a single crystal of stoichiometric quartz in the desired device geometry (with some fraction of the atoms removed at random, if vacancies are to be modeled). The system is brought to thermal equilibrium in 100,000 time steps. Then the resonator is deflected into its fundamental flexural mode of oscillation, and released. Once released, the thermostat is turned off, and no further energy is put into the system.
Various properties of the resonator have been studied. Fig. 7 shows the Young’s modulus as a function of size. The resonant frequency of the oscillator is proportional to the square root of the Young’s modulus. The dashed lines in the figure indicate the bulk value, whereas the solid lines indicate the best fit to a constitutive equation which includes a surface term. Atomistic effects are clearly evident for devices less than 0.2 microns in length. Note that this is true even for the single crystal device at T=10K. This device is essentially a perfect crystal, disrupted only by the surfaces. But it is the surface relaxation which produces deviations from the bulk behavior.
Fig. 8 shows how the oscillator rings as a function of time when plucked in flexural mode. Note that relatively large deflections of the resonator are possible, as great as 0.2%, due to the increased compliance of the microscopic devices. The response of the oscillator at 300K shows marked effects of anharmonicity . There is a pronounced frequency doubling effect in the smaller oscillator, and even in the first few periods of the larger oscillator there are clear departures from a sinusoidal oscillation. The mode mixing is most apparent in the Fourier transform of the oscillations of the small device shown in Fig. 9. A significant amount of the first harmonic (as well as a bit of the second) has mixed into the spectrum. This is not the case for an identical simulation run at T=10K, where only the fundamental mode is present (not shown). Fig. 10 shows that the response of the oscillator with 1% vacancies is also anomalous. A substantial plastic deformation has resulted from the presence of vacancies at both temperatures, and again at 300K the response is highly anharmonic.
We can use simulations such as these to calculate the Q-value for the various resonators. The Q-value for the small devices at room temperature shown in Figs. 9 and 10 are Q=300 and Q=200, respectively, at their resonant frequency of 24 GHz. The simulation time for the large system captures too few oscillations for a direct computation of the Q-value, but a scaling analysis together with a fit to the dominant dissipative processes enables us to estimate this Q-value as well. The results will be presented elsewhere.
This behavior could not be predicted from continuum elastic theory. The anharmonic response has been shown to be the result of surface effects. The degradation of the Q-value of small resonators is also due to a surface effect. In large single crystal resonators, the primary sources of dissipation are bulk thermoelastic and phonon-phonon processes. Our results show that for small resonators at room temperature there is a new dominant source of dissipation at the surface. The reduction of Q for small resonators has been observed in experiments, where it has been attributed to flawed surfaces. Our results show that even for initially perfect single crystal devices, atomistic effects cause a significant degradation in Q at room temperature.
The plastic behavior is due to relaxation of the lattice about the vacancies . The only way these effects could be addressed using continuum elastic theory would be to construct empirical models that would extend the standard finite element analysis. However, this type of model simply does not afford the confidence necessary to push the frontiers of device design. Our methodology, coupling of length scales, is able to make definitive predictions of device performance.
## 5 Work in Progress
We are in the process of setting up the micro-gear simulation. We have chosen to study silicon gear teeth because (a) a good TB parameterization exists and (b) because gear teeth are often made of polysilicon. The issues to be investigated are: (a) When two clean surfaces are brought together they “cold weld” - thus when two teeth are in contact, bonds will form across the opposing interfaces. As the gears rotate, these bonds will break. How much matter is transferred? How rough is the newly exposed surface? (b) Suppose we affix linear polymer chains to the surfaces of the gears (self-assembled monolayers) of length (say) 12 carbon atoms, does this significantly reduce wear and friction? The monolayer molecules in question are the alkyltrichlorosilanes. Interatomic potentials are available for similar polymeric systems . Technologically, such processing of the surfaces is very doable , and in fact has already been applied to micro-gears . How much energy is dissipated into the body of the gears when such surfaces rub? At what speed can we run a gear train before entanglement and relaxation times in the SAM polymer become an issue?
## 6 Conclusion
The algorithmic and computational avenue applied here represents a significant departure from the usual finite element approach based on continuum elastic theory. We have shown that atomistic simulation, and in particular multiscale atomistic simulation, offers significant improvements for the modeling of MEMS on sub-micron length scales. It is at these scales that some of the assumptions of continuum mechanics fail. The issue is at what system size and in what way they fail. These are issues that we are able to answer unambiguously using atomistic simulation and coupling of length scales.
## ACKNOWLEDGMENTS
This work was supported by ONR and DARPA. Supercomputing resources were provided through a DOD HPC ‘Grand Challenge’ award at the Maui and ASC Supercomputing Centers.
|
no-problem/9812/astro-ph9812402.html
|
ar5iv
|
text
|
# Progress in Model Atmosphere Studies of WR stars
## 1. Introduction
It is only through understanding the physics of massive stars, their atmospheres, radiation, and evolution, that we will be able to make progress in many aspects of astrophysics. Particularly important is the quantitative study of young starbursts, which are dominated by the effects of O-type and Wolf-Rayet (WR) stars. WR stars comprise only 10% of the massive stellar content in the Galactic mini-starburst, NGC 3603, yet they contribute 20% of the total ionizing flux and 60% of the total kinetic energy injected into the ISM (Crowther & Dessart 1998). In order that young starbursts can be properly studied, both nearby, and at high-redshift, it is crucial that the properties and evolution of O-type and WR stars spanning a range of initial metallicities are determined. In this review, I will consider recent theoretical progress in WR analyses that has been made towards this goal, focusing especially on spectroscopic and ionizing properties.
## 2. Quantitative spectroscopy of WR stars
Quantitative analysis of W-R stars represents a formidable challenge, since their stellar winds are so dense that their photospheres are invisible, so that the usual assumptions of plane-parallel geometry and local thermodynamic equilibrium (LTE) are wholly inadequate. A minimum requirement is to consider non-LTE in an extended, expanding atmosphere for multi-level atoms. At present, three independent model atmosphere codes are capable of routinely analysing the spectra of WR stars, considering complex model atoms of hydrogen, helium, nitrogen, carbon and oxygen, developed by W.-R. Hamann (Potsdam), W. Schmutz (Zurich) and D.J. Hillier (Pittsburgh), the latter also implemented by P.A. Crowther (London) and F. Najarro (Madrid). Each code solves the radiative transfer problem in the co-moving frame, subject to statistical and radiative equilibrium, including the effects of electron scattering and clumping. Overall, consistency between results from these codes is very good. Schmutz and Hillier have also accounted for line blanketing by heavy elements.
Individual calculations are computationally demanding, so that a large parameter space can not be readily explored. Consequently, computationally quick codes have been developed which solve the transfer problem in the Sobolev approximation and assume a wind temperature distribution (e.g. de Koter et al. 1997; Machado, these proc.) De Marco et al. (these proc.) compare the predictions of the code by de Koter with that of Hamann for WC stars.
## 3. Progress in the determination of stellar properties
In the simplest case, the stellar properties of WR stars are derived from the following diagnostics: two spectral lines from adjacent ionization stages of helium (He ii $`\lambda `$5412 and He i $`\lambda `$5876 most commonly), plus the absolute magnitude in a standard filter (typically $`M_v`$) and the terminal wind velocity ($`v_{\mathrm{}}`$), often measured from UV resonance lines. The default number of model parameters available is therefore four, $`R_{}`$, $`T_{}`$, $`\dot{M}`$ and $`v_{\mathrm{}}`$. Stellar temperatures for extended atmospheres are related to the inner boundary of the model atmosphere (generally around Rosseland optical depth $`\tau _{\mathrm{Ross}}`$$``$20), which often deviates significantly from the ‘effective’ temperature, at $`\tau _{\mathrm{Ross}}`$=2/3 (Hamann 1994). Schmutz et al. (1989) identified the so-called transformed radius ($`R_t`$), a measure of wind density, which relates $`R_{}`$, $`\dot{M}`$ and $`v_{\mathrm{}}`$ so that almost identical spectra are produced for fixed $`R_t`$, reducing the number of free parameters. In this way, a large number of WR stars in the Galaxy and Magellanic Clouds have been analysed by Hamann and co-workers, comparing observed line equivalent widths to interpolations of large model grids (see Hamann, these proc.). However, actual WR stars are not pure helium stars, so that the contribution of other elements is necessary.
### 3.1. The effect of including metals
It was soon established that the wind properties of WR stars were affected by the presence of metals, notably carbon and nitrogen (Hillier 1988; 1989). For WN stars, metals are trace elements ($``$1% in nitrogen by mass), so pure He analyses compare relatively well with those additionally including carbon and nitrogen, which control the outer wind properties. In late-type WN (WNL) stars hydrogen contributes significantly (up to $``$50% in extreme cases). Determination of atmospheric contents requires detailed analysis of individual stars through a comparison between theoretical line profiles (e.g. H$`\alpha `$ for hydrogen) and spectroscopic observations (e.g. Crowther et al. 1995a). In WC stars, it was soon realised that pure He studies were inadequate to obtain reliable stellar properties, since carbon mass fractions are $``$40–50% (Hillier 1989). WC analyses need to use He and C diagnostics simultaneously in order that the stellar and chemical properties may be determined. The degree of complexity in atomic data handled for carbon has a great influence on predicted line strengths (Hillier & Miller 1999).
This technique has a major disadvantage in that it is time consuming and computationally expensive. Nevertheless, derived stellar properties can be considered robust, while model deficiencies may be readily identified. Studies of individual stars have now been performed for Galactic, Magellanic Cloud and other Local Group WR stars using 4m class ground-based telescopes (e.g. Smith et al. 1995 in M33). Source confusion becomes problematic at large distances from ground-based observations (0.<sup>′′</sup>8 corresponds to a scale of 3 parsec at the modest distance of M33).
### 3.2. Choice of spectral diagnostics
Although helium and carbon diagnostics are combined to derive the properties of WC stars, the majority of WN studies use solely helium. Early results for weak-lined, early-type WN (WNE) stars led to surprisingly low stellar temperatures (e.g. Hamann et al. 1995), which were comparable with WNL stars instead of strong-lined WNE stars. Crowther et al. (1995b) demonstrated that the stellar temperatures of weak-lined WNE stars are in line with strong-lined examples, if nitrogen diagnostics (e.g. N iv $`\lambda `$4058, N v $`\lambda `$4603–20) are used instead of helium. Results from helium are more straightforward, since the availability and quality of its atomic data is superior to nitrogen. However, He i lines are typically formed at large radii from the core, and are extremely weak in hot WR stars, with the exception of 10830Å that is observationally challenging. In contrast, N iv-v lines originate from the inner wind and are readily observed. Consequently, nitrogen ought to serve as a more sensitive diagnostic of the stellar temperature, and circumvent the problem identified by Moffat & Marchenko (1996). They noted that stellar radii derived from He analyses were greater than the orbital radii of some short period WR+O binaries. Discrepancies for WN stars are not restricted to He and nitrogen diagnostics (see e.g. Crowther & Dessart 1998).
Hamann & Koesterke (1998a) have recently re-analysed a large sample of Galactic WN stars and arrived at similar conclusions to Crowther et al. (1995b). In Fig. 1, the stellar temperatures and luminosities of WN stars obtained by alternative helium and nitrogen diagnostics are compared. Consistent results are obtained for WNL stars, while higher temperatures and luminosities are obtained from nitrogen diagnostics for WNE stars, especially weak-lined stars. Differences in derived temperatures may be large, increasing by a factor of up to two (from 36kK to 71kK for the WN5(h) star WR49). The change in luminosity is greater still – increasing by a factor of six in this star because of the sensitive dependence of bolometric correction (B.C.) with temperature. Parameters derived from helium or nitrogen lines should be fully consistent, so that discrepancies indicate that something is missing in current models. Perhaps clumps in the wind affect the ionization balance in the He i line forming region of WNE stars – this may also be relevant to the (poorly predicted) strength of He i P Cygni absorption components. Whatever the cause, care should be taken when comparing results obtained from different spectral diagnostics.
Detailed analyses of WC-type stars have also been carried out (e.g. Koesterke & Hamann 1995; Gräfener et al. 1998) using helium, carbon and occasionally oxygen diagnostics. However, the additional number of free parameters (C/He, O/He) has restricted the sample analysed to date, and oxygen diagnostics lie in the near-UV, requiring space based observations (Hillier & Miller 1999). Since the UV and optical spectra of WC stars are dominated by overlapping broad emission lines, it is difficult to assign suitable continuum windows. Analyses typically consider the continuum and line spectra in isolation, namely that interstellar extinctions are obtained by matching continua to de-reddened observations, while theoretical line profiles are compared to normalized spectra. A less error-prone approach is the comparison between de-reddened fluxed observations and synthetic spectra, accounting for line overlap. In this way, erroneously defined continuum windows (e.g. at He ii $`\lambda `$5412, Hillier & Miller 1999), and unusual UV extinction laws, may be identified.
### 3.3. IR analyses
Studies discussed above rely exclusively on optical (or occasionally UV) spectral diagnostics. The first infra-red (IR) spectroscopy of WR stars was obtained by Williams (1982), although recent advances mean that this wavelength region can now be used to observe a large sample of WR stars, particularly those obscured at shorter wavelengths. Crowther & Smith (1996) have assessed the reliability of IR analyses of WR stars by studying two WNE stars for which UV and optical data sets were also available. They found that results from exclusively near-IR observations were in good agreement with optical studies, and with later Infra-red Space Observatory (ISO) spectroscopy for WR136 (R. Ignace, priv. comm.). Bohannan & Crowther (1999) have recently compared optical and IR analyses of Of and WNL stars.
Quantitative IR studies of WN-like stars at our Galactic Centre have recently been presented (e.g. Najarro et al. 1997a). Unfortunately, the majority of these stars are relatively cool, so that the sole K-band He ii diagnostic at 2.189$`\mu `$m is unavailable. Without a second ionization stage, a unique temperature may not be obtained, so that mass-loss rates and abundances are uncertain. Dessart et al. (these proc.) attempt to solve this by using the stronger He ii 3.09$`\mu `$m line in the thermal IR as a temperature diagnostic. Another limitation with the K-band is that the prominent He i line at $`\lambda `$2.058$`\mu `$m is strongly affected by (metallicity dependent) line blanketing effects, as shown by Crowther et al. (1995a, 1998). Problems with the quantitative analysis of low temperature stars are neatly summarised by Hillier et al. (1999) for the Galactic early B-type supergiant HDE 316285. They obtained a wide range of possible mass-loss rates and surface H/He abundances for this star, despite the availability of high quality optical and near-IR spectroscopy.
## 4. Relaxing the standard assumptions
Model calculations so far discussed use $`R_{}`$, $`T_{}`$, $`\dot{M}`$ and $`v_{\mathrm{}}`$, plus elemental abundances as free parameters. However, observational evidence suggests that presently assumed quantities, such as the velocity law and homogeneity may be inappropriate. In addition, it is well known that line blanketing by thousands of transitions in the ultraviolet (UV) and extreme ultraviolet (EUV) need to be incorporated into calculations. Each additional relaxation adds (at least) one new variable parameter to the existing set. Consequently, of the several hundred WR stars thus far analysed quantitatively, to date studies of only two have included assorted elements, a variety of velocity laws, clumping and line blanketing (Schmutz 1997; Hillier & Miller 1999).
### 4.1. Variations in velocity law
Generally, a uniform form of the radial velocity field ($`vv_{\mathrm{}}(1R/r)^\beta `$) is assumed, of exponent $`\beta `$=1. Tailored analysis are required to test alternative laws. Unfortunately, different velocity forms are frequently able to reproduce the observed spectrum equally well (Hillier 1991a). In some cases, specific exponents produce optimum agreement, provided with a suitably large range of spectroscopic observations. From a careful comparison of the optical and far-red appearance of the Luminous Blue Variable (LBV) P Cygni, Najarro et al. (1997b) found that a $`\beta `$=4.5 law provided the best match. Including mid-IR ISO observations led to a revision to $`\beta `$=2.5 (Najarro et al. 1998). Unfortunately, a long wavelength observational baseline is rarely available. Schmutz (1997) went a stage further by deriving the form of the velocity law in WR6 from hydrodynamics, at least in the outer visible part of the wind, with $`\beta `$=3. As a indication of the reliability of this approach, the emission profile of He i $`\lambda `$10830 was reproduced better than in previous studies.
### 4.2. Wind inhomogeneities and departures from spherical symmetry
WR winds are known to be inhomogeneous, from both observational and theoretical arguments (Willis, these proc.). However, homogeneity has been assumed by the majority of atmosphere studies to date. Consideration of electron scattering – causing a frequency redistribution of line photons – provides the key to spectral synthesis (Hillier 1984; 1991b). Homogeneous models often overestimate the strength of electron scattering wings relative to the overall emission line intensity. Since free-free emission and radiative recombination both scale as the square of the density, whereas the electron scattering opacity scales linearly with density, estimates of wind inhomogenities may be estimated by varying volume filling factors and mass-loss rates so that line profiles and electron scattering wings are simultaneously reproduced. In line transfer calculations performed to date, several simplifying assumptions are made, namely that models are composed of radial shells of material, with no inter-clump medium. The variation of clumping factor with radius taken into consideration in some cases since radiative instabilities are not expected to be important in the inner wind.
Schmutz (1997) and Hillier & Miller (1999) have estimated mass-loss rates of WR6 and WR111 which are a factor of 3–4 lower than those resulting from homogeneous models. Hamann & Koesterke (1998b) have also applied an identical approach to a sample of four WR stars, with similar results obtained. Substantially lower mass-loss rates of WR stars has importance in evolutionary model calculations and in reducing the momentum (alternatively opacity) problem of driving WR winds (Gayley et al. 1995).
To date, all spectroscopic studies have assumed spherical symmetry. Evidence from spectropolarimetry indicates that this is appropriate for $``$85% of cases (Harries et al. 1998). For the remaining stars, density ratios of 2–3 are implied from observations. Future calculations will need to consider departures from spherical symmetry. Indeed, the wind of the prototypical WNE star WR6 is grossly asymmetric.
### 4.3. Influence of line blanketing
Observations of the forest of iron lines in the UV spectra of WR stars, demonstrate the large influence that line blanketing by Fe-group elements has on the emergent spectrum. The neglect of blanketing reveals itself through inconsistencies of model fits, and results from comparison with ionized nebulae. The principal problem in accounting for line blanketing is being able to handle the effect of tens of thousands of line transitions in the radiative transfer calculations. To date, solely Schmutz and Hillier have made allowance for blanketing, albeit using different techniques, each with their own advantages and disadvantages. Monte Carlo sampling by Schmutz (1997) allows the opacity of a huge number of lines to be considered, although spectral synthesis of individual features in the UV is not possible, while the reverse is true for Hillier & Miller (1998) who use a ‘super-level’ approach, treating the transfer problem correctly.
In Fig. 2 models for a WCE star are compared, in which increasing number of elements are included, He, C, O, and Fe. Carbon and oxygen have a considerable effect on the UV and optical energy distribution of the models, with Fe modifying the energent UV flux distribution (Hillier & Miller 1998, 1999). What effect does allowing for clumping and line blanketing have on the resulting stellar properties? In Table 1 the results of Schmutz (1997) and Hillier & Miller (1999) for WR6 (WN4b) and WR111 (WC5) are compared with earlier studies. Stellar temperatures and bolometric luminosities of the blanketed analyses are considerably greater than those from unblanketed models, with a significant EUV excess (and corresponding increase of B.C.), while mass-loss rates are significantly lower, as a result of considering clumped winds. For the case of WNL stars, Crowther et al. (1998) and Herald et al. (these proc.) find that blanketing has a minor influence on stellar temperatures (though the EUV energy distribution is affected). This result is in apparent contradiction with the analysis of LMC WN9–11 stars by Pasquali et al. (1997) using grids of line blanketed models. Pasquali et al. revealed considerably higher temperatures relative to earlier unblanketed tailored analyses (Crowther et al. 1995a; Crowther & Smith 1997). Subsequent tailored spectroscopic analyses including blanketing by Pasquali (priv. comm.), agree well with the parameters obtained by Crowther & Smith.
Our ability to synthesise individual and groups of Fe lines in the spectrum of WR stars suggests that they can be used to derive Fe-group abundances. UV spectra of O stars (Haser et al. 1998) and optical spectra of A supergiants (McCarthy et al. 1995) have previously been used to determine Fe-contents in extra-galactic environments, though few detailed attempts have been made using WR stars (see Hillier & Miller 1999). As an indication of the potential for the future, Crowther et al. (1999) have recently used Hubble Space Telescope (HST) spectroscopy of the erupting LBV V1 in the giant H ii region NGC 2363, within the Magellanic irregular galaxy NGC 2366 (3.5 Mpc) to determine its Fe-abundance.
## 5. What are the ionizing spectra of Wolf-Rayet stars?
The ionizing flux distribution of WR stars has importance in the study of extra-galactic regions containing young massive stars (giant H ii regions, WR galaxies etc.). Recent results for O stars incorporating non-LTE and wind effects have resulted in improved agreement with observations of associated H ii regions (e.g., Stasińska & Schaerer 1997). It is equally important that suitable ionizing distributions for WR stars are used, which affect determinations of IMFs and ages. Since the Lyman continuum distributions of WR stars is not directly visible (due to absorption by intervening hydrogen), indirect methods need to be used to verify current models.
### 5.1. Nebulae as probes of the Lyman continuum flux distribution
The principal work in this field was that of Esteban et al. (1993) who combined pure helium, unblanketed WR model fluxes (Schmutz et al. 1992) with observed properties of WR ring nebulae, to investigate the properties of the central stars. Esteban et al. varied stellar temperatures until agreement was reached between the observed and predicted nebular properties. Comparisons with (independent) stellar analyses of the central stars was found to be reasonable, except that lower temperatures were required from the photo-ionization models for WNL stars. Recently, Crowther et al. (1998) and Pasquali et al. (these proc.) have returned to this technique, newly considering the influence of line blanketing using the Hillier and Schmutz codes. They depart from Esteban et al. in that ionizing flux distributions obtained from a stellar analysis of the central star are used in the photo-ionization modelling.
Crowther et al. (1998) compared line blanketed and unblanketed flux distributions resulting from stellar analyses of the WN8 star WR124, with observations of its associated nebula, M1–67. They found that the blanketed model predicted the nebular temperature and ionization balance much better than the unblanketed case. Allowance for improved nebular properties of M1–67 from Grosdidier et al. (1998), particularly the radial density distribution, leads to even better agreement with observations. Pasquali et al. (these proc.) find good agreement between the predicted and observed nebular properties of NGC 3199, using stellar flux distributions from analyses of its central WNE star WR18 with the Schmutz and Hillier codes. Unfortunately, the observed properties ($`T_e`$, $`N_e`$, $`\mathrm{\Delta }R`$, abundances etc.) of most WR nebulae at present are insufficiently well determined to use as tests of stellar models.
### 5.2. The effect of blanketing on ionizing fluxes
Overall, spectral synthesis and photo-ionization modelling results give us confidence in the validity of current line blanketed Wolf-Rayet codes. Since the only generally available WR models are unblanketed, pure helium energy distributions of Schmutz et al. (1992), how do new results compare? The calculation of a large multi-parameter grid of line blanketed models is a formidable computational challenge. For the moment, I have obtained models for WR stars with the Hillier & Miller (1998) code, varying solely temperatures (30 to 150kK). WN models span WN4 to WN9 spectral types and include the effects of complex model atoms of H i, He i-ii, C ii-iv, N ii-v, Si iii-iv and Fe iii-vii. In Fig. 3, selected synthetic UV, optical and near-IR spectra are presented. Similar calculations for WC stars spanned WC4 to WC9 and included He i-ii, C ii-iv, O ii-vi and Fe iii-vii in detail. Their predicted Lyman continuum distributions are fairly soft in all cases, with negligible emission above the He<sup>+</sup> edge at 54eV.
In Fig. 4 the ionizing fluxes of these models in the H<sup>0</sup> and He<sup>0</sup> continua (in units of photon s<sup>-1</sup> cm<sup>-2</sup>) are compared with recent solar metallicity O-star models (Schaerer & de Koter 1997), plus the pure helium Schmutz et al. (1992) models. The line blanketed WN flux distributions support the pure helium Schmutz et al. (1992) predictions, although the additional blanketing from C and O in WC stars produces a softer ionizing spectrum at an identical temperature, with negligible flux emitted $`\lambda `$300Å. WR stars also compare closely with comparable temperature O stars in their ionizing flux per unit area.
### 5.3. The effect of wind density
Schmutz et al. (1992) stress the importance of stellar wind density on the ionizing flux distributions of WR stars, such that emission at $`\lambda `$228Å relies on the WR wind being relatively transparent. Denser winds, such as those presented above for representative Galactic WR stars, destroy photons beyond this edge. To illustrate this, additional calculations have been performed for lower wind densities. Although a mass-loss versus metallicity ($`Z`$) scaling for WR stars has not been identified, let us assume that their winds are radiatively driven with a dependance of $`\dot{M}Z^{0.5}`$ (as obtained for radiatively driven O-type stars by Kudritzki et al. 1989).
I have taken the 150kK WC model, whose synthetic spectrum approximates a WC4-type star, and solely reduced its mass-loss rate (by a factor of three) and Fe-content (by a factor of ten). The optical and ionizing spectrum of the low wind density model are compared with the WC4 model in Fig. 5, revealing a harder flux distribution (increasing the B.C. by 1.2 mag), and a dramatic change in the emergent optical spectrum. O vi emission is very strong so the low wind density case resembles a WO-type star. Consequently, a modest change in mass-loss rate has a major influence on the ionizing energy distribution and observed spectral appearance. Strong O vi emission in a WR spectrum is connected principally with the wind density, rather than elemental abundance (Smith & Maeder 1991 identified WO stars as the chemically evolved descendants of WC stars). In WC4 stars, the high wind density and consequently very efficient wind cooling means that O<sup>6+</sup> recombines to O<sup>5+</sup> and subsequently O<sup>4+</sup> interior to the optical line formation region, producing observed O iv-v lines. The less efficient cooling of WO winds, through a lower wind density (because of lower mass-loss rates and higher wind velocities) permits O<sup>6+</sup> recombination in the optical line formation region, producing strong O vi emission. In support of this, recall that WO stars outnumber the WC population at low metallicities (SMC, IC1613).
### 5.4. Nebular He ii $`\lambda `$4686 and bolometric corrections
For my second case, I have taken the earlier 130kK WN model, with a synthetic spectrum of a strong-lined WN4 star, and reduced its mass-loss rate by a factor of ten (scaling its metal content to 0.01$`Z_{}`$). The resulting optical spectrum would be classified as a weak-lined WN2 star, as shown in Fig. 6. Its ionizing flux distribution is extremely hard, with a very strong flux above 54eV ($``$40% of its entire luminosity!). If mass-loss rates of WR stars are driven by radiation pressure, their spectral appearance and ionizing properties will be very sensitive to metallicity. The low metallicity WR model presented here may have application in very metal-poor starbursts, such as I Zw 18 which is thought to contain WR stars (de Mello et al. 1998).
The above results suggest that solely hot WR stars with weak winds produce a significant flux in the He<sup>+</sup> continuum, most likely at low metallicities. This is supported by the known sample of WRs whose nebulae show strong He ii $`\lambda `$4686 emission by Garnett et al. (1991), namely WR102 (WO, Galaxy), Brey 2 (WNE, LMC), Brey 40a (WNE+O, LMC), AB7 (WNE+O, SMC), DR1 (WO, ICI613). Young, low metallicity starbursts would be expected to exhibit strong nebular He ii $`\lambda `$4686 emission from WR stars, in contrast with high metallicity starbursts.
For the grid of high wind density models, representative of strong-lined Galactic WR stars, B.C’s in the range $``$2.6 to $``$4.4 mag (WN), and $``$3.1 to $``$4.6 mag (WC) are obtained. Since wind density affects the ionizing spectrum of WR models, bolometric corrections are also affected. B.C’s for the WO and WN2 models are much higher and very wind density sensitive, ($``$5.8 and $``$7.1 mag, respectively). Smith et al. (1994) used observations of clusters in the Galaxy to estimate WR masses and B.C’s, namely $``$4.5 mag for WC stars, and $``$4 to $``$6 mag for WN stars, in fair accord with predictions. Massey (these proc.) has repeated this for the LMC, and finds B.C’s of $`6`$ to $``$8 mag for cluster WNE stars. From calculations performed here, such stars would be expected to have low wind densities and emit strongly above 54eV. Detailed analysis of individual LMC WNE stars are sought in order to verify these predictions.
Overall, I have discussed the techniques used to derive stellar and chemical properties of WR stars, highlighting the importance of clumping, line blanketing on derived parameters, and the role of wind density and metallicity on the emergent spectrum and ionizing properties of WR stars.
#### Acknowledgments.
I would like to thank my collaborators, especially John Hillier. Bill Vacca brought the importance of reliable ionizing fluxes of WR stars to my attention. PAC is a Royal Society University Research Fellow.
## References
Bohannan, B., Crowther, P.A. 1999, ApJ 511 (Jan 20th) in press
Crowther, P.A., Smith, L.J. 1996, A&A 305, 541
Crowther, P.A., Smith, L.J. 1997, A&A 320, 500
Crowther, P.A., Dessart, L. 1998, MNRAS 296, 622
Crowther, P.A., Hillier, D.J., Smith, L.J. 1995a, A&A 293, 403
Crowther, P.A., Smith, L.J., Hillier, D.J. 1995b, A&A 302, 457
Crowther, P.A., Bohannan, B., Pasquali, A. 1998, in: Proc. ‘Boulder-Munich II: Properties of Hot, Luminous Stars’, (Howarth I.D. ed.), ASP Conf. Series, 131, p.38
Crowther, P.A., Drissen, L., Smith, L.J. et al. 1999, in: Proc. Unsolved Problems in Stellar Evolution, CUP in press
Esteban, C., Smith, L.J., Vílchez, J.M., Clegg, R.E.S. 1993, A&A 272, 299
Garnett, D.R., Kennicutt, R.C., Chu, Y-.H., Skillman, E.D. 1991, PASP 103, 850
Gayley, K.G., Owocki, S.P., Cranmer, S.R. 1995, ApJ 442, 296
Gräfener, G., Hamann, W.-R., Hillier, D.J., Koesterke, L. 1998, A&A 329, 190
Grosdidier, Y., Moffat, A.F.J., Joncas, G., Acker, A. 1998, ApJ 506, 127
Hamann, W.-R. 1994, Space Sci. Rev., 66, 237
Hamann, W.-R., Koesterke, L., Wessolowski, U. 1995, A&A, 299 151
Hamann, W.-R., Koesterke, L. 1998a, A&A 333, 251
Hamann, W.-R., Koesterke, L. 1998b, A&A 335, 1003
Harries, T.J., Hillier, D.J., Howarth, I.D. 1998 MNRAS 296, 1072
Haser, S.M., Pauldrach, A.W.A., Lennon, D.J. et al. 1998, A&A 330, 285
Hillier, D.J. 1984, ApJ 280, 744
Hillier, D.J. 1988, ApJ 327, 822
Hillier, D.J. 1989, ApJ 347, 392
Hillier, D.J. 1991a, in: Proc IAU Symp 143, Wolf-Rayet stars and Interrelations with other Massive Stars in Galaxies, (K.A. van der Hucht, B. Hidayat eds.), Kluwer, p.59
Hillier, D.J. 1991b, A&A 247, 455
Hillier, D.J., Miller, D.L. 1998, ApJ 496, 407
Hillier, D.J., Miller, D.L. 1999, ApJ in press
Hillier, D.J., Crowther, P.A., Najarro, F., Fullerton, A.W.A. 1999, A&A in press
Koesterke, L., Hamann, W.-R., Wessolowski, U. 1992, 261, 535
Koesterke, L., Hamann, W.-R. 1995 A&A 299, 503
de Koter, A., Heap, S.R., Hubeny, I. 1997, ApJ 477, 792
Kudridzki, R.-P., Pauldrach, A.W.A., Puls, J., Abbott, D.C. 1989, A&A 219, 205
McCarthy, J.K., Lennon, D.J., Venn, K.A. et al. 1995, ApJ 455, L135
de Mello, D.F., Schaerer, D., Heldman, J., Leitherer, C. 1998, ApJ in press
Moffat, A.F.J., Marchenko, S.V. 1996, A&A 305, L29
Najarro, F. Krabbe, A., Genzel, R., et al. 1997a, A&A 325, 700
Najarro, F., Hillier, D.J., Stahl, O. 1997b, A&A 326, 1117
Najarro, F., Kudritzki, R.-P., Hillier, D.J., et al. 1998 in: Proc. ‘Boulder-Munich II: Properties of Hot, Luminous Stars’, (Howarth I.D. ed.), ASP Conf. Series, 131, p.357
Pasquali, A., Langer, N., Schmutz, W. et al. 1997, ApJ 478, 340
Schaerer, D., de Koter, A. 1997, A&A 322, 615
Schmutz, W. 1997, A&A 321, 268
Schmutz, W., Hamann, W.-R., Wessolowski, U. 1989, A&A 210, 236
Schmutz, W., Leitherer, C., Gruenwald, R. 1992, PASP 104, 1164
Smith, L.F., Maeder, A. 1991, A&A 241, 77
Smith, L.F., Meynet, G., Mermilliod, J-.C. 1994, A&A 287, 835
Smith, L.J., Crowther, P.A., Willis, A.J. 1995, A&A 302, 830
Stasińska, G., Schaerer, D. 1997, A&A 322, 615
Williams, P.M. 1982, in: Proc IAU Symp. 99, Wolf-Rayet Stars : Observations, Physics, Evolution, (de Loore, C.W.H., Willis, A.J., eds.) Reidel, Dordrecht, p. 73
|
no-problem/9812/math9812006.html
|
ar5iv
|
text
|
# On the cohomology rings of Hamiltonian T-spaces
## 1. Introduction
The classification of manifolds equipped with group actions presents difficulties beyond those inherent in the study of manifolds per se. Even the basic questions—What is the equivariant cohomology ring of the manifold? What can be said about the fixed manifolds of the group action? What is the cohomology ring of the quotient?—turn out to be delicate and involved.
Much more is known in the case of a symplectic manifold $`(M^{2m},\omega )`$ equipped with a Hamiltonian action of a torus $`T`$. Let $`H_T^{}(M)`$ denote the rational equivariant cohomology ring of $`M`$.
The following theorem of F. Kirwan relates the equivariant cohomology of $`M`$ with the equivariant cohomology of its fixed point set:
###### Theorem 1.1.
(Kirwan)\[K\] Let a torus $`T`$ act on a compact symplectic manifold $`(M,\omega )`$ in a Hamiltonian fashion. Denote the fixed point set of the action by $`F`$. The natural inclusion $`i:FM`$ of the fixed point set in the manifold induces an injection $`i^{}:H_T^{}(M)H_T^{}(F)`$.
A related result, also due to Kirwan, relates the equivariant cohomology of $`M`$ and the cohomology of the symplectic quotient of $`M`$.
###### Theorem 1.2.
(Kirwan)\[K\] Let a torus $`T`$ act on a compact symplectic manifold $`(M,\omega )`$ with moment map $`\mu :M𝔱^{}=\mathrm{Lie}(\mathrm{T})^{}`$. Suppose that $`\xi 𝔱^{}`$ is a regular value of $`\mu `$ and let $`M_\xi :=\mu ^1(\xi )/T`$ be the symplectic quotient of $`M`$. The inclusion map $`K:\mu ^1(\xi )M`$ induces a surjection $`K^{}:H_T^{}(M)H_T^{}(\mu ^1(\xi ))=H^{}(M_\xi )`$.
These two results give rise to two natural questions. What is the image of $`i^{}`$? And what is the kernel of $`K^{}`$? The purpose of this paper is to address the first question. In a companion paper \[TW\], we answer the second.
The description we give of the image of the injective map $`i^{}`$ is due to Goresky, Kottwitz, and MacPherson \[GKM\], using the work of Chang and Skjelbred (\[CS\]; see also \[AP, BV, H\]).
###### Definition 1.3.
Let $`NM`$ denote the subset of $`M`$ consisting of points whose $`T`$-orbits are one-dimensional; i.e.,
$$N:=\{pMTpS^1\}$$
By the local normal form theorem, each connected component $`N_\alpha `$ of $`N`$ is an open symplectic manifold whose closure $`\overline{N_\alpha }`$ is a compact, symplectic submanifold of $`M`$. The restriction of $`\mu `$ to $`\overline{N_\alpha }`$ is a moment map for the restricted torus action. Furthermore the closure of $`N`$ is given by $`\overline{N}=NF`$; this is referred to as the one-skeleton of $`M`$.
###### Example 1.4.
Consider the case where $`M^{2m}`$ is a $`2m`$ dimensional toric variety, equipped with the appropriate Hamiltonian action of a torus $`T`$ of rank $`m`$. The image of the moment map is the moment polytope $`\mathrm{\Delta }=\mu (M)`$. Let $`v(\mathrm{\Delta })`$ denote the union of of the vertices of $`\mathrm{\Delta }`$, and $`e(\mathrm{\Delta })`$ the union of the interior of the edges. Then $`F=\mu ^1(v(\mathrm{\Delta }))`$, while $`N=\mu ^1(e(\mathrm{\Delta }))`$.
The main result of this paper is the following
###### Theorem 1.
Let $`(M,\omega )`$ be a compact symplectic manifold equipped with a Hamiltonian action of a torus $`T`$. Let $`F`$ be the fixed point set and let $`\overline{N}`$ be the one-skeleton. Let $`i:FM`$ and $`j:F\overline{N}`$ be the natural inclusions, and $`H_T^{}(M)\stackrel{i^{}}{}H_T^{}(F)`$ and $`H_T^{}(\overline{N})\stackrel{j^{}}{}H_T^{}(F)`$ be the pull-back maps in equivariant cohomology. Then the images of $`i^{}`$ and $`j^{}`$ are the same.
This theorem is proved in considerable generality in \[GKM\]. The purpose of this paper is to give a simple proof of this result in the symplectic setting, which will enable us to obtain a description of the cohomology ring of $`M`$ in a form that will make the structure of the map $`K^{}`$ to the cohomology ring of the symplectic quotient transparent. Our methods should yield similar statements in integral cohomology. Additionally, our methods give an algorithm for turning this description of the cohomology ring into an explicit set of generators and relations.
Before outlining the proof of Theorem 1, let us consider a few special examples.
###### Example 1.5.
If the torus $`T`$ is one-dimensional, $`\overline{N}=M`$, so the theorem is obviously true but trivial.
###### Example 1.6.
Suppose that the closure $`\overline{N_i}`$ of each of the components of $`N`$ is a copy of the two-sphere $`P^1`$. For each such component there exists a corank-one subgroup $`K_iT`$ which acts trivially on $`\overline{N_i}`$. The quotient $`T/K_i`$ is isomorphic to $`S^1`$, and the corresponding action on $`P^1`$ must be the usual action, so that $`H_T^{}(\overline{N_i})`$ is given as follows. Let $`\gamma _i:K_iT`$ denote the inclusion, and let $`\gamma _i^{}:H_T^{}(\mathrm{pt})H_{K_i}^{}(\mathrm{pt})`$ denote the induced map in equivariant cohomology. For each $`i`$, the set $`\overline{N_i}F`$ consists of two points $`n_i,s_i`$; and the image of $`H_T^{}(\overline{N_i})`$ in $`H_T^{}(\overline{N_i}F)=H_T^{}(n_i)H_T^{}(s_i)`$ consists of those elements $`(a,b)H_T^{}(n_i)H_T^{}(s_i)`$ such that $`\gamma _i^{}a=\gamma _i^{}b`$. Let the fixed points of the $`T`$-action be given by $`F_i`$, $`i=1,\mathrm{},N`$. Then the image of $`H_T^{}(M)`$ in $`H_T^{}(F)=H_T^{}(F_i)`$ consists of
$$(a_1,\mathrm{},a_N)H_T^{}(F_1)\mathrm{}H_T^{}(F_N)$$
(1.7)
such that, for each $`\overline{N_i}`$,
$$\gamma _i^{}a_{n_i}=\gamma _i^{}a_{s_i}$$
(See \[GKM\], \[GS\]). This gives a completely combinatorial algorithm for computing $`H_T^{}(M)`$.
The main tool needed to prove Theorem 1, as well as the injectivity and surjectivity theorems 1.1, 1.2, is a repeated use of equivariant Morse theory. The key fact in all these cases is that components of the moment map $`\mu `$ give equivariantly self-perfecting Morse functions whose critical set is precisely $`F`$. As the same is true for each of the $`\overline{N_i}`$’s, similar statements can be made for the one-skeleton $`\overline{N}.`$
These self-perfecting Morse functions give us a very useful way of constructing the equivariant cohomology ring of $`M`$ from the cohomology rings of the fixed manifolds $`F_i`$: roughly speaking, the contribution of each fixed manifold to the cohomology ring of $`M`$ consists of those classes in $`H_T^{}(M)`$ which vanish on all fixed points “below” $`F_i`$, and whose value on $`F_i`$ is a multiple of the downward Euler class of the Morse flow. As a similar statement can be made about the cohomology ring of $`\overline{N}`$, we may compare the images of $`H_T^{}(M)`$ and $`H_T^{}(\overline{N})`$ to prove our result.
## 2. Morse Theory and the Moment Map
In this section we state several results which will be the key steps in the proof of Theorem 1. Among them is Kirwan’s injectivity theorem (of which we supply a proof). All of these results follow directly from the equivariant Morse complex associated to the choice of a Hamiltonian as a Morse function. We note that several of the results of this section have analogs in integral cohomology; however we are only concerned with rational cohomology in this paper.
Let us recall our set-up. Let $`(M,\omega )`$ be a compact symplectic manifold with a moment map $`\mu `$ for the action of a torus $`T^n`$. Let $`FM`$ denote the fixed point set of the torus action. Given a generic element $`\xi 𝔱`$, the function $`f=\mu ,\xi :M𝐑`$ is a Morse function on $`M`$ whose critical set coincides with the fixed point set $`F`$.
Let us consider the fundamental exact sequence corresponding to the Morse function $`f`$. Let us denote by $`C`$ the critical set of $`f`$, and choose $`cC`$. We may assume that an interval $`[cϵ,c+ϵ]`$ contains no critical values of $`f`$ other than $`c`$. Let $`M_c^+=f^1(\mathrm{},c+ϵ)`$, $`M_c^{}=f^1(\mathrm{},cϵ)`$. Then we have the following lemma, which is the main technical fact behind our results:
###### Proposition 2.1.
Let a torus $`T`$ act on a compact manifold $`M`$ with moment map $`\mu :M𝔱^{}`$. Given $`\xi 𝔱`$, choose a critical value $`c`$ of the projection $`f:=\mu ^\xi `$. Let $`F`$ be the set of fixed points, and let $`F_c`$ be the component of $`F`$ with value $`c`$.
The long exact sequence in equivariant cohomology for the pair $`(M_c^+,M_c^{})`$ splits into short exact sequences:
$$0H_T^{}(M_c^+,M_c^{})H_T^{}(M_c^+)\stackrel{k^{}}{}H_T^{}(M_c^{})0.$$
(2.2)
Moreover, the restriction from $`H_T^{}(M_c^+)`$ to $`H_T^{}(F_c)`$ induces an isomorphism from the kernel of $`k^{}`$ to those classes in $`H_T^{}(F_c)`$ which are multiples of $`e_c`$, the equivariant Euler class of the negative normal bundle of $`F_c`$.
Proof: By our assumptions, $`f`$ is a Morse function, and there is a unique critical value of $`f`$ contained in the interval $`[cϵ,c+ϵ]`$. Denote the corresponding critical manifold by $`F_c`$, and the negative disc and sphere bundles of $`F_c`$ by $`D_c`$, $`S_c`$ respectively. The pair $`(M_c^+,M_c^{})`$ can be retracted onto the pair $`(D_c,S_c)`$, so there is an isomorphism
$$H_T^{}(M_c^+,M_c^{})H_T^{}(D_c,S_c)$$
(2.3)
By the Thom isomorphism theorem, we have
$$H_T^{}(D_c,S_c)H_T^{\lambda _c}(D_c)=H_T^{\lambda _c}(F_c)$$
(2.4)
where $`\lambda _c`$ is the Morse index of the critical manifold $`F_c`$; so we obtain a commutative diagram
(2.5)
where $`e_c=e(D_c)`$ is the equivariant Euler class of the bundle $`D_cF_c`$. The cup product map $`e_c`$ is injective; the same therefore is true of the maps $`\delta _c`$ and $`\gamma _c`$, proving the lemma.
The following corollary is then immediate:
###### Corollary 1.
(see \[AB1, AB2, K\]) The function $`f`$ is an equivariantly perfect Morse function on $`M`$.
Another application of this proposition is the proof of the following theorem of Kirwan.
###### Theorem 2.
Let a torus $`T`$ act on a symplectic manifold $`(M,\omega )`$ with proper bounded below moment map $`\mu :M𝔱^{}`$. Let $`i:FM`$ denote the natural inclusion of the set $`F`$ of fixed points. The pullback map $`i^{}:H_T^{}(M)H_T^{}(F)`$ is injective.
###### Proof.
Order the critical values of $`f`$ as $`c_1<c_2<\mathrm{}<c_n`$. The theorem obviously holds for $`f^1(\mathrm{},c_1)=\mathrm{}.`$ Assume the proposition holds for the manifold $`M^{}:=f^1(\mathrm{},c_i)`$. We will show that it will hold for the manifold $`M^+:=f^1(\mathrm{},c_{i+1})`$; the result follows then by induction.
By Lemma 2.1, we have a map of short exact sequences
(2.6)
where $`F_i`$ denotes the critical set with value $`c_i`$. By induction, the inclusion $`i_{}`$ of $`FM^{}`$ into $`M^{}`$ induces an injection in cohomology. By Proposition 2.1 the image of $`H_T^{}(M^+,M^{})`$ in $`H_T^{}(M^+)`$ is embedded injectively in $`H_T^{}(F_i)`$. The theorem then follows by diagram chasing. ∎
## 3. Proof of the Main Theorem
We are now ready to prove our main theorem: the restriction map to the fixed point set induces an isomorphism from the equivariant cohomology of the original Hamiltonian manifold to the image of the equivariant cohomology of the one-skeleton, under its restriction map to the fixed point set. The key idea is to use the tools developed in the last section to compare the graded rings associated to these images using the filtration given by the Morse function obtained by choosing a projection of the moment map. The result will then follow from the naturality of these objects, induction on the critical points, and the injectivity of the restriction map.
Recall that the one-skeleton is given by
$$\overline{N}:=\{pMTp\text{ is one-dimensional or zero-dimensional}\}.$$
Clearly, the image of $`i^{}:H_T^{}(M)H_T^{}(F)`$ is a subset of the image of $`j^{}:H_T^{}(\overline{N})H_T^{}(F)`$. Therefore, $`i^{}`$ induces a map, which we will also call $`i^{}`$, from $`H_T^{}(M)`$ to $`\mathrm{im}j^{}H_T^{}(F)`$. By Theorem 1.1, this map is injective. Therefore, to prove the theorem, it suffices to show that this map is surjective.
On the level of the graded rings associated to the Morse filtration, surjectivity will follow from comparing the proposition below with Proposition 2.1.
###### Proposition 3.1.
Let a torus $`T`$ act on a compact symplectic manifold $`M`$ with moment map $`\mu :M𝔱^{}`$. Given $`\xi 𝔱`$, choose a critical point $`c`$ of the projection $`f:=\mu ^\xi `$. Let $`F`$ denote the fixed point set and let $`F_c`$ denote the component of $`F`$ with value $`c`$. Define $`F^{}:=Ff^1(\mathrm{},cϵ)`$ and $`\overline{N^+}:=\overline{N}f^1(\mathrm{},c+ϵ)`$.
Let $`\eta `$ be a cohomology class in $`H_T^{}(\overline{N^+})`$ which vanishes when restricted to $`H_T^{}(F^{})`$. Its restriction to $`H_T^{}(F_c)`$ is a multiple of $`e_c=e(D_c)`$, the equivariant Euler class of the downward normal bundle $`D_c`$ of $`F_c`$ (in $`M`$).
###### Proof.
Consider any component $`N_\alpha `$ of the set $`N`$ of one-dimensional orbits such that the closure $`\overline{N_\alpha }`$ contains $`F_c`$. The closure $`\overline{N_\alpha }`$ is a smooth $`T`$-invariant symplectic manifold with moment map $`\mu `$. The class $`\eta `$ induces a cohomology class on $`\overline{N_\alpha ^+}:=\overline{N_\alpha }f^1(\mathrm{},c+ϵ)`$ which vanishes when restricted to $`\overline{N_\alpha }F^{}`$, and hence (by injectivity which we need to state in this version), when restricted to $`\overline{N_\alpha ^{}}:=\overline{N_\alpha }f^1(\mathrm{},cϵ)`$. Thus, by Proposition 2.1, any element of the kernel of the natural map $`H_T^{}(N_\alpha ^+)H_T^{}(N_\alpha ^{})`$ is, when restricted to $`H_T^{}(F_c)`$, a multiple of the equivariant Euler class $`e_\alpha `$ (here $`e_\alpha =e(D_c\overline{N_\alpha })`$ is the equivariant Euler class $`e_c`$ of the downward normal bundle $`D_c\overline{N_\alpha }`$ of $`F_c`$ in $`\overline{N_\alpha }`$).
So the restriction of $`\eta `$ to $`F_c`$ is a multiple of the equivariant Euler class of the downward normal bundle of $`F_c`$ in $`N_\alpha `$. Since this holds for each component $`N_\alpha `$, and each of these components must have a different stabilizer, we may apply Lemma 3.2 below. Therefore, the class $`\eta `$ must be multiple of the product of the equivariant Euler classes the negative normal bundles to $`F_c`$ in all the components of $`N`$ whose closure contains $`F_c`$. But this is precisely the equivariant Euler class of the negative normal bundle to $`F_c`$ in $`M`$. ∎
###### Lemma 3.2.
Let a torus $`T`$ act on a complex vector bundle $`E`$ over a manifold $`F`$, so that the fixed set is precisely $`F`$. Decompose $`E`$ into the direct sum of bundles $`E_\alpha `$, where each $`E_\alpha `$ is acted on with a different weight $`\alpha 𝔱^{}`$. Let $`e_\alpha `$ be the Euler class of the sub-bundle $`E_\alpha `$.
Then if $`yH_T^{}(F)`$ is a multiple of $`e_\alpha `$ for each $`\alpha `$, then $`y`$ is a multiple of the product of the $`e_\alpha `$.
###### Proof.
Assume first that $`F`$ is a single point.
Let $`\alpha 𝔱^{}`$ be the weight with which $`T`$ acts on the sub-bundle $`E_\alpha `$. Since $`\alpha `$ is a linear function on $`𝔱`$, it lies naturally in $`H^{}(BT)`$ = $`H_T^{}(F)=\mathrm{Sym}(𝔱^{})`$, the algebra of symmetric polynomials on $`𝔱`$. The equivariant euler class of $`E_\alpha `$ is given by $`e_\alpha =\alpha ^{n_\alpha }`$, where $`n_\alpha `$ is the complex dimension of $`E_\alpha `$. The $`\alpha `$ are distinct by assumption, and non-zero since no point not in the zero section is fixed by $`T`$. Therefore, the $`e_\alpha `$ are pairwise relatively prime. (Recall that every polynomial ring over $`𝐐`$ is a unique factorization domain.)
More generally, since $`F`$ is fixed by $`T`$, $`H_T^{}(F)=H^{}(F)H^{}(BT)`$. Thus, $`H_T^{}(F)`$ is bigraded. In particular, given any integer $`i`$, any cohomology class $`aH_T^{}(F)`$ has a well-defined component $`a_iH^i(F)H^{}(BT)`$, and the sum of all such components is $`a`$ itself; we will call $`a_i`$ the component of $`a`$ with $`F`$-degree $`i`$.
Note that the component of $`e_\alpha `$ with $`F`$-degree $`0`$ is precisely $`\alpha ^{n_\alpha }`$. By the previous discussion, these are non-zero and pairwise relatively prime. Therefore it is enough to prove that if $`e`$ and $`f`$ are two cohomology classes whose components $`e_0`$ and $`f_0`$ with $`F`$-degree zero are relatively prime, and if $`e`$ and $`f`$ both divide $`\alpha `$, then so does $`ef`$. We will prove this by induction.
We claim that if $`e(fw+x)=f(ew+y)`$, where the components of $`x`$ and $`y`$ with $`F`$-degree $`i`$ vanish for all $`i<k`$, then there exist $`x^{},y^{}`$ such that $`e(fw+x^{})=f(ew+y^{})`$, and such that the components of $`x^{}`$ and $`y^{}`$ with $`F`$-degree $`i`$ vanish for all $`i<k+1`$. To see this, compare the component of $`F`$-degree $`k`$ on the two sides of the equation $`e(fw+x)=f(ew+y)`$. Cancelling out terms which appear on both sides, we get $`e_0x_k=f_0y_k`$. Since $`e_0`$ and $`f_0`$ are relatively prime polynomials, this shows that there exists $`z_k`$ such that $`x_k=f_0z_k`$ and $`y_k=e_0z_k`$. ∎
We now proceed to prove that the map $`i^{}:H_T^{}(M)\mathrm{im}j^{}H_T^{}(F)`$ is surjective. We proceed, as usual, by induction:
Consider any critical point $`c`$ of the projection $`f:=\mu ^\xi `$. Define $`M^+:=f^1(\mathrm{},c+ϵ)`$, and $`M^{}:=f^1(\mathrm{},cϵ)`$ for any sufficiently small $`ϵ`$. Let $`\overline{N^+}:=\overline{N}M^+`$, $`\overline{N^{}}:=\overline{N}M^{}`$, $`F^+:=FM^+`$, and $`F^{}:=FM^{}`$. Let $`i^+:F^+M^+`$, $`i^{}:F^{}M^{}`$, $`j^+:F^+\overline{N^+}`$ and $`j^{}:F^{}\overline{N^{}}`$ denote the corresponding inclusion maps. It is enough to assume that the induced map $`i^{}:H_T^{}(M^{})\mathrm{im}j^{}H_T^{}(F^{})`$ is surjective, and to prove that the induced map $`i^+:H_T^{}(M^+)\mathrm{im}j^+H_T^{}(F^+)`$ is also surjective.
Since the images of $`H_T^{}(M^{})`$ and $`H_T^{}(N^{})`$ inside $`H_T^{}(F^{})`$ are the same, it follows that the natural restriction map $`r`$ from $`\mathrm{im}j^+H_T^{}(F^+)`$ to $`\mathrm{im}j^{}H_T^{}(F^{})`$ is surjective. Thus, taking the exact sequence (2.2) of Proposition 2.1, we have a map of short exact sequences
(3.3)
By our inductive assumption, $`i^{}`$ is surjective.
By Proposition 3.1, every element in $`\mathrm{ker}r`$ is a multiple of $`e_c`$, the equivariant Euler class of the negative normal bundle of $`F_c`$, the component of the fixed point set with value $`c`$. On the other hand, by Proposition 2.1, every multiple of $`e_c`$ is in the image of the restriction $`H_T^{}(M^+,M^{})`$ to $`H_T^{}(F_c)`$ Thus, the arrow from $`H_T^{}(M^+,M^{})`$ to $`\mathrm{ker}r`$ is surjective too.
The result follows by a diagram chase.
## 4. Some comments about integral cohomology
We close with some comments about integral cohomology. Unfortunately, the integer version of Theorem 1 is not true in general. In fact even the injectivity theorem 1.1 will not hold in integral cohomology without some assumptions. However, both injectivity and a version of Theorem 1 can be proved over the integers where certain restrictions are placed on the allowable stabilizer subgroups.
Perhaps the easiest example where Theorem 1 is not true for integer cohomology is that of $`S^1\times S^1`$ acting on $`S^2\times S^2`$, with speed two on each sphere.<sup>1</sup><sup>1</sup>1If the reader is disturbed by the fact that this action is not effective, she or he may tack on another couple of $`S^2`$’s spinning at speed one. Using the moment map for the diagonal action as our Morse function, we see that every cohomology class which vanishes outside a neighborhood of the north pole $`\times `$ the north pole must be a multiple of $`4x^2`$ when restricted to that point. In contrast, there exists a cohomology class on the one-skeleton which vanishes outside this neighborhood but is only a multiple of $`2x^2`$. Essentially, the problem is that the weights are not relatively prime. It is easy to place a condition on the stabilizer groups at each fixed point in a way that negates this possibility. This is essentially all that can go wrong, and a version of Theorem 1 can be expected to hold if such an assumption of relative primality is made.
|
no-problem/9812/cond-mat9812092.html
|
ar5iv
|
text
|
# Controlling Josephson transport by manipulation of Andreev levels in ballistic mesoscopic junctions
## 1 Introduction
The ability to control Josephson current in SNS and SIS junctions is certainly of fundamental interest, and may also become important in applications to superconducting electronics with multiterminal devices - superconducting transistors . The fundamental mechanism involves coherent transport of quasiparticles through the normal region between superconducting electrodes, influencing the more or less complex normal region via electrostatic gates, current injection and electromagnetic radiation. From an applied point of view, the goal is to switch the Josephson current with sufficient gain to be able to drive networks of components. In recent experiments on diffusive and ballistic short multiterminal junctions, the Josephson current has been supressed by means of electron injection. All this makes it highly interesting to get a complete picture of nonequilibrium current transport in ballistic superconducting junctions.
In the following we will use the more general terminology of SXS to characterize normal regions X with energy dependent transmission amplitudes for electrons and holes, e.g. describing situations intermediate between SNS and SIS, or entirely different situations involving resonant double barrier structures (Fig. 1).
Within the framework of the 1D Bogoliubov-de Gennes equation a general dispersion equation for the bound Andreev level spectrum $`E=E(\varphi )`$ as a function of the superconducting phase difference $`\varphi `$ can be derived in terms of the normal energy dependent scattering amplitudes of the junction . Knowing the Andreev level spectrum, the Josephson current flowing through the junction through these bound states is then given by
$$I=\underset{bound}{}I(E)n(E),I(E)=\frac{2e}{\mathrm{}}(dE/d\varphi ),$$
(1)
where $`n(E)`$ is the Andreev level filling factor .
Control of dc Josephson current in SXS junctions can be achieved either (i) by modifying the Andreev level structure by electrostatic gates, or (ii) by modifying the quasiparticle distribution and Andreev levels via injection of current or electromagnetic (EM) radiation into the normal region. In order to achieve strong effects and efficient control, it is essential to taylor the superconducting junctions to have Andreev level spectra which are sensitive to external influence. One natural idea is then to utilize various resonance effects.
In this paper we will emphasize two types of coherent resonant transport through Andreev levels in ballistic junctions: (i) coupling of degenerate SN-interface Andreev levels through a tunnel barrier, and (ii) transport through a normal resonant level in a double barrier structure (DBS).
Let us first consider an intermediate junction on the scale of the coherence length, with a barrier long enough to be completely non-transparent when it exceeds the Fermi level. In this case, the gate-controlled long potential barrier is always below the Fermi level, influencing the wave functions over a wide region with many $`k_F`$-oscillations, $`k_FL1`$, with ”free” electron-hole propagation and dephasing in the normal X-region under the gate.
When the gate potential $`V_g`$ increases towards the Fermi level, the Fermi wave vector $`k_F`$, the electron density and the coherence length $`\xi _N`$ decrease. Simultaneously, the electron-hole dephasing increases, and the junction therefore effectively becomes long. This is reflected in the Andreev level structure, the gap successively filling up with Andreev levels, similar to the case of actually increasing the physical length $`L`$ of the SNS junction.
If this barrier is pushed above the Fermi level the transparency will vanish, $`D0`$, and the junction will essentially become divided into two isolated SI systems, each having a localized Andreev surface/interface state sitting in the gap close to each gap edge . If we now decrease the length of the barrier to allow tunneling, $`D1`$ but finite, the degenerate SI interface states begin to interact through the tunnel barrier, and we get an SIS junction.
This then takes us to the next case: an SXS junction with $`N<X<I`$, using a gate-controlled ”tunnel” barrier to tune the junction from SNS to SIS. As long as the top of the low-transparency barrier does not pass above the Fermi level $`\mu `$, the junction will essentially behave as a short SNS junction, with a symmetric pair of dispersing Andreev levels around the Fermi level. Not until the barrier becomes significantly higher than the Fermi level does the junction approach SIS behaviour .
If the barrrier is long enough, for $`D0`$ it can split the Andreev envelope into two degenerate parts localized around the respective SI (or SNI) interfaces - Andreev interface/surface states - a double-well situation. Weak coupling splits the degenerate Andreev levels into two bonding-antibonding levels, with level splitting $`\sqrt{D}`$ \- corresponding to resonance coupling of the superconducting electrodes. The central aspect is that even in SIS junctions the development of the Andreev level structure depends not only on the strength of the barrier (via the transmission probability D), but also on the length of the barrier via electron-hole dephasing.
A second method of controlling dc Josephson current is connected with creating nonequilibrium population of Andreev levels. One way to achive this is by connecting the normal region to a normal reservoir with a shifted chemical potential (terminal 1 in Fig. 1a) . This opens up the junction, broadens the Andreev levels and modifies the wave functions of the Andreev states. Seen from the normal reservoir, the junction looks like a split NS Andreev mirror with resonances. This creates an anomalous interference current which in long junctions is independent of junction length, scales linearly with voltage $`V`$ up to $`\mathrm{\Delta }`$ and saturates at a magnitude that is of the order of the short junction current $`Ie\mathrm{\Delta }/\mathrm{}`$.
Another way to achieve nonequilibrium population of Andreev levels is via optical pumping using EM radiation. Resonant pumping between two Andreev levels will strongly mix the wave functions of these two levels, inducing highly coherent combinations of static Andreev states with pronounced interference effects, leading to Rabi oscillation of the Josephson current . This makes it possible selectively to control the population $`n(E)`$ of the resonant levels. By saturating (equalizing) the level populations one may e.g. induce suppression of the contribution of the resonant levels to the total current.
## 2 Gate-potential control of Andreev levels - JOFET
Takyayanagi and coworkers have recently demonstrated that the dc Josephson current in gate-controlled ballistic S-2DEG-S junctions can be suppressed using a gate voltage of only $`V_g=1V`$, the entire current-voltage characteristics (IVC) transforming from RSJ to ohmic behaviour.
To describe this situation, consider a gate-controlled SNS-type junction with length $`L\xi _N`$. This can be illustrated by Fig. 1b with the DBS removed, denoting the remaining flat potential barrier by $`V_g`$. Since the potential barrier is broad, $`k_FL1`$, it influences the wave functions over a wide region, defining a local Fermi wave vector and a coherence length for the normal region.
To begin with, we set the effective mass $`m^{}=1`$ everywhere, and the Fermi velocities are taken to be equal when $`V_g=0`$ (perfectly matched SNS). The parameters are chosen so that the pure SNS junction ($`V_g=0`$) is fairly close to the short limit ($`\xi _NL`$), with 1-2 bound Andreev levels localized in the N-region.
The coherence length in the normal region is defined as $`\xi _N^1=\delta k=(k^+k^{})`$ where $`k^\pm =\sqrt{2m^{}(\mu V_g\pm E)}`$. In the present case of fixed length L, the electron-hole dephasing is directly controlled by the Fermi vave vector in the gate region. For $`E\mu V_g`$, the dephasing $`\beta =\delta kL`$ in the normal region is given by
$$\beta =EL/\sqrt{2m^{}(\mu V_g)};\mathrm{\hspace{0.33em}\hspace{0.33em}0}<E\sqrt{\mathrm{\Delta }}$$
(2)
When the gate potential increases towards the Fermi level the electron-hole phase difference keeps increasing, allowing more and more Andreev levels to enter the gap from the continuum. This is precisely also what happens if one increases the length of the junction $`L`$ while keeping the potential fixed. Therefore, increasing the gate potential decreases the coherence length and effectiely makes the junction long.
We find that for increasing gate voltage $`V_g`$, the junction remains fairly short (2-3 Andreev levels) until $`V_g/\mu 0.9`$, the critical current $`I_c`$ and the $`I_cR_N`$ product decreasing slowly. However, for increasing gate voltage in the range $`0.9V_g/\mu 1\mathrm{\Delta }/\mu `$, the gap region becomes raidly filled up by an increasing number of Andreev levels, causing the critical current $`I_c`$ and the $`I_cR_N`$ product rapidly to drop, the coherence length $`\xi _N`$ becoming short and the junction effectively becoming long, similar to the dirty case .
When $`1\mathrm{\Delta }/\mu V_g/\mu 1`$, the top of the barrier enters the bottom of the superconducting gap $`\mathrm{\Delta }`$ and approches the Fermi level, gradually depleting the Andreev levels in the gap (symmetrically around the Fermi level), making the critical current $`I_c`$ and the $`I_cR_N`$ product approach zero.
Finally, when $`V_g/\mu 1`$, the barrier passes the Fermi level, the junction goes into the SIS tunnel regime and the critical current $`I_c`$ goes exponentially to zero. Since the barrrier is very long, in practice there is an abrupt cutoff. Here is a difference in principle with the physically long junction where the gap keeps filling up, the bound-state current decaying like $`\xi _N/L`$.
In numerical studies of S-Semiconductor-S junctions , appropriate for the S-2DEG-S JOFET , the small effective mass ($`m^{}=0.04`$) in the 2DEG normal region introduces a big potential barrier already in the absense of gate voltage. As a result there are pronounced normal resonances in the X-region which strongly influence the Andreev level spectrum and the Josephson current , as described in Sect. 3. In particular, the $`I_cR_N`$ product is reduced already for zero gate voltage, $`V_g=0`$, in qualitative agreement with experiment .
## 3 Resonance Josephson coupling in complex junctions
A general question we would like to address is: How can the Andreev level spectrum in an S-X($`V_g`$)-S junction be influenced by a more or less complex barrier structure created by a split-gate arrangement, such as illustrated in Fig. 1?
The different parts of the potential barrier structure (Fig. 1b) are imagined to be controlled by by a split-gate combination such that one can independently vary the broad potential $`V_g`$ (length L) and the position of the resonance in double barrier structure (DBS) via $`V_0`$. While we are at it, we could just as well imagine to be able also to control the barrier heights of the double barrier structure. As a result we can, in principle, control both the position and the width of the normal resonant level in the DBS.
Let us limit our discussion to short junctions. If we assume the narrow normal resonance to be far away from the Fermi level, $`E_0\mathrm{\Delta },\mathrm{\Gamma }`$, the DBS acts a low-transmission tunnel barrier. If the structure is symmetric, the Andreev level energies are given by ($`|\varphi 2\pi n|\sqrt{\mathrm{\Delta }/\mu }`$)
$$E=\pm \mathrm{\Delta }\sqrt{1\left(\frac{\beta }{2}\pm \sqrt{D}\mathrm{sin}\frac{\varphi }{2}\right)^2},$$
(3)
where $`\beta =\beta (E)=\beta _{DBS}+\beta _2+\beta _3`$ is the total electron-hole dephasing due to transmission through the entire normal X-region (the double barrier structure (DBS) plus the two semiconducting or metallic regions; $`\beta _i=2d_i/\xi _N,i=2,3`$).
It follows from Eq. (3) that the current through individual Andreev levels is given by
$$I(E_\pm )=sign(E_\pm )\frac{e\mathrm{\Delta }}{2\mathrm{}}(D\mathrm{sin}\varphi \pm \frac{\beta \sqrt{D}}{2}\mathrm{cos}\frac{\varphi }{2})$$
(4)
If $`\beta /2\sqrt{D}`$, the second term in Eq. (4), originating from two Andreev bands with anomalously large dispersion, $`E(\varphi )\sqrt{D}`$, will dominate the current . Because of this large dispersion, the currents of the individual Andreev bound states will exceed the Ambegaokar-Baratoff critical current , the sum current in Eq. (4), by a factor $`(\beta /2)/\sqrt{D}1`$.
If there are no normal metal or semiconductor regions ($`d_2=d_3=0`$), the junction behaves like a SIS junction with transparency $`D1`$. For simplicity and rigour, let us describe this situation by a clean tunnel barrier . The dephasing parameter $`\beta `$ can then be directly related to the electron-hole wave vector difference $`\delta \kappa =\kappa ^+\kappa ^{}`$, where $`\kappa ^\pm =\sqrt{2m^{}(V_g\mu \pm E)}`$. To illustrate a typical situation, let us assume that the tunnel barrier is about twice as high as the Fermi energy, $`V_g2\mu `$. The dephasing parameter $`\beta `$ is then given by
$$\beta \frac{\delta \kappa }{k_F}1$$
(5)
which finally leads to the very simple condition for Andreev level splitting:
$$\sqrt{D}\frac{\delta \kappa }{k_F}\frac{\mathrm{\Delta }}{\mu }1.$$
(6)
Let us emphasize that the form of the spectrum in Eq. (3) results from the coupling of degenerate Andreev surface/interface states . If the DBS in Fig. 1b is made completely opaque, the junction will split into two isolated SNI/INS interfaces. For $`D=0`$, the energies of the degenerate Andreev SNI surface levels are given by $`E=\pm \mathrm{\Delta }\sqrt{1\left(\beta /2\right)^2}`$, where $`\beta \mathrm{\Delta }/\mu +2d/\xi _N`$. Consider first the extreme SIS case: then $`d=0`$ and $`\beta \mathrm{\Delta }/\mu 1`$, and the Andreev surface levels will be sitting very close to the gap edges. Increasing the width $`d`$, dephasing will be dominated by the normal interface regions: in this case $`\beta =2d/\xi _N`$ can be large, and the Andreev surface state will be pulled down into the gap.
Bringing together the degenerate SI surface states causes resonance tunneling through the barrier and the formation of a two-level system of ”bonding-antibonding” Andreev levels split by $`\mathrm{\Delta }\beta \sqrt{D}`$, the Josephson resonance coupling. The enhanced currents in Eq. (4) will appear as nonequilibrium currents in the junction: in equilibrium they will be hidden due to mutual cancellations.
Bringing the SI/IS interfaces closer together will finally increase the transparency $`D`$ so much that the Josephson coupling will dominate over electron-hole dephasing: the upper level becomes pushed into the continuum and the Andreev level begins to disperse on the scale of $`\mathrm{\Delta }D`$: there will then only be a single level, $`E=\pm \mathrm{\Delta }\sqrt{1D\mathrm{sin}^2\varphi /2}`$ .
Asymmetry in position of the Andreev surface levels due to asymmetry of the junction geometry, $`\delta \beta =\beta _1\beta _2=2|d_1d_2|/\xi _N>\sqrt{D}`$, leads to a modificaiton of the Andreev level energy,
$$E=\pm \mathrm{\Delta }\sqrt{1\left(\frac{\beta }{2}\pm \sqrt{(\delta \beta )^2+D\mathrm{sin}^2\frac{\varphi }{2}}\right)^2},$$
(7)
This type of asymmetry, or asymmetry due to fluctuations of the superconducting order parameter in the electrodes, $`\mathrm{\Delta }_1\mathrm{\Delta }_2>\beta \sqrt{D}\mathrm{\Delta }`$, will destroy the resonance coupling and give back the usual expression $`E=\pm \mathrm{\Delta }\sqrt{1D\mathrm{sin}^2\varphi /2}`$.
The discussion above concerned a narrow level far away from the Fermi level, $`E_0\mathrm{\Gamma }`$, describing non-reonant tunneling through the DBS structure. In the case of a narrow resonance near the Fermi level, $`E_0\mathrm{\Gamma }`$, we instead obtain
$$E=\pm \sqrt{E_0^2+\mathrm{\Gamma }^2\mathrm{cos}^2(\varphi /2)}.$$
(8)
When the resonance approaches the Fermi level, $`E_0=0`$, and the structure becomes completely transparent for normal electrons. The Andreev energy and Josephson current
$$E=\pm \mathrm{\Gamma }\mathrm{cos}(\varphi /2);I=e\mathrm{\Gamma }\mathrm{sin}(\varphi /2)$$
(9)
then have the form typical for an SNS junction, but with $`\mathrm{\Delta }`$ replaced by $`\mathrm{\Gamma }`$.
In a wave function picture, in the SNINS structure the Andreev bound state wave function is localized in the two N-regions, and hybridization through the I-barrier (non-resonant DBS) causes resonant hopping of quasiparticles and determines the bandwidth (Andreev level splitting). When a normal Breit-Wigner resonance in the DBS region is pulled towards the Fermi level, the Andreev bound state wave function moves into the resonant central N-region: resonance hopping of quasiparticles is then suppressed, and the bandwidth is determined by the width of the normal resonance.
Finally, it should be noted that d-wave superconductors with a nodeline along the SI surface normal are characterized by midgap (zero energy) Andreev surface states . In a SIS (or SNINS) junction, these Andreev states will interact, leading to resonance splitting $`\sqrt{D}`$ and resonant transport , as discussed above.
## 4 Non-equilibrium population of Andreev levels via current injection - JOINT
So far we have only considered equilibrium Josephson current in the 1D ballistic SXS junction in Fig. 1, because the injection lead 1 has been unbiased and essentially disconnected.
We now connect the injection lead 1 to a normal reservoir and apply a bias voltage $`V_1=V`$, thus broadening the Andreev levels and creating a nonequilibrium quasiparticle distribution. The assumption is that the normal X-region of the junction will be in equilibrium with the normal reservoir, and therefore in nonequilibrium relative to the superconducting electrodes.
When the junction is open, the Andreev levels not only broaden but also split into two quasibound states associated with injected electrons and holes . The total nonequilibrium current in the junction is given by
$$I=𝑑E\left[i^en^e+i^hn^h+i^sn\right]$$
(10)
where $`i^s`$ is the quasiparticle current injected from the superconducting electrodes, and $`i^e`$, $`i^h`$ are currents of electrons and holes injected from the normal reservoir. These currents are controlled by the reservoir, $`n^{e,h}=n(E\pm eV)`$. To maintain this control, the coupling $`\epsilon `$ between the junction and the reservoir must be sufficiently strong: the lifetime within the junction $`d/(\epsilon v_F)`$ must be smaller than inelastic relaxation time $`\tau _i=l_i/v_F`$, i.e. $`\epsilon >d/l_i`$.
If we introduce sum and difference currents $`i^\pm =i^e\pm i^h`$, the total current may be written as
$$I=I_{eq}+I_{neq}=I_{eq}+I_r+I_a$$
(11)
in terms of the equilibrium current $`I_{eq}=(i^++i^s)n𝑑E`$, the regular nonequilibrium current $`I_r=(i^+/2)(n^e+n^h2n)𝑑E`$ and the anomalous nonequilibrium current $`I_a=(i^{}/2)(n^en^h)𝑑E`$.
In the nonequilibrium current $`I_{neq}`$, the regular term $`I_r`$ describes the effects of nonequilibrium population of the the Andreev levels, while the anomalous term $`I_a`$ describes modification of the Andreev states themselves due to the open injection lead. This modification is not primarily the broadening of the levels: much more important are the changes in the transition amplitudes for electrons and holes, leading to a difference current $`i^{}`$ which alwas has the same sign, as seen in Fig. 2a.
If we assume the normal reservoir to be weakly coupled to the SNS-junction via a tunnel barrier, with $`\epsilon `$ a small coupling constant, in the limit $`\epsilon 0`$ the quasibound states develop into true bound states, Andreev levels, responsible for the current transport in the junction. In this limit we obtain the sum current
$$i^+=\frac{2e}{\mathrm{}}\frac{dE}{d\varphi }\delta (EE_n^\pm ),$$
(12)
and the difference current
$$i^{}=\frac{e}{2\mathrm{}}\frac{\text{Im}(rd^{})\mathrm{sin}\varphi }{\sqrt{D}|\mathrm{cos}(\varphi /2)|\sqrt{1D\mathrm{sin}^2(\varphi /2)}}\left|\frac{dE}{d\varphi }\right|\delta (EE_n^\pm ).$$
(13)
We emphasize that the difference current arises due to essential modification of the interference effects building up the Andreev states.
In the weak coupling limit $`\epsilon 0`$, the Andreev resonances become very narrow, and we may use the bound state expression $`dE/d\varphi `$ to calculate all currents. An important observation is that the anomalous current is proportional to $`|dE/d\varphi |`$, i.e the modulus of the current carried by each level, as shown numerically in Fig. 2a. This means that the absolute values of all currents add up, preventing the cancellation between consecutive Andreev levels responsible for the exponentially small equilibrium current in long junctions.
The anomalous current is proportional to Im$`(rd^{})`$, i.e. is sensitive to the phase of the midpoint scatterer. The anomalous current therefore has interference origin and can be attributed to modification of the bound Andreev states themselves (i.e., not only to nonequilibrium population, responsible for the regular current) . The effect arises because the midpoint scatterer acts as a beam splitter at the injection point, creating left and right moving beams of inected electrons and holes. The current through the SXS junction then becomes a coherent superposition of split beams of electron and holes having scattered at the two SN interfaces, picking up superconducting phases $`\pm \varphi /2`$ at the SN interfaces and transmission and reflection phase shifts at the midpoint scatterer (and at the SN interfaces if the normal interface transmission is not perfect). As a result, the difference current keeps the same direction for each Andreev level (for fixed phase of the scatterer).
Figure 2b shows the IVCs of the regular, anomalous and total nonequilibrium currents as as functions of bias voltage of the injection lead. Of particular interest is that the anomalous current in long junctions at finite temperatures scales linearly with voltage and saturates at $`eV/\mathrm{\Delta }`$ at a current of the same magnitude as the short junction current. The equilibrium current of a short junction $`L<<\xi _0`$ can be supressed by the nonequilibrium regular current for $`eV>\mathrm{\Delta }`$ (Fig. 2b). For a long junction $`L>>\xi _0`$ at zero temperature, the equilibrium current oscillates for $`eV<\mathrm{\Delta }`$, and is suppressed for $`eV>\mathrm{\Delta }`$. The nonequilibrium current has a staircase-like voltage dependence and saturates at $`eV>\mathrm{\Delta }`$ at a level typical for the current of a short junction. At finite temperature the equilibrium current is exponentially small, and the nonequilibrium current scales linearly with voltage for $`eV<\mathrm{\Delta }`$.
Note that the injection current is proportional to the anomalous current, and scales with the coupling constant: $`I_1\epsilon I_a`$. The anomalous current can therefore in principle be controlled with a very small injection current. The injection current is only needed for establishing the local equilibrium in the normal region of the junction. If the coupling is very weak, this will influence the transient time $`t=d/(\epsilon v_F)`$ to reestablish equilibrium if the bias voltage is changed, i.e. influence the switching speed. The current gain in the JOINT can be defined as $`G=I_r/I_1D/\epsilon `$ for the regular current (see also , and $`G=I_a/I_1\sqrt{DR}/\epsilon `$ for the anomalous nonequilibrium current . For $`\epsilon 1`$ this current gain can be become very large (recently demonstrated experimentally for the regular current ).
To build a transistor utilizing the anomalous current, a Josephson interference transistor (JOINT) , is obviously going to be very demanding, and will require delicate control of the junction parameters. At least in principle, by properly gating the junction one might be able to tune the effective length of the junction, the scattering phase shifts and the coupling to the injection lead in order to optimize the junction to reveal the anomalous Josephson effect.
## 5 Non-equilibrium population of Andreev levels via EM radiation.
In optical physics it is common practice to control level populations by pumping with resonant electromagnetic (EM) radiation.
Similarly, the connection between the Andreev level energy $`E(\varphi ,D,\beta )`$ and the Josephson current $`I(E)=(2e/\mathrm{})(dE/d\varphi )`$ should make it possible to control the Josephson current by using resonant EM radiation to pump between Andreev levels (Fig. 3), modulating the energy $`E(\varphi ,D,\beta )`$ and controlling the level populations $`n(E)`$. In principle one only has to modulate the phase difference $`\varphi `$ or the normal transparency $`D`$ of the junction.
There are different mechanisms of coupling Andreev levels to external EM fields. In general, a longitudinal electric field component oriented along the junction will create a time-dependent potential difference $`V(t)`$ across the junction and induce a time-dependent phase difference $`\varphi (t)=\varphi +2eV(t)𝑑t`$ by virtue of the Josephson relation. Another coupling mechanism, using e.g. a transverse electric field component, involves manipulation of the normal electron scattering properties of the junction, i.e. junction transparency $`D`$, electron density, etc. Such possibilites should exist in gate-controlled S-2DEG-S devices and mechanically controllable break junctions.
The resulting level vibration will induce interlevel transitions, and the redistribution of excitations among the bound Andreev levels will lead to a change of the Josephson current. The matrix element for the interlevel transitions depends on the mechanism of the coupling and the junction geometry. In the simplest case $``$ voltage driven interlevel transitions in a short one-mode constriction with a single pair of Andreev levels (Fig. 3a) $``$ the matrix element has the form $`\lambda =(\mathrm{\Delta }/\omega )^2\mathrm{sin}^2(\phi /2)D\sqrt{R}`$. The matrix element for junctions with more complex geometry is given in . It is important to mention that normal electron back scattering by the junction is essential for transitions between the Andreev levels: in purely ballistic constrictions ($`R=0`$) the interlevel transitions are forbidden due to conservation of the momentum of electrons at the Fermi level.
The current response strongly depends on the way of switching on the radiation. Adiabatic turning on a flat-top EM pulse will suppress the Josephson current $`I=I(E)n(E)`$ in a short constriction due to equalization of the level populations $`n(E)`$,
$$I/I_0=1\left[(\lambda eV)/\mathrm{}(\delta /2+\mathrm{\Omega })\right]^2$$
(14)
where $`\delta `$ is the deviation from resonance (detuning), and $`\mathrm{\Omega }=\sqrt{(\lambda eV)^2+\delta ^2/4}`$ is the Rabi frequency. Exactly at resonance, $`\delta =0`$, the current will be entirely suppressed.
Since EM radiation induces a highly coherent mixture of Andreev states, the off-diagonal part of the density matrix may give rise to oscillating contributions to the current (Rabi oscillations). Therefore, if the EM power is turned on suddenly, the level populations $`n(E)`$ will oscillate between the lower level (the initial state) and upper level, causing Rabi oscillations in the Josephson current,
$$I(t)=I_0\left[12\left(\lambda eV/\mathrm{}\mathrm{\Omega }\right)^2\mathrm{sin}^2\mathrm{\Omega }t\right]$$
(15)
The regime of current oscillation is sensitive to dephasing, which makes it possible to measure dephasing times.
Due to the selective character of the control, the lower bound for the power is only limited by the level width, originating e.g. from inelastic interaction with phonons. Since harmonic EM radiation only affects the resonant pair of Andreev states, the current control is most efficient if the number of Andreev states involved in the Josephson transport is small, i.e. short junction length and small number of normal transport electron modes.
In order to suppress the Josephson current, the members of the resonant pair of Andreev levels must carry current in opposite directions (Fig. 3a); otherwise nothing will happen. One may however ask whether it is possible to enhance the Josephson current by pumping with EM radiation. The answer is in principle yes: by using a suitably tuned SN-DBS-NS junction to establish an Andreev split-level scheme with $`E\sqrt{D}`$ dispersion, as shown in Fig. 3b, one can kill the current from one of the levels by optical pumping, leaving an uncompensated ”giant” Josephson current from the unpumped level.
Morever, in a long junction filled with Andreev levels, it should be possible by resonant optical pumping to create a Josehpson current. In junctions where $`l_\varphi >>\xi _T`$, the Josephson current dies on a lengthscale much shorter than the phase breaking length of the electrons, due to cancellation among bound states and between the net bound state current and the continuum current. In such a long junction with vanishing equilibrium Josephson current, by pumping between two Andreev levels, and therefore effectively killing the current from one bound state, it should be possible to unbalance the cancellation and create a mesurable Josephson current of one bound state arising from uncompensated continuum current (see also Ref. ).
## 6 Concluding remarks
The recent progress in the field of superconducting mesoscopic junctions, with theory and experiment going hand in hand, has lead to the realization of elementary Josephson transistors which demonstrate the feasibility of controlling Josephson current. So far, only a few possibilities have been accomplished, mainly involving quenching of equilibrium Josephson current using electrostatric gates or current injection. Some interesting new possibilities involve switching on a Josephson current, using injection of normal current or electromagnetic radiation to create nonequlibrium conditions. These possibilities will require experimental control over multiterminal junctions with very few transport modes, and will present a major challenge for experimental investigations.
Acknowledgement$``$This work has been supported by the Swedish grant agencies NFR, TFR, NUTEK, and the Japanese grant agency NEDO.
|
no-problem/9812/cond-mat9812086.html
|
ar5iv
|
text
|
# Hidden Integrability of a Kondo Impurity in an Unconventional Host
## Abstract
We study a spin-$`\frac{1}{2}`$ Kondo impurity coupled to an unconventional host in which the density of band states vanishes either precisely at (“gapless” systems) or on some interval around the Fermi level (“gapped”systems). Despite an essentially nonlinear band dispersion, the system is proven to exhibit hidden integrability and is diagonalized exactly by the Bethe ansatz.
In the Bethe ansatz (BA) approach to the theory of dilute magnetic alloys , initiated by Wiegmann and Andrei , the conditions of (i) a linear dispersion of band electrons near the Fermi level and (ii) an energy independent electron-impurity coupling play such a crucial role that up to recent time they have been considered as necessary mathematical conditions for integrability of impurity models. However, it has recently been found that integrability of the degenerate and $`U\mathrm{}`$ nondegenerate Anderson models is not destroyed by a nonlinear dispersion of particles and an energy dependent hybridization, but it becomes only hidden . The approach developed has allowed one to study the ground state properties of a $`U\mathrm{}`$ Anderson impurity embedded in unconventional Fermi systems, such as a BCS superconductor and a “gapless” host . In the latter, a density of band states is assumed to vanish precisely at the Fermi level .
In the BA approach, the spectrum of a metal host is alternatively described in terms of interacting charge and spin excitations rather than in terms of free particles with spin “up” and “down”. A choice of an appropriate Bethe basis of a host is dictated by physical properties of an impurity. In the Anderson model, spin and charge excitations strongly interact, while in exchange models, they are decoupled from each other. The latter stems from a momentum independence of electron-impurity scattering in exchange models. The Kondo model is thus “simpler” than the Anderson one, and its solution is derived from that of the Anderson model in the limit where an impurity energy lies much below the Fermi level of a host.
However, in an unconventional host electron-impurity scattering amplitudes even in exchange models are clear to depend essentially on an electron momentum due to an energy dependent density of band states. Therefore, charge and spin degrees of freedom in exchange models should be coupled to each other, that could lead to novel features of the Kondo physics of unconventional Fermi systems.
In this Letter, we report first results for the Kondo model in an unconventional host. Despite an essentially nonlinear dispersion of band electrons near the Fermi level, the model is proven to exhibit hidden integrability. As in the unconventional Anderson model , an auxiliary $`\tau `$-space related to the particle energy is introduced to show that the multiparticle scattering process is factorized into two-particle ones. Multiparticle functions in an auxiliary $`x`$-space related to the particle momentum are then found as a result of an integral “dressing” procedure of the BA functions in the $`\tau `$-space. Thus, factorizability of scattering in the $`x`$-space is hidden and becomes visible only in the limit of large interparticle separations.
Integrability of the unconventional Kondo model is clear to provide us a powerful theoretical tool in studies of the Kondo physics in such systems of fundamental interest as superconductors and unconventional Fermi systems, where a band particle dispersion cannot be linearized near the Fermi level.
We start with the standard s-d exchange (Kondo) model with an arbitrary band electron dispersion $`ϵ(k)`$. An effective 1D Hamiltonian of the model is written in terms of the Fermi operators $`c_\sigma ^{}(ϵ)`$ \[$`c_\sigma (ϵ)`$\] which create (annihilate) an electron with a spin $`\sigma =,`$ in an $`s`$-wave state of energy $`ϵ`$,
$$H=\underset{\sigma }{}_C\frac{dϵ}{2\pi }ϵc_\sigma ^{}(ϵ)c_\sigma (ϵ)+\underset{\sigma ,\sigma ^{}}{}_C\frac{dϵ}{2\pi }\frac{dϵ^{}}{2\pi }I(ϵ,ϵ^{})c_\sigma ^{}(ϵ)\left(\stackrel{}{\sigma }_{\sigma \sigma ^{}}\stackrel{}{S}\right)c_\sigma ^{}(ϵ^{}).$$
(1)
Here, $`\stackrel{}{\sigma }`$ are the Pauli matrices, $`\stackrel{}{S}`$ is the impurity spin operator. An effective electron-impurity coupling $`I(ϵ,ϵ^{})=\frac{1}{2}I\sqrt{\rho (ϵ)\rho (ϵ^{})}`$ combines the exchange coupling constant $`I`$ and the density of band states $`\rho (ϵ)=dk/dϵ`$. The integration contour $`C`$ contains two intervals, $`C=(D,\mathrm{\Delta })(\mathrm{\Delta },D)`$, where $`D`$ is the half band width, and $`2\mathrm{\Delta }`$ is the energy gap in a gapped host, while in a gapless host $`\mathrm{\Delta }=0`$. In what follows, we restrict our consideration to the case of $`S=\frac{1}{2}`$. The electron energies and momenta in Eq. (1) and hereafter are taken relative to the Fermi values, which are set to be equal to zero.
We look for one-particle eigenstates of the system in the form
$$|\mathrm{\Psi }_1>=\underset{\sigma }{}\underset{s=0,1}{}_C\frac{dϵ}{2\pi }\psi _{\sigma ;s}(ϵ)c_\sigma ^{}(ϵ)\left(S^+\right)^s|0>,$$
(2)
where the vacuum state $`|0>`$ contains no electrons and the impurity spin is “down”. The Schrödinger equation for the auxiliary wave function $`\varphi _{\sigma ;s}(ϵ)=\psi _{\sigma ;s}(ϵ)/\sqrt{\rho (ϵ)}`$ is easily found to be
$`(ϵ\omega )\varphi _{\sigma ;s}(ϵ|\omega )+{\displaystyle \frac{1}{2}}I\left(\stackrel{}{\sigma }_{\sigma \sigma ^{}}\stackrel{}{S}_{ss^{}}\right)A_{\sigma ^{};s^{}}(\omega )=0`$ (4)
$`A_{\sigma ;s}(\omega )={\displaystyle _C}{\displaystyle \frac{dϵ}{2\pi }}\rho (ϵ)\varphi _{\sigma ;s}(ϵ|\omega ),`$ (5)
where $`\omega `$ is the eigenenergy.
To simplify notation we omit hereafter the spin indexes. Inserting the general solution of Eq. (4a),
$$\varphi (ϵ|\omega )=2\pi \delta (ϵ\omega )\chi \frac{\frac{1}{2}I}{ϵ\omega i0}\left(\stackrel{}{\sigma }\stackrel{}{S}\right)A(\omega ),$$
(7)
with an arbitrary spinor $`\chi `$, into Eq. (4b), one obtains
$$\left[1+\frac{1}{2}I\mathrm{\Sigma }(\omega )\left(\stackrel{}{\sigma }\stackrel{}{S}\right)\right]A(\omega )=\rho (\omega )\chi .$$
(8)
Here, the self-energy $`\mathrm{\Sigma }(\omega )`$ is found to be
$$\mathrm{\Sigma }(\omega )=_C\frac{dϵ}{2\pi }\frac{\rho (ϵ)}{ϵ\omega i0}=P_C\frac{dϵ}{2\pi }\frac{\rho (ϵ)}{ϵ\omega }+\frac{i}{2}\rho (\omega )=\mathrm{\Sigma }^{}(\omega )+i\mathrm{\Sigma }^{\prime \prime }(\omega ),$$
(9)
where $`P`$ stands for the principal part.
It is convenient to rewrite Eq. (3a) for the Fourier image of the function $`\varphi (ϵ|\omega )`$,
$$\varphi (\tau |\omega )=_{\mathrm{}}^{\mathrm{}}\frac{dϵ}{2\pi }\varphi (ϵ|\omega )\mathrm{exp}(iϵ\tau ).$$
(10)
Then, the equation (4a) takes the form
$$\left(i\frac{d}{d\tau }\omega \right)\varphi (\tau |\omega )+\frac{1}{2}I\delta (\tau )\left(\stackrel{}{\sigma }\stackrel{}{S}\right)A=0.$$
(11)
Its solution is easily found to be
$$\varphi (\tau |\omega )=e^{i\omega \tau }\{\begin{array}{cc}\hfill \chi ,& \tau <0\hfill \\ \hfill 𝐑(\omega )\chi ,& \tau >0\hfill \end{array}$$
(12)
with the electron-impurity scattering matrix
$$𝐑(\omega )=u(\omega )+2v(\omega )\left(\stackrel{}{\sigma }\stackrel{}{S}\right),$$
(14)
where the parameters $`u(\omega )`$ and $`v(\omega )`$ are determined by the equations
$`u+v`$ $`=`$ $`1i{\displaystyle \frac{\frac{1}{4}I\rho (\omega )}{1+\frac{1}{4}I\mathrm{\Sigma }(\omega )}}`$ (15)
$`u3v`$ $`=`$ $`1+i{\displaystyle \frac{\frac{3}{4}I\rho (\omega )}{1\frac{3}{4}I\mathrm{\Sigma }(\omega )}}.`$ (16)
In a metal, where an electron-impurity scattering is energy independent, one may introduce the permutation operator of electron spins, $`𝐏=\delta _{\sigma _1,\sigma _2^{}}\delta _{\sigma _2,\sigma _1^{}}`$ as an electron-electron scattering matrix to factorize multiparticle scattering and to construct thus Bethe ansatz eigenfunctions of the system. In an unconventional host, multiparticle scattering is not factorized by the permutation operator because of an energy dependence of electron-impurity scattering amplitudes.
Let us introduce a two-particle scattering matrix of electrons with the energies $`\omega _1`$ and $`\omega _2`$,
$$\varphi (\tau _1>\tau _2|\omega _1,\omega _2)=𝐫_{12}(\omega _1,\omega _2)\varphi (\tau _1<\tau _2|\omega _1,\omega _2),$$
(18)
by the most general $`SU(2)`$-symmetric expression
$$𝐫(\omega _1,\omega _2)=\frac{h(\omega _1)h(\omega _2)i𝐏}{h(\omega _1)h(\omega _2)i}.$$
(19)
Then multiparticle scattering in the system of host electrons is well known to be factorized at an arbitrary function $`h(\omega )`$, since the matrix $`𝐫_{ij}`$, where the particle index $`j=1,2,3`$ is assumed to incorporate also its energy, obeys the Yang-Baxter factorization conditions
$$𝐫_{12}𝐫_{13}𝐫_{23}=𝐫_{23}𝐫_{13}𝐫_{12}.$$
(21)
Introducing an impurity fixes an expression for the function $`h(\omega )`$. To factorize multiparticle scattering in the presence of the impurity with $`𝐑`$ matrix derived in Eq. (10), one needs to solve the Yang-Baxter equations involving two electrons and the impurity,
$$𝐫_{12}𝐑_{10}𝐑_{20}=𝐑_{20}𝐑_{10}𝐫_{12},$$
(22)
that gives
$$h(\omega )=\frac{u(\omega )}{v(\omega )}.$$
(23)
Thus, the multiparticle scattering in the unconventional Kondo model is proven to be factorized in the auxiliary $`\tau `$-space related to the particle energy, and the $`N`$-particle eigenfunctions of the model, $`\mathrm{\Phi }(\tau _1,\mathrm{},\tau _N)`$ are written in the standard BA form. The $`N`$-particle eigenfunctions in the auxiliary $`x`$-space related to the particle momentum, $`\mathrm{\Psi }(x_1,\mathrm{},x_N)`$, are found as a result of a “dressing” procedure
$$\mathrm{\Psi }(x_1,\mathrm{},x_N)=_{\mathrm{}}^{\mathrm{}}\mathrm{\Phi }(\tau _1,\mathrm{},\tau _N)\underset{j=1}{\overset{N}{}}u(x_j|\tau _j)d\tau _j,$$
(25)
with the dressing function
$$u(x|\tau )=_C\frac{dϵ}{2\pi }\rho (ϵ)e^{i[k(ϵ)xϵ\tau ]},$$
(26)
where $`k(ϵ)`$ is the inverse dispersion. In a metal host, where $`ϵ(k)=k`$, the dressing function is nothing but the Dirac delta function, $`u(x|\tau )=\delta (x\tau )`$, and hence the $`x`$ and $`\tau `$ representations coincide.
Imposing periodic boundary conditions on the wave function $`\mathrm{\Psi }(x_1,\mathrm{},x_N)`$ on the interval of size $`L`$ leads in the standard manner to the BA equations
$`\mathrm{exp}(ik_jL)\phi _c(\omega _j)`$ $`=`$ $`{\displaystyle \underset{\alpha =1}{\overset{M}{}}}{\displaystyle \frac{h_j\lambda _\alpha \frac{i}{2}}{h_j\lambda _\alpha +\frac{i}{2}}}`$ (28)
$`\phi _s(\lambda _\alpha ){\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{\lambda _\alpha h_j\frac{i}{2}}{\lambda _\alpha h_j+\frac{i}{2}}}`$ $`=`$ $`{\displaystyle \underset{\beta =1}{\overset{M}{}}}{\displaystyle \frac{\lambda _\alpha \lambda _\beta i}{\lambda _\alpha \lambda _\beta +i}}`$ (29)
where $`M`$ is the number of particles with spin “down”, $`k_jk(\omega _j)`$, and $`h_jh(\omega _j)`$. The eigenenergy $`E`$ and the $`z`$ component of the total spin of the system $`S^z`$ are found to be
$$E=\underset{j=1}{\overset{N}{}}\omega _j,S^z=\frac{1}{2}+\frac{N}{2}M.$$
(30)
In Eqs. (13) the phase factors
$`\phi _c(\omega )`$ $`=`$ $`{\displaystyle \frac{1+\frac{1}{4}I\mathrm{\Sigma }^{}(\omega )\frac{i}{2}\frac{1}{4}I\rho (\omega )}{1+\frac{1}{4}I\mathrm{\Sigma }^{}(\omega )+\frac{i}{2}\frac{1}{4}I\rho (\omega )}}`$ (32)
$`\phi _s(\lambda )`$ $`=`$ $`{\displaystyle \frac{\lambda \frac{i}{2}}{\lambda +\frac{i}{2}}}`$ (33)
describe the scattering of charge and spin excitation of the host on the impurity. The equations (13) solve exactly the problem of diagonalization of the s-d exchange (Kondo) model in an unconventional host.
It should be emphasized that these equations are incomplete yet in the case of a gapped host, since they do not account for electron-impurity bound states with eigenenergies lying inside the gap. However, for a further analysis of the problem one needs to specify physical characteristic of a host, therefore we are going to address the thermodynamic properties of a Kondo impurity in unconventional Fermi systems in separated publications.
|
no-problem/9812/cond-mat9812033.html
|
ar5iv
|
text
|
# Effects of Charge Density Modulation on Incommensurate Antiferromagnetism: Ginzburg-Landau Study
## Acknowledgments
We thank Prof. Y. Endoh for providing us his experimental results prior to publication. H. K. and H. Y. thank A. Ino for instructive discussion. H. Y. also thanks T. Adachi, S. Fujiyama, K. Yokoyama and M. Yumoto for stimulating discussions. This work is supported by a Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Sports and Culture of Japan.
|
no-problem/9812/cond-mat9812204.html
|
ar5iv
|
text
|
# On Phase Transition and Self-Organized Critical State in Granular Packings
## Abstract
We model two-dimensional systems of granular aggregates confined between two planes and demonstrate that at a critical grain volume fraction an abrupt rigidity transition occurs. This transition is observed both in static and shear tests. The grain volume fraction at which the transition occurs, $`\nu _c`$, decreases with increasing friction between the grains. Densely packed grains, with a volume fraction $`\nu >\nu _c`$, display an elastic-plastic rheology. Dilute packings, with $`\nu <\nu _c`$, display gas-like characteristics. It is shown that when volume fraction is allowed to change freely (using constant normal stress boundary condition), it evolves spontaneously to $`\nu _c`$ under a wide range of boundary conditions, exhibiting ’self-organized criticality’.
Granular media is a fundamental, yet not well understood, complex system with wide ranging applications to technological and natural systems. In recent years there has been much work on granular dynamics, with emphasis on how behavior of grain aggregates may resemble solids, liquids or gases . Already Reynolds in 1885 noted that loosely packed sands deform easily as fluids while dense packings resist shear as solids. The two phases are traditionally treated separately by ’kinetic-gas’ approaches for loosely packed grains and elasto-plastic (often using associated plasticity ) theories for dense soils. In this Letter we numerically investigate the transition between solid and gas behaviors in granular aggregates, and show that the phase boundary between gas and solid is also an attracting point, to which systems naturally evolve. We numerically model grain aggregates using a version of the popular ’discrete element method’ which treats grains as inelastic disks with rotational and translational degrees of freedom. Two grains of radius $`R_i`$ and $`R_j`$ undergo an inelastic interaction when the distance separating them $`r_{ij}`$ is less than the sum of their radii. During the interaction the $`i`$th grain experiences a contact force that has both shear and normal components:
$$F_{ij}(t)=[k_n(R_i+R_jr_{ij})\gamma m(\dot{𝐫}_{\mathrm{𝐢𝐣}}\widehat{𝐧})]\widehat{𝐧}+[min(k_s\mathrm{\Delta }s,\mu (𝐅\widehat{𝐧}))]\widehat{𝐬}$$
(1)
where $`\widehat{𝐧}=(𝐫_{ij}\widehat{𝐱},𝐫_{ij}\widehat{𝐲})/r_{ij}`$, $`\widehat{𝐬}=(𝐫_{ij}\widehat{𝐲},𝐫_{ij}\widehat{𝐱})/r_{ij}`$, are the unit vectors in the normal and tangential directions respectively. $`k_n,k_s`$ are the normal and shear elastic constants, $`m`$ is the grain mass, $`\gamma `$ is a damping coefficient ensuring inelasticity of the interaction, $`\mu `$ is the surface friction coefficient, and $`\mathrm{\Delta }s`$ is the shear displacement since the initial contact of two particles. Force is integrated through time to calculate grain position and velocity. For collisions governed by equation (1) energy loss is governed by a normal restitution coefficient $`e_n=\mathrm{exp}(\gamma t_{col}/2)`$, where $`t_{col}=\pi (2k_n/m\gamma ^2/4)^{1/2}`$ is the collision time, and by frictional work which depends on the amount of real slip (after the elastic limit is reached) and the frictional shear force, $`\mu F_n`$. Time is measured in units of undissipated elastic wave travel time $`t_0=\sqrt{m/k_n}`$, and distance in units of average disk diameter $`x_0=2\overline{R}`$. In simulations presented here $`k_n=1,\gamma =1,k_s=0.5`$ and $`m=1`$.
In this Letter we investigate three related problems. The first problem is the general characteristics of static granular aggregates after compaction between two parallel plates. The second and third problems are the behavior of the same confined aggregates during shear using two different boundary conditions: constant distance between shearing walls, termed constant volume boundary conditions (CVBC), and constant normal stress applied to the wall, termed constant force boundary conditions (CFBC). Simulations are performed on square systems with $`n`$ disks. The top and bottom edges of the box are composed of grains glued together to form rigid rough walls of length $`l`$ (Figure 1). The box is periodic in the horizontal direction. Grain radii are randomly drawn from a Gaussian distribution that peaks at $`\overline{R}`$, with a standard deviation of $`0.5\overline{R}`$. Polydispersivity is introduced to discourage ordering effects. The system is initiated as tall loosely packed box, which is compacted vertically by normal stresses to a height $`l`$. After compaction the horizontal walls are allowed to move freely in the horizontal direction to relax forces. However, unrelaxed normal forces can be maintained on the walls in this case, since we do not allow the walls to move vertically, using CVBC. Global rearrangements during compaction and relaxation ensure (local) minimal energy configuration, as would occur during natural compaction.
After compaction and relaxation we measure properties of static configurations at different solid volume fractions, $`\nu =_{i=1}^n\pi R_i^2/l^2`$, ranging between 0.75 to 0.96. Measurements of three parameters, (a) the average number of grains touching (i.e. exerting a force on) a grain in the interior of the box, termed the coordination number $`Z`$, (b) The normal stress (normal force per unit length) $`N`$ operating on the upper and lower confining walls, and (c) the systems shear modulus, G, are presented in Figure 2. All measures show an abrupt change in behavior at a critical volume fraction $`\nu _c`$, which depends on the value of friction prescribed between the grains, but not on the system size. The coordination number is approximately zero for $`\nu <\nu _c`$. At $`\nu _c`$ the coordination number abruptly jumps and then increases as an empirical power-law, $`Z(\nu \nu _{c}^{}{}_{\mu }{}^{})^{\alpha _\mu }`$, where the subscript $`\mu `$ denotes different friction values used in simulations. Power-law fits (in solid curves) yields $`\nu _{c}^{}{}_{0}{}^{}=0.83\pm 0.01,\alpha _0=0.5\pm 0.1`$ and $`\nu _{c}^{}{}_{0.5}{}^{}=0.805\pm 0.01,\alpha _{0.5}=0.3\pm 0.05`$. Critical behavior with values of $`\nu _{c}^{}{}_{0}{}^{}=0.82\pm 0.02`$ were obtained for hard smooth disks , and visco-elastic 2D bubbles for mono-dispersed and polydispersed systems, demonstrating that $`\nu _c`$ is fairly independent of the disk size distribution, and the interaction law in absence of friction. The difference in $`\nu _c`$ and $`\alpha `$ between frictional and smooth grains occurs because frictional grains tend to ’stick’, and thus cannot achieve the lower energy configuration of smooth disks.
The fact that $`N`$ and $`Z`$ (Figure 2a,b) are zero for $`\nu <\nu _c`$, indicates that below $`\nu _c`$ grains do not touch. At $`\nu _c`$ grains first touch and elastically repel each other, exerting normal stresses on the walls. For $`\nu >\nu _c`$, normal stress follows $`NZ(\nu \nu _c)`$ (shown in solid lines), a slightly modified form (accounting for the phase transition) of predictions from standard models of densely packed elastic disks . The observed transition is identified as a macroscopic rigidity transition in Figure 2c. The system’s elastic shear modulus $`G`$ is obtained by imposing a small homogeneous shear step strain of a magnitude $`ϵ`$ and measuring the resulting shear stress on the walls $`\sigma `$ ($`G=\sigma /ϵ`$ is independent of $`ϵ`$ for $`ϵ<10^4`$ of the grain radius, here we use $`ϵ=10^5`$). The procedure follows that outlined in . Figure 2c shows $`G`$ increasing from zero to a finite value, as the system passes through the phase transition at $`\nu =\nu _c`$. Physically, the rigidity should depend on average number of springs per disk, yet not on their compression, i.e. $`GZ`$ (solid curves). The rigidity-transition thus identified is a result of global geometrical constraints of grain packings, with the coordination number identified as the order parameter for the transition. It was previously shown that $`\nu =0.83\pm 0.02`$ marks the upper limit of compacity of disordered packings of smooth hard mono-sized and poly-sized disks , where long-range order in disk positions appearing for $`\nu >\nu _c`$.
To investigate the transition in dynamic behavior of grains we conducted a set of simulation shearing the static 24x24 configurations, in couette flow. Here we show results of simulations with $`\mu =0.5`$. (Non-frictional grains and other values of friction show similar behavior with the transition occuring at $`\nu _{c}^{}{}_{\mu }{}^{}`$, and are thus not presented). The upper wall was moved at a constant velocity ($`v=10^3x_0/t_0`$) while $`l`$ was kept constant, maintaining CVBC. We observe different behavior as a function of the solid fraction: In configurations with $`\nu <\nu _c`$, momentum is transferred from the wall to interior grains mostly via short-lived collisions, observed in spiky fluctuations in stress measured on the wall and in coordination number, Fig. 3. Stresses are transmitted only within local clusters (as in Figure 1(t2)). The power spectra of the stress fluctuations time series approaches white noise demonstrating the uncorrelated nature of stress-transfer. For dense packings, with $`\nu >\nu _c`$, grains interact via long-lasting contacts. Global motion is characterized by elastic-plastic cycles: clusters of grains in contact accumulate recoverable elastic strain, but when stresses become too great, grains suddenly rearrange to relieve the stress. Continued shearing then begins accumulation of elastic strain on the new particle arrangement. This behavior leads to a stick-slip stress time series (Figure 3). Long clusters of these grains in contact form system-spanning ’stress chains’ that transmit forces from the boundaries into the interior (as in Figure 1(t1)). The stress fluctuation power spectra for $`\nu >\nu _c`$ follows $`f^\eta `$, where $`\eta 2`$, indicating long-time correlations and in agreement with experimental results conducted in densely packed systems. Most interesting is the behavior at the transition point: When $`\nu =\nu _c`$ the system oscillates between a solid-like ’jammed’ state (Figure 1(t1)) and a gas-like behavior (Figure 1(t2)). At $`\nu =\nu _c`$, values of $`Z,\sigma ,N`$, fluctuate between values characteristic of the ’solid-phase’ and those characteristic of ’gas-phase’ (Figure 3). The transition density thus bridges the gap between liquid and solid by coexistence of the two phases. We note the relation to Kirkwood-Alder transition (KAT), a disorder-order transition of repulsive hard-spheres studied extensively in the context of collidol suspensions . The relation to KAT, as well as the transition characteristics, suggest that the granular rigidity transition is a first-order transition.
The relation between stress and strain rate (to be published elsewhere) is another property that changes across the phase boundary. When $`\nu <\nu _c`$ we measure $`(\sigma ,N)v^2`$, as expected from theory and experiments . When $`\nu >\nu _c`$ measured average stresses are nearly independent of strain rate, as expected for elastic-plastic materials and as seen in experiments in densely packed granular systems . For $`\nu =\nu _c`$ stress-strain-rate curves resemble those of plastic materials, since high stresses in jammed states dominate time-averaged behavior.
We finally demonstrate the role of the critical solid fraction when the system is allowed to evolve to it’s own prefered solid fraction. Soil mechanists have long known of a ’critical density’. If a granular aggregate is over-compacted and sheared under CFBC, it will deform while shearing and expand to this ’critical density’. If it is initially under-consolidated it will compact while shearing until it reaches the same ’critical density’. We perform simulations, similar to these experimental conditions, by applying a constant normal stresses to the rigid boundaries (CFBC), while shearing the upper wall at velocity $`v`$. The systems expand or contract, depending on the initial porosity, and finally reach a steady state where variables ($`\nu ,\sigma ,Z`$) fluctuate around a constant value. Steady state solid fractions, $`<\nu >`$, are shown in Figure 4(a) as function of $`N`$ and $`v`$, for simulations with $`\mu =0.5`$. For a wide range of applied normal stresses (range widens with decreasing velocity) systems attain the critical volume fraction for frictional grains, i.e. $`<\nu >\nu _{c}^{}{}_{0.5}{}^{}=0.805`$. For smooth disks, simulations converge on $`<\nu >\nu _{c}^{}{}_{0}{}^{}=0.835`$, in a similar manner. Though $`<\nu >`$ is fairly independent of $`N`$ in the ’critical regime’, the average coordination number $`<Z>`$ increases (Figure 4(b1)(b2)). In those systems having $`<\nu >\nu _c`$ system-spanning stress-chains coexist with unstressed ’gas islands’. With increasing $`N`$, the number and size of gas islands decreases, and the connectivity of chains increases, and thus $`<Z>`$ increases. Deviations from the critical state occur in two cases: 1) At high enough normal stresses, ’solidification’ occurs: $`<\nu >`$ increases above $`\nu _c`$, islands virtually vanish and chains become fully connected ($`N=10^2`$ in Figure 4a). 2) When the ratio of inertial to normal forces is high inertial forces cause decompaction, $`<\nu >`$ becomes smaller than $`\nu _c`$ (simulations performed at higher velocities and lower normal stresses: Figure 4a, $`v=10^3`$ and $`N10^5`$), with islands growing to divide stress chains, resulting in ’liquification’. We also observed the gas to solid transition in the power spectra of stress fluctuations, where power follows $`f^\eta `$ with $`\eta `$ increasing continuously from $`\eta 0`$ for $`N=10^7`$ to $`\eta 2`$ for $`N10^2`$. In these CFBC simulations the system, although at the critical state, does not flip between gas and solid phases as in CVBC, since ’jamming’ episodes may be avoided by slight dilation (producing fluctuations around mean porosity and height). Instead of temporal coexistence seen in CVBC the critical state in CFBC is marked by spatial coexistence of two phases: gas-islands and stress-bearing chains.
Why is the transition between gas and solid also the ’critical-density’ to which the system is attracted? any finite normal stress acting on grains would cause grain compaction and contact, and thus loss of ’gas-like’ properties, approaching $`\nu _c`$ from below. Solid fractions of $`\nu >\nu _c`$ are characterized by finite elastic deformation ($`>1\%`$) of grains, which require extremely high normal stresses for stiff natural granular material as rocks. Thus $`\nu _c`$ is the rigid limit, where stresses are accommodated by efficient load bearing structures. Experimentally, a coexistence of liquid and solid regimes has been observed in shearing granular systems, and oscillations between ’jamming’ and ’flowing’ states occur spontaneously in a variety of systems, from hoppers to natural and experimental land-slides , suggesting proximity of these systems to the phase boundary. An attracting phase boundary may also explain ’fragile materials’, a recent term used to describe ’jammed’ states which may be unjammed by small fluctuations .
To summarize, we identified the rigidity transition for granular media in $`2D`$ at $`\nu _c=0.800.84`$. Two-dimensional results can be mapped to corresponding 3D volume fractions using a relation for inter-particle voids, $`\nu _{3D}=4\nu _{2D}^{3/2}/3\pi ^{1/2}`$ , predicting that in 3D $`\nu _c=0.540.58`$ with the lower number representing frictional grains. Experiments confirm that immense stiffening of rapidly shearing grain aggregates occurs at $`\nu _c0.54`$. The critical solid-fraction is a phase-boundary between gas and solid regimes of behavior, where the two phases were observed to coexist in space and/or time. The critical state is self-organizing, and corresponds to the ’critical density’ known in soil mechanics: Sheared aggregates tend to this critical volume fraction under a wide range of normal loads. We used normal stresses ranging between $`10^210^7`$, which using characteristic young modulus for Earth materials (e.g. rock), translates to $`N=110^5`$ KPa, (corresponding for soils to burial depth of $`110^5m`$). Based on this we suggest that many natural granular deformation processes will occur at a solid fraction which constitutes a phase boundary, and neither gas nor elasto-plastic descriptions will fully capture these systems behavior. However, such criticality in natural systems is easily missed, since spatial and temporal averages of stresses tend to be dominated by stress chains, and thus have the mark of solid deformation. It is clear that much more theoretical and experimental work needs to be done, including investigating much larger systems, and examining the critical state more closely.
|
no-problem/9812/math9812131.html
|
ar5iv
|
text
|
# A Note on the Gauss Map of Complete Nonorientable Minimal Surfaces.
## 1 Introduction and Preliminaries
The study of the Gauss map of complete orientable minimal surfaces in $`\text{}^3`$ has achieved many important advances and also has given rise to many problems in recent decades. The most interesting question is to determine the size of the spherical image of such a surface under its Gauss map.
R. Osserman was the person who started the systematic development of this theory, and so, in 1961 he proved that the set omitted by the image of a complete non flat orientable minimal surface by the Gauss map has logarithmic capacity zero. In 1981 F. Xavier proved that this set covers the sphere except six values at the most, and finally in 1988 H. Fujimoto obtained the best possible theorem, and proved that the number of exceptional values of the Gauss map is four at the most. An interesting extension of Fujimoto’s theorem was proved in 1990 by X. Mo and R. Osserman . They showed that if the Gauss map of a complete orientable minimal surface takes on five distinct values only a finite number of times, then the surface has finite total curvature.
There are many kinds of complete orientable minimal surfaces whose Gauss map omits four points of the sphere. Among these examples we emphasize the classical Sherk’s doubly periodic surface and those described by K. Voss in (see also ). The first author of this paper in constructs orientable examples with non trivial topology.
Under the additional hypothesis of finite total curvature, R. Osserman proved that the number of exceptional values is three at the most.
In the nonorientable case, the Gauss map of the two sheeted orientable covering surface induces, in a natural way, a generalized Gauss map from the nonorientable surface on the projective plane. From Fujimoto’s theorem applied to the two sheeted orientable covering, this generalized Gauss map omits two points of $`\text{}\text{}^2`$ at the most.
It left open the following questions:
1. Are there complete nonorientable minimal surfaces in $`\text{}^3`$ whose generalized Gauss map omits two points of $`\text{}\text{}^2`$?
2. Are there complete non flat orientable minimal surfaces in $`\text{}^3`$ with finite total curvature whose Gauss map omits three points of $`\text{𝕊}^2`$?
3. Are there complete nonorientable minimal surfaces in $`\text{}^3`$ with finite total curvature whose generalized Gauss map omits one point of $`\text{}\text{}^2`$?
Concerning the second problem, A. Weitsman and F. Xavier in and Y. Fang in obtained nonexistence results, provided that the absolute value of the total curvature is less than or equal to $`16\pi `$ and $`20\pi `$, respectively.
In this paper we give an affirmative answer to the first question, and prove:
> Theorem There are complete nonorientable minimal surfaces in $`\text{}^3`$ whose generalized Gauss map omits two points of the projective plane.
Our method of construction is somewhat explicit and very simple, and it is based on a more elaborate use of the Voss technique.
Finally, we briefly summarize some of the basic facts we will need in this paper.
Let $`X:M\text{}^3`$ be a minimal immersion of a surface $`M`$ in three dimensional Euclidean space. Using isothermal parameters, $`M`$ has in a natural way a conformal structure. When $`M`$ is orientable, we label $`(g,\eta )`$ as the Weierstrass data of $`X`$. Remember that the stereographic projection $`g`$ of the Gauss map of $`X`$ is a meromorphic function on $`M`$, and $`\eta `$ is a holomorphic 1-form on $`M`$.
Moreover,
$$X=\text{Real}(\mathrm{\Phi }_1,\mathrm{\Phi }_2,\mathrm{\Phi }_3),$$
where $`\mathrm{\Phi }_1=\frac{1}{2}\eta (1g^2),\mathrm{\Phi }_2=\frac{i}{2}\eta (1+g^2),\mathrm{\Phi }_3=\eta g`$ are holomorphic 1-forms on $`M`$ satisfying:
$$\left(\underset{j=1}{\overset{3}{}}|\mathrm{\Phi }_j|^2\right)(P)0,PM.$$
In particular, $`\mathrm{\Phi }_j`$, $`j=1,2,3,`$ have no real periods on $`M`$. Furthermore, the Riemannian metric $`ds^2`$ induced by $`X`$ on $`M`$ is given by:
$$ds^2=\underset{j=1}{\overset{3}{}}|\mathrm{\Phi }_j|^2.$$
For more details see .
Consider now $`X^{}:M^{}\text{}^3`$ a conformal minimal immersion of a nonorientable surface $`M^{}`$ in $`\text{}^3`$. Let $`\pi _0:MM^{}`$, $`I:MM`$ denote the conformal oriented two sheeted covering of $`M^{}`$ and the antiholomorphic order two deck transformation for this covering, respectively.
If $`(g,\eta )`$ represents the Weierstrass data of $`X=X^{}\pi _0`$, then it is not hard to deduce that :
$$I^{}(\mathrm{\Phi }_j)=\overline{\mathrm{\Phi }_j},j=1,2,3.$$
(1)
In particular, $`gI=I_0g`$, where $`I_0(z)=1/\overline{z}`$, and so there is a unique map
$$G:M^{}\text{}\text{}^2\overline{\text{}}/I_0$$
satisfying
$$G\pi _0=gp_0,$$
where $`p_0:\overline{\text{}}\overline{\text{}}/I_0`$ is the natural projection. We call $`G`$ the generalized Gauss map of $`X^{}`$.
Conversely, given $`(M,g,\eta )`$ the Weierstrass representation of a minimal immersion $`X`$ of an orientable surface $`M`$ in $`\text{}^3`$, and given $`I:MM`$ an antiholomorphic involution without fixed points on $`M`$ satisfying (1), then $`X`$ induces a minimal immersion $`X^{}`$ of $`M^{}=M/I`$ in $`\text{}^3`$ such that $`X=X^{}\pi _0`$. For more details see .
Finally, denote:
* $`\text{𝔻}=\{z\text{}:|z|<1\}`$,
* $`\text{𝔻}^{}=\text{𝔻}\{0\}`$,
* for each $`R>1`$, $`A(R)=\{z\text{}:\mathrm{\hspace{0.33em}1}/R<|z|<R\}`$.
Throughout the proof of Theorem 2, we will use the following result:
###### Theorem 1
Let $`M`$ be a Riemann surface with holomorphic universal covering space 𝔻. Then $`M\text{𝔻}`$, $`\text{𝔻}^{}`$, or $`A(R)`$, provided $`\mathrm{\Pi }_1(M)`$ is commutative.
The proof of this theorem can be found in \[2, Chapter IV\].
## 2 Main Theorem
To obtain the result we have stated in the introduction, we need the following two Lemmas.
###### Lemma 1
There exist $`R>1`$ and holomorphic $`1`$-forms $`\mathrm{\Phi }_j`$, $`j=1,2,3`$, on $`A(R)`$ such that:
1. $`\mathrm{\Phi }_1^2+\mathrm{\Phi }_2^2+\mathrm{\Phi }_3^20.`$
2. $`|\mathrm{\Phi }_1|^2+|\mathrm{\Phi }_2|^2+|\mathrm{\Phi }_3|^20.`$
3. The metric $`ds^2\stackrel{\mathrm{def}}{=}|\mathrm{\Phi }_1|^2+|\mathrm{\Phi }_2|^2+|\mathrm{\Phi }_3|^2`$ is complete.
4. The Gauss map
$$g=\frac{\mathrm{\Phi }_1+i\mathrm{\Phi }_2}{\mathrm{\Phi }_3}$$
omits four points of the Riemann sphere $`\overline{\text{}}`$.
5. $`I^{}(\mathrm{\Phi }_j)=\overline{\mathrm{\Phi }_j},`$ $`j=1,2,3,`$ where $`I:A(R)A(R)`$ is given by $`I(z)=1/\overline{z}`$.
* Let $`\alpha ,\beta \text{}^{}`$, $`\alpha \{\beta ,1/\overline{\beta }\}`$, label
$$M=\overline{\text{}}\{\alpha ,\beta ,\frac{1}{\overline{\alpha }},\frac{1}{\overline{\beta }}\}.$$
and consider the following Weierstrass representation on $`M`$:
$$\widehat{g}=z,\widehat{\eta }=\frac{idz}{(z\alpha )(z\beta )(\overline{\alpha }z+1)(\overline{\beta }z+1)}.$$
(2)
If we define $`\widehat{I}:MM`$, $`\widehat{I}(z)=1/\overline{z}`$, then $`\widehat{I}`$ is an antiholomorphic involution without fixed points, verifying:
$$\widehat{g}\widehat{I}=\frac{1}{\overline{\widehat{g}}},\widehat{I}^{}(\widehat{\eta })=\overline{\widehat{\eta }\widehat{g}^2}.$$
(3)
Thus, if we define:
$`\widehat{\mathrm{\Phi }}_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}(1\widehat{g}^2)\widehat{\eta },`$
$`\widehat{\mathrm{\Phi }}_2`$ $`=`$ $`{\displaystyle \frac{i}{2}}(1+\widehat{g}^2)\widehat{\eta },`$
$`\widehat{\mathrm{\Phi }}_3`$ $`=`$ $`\widehat{g}\widehat{\eta }.`$
then it is obvious, from (3), that $`\widehat{I}^{}(\widehat{\mathrm{\Phi }}_j)=\overline{\widehat{\mathrm{\Phi }}_j}`$. Furthermore, these holomorphic 1-forms satisfy:
+ $`\widehat{\mathrm{\Phi }}_1^2+\widehat{\mathrm{\Phi }}_2^2+\widehat{\mathrm{\Phi }}_3^20`$,
+ $`|\widehat{\mathrm{\Phi }}_1|^2+|\widehat{\mathrm{\Phi }}_2|^2+|\widehat{\mathrm{\Phi }}_3|^20`$,
+ The Riemannian metric $`d\widehat{s}^2=|\widehat{\mathrm{\Phi }}_1|^2+|\widehat{\mathrm{\Phi }}_2|^2+|\widehat{\mathrm{\Phi }}_3|^2`$ is complete in $`M`$.
On the other hand, the Uniformization Theorem says us that the holomorphic universal covering of $`M`$ is either or the unit disc, 𝔻 (see \[2, §IV.4\]). However, is the conformal covering of only two non compact Riemann surfaces: and $`\text{}^{}`$ (see \[2, §IV.6\]). Thus, the holomorphic universal covering of $`M`$ is 𝔻. We label $`\pi :\text{𝔻}M`$ as the conformal covering map.
Let $`\stackrel{~}{I}`$ be a lift of $`\widehat{I}`$ to 𝔻, and denote $`\stackrel{~}{\mathrm{\Phi }}_j=\pi ^{}(\widehat{\mathrm{\Phi }}_j)`$, $`j=1,2,3`$. It is clear that $`\stackrel{~}{I}^{}(\stackrel{~}{\mathrm{\Phi }}_j)=\overline{\stackrel{~}{\mathrm{\Phi }}_j}`$, $`j=1,2,3`$.
Since $`\widehat{I}`$ is an antiholomorphic involution in $`M`$ without fixed points, then $`\stackrel{~}{I}^{2k+1}`$, $`k\text{}`$, is an antiholomorphic transformation in 𝔻 without fixed points too.
Let us see that $`\stackrel{~}{I}^{2k}`$, $`k\text{}^{},`$ has no fixed points in 𝔻. Indeed, note that $`\stackrel{~}{I}^{2k}`$, $`k\text{}^{}`$, is a lift of the identity mapping in $`M`$. Thus, if $`\stackrel{~}{I}^{2k}`$ fixes a point of 𝔻, we infer that $`\stackrel{~}{I}^{2k}`$ is the identity mapping $`\text{1}_\text{𝔻}`$ in 𝔻.
Assume that there is $`k>0`$ such that $`\stackrel{~}{I}^{2k}=\text{1}_\text{𝔻}`$. Let
$$k_0=\text{Minimum}\{k\text{}^{}:\stackrel{~}{I}^{2k}=\text{1}_\text{𝔻}\},$$
and observe that $`k_0`$ is the finite order of $`\stackrel{~}{I}^2`$. It is clear that $`k_0>1`$. Otherwise, $`k_0=1`$ and so there would be antiholomorphic involutions without fixed points in 𝔻, which is absurd. Furthermore, from the definition of $`k_0`$, it is obvious that $`\stackrel{~}{I}^{2k}`$ has no fixed points, $`0<k<k_0.`$
Therefore, the quotient $`\text{𝔻}/\stackrel{~}{I}^2`$ is a Riemann surface with fundamental group isomorphic to $`\text{}_{k_0}`$. No such surface exists (see for instance Theorem 1).
This contradiction implies that $`\stackrel{~}{I}^{2k}`$, $`k\text{}^{},`$ has no fixed points and $`\stackrel{~}{I}^2\text{}`$. In other words, the map
$$\zeta :\text{𝔻}\text{𝔻}/\stackrel{~}{I}^2$$
is a cyclic conformal covering and the fundamental group of $`\text{𝔻}/\stackrel{~}{I}^2`$ is isomorphic to .
Using Theorem 1 we deduce that $`\text{𝔻}/\stackrel{~}{I}^2`$ is conformally equivalent to either $`\text{𝔻}^{}`$ or $`A(R)`$, for a suitable $`R>1`$.
The map $`\stackrel{~}{I}`$ induces on $`\text{𝔻}/\stackrel{~}{I}^2`$ an antiholomorphic involution, $`I`$. Moreover, $`\text{𝔻}/\stackrel{~}{I}^2`$ is in a natural way a covering of $`M`$, and $`I`$ is projected under this covering map on the original involution $`\widehat{I}`$ on $`M`$. Since $`\widehat{I}`$ has no fixed points in $`M`$, the same occurs for $`I`$ in $`\text{𝔻}/\stackrel{~}{I}^2`$.
However, any antiholomorphic involution in $`\text{𝔻}^{}`$ extends to 𝔻, and is the conjugate of a Möbius transformation leaving 𝔻 invariant and fixing $`0`$. In particular, any such map has infinitely many fixed points in 𝔻. Hence, we conclude that $`\text{𝔻}/\stackrel{~}{I}^2`$ can not be conformally equivalent to $`\text{𝔻}^{}`$, i.e., $`\text{𝔻}/\stackrel{~}{I}^2`$ is conformally diffeomorphic to $`A(R)`$, for a suitable $`R>1`$.
If we look at $`I`$ as an antiholomorphic involution in $`A(R)`$, then elementary arguments of complex analysis give that $`I(z)=1/\overline{z}`$, $`zA(R)`$.
On the other hand, as $`(\stackrel{~}{I}^2)^{}(\stackrel{~}{\mathrm{\Phi }}_j)=\stackrel{~}{\mathrm{\Phi }}_j`$, then $`\stackrel{~}{\mathrm{\Phi }}_j`$ can be induced in the quotient $`\text{𝔻}/\stackrel{~}{I}^2`$, $`j=1,2,3`$. The corresponding holomorphic 1-forms on $`\text{𝔻}/\stackrel{~}{I}^2`$ are denoted as $`\mathrm{\Phi }_1`$, $`\mathrm{\Phi }_2`$, and $`\mathrm{\Phi }_3`$, and they obviously satisfy 1, 2, 3 and 5 in the lemma statement.
Finally, the meromorphic function
$$g=\frac{\mathrm{\Phi }_1+i\mathrm{\Phi }_2}{\mathrm{\Phi }_3},$$
clearly omits the points $`\alpha `$, $`\beta `$, $`1/\overline{\alpha }`$, and $`1/\overline{\beta }`$, and 4 holds. This concludes the proof.
$`\mathrm{}`$
###### Lemma 2
There exists a rational function $`f:\overline{\text{}}\overline{\text{}}`$ satisfying:
1. The only poles of $`f`$ are $`0`$ and $`\mathrm{}.`$
2. $`fI=\overline{f}.`$
3. $`f(z)0`$, provided that $`|z|=1.`$
4. $`\text{Residue}({\displaystyle \frac{f(z)}{z}}dz,0)=0.`$
* Define
$$f:\text{}\text{},$$
$$f(z)=\frac{(zm_1)(zm_2)(m_1z+1)(m_2z+1)}{z^2},$$
where $`m_1,m_2\text{}.`$
We have
$$\text{Residue}(\frac{f(z)}{z}dz,0)=(1m_1^2)(1m_2^2)2m_1m_2.$$
The choice $`m_1=2`$ and $`m_2=\frac{2+\sqrt{13}}{3}`$ completes the proof. $`\mathrm{}`$
Now we are able to prove the main result of this paper.
###### Theorem 2
There exist complete nonorientable minimal surfaces in $`\text{}^3`$ whose generalized Gauss map omits two points of $`\text{}\text{}^2`$.
* Take $`A(R)`$, $`\mathrm{\Phi }_1`$, $`\mathrm{\Phi }_2`$, and $`\mathrm{\Phi }_3`$ as in Lemma 1, and $`f`$ as in Lemma 2. Put
$$\mathrm{\Phi }_j=\phi _j(z)\frac{dz}{z},$$
and write
$$\phi _j(z)=a_{j\mathrm{\hspace{0.17em}0}}+\underset{n>0}{}\left(a_{jn}z^n+(1)^{n+1}\overline{a_{jn}}z^n\right),a_{j\mathrm{\hspace{0.17em}0}}i\text{},$$
the Laurent series expansion of $`\phi _j`$, $`j=1,2,3`$.
Observe that
$$f(z)=\underset{n=1}{\overset{m}{}}\left(b_nz^n+(1)^n\overline{b_n}z^n\right),$$
where $`m\text{}^{}.`$ Let $`k\text{}`$, $`k`$ odd, $`k>m`$, and notice that:
$$\text{Residue}(\left[\underset{n>0}{}\left(a_{jn}z^{kn}+(1)^{n+1}\overline{a_{jn}}z^{kn}\right)\right]f(z)\frac{dz}{z},0)=0,j=1,2,3.$$
(4)
Furthermore, it is obvious from Lemma 2
$$\text{Residue}(a_{j\mathrm{\hspace{0.17em}0}}f(z)\frac{dz}{z},0)=0,j=1,2,3.$$
(5)
Consider the covering $`T_k:A(\sqrt[k]{R})A(R)`$, $`T_k(z)=z^k,`$ and define the holomorphic $`1`$-forms on $`A(\sqrt[k]{R})`$:
$$\mathrm{\Psi }_j=f(z)T_k^{}(\mathrm{\Phi }_j)=kf(z)\phi _j(z^k)\frac{dz}{z},j=1,2,3.$$
Taking into account (4) and (5), we deduce that $`\mathrm{\Psi }_j`$ is exact, $`j=1,2,3.`$
Moreover, it is clear that:
$$\underset{j=1}{\overset{3}{}}\mathrm{\Psi }_j^20,$$
and since $`k`$ is odd,
$$I^{}(\mathrm{\Psi }_1,\mathrm{\Psi }_2,\mathrm{\Psi }_3)=(\overline{\mathrm{\Psi }_1},\overline{\mathrm{\Psi }_2},\overline{\mathrm{\Psi }_3}),$$
(6)
where $`I:A(\sqrt[k]{R})A(\sqrt[k]{R})`$ is the lift of the former involution in $`A(R)`$, that keeps being the map $`I(z)=1/\overline{z}.`$
Note that $`lim_k\mathrm{}\sqrt[k]{R}=1`$, and remember that the zeroes of $`f`$ are not in $`\text{𝕊}^1`$. Then, taking $`k`$ large enough, we can guarantee that $`f`$ never vanishes in the closure of $`A(\sqrt[k]{R}).`$ So, as the only poles of $`f`$ are $`0`$ and $`\mathrm{}`$, there exist $`c>1`$ such that
$$\frac{1}{c}<|f(z)|<c,zA(\sqrt[k]{R}).$$
Therefore, $`_{j=1}^3|\mathrm{\Psi }_j|^20,`$ and if we define $`ds_0^2=|\mathrm{\Psi }_1|^2+|\mathrm{\Psi }_2|^2+|\mathrm{\Psi }_3|^2`$, one has:
$$\frac{1}{c^2}T_k^{}(ds^2)ds_0^2c^2T_k^{}(ds^2).$$
Since $`ds^2`$ is complete, the same occurs for the metrics $`T_k^{}(ds^2)`$ and $`ds_0^2`$.
Summarizing, the minimal immersion
$$X:A(\sqrt[k]{R})\text{}^3,$$
$$X(z)=\text{Real}\left(_1^z(\mathrm{\Psi }_1,\mathrm{\Psi }_2,\mathrm{\Psi }_3)\right),$$
is well defined, complete, and its Gauss map $`gT_k`$ omits four points of $`\overline{\text{}}`$.
From (6), $`X`$ induces a minimal immersion of the Möbius strip $`A(\sqrt[k]{R})/I`$ in $`\text{}^3`$, and so the Theorem is proved.
$`\mathrm{}`$
|
no-problem/9812/physics9812006.html
|
ar5iv
|
text
|
# Causal, psychological, and electrodynamic time arrows as consequences of the thermodynamic time arrow
## Acknowledgement
This work was supported by the Ministry of Science and Technology of the Republic of Croatia under Contract No. 00980102.
|
no-problem/9812/cond-mat9812385.html
|
ar5iv
|
text
|
# Coulomb Charging Effects for Finite Channel Number
## Abstract
We consider quantum fluctuations of the charge on a small metallic grain caused by virtual electron tunneling to a nearby electrode. The average electron number and the effective charging energy are determined by means of perturbation theory in the tunneling Hamiltonian. In particular we discuss the dependence of charging effects on the number $`N`$ of tunneling channels. Earlier results for $`N1`$ are found to be approached rather rapidly with increasing $`N`$.
Single electron effects are well studied for metallic junctions in the region of weak tunneling . In the case of a large tunneling conductance $`G_T`$ comparable or even larger than the conductance quantum $`G_K=e^2/h`$, it is necessary to go beyond lowest order perturbation theory in the tunneling Hamiltonian \- . In this case the theory involves contributions of order $`1/N`$ where $`N`$ is the number of transport channels. Since the junction area is typically much larger than $`\lambda _F^2`$, where $`\lambda _F`$ is the Fermi wavelength, the number of tunneling channels $`N`$ is often very large for metallic junctions, typically about $`10^4`$. Terms of order $`1/N`$ are thus dropped in most former approaches. In contrast to lithographically fabricated metallic junctions, in break junctions only a small number of channels may be available and terms proportional to $`1/N`$ cannot be neglected. Also for tunnel barriers in a two-dimensional electron gas there are typically only $`2`$ spin degenerate transport channels contributing to the tunneling current. Thus, it is interesting to investigate how Coulomb blockade effects are modified by such $`1/N`$ corrections. Here we focus on systems where the dimensionless conductance per channel, $`g_0=G_T/4\pi ^2NG_K`$, remains small, that is to cases where the channels contributing to charge transfer are weakly transmitting. Further we restrict ourselves to the zero temperature case. This covers only partly the range of experimental interest but the results indicate how relevant $`1/N`$ corrections are.
Specifically, we consider a small metallic grain in between two bulk electrodes biased by a voltage source $`U_{\mathrm{ex}}`$. The circuit diagram of the system is depicted in Fig. 1a). Between the grain and the left bulk electrode electrons can tunnel while the right electrode couples purely capacitively. Thus no dc-current flows through the system. Due to the discreteness of the tunneling process, the excess charge of the grain can be shifted only by multiples of the elementary charge $`e`$, and we introduce an excess charge number $`n`$ which characterizes the island charge $`q=ne`$. As far as the tunneling conductance per channel is small compared to the conductance quantum $`G_K`$, electron tunneling can be described in terms of a tunneling Hamiltonian which will be treated as a perturbation. The unperturbed Hamiltonian for the Fermi liquids and the Coulomb energy reads
$$H_0=E_c(nn_{\mathrm{ex}})^2+\underset{k\sigma }{}ϵ_{k\sigma }a_{k\sigma }^{}a_{k\sigma }+\underset{q\sigma }{}ϵ_{q\sigma }a_{q\sigma }^{}a_{q\sigma },$$
(1)
where $`n_{\mathrm{ex}}=C_GU_{\mathrm{ex}}/e`$ is a dimensionless voltage and
$$E_c=\frac{e^2}{2(C_T+C_G)}$$
(2)
is the charging energy needed to transfer one electron to the uncharged island at $`n_{\mathrm{ex}}=0`$. $`ϵ_{k\sigma }`$ and $`ϵ_{q\sigma }`$ are one-particle-energies for channel index $`\sigma `$ and longitudinal quantum number $`k`$ and $`q`$, respectively. The operator $`a_{k\sigma }^{}`$ ($`a_{k\sigma }`$) describes creation (annihilation) of a quasiparticle with quantum numbers $`k\sigma `$, and $`a_{q\sigma }^{}`$, $`a_{q\sigma }`$ are defined likewise. With these operators the tunneling Hamiltonian may be written as
$$H_T=\underset{kq\sigma }{}\left(t_\sigma a_{k\sigma }^{}a_{q\sigma }\mathrm{\Lambda }+\text{h.c.}\right)$$
(3)
where $`t_\sigma `$ is the tunneling matrix element and $`\mathrm{\Lambda }`$ shifts the island charge number $`n`$ thereby changing the Coulomb energy.
At zero temperature the system is described by the ground state energy $``$ and in view of eq. $`(\text{1})`$ the expectation value of the island charge may be written
$$n=n_{\mathrm{ex}}\frac{1}{2E_c}\frac{}{n_{\mathrm{ex}}}.$$
(4)
To calculate $``$ we make use of Rayleigh-Schrödinger perturbation theory. The perturbation $`(\text{3})`$ contains products of one creation and one annihilation operator, thus only even powers in $`t_\sigma `$ contribute to $``$, and it may be written
$$=^{(0)}+^{(2)}+^{(4)}+𝒪(t_\sigma ^6).$$
(5)
The zeroth order term $`^{(0)}`$ without tunneling is given by the minimum of the electrostatic energy and reads $`E_c(n_0n_{\mathrm{ex}})^2`$ where $`n_0`$ is the integer closest to $`n_{\mathrm{ex}}`$. Hence, the averaged island charge in zeroth order perturbation theory is given by the well known Coulomb staircase. $``$ depends on $`n_{\mathrm{ex}}`$ only via the electrostatic energy. It is thus an antisymmetric and quasiperiodic function of $`n_{\mathrm{ex}}`$ which allows us to confine ourselves to $`0n_{\mathrm{ex}}<{\scriptscriptstyle \frac{1}{2}}`$. For the second order term we get formally
$$^{(2)}=0|H_TQH_T|0$$
(6)
with the auxiliary operator
$$Q=\frac{1|00|}{_0H_0}.$$
(7)
In the same way the fourth order term reads
$$^{(4)}=0|H_TQH_TQH_TQH_T|00|H_TQ^2H_T|00|H_TQH_T|0$$
(8)
where terms with the energy denominator squared arise from the normalization of the ground state wave function. Inserting the tunneling Hamiltonian $`(\text{3})`$ into the second order contribution $`(\text{6})`$ one gets
$$^{(2)}=\underset{kq\sigma }{}t_\sigma ^2\left[\frac{\mathrm{\Theta }(ϵ_{q\sigma })\mathrm{\Theta }(ϵ_{k\sigma })}{\delta E_1+ϵ_{k\sigma }ϵ_{q\sigma }}+\frac{\mathrm{\Theta }(ϵ_{q\sigma })\mathrm{\Theta }(ϵ_{k\sigma })}{\delta E_1ϵ_{k\sigma }+ϵ_{q\sigma }}\right]$$
(9)
with the Coulomb energy differences $`\delta E_n=E_c(n^22nn_{\mathrm{ex}})`$. Both summands correspond to the virtual creation of an electron-hole pair with electron and hole sitting on different electrodes, cf. Fig. 1b). This can be represented in terms of Goldstone graphs depicted in Fig. 2. The contributions of second order correspond to the diagrams in Fig. 2a) where the upper arc with an arrow to the right (left) electrode represents the creation of an electron-hole pair with the electron sitting on the right (left) and the hole sitting on the other electrode of the junction. The lower arc destroys this pair and we may omit the arrow since the process is uniquely determined by the upper one.
If we assume that the one particle energy is separable into a longitudinal and channel part, the summand in eq. $`(\text{9})`$ is independent of the channel $`\sigma `$ apart from the factor $`t_\sigma ^2`$. The sum over the transversal and spin quantum numbers leads to an overall factor $`N`$ multiplied by the average $`t^2=_\sigma t_\sigma ^2/N`$ of the tunneling matrix elements. We assume the bandwidth $`D`$ to be large compared to $`E_c`$ and take the limit $`D/E_c\mathrm{}`$ at the end of the calculation. To keep the finite bandwidth in intermediate formulas, we introduce an exponential cutoff and replace the sum over the longitudinal quantum numbers by
$$\underset{k}{}F(ϵ_{k\sigma })\rho _{\mathrm{}}^{\mathrm{}}dϵe^{|ϵ|/D}F(ϵ)$$
(10)
where $`\rho `$ is the density of states at the Fermi level. Analogously, we introduce an integral with $`\rho ^{}`$ and $`D`$ for the $`q`$ sum. Formula $`(\text{9})`$ then takes the form
$$^{(2)}=t^2\rho \rho ^{}N_0^{\mathrm{}}dϵϵe^{|ϵ|/D}\left(\frac{1}{\delta E_1+ϵ}+\frac{1}{\delta E_1+ϵ}\right).$$
(11)
We now define the dimensionless tunneling conductance per channel by $`g_0=t^2\rho \rho ^{}`$. Whereas the integral $`(\text{11})`$ is divergent for $`D\mathrm{}`$, the average island charge to first order in $`g_0`$ remains finite leading to
$$n=g_0N\mathrm{ln}\left(\frac{1+2n_{\mathrm{ex}}}{12n_{\mathrm{ex}}}\right)+𝒪(g^2)$$
(12)
in accordance with previous results .
Along these lines the fourth order term $`(\text{8})`$ is given in terms of double sums over $`kq\sigma `$. These contributions can also be represented graphically as twofold virtual electron-hole pair creation and annihilation. We distinguish three groups of graphs depicted in Figs. 2b), c) and d). The first set of the graphs represents virtual electron hole pair creation and annihilation in two distinct channels $`\sigma `$ and $`\sigma ^{}`$ which are only coupled by the energy denominator. Fig. 2c) represents graphs arising from the normalization of the wave function where the dash signs that the energy denominator is squared, cf. eq. $`(\text{8})`$. If there were no Coulomb interaction, i.e. $`E_c=0`$, the graphs b) and c) would cancel in accordance with the linked cluster theorem for uncorrelated fermions. Here we have to take these terms into account since the Coulomb energy correlates the electrons in the two electrodes. The channels of the graphs in b) and c) are not restricted and we get by summation over the transversal quantum numbers a factor $`N^2`$. On the other hand, the graphs in Fig 2d) describe processes within one channel because the electron created recombines not with the hole created at the same time but with a different one which has to have the same channel number. Thus we get by summation over the channels only a factor $`N`$. Therefore, we write the average island charge in the form
$$n=g_0N\mathrm{ln}\left(\frac{1+2n_{\mathrm{ex}}}{12n_{\mathrm{ex}}}\right)+(g_0N)^2[c(n_{\mathrm{ex}})c(n_{\mathrm{ex}})]+g_0^2N[d(n_{\mathrm{ex}})d(n_{\mathrm{ex}})]+𝒪(g_0^3).$$
(13)
Previous theories for $`N1`$ introduce the dimensionless conductance $`g=g_0N`$ of the junction and the terms proportional to $`g_0^2N=g/N`$ are dropped.
The graphs in Fig. 2 b) and c) do not include $`1/N`$-corrections and they lead to the known second order result for the island charge
$`c(u)`$ $`=`$ $`u\left[{\displaystyle \frac{4\pi ^2}{3}}+\mathrm{ln}^2\left({\displaystyle \frac{12u}{1+2u}}\right)\right]{\displaystyle \frac{16(1+2u2u^2)}{(32u)(1+2u)}}\mathrm{ln}(12u)`$ (15)
$`2(1u)\left\{\mathrm{ln}^2\left[{\displaystyle \frac{12u}{4(1u)}}\right]+2\text{Li}_2\left[{\displaystyle \frac{32u}{4(1u)}}\right]{\displaystyle \frac{8(1u)}{(12u)(32u)}}\mathrm{ln}[4(1u)]\right\}.`$
The graphs in Fig. 2 d) can also be integrated out analytically leading to a new contribution of order $`1/N`$. We find
$`d(u)`$ $`=`$ $`{\displaystyle \frac{8}{3}}\mathrm{ln}^3\left({\displaystyle \frac{44u}{12u}}\right){\displaystyle \frac{26}{3}}\mathrm{ln}^3(12u)+15\mathrm{ln}^3(32u)`$ (20)
$`+[15\mathrm{ln}(2)+16\mathrm{ln}(44u)]\mathrm{ln}^2\left({\displaystyle \frac{12u}{32u}}\right)+[2\mathrm{ln}(1+2u)7\mathrm{ln}(32u)]\mathrm{ln}^2(12u)`$
$`+\left[10\mathrm{ln}^2\left({\displaystyle \frac{44u}{12u}}\right){\displaystyle \frac{10\pi ^2}{3}}+38\mathrm{ln}^2(32u)\right]\mathrm{ln}\left({\displaystyle \frac{12u}{32u}}\right)+{\displaystyle \frac{4}{3}}\pi ^2\mathrm{ln}\left({\displaystyle \frac{32u}{44u}}\right)`$
$`+4\mathrm{ln}\left({\displaystyle \frac{12u}{32u}}\right)\left\{3\mathrm{Li}_2\left({\displaystyle \frac{12u}{32u}}\right)+3\mathrm{Li}_2\left({\displaystyle \frac{32u}{44u}}\right)+2\mathrm{Li}_2\left[{\displaystyle \frac{8(1u)}{(32u)^2}}\right]\right\}`$
$`+6\mathrm{Li}_3\left({\displaystyle \frac{12u}{32u}}\right)8\mathrm{Li}_3\left({\displaystyle \frac{32u}{44u}}\right)8\mathrm{Li}_3\left({\displaystyle \frac{12u}{44u}}\right)`$
where $`\mathrm{Li}_2(z)`$ is the dilogarithm and $`\mathrm{Li}_3(z)`$ the trilogarithm function . Within second order perturbation theory the result (13) is valid for arbitrary channel numbers including $`N=1`$.
We now compare our result to earlier findings. One of the mostly frequently discussed quantities is the effective charging energy $`E_c^{}`$ characterizing the effective strength of the Coulomb blockade effect. This quantity is defined by
$$E_c^{}/E_c=\frac{1}{2}\frac{^2}{n_{\mathrm{ex}}^2}|_{n_{\mathrm{ex}}=0}.$$
(21)
For small tunneling conductance, $`g_00`$, the effective charging energy approaches the bare charging energy $`E_c`$ whereas for strong electron tunneling $`E_c^{}`$ vanishes. Our analytic expression leads to
$$E_c^{}/E_c=14g+5.066\mathrm{}g^27.167\mathrm{}g^2/N$$
(22)
where the constants are analytically known but too lengthy to present here. In Fig. 3 we show the normalized effective charging energy in first and second order perturbation theory without $`1/N`$ corrections compared to the complete second order result as a function of the dimensionless conductance $`G_T/G_K`$. We see that the $`1/N`$ corrections become more significant for larger tunneling conductance.
In Fig. 4 the average island charge $`n`$ in first and second order in $`g_0`$ with and without the $`1/N`$ corrections is depicted for the case $`N=5`$ and $`G_T/G_K=5`$. We see that the $`1/N`$ corrections become significant especially for larger external voltages. From a comparison with Monte Carlo data one estimates that results of second order perturbation theory are reliable for $`n_{\mathrm{ex}}`$ up to $`0.3`$. In the vicinity of the step, i.e., for $`n_{\mathrm{ex}}{\scriptscriptstyle \frac{1}{2}}`$, finite order perturbation theory diverges and one has to sum diagrams of all order . From our eq$`(\text{20})`$ one sees that the $`1/N`$-corrections become relevant near $`n_{\mathrm{ex}}={\scriptscriptstyle \frac{1}{2}}`$ even for large $`N`$ since the qualitative behavior is changed. For $`n_{\mathrm{ex}}{\scriptscriptstyle \frac{1}{2}}`$, the leading $`N`$ terms of second order show the logarithmic divergence $`2(g_0N)^2\mathrm{ln}^2\delta `$ where $`\delta ={\scriptscriptstyle \frac{1}{2}}n_{\mathrm{ex}}`$, whereas the $`1/N`$ corrections lead to $`\frac{4}{3}g_0^2N\mathrm{ln}^3\delta `$ which eventually dominates the asymptotic behavior. For the special case $`N=1`$, this divergence is in accordance with earlier findings by Matveev . In general, for the finite $`N`$ corrections of order $`g_0^k`$ we get the leading asymptotic behavior
$$ng_0^kN\mathrm{ln}^{2k1}\delta +𝒪(g^{k+1})$$
(23)
arising from the diagram depicted in Fig. 5. These terms dominate the asymptotics for $`n_{\mathrm{ex}}{\scriptscriptstyle \frac{1}{2}}`$, even for large $`N`$.
In summary, we have studied the influence of finite channel numbers on the Coulomb staircase. We found that for small voltages, $`n_{\mathrm{ex}}0`$, the $`1/N`$ corrections are significant up to $`N6`$ and become increasingly important as $`n_{\mathrm{ex}}={\scriptscriptstyle \frac{1}{2}}`$ is approached. While near the step finite order perturbative results as derived here are not sufficient, the expression obtained for the effective charging energy is valid within an experimentally relevant range of parameters.
This work was supported by the Deutsche Forschungsgemeinschaft (Bonn) and the Deutscher Akademischer Austauschdienst (DAAD).
|
no-problem/9812/cond-mat9812244.html
|
ar5iv
|
text
|
# The peculiarities of phase transitions in layered ferroelectrics-semiconductors 𝑇𝑙𝐼𝑛_{1-𝑥}𝐹𝑒_𝑥𝑆₂
## 1 Introduction
According to data of neutron diffraction measurements and x-ray investigations the monoclinic modification of layered ferroelectrics -semiconductors $`TlInS_2`$ undergoes a sequence of structural phase transitions with the intermediate incommensurate phase having the modulation vector $`\stackrel{}{q}_{inc}=(\delta ;\delta ;0.25),`$ where $`\delta =0.012`$.
$`TlInS_2`$ crystals have the space group symmetry $`C_{2h}^6`$ at the high-symmetric phase $`\left(T>220K\right)`$. The ionic-covalent bonds in layers and ones of Van-der-Waals type between layers are competive interactions in the structure of these crystals . Inproper ferroelectric phase transitions in $`TlInS_2`$ are intermediate between types of mixing and order-disorder with the characteristic value of Curie’s constant $`C10^3K`$ .
An analysis of experimental data by the temperature dependence of dielectric constant for various samples $`TlInS_2`$ with monoclinic structure shows that the $`ϵ\left(T\right)`$ curves characterized for proper ferroelectrics with incommensurate phase acquire the form characteristic for inproper ferroelectrics with incommensurate phase (Fig.1, curves 1,2,3). The detailed study of results of neutron diffraction measurements , x-ray , nuclear magnetic resonance and nuclear quadrupole resonance investigations allows to conclude that layered compounds $`TlInS_2`$ depending on concentration of defects in the structure have different character of phase transitions. Besides the considerable mobility of atoms in an interlayer space imparts to the structure the properties of weak glass .
At certain value of concentration of defects in the structure the system losing the property of layerity becomes isotropic. The range of existence of incommensurate phase and the position of Lifshits’ points on the diagram of states depend on concentration of defects in the structure. For the most widespread group of samples $`TlInS_2`$ the dependence $`ϵ\left(T\right)`$ is characterized by the evistenge of two maxima and the value of low temperature peak varies in the large range (Fig.1, curve 2) .
The anomaly shows itself on the dependence $`ϵ\left(T\right)`$ in the range of existence of incommensurate phase at $`T_{ic}204K`$ in addition to maxima at $`T_i`$ and $`T_c`$ for frequently occurred and detailly investigated samples $`TlInS_2`$. Experimental investigations show that the incommensurate phase springs up in the temperature range in which the spontaneous polarization differs from zero .
In seldom occurred samples $`TlInS_2`$ (with large concentration of defects) the dependence $`ϵ\left(T\right)`$ (Fig.1, curve 1) is typical for isotropic inproper ferroelectrics like $`Rb_2Znbr_4`$ . In this work the results of differential-thermal analysis (DTA) of various samples $`TlInS_2`$ as well as dielectric measurements of $`TlIn_{1x}Fe_x^{56,57}S_2`$ are presented.
## 2 The Results of Differential-Thermal Analysis and Dielectric Measurements
Single crystals of $`TlInS_2`$ compound were grown by modificated Bridgman method at small variation of growth conditions. For dielectric measurements the samples had the form of parallelepiped with parallel edges orientated along crystallographic directions. The silver paste was used as contacts put on polished surfaces. The measurements of dielectric constant $`ϵ`$ were carried out on the frequency $`1kHz`$ in the temperature range $`100300K`$. The rate of temperature change was $`0.1K/min`$.
The results of DTA of various samples $`TlInS_2`$ are presented on Fig. 2. The samples with one maximum on the $`ϵ(T)`$ dependence show only one endothermal effect at $`1020K`$ on the DTA curve stipulated by melting with the melting heat $`155.75KJ/mol`$. The samples $`TlInS_2`$ with two maxima on the $`ϵ(T)`$ dependence have two endothermal effects at temperatures $`920`$ and $`1060K`$ with the melting heats $`80KJ/mol`$ and $`125KJ/mol`$, respectively. Two endothermal effects at temperatures $`920`$ and $`1020K`$ with the melting heats $`78`$ and $`135KJ/mol`$, respectively, are detected on the DTA curve for samples displaying three maxima in the temperature dependence of dielectric constant.
The results of dielectric measurements for compounds $`TlIn_{1x}Fe_x^{56,57}S_2`$, where $`x=0;0.01;0.1`$, at both the regime of cooling and the regime of heating are presented on Fig. 3. An analysis of experimental data of temperature dependence of dielectric constant $`ϵ(T)`$ allows to conclude:
-at any $`x`$ the values of $`ϵ(T)`$ in the regime of heating are more than in the regime of cooling;
-at any $`x`$ the $`ϵ(T)`$ dependence displays two maxima, and the temperature interval between them does not almost depend on temperature;
-at increasing $`x`$ the temperature interval of existence of incommensurate phase shifts to low-temperature region;
-at increasing $`x`$ the response of system decreases;
-at increasing $`x`$ the transitions at $`T_i`$ and $`T_c`$ acquire the character of blurred phase transition;
-with increasing $`x`$ the $`ϵ(T)`$ curve loses the shape and acquires the form characterized for inproper ferroelectrics with incommensurate phase;
-substitution of $`Fe^{57}`$ on $`Fe^{56}`$ does not change the temperature width of existence of incommensurate phase and shifts it to high temperatures.
## 3 Discussions
According to data of $`x`$-ray investigations the atoms $`Tl`$ in the elementary cell of crystals $`TlInS_2`$ occupy the general position . Besides these atoms have high mobility in the interplane space . Nuclear quadrupole resonance investigations indicate on the anomaly at temperature $`250K`$ . Evidently, the relative displacement of atoms $`Tl`$ in the elementary cell occurs at $`250K`$ and this deformation does not change the space group symmetry of the crystal. In addition, the deformation of the band structure of $`TlInS_2`$ and levels of adhesion take place. An increase of concentration of carriers decreases the temperature of transition to the incommensurate phase and increases the response of this system on the external influence. These facts explain the temperature dependence $`ϵ\left(T\right)`$ in the high-symmetric phase. Although experimental data about details of disordering in the compounds with blurred phase transitions are absent , the broadening of the phase transition is general phenomenon for solid solutions and other disordered structures.
With increasing of concentration of defects in the layered structure the blurring of phase transitions at $`T_i`$ and $`T_c`$ and their shift to low temperatures happen. At some threshold value the layerity of structure loses and the crystal becomes isotropic and distribution of defects does homogeneous. As known, if the disordering is homogeneous, a clear phase transitions observed even in amorphous structures .
An increase of a depth of minimum in the temperature dependence $`ϵ\left(T\right)`$ is the result of intensification of the constant of interaction of polarization with the amplitude of the order parameter. Consequently, with increasing $`x`$ the interaction of polarization with the amplitude of the order parameter rises while it decreases with the phase. Incidentally, the incommensurate phase behaves inself as modulated one in which the soliton regime is absent. The temperature hysteresis of $`ϵ\left(T\right)`$ connects with the temperature dependence of the band gap with accuracy to the energy of adhesive level and it connects with the inner mobility of impurities of the structure.
In the perfect layered samples $`TlInS_2`$ a sequence of structural phase transitions: highsymmetrical-incommensurate $`I`$-incommensurate $`II`$-commensurate phase, are sprung up.
With increasing of concentration of defects in the layered structure of $`TlInS_2`$ the line of Lifshits’ points on the states diagram shifts and the incommensurate phase $`II`$ in dielectric measurements is not observed.
With further rise of concentration of defects in the structure the compound $`TlInS_2`$ loses the property of layerity and the sample becomes isotropic. The temperature dependence of dielectric constant of isotropic samples at the transition from the high-symmetric phase to the incommensurate phase remains continuous, the incommensurate phase has not the minimum and behaves itself as $`ϵ\left(TT_c\right)^1`$. Consequently, the dependence of physical properties of $`TlInS_2`$ on concentration of defects is the result of existence of competing interactions in the structure.
The principal parameter determining the mechanism of melting is the energy required for the rupture of bonds in the structure . That is why, only one endothermal effect shows itself in the isotropic samples of $`TlInS_2`$. Layered samples $`TlInS_2`$, having the temperature dependence with two or three maxima, show themselves two endothermal effects on the DTA curve.
The opposite isotropic effect connected with substitution of $`Fe^{57}`$ on $`Fe^{56}`$ in the $`TlIn_{1x}Fe_xS_2`$ compounds is not the result of the rise of the tunneling frequency between two positions of equilibrium. The existence of blurred phase transition and the fact of a decrease of temperature of phase transition with a rise of concentration of defects testifies that the samples with $`Fe^{57}`$ more regulated than ones with $`Fe^{56}`$.
FIGURE CAPTIONS:
Fig.1. Temperature dependence of dielectric constant $`ϵ\left(T\right)`$ of crystals $`TlInS_2`$ (obtained from various technologic batches) in the regime of cooling
Fig.2. The DTA curves of crystals $`TlInS_2`$ (obtained from various technologic batches)
Fig.3. Temperature dependence of dielectric constant of solid solutions $`TlIn_{1x}Fe_xS_2`$:
curves $`1,2`$ corresponds to the value $`x=0`$
curves $`3,4`$ \- $`x=0.1`$ $`(Fe^{56})`$
curve $`5`$ \- $`x=0.01`$ $`(Fe^{57})`$
curves $`6,7`$ \- $`x=0.1`$ $`(Fe^{57})`$
curves $`1,3,5,7`$ are measured in the regime of cooling, curves $`2,4,6`$ -in the regime of heating.
|
no-problem/9812/cond-mat9812316.html
|
ar5iv
|
text
|
# The Origin of the Designability of Protein Structures
## Introduction
Natural proteins fold into unique compact structures in spite of the huge number of possible conformations. For most single domain proteins, each of these native structures corresponds to the global minimum of the free energy.
It has been proposed phenomenologically that the number of possible structures of natural proteins is only about one thousand, which suggests that many sequences can fold into one preferred structure. There have been theoretical studies for the existence of such preferred structures .
In many of theoretical studies for the protein folding, a simplified model called HP model is adopted. HP model is one of 2-letter codes lattice models where a protein is represented by a self-avoiding chain of beads placed on a lattice, with two types of beads, hydrophobic(H) and polar(P). In the HP model, the energy of a structure is given by the nearest-neighbor topological contact interactions as
$$H=\underset{i<j}{}E_{\sigma _i\sigma _j}\mathrm{\Delta }(r_ir_j)$$
(1)
where $`i`$ and $`j`$ are monomer indexes, $`\{\sigma _i\}`$ are monomer types ($`\sigma =`$ H or P); $`\mathrm{\Delta }(r_ir_j)=1`$ if $`r_i`$ and $`r_j`$ are topological nearest neighbors not along the sequence, and $`\mathrm{\Delta }(r_ir_j)=0`$ otherwise.
Based on the HP model, a concept of designability has recently been introduced; the number of sequences that have a given structure as their non-degenerate ground state (native state) is called the designability of this structure. When many sequences have a common native structure, one say that the structure is highly designable. Adding to the importance in the protein design problem, the designability also have evolutional significance because highly designable structures are found to be relatively stable against mutations.
In the original study of H. Li et al., HP models on the square and cubic lattices are employed, with the energy parameters in eq.(1) being $`E_{HH}=2.3,`$ $`E_{HP}=1.0,`$ $`E_{PP}=0.0`$. For each sequence, they calculated the energy over all maximally compact structures and picked up the native structure. The results indicated that highly designable structures actually exist on both lattices.
A. Irbäck and E. Sandelin studied the HP models on the square and triangular lattices. They adopted different energy parameters from H. Li et al. , namely, $`E_{HH}=1,`$ $`E_{HP}=E_{PP}=0`$. In the calculation of the designability, they considered all the possible structures, not restricting to the maximally compact ones. For the square lattice, they confirmed the existence of the highly designable structures as in Ref.. For the triangular lattice, however, no such structures were found. In addition to the nearest-neighbor topological contact interactions, they considered local interactions represented by the bend angle and calculated the designability. Indeed the local interactions reduced degeneracy (i.e., the number of sequences which have non-degenerate ground state increased) and made the designability higher. But they found that the designability on the square lattice was still much higher than that on the triangular lattice. They concluded that the difference in the designability for these two lattices are related to the even-odd problem, that is, whether the lattice structure is bipartite or not.
Quite recently, H. Li et al. proposed a new model based on the HP model on the square lattice . In the model, the hydrophobic interaction is treated in such a way that the energy decreases if the hydrophobic residue is buried in the core. They justify this treatment in two reasons: (1) the hydrophobic force which is dominant in folding originates from aversion of hydrophobic residues from water. (2) the Miyazawa-Jernigan matrix contains a dominant hydrophobic interaction of the linear form $`E_{\alpha \beta }=h_\alpha +h_\beta `$ . They took
$$H=\underset{i=1}{\overset{N}{}}s_ih_i$$
(2)
where $`\{h_i\}`$ represent a sequence : $`h_i=1`$ if the i-th amino acid is H-type and $`h_i=0`$ if it is P-type. And $`\{s_i\}`$ represent a structure : $`s_i=0`$ if the i-th amino acid is on the surface and $`s_i=1`$ if it is in the core. They calculated the designability over all maximally compact structures, whose result is consistent with their former study \[See Table. I \].
In our view, there are many points to be explored further for the designability problem. First, since the structures of natural proteins are compact but not necessarily “maximally compact” in general, how can we justify the discussion where only the maximally compact structures are taken into account? Second, is it adequate to consider only nearest-neighbor interactions? Properties of a system with only nearest-neighbor interactions are directly affected by the lattice structure, in particular, whether the lattice is bipartite or not. Is it good, only from these facts, to conclude immediately that the absence of the highly designable structures on the triangular lattice should be ascribed to the even-odd problem associated with the triangular lattice? One should discuss the problem on the triangular lattice by using a model like the one in Ref. where the interactions do not depend on the contact between monomers, hence, do not directly reflect the non-bipartiteness.
Our aim of this paper is to examine the above points and clarify what determines the designability of protein structures. We use a new model with a 2-letter codes (H and P) on the square and triangular lattices and calculate the designability over all possible structures. In our model, based on Ref., the energy increases if the hydrophobic residue is exposed to the solvent. We will call this model “solvation model”. In brief, the solvation model is a 2-letter codes lattice model where the hydrophobic force to form a core is dominant and the interactions do not directly reflect the bipartiteness. Using the solvation model and the HP model, we investigate model-independent properties of designability.
## The Models
In the solvation model, based on Ref. , a protein is represented by a self-avoiding chain of beads with two types H and P, placed on a lattice. A sequence is specified by a choice of monomer types at each position on the chain.
We used two-dimensional lattice models because a computable length by numerical enumeration of the full conformational space is limited (square lattice : 18, triangular and cubic lattices : 13). Even with this chain-length limitation, we can make a “hydrophobic core” in two dimensions, in contrast with the three-dimensional case.
A structure is specified by a set of coordinates for all the monomers and is mapped into the number of contacts with the solvent. In our model, the total energy is given in terms of the monomer-solvent interactions, and depends only on the number of contacts with the solvent:
$$H=\underset{i=1}{\overset{N}{}}E_{s_i}h_i$$
(3)
where $`\{h_i\}`$ represent a sequence : $`h_i=1`$ if the i-th monomer is the H-type and $`h_i=0`$ if it is P-type. The variable $`s_i`$ denotes the number of contacts with the solvent, for example, $`s_i=\{0,1,2,3\}`$ on the square lattice and $`s_i=\{0,1,2,3,4,5\}`$ on the triangular lattice. In other words, $`s_i=0`$ means that the $`i`$-th monomer is buried away from the solvent. We take $`E_0=0,`$ $`E_1=\sqrt{2},`$ $`E_2=\sqrt{7},`$ $`E_3=\sqrt{13},`$ $`E_4=\sqrt{19},`$ $`E_5=\sqrt{23}`$. That is, the possible minimum energy is zero. And these parameters are selected so that the larger the number of contacts with the solvent is, the more the degree of energy increase is; the hydrophobic residue is energetically unfavorable to be at the corner. Although the choice of these values is somewhat arbitrary, we have considered the following points: (1) these values should not increase too much rapidly with the increase in the number of contacts with the solvent, and (2) the way of choosing these values must not bring about nonessential accidental degeneracies (due to simple rational ratios between the parameters) .
Using the model on the square and triangular lattices, we calculate the designability for all the $`2^N`$ sequences, where $`N`$ is the number of monomers, by exact computer-enumeration method over the full conformational space. To get correct data, we exclude overcounting coming from redundant structures which are mutually related by rotation, reflection and reverse-labeling.
On the basis of data obtained by the solvation model and the HP model, we examine what determines the designability from three points of view: (1) the effect of the search-space restriction, namely, the search within maximally compact structures (in this paper, we just used maximally compact structures as a simplest example of the search-space restriction, and we may consider other one, e.g., structures with the biggest core), (2) the effect of the lattice structure, namely, whether the lattice is bipartite or not (or, equivalently, the even-odd problem), (3) the effect of the number of monomers (or, the length of the chain).
## Results and Discussion
Let us now give results of calculations.
(1) The effect of the search within maximally compact structures
In Fig. 1, we show the designability calculated on the square lattice for $`N=16`$, using maximally compact structures. For comparison, in Fig. 2, we show the designability of the same system without the search-space restriction (i.e., search over all possible structures). In both cases, there are some highly designable structures. However, these structures are not common to both cases. In Fig. 2, the number of sequences that have native structures is 8277, but the number of sequences that have maximally compact structures as native is only 1087 out of 8277. That is, most sequences that have native structures have non-maximally compact structures as native. The importance of non-maximally compact structures has also been pointed out for the HP model . These facts imply that it is not good to calculate the designability over only maximally compact structures. Such calculation picking up a “native” structure out of maximally compact structures, is not correct if the true native structure is non-maximally compact. Further, when the lowest-energy non-maximally compact structure and the lowest-energy maximally compact structure are degenerate, there is no native structure (native structure must be non-degenerate), but the restricted-search-space calculation gives a false result that there is a native (and maximally compact) structure. We should say that the designability calculated over only maximally compact structures may be erroneous.
(2) The effect of the lattice structure: bipartite or non-bipartite
In two previous studies using the HP model , interactions of the system directly reflected whether the lattice is bipartite or not. Moreover the designability on the triangular lattice was calculated with the energy parameters in eq.(1) being $`E_{HH}=1,E_{HP}=E_{PP}=0`$, which would cause accidental degeneracies. In their results, highly designable structures were not found for the triangular lattice. Also, it seemed that native structures are likely to contain the hydrophobic core where a group of hydrophobic monomers contact with each other; such contact can be made only if the distance between the monomers along the sequence is odd. Therefore, the bipartiteness has been thought to be a main source of the designability.. If so, highly designable structures do not actually exist, i.e., the concept of designability itself could be meaningless. On the other hand, if such preferred structures should exist on the basis of the proposal by C. Chothia , the use of the lattice model would be inadequate. Then, we used the solvation model, which does not directly reflect the bipartiteness, and calculated the designability on the square and triangular lattices. Besides, we also calculated the designability on the triangular lattice using the HP model, with the energy parameters being $`E_{HH}=2.3,E_{HP}=1.0,E_{PP}=0.0`$.
In Table. II, we show the total number of sequences that have non-degenerate ground state ($`S_n`$) and the highest designabilities ($`D_h`$) on the triangular lattice for $`N=13`$, obtained by using different interactions. This result shows that, even if we take different values of energy parameters, or even if we use the solvation model, the triangular lattice is still unfavorable for the designability although $`S_n`$ varies largely. On the other hand, for the square lattice, highly designable structures are found in the solvation model as well as in the HP model (Fig. 2). These results imply that the absence of the highly designable structures for the triangular lattice should not be ascribed to the even-odd problem (or, the non-bipartiteness), but to other reasons. The properties that highly designable structures are found on the square lattice and no such structures are found on the triangular lattice might be general in 2-letter codes lattice models where the hydrophobic force is dominant.
(3) The effect of the number of monomers
Then, why are the highly designable structures absent for the triangular lattice? Smallness of number of monomers (in other words, the length of a chain is too short), may be a possible reason. Important object in the protein structure is the hydrophobic core which consists of buried monomers in no contact with the solvent. Recall that the limit of a computable length by exact enumeration of the full conformational space on the triangular lattice is 13. The biggest core which we can make by using this limited length is the one which consists of only three monomers; the length is too short for the hydrophobic force to form a core. This monomer-number effect is also found on the square lattice. Consider the following conditions: at least ten sequences have a given structure as their native state, and at the same time, there are at least five such structures. Only if these conditions are satisfied, let us say that “there are highly designable structures.” Then, at $`N=10`$ or less, there are no highly designable structures even for the square lattice \[Table. III, Table. IV\]. This result implies that, when we discuss whether there are highly designable structures or not, we need a long chain enough to make a core of enough size. This further implies that, in three-dimensional case, we will need a chain of longer length than that in two-dimensional case to make a core.
Let us see Table. III, Table. IV and Table. V. In Table. V, we show the designability calculated on the triangular lattice for $`N=13`$. On the square lattice for $`N=10`$, the biggest core consists of two monomers. Both on the triangular lattice for $`N=13`$ and on the square lattice for $`N=11`$, the biggest core consists of three monomers. We see that the triangular lattice is unfavorable for designability compared with square lattice, even when the biggest possible core size is same or a little larger. A possible reason would be the number of all possible structures, particularly the number of structures with the biggest core. As the length of a chain becomes long, the number of all possible structures increases almost exponentially as $`\mu ^N`$ ( $`2<\mu <3`$ for the square lattice, and $`4<\mu <5`$ for the triangular lattice). On the triangular lattice for $`N=13`$, the number of all possible structures is 6,279,601 and the number of structures with the biggest core is 4,110 out of them. On the other hand, on the square lattice for $`N=10,11`$, the number of all possible structures is 2,034, 5,513 and the number of structures with the biggest core is 23, 5, respectively. Thus the number of all possible structures and the number of structures with the biggest core on the triangular lattice are much larger than those on the square lattice . In consequence, the degeneracy tends to grow, which is unfavorable for designability. In this view, designable structures on the triangular lattice would be more difficult to appear than on the square lattice.
## Summary
We have calculated the designability of the protein structure using the solvation model and the HP model, to deduce model-independent properties of designability. The solvation model introduced in this paper satisfies two conditions: (1)the hydrophobic force is dominant, (2) the model does not directly reflect the bipartiteness. We have examined what determines the designability from three points of view: effect of restricted search within maximally compact structures, the bipartite/non-bipartite effect, the length of the chain.
In result, we have found that it is inadequate to calculate the designability within maximally compact structures. Our results imply that the reason why no highly designable structures on the triangular lattice have been found is not the non-bipartiteness. We suppose that the main factor which affects the designability is the chain length, because for sufficiently large hydrophobic core to form, long enough chains are required. Triangular lattice is more unfavorable for the designability than square lattice irrespective of models or energy parameters, probably because the number of all possible structures is large. However, if we can deal with longer chain than in the present study, it is possible that we find highly designable structures even on the triangular lattice. The calculations of the designability for longer chains on the triangular lattice are highly desirable. These conclusions would apply to a wide variety of 2-letter codes lattice models, where the hydrophobic force is dominant, regardless of energy parameters and further details of the model.
Though a concept of designability is currently defined for a 2-letter codes lattice model, our final goal is to examine whether natural proteins have highly designable structures. Therefore it is an interesting problem to extend the study of the designability for a 20-letter codes model (e.g., MJ model, KGS model) and an off-lattice model. Substituting 20-letter codes for 2-letter codes certainly reduces degeneracy, and most of all sequences come to have a structure as non-degenerate ground state (i.e., native structure).
## Acknowledgements
We would like to thank Y. Akutsu and M. Kikuchi for useful discussions and careful reading of the manuscript.
|
no-problem/9812/hep-th9812015.html
|
ar5iv
|
text
|
# Untitled Document
hep-th/9812015
Molien Function for Duality
Philippe Pouliot
Department of Physics
University of California
Santa Barbara, CA 93106
Abstract
The Molien function counts the number of independent group invariants of a representation. For chiral superfields, it is invariant under duality by construction. We illustrate how it calculates the spectrum of supersymmetric gauge theories.
1. Introduction
It is quite remarkable that certain four dimensional gauge theories can be solved exactly. The examples that have been solved so far ( for a review) are quite special: they have lots of symmetries.
A generic theory does not have so many symmetries, so here I introduce a tool which I hope will be useful to the study of more general theories. To be concrete, I will consider only supersymmetric theories.
2. The Molien function
Consider a supersymmetric gauge theory with chiral superfields transforming as a representation $`R`$ of a group $`G`$. I make no restrictions on $`R`$: it can be reducible, and also contain singlets (mesons). Similarly, $`G`$ can be a product of groups, or it can be the identity, for a confining theory.
The Molien generating function for the representation $`R`$ is
$$M(z)=\underset{k=0}{\overset{\mathrm{}}{}}c_kz^k$$
where $`c_k`$ is the number of independent group invariant polynomials of order $`k`$. It is a holomorphic function.
It turns out that there is a nice way to write down $`M(z)`$ (see p. 204 for an easy proof):
$$M(z)=\frac{d\mu (g)}{det(1zR(g))}.$$
The idea of the proof is that one can diagonalize the unitary representation $`R`$ for any fixed group element $`g`$; then integration ($`𝑑\mu (g)`$) over the whole group picks out only the singlets in the tensor products $`R^k`$.
The function $`M`$ can be evaluated more explicitly (see the nice paper by Forger for much useful and readable complementary details.)
$$M(z)=\frac{1}{|W|}\mathrm{}\frac{dw_1}{2\pi iw_1}\mathrm{}\frac{dw_l}{2\pi iw_l}\frac{\mathrm{\Pi }_\alpha (1w^{h(\alpha )})}{\mathrm{\Pi }_\lambda (1zw^{h(\lambda )})}.$$
$``$ $`|W|`$ is the number of elements in the Weyl group
$``$ $`l`$ is the rank of the group;
$``$ the products are over all the roots $`\alpha `$ of the group and over all the weights $`\lambda `$ of the representation $`R`$;
$``$ and finally, the concise notation $`w^{h(\alpha )}`$ means $`w_1^{h(\alpha _1)}\mathrm{}w_l^{h(\alpha _l)}`$, where $`h(\alpha _i)`$ is the eigenvalue of the root $`\alpha `$ under the Cartan generator $`H_i`$.
Another representation of the Molien function coefficients is given in terms of an index:
$$c_k=\frac{1}{|W|}\underset{\stackrel{~}{\lambda }}{}i(\stackrel{~}{\lambda })m_k(\stackrel{~}{\lambda })$$
with the $`\stackrel{~}{\lambda }`$ denoting the extended weights and $`m_k`$ the $`k`$-extended multiplicities. This terminology is defined in . According to , the index might be more efficient for explicit calculations.
3. Duality<sup>1</sup> I wish to thank O. Aharony and A. Schwimmer for inquiries that led to a clarification of this section.
For a $`𝒩=1`$ supersymmetric gauge theory with a vanishing superpotential, the Molien function calculated for the gauge group of the theory encodes much information about the low-energy spectrum. It calculates how many gauge invariant independent chiral (holomorphic) operators there are of a given degree in the number of elementary fields. In other words, it contains much about the structure of the chiral ring.
Consider now a dual “magnetic” description to this theory. Since by assumption it has the same low-energy spectrum, there must be a way, expected to be complicated (i.e. not a Molien function!), to calculate the Molien function of the “electric” theory in terms of the data of the magnetic theory. In this sense the Molien function is duality invariant.
Since it is a holomorphic function, one would hope that the powerful tools of complex analysis can be useful to study its properties. This, however, is highly speculative, as well as the remaining of this paragraph. Even if it is not enough to fully characterize a supersymmetric conformally invariant gauge theory, the Molien function could be “the” characteristic function of $`𝒩=1`$ duality <sup>2</sup> (However, there is no claim of uniqueness here: with the meaning of duality invariance above, any formal function of variables $`z_\alpha `$ (one variable for each chiral invariant operator $`O_\alpha `$) will be duality invariant, as long as the constraints among the operators are suitably implemented by constraints among the $`z_\alpha `$, a feat that the Molien function accomplishes naturally. See section 4.). It is definitely interesting because it does not rely on global symmetries. For a generic theory, global symmetries are small and the constraints one can get from them have a limited power: it is well known that satisfying the ’t Hooft anomaly matching conditions is not enough. In string theory, there are no global symmetries anyway. Perhaps dynamical properties of the low-energy theory can be inferred from the chiral spectrum. This would come about by making the following statement more precise: the Molien function of a confining theory is simple, while the Molien function of a gauge theory which is not asymptotically free is complicated (it has high order syzygies among its invariants.)
If the theory has a non-zero superpotential, extra constraints are introduced among the invariants. The definition of the Molien function stays the same (namely $`M(z)=_{k=0}^{\mathrm{}}c_kz^k`$ where $`c_k`$ is the number of independent group invariant polynomials of order $`k`$), but the integral representation ($`\frac{d\mu (g)}{det(1zR(g))}`$) should be generalized to include the effect of the superpotential (I don’t know how to write it down).
4. Generalized Molien Function
It would be nice to have an explicit way to construct the Molien function of the electric from the magnetic theory. One might hope that there is a generalization of the Molien function, $`\stackrel{~}{M}`$, which is such that calculating $`\stackrel{~}{M}`$ in the electric and in the magnetic theory would give the same result. I do not know if this is possible.
As a step in this direction, it is convenient to define a generalized Molien function, still assuming that the superpotential is zero, by choosing a global $`U(1)`$ charge, under which the elementary fields in irreducible representation matrices $`R_i`$ transform with charges $`q_i`$, $`i=1,\mathrm{},n`$ (which can all be taken to be integers by suitable rescaling). With this, the generalized Molien function
$$M_{\{q_i\}}=𝑑\mu (g)/det\left(\begin{array}{c}1z^{q_1}R_1(g))\\ & 1z^{q_2}R_2(g))\\ & & \mathrm{}\\ & & & 1z^{q_n}R_n(g))\end{array}\right)$$
has the property that the invariant operators are counted with the same power of $`k`$ as coefficients of $`z^k`$ in the electric and in the magnetic theories. The coefficients will still disagree of course because the constraints from the superpotentials have not been included.
5. Illustration
Aside from the duality application, the Molien function provides a technique to grind out the spectrum of a theory, along with plethysms, branching rules and other counting arguments. I will illustrate some of the uses with the simplest examples. Start with $`𝒩=1`$ supersymmetric $`SU(2)`$ gauge theory with one flavor of fundamentals $`Q_i`$ (two doublets) .
Evaluating $`M(z)`$ with the integral representation readily gives
$$M=\frac{1}{1z^2}=1+z^2+z^4+\mathrm{}.$$
This generating function is characteristic of a freely generated ring with one invariant: there’s one polynomial of order 2, namely $`Q_1Q_2`$, and one of order 4, $`(Q_1Q_2)^2`$, and so on.
With 4 doublets,
$$M(z)=\frac{1z^4}{(1z^2)^6}=1+6z^2+20z^4+50z^6+\mathrm{}.$$
The coefficient $`6`$ indicates that the ring is generated by the invariants $`V_{ij}=Q_iQ_j`$. At order $`z^4`$, we learn that the $`V_{ij}`$ are not independent, but there is one constraint among them, the famed $`\mathrm{pf}\mathrm{V}=\mathrm{\Lambda }^4`$. Studying the following coefficients shows that there are no more constraints.
With 6 doublets,
$$M(z)=\frac{1+6z^2+6z^4+z^6}{(1z^2)^9}=1+15z^2+(12015)z^4+(6801891)z^6+\mathrm{}.$$
This is already more complicated. There are 15 invariants $`V_{ij}`$, and $`15`$ constraints (syzygies) $`ϵ^{ijklmn}V_{kl}V_{mn}`$, but there are constraints amongst the constraints and so on.
More generally, with $`d`$ doublets,
$$M_d=\underset{k=0}{\overset{\mathrm{}}{}}dim\left(\text{ }\text{ }\mathrm{}\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\text{ }\right)z^k$$
where the tensor under the $`SU(2d)`$ symmetry has $`k`$ horizontal boxes.
6. New Example
As the rank of the group increases, the formulas for the Molien function become rapidly cumbersome to evaluate. For the integral representation, one is faced with high order poles to be evaluated by the residue theorem. A trick is to settle for less than the full generating function, and get only the first few $`c_k`$: one takes derivatives with respect to $`z`$, and then set $`z=0`$, before evaluating the residues at the $`w_i`$, for which the poles are now automatically all at $`w_i=0`$. To go beyond that, perhaps one could reexpress these integrals using Littlewood’s Schur functions, or use the index formula of . Another possibility to evaluate the Molien function more effectively is to use the MacMahon algorithm <sup>3</sup> I thank C. Cummins for pointing out the usefulness of in this respect..
Here for simplicity, I will only calculate the spectrum of the supersymmetric $`SU(2)`$ gauge theories with one matter field in the 4-dimensional representation $`S`$ and $`2k`$ doublet fields $`Q_i`$.
$``$ $`k=0`$. This theory was studied in . There is just one invariant, quartic, so $`M=1/(1z^4)`$.
When $`k>0`$, the theories are not asymptotically free. That does not make them uninteresting, because they can still be the free duals of strongly coupled theories.
$``$ $`k=1`$
$$M=\frac{1z^2+5z^4z^6+z^8}{(1z^4)^3(1z^2)^2}$$
At this stage, we see the invariants $`Q^2`$, $`SQ^3`$, $`S^2Q^2`$, $`S^3Q^3`$ and $`S^4`$. The invariants $`SQ^3`$, $`S^2Q^2`$, $`S^3Q^3`$ are fully symmetric in their flavor indices. They generate the full ring, but they are not independent. Checking this result for $`k>1`$, we see that these invariants still form a full set, but there are more constraints.
$``$ $`k=2`$
$$M=\frac{1+2z^2+28z^4+23z^6+73z^8+23z^{10}+28z^{12}+2z^{14}+z^{16}}{(1z^4)^5(1z^2)^4}.$$
$``$ $`k=3`$
$`M=\frac{1+9z^2+101z^4+319z^6+1020z^8+1475z^{10}+2091z^{12}+1475z^{14}+1020z^{16}+319z^{18}+101z^{20}+9z^{22}+z^{24}}{(1z^4)^7(1z^2)^6}.`$
Acknowledments
This work is supported by NSF Grant PHY97-22022. I wish to thank O. Aharony, C. Cummins and A. Schwimmer for useful correspondence.
References
relax K. Intriligator and N. Seiberg, Lectures on Supersymmetric Gauge Theories and Electro-Magnetic Duality, Nucl. Phys. suppl. 45BC (1996) 1, hep-th/9509066. relax D. Sattinger and O. Weaver, Lie Groups and Algebras with Applications to Physics, Geometry and Mechanics, 1986 Springer-Verlag. relax M. Forger, Invariant Polynomials and Molien Functions, J. Math. Phys. 39 (1998) 1107. relax L. Begin, C. Cummins and P. Mathieu, Generating Functions for Tensor Products, hep-th/9811113. relax K. Intriligator, N. Seiberg and S. Shenker, Proposal for a Simple Model of Dynamical SUSY Breaking, Phys. Lett. B 342 (1995) 152, hep-th/9410203.
|
no-problem/9812/astro-ph9812031.html
|
ar5iv
|
text
|
# An RXTE study of M87 and the core of the Virgo cluster
## 1 Introduction
The nearest giant elliptical galaxy, M87 (NGC 4486), holds a central place in the study of low-luminosity radio galaxies and extragalactic radio jets. This galaxy, situated at the center of the Virgo cluster of galaxies, is associated with the Faranoff-Riley class I radio source Virgo-A and displays the most prominent extragalactic radio jet in the northern sky. It was the first extragalactic jet to be discovered (Curtis 1918) and has since been subjected to intense observational study at all available wavelengths (see Biretta 1993 for a review). The close proximity of this source, about 16 Mpc (e.g., Tonry 1991), makes it a crucial laboratory for testing our understanding of both extragalactic jets and the central engine structure of radio-loud AGN.
In the soft X-ray band, imaging with the high-resolution imagers (HRIs) on both the Einstein and ROSAT satellites have resolved emission from the core of M87 and knots A, B and D of its optical jet (Schreier, Gorenstein & Feigelson 1982; Biretta, Stern & Harris 1991; hereafter BSH91). The mechanisms underlying any of these emission components is unknown. Suggestions for the jet emission mechanism include synchrotron emission from ultra-relativistic ($`\gamma 10^7`$) electrons in the jet plasma, inverse Compton scattering of infra-red/optical photons by a population of $`\gamma 100`$ electrons in the jet plasma, and thermal bremsstrahlung from shock heated gas surrounding the jet. The observed core emissions could represent the inner jet with one of the above mechanisms producing the X-rays. On the other hand, emission from an accretion disk corona (as in the Seyfert case) or a hot accretion disk (such as an Advection Dominated Accretion Flow; Reynolds et al. 1996) might also be important for understanding the core emissions.
For years there was a mystery surrounding the hard X-ray emission from M87 and the Virgo cluster. Whereas the centroid of the low-energy emission lies on M87, Davison (1978) used Ariel-V data to show that the higher-energy emissions possessed a different centroid (displaced to the north-west by $`1^{}`$). This puzzle was resolved by the coded-mask imaging from Spacelab-2 which found that the high-energy emissions ($`10\mathrm{keV}`$ or greater) of the Virgo cluster are dominated by the Seyfert 2 galaxy NGC 4388 (Hanson et al. 1990). NGC 4388 is displaced from M87 by just over a degree to the north-west. An unambiguous measurement of the hard X-ray flux, spectrum and variability properties of M87 has proven difficult due to the presence of this confusing source. Takano & Koyama (1991) analyzed Ginga scanning data and determined a photon index of $`\mathrm{\Gamma }=1.9\pm 0.02`$ and a 10–20 keV flux of $`F_{1020\mathrm{keV}}=1.6\times 10^{11}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$, corresponding to a 2–10 keV flux of $`F_{210\mathrm{keV}}=3.4\times 10^{11}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$ (assuming a simple extrapolation of the 10–20 keV powerlaw to lower energies). Hanson et al. (1990) report upper limits that are 5 times weaker. Takano & Koyama (1991) take this as evidence for variability.
With the exception of the EXOSAT ME which suffered from severe background subtraction issues at high energies, the Rossi X-ray Timing Explorer (RXTE) is the first hard X-ray observatory with a sufficiently small field of view to avoid contamination by NGC 4388. In this letter we report four RXTE observations of M87. The AGN/jet is not detected above the thermal emission of the Virgo cluster, and upper limits on its flux are given. We include the effect of unknown fluctuations in the Cosmic X-ray Background (CXB) when deriving our limits on the non-thermal emission and these, indeed, turn out to be the limiting factor for these data. Astrophysical implications of this non-detection are discussed in Section 4.
## 2 Observations and basic data reduction
We observed M87 with RXTE four times. The dates (and good on-source exposure times) were 30-Dec-1997 to 2-Jan-1998 (42600 s), 9-Jan-1998 to 12-Jan-1998 (43500 s), 19-Jan-1998 to 22-Jan-1998 (35500 s), and 30-Jan-1998 to 3-Feb-1998 (44900 s). The motivation of the project was to search for the hard X-rays from M87 and its jet, and study their spectrum and variability.
RXTE is a hard X-ray observatory possessing two pointed instruments, the Proportional Counter Array (PCA) and the High Energy X-ray Timing Experiment (HEXTE), as well as an all-sky monitor (ASM). In this work, we use data from the PCA and HEXTE.
The PCA consists of five nearly identical co-aligned Xenon proportional counter units (PCUs) with a total effective area of about 6500 cm<sup>2</sup> and is sensitive in the energy range from 2 keV to $`60`$ keV (Jahoda et al. 1996). Data taken in Standard-2 mode were extracted into spectra and lightcurves using Ftools v4.1 supplemented with the RXTE patch (A. Smale, private communication). For spectral fitting, response matrices were generated using the ftools routine pcarmf v3.3 (corrected for the 1998-Aug-29 bug in which pcarmf failed to account for temporal variations of the response matrix). To take into account the remaining uncertainties of the matrix we added 1% systematic errors to our data. This value was determined from the deviations from a pure power-law in a fit to the Crab Nebula and pulsar spectrum. Background subtraction was performed using the latest background models (released in June 1998 as part of the RXTE patch to ftools v4.1). To increase signal-to-noise, only data from the top layer of the PCUs are considered here. We limit our consideration to the energy range 4–15 keV — the lower bound is determined by the lower limit of the well calibrated energies, whereas the upper bound is the energy at which the data become background dominated.
The HEXTE consists of two clusters of four NaI/CsI-phoswich scintillation counters, sensitive from 15 to 250 keV (Rothschild et al. 1998). Background subtraction is done by source-background rocking of the two clusters. No signal from M87 was detected in HEXTE for either the individual observations or the co-added data of all of our datasets. However, the limits set by these non-detections are uninteresting compared with the PCA limits that we shall address below, and hence we shall not discuss HEXTE data any further.
The total background subtracted PCA count rate was $`113\mathrm{counts}\mathrm{s}^1`$ for all 5 PCUs, with no evidence for variability either between the four observations or within individual observations. In the absence of any temporal structure, we focus on the spectral aspects. We have extracted 4 PCA spectra, one for each individual observation, and rebinned to at least 20 photons per energy bin. This is a requirement for the $`\chi ^2`$ analysis of the following section to be appropriate. Spectral fitting was then performed using the xspec v10.0 package.
## 3 Spectral results
In this section, the 4–15,keV spectra resulting from the four observations are analyzed individually in order to assess spectral variability. Since we fail to detect spectral variability, we also fit the combined spectrum jointly. We expect each spectrum to be dominated by thermal plasma emission from the ICM of the Virgo cluster. Non-thermal emission from the AGN or jet, if present, would be revealed as a hard tail above and beyond the thermal emission.
For these data, unknown fluctuations in the CXB are the limiting factor in our ability to study the non-thermal emission from M87. The spectral fitting presented in section 3.1 (and in Table 1) incorporates this uncertainty by including such fluctuations as an extra model component. In detail, we include a power-law component with photon index $`\mathrm{\Gamma }=1.8`$ which is allowed to vary in normalization (measured at 1 keV) between $`N_{\mathrm{CXB}}=\pm 2\times 10^4\mathrm{ph}\mathrm{keV}^1\mathrm{cm}^2\mathrm{s}^1`$ (i.e., the 1-$`\sigma `$ fluctuations of the CXB estimated by scaling the Ginga results of Butcher et al. to account for the instrumental responses, effective areas and fields of view).
### 3.1 The thermal plasma and limits on a $`\mathrm{\Gamma }=2`$ power-law
Initially, these spectra were fitted with a single temperature thermal plasma model (Mewe, Gronenschild, & van den Oord 1985; Kaastra 1992) modified by Galactic absorption ($`N_\mathrm{H}=2.5\times 10^{20}\mathrm{cm}^2`$) and redshifted appropriately for the Virgo/M87 system ($`z=0.003`$). The resulting fits are acceptable ($`\chi _\nu ^21.1`$ for 25 degrees of freedom) and are shown in Table 1. As can be seen from Table 1, all spectra are adequately described by a $`kT2.5\mathrm{keV}`$ plasma with an abundance of $`Z0.26Z_{}`$. Extrapolating the fits in the 2–4 keV range, the total inferred 2–10 keV flux is $`2.8\times 10^{10}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$, although there may well be a cooler cluster component contributing to the cluster X-ray emission below 4 keV (e.g. Lea et al. 1982; Matsumoto et al. 1996). The corresponding 2–10 keV luminosity is $`8.7\times 10^{42}\mathrm{erg}\mathrm{s}^1`$.
The results of the thermal plasma fit to all four datasets simultaneously (i.e., the combined fit) are shown in Fig 1. A ‘line-like’ feature can be seen at $`6\mathrm{keV}`$. This feature is seen in all PCU detectors individually and in each observing period. An additional narrow Gaussian line with energy $`E=6.22\pm 0.12\mathrm{keV}`$, slightly less than the cold iron fluorescent line energy of iron, and equivalent width $`84\mathrm{eV}`$ describes this feature well and produces a large improvement in the goodness of fit ($`\chi _\nu ^2=56.0/107`$). There are several reasons for suspecting that this line is not real. Firstly, one must be immediately suspicious since this is a weak line-feature embedded in the wings of a much stronger line (i.e., the ionized emission line from the thermal cluster gas). Secondly, such a line is not seen in ASCA data of the Virgo cluster (an examination of archival ASCA data gives an upper limit of $`20\mathrm{eV}`$ on the equivalent width of a line at these energies). Thirdly, there is no astrophysical precedent for observing such a line from a cluster dominated system. While we cannot firmly reject the hypothesis that this line is real, we suspect that it is due to a small mismodelling of the ionized emission lines. Furthermore, the presence of this line in the spectrum does not affect the best fit values or uncertainties of the other spectral parameters. Thus, we shall not discuss this feature any further, and shall not include it in the spectral fitting described below.
In order to assess the presence of hard non-thermal emission from the AGN or jet, a power-law component was added to the thermal plasma model. In no case did the addition of the power-law component lead to any significant improvement in the goodness of fit. Table 1 quotes the 90 per cent limits on the normalization of the power law (at $`1\mathrm{keV}`$) assuming a photon index of $`\mathrm{\Gamma }=2`$ (close to the canonical value for Type-1 AGN).
### 3.2 Detailed limits on non-thermal emission
Given the lack of any spectral variability, we choose to refine our limits on the non-thermal emission using the combined spectrum (i.e., summing the spectra from our individual observations). Here, we compute confidence contours on the $`(\mathrm{\Gamma },N_{\mathrm{pow}})`$-plane using a more rigorous treatment of the effect of CXB fluctuations on our spectral fitting.
Since we know the probability distribution of the CXB fluctuations (Butcher et al. 1997), we can integrate over possible fluctuations and obtain ‘averaged’ confidence contours. We define a likelihood function given a particular CXB fluctuation,
$$(\mathrm{\Gamma },N_{\mathrm{pow}}|N_{\mathrm{CXB}})\mathrm{exp}\left(\frac{\chi ^2(\mathrm{\Gamma },N_{\mathrm{pow}}|N_{\mathrm{CXB}})}{2}\right),$$
(1)
where $`\chi ^2(\mathrm{\Gamma },N_{\mathrm{pow}}|N_{\mathrm{CXB}})`$ is derived from fitting the spectral model of the previous section to the combined data with $`\mathrm{\Gamma },N_{\mathrm{pow}}`$ and $`N_{\mathrm{CXB}}`$ fixed at given values. Defining the probability of a given CXB fluctuation as $`p(N_{\mathrm{CXB}})`$, we can integrate the likelihood function over this parameter,
$$(\mathrm{\Gamma },N_{\mathrm{pow}})=_{\mathrm{}}^+\mathrm{}(\mathrm{\Gamma },N_{\mathrm{pow}}|N_{\mathrm{CXB}})p(N_{\mathrm{CXB}})dN_{\mathrm{CXB}}.$$
(2)
Operationally, we compute the $`\chi ^2`$ surfaces over the $`(\mathrm{\Gamma },N_{\mathrm{pow}})`$-plane with given values of $`N_{\mathrm{CXB}}`$ using the xspec package (as would be done if we were to compute regular confidence contours). Given these $`\chi ^2`$ surfaces, we approximate eqn (2) by a simple sum weighted according to the Gaussian probability function $`p(N_{\mathrm{CXB}})`$.
From $``$, we can produce confidence contours on the $`(\mathrm{\Gamma },N_{\mathrm{pow}})`$-plane in which the CXB fluctuations have been averaged over. Figure 2 shows the resulting 68%, 90%, and 95% contours.
The resulting upper limit on the hard X-ray power-law from M87 assuming $`\mathrm{\Gamma }=2`$ is
$$F_{\mathrm{pow}}(210\mathrm{keV})<4.1\times 10^{12}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1,$$
(3)
corresponding to an isotropic luminosity of
$$L_{\mathrm{pow}}(210\mathrm{keV})<1.0\times 10^{41}\mathrm{erg}\mathrm{s}^1.$$
(4)
using a distance of 16 Mpc. These limits are quoted at the 90 per cent confidence level.
## 4 Discussion
Our upper limit on the hard X-ray power-law component of M87 are the most stringent to date. They are superior to the upper limits derived from Spacelab-2 data \[$`F_{\mathrm{pow}}(210\mathrm{keV})<8\times 10^{12}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$; Hanson et al. 1990\] and ASCA data \[$`F_{\mathrm{pow}}(210\mathrm{keV})<8\times 10^{12}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$; Reynolds et al. 1996\]. Note that Matsumoto et al. (1996) claim a detection of the M87 power-law with $`F_{\mathrm{pow}}(210\mathrm{keV})8\times 10^{12}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$ using the same ASCA data as Reynolds et al. (1996). However, the complex structure of the thermal ICM, which possesses temperature and abundance gradients as well as a possible multiphase structure, makes the thermal ICM spectrum difficult to model and so renders such conclusions about superposed non-thermal emission open to suspicion. Our limits are also inconsistent with the value of $`F_{\mathrm{pow}}(210\mathrm{keV})3.4\times 10^{11}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$ derived from Ginga scanning data (Tanako & Koyama 1991). While this may be evidence for variability, it is also possible that scattered flux from NGC 4388 coupled with the complex thermal ICM influences the subtle analysis of Tanako & Koyama (1991). In summary, we would argue that there has never been an irrefutable detection of a hard X-ray power-law from M87.
However, both the Einstein and ROSAT HRIs have imaged variable soft X-ray point sources coincident with the core of M87 and knot-A (Schreier, Gorenstein & Feigelson 1982; BSH91; Harris, Biretta & Junor 1998; hereafter HBJ98). The last ROSAT measurement shown by HBJ98 was taken on 5-Jan-1998 and hence lies within these RXTE observations. Assuming a single (absorbed) power-law form extending from the ROSAT band into the RXTE band, the 1998 ROSAT fluxes can be converted into loci on the $`\mathrm{\Gamma }`$-$`N_{\mathrm{pow}}`$ plane, as shown in Fig. 2. It can be seen that the ROSAT sources must possess a fairly steep high-energy spectrum ($`\mathrm{\Gamma }>1.90`$ and $`\mathrm{\Gamma }>1.75`$ for the core and jet respectively) in order to avoid detection by RXTE. If the core and jet components have the same hard X-ray spectrum, then the spectral slope must exceed $`\mathrm{\Gamma }>2.05`$ in order for the combined hard X-rays from these two sources to remain undetected. If there is significant intrinsic absorption in this system, even steeper intrinsic spectra are required.
It is informative to compare these limits with the photon indices found in various classes of AGN. In a recent ASCA study, Reeves et al. (1997) found that radio-loud quasars possess ASCA band photon indices of $`\mathrm{\Gamma }=1.63\pm 0.04`$ (see also studies by Lawson et al. 1992, and Williams et al. 1992). Our data rule out the possibility that the ROSAT source seen in the core of M87 possesses such a flat X-ray spectrum when extrapolated to the RXTE band. However, Tsvetanov et al. (1998) have recently suggested that M87 is a misaligned example of a BL-Lac object. These AGN typically possess steep ASCA-band X-ray spectra with $`\mathrm{\Gamma }1.8`$ for the low-energy peaked BL-Lacs (LBL) and $`\mathrm{\Gamma }23`$ for the high-energy peaked BL-Lacs (HBL; e.g. Kubo et al. 1998). The lack of an observed break in the optical-UV spectrum of the jet (Tsvetanov et al. 1998), and the steep X-ray spectral index, would suggest that it belong to the HBL catagory. Thus, our data are entirely consistent with the suggestion that M87 is a misaligned HBL.
## 5 Conclusions
We have presented four RXTE observations of M87 spread over the month of Jan-1998 and totaling approximately 167 ksec of on-source exposure time. Thermal plasma emission from the Virgo cluster ICM is clearly detected in the PCA with a temperature of $`kT2.5keV`$. The metal abundance of the plasma is inferred to be $`Z0.26\mathrm{Z}_{}`$, although this must be treated as an average abundance given the known abundance gradients in this system (Matsumoto et al. 1996). There was no detection in the HEXTE.
Once the thermal ICM emission has been modeled, there is no detection of any hard X-ray non-thermal (power-law) emission in the PCA spectrum. Our upper limit on the flux of a $`\mathrm{\Gamma }=2`$ power-law component is $`F_{\mathrm{pow}}<4.1\times 10^{12}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$. Fluctuations in the CXB are the limiting factor in our ability to set upper limits on the non-thermal emission. If the core and jet sources detected by ROSAT possess a power-law spectrum into the RXTE band, the photon indices of this sources must be $`\mathrm{\Gamma }_{\mathrm{core}}>1.90`$ and $`\mathrm{\Gamma }_{\mathrm{jet}}>1.75`$ respectively. This is entirely consistent with the hypothesis that M87 is a misaligned HBL.
## Acknowledgments
We thank Julia Lee, Michael Nowak, Beverly Smith, and Jörn Wilms for useful discussions. We also thank Andrew Hamilton for sharing his statistical expertise, and Craig Sarazin for pointing out an error in a previous draft of this manuscript. CSR thanks support from NASA under LTSA grant NAG5-6337. CSR also thanks support from Hubble Fellowship grant HF-01113.01-98A awarded bythe Space Telescope Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA under contract NAS 5-26555. ACF thanks the Royal Society for support.
|
no-problem/9812/cond-mat9812052.html
|
ar5iv
|
text
|
# Residual resistivity ratio and its relation to the magnetoresistance behavior in LaMn2Ge2-derived alloys
## Abstract
Results of low temperature magnetoresistance ($`\mathrm{\Delta }\rho /\rho `$) and isothermal magnetization (M) measurements on polycrystalline $`\mathrm{𝑓𝑒𝑟𝑟𝑜𝑚𝑎𝑔𝑛𝑒𝑡𝑖𝑐}`$ (T<sub>C</sub> close to 300 K) natural multilayers, LaMn<sub>2+x</sub>Ge<sub>2-y</sub>Si<sub>y</sub>, are reported. It is found that the samples with large residual resistivity ratio, $`\rho `$(300K)/$`\rho `$(4.2K), exhibit large positive magnetoresistance at high magnetic fields. The Kohler’s rule is not obeyed in these alloys. In addition, at 4.5 K, there is a tendency towards linear variation of $`\mathrm{\Delta }\rho /\rho `$ with magnetic field with increasing $`\rho `$(300K)/$`\rho `$(4.2K); however, the field dependence of $`\mathrm{\Delta }\rho /\rho `$ does not track that of M, thereby suggesting that the magnetoresistance originates from non-magnetic layers. It is interesting that these experimental findings on bulk polycrystals are qualitatively similar to what is seen in artificially grown multilayer systems recently.
The investigation of magnetic multilayers has been a field of active interest in view of the giant magnetoresistance behaviour (GMR) exhibited by these systems. The physics behind electron transport in these systems is not yet fully understood. Though the magnetic multilayers typically show large negative magnetoresistance, $`\mathrm{\Delta }\rho /\rho `$, (defined as \[$`\rho `$(H)-$`\rho `$(0)\]/$`\rho `$(0) where H is the applied magnetic field), there have been recent reports of positive $`\mathrm{\Delta }\rho /\rho `$ of equally large magnitude in some systems, e.g. Dy/Sc superlattices with Dy being magnetic, and non magnetic Cr/Ag/Cr trilayers. In these systems, the application of an external magnetic field modifies the electron scattering/reflection across the interfaces; hence interfacial contribution to resistivity plays an important role in determining the $`\mathrm{\Delta }\rho /\rho `$ behavior as evidenced by higher $`\mathrm{\Delta }\rho /\rho `$ values if the mean free path ($`\mathrm{}`$) is large. There is a breakdown of Kohler’s rule and there is an inverse correlation between the residual resistivity ratio (RRR= $`\rho `$(300 K)/$`\rho `$(4.2 K)) and large positive magnetoresistance (LPMR) in Ref. 5.
We have earlier reported LPMR behavior, qualitatively similar to the Dy/Sc and Cr/Ag/Cr systems, on natural multilayer systems, viz., polycrystalline materials with a layered structure. These systems are: RMn<sub>2</sub>Ge<sub>2</sub> (R= Sm, La) crystallizing in the ThCr<sub>2</sub>Si<sub>2</sub>-type tetragonal structure, containing atoms in the layered sequence Th-Si-Cr-Si-Th-Si-Cr-Si-Th along the c-axis. The finding of note is that the sign of $`\mathrm{\Delta }\rho /\rho `$ is positive at low temperatures inspite of the fact that the compounds are ferromagnetic with the Curie temperature around 300 K. Here, we report the results of our studies on several annealed and as-cast alloys derived from the compound, LaMn<sub>2</sub>Ge<sub>2</sub>, to look for correlations between RRR and the magnitude as well as the field dependence of $`\mathrm{\Delta }\rho /\rho `$, seen in those multilayers. We induce a change in RRR by small variations in Mn-Ge ratio, small chemical substitutions at Mn/Ge sites and in the sample-preparative conditions.
The polycrystalline samples, LaMn<sub>2+x</sub>Ge<sub>2-y</sub>Si<sub>y</sub>, mentioned in table 1 (except the sample 1, which was prepared by arc melting), were prepared by melting together constituent elements in an induction furnace. Slight excess of Mn was added to compensate for weight loss while melting. In the present studies, we have adopted an induction melting method for the preparation of samples since this method is found to allow better control over Mn loss compared to the arc melting method. For each sample, a part of the as-cast ingot was homogenised in an evacuated sealed quartz tube at 800 C for 1 week. X-ray diffraction patterns (Cu K<sub>α</sub>) confirm single phase nature of the alloys. The basal plane parameter is close to 4.2Å and the c-parameter is close to 11Å in all the alloys. The electrical resistivity ($`\rho `$) measurements were performed by a conventional four-probe method employing silver paint for making electrical contacts of the leads with the samples, with the direction of the current being the same as that of H (longitudinal mode). The isothermal magnetization (M) was measured at 4.5 K using a commercial (Quantum design Inc.) superconducting quantum interference device.
In Table 1, we summarize the values of RRR and $`\mathrm{\Delta }\rho /\rho `$ to highlight the disorder effect on the $`\mathrm{\Delta }\rho /\rho `$ of the LaMn<sub>2</sub>Ge<sub>2</sub>-derived alloys. It is interesting to note that the LaMn<sub>2</sub>Ge<sub>2</sub> sample prepared by induction method (sample 6) is characterized by a smaller magnetoresistance compared to the previous sample prepared by arc melting (sample 1). Thus different preparative conditions introduce different degree of defects and disorder in the samples. It is clear that, as the RRR decreases, the LPMR effect as seen at 4.5 K is gradually washed out, viz., for RRR $`<`$ 10, the low temperature $`\mathrm{\Delta }\rho /\rho `$, though positive, is of small magnitude, showing that disorder, as inferred from the magnitude of RRR, kills the LPMR behavior in these systems. It appears that the ferromagnetism of Mn layer in LaMn<sub>2</sub>Ge<sub>2</sub> plays no direct role on the magnetoresistance, as otherwise one would have seen negative magnetoresistance as is usually the case with polycrystalline ferromagnetic materials; this is further corroborated by a comparison of $`\mathrm{\Delta }\rho /\rho `$ with the isothermal magnetization behavior at 4.5 K: M saturates for small values of H for all samples (typical behavior is shown for a few samples in Fig. 2c). Thus, as in the case of Dy/Sc multilayer, M and $`\mathrm{\Delta }\rho /\rho `$ do not track each other. At this juncture, it may be stated that the electrical conduction may be favored along the basal plane as indicated by single crystal investigations on SmMn<sub>2</sub>Ge<sub>2</sub>. However, the non-magnetic layers, but not the Mn layer, are the ones which contribute to magnetic response from the above arguments. The fact that we have observed LPMR even in nonmagnetic alloys belonging to the same structure if RRR is large, supports the idea that the magnetic Mn layer is not responsible for the observed behavior. However, the choice of a ferromagnetic layered compound enables us to draw definitive conclusions on the selective role of nonmagnetic layers on the magnetoresistance behavior; though we have seen large values in layered antiferromagnets in the past, such a conclusion was not possible from those data, as antiferromagnetism can also give rise to positive values. We further speculate that the non-magnetic layer acting as an interface is the one that contains Ge, since low temperature positive magnetoresistance could be observed even in the ferromagnetic, isomorphous Sm sample in which Sm layer is ferromagnetic at 4.2K.
In Fig. 1b and 1d we show the temperature dependence of $`\mathrm{\Delta }\rho /\rho `$, derived from the data shown in Fig 1a and 1c respectively. The values of $`\mathrm{}`$ have also been evaluated from the $`\rho `$ data from the knowledge of the unit-cell density (6.8 gm/cc) and the number of conduction electrons (estimated to be 19, assuming tri, tetra, and tetra positive La, Mn and Si/Ge respectively) within the Sommerfeld theory of conduction. The assumption of theoretical mass density and uncertainties in the absolute determination of $`\rho `$, for instance, due to microcracks in the sample, and in the number of conduction electrons introduce errors in the derived values of $`\mathrm{}`$. Therefore, not much significance may be attached to these values of $`\mathrm{}`$. For qualitative arguments, $`\mathrm{}`$ thus obtained for the two most magnetoresistive samples are shown in the figures 1b and 1d (continuous lines through the data points). The point to be noted is that the $`\mathrm{}`$ closely follows $`\mathrm{\Delta }\rho /\rho `$ as a function of temperature. It may be remarked that the value of $`\mathrm{}`$ at which significant positive $`\mathrm{\Delta }\rho /\rho `$ starts appearing is above 3Å. We derived the same cut-off limit from the low temperature $`\rho `$ data of all the samples (see table 1) as well. Therefore, though the absolute values of $`\mathrm{}`$ may not be reliable, on this basis, one can draw a conclusion that there is a close relationship between the magnitude of the magnetoresistance and $`\mathrm{}`$. However, for the reason mentioned above, we refrain from drawing more conclusions from the derived $`\mathrm{}`$ values.
In Fig. 2a and 2b, we present the results of $`\mathrm{\Delta }\rho /\rho `$ as a function of H at 4.5 K for some of the samples which exhibit significant magnetoresistance. It is to be remarked that the plot of $`\mathrm{\Delta }\rho /\rho `$ versus H/$`\rho `$ is not a single line (shown for one sample in Fig. 3) and thus the Kohler’s rule is found to break down - this is similar to the observations in Cr/Ag/Cr trilayers. For the most magnetoresistive alloy (sample 1), the field dependence is linear in the entire field range of measurement, instead of a classical quadratic dependence. For samples 2 and 3, we observe a deviation from this linearity at low fields (shown in the inset of Fig. 2a for sample 2) beyond which there is a linear behaviour. The low field quadratic dependence could be observable for samples 3 and 4 with still smaller RRR values. Thus there is a tendency towards classical quadratic dependence with a decrease of RRR. (In one of the samples, LaMn<sub>1.9</sub>Ge<sub>2.1</sub> (I) in Fig. 2b, we see a convex curvature around a higher field of 20 kOe, the reason for which is unclear). Thus, the low field linear dependence of magnetoresistance appears to be a signature of relatively better crystallographic order in these alloys. In the same way, in the Cr/Ag/Cr multilayers, an increase in RRR brings about a transition from quadratic to linear behavior in the $`\mathrm{\Delta }\rho /\rho `$ vs H plots, in addition to increasing the magnitude of observed $`\mathrm{\Delta }\rho /\rho `$. At this juncture, we would like to recall a recent report on nonmagnetic bulk silver chalcogenides of the type Ag<sub>2+δ</sub>(Se,Te): For stoichiometric ($`\delta `$= 0) samples, there is no significant magnetoresistance; however, possible introduction of electronic defects for $`\delta `$ $`>`$ 0 gives rise to LPMR behavior. In these systems also, the magnetoresistance is found to exhibit an almost linear dependence on the applied field. Our samples being magnetic, the following point may have to be considered in addition to various other explanations which can be offerred to the above non-magnetic systems; that is, in large RRR alloys, the internal magnetic fields due to ferromagnetism of Mn is large enough that one is well above the quadratic region already in zero external field; however, the deviation from the linearity in smaller RRR samples may possibly be due to randomness of this local field caused by disorder.
In conclusion, we have brought out the effect of residual resistivity ratio on the magnitude of the magnetoresistance and its dependence on H at low fields in LaMn<sub>2</sub>Ge<sub>2</sub>-derived alloys, resembling those of some multilayers. The expected negative sign of $`\mathrm{\Delta }\rho /\rho `$ is not seen in any of these ferromagnetic alloys and this observation, alongwith the isothermal magnetization behavior which is different from that of $`\mathrm{\Delta }\rho /\rho `$, is taken as the evidence for selective magnetic field response from the non-magnetic layers. The present results appear to render an experimental support to a proposal that these layered compounds may serve as model systems for artificial multilayers. It is however not clear whether there are unusual domain wall effects in such ferromagnetic alloys, an aspect which is yet to be settled even in simple ferromagnetic systems. In this sense, it is worthwhile to study the $`\mathrm{\Delta }\rho /\rho `$ behaviour on clean single crystal samples of RMn<sub>2</sub>Ge<sub>2</sub>(Si<sub>2</sub>); this will also enable to comfortably probe the magnetoresistance in a geometry with the current perpendicular to the plane, an important investigation difficult to make on multilayers. In addition, the studies on artificially grown layers of the constituent elements in the sequence present in the crystal structure may be rewarding. We hope that this work will motivate further investigations of such natural multilayer systems.
|
no-problem/9812/astro-ph9812110.html
|
ar5iv
|
text
|
# A Polarimetric Search for Hidden Quasars in Three Radio Selected Ultraluminous Infrared Galaxies
## 1 Introduction
Many sources in the $`IRASFaintSourceCatalog`$ ($`FSC`$) exhibit enormous infrared luminosities ($`L_{IR}>10^{12}`$ $`L_{\mathrm{}}`$), and are known as ultraluminous infrared galaxies (ULIRGs, Sanders et al. (1988)). These objects are an important constituent of our local Universe, with luminosities and space densities similar to those of quasi-stellar objects (Soifer et al. (1987)). This led to the suggestion that ULIRGs could contain infant quasars enshrouded in a large amount of dust (Sanders et al. (1988)). On the other hand, they may also represent energetic, compact starbursts (Condon et al. (1991)). Understanding the dominant energy input mechanism in these ULIRGs – whether it is obscured quasars or intense bursts of star formation – has remained the main issue concerning their nature. Of particular interest are the “warm” ULIRGs (i.e., those having $`f_{25}/f_{60}>0.2`$, Low et al. (1988); Sanders et al. (1988))<sup>1</sup><sup>1</sup>1$`f_{25}`$ and $`f_{60}`$ are the $`IRAS`$ flux densities in units of Jy at 12 $`\mu `$m and 60 $`\mu `$m, respectively.. They include a few of the exceptionally luminous sources, sometimes referred to as the “hyperluminous” ($`L_{IR}\stackrel{>}{}10^{13}`$ $`L_{\mathrm{}}`$) infrared galaxies, such as P09104+4109 ($`z`$ = 0.44; Kleinmann et al. (1988)), F15307+3252 ($`z`$ = 0.926; Cutri et al. (1994)), and F10214+4724 ($`z`$ = 2.286; Rowan-Robinson et al. (1991)). All of these galaxies have been shown to harbor a quasar nucleus obscured from direct view through the detection of hidden broad line regions (HBLRs) in polarized, scattered light (P09104+4109: Hines & Wills (1993); Tran 1998; F15307+3252: Hines et al. (1995); F10214+4724: Goodrich et al. (1996)). This has led to the suggestion that perhaps all warm ULIRGs contain buried QSOs and that they may be the misdirected type 2 QSOs (AGN with Seyfert 2-like emission-line characteristics but QSO-like luminosities; Wills & Hines (1998)).
With the exception of the above extreme $`IRAS`$ sources, most of the known ULIRGs are relatively nearby ($`z`$ $`\stackrel{<}{}`$0.1). In order to identify more ULIRGs at higher redshifts and study their population, several searches have been undertaken, using the positional correlation of sources in the $`FSC`$ with those detected in deep radio surveys. Stanford et al. (1998) correlated objects in the $`FSC`$ with those in the Faint Images of the Radio Sky at Twenty-cm (FIRST) catalog (Becker, White, & Helfand 1995), which has a flux density limit of $``$ 1 mJy. The results show that nearly all matches that are faint on the Palomar Observatory Sky Survey (POSS) turn out to be ULIRGs, if they lie on the well-known radio-to-far infrared (FIR) flux correlation (Condon et al. (1991)) established by normal late-type galaxies and starburst galaxies at low redshifts. Spectroscopically, the vast majority of the sources ($``$ 85%) in this “radio-quiet” FIR sample (hereafter FF sources) can be classified as starburst systems (Stanford et al. (1998)).
Dey & van Breugel (1994) correlated the Texas 365 MHz radio catalog (Douglas et al. (1996)) with an early version of the $`FSC`$, and selected from this sample those sources that lie above the radio-to-FIR flux-flux correlation (see Sopp & Alexander (1991); Colina & Pérez-Olea (1995); Stanford et al. (1998)). From this sample (hereafter TF sources) they discovered a significant class of gas and dust-rich AGN which are characterized by large FIR luminosities, and which are intermediate in radio luminosity, FIR color, and optical spectral class between the $`IRAS`$ ultraluminous galaxies and powerful quasars (Dey & van Breugel (1994); Stanford et al. (1998)). Because of the much higher detection threshold of the Texas survey ($``$ 150 mJy, Douglas et al. (1996)), the majority of these sources are “radio-loud” LINERs, Seyfert 2s and post–starburst active galactic nuclei (PSAGN). The latter are characterized by the strong Balmer absorption lines, Seyfert 2-like emission lines and compact, often almost stellar, optical morphologies.
A notable feature of both the radio-quiet and radio-loud FIR starburst and PSAGN objects, visible in high signal-to-noise spectra, is the prevalence of many strong high-$`n`$ Balmer absorption lines (EW $``$ 10 Å), as well as a strong Balmer discontinuity in their spectra. These features signify a large population of hot, relatively young stars (10<sup>8</sup> – 10<sup>9</sup> yrs). In addition, their emission-line spectra tend to exhibit a very low ionization state, as characterized by the strength of \[O II\] $`\lambda `$3727 relative to \[O III\] $`\lambda `$5007 (\[O II\] $`\lambda `$3727/\[O III\] $`\lambda `$5007 $`\stackrel{>}{}`$2), and the lack or weakness of high-ionization lines such as \[Ne V\] $`\lambda `$3426. Qualitatively, the spectra of these galaxies appear very similar to those of the class of galaxies called “E + A” or “post-starburst” galaxies (Dressler & Gunn (1983); Oegerle, Hill, & Hoessel 1991; Wirth, Koo, & Kron 1994; Liu & Green (1996)), whose spectra display not only the 4000 Å break, G-band, and MgI b features that are indicative of an old ($``$ 10 Gyr) stellar population (the “E” component), but also the Balmer absorption lines characteristic of much younger stars ($`<`$ 1 Gyr, the “A” component). Many FF and TF sources appear to differ from E + A galaxies only in the additional presence of emission lines, indicating the presence of on-going star formation or active galactic nuclei (see spectra of TF J1020+6436 and FF J1614+3234 below). These objects may form an important evolutionary link between starbursts and quasar activity (Canalizo & Stockton (1997); Stockton, Canalizo, & Close 1998).
In order to better understand the energy sources of ULIRGs and the relationship between AGN and starburst activity, we started a spectropolarimetric survey to search for hidden broad emission lines in both types of objects: the radio-quiet (FF) and radio-loud (TF) ULIRGs. This would indicate whether obscured AGNs might exist, and the high signal-to-noise data might be used to investigate the nature of their stellar populations. This might be used to help quantify the relative importance of quasar and starburst activity, and to search for possible correlations with the age of the starbursts, luminosity, and morphological evolution.
This paper reports the first results for three objects with $`L_{IR}\stackrel{>}{}10^{12}`$ $`L_{\mathrm{}}`$<sup>2</sup><sup>2</sup>2$`L_{IR}`$ is calculated by following the prescription of Sanders & Mirabel (1996).: TF J1020+6436, TF J1736+1122, and FF J1614+3234. Only TF J1736+1122 is certainly IR “warm” with $`f_{25}/f_{60}`$ = 0.404. For TF J1020+6436 and FF J1614+3234, only an upper limit of the $`f_{25}`$ flux density was obtained. Table 1 lists the infrared properties for the three objects. Throughout this paper, we assume $`H_o`$ = 75 km s<sup>-1</sup> Mpc<sup>-1</sup>, $`q_o`$ = 0 and $`\mathrm{\Lambda }`$ = 0.
## 2 Observations and Reductions
### 2.1 Spectropolarimetry
Spectropolarimetric observations were made with the polarimeter (Cohen et al. (1997)) installed in the Low Resolution Imaging Spectrometer (LRIS, Oke et al. (1995)) on the 10-m Keck II telescope on the night of 1997 April 10 (UT). A 1″ wide, long slit oriented at the parralactic angle was centered on the nucleus of each galaxy. We used a 300 grooves mm<sup>-1</sup> grating which provided a dispersion of 2.46 Å pixel<sup>-1</sup> and a resolution of $``$ 10 Å (FWHM). The observations were made by following standard procedures of rotating the half waveplate to four position angles (0°, 22.5°, 45°, and 67.5°), and dividing the exposure times equally among them. The total exposure times were one hour each for TF J1020+6436 and FF J1614+3234, and 40 minutes for TF J1736+1122. We followed standard polarimetric reduction techniques as described, for example, by Cohen et al. (1997).
### 2.2 Near-Infrared Imaging
Near-IR observations of FF 1614+3234 were made using the Near Infrared Camera (NIRC, Matthews & Soifer (1994)) at the Keck I Telescope. NIRC employs a 256$`\times `$256 InSb array with a scale of 0.15 arcsec pixel<sup>-1</sup>. The data were obtained through a $`K_s`$ filter on 1998 April 18 (UT), under photometric conditions with $``$0$`\stackrel{}{\mathrm{.}}`$5 seeing. The target exposures were taken using a nonredundant dithering pattern, with individual shifts between pointings of 3″. Integration times for each pointing were typically 60 s comprised of 3 coadds; the total on-source exposure time was 1920 s. After bias subtraction, linearization, and flatfielding (using sky flats), the data were sky–subtracted, registered and summed using the DIMSUM<sup>3</sup><sup>3</sup>3DIMSUM is the Deep Infrared Mosaicing Software package, developed by P. Eisenhardt, M. Dickinson, A. Stanford, and J. Ward, and is available as a contributed package in IRAF. near–IR data reduction package. Subsequent photometry was flux calibrated via short observations of UKIRT faint standards (Casali & Hawarden (1992)), which, after a suitable transformation, yields magnitudes on the CIT system.
Figure 1 shows the $`K_s`$-band image of FF J1614+3234 obtained with Keck I Telescope. As can be seen, the near-IR image shows a disturbed morphology, with an extension separated $``$ 1.8 from the nucleus to the south-west. At a lower surface brightness level, extended filamentary arc-like structure is present to the north and west. This feature is suggestive of a tidal tail due to a merger event or interaction. There is also an apparent companion $``$ 2.2 to the north-west. These features are consistent with FF J1614+3234 being similar to an E + A galaxy, which often occurs in merging systems (e.g., Liu & Kennicutt (1995)). They also provide strong circumstantial evidence that any starburst and AGN activity present is triggered by galaxy interaction. No indication of gravitational lensing is evident from the image.
### 2.3 Radio Observations
TF J1020+6436 and TF J1736+1122 were both observed with the VLA, using standard bandwidths and calibrations, to obtain more accurate positions, and morphological and spectral index information. TF J1020+6436 was observed at 4860 MHz in the C-array and was found to be unresolved with an angular size $`<`$ 1.1. TF J1736+1122 was observed at 4860 MHz and 8440 MHz in the B and A-arrays respectively. The B-array observations showed that this source was slightly resolved, with an angular size of 0.$`\times `$ 0.4, at PA = 89°. In the A-array, and at the higher frequency of 8440 MHz, the source shows a resolved core with a clear double structure (see Fig. 2) elongated along PA $``$ 97°. This structure axis is essentially perpendicular ($`\mathrm{\Delta }`$PA $``$ 80°) to the polarization PA of 18° (see §4.1), as has been found for many polarized narrow-line radio galaxies and Seyfert 2 galaxies (e.g., Tran et al. (1998) and references therein). For FF J1614+3234 the radio data was taken from the FIRST catalog (Becker et al. 1995). Only the 1400 MHz observations are available for this source, which is unresolved with an angular size of $`<`$ 4″. The positions and flux densities of the three sources are listed in Table 2.
It is of interest to note that the radio continua of the two radio-loud TF sources have quite steep spectra, with spectral indices $`\alpha 1`$ ($`f_\nu \nu ^\alpha `$). In the case of TF J1736+1122 the radio spectrum steepens significantly between 365 MHz and 8440 MHz, with $`\alpha `$ changing from $``$0.75 to $``$1.45 (see Table 2). In addition, both objects are quite small ($`<`$ 2.6 kpc). Such compact steep spectrum (CSS) radio sources comprise a significant fraction of complete radio source surveys, and are thought to be “frustrated” or young radio sources whose jets are plowing through the dense ISM of their parent galaxies (see O’Dea 1998 for a recent discussion).
## 3 Results
### 3.1 Line Ratios, Diagnostics, and Classification
Table 3 presents the integrated emission-line flux ratios and their rest-wavelength equivalent widths. The emission-line fluxes were measured from the starlight-subtracted spectra. For FF J1614+3234 and TF J1020+6436, isochrone population synthesis models of the underlying young starbursts (see §3.2) have been removed. For TF J1736+1122, the contribution from an old stellar population represented by the elliptical galaxy NGC 4478 was used (§3.3). In TF J1736+1122, the narrow emission lines display extended wings that cannot be adequately fit with a single Gaussian profile. A sum of two Gaussian profiles, one with FWHM $``$ 500 km s<sup>-1</sup> and another with FWHM $``$ 1300 – 1800 km s<sup>-1</sup> but identical center, is able to reproduce the observed profiles very well. The observed spectrum of TF J1020+6436 appears to suffer significant internal reddening. The measured narrow-line Balmer decrement is H$`\alpha `$/H$`\beta `$ = 8.8, giving an extinction of $`E(BV)=1.05`$, assuming an intrinsic ratio of H$`\alpha `$/H$`\beta `$ = 3.1 and using the extinction law of Cardelli, Clayton, & Mathis (1989).
Plotting the line flux ratios (involving \[O II\] $`\lambda `$3727, H$`\beta `$, \[O III\] $`\lambda `$5007, \[N II\] $`\lambda \lambda `$6548, 6583, H$`\alpha `$, \[S II\] $`\lambda `$6724, and \[O I\] $`\lambda `$6300) on the diagnostics diagrams of Baldwin, Phillips, & Terlevich (1981) and Veilleux & Osterbrock (1987) shows that both TF J1020+6436 and TF J1736+1122 lie in regions typically occupied by narrow-line AGNs. This suggests that the dominant energy input in the narrow line regions (NLRs) is photoionization by a hard continuum. These objects can be classified as Seyfert 2 galaxies. However, only TF J1736+1122 can be called a bona fide high-ionization Seyfert 2. Although TF J1020+6436 formally qualifies as a type 2 Seyfert, its ionization level is quite low (\[O II\] $`\lambda `$3727/\[O III\] $`\lambda `$5007 = 1.4). What clearly separates it from TF J1736+1122 is the domination of early-type stellar spectral features in the blue end of the spectra. TF J1020+6436 satisfies half of the definition of Heckman (1980) for LINERs<sup>4</sup><sup>4</sup>4Criteria for LINERs as defined by Heckman (1980) are: \[O II\] $`\lambda `$3727/\[O III\] $`\lambda `$5007 $`>`$ 1 and \[O I\] $`\lambda `$6300/\[O III\] $`\lambda `$5007 $`>`$ 1/3.: \[O II\]/\[O III\] $`>`$ 1. The other half of the criterion is not satisfied (\[O I\]/\[O III\] = 0.15 $`<`$ 0.33, Table 3), and thus it is not a LINER, and we adopt the Seyfert 2 classification for it. Since our spectrum of FF J1614+3234 does not cover wavelengths longward of \[O III\] $`\lambda `$5007, an unambiguous classification of this galaxy based on the line ratios of \[N II\], H$`\alpha `$, and \[S II\] is not possible. Using the available line strengths, we classify FF J1614+3234 as a LINER based on the low ionization level (\[O II\] $`\lambda `$3727/\[O III\] $`\lambda `$5007 = 1.75; \[O III\] $`\lambda `$5007/H$`\beta `$ = 2.0) and the strong appearance of hot-star spectral features.
### 3.2 No Detection of Hidden Quasars in FF J1614+3234 and TF J1020+6436
FF J1614+3234 and TF J1020+6436, whose total flux spectra and $`Q`$ and $`U`$ Stokes parameters are shown in Figure 3, display no significant detectable polarizations. A formal averaging of the continuum polarization over the observed wavelength range 4500 Å – 8000 Å gives $`P`$ = 0.24% $`\pm `$ 0.18% and 0.15% $`\pm `$ 0.15% (uncorrected for biasing) in FF J1614+3234 and TF J1020+6436, respectively. As a result, we detected no polarized broad lines in the spectra of two of the more luminous ULIRGs in our sample. There is also little sign of Mg II $`\lambda `$2800 in the spectrum of FF J1614+3234. It shows a very weak feature at the wavelength expected for Mg II $`\lambda `$2800, but the profile is highly irregular. Although the non-detection of hidden broad lines, by itself, does not rule out the presence of a quasar, as the geometry may not be propitious for scattered light to reach our line of sight (e.g., blockage or lack of scattering region, high viewing inclination), the low excitation state and the presence of strong Balmer absorption lines seem to indicate that the energy output is dominated by a young starburst. The enormous IR luminosity is likely the result of this starburst activity.
We used the GISSEL models of Bruzual & Charlot (1996, BC96) to characterize the young stellar population dominating the spectra of FF J1614+3234 and TF J1020+6436. We assumed an instantaneous burst of star formation with a Salpeter (1955) initial mass function having a mass range $`0.1M_{\mathrm{}}<M<125M_{\mathrm{}}`$ and solar metallicity. The fit is a $`\chi ^2`$ minimization of the sum of a blue featureless continuum (representing the Seyfert 2/LINER ionizing continuum) and the young starburst model. Our analysis indicates that some additional reddening is also required for a good fit. Using the starburst extinction curve of Calzetti, Kinney, & Storchi-Bergmann (1994), the required reddening is $`E(BV)0.3`$ for FF J1614+3234 and $``$ 0.35 for TF J1020+6436. Figure 4 shows the best fits to the observed spectra, which were achieved for a starburst with an age of 320 Myr. This is consistent with the median age derived for the starburst knots found in all of the ULIRGs studied by Surace et al. (1998) using $`HST`$ imaging. Our results show that the light from the 320 Myr-old starburst component contributes about 68% and 76% of the total light at 4200 Å (rest frame) for FF J1614+3234 and TF J1020+6436, respectively. The rest of the light is due to the blue ionizing continuum of the Seyfert 2/LINER nucleus and perhaps an even younger population of hot (OB) stars.
### 3.3 TF J1736+1122: Hidden Broad Lines and Starburst Activity
Unlike the two ULIRGs discussed above, TF J1736+1122 is highly polarized, having a continuum $`P`$ reaching up to $``$ 23% in the near-UV. The spectropolarimetry is shown in Figure 5. The polarized flux spectrum, $`P\times F_\lambda `$, clearly displays not only broad Balmer emission lines of H$`\alpha `$, H$`\beta `$, H$`\gamma `$, but also the Fe II blends near 4570 Å and 5250 Å. In this respect, it is similar to the $`P\times F_\lambda `$ spectrum of the hidden Seyfert 1 galaxy NGC 1068 (Miller, Goodrich & Mathews 1991; Tran (1995)).
We corrected the polarization for a small amount of interstellar polarization (ISpol), based on the polarization of the narrow \[O II\] emission line and the polarization of several stars near the line of sight to TF J1736+1122. The ISpol we adopt is $`P`$ = 0.4 % at PA = 80°. This is consistent with the maximum polarization expected ($`P_{max}<9.0E(BV)`$; Serkowski, Mathewson, & Ford 1975) from the Galactic reddening toward TF J1736+1122 ($`E(BV)`$ = 0.144; Burstein & Heiles (1984)). Since the correction is small relative to the polarization of TF J1736+1122, it does not significantly affect the results or alter our conclusions.
We also corrected the observed polarization for the unpolarized starlight in the host galaxy (e.g., Tran (1995)). Unlike FF J1614+3234 or TF J1020+6436, the underlying stellar population of TF J1736+1122 is dominated by old ($`>`$ a few Gyr) stars, as indicated by the lack of Balmer absorption lines. Accordingly, a selection of normal elliptical galaxies was tried in the fit, and the best fitting template galaxy is NGC 4478. We derived a galaxy fraction of 0.52 at 5500 Å. Before starlight subtraction the percentage of polarization rises rapidly with shorter wavelength, reaching $``$ 23% at 3250 Å. This rise in $`P`$ is due to the decreasing dilution of starlight at shorter wavelengths, as the starlight-corrected polarization is essentially flat. There is a rise in polarization in the broad wings of H$`\alpha `$ and H$`\beta `$, signifying the presence of a second diluting “featureless continuum” (FC2), perhaps due to hot stars (Cid Fernandes & Terlevich (1995); Heckman et al. (1997)) or thermal radiation from hot electrons (Tran (1995)).
The presence of hot stars, perhaps arising in a starburst, is suggested by the broad feature underlying He II $`\lambda `$4686. This broad feature is shown in Figure 6 as a residual after the narrow line components are fitted and subtracted. This feature cannot be a broad scattered He II component since it appears very strongly in the total flux spectrum, suggesting that the light is viewed directly, and it is not detected in the $`P\times F_\lambda `$ spectrum (Fig. 5). It can be attributed to Wolf-Rayet (W-R) stars, as has been identified by Heckman et al. (1997) in Mrk 477 and Storchi-Bergmann, Cid Fernandes, & Schmitt (1998) in Mrk 1210, both of which are Seyfert 2 galaxies with hidden broad-line regions, and for which a large fraction of FC2 has been inferred (Tran (1995)). It seems, therefore, that there is a large number of young and/or W-R stars in TF J1736+1122, which contributes to the observed FC2. As discussed by Schaerer & Vacca (1998), the presence of this feature indicates that the time elapsed since the last burst of star formation cannot be more than a few Myr. If we assume that the broad feature is all due to the “W-R bump” and that H$`\beta `$ emission arises entirely from massive stellar activity (clearly an overestimate since most of the H$`\beta `$ must be due to the AGN), we can derive an absolute lower limit to the ratio of W-R to O stars. Using the relation derived by Schaerer & Vacca (1998) (their Fig. 22 and equ. 17), the observed W-R bump/H$`\beta `$ ratio of 0.14 (Table 3) gives a W-R/O star number ratio $`>`$ 0.2. By contrast, there is little sign of the W-R bump in either FF J1614+3234 or TF J1020+6436, suggesting that the number of W-R stars relative to O stars must be low. This may indicate that few of the most massive stars ($`\stackrel{>}{}`$ 35 $`M_{\mathrm{}}`$, which turn into W-R stars) are formed in these galaxies, or their star formation episodes, which may still be ongoing, last longer than in TF J1736+1122.
## 4 Discussion
### 4.1 Modelling of the Continuum and Line Polarization in TF J1736+1122
In order to separate the line and continuum components of polarization of TF J1736+1122, we modelled the observed spectropolarimetric observations. Following a method similar to that described by Tran et al. (1997), we model the spectropolarimetry surrounding the H$`\alpha `$ \+ \[N II\] complex by simultaneously fitting the continuum and emission-line components in the total flux, $`Q\times F_\lambda `$, and $`U\times F_\lambda `$ spectra. The broad H$`\alpha `$ in $`P\times F_\lambda `$ is best fitted by two Gaussian profiles with FWHMs of 18000 km s<sup>-1</sup> and 4900 km s<sup>-1</sup>, giving a combined FWHM of 5580 km s<sup>-1</sup>. We use this same 2-component model to fit the H$`\alpha `$ line profile in the total flux spectrum. The narrow lines of H$`\alpha `$ \+ \[N II\], \[O I\], \[S II\] were fitted assuming that they are similarly polarized, and having Gaussian profiles with FWHMs similar to that of the \[O III\] lines ($``$ 840 km s<sup>-1</sup>). The results of our polarization modelling are shown in Figure 7.
The results indicate that the broad-line polarization is very high, $``$ 50%, while the continuum $`P`$ is much lower at $``$ 22 %, both at a similar PA of 20°. This indicates that the amount of unpolarized FC2 flux relative to the total continuum (i.e., FC1 + FC2) would be about 60% in order to make the continuum and broad-line polarizations equal. This is also just the amount required in order to scale up the polarized flux spectrum and subtract off the broad-line component from the total flux spectrum. Our fitting also shows that the narrow lines possess a small amount of polarization $``$ 0.4 % at PA = 162°. They are clearly polarized at a PA different from that of the broad line and continuum, as evidenced by the apparent absorption in the $`P\times F_\lambda `$ spectrum at the wavelengths of strong narrow emission lines such as \[O III\] $`\lambda \lambda `$4959, 5007 (Fig. 5). The results found here are similar to all prior results for Seyfert 2 and narrow line radio galaxies, (e.g., Tran (1995); Tran, Cohen, & Goodrich 1995; Cimatti et al. (1996); Dey et al. (1996); Tran et al. (1998)) in that they all show the presence of an FC2, an old stellar population, high polarization in the continuum and broad lines but little or no polarization in the narrow lines.
Polarization modelling at H$`\beta `$ \+ \[O III\] shows similar results for the broad-line and continuum polarization. However, the fitting here is complicated by the presence of strong underlying optical Fe II emission, and thus the results are not as well constrained. However, the similar polarizations obtained suggest that the polarization is independent of wavelength.
The continuum polarization PA shows a slight rotation of $``$ 4° from 16° to 20° across the wavelength range covered. This may mean that FC2 is slightly polarized or that there is some small interstellar polarization internal to the host galaxy. Any of these effects is expected to be small since there is little or no difference in the polarization PA between the continuum and broad lines at H$`\alpha `$ or H$`\beta `$.
### 4.2 Polarized Flux Spectrum and Fe II emission in TF J1736+1122
The luminosity of the hidden broad-line AGN indicates that it is consistent with being a quasar. Using the formulation of Véron-Cetty & Véron (1996), the observed absolute magnitude of TF J1736+1122 is $`M_B21.2`$. Assuming an optically thin, uniformly filled scattering cone with half-opening angle of 40° oriented at an inclination of 70° (§4.4), producing an intrinsic polarization of 50%, we derive $`f_{direct}/f_{scattered}10`$ (see e.g., Brotherton et al. (1998)). Thus the luminosity of the obscured source is $`M_B23.7`$, marginally qualifying its status as a quasar. With an optical spectral index of $`\alpha 0.24`$ ($`f_\nu \nu ^\alpha `$), the scattered ($`P\times F_\lambda `$) spectrum of TF J1736+1122 also appears very similar to the composite quasar spectrum of Francis et al. (1991), which has $`\alpha =0.32`$. The Balmer decrement in $`P\times F_\lambda `$ is 4.2 $`\pm `$ 1. These results indicate that there is no significant “bluening” or reddening of the scattered spectrum. Thus it appears that the scatterer is gray, as would be the case for electrons. For comparison, the Balmer decrement in the NLR is H$`\alpha `$/H$`\beta `$ = 2.9, also essentially unreddened by dust.
As has been mentioned, an interesting feature in the $`P\times F_\lambda `$ spectrum of TF J1736+1122 is the presence of strong polarized Fe II emission. The rest-frame EWs of the Fe II blends at 4570 Å and 5250 Å are $``$ 35 Å and 50 Å, respectively<sup>5</sup><sup>5</sup>5The Fe II EWs were measured by scaling and broadening the Fe II template I Zw 1 to fit the observed blends, as described by Boroson & Green (1992).. This is comparable to those seen in normal quasars, Seyfert 1, and broad-line radio galaxies (e.g., Joly (1991)). The detection of Fe II in $`P\times F_\lambda `$ is rather surprisingly rare among the class of polarized broad-line AGNs. Of all the Seyfert 2s and narrow-line radio galaxies known to show HBLRs (Tran (1995); Tran et al. (1995); Young et al. (1996); Cohen et al. (1998)) only NGC 1068 and Mrk 463E have been observed previously to show optical Fe II in their $`P\times F_\lambda `$ spectra.
Although optical Fe II emission in AGNs has been known for many years, its exact nature and origin are still not fully understood. The strength of Fe II emission tends to be weaker in broad-line radio galaxies compared to radio-quiet Seyfert 1 galaxies (e.g., Osterbrock (1977)). This general tendency between Fe II strength and radio-loudness also holds for QSO and quasars (Boroson & Green (1992)), but radio-loud objects are by no means always weak Fe II emitters (e.g., Joly (1991)). Among the radio-loud quasars, it has been found (e.g., Heckman (1983); Jackson & Browne (1991); Brotherton (1996)) that Fe II strength is correlated with radio spectral index, in the sense that flatter-spectrum sources display stronger optical Fe II. TF J1736+1122 has a radio spectrum that is rather steep (Table 2). Its radio-to-optical flux ratio and radio luminosity (Table 2) indicate that TF J1736+1122 is formally radio-loud, albeit only marginally. In the orientation-dependent model of quasars, flat-spectrum radio sources are thought to be those viewed at small inclination angle where the core component dominates, while steep-spectrum radio sources are those viewed at larger viewing angle where the lobes dominate (e.g., Orr & Browne (1982); Boroson & Oke (1984)). Thus, TF J1736+1122 could be a radio-loud quasar being viewed at large inclination angle, with heavy obscuration by dust, perhaps in the form of a putative torus, dominating the central nucleus.
Because Fe II emission appears prominently in broad-line (type 1) AGNs, in which the nuclear and BLR radiation escape freely in our direction, it is thought to be associated closely with the ionizing continuum, BLR and radio core (e.g., Zheng & Keel (1991)). However, several observations (Hickson & Hutchings (1987); Canalizo & Stockton (1997)) have shown that Fe II emission in the QSO PG 1700+518 varies spatially, suggesting that it could come from more extended regions, perhaps in a superwind and shocks generated by supernovae and supernova remnants of a starburst (Terlevich et al. (1992)). Since the polarization magnitude and position angle of the continuum, broad emission lines, and Fe II emission are virtually identical in TF J1736+1122, indicating that they must all follow the same scattering geometry, our results support the picture in which Fe II emission originates from a compact source very close to the nucleus, arguing against an extended supernova origin for Fe II.
### 4.3 Detection of HBLRs and Nature of ULIRGs: Starburst versus AGN
The original motivation for the HBLR search of ULIRGs in the present study was initiated by their positions in the radio-FIR correlation diagram. Based solely on their general coincidence with the hyperluminous $`IRAS`$ galaxies in this diagram, and Sanders et al. (1988) ULIRG-quasar evolutionary scheme, we would expect them to harbor obscured quasars. This expectation has not been borne out. The lack of polarization in TF J1020+6436 and FF J1614+3234 raises an important question: if a BLR is not detected in polarized light, does it truly not exist? While we cannot rule out the presence of an obscured quasar in these sources, it appears that if a quasar BLR exists, its presence ought to betray itself by some tell-tale signs. With sufficiently high signal-to-noise ratio (S/N), a BLR should be readily observed if it is present. It could be that the polarized signal is swamped by unpolarized light and is undetectable, or perhaps other factors such as the geometry and distribution of the surrounding medium are less than optimal for scattered light to reach our line of sight. In such cases, little or no light from the central BLR could be observed even if it existed. There are reasons to suggest that although this situation may well be possible, it is unlikely. First, if the quasar “monster” is present, deep, high S/N observation at Keck should be able to reveal it, even in the presence of a relatively large starlight fraction. This has been demonstrated by the spectacular detection of broad Balmer lines in Cygnus A (Ogle et al. (1997)), which have long been sought after in previous fruitless attempts (Goodrich & Miller (1989); Jackson & Tadhunter (1993)). At the level of S/N of the present observations, we should have easily detected the broad lines if they were present at a polarized flux level similar to that in Cygnus A. Second, a detection of broad polarized emission lines in an object dominated by a strong Balmer absorption-line (A star) spectrum would be unprecedented. Of the approximately two dozen HBLR type 2 AGNs detected by spectropolarimetry to date (e.g., Tran (1995); Tran et al. (1995); Young et al. (1996); Tran et al. (1998); Cohen et al. (1998)), none fits this description. The most likely, and reasonable explanation for this is that such objects do not exist or are very rare. The similar galaxy fractions detected in systems both with and without strong Balmer absorption lines (§3.2, 3.3) make it unlikely that the lack of detectable polarization (and hence of hidden quasar) in the former is simply due to the large dilution of scattered light by hot stars. If we believe in a scenario in which both the QSO and starburst activities are triggered by close interaction and can coexist (Stockton (1998)), then this may imply that the life times of the “starburst phase” of QSOs are very short, comparable to those of the most massive stars. Third, our finding is consistent with the conclusions reached by the near-infrared spectroscopic study of Veilleux et al. (1997), who showed that the optically LINER-like ULIRGs do not show any evidence for broad Paschen lines or strong high-ionization \[Si VI\] emission (which would signify the presence of a genuine monster). By contrast, most, if not all (7–9 out of 10) of the Seyfert 2 ULIRGs in their sample were found to display obscured BLR and strong \[Si VI\]. Since TF J1020+6436, and perhaps FF J1614+3234 are marginal Seyfert 2s, borderlining LINERs, they are not expected to contain hidden BLRs.
Since the ULIRG phenomenon appears to be a mix of either AGN-dominated and starburst-dominated events, it is of interest to see if it could be determined, from optical spectroscopic and $`IRAS`$ photometric data alone, whether an ULIRG harbors a genuine quasar or is powered by a starburst. The above results suggest that the emission-line ionization level of an ULIRG plays an important role in identifying the main power source. Another important factor is the IR color $`f_{25}/f_{60}`$, as it has been shown that the success of detecting HBLRs is well correlated with this ratio in both Seyfert 2 galaxies (Heisler, Lumnsden, & Bailey 1997) and ULIRGs (Veilleux et al. (1997)). Figure 8 shows a plot of the line ratio \[O III\] $`\lambda `$5007/H$`\beta `$(narrow), which can serve as an indicator of the ionization level, versus $`f_{25}/f_{60}`$. Shown are those narrow-line ULIRGs in which HBLRs have been actively searched for either in near-IR spectroscopy (Veilleux et al. (1997)), or optical spectropolarimetry (this paper). They include not only Seyfert 2 galaxies, but also sources classified as H II and LINERs with available line ratios and IR colors. Most optical spectroscopic data are obtained from Veilleux et al. (1995) and Kim, Veilleux, & Sanders (1998). We also include a few other well-known ULIRGs (Arp 220, Mrk 273, NGC 6240, UGC 5101), the three “hyperluminous” sources (F10214+4724, P09104+4109, F15307+3252), as well as Mrk 463E, a Seyfert 2 galaxy harboring HBLR with $`L_{IR}=10^{11.8}`$ $`L_{\mathrm{}}`$. As can be seen, there is a clear tendency for higher-ionization and “warmer” Seyfert 2 ULIRGs to show HBLR indicative of a “buried quasar.” Furthermore, Seyfert ULIRGs without HBLRs lie in a region of the diagram similar to that occupied by H II and LINER galaxies, none of which has been found to have HBLRs.
Figure 8 is analogous to the diagnostic diagram of Genzel et al. (1998), who use a plot of the mid-IR line ratio \[O IV\] 25.9 $`\mu `$m/\[Ne II\] 12.8 $`\mu `$m versus the strength of the 7.7 $`\mu `$m PAH (polycyclic aromatic hydrocarbons) feature, and to that of Lutz et al. (1998), who use the same PAH feature and the 5.9$`\mu `$m/60$`\mu `$m continuum flux ratio, to separate out ULIRGs that are predominantly powered by AGN or starburst. Unlike these diagrams, however, Figure 8 not only involves the much more accessible (for moderate redshifts) optical lines of \[O III\] and H$`\beta `$, but can also identify those AGNs harboring hidden quasars.
Thus, it appears that when an IR-luminous object displays a very high IR luminosity (ultraluminous), is relatively “warm”, and shows a high-ionization spectrum characteristic of Seyfert 2 galaxies, it would most certainly contain an HBLR, (i.e., it is truly a buried quasar). TF J1736+1122, for example, shows a classic strong-ionization Seyfert 2 spectrum, a warm $`f_{25}/f_{60}`$ color of 0.404, an IR luminosity of $`L_{IR}10^{12}`$ $`L_{\mathrm{}}`$, and is observed to show a spectacular HBLR. It is highly polarized, with properties that are consistent with there being a quasar. FF J1614+3234, on the other hand, has an extremely high $`L_{IR}`$ ($`>10^{12.6}`$ $`L_{\mathrm{}}`$), placing it in the company of the hyperluminous IR galaxies, and may be relatively warm ($`f_{25}/f_{60}`$ $`<`$ 0.31), but displays a feeble LINER spectrum. It is not polarized and there is no evidence of any hidden broad lines. The one characteristic that the ULIRGs with HBLR have in common is the display of classic high-ionization optical spectrum of a Seyfert 2. This then appears to be a necessary condition for an ULIRG to harbor a buried quasar.
### 4.4 The Influence of Quasars on Diagnostic Properties of ULIRGs
We now discuss how these three conditions – ionization level, IR luminosity, and IR color – are related to the existence of an obscured quasar in these AGNs.
a) High-ionization Seyfert 2 spectrum: This is necessary because a hard power-law input spectrum is needed to produce the high-ionization emission lines characteristic of an AGN. This is most likely produced by an accretion disk feeding a central black hole rather than by a population of hot stars, or bursts of star formation (with possible additional contributions from shocks).
b) High IR luminosity: Veilleux et al. (1995) and Kim et al. (1998) have found that the fraction of Seyfert galaxies among luminous infrared galaxies increases dramatically for $`L_{IR}>10^{12}`$ $`L_{\mathrm{}}`$. It follows naturally then that the number of galaxies with HBLRs also increases with IR luminosity. The enormous $`L_{IR}`$ in these objects likely arises from reprocessed optical/UV/X-ray radiation of the ionizing continuum which was absorbed by the putative dusty molecular torus surrounding the central engine (e.g., Granato, Danese, & Franceschini 1996).
c) Warm IR color: Hutchings & Neff (1991) and Kay (1994) have suggested that HBLR Seyfert galaxies tend to show warmer IR color. This has recently been confirmed by Heisler et al. (1997) in their survey of southern Seyfert 2 galaxies. They suggest that this trend could be explained in the context of the obscuring torus paradigm, where Seyfert 2s with HBLRs are those viewed at intermediate but not too high an inclination angle, so that our line of sight encompasses much of the inner, warmer regions of the dusty torus. Since the dusty torus could be optically thick even to mid-IR photons (12$`\mu `$m–25$`\mu `$m, Dopita et al. (1998)), at larger viewing angle, our line of sight intercepts mostly the outer, cooler gas and dust in the torus, giving rise to a Seyfert 2 with cool IR color and no detectable HBLR. This implies that the scattering takes place very close to the central region. The same picture may also be true in the ULIRGs, leading to a similar correlation between the visibility of HBLR and warmness (Veilleux et al. (1997); Fig. 8).
The viewing inclination to these objects can in principle be constrained by the inferred intrinsic polarization of these ULIRGs, and thus provides a way to test the above interpretation. The polarizations from ULIRGs with HBLRs are generally very high ($``$ 20–30%), indicating that the viewing angle is also large. The highest intrinsic polarization observed is perhaps that of TF J1736+1122 presented in this paper with $`P`$ $``$ 50%. For a uniformly filled cone of optically thin scattering material with a typical half-opening angle of $``$ 40° (e.g., Wilson & Tsvetanov (1994)), we obtain an inclination angle of $`i`$ 70°, based on the model of Brown & McLean (1977) (see also Miller & Goodrich (1990); Miller et al. (1991); Wills et al. (1992); Balsara & Krolik (1993); Hines & Wills (1993); Brotherton et al. (1998)). This seems too high for the picture proposed by Heisler et al. If multiple scattering in clumpy clouds as well as dichroic extinction by optically thick gas contribute, the intrinsic polarization would be lower (Kartje (1995); Kishimoto (1996)), and thus $`i`$ would have to be even higher. Similarly, if the cone opening angle is larger than 40°, the inferred $`i`$ would be higher for a given $`P`$. For smaller cone or torus opening angle, the required $`i`$ would be reduced, but the fraction of scattered light intercepted by the observer would also be much less. For example, if the half-opening angle were 20°, $`i`$ would be $``$ 58°, but the fraction of light scattered would be reduced by a factor of 3.5, making it much harder to detect the polarization. In any case, for $`P`$ = 50%, $`i`$ cannot be smaller than $``$ 55°, at which point virtually no light gets scattered from the hidden quasar since the cone would be infinitely narrow. Thus it seems likely that our viewing angle to TF J1736+1122 is $`\stackrel{>}{}`$ 70°, and yet broad polarized emission lines are easily seen and a warm IR color is detected, in contrast to the expectations of the Heisler et al. interpretation.
An alternative picture is that, in the case of the ULIRGs, the warmness of the source, as well as the high-ionization level, is a direct result of the very presence of a quasar “monster” in the nucleus, and not solely a consequence of our viewing angle. Apparently, in order for dust grains to radiate efficiently at 25 $`\mu `$m, heating sources with very high energy densities (e.g., AGN) may be required (cf., Rowan-Robinson & Crawford 1989; Condon, Anderson, & Broderick 1995). In most “cold” ULIRGs the quasar simply does not exist, has not yet had sufficient time to form, or has burned out; the energetics are instead dominated by starburst, H II region, or LINER activity. The lack of HBLRs in these types of objects is unlikely to be due to dust obscuration (e.g., Veilleux et al. (1997)), since even in such cases scattered broad lines could still be observed.
## 5 Conclusions
Although evidence of starburst activity is found in all three objects of the present study, some ULIRGs are powered predominantly by quasars, and some are mainly energized by starbursts with no indication of an obscured AGN. Thus their high IR luminosity of $``$ 10<sup>12</sup>–10<sup>13</sup> $`L_{\mathrm{}}`$ does not guarantee the presence of a quasar in the center of these objects, such as those found in other well-known ULIRGs (e.g., F10214+4724, P09104+4109, and F15307+3252). Apparently, there is a class of galaxies which have IR luminosities similar to those of the “classical” $`IRAS`$-warm galaxies but there are no detectable quasars residing in them. The ULIRG phenomenon, therefore, can be both explained by the presence of a quasar or a powerful starburst.
We discovered a new hidden broad-line $`IRAS`$ galaxy TF J1736+1122, which is intrinsically highly polarized ($``$ 50% in broad H$`\alpha `$), independent of wavelength. The obscured source has characteristics consistent with it being a quasar. The detection of HBLR in this galaxy and the non-detections in FF J1614+3234 and TF J1020+6436 are consistent with the idea that buried quasars preferentially reside in ULIRGs having warm colors, but with the necessary condition that their spectra must show characteristics of a high-ionization Seyfert 2 galaxy. Quasar-hosting galaxies also seem to have a preference for a dominant stellar population that is old ($`>`$ 1 Gyr), while galaxies without detectable sign of a quasar display a younger stellar population. If the AGN and starburst activities in these objects are triggered by the same mechanism at the same time (cf. Stockton (1998)), then our findings indicate that quasars are likely to be found only in evolved systems with age $`\stackrel{>}{}`$ 300 Myr. More observations of a larger sample of ULIRGs, and especially of quasar host galaxies are needed to further examine this issue.
We thank Carlos “the mad Belgian” De Breuck for fruitful discussions, and Jan Tweed for her assistance with Figure 2. The W. M. Keck Observatory is operated as a scientific partnership between the California Institute of Technology and the University of California, made possible by the generous financial support of the W. M. Keck Foundation. Work performed at the Lawrence Livermore National Laboratory is supported by the DOE under contract W7405-ENG-48. R.A. acknowledges the support of NSF grant AST-9617160. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.
|
no-problem/9812/astro-ph9812111.html
|
ar5iv
|
text
|
# Theoretical stellar models for old galactic clusters
## 1 Introduction
Evolutionary stellar models represent a key ingredient for any investigation about the evolution of galaxies. Theoretical predictions concerning cluster isochrones represent the only available clock for assessing the age of stellar clusters in our Galaxy and beyond, allowing a meaningful approach to the star formation history in local group galaxies. Moreover, the body of theoretical prescriptions concerning the evolutionary status of stars with different ages and chemical compositions is at the basis of the present effort for understanding the radiative properties of galaxies in terms of stellar populations. According to such evidence, it appears of obvious relevance to rely on evolutionary scenarios as detailed and reliable as possible.
In a recent paper (Cassisi et al. 1998, hereinafter Paper I) we have revisited stellar models for old Population II stars, discussing the effects of both the most recent improvements in the input physics and of element diffusion within the stellar structures. In this paper we will use the same evolutionary scenario to extend theoretical predictions to stellar metallicities and ages suitable for Population I stars. As in Paper I, we will discuss in some detail the reliability of evolutionary predictions vis-a-vis well known uncertainties of the theoretical scenario, as a useful warning against an incorrect use of theoretical predictions.
The results of selected evolutionary computations will be presented and discussed in the next section. Sections 3 will deal with a comparison with observational data for the two clusters NGC2420 and M67, taken as representatives of the ”best observed” galactic clusters in selected classes of age. On this basis we will collect observational suggestions concerning a suitable choice of the mixing length parameter to be used in constructing models of cool stars, together with some warnings about the use of theoretical results. Section 4 will present and discuss cluster isochrones for selected choices about star metallicity. General conclusions are presented in the subsequent section.
## 2 Theoretical models
For the sake of discussion, let us first neglect element diffusion to recall some relevant points concerning the evolutionary scenario for Population I stars. The upper panel of Figure 1 shows selected tracks in the logL, logTe diagram, as evaluated by adopting a ratio of the mixing length to the local pressure scale height $`\alpha `$=1.6, and under the alternative assumptions Z=0.007, 0.010, 0.015 and 0.020. Data in this figure show the remarkable similarity of tracks at fixed mass but different metallicities. As already known, an increase in the metal content moves ZAMS models toward redder colors and fainter magnitudes, whereas the difference in temperature between the ZAMS and the Red Giant Branch (RGB) decreases.
However, one knows that theory gives no firm constraints about the most appropriate value of $`\alpha `$, so that it appears of some relevance to bring to the light the consequences of a variation of the mixing length parameter within reasonable limits. This is shown in the lower panel of the same Figure 1, which reports evolutionary tracks given in the upper panel for Z=0.015 but under the three alternative choices $`\alpha `$= 1.0, 1.6 or 2.2. As already known, one finds that theory gives firm predictions about the temperature of ZAMS stars only for stars with masses larger or of the order of 1.5 M. Below this limit the uncertainty in the temperature appears roughly of the same order of magnitude of the difference produced in the upper panel of Figure 1 by the variation in metallicity. Moreover, the figure reinforces the evidence that theory gives only marginal constraints on the temperature (and thus on the color) of the Red Giant Branch (RGB), which appears as a free parameter to be adjusted -in the case- by tuning the assumption on $`\alpha `$.
To allow the comparison with the observed C-M diagrams, one has to use evolutionary tracks to predict cluster isochrones in the chosen bands of magnitudes and colors. This has been done by first producing ”theoretical” isochrones in the logL, logTe diagram, to be finally transformed into C-M diagrams according to suitable assumptions about the bolometric corrections and the color temperature relations. Let us notice that the intermediate step represent the ”true” evolutionary result, to be compared with the result of similar evolutionary evaluations before the intervention of further assumptions about model atmosphere computations.
Figure 2 shows a selected sample of the new theoretical isochrones, as computed for the labeled assumptions about the cluster ages, $`\alpha `$=1.6 and Z=0.02. One may notice the mild variation of the luminosity of the clump of He burning stars for cluster ages larger than 1 Gyr, already discussed in Castellani, Chieffi & Straniero (1992, hereinafter CCS). As a consequence, for each given chemical composition and in the above quoted range of ages the difference in magnitude between the clump and the MS turn off appears a good indicator of the cluster ages and, for each age, the clump luminosity appears as a relevant standard candle for deriving the cluster distance modulus. Note that for ages lower than 1 Gyr the luminosity of the clump follows the variation of the size of the He core through and beyond the ”Red Giant Transition” as already exhaustively discussed in the current literature (see, e.g., Sweigart, Greggio & Renzini 1990).
Isochrones with solar composition (Z=0.02) allow a comparison with previous CCS results, as obtained from the same code with the same assumption about the mixing length, but with ”old” input physics. As a most significant comparison we will chose a cluster age large enough for red giants undergoing electron degeneracy. The comparison, as given in Figure 3 (upper panel) for the 4 Gyr isochrones, shows only marginal differences. However, the use of more recent evaluations of colors gives larger differences in the predicted CM diagrams. This is shown in the lower panel of the same Figure 3, which shows the new 4 Gyr isochrone as translated into the V, B-V diagram adopting for MS stars the empirical relations given by Alonso et al. (1996), implemented with Castelli et al. (1997a,b) model atmospheres, shifted by B-V$``$-0.04 to match the empirical relations at their limit of validity ($`\mathrm{log}g4`$, $`T_e6700`$ K). In this way we attempted to match model atmospheres to current observational constraints, though at the cost of some marginal discontinuities in the predicted run of colors. Comparison with the isochrone presented by CCS, as given in the same lower panel, reveals -in particular- that the new transformations predict a steeper MS, with consequences that will be discussed later on in this paper.
Figure 4 discloses the effect on the isochrones of different assumptions about the mixing length. As expected on the basis of evolutionary tracks in Figure 1, one finds that cooler stars are strongly affected by the mixing length. MS and TO stars are free from such a theoretical uncertainty only for B-V smaller than B-V$``$0.4. For color redder than this limit the uncertainty in color grows, reaching approximately $`\delta (BV)`$0.1 mag. in the interval Mv=4.0-6.0 mag., slowly decreasing at even lower MS luminosities.
The above evolutionary evidence is a consequence of the well known occurrence for which above a given effective temperature stellar envelopes become free of convection and, thus, not affected by the treatment of such a mechanism. Note that decreasing the MS mass, the density of the envelope increases and convection becomes less and less superadiabatic; as a consequence, the (convective) less massive MS structures are only marginally affected by assumptions about the mixing length (vandenBerg, Hartwick & Dawson 1983). This occurrence explains the decreasing shift produced by variation in the mixing length on the lower portion of the MS in Figure 4. The different dependence of RG and MS stars at a given color is again related to the density of the envelopes: RG have rarefied envelope, thus experiencing strong superadiabaticity, whereas MS stars have much denser envelopes, thus less affected by superadiabatic convection.
As already noticed by Chaboyer (1995) for Population II systems, one finds that the mixing length does affect the calibration of the TO magnitudes in terms of cluster ages. According to the discussion given in Paper I, the effect of mixing length becomes larger for TO points located at redder colors, i.e., when increasing either the age or the star metallicity. Figure 4 shows that such an error is rather negligible for the 3 Gyr isochrone, but it increases to about 1 Gyr for the 6 Gyr isochrone, growing up for larger ages. Thus the TO magnitude cannot supply age evaluations for old metal rich clusters more precise than this limit, until an improved treatment of convection becomes available for stellar evolutionary models.
Concerning He burning stars, the previous Figure 2 shows that for ages larger than, about, 1 Gyr, He burning occurs near the RG branch, in agreement with observational evidences. However, the beautiful theoretical constraint given by the rather constant luminosity of the clump of He burning stars, will give less firm predictions when transferred into the CM diagram, since magnitudes depend on the bolometric correction which, in turn, depends on the color of the clump and, thus, on the assumptions adopted for the mixing length parameter.
Before closing the discussion about the mixing length, let us here advise the reader that the problem is not, or at least it could not be, to find out the ”right value” of the mixing length. In fact, the mixing length theory is only a rough (though useful) parametrization of the efficiency of convection, and there is no reason for constraining the mixing length parameter to be the same in stellar structures characterized not only by different stellar masses or chemical compositions, but also in different evolutionary phases of a given star. Thus the above discussion about the effects of the mixing length has to be taken as an investigation of the range of theoretical uncertainty on the various evolutionary phases, without necessarily assuming that, e.g., MS and RG should have effective temperatures corresponding to a common value of $`\alpha `$.
Finally, one may investigate how far the efficiency of element sedimentation affects the above results. Figure 5 shows selected isochrones for Z=0.015 as compared with similar computations but with element diffusion taken into account. As expected, one finds that diffusion plays a role only at larger ages, decreasing the color of the turn-off (TO), with minor influence on other evolutionary phases. It turns out that the diffusion sensitively affects only the isochrones which are already affected by much larger uncertainties in the mixing length, masking the effect of diffusion. Thus one can safely use canonical, no-diffusion models, bearing in mind the above discussed theoretical indeterminations. Note that this is not the case for old metal poor globular cluster stars, where diffusion has been efficient for much larger times (see Paper I).
## 3 From observations to theory and back
According to the discussion given in the previous section one should feel reluctant to present a theoretical scenario based on a given assumption about the value of the mixing length parameter. In this section we will follow a different approach, comparing the evolutionary scenario depicted in the previous section with observational data, looking for more light about reasonable theoretical predictions. For accomplishing this goal, we focused our attention on two ”best studied” galactic clusters, namely NGC2420 and M67, for which good C-M diagrams have been already presented in the literature. In the following we will investigate the agreement between observational data and theoretical predictions, as computed by including element diffusion, searching for observational constraints to the theory.
#### 3.0.1 NGC 2420
Figure 6 shows the beautiful CM diagram of NGC2420 presented by Antony Twarog et al. (1990). Following these authors the cluster should have a reddening not smaller than E(B-V)=0.05, with some evaluations as large as E(B-V)= 0.14 (Cohen 1980). Thus the color of TO stars should be lower than B-V=0.35. According to the discussion given in the previous section, we note that the cluster appears young enough to have TO stars only marginally affected by the mixing length parameter. Such a fortunate occurrence decreases the degrees of freedom and made the cluster our first choice for fitting theory to observations.
According to Antony Twarog et al. (1990) metallicity estimates give for the cluster \[Fe/H\]=-0.35 $`\pm `$0.10, but with evaluations reaching \[Fe/H\]=-0.6 (Canterna et al. 1986). We will assume Z=0.007 (\[Fe/H=-0.4\]). By taking Y=0.23 for old metal poor stars and Y=0.27 for the Sun (Z=0.02), a linear interpolation on metallicity finally gives for NGC2420 Z=0.007 and Y=0.244.
For fitting theory to observations, taking into account the already discussed uncertainty on the location of cool stars, we devised a procedure which runs as follows. We first determined the predicted location in the CM diagram of the ”He clump” for a suitable range of assumptions about the mixing length parameter, namely $`\alpha `$= 1.0-2.2 As an exemple, this is shown in Figure 7 for a 1.5 Gyr isochrone, with Y=0.27, Z=0.007. By forcing the match between the predicted and the observed location of both TO and He burning stars one eventually finds the cluster reddening, the distance modulus, and the mixing length value suitable for the observed color of luminous red stars.
As disclosed by the same Figure 6, the fitting is not only possible, but it appears remarkably good, since it reproduces to a high degree of accuracy the shape of H burning stars near and beyond the overall contraction phases. In this way one would derive E(B-V)$``$0.16, DM$``$12.4, $`\alpha `$=1.9 and a cluster age of about 1.5 Gyr, in reasonable agreement with previous CCS results, as obtained for a solar composition. Comparison with the isochrones for the ages 1.25 and 1.75 Gyr, as given in the same figure, shows that the formal error in age should not exceed 0.1 Gyr! Of course, reddening depends on the adopted relation between temperatures and colors. Adopting only Castelli et al. (1997) models, one would derive a reddening smaller by 0.05, thus E(B-V)$``$0.11. Similarly, the distance modulus depends on the amount of bolometric corrections which, however, appears much less model dependent than colors do.
One has finally to notice a disagreement between the theoretical and the observed location of the lower portion of the main sequence, as due to the evident inadequacy of either stellar models or color temperature relations. The agreement would be improved adopting for these models smaller values of the mixing length, which means to assume external convection to be less efficient in MS stars than expected adopting a common mixing length ($`\alpha `$=1.9) for both MS and RG stars. However, Figure 7 shows that low masses MS stars are less and less affected by the mixing length. Thus it appears difficult to reconcile theory with observation only by tuning the mixing length. As we will discuss in the case of M67, we suspect that such a disagreement is due to the inadequacy of adopted color-temperature relations.
Theoretical predictions can be submitted to a further independent test. As largely discussed in CCS, in relatively young clusters the distribution of stars in the advanced evolutionary phases is a rather sensitive function of the cluster age. Thus beyond the agreement between the CM diagram loci, one can test the theoretical distribution vis-a-vis the observed distribution. This has been done by computing, with a Monte Carlo technique, the synthetic CM diagram of the cluster, with the same number of off main sequence stars as observed. Theoretical predictions, as shown in Figure 8, middle panel, appear in reasonable agreement with observations, giving further support to the predicted cluster age. The lower panel finally compares the synthetic cluster with the theoretical isochrone, giving light on the effect of binary stars on the topology of the overall contraction gap we will discuss in the final section.
### 3.1 M67
Figure 9 shows the CM diagram of M67 presented by Montgomery, Marschall & Janes (1993). According to these authors, the cluster is characterized by a metallicity \[Fe/H\]=-0.05. Correspondingly we will refer to an evolutionary scenario as computed for Z=0.015. Reddening evaluations range in the interval E(B-V)=0.03 - 0.10 (Antony Twarog et al. 1990); in particular Cohen (1980) analyzed absorption lines of sodium of the interstellar gas giving E(B-V)=0.09.
Inspection of Figure 9 discloses that now all the cluster stars lie in a region affected by the uncertainty about the mixing length. Moreover, the MS of the cluster is now largely in the range of colors where we already found a mismatch between theory and observation. The same figure 9 shows that the most luminous portion of cluster stars can be reasonably fitted by theory for an age of 3.25 Gyr, assuming again $`\alpha `$=1.9, and with the labeled values of cluster distance modulus and reddening. However the location of the theoretical MS appears far from being satisfactory.
In this context, we have already noticed that new computations appear in satisfactory agreement with previous CCS evaluation. Degl’Innocenti & Marconi (1998) have also shown that the new theoretical 4 Gyr isochrone already presented in the upper panel of Figure 2 appears also in good agreement with similar predictions by Bertelli et al. (1994). Thus theory appears rather solid in predicting the location in the theoretical HR diagram of similar stars, with only a minor influence of the updated physics inputs. Curiously enough, one finds that these isochrones were already used to give a good fitting of the M67 MS not only by CCS but also by Carraro et al. (1996). If we add the evidence that CCS were also nicely fitting the MS of NGC2420, one is driven to conclude that the fitting of these MS is largely a matter of the adopted color-temperature relations and that, unfortunately, more recent predictions of model atmospheres are moving theory away from observations.
In this respect one has to notice that recent empiric calibrations of V-K colors in terms of stellar temperature, as given by stellar interferometry, have revealed that the results of recent model atmospheres given by Gratton et al. (1996) are supported by observation only at temperatures around 7000 <sup>o</sup>K, whereas at lower temperature theoretical predictions appear in increasing disagreement with the calibration (Di Benedetto, private communication). This suggests to us that the temperature-color relation is perhaps the weak link in the connection theory-observation. According to these evidences, we are inclined to conclude that the transformation of temperatures into colors is probably the most relevant problem in using evolutionary theory. In this context, theoretical fittings as given in Figure 9 should probably be regarded only as an exercise, whereas a reliable investigation of cluster evolutionary parameters should probably wait for a much firmer assessment of this problem
Let us finally notice that Montgomery et al. (1993) by fitting CCS isochrones to the same observational data already noticed the good agreement between theory and observations in the MS and TO regions. However, the missed fitting of the RG colors induced the same authors to conclude that ”the current generation of theoretical isochrones cannot be fit to the observed sequence within the observational errors”. Even tough one cannot disagree with such a statement, we wish here to stress once again that the RG colors is largely matter of a cosmetic adjustment of theoretical results, as we have done in this section.
As already discussed, the computations of the synthetic clusters can give an independent indication at least of the compatibility of the theoretical scenario one is dealing with. Figure 10 shows that assuming 10% of binary stars one finds a satisfactory agreement of theoretical predictions with observational data.
## 4 Cluster isochrones
The discussion given in the previous sections has shown that $`\alpha `$=1.9 appears the adequate (cosmetic) choice to produce cluster isochrones for old open clusters, all over the range of metallicity Z=0.007, 0.015, at least. According to such a result, Figure 11 shows cluster isochrones computed under the above quoted assumptions and for four selected metallicities. Data for all the computed isochrones are available at the anonymous ftp at astr18pi.difi.unipi.it (/pub/open), where we give theoretical (logL, log Te) isochrones together with V, B-V and V-I magnitudes as predicted according to the already quoted match between Alonso et al. (1996) and Castelli et al. (1997 a,b) evaluations. Bearing in mind the caveat concerning theoretical colors, let us now discuss some theoretical predictions of general relevance.
Figure 12 gives the bottom luminosity of the clump of He burning stars as a function of the cluster age in the interval 1-6 Gyr and for the five explored metallicities. For each given metallicity, the figure discloses the good constancy of this parameter, which should allow the use of He burning stars as useful standard candles.
However, one finds that for ages larger than 2 Gyr, the luminosity of the clump is slowly decreasing when the age increases. Since the clump is in the meantime becoming redder, the bolometric correction increases, increasing the variation in magnitudes. This is shown in Figure 13, where we report the predicted (bottom) magnitudes of the clump as a function of the metallicity Z for selected assumptions about the cluster ages. One finds that the linear relation:
$$\mathrm{Mv}=1.59+0.59\mathrm{logZ}$$
(1)
reproduces the theoretical results over the whole range of ages within 0.1 mag.
As another relevant prediction, Figure 14 gives the difference in magnitude between the clump and the top luminosity of the H burning sequence (a parameter already introduced in CCS) as a function of the cluster ages for the chosen metallicities.
## 5 Conclusions
Galactic stellar clusters have covered a central role in the progress of stellar evolutionary theories, giving direct evidences for the evolution of metal rich, Population I stars for a fairly large variety of cluster ages and metallicities. However, throughout this paper we have shown that the situation is far from being completely satisfactory. Theoretical uncertainties on the efficiency of external convection play a major role in shading a disturbing degree of freedom in relevant theoretical predictions. Indetermination on both the temperature-color relations and in the actual cluster reddening add further problems into this scenario.
In this paper we have attempted an empirical calibration of the theory, reaching what we regard as a satisfactory agreement between theory and observation. It remains the disturbing discordance of the MS slope at the larger explored B-V value, whose origin is far from firmly established. Further improvements in color-temperature relations and in the determination of cluster reddening are possible and expected. However external convection keep being the main outstanding problem. In this sense new approaches to the theory of convection, as the one presented by Canuto & Mazzitelli (1992) or Lydon, Fox & Sofia (1993), should be carefully tested to well studied galactic clusters and, if necessary, improved in the hope of reaching a real knowledge of such a fundamental ingredient of stellar evolutionary theories.
As a final point we note that that the predicted luminosity of He burning clumps is directly correlated with the size of the He cores at the end of central H burning. Both NGC2420 and M67 are predicted with Red Giants undergoing electron degeneracy. The good correspondence between observations and theory found in the previous section supports current evaluations of the physical mechanisms affecting these structures. Note that in this case, the predicted He core are only marginally affected by the possible efficiency of convective overshooting, which however should modulate the shape of the CM diagram in the Turn Off region.
In this context Demarque, Sarajedini & Guo (1994) have recently proposed an alternative approach to constrain the efficiency of overshooting, as based on the topology of the CM diagram in the region of the overall contraction gap. In our feeling that approach, tough ingenious and theoretically ywell founded, suffers of some limitations mainly due to the occurrence of binary stars within the gap (see bottom panels in the previous Figures 8 and 10), preventing from firm observational constraints on the suggested CM parameters.
## Acknowledgments
It is a pleasure to thank Giuseppe Bono for a critical reading of the manuscript and for valuable suggestions. One of the authors, M.M., acknowledges the grant from C.N.A.A.
|
no-problem/9812/cond-mat9812354.html
|
ar5iv
|
text
|
# Reduction of the Superfluid Density in the Vortex-Liquid Phase of 𝐁𝐢_𝟐𝐒𝐫_𝟐𝐂𝐚𝐂𝐮_𝟐𝐎_𝐲
\[
## Abstract
In-plane complex surface impedance of a $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$ single crystal was measured in the mixed state at 40.8 GHz. The surface reactance, which is proportional to the real part of the effective penetration depth, increased rapidly just above the first-order vortex-lattice melting transition field and the second magnetization peak field. This increase is ascribed to the decrease in the superfluid density rather than the loss of pinning. This result indicates that the vortex melting transition changes the electronic structure as well as the vortex structure.
\]
The electronic structure of high-$`T_c`$ superconductors (HTSC’s) in the mixed state attracts much attention. In conventional superconductors (CSC’s) with $`s`$-wave gap, quasiparticle (QP) states in the mixed state are localized in the vortex cores where the superconducting gap is suppressed. The QP excitation spectrum inside the vortex core consists of quantized energy levels separated by $`\mathrm{\Delta }E\mathrm{\Delta }_0/k_F\xi `$, where $`\mathrm{\Delta }_0`$ is the bulk gap, $`k_F`$ is the Fermi momentum and $`\xi `$ is the coherence length . Since $`\mathrm{\Delta }E`$ in CSC’s is much smaller than the scattering energies, the vortex core can be regarded as a normal metal. In HTSC’s, above picture can not be applied because of the following reasons. First, the symmetry of the superconducting gap in HTSC’s is not an $`s`$-wave but is most likely a $`d`$-wave . Since the amplitude of the $`d`$-wave gap is zero at the node, QP’s are not localized in the vortex core but extend along the node direction . Calculations based on the Bogoliubov-de Gennes equations suggest that there are no truly localized states in the vortex core of pure $`d`$-wave superconductors . Secondly, even if there are localized QP states inside the vortex core owing to the mixing of the different gap symmetry , $`\mathrm{\Delta }E`$ may exceed any other energy scales, since $`k_F\xi `$ of HTSC’s is very small. Therefore the vortex core in HTSC’s should be different from that of CSC’s in either cases. Finally, in HTSC’s, the greater part of the mixed state is a vortex liquid phase which is practically lost in CSC’s. It may be possible that the difference in the vortex structure brings the difference in the electronic state.
Experimentally, scanning tunneling spectroscopies and thermal conductivity measurements have been used to study the QP state of HTSC’s in the mixed states. High frequency surface impedance $`Z_s=R_s+iX_s`$ also provides useful information on the electronic state. Using high enough frequencies, one can be free from the vortex pinning and the information not only on QP’s but also on superfluids can be deduced from $`R_s`$ and $`X_s`$, respectively. So far, a vortex flow resistivity and a QP lifetime inside the vortex core were discussed from the $`R_s`$ measurements in $`\mathrm{YBa}_2\mathrm{Cu}_3\mathrm{O}_\mathrm{y}`$ . Recently, Mallozzi et al. measured the complex resistivity in the mixed state of $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$ and argued a possible $`d`$-wave effect . However, measurements of $`Z_s`$ near the vortex-melting transition are still lacking. In this paper, we report both $`R_s`$ and $`X_s`$ measurements on a $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$ single crystal and show that the high-frequency response can not be explained in terms of simple vortex flow. Moreover, we succeeded in measuring $`Z_s`$ across the first-order vortex melting transition line and found that the reactive part, $`X_s`$, increases rapidly above the transition while there was little change in $`R_s`$. This result indicates that the superfluid density or the amplitude of the order parameter decreases in the vortex liquid phase.
A $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$ single crystal was grown by the floating zone method. The as grown crystal was annealed in air at 800 $`{}_{}{}^{}\mathrm{C}`$ for 3 days and was quenched to room temperature to achieve an optimum oxygen content. A superconducting transition temperature $`T_c`$ defined at zero resistivity was 91 K. Prior to the surface impedance measurements, we measured the magnetic-field dependence of the local magnetization of the same crystal using the micro Hall probe magnetometry and determined the first-order vortex melting transition field from the position of the magnetization jump. Surface impedance was measured by the cavity perturbation technique with a cylindrical Cu cavity operated at 40.8 GHz in the TE<sub>011</sub> mode. The sample was located at the center of the cavity which is the anti-node of the microwave magnetic field $`H_{rf}`$ being parallel to the $`c`$-axis of the sample. The dimensions of the sample were 0.5$`\times `$0.5$`\times `$0.02 mm<sup>3</sup>, which is appreciably larger than the normal-state skin depth ($``$$`\mu `$m at 40.8 GHz). The surface resistance $`R_s`$ and the surface reactance $`X_s`$ can be obtained from the changes in the quality factor of the cavity and the resonance frequency, respectively. We determined absolute values of $`R_s`$ and $`X_s`$ from comparison with the dc resistivity and assuming that $`R_s=X_s`$ (Hagen-Rubens limit) in the normal state. Using this procedure, we obtained the reasonable zero-temperature penetration depth of $``$2$`\times `$10<sup>2</sup> nm. Surface impedance measurements were performed with swept temperature $`T`$ under field cooled conditions to avoid any extrinsic effects associated with pinning e.g. giant magnetostriction . In all the measurements, the static magnetic fields $`H_{dc}`$ were applied along the $`c`$-axis.
First, we briefly introduce the general behavior of the surface impedance of type-II superconductors . The surface impedance $`Z_s`$ is related to the complex effective penetration depth $`\stackrel{~}{\lambda }`$ as $`Z_s=i\mu _0\omega \stackrel{~}{\lambda }`$, where $`\mu _0`$ is the vacuum permeability and $`\omega `$ is the angular frequency. The complex resistivity $`\stackrel{~}{\rho }`$ is also expressed by $`\stackrel{~}{\lambda }`$ as $`\stackrel{~}{\rho }=i\mu _0\omega \stackrel{~}{\lambda }^2`$. In the Meissner state, the response is reactive and $`\stackrel{~}{\lambda }`$ is purely real. Therefore, $`R_s0`$ and $`X_s=\mu _0\omega \lambda _L`$ where $`\lambda _L`$ is the London depth. In the normal state, the response is dissipative and $`\stackrel{~}{\lambda }^2`$ is purely imaginary. Therefore, $`R_s=X_s=\mu _0\omega \delta /2`$ where $`\delta =(2\rho _n/\mu _0\omega )^{1/2}`$ is the skin depth and $`\rho _n`$ is the normal resistivity. In the mixed state, vortex dynamics affects $`\stackrel{~}{\lambda }`$. If the frequency is low, vortices are effectively pinned and a response is similar to that of the Meissner state except that $`\stackrel{~}{\lambda }^2\lambda _L^2+B\mathrm{\Phi }_0/\mu _0\kappa _p`$, where $`B`$ is the magnetic induction, $`\varphi _0`$ is the flux quantum and $`\kappa _p`$ is the Labusch parameter which denotes the pinning strength. On the other hand, if the frequency is high enough, the viscous loss becomes dominant and a response is similar to that of the normal state except that $`\delta `$ is replaced by the vortex flow skin depth $`\delta _f(2B\mathrm{\Phi }_0/\mu _0\omega \eta )^{1/2}`$, where $`\eta `$ is the viscosity of the vortex motion being related to the QP excitation inside the vortex core. Within the vortex flow theory of Bardeen and Stephen , $`\eta `$ is independent of $`B`$ and $`R_sX_sB^{1/2}`$ at high enough fields where $`\delta _f\lambda _L`$ and $`R_sB`$, $`X_s\mu _0\omega \lambda _L`$ at low fields. A crossover from the pinned regime to the dissipative regime occurs at the pinning frequency $`\omega _p=\kappa _p/\eta `$.
The field dependence of $`Z_s`$ of the $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$ single crystal is shown in Fig. 1. At high fields above 0.2 T, $`R_s`$ is almost proportional to $`B^{1/2}`$. However, $`X_s`$ is larger than $`R_s`$ up to the highest field we used. This result means that the contribution from the reactive parts, superfluid and/or pinned vortices, can not be neglected. Therefore, the observed $`B^{1/2}`$ dependence of $`Z_s`$ can not be attributed to the high-field region of the Bardeen-Stephen type vortex flow.
At low fields, a prominent anomaly can be seen. Above a certain field, $`X_s`$ increases rapidly and saturates at a higher field while there is little change in $`R_s`$ in the same field region. To examine this anomaly in more detail, we plot the field induced changes in the real part of the effective penetration depth Re$`\mathrm{\Delta }\stackrel{~}{\lambda }(H_{dc})(X_s(H_{dc})X_s(0))/\mu _0\omega `$ in Fig. 2. At low fields and low temperatures, Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ is independent of $`T`$ and linear in $`H_{dc}`$. At high temperatures, Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ begins to deviate from the linear $`H_{dc}`$ dependence and rapidly increases above a certain field $`H_{kink}`$. In Fig. 3, $`H_{kink}`$ is plotted on the magnetic phase diagram together with the vortex melting transition field $`H_m`$ and the second magnetization peak field $`B_{sp}`$. Here, $`H_{kink}`$ is defined at the onset of the rapid increase in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }(T)`$ as shown in the inset of Fig. 3. The agreement between $`H_m`$ and $`H_{kink}`$ is excellent, indicating that the vortex melting affects Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$. In the present crystal, the “critical point” is located at 45 K and 33 mT. Above this field, the sharp feature in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }(T)`$ disappears as shown in the inset of Fig. 3. However, as shown in Fig. 2, the small deviation from the linear $`H_{dc}`$ dependence is still observed in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }(H_{dc})`$ around 40 mT, somewhat higher than $`B_{sp}`$. Considering the difference between the applied field and the magnetic induction inside the sample and the different processes of the field applications in the local magnetization and the microwave measurements, it is reasonable to regard that the increase in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }(H_{dc})`$ below 45 K and the second magnetization peak have the same origin, most likely the field induced vortex decoupling . To sum up, Re$`\mathrm{\Delta }\stackrel{~}{\lambda }(H_{dc})`$ increases when the 3-dimensional vortex lattice model is no longer valid. Note that this anomaly only appears in the reactive response (Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ or $`X_s`$) and almost no anomaly appears in the dissipative response ($`R_s`$).
Now we discuss the origin of the change in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$. Since $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$ does not contain magnetic ions, we can recognize that Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ consists of two components: the London penetration depth $`\lambda _L`$ and the vortex motion. First we interpret our results in terms of the change in the vortex motion, which has been widely assumed so far , and will show that this scenario is inappropriate at least near the vortex melting transition of $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$. From the vortex-motion point of view, the increase in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ should be attributed to the loss of pinning at the vortex melting transition. Since contributions from $`\lambda _L`$ and the vortex motion can be separated if the response is written by the complex resistivity . When the effect of vortex creep, which is important only near $`T_c`$, is neglected, the real ($`\rho _1`$) and the imaginary ($`\rho _2`$) parts of the complex resistivity are given by:
$$\frac{\rho _1}{\mu _0\omega \lambda _L^2}=\frac{s}{1+s^2}+\frac{r^2+sr}{(1+s^2)(1+r^2)}b,$$
(2)
$$\frac{\rho _2}{\mu _0\omega \lambda _L^2}=\frac{1}{1+s^2}+\frac{rsr^2}{(1+s^2)(1+r^2)}b.$$
(3)
Here, $`s=(2\lambda _L^2/\delta _{nf}^2)`$, $`\delta _{nf}`$ is the normal-fluid skin depth, $`r=\omega /\omega _p`$ and $`b=B\varphi _0/\mu _0\omega \eta \lambda _L^2`$. The parameters $`s`$ and $`r`$ denote the normal-fluid fraction relative to the superfluid density and the weakness of the pinning, respectively. Note that all the parameters related to the vortex motion ($`r`$ and $`b`$) are included only in the second terms of the right hand sides of Eq. Reduction of the Superfluid Density in the Vortex-Liquid Phase of $`\mathrm{𝐁𝐢}_\mathrm{𝟐}\mathrm{𝐒𝐫}_\mathrm{𝟐}\mathrm{𝐂𝐚𝐂𝐮}_\mathrm{𝟐}𝐎_𝐲`$. Therefore, if only the vortex motion is considered, magnetic field dependent parts of the complex resistivity $`\mathrm{\Delta }\rho _1`$, $`\mathrm{\Delta }\rho _2`$ are nothing but these terms. Here, we introduce the ratio of these two terms: $`\mathrm{\Delta }\rho _2/\mathrm{\Delta }\rho _1=(rsr^2)/(r^2+sr)`$. Let us consider the behavior of this ratio at the vortex melting transition. If the vortices melt, the effective pinning strength must be weakened. Accordingly, $`r1/\kappa _p`$ should increase above the transition. Since $`\mathrm{\Delta }\rho _2/\mathrm{\Delta }\rho _1`$ monotonically decreases with increasing $`r`$, $`\mathrm{\Delta }\rho _2`$ should decrease relative to $`\mathrm{\Delta }\rho _1`$ at the transition. This behavior is expected to be irrespective of the nature of pinning, surface or bulk, because any depinning processes cause energy losses which mainly affect $`\rho _1`$. To compare the above arguments with the experimental results, we calculated $`\stackrel{~}{\rho }`$ from $`Z_s`$. As shown in Fig. 4, the experimentally obtained $`\mathrm{\Delta }\rho _2`$ increases relative to $`\mathrm{\Delta }\rho _1`$ at the vortex melting transition and the second magnetization peak field. Therefore, the increase in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ can not be attributed to the loss of pinning. Instead, it should be originated from the increase in $`\lambda _L`$ itself. Since $`1/\lambda _L^2`$ is proportional to the superfluid density, this result strongly suggest that the superfluid density decreases in the vortex liquid phase of $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$ due to the additional pair breaking. If all the QP’s are localized in the vortex core as in $`s`$-wave superconductors, the electronic state might be insensitive to the change in the vortex structure. Therefore, we speculate that the reduction of the superfluid density at the vortex melting transition is related to the $`d`$-wave superconductivity.
Next we discuss the behavior of Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ in the vortex solid phase. Since $`R_s0`$ in this regime, we can recognize that vortices are effectively pinned so that $`r1`$, and $`\stackrel{~}{\lambda }(\lambda _L^2+B\mathrm{\Phi }_0/\mu _0\kappa _p)^{1/2}\lambda _L+B\mathrm{\Phi }_0/2\lambda _L\mu _0\kappa _p`$. If $`\lambda _L`$ is independent of $`H_{dc}`$, $`\kappa _p`$ should also be independent of $`H_{dc}`$ and $`T`$, since Re$`\mathrm{\Delta }\stackrel{~}{\lambda }B`$ and independent of $`T`$ as shown in Fig. 2. Field independent $`\kappa _p`$ is only expected when the vortices are individually pinned by strong pinning centers. This is unreasonable because we observed the sharp first-order magnetization jump in our sample. Therefore, $`\lambda _L`$ as well as $`\kappa _p`$ should be $`H_{dc}`$ dependent even in the vortex solid phase. Let us assume that the field dependence of $`\lambda _L`$ is dominant, as has been argued by Mallozzi et al. in the vortex liquid phase . The field dependent $`\lambda _L`$ in the mixed state of $`d`$-wave superconductors is discussed in terms of the magnetic-field effect on the extended QP state by introducing the Doppler energy shift due to the circulating current around the vortices . According to these theories, normal-fluid density, which is proportional to the change in $`\lambda _L`$, varies as $`H_{dc}^{1/2}`$ when $`E_H>\mathrm{\Gamma }`$ and varies as linear in $`H_{dc}`$ when $`E_H<\mathrm{\Gamma }`$, where $`E_H`$ is the averaged QP energy shift given by $`E_H\mathrm{\Delta }_0(H_{dc}/H_{c2})^{1/2}`$ and $`\mathrm{\Gamma }`$ is the QP scattering rate by impurities or the thermal agitation. In terms of this scenario, since the observed Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ in the vortex solid phase is linear in $`H_{dc}`$ and independent of $`T`$, $`E_H<\mathrm{\Gamma }`$ and $`\mathrm{\Gamma }`$ should be dominated by the impurity scattering. However, the zero-field penetration depth of the same sample is found to be linear in $`T`$ at low temperatures. This suggests that the dominant pair breaking mechanism should not be the impurity scattering but the thermal smearing. To resolve the discrepancy between Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ in the vortex solid phase and the zero-field $`\lambda _L`$, knowledge of the field dependence of $`\kappa _p`$ and/or more detailed theoretical treatment must be needed.
Finally, we should briefly consider Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ above $`H_m`$. As shown in Fig. 2, change in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ just above the transition is rather gradual one. Considering that the melting transition is of the first order, there should be a small discontinuity in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ at $`H_m`$, which, we believe, is not observed because of sparse data points. The succeeding gradual change at higher fields may come from the increasing pair breaking and/or the vortex motion. If all the changes in Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$ in the vortex liquid phase arise from the pair breaking , the quantity $`\lambda _L(T=0)^2/(\mathrm{Re}\stackrel{~}{\lambda })^2`$ at low temperatures represents the superfluid fraction $`f_s`$. We calculated this quantity at 10 K and found that $`\lambda _L(T=0)^2/(\mathrm{Re}\stackrel{~}{\lambda })^20.3`$ at $`H_{dc}`$=1 T. Even if $`d`$-wave effect is taken into account, this value is too small since $`f_s`$ is given by $`f_s(H_{dc})1(H_{dc}/H_{c2})^{1/2}`$ and $`H_{c2}`$ is considered to be around 100 T. Therefore, both the increase in $`\lambda _L`$ and the contribution from the vortex motion should be considered in the vortex liquid phase. We stress here again that the behavior of $`\mathrm{\Delta }\rho _2/\mathrm{\Delta }\rho _1`$ at the transition can not be explained in terms of the vortex motion alone, even though the two effects contribute to Re$`\mathrm{\Delta }\stackrel{~}{\lambda }`$. To separate both contributions, detailed frequency dependence measurements are indispensable. This experiment is now underway.
In conclusion, we have measured the complex surface impedance $`Z_s=R_s+iX_s`$ of a $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_\mathrm{y}`$ single crystal in the mixed state. We succeeded in detecting the change in $`Z_s`$ at the first-order vortex melting transition and the second magnetization peak field for the first time. Above the transition, $`X_s`$ which is proportional to the real part of the effective penetration depth increases while there was little change in $`R_s`$. From the analysis of the complex resistivity deduced from $`Z_s`$, we showed that the increase in $`X_s`$ can not be ascribed to the loss of pinning but arises from the reduction of the superfluid density. Namely, the additional pair breaking mechanism may exist in the vortex liquid phase. Our results indicate that not only the phase but also the amplitude of the order parameter take different values in different vortex phases. We speculate that this effect is related to the $`d`$-wave superconductivity.
The authors thank to Y. Matsuda and H. Kitano for helpful discussions. This work was partly supported by the Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Sports and Culture of Japan.
|
no-problem/9812/astro-ph9812048.html
|
ar5iv
|
text
|
# NGC 2613, 3198, 6503, 7184: Case studies against ‘maximum’ disks
## 1. Introduction
‘Maximum’ disk versus submaximal disk decompositions of the rotation curves of spiral galaxies have been discussed at great length in the literature (cf. the articles by Bosma and Sellwood in this volume). The aim of this paper is to draw attention to the implications of such models of the rotation curves for the internal dynamics of the disks.
The sample of galaxies studied here has been drawn from the list of Bottema (1993) of galaxies with measured stellar velocity dispersions. The criteria were: (a) the rotation curve of each galaxy, preferentially in HI, is observed, (b) each galaxy is so inclined that the planar velocity dispersions are measured, but (c) that its morphology is still discernible.
## 2. Decomposition of the rotation curves and diagnostic tools
The rotation curve of each galaxy is fitted by the superposition of contributions due to the stellar and gaseous disks, both modelled by thin exponential disks, the bulge (NGC 2613 only), modelled by a softened $`r^{3.5}`$ density law, and the dark halo, modelled by a quasi-isothermal sphere,
$$v_\mathrm{c}^2(R)=v_{\mathrm{c},\mathrm{d}}^2(R)+v_{\mathrm{c},\mathrm{g}}^2(R)+v_{\mathrm{c},\mathrm{b}}^2(R)+v_{\mathrm{c},\mathrm{h}}^2(R).$$
(1)
The radial scale lengths of the disks, $`h`$, and core radii of the bulges, $`r_{\mathrm{c},\mathrm{b}}`$, as well as the bulge to disk ratios have been adopted from published photometry of the galaxies (cf. Bottema (1993), Broeils (1992) and references therein). Only in the cases of NGC 3198 and 6503 HI data were available, which allowed the determination of the $`v_{\mathrm{c},\mathrm{g}}`$ contribution in Eq. (1). No quantitative photometry of the bulge of NGC 7184 is available.
The diagnostic tools, which I use to analyze the rotation curve models, are the Toomre stability parameter of the disks and, following Athanassoula et al. (1987), the predicted multiplicity of the spiral structures. The Toomre stability parameter is given by
$$Q=\frac{\kappa \sigma _\mathrm{U}}{3.36G\mathrm{\Sigma }_\mathrm{d}}.$$
(2)
In Eq. (2) $`\kappa `$ denotes the epicyclic frequency, which can be directly derived from the rotation curve, $`\sigma _\mathrm{U}`$ the – measured – radial velocity dispersion of the stars, $`G`$ the constant of gravitation, and $`\mathrm{\Sigma }_\mathrm{d}`$ the surface density of the disk, which follows from the fits to the rotation curves. The stability parameter must lie in the range 1 $`<Q<`$ 2, in order to prevent Jeans instability of the disk, on one hand, and to allow the disks to develop spiral structures, on the other hand. All the galaxies studied here are not grand-design spirals. In these galaxies the spiral structures are almost certainly due to ‘swing amplification’ of perturbations of the disks (Toomre 1981). This mechanism is most effective, if the circumferential wave length of the density waves is
$$\lambda =X\left(\frac{dv_\mathrm{c}(R)}{dR}\right)\lambda _{\mathrm{crit}}=X\left(\frac{dv_\mathrm{c}(R)}{dR}\right)\frac{4\pi ^2G\mathrm{\Sigma }_\mathrm{d}}{\kappa ^2}.$$
(3)
The value of the $`X`$ parameter is about 2 in the case of a flat rotation curve, but less in the rising parts of the rotation curve (Athanassoula et al. 1987). I apply in Eq. (3) a relation for $`X(\frac{dv_\mathrm{c}(R)}{dR})`$ found by analyzing the stellardynamical equivalent of the Goldreich & Lynden-Bell sheet (Fuchs 1991). The expected number of spiral arms is given by $`m=2\pi R/\lambda `$. Eq. (3) is derived from local density wave theory. Recent alternative approaches based on global analyses of non-axisymmetric perturbations of galactic disks are described by Haga & Iye (1994), Evans & Read (1998a, b), and Pichon & Cannon (1998).
Decompositions of the rotation curves of the galaxies, which maximise the disk contribution in Eq. (1), are shown in Figs. 1a and b together with the resulting stability parameters and expected multiplicities of spiral arms. As can be seen from Figs. 1 the $`Q`$ parameters are systematically close to or even less than one. That is impossible in real galactic disks. As is well known since the classical paper by Sellwood & Carlberg (1984), the disks would evolve fiercely under such conditions and heat up dynamically on short time scales. If the model of Sellwood & Carlberg is scaled to the dimensions of NGC 6503, the numerical simulations indicate that the disk would heat up within a Gyr from $`Q`$ = 1 to 2.2 and any spiral structure would be suppressed. The amount of young stars on low velocity dispersion orbits, which would have to be added to the disk in order to cool it dynamically back to $`Q`$ = 1, can be estimated from Eq. (2). In NGC 6503 a star formation rate of 40 $`M_{}/pc^2/Gyr`$ would be needed, while actually a star formation rate of 1.5 $`M_{}/pc^2/Gyr`$, as deduced from the H<sub>α</sub> flux (Kennicutt et al. 1994), is observed. Thus ‘maximum’ disks seem to be unrealistic under this aspect.
Furthermore, according to the ‘maximum’ disk models the galaxies should be two-armed spirals. This is in agreement with Athanassoula et al. (1987) and Haga (1998, private communication), who concluded that in NGC 3198 and 6503 $`m`$ is at least two. However, as can be seen on images of the galaxies (cf. The Carnegie Atlas of Galaxies), all the galaxies discussed here have a multi-armed, irregular morphological appearrance.
Both deficiencies can be remedied simultaneously, if submaximal disks are assumed. This is illustrated in Fig. 2 for NGC 3198, where the mass-to-light ratio of the disk has been reduced from $`M/L_\mathrm{B}`$ = 3.5 to 2.2 $`M_{}/L_{\mathrm{B},}`$. Within the optical radius the dark halo contributes twice the mass of the disk and its core radius is of the order of the radial scale length of the disk. As can be seen from Fig. 2, the $`Q`$ parameter lies in a more realistic range and the predicted multiplicity of spiral arms fits better to the observed morphology of the galaxy than in the ‘maximum’ disk model.
I am grateful to E. Athanssoula, A. Bosma, and J. Kormendy for valuable hints and discussions.
## Discussion
Kormendy: I’m worried about the interpretation of velocity dispersion measurements in late-type disks that contain young stars. There is a danger that stars that dominate the spectra are relatively bright, young, and low in velocity dispersion and that the stars that dominate the disk mass are older, fainter and hotter. Our Galactic disk shows just such an effect. It would imply that measured dispersion values may be too small to correctly represent the disk mass in calculations of $`Q`$.
Fuchs: Yes, one has to keep this in mind. For the Galactic disk the effect can be estimated using data from the solar neighbourhood. A detailed analysis shows that the luminosity-weighted, scale-height corrected radial velocity dispersion of stars in the Galactic disk is $`\sigma _\mathrm{U}`$ = 36 km/s, which has to be compared with 44 km/s of the old disk stars. The weight of young stars is 25% of the total weight. In Sc galaxies, which are bluer than the Galaxy with an averaged $`<BV>`$ of 0.66 mag, this might be shifted even more towards young stars. On the other hand, Sc galaxies are more gas rich, which has a destabilizing effect. Taken all together, the $`Q`$ argument seems to be quite robust.
## References
Athanassoula, E., Bottema, R., & Papaioannou, S. 1987, A&A, 179, 23
Bottema, R. 1993, A&A, 275, 16
Broeils, A. 1992, thesis Univ. of Groningen
Evans N. W., Read, J. C. A. 1998a, b, MNRAS, in press
Fuchs, B. 1991, in Dynamics of Disc Galaxies, B. Sundelius, Göteborg, 359
Haga, M., Iye, M. 1994, MNRAS, 271, 427
Kennicutt, R. C., Tamblyn, P., & Congdon, C. W. 1994, ApJ, 435, 22
Pichon, C., Cannon, R. C. 1998, MNRAS, in press
Sellwood, J. A., Carlberg, R. G. 1984, ApJ, 282, 61
Toomre A. 1981, in The Structure and Evolution of Normal Galaxies, S. M. Fall & D. Lynden-Bell, Cambridge Univ. Press, 111
|
no-problem/9812/cond-mat9812260.html
|
ar5iv
|
text
|
# REFERENCES
Phase coherence and “fragmented” Bose condensates
D.S. Rokhsar
Department of Physics, University of California, Berkeley, CA 94720-7300, USA and Physical Biosciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
submitted November 27, 1998; PACS 03.75 Fi
We show that a “fragmented Bose condensate” in which two or more distinct single-particle states are macroscopically occupied by the same species of boson is inherently unstable to the formation of a conventional Bose condensate whose macroscopically occupied state is a linear combination of the “fragments” with definite relative phases. A related analysis shows that a reproducible relative phase develops when two initially decoupled condensates are placed in contact.
The essential property of a Bose condensate is the existence of a single macroscopically large eigenvalue of its one-body density matrix. This property reveals the existence of long-range phase coherence across the fluid, which leads to the two-fluid description of Bose condensates and implies the quantization of circulation in superfluid flow. Hypothetical Bose states without this property were discussed long ago by Nozieres and St. James , and have recently received renewed theoretical attention in the case of trapped bosons. In particular, Wilkin et al. found that the azimuthally symmetric rotating states of a dilute Bose gas with attractive interactions are “fragmented” Bose condensates in which many single-particle states (of differing angular momenta) are macroscopically occupied. Such multi-condensate Bose fluids would be expected to have unusual properties.
A related set of issues stems from a gedankenexperiment proposed by Anderson, in which two previously isolated Bose condensates are brought into particle-exchanging contact. Before contact is initiated, each isolated subsystem has its own macroscopically occupied state. The two isolated Bose subsystems considered together therefore provide a simple example of a system with a “fragmented” condensate. What happens when the condensates are allowed to exchange particles – will the phase difference be random from experiment to experiment, or be reproducibly zero?
Here we show that when fragmented condensates overlap, they are inherently unstable to the formation of a single, conventional condensate of well-defined relative phase. The essential idea is that even weak Josephson coupling between the fragments rapidly generates phase coherence, modifying the density matrix deterministically to give a unique macroscopically occupied single particle state. The proof proceeds by first assuming the existence of a Bose state with multiple condensates, and then explicitly constructing a lower energy state. The argument is very general, and applies to strongly as well as weakly interacting systems, and for either sign of the scattering length. As a corollary, we show that the relative phase that emerges when two previously uncoupled condensates are placed in contact is reproducible, and is determined by the physics of coherent particle exchange between them.
Multiple condensates. Consider a fluid of $`N`$ identical Bose particles. Let $`|0`$ be a many-body state that is presumed to be both (a) the ground state of a many-body Hamiltonian $``$ and (b) a “fragmented condensate,” i.e., a state whose reduced one-body density matrix
$$\rho (𝐫,𝐫^{})\widehat{\psi }(𝐫^{})^{}\widehat{\psi }(𝐫)$$
(1)
has two or more macroscopic eigenvalues $`N_i`$, with corresponding eigenfunctions $`\varphi _i(𝐫)`$. The $`N_i`$ have the physical interpretation of the mean occupancy of the single-particle state $`\varphi _i`$ in the many-body state $`|0`$. In general the macroscopic $`N_i`$ are not integers, and they need not add up to the total particle number $`N`$. For simplicity, we will assume that ony two such macroscopic eigenvalues exist, but our arguments are easily generalized to any finite number.
We define the condensate annihilation operators
$$b_i𝑑𝐫\varphi _i(𝐫)\widehat{\psi }(𝐫)$$
(2)
where $`\widehat{\psi }(𝐫)`$ is the particle annihilation operator at position $`𝐫`$. Then in the fragmented state,
$$[\widehat{\rho }_0]_{ij}=0|b_j^{}b_i|0=N_i\delta _{ij},$$
(3)
which follows from the assumption that $`\varphi _i`$ are eigenstates of $`\widehat{\rho }`$. For $`ij`$ the density matrix element vanishes, indicating the absence of coherence between the condensates.
A family of states. We now construct a family of $`N`$-particle states labeled by the integers $`q=0,\pm 1,\pm 2,\mathrm{}`$ in which $`q`$ particles are transferred between the two condensates. For positive $`q`$, the transfer is from 1 to 2; for negative $`q`$, the transfer is from 2 to 1; the $`q=0`$ state is simply the original fragmented state $`|0`$. These states can be formed by repeated applications of the operator
$$B^{}\frac{b_2^{}b_1}{\sqrt{N_2N_1}}$$
(4)
for positive $`q`$, and its Hermitian conjugate $`B`$ for negative $`q`$. The factors in the denominator make the states
$`|q`$ $``$ $`(B^{})^q|0q0`$ (5)
$`|q`$ $``$ $`(B)^{|q|}|0q<0`$ (6)
normalized for $`qN_i`$, which will be the case of interest below. The operator $`\widehat{Q}`$ can then be defined as $`q|qq|`$. Our strategy will be to construct a linear combination of the states $`|q`$ that is lower in energy than the putative ground state $`|0`$.
An effective tight-binding model. We consider a many-body Hamiltonian $``$ given by the sum of a one-body part $`\widehat{h}_o`$ – kinetic energy plus a single-particle potential $`V(𝐫)`$ – and a two-body interaction $`U(𝐫𝐫^{})`$, which for notational simplicity we will take to be a $`\delta `$-function pseudopotential of strength $`U`$. (Our results are easily extended to more general interactions.) Within the subspace of states $`\{|q\}`$, $``$ can be written
$$=[t_1B^{}+t_1^{}B][t_2(B^{})^2+t_2^{}B^2]+\frac{1}{2}\kappa Q^2,$$
(7)
which has the form of a one-dimensional tight-binding model with nearest and next-nearest neighbor hopping matrix elements $`t_1`$ and $`t_2`$, and a harmonic potential of spring constant $`\kappa `$.
The generally complex hopping matrix elements are given by
$`t_1`$ $`=`$ $`\sqrt{N_1N_2}{\displaystyle 𝑑𝐫\varphi _2^{}(𝐫)\widehat{h}_{\mathrm{eff}}\varphi _1(𝐫)},`$ (8)
$`t_2`$ $`=`$ $`{\displaystyle \frac{U}{2}}N_1N_2{\displaystyle 𝑑𝐫\varphi _2^{}(𝐫)\varphi _2^{}(𝐫)\varphi _1(𝐫)\varphi _1(𝐫)},`$ (9)
where $`\widehat{h}_{\mathrm{eff}}\widehat{h}_o+U_iN_i|\varphi _i(𝐫)|^2`$ is an effective one-body Hamiltonian that includes mean field interactions.
If the condensates $`\varphi _i(𝐫)`$ extend over the same $`d`$-dimensional volume $`R^d`$, where $`R`$ is a characteristic length scale, then the matrix elements of eq. (89) scale as $`t_1Nϵ_1`$ and $`t_2Nϵ_2`$, where the $`ϵ_i`$ are functions of the mean density $`N/R^d`$ with the units of energy per particle. Typically, they will be of the same order of magnitude as the chemical potential. We consider below the special case in which the matrix elements $`t_1`$ and $`t_2`$ vanish for symmetry reasons.
The hopping terms of the Hamiltonian (7) give rise to a tight-binding band with dispersion
$$E(k)=[t_1e^{ik}+t_1^{}e^{ik}][t_2e^{i2k}+t_2^{}e^{i2k}],$$
(10)
which can be expanded about its minimum to yield
$$E(k)=\frac{TN}{2}+\frac{T^{}N}{2}(k\chi )^2+\mathrm{}$$
(11)
where the real numbers $`T`$ and $`T^{}`$ depend on the $`ϵ_i`$, and $`\chi `$ can be found by minimizing eq. (10). The effective mass for this band is given by $`\mathrm{}^2/M_{\mathrm{eff}}=T^{}N`$.
The “spring constant” $`\kappa `$ in eq. (7) can be found quite generally from the assumption that the condensates are compressible, which is expected on general grounds for a Bose fluid. Then the energy of a state with $`q_i`$ particles added to condensate $`i`$ (by repeated application of $`b_i^{}`$) will be a smooth function that can be Taylor expanded for small $`q_iN_i`$, so that
$$q||q=E_0+q(\mu _1\mu _2)+\frac{\kappa }{2}q^2+\mathrm{}$$
(12)
where $`\mu _iE/N_i`$ is the chemical potential of the $`i`$-th fragment and
$$\kappa =\frac{K}{N}[\frac{^2E}{N_1^2}2\frac{^2E}{N_1N_2}+\frac{^2E}{N_2^2}].$$
(13)
For condensates $`\varphi _i`$ of volume $`R^d`$, $`K`$ is a number of order $`U(N/R^d)`$ that is again comparable to the chemical potential, and hence $`T`$ and $`T^{}`$. Since by assumption $`q=0`$ is the ground state the chemical potentials must be equal and the linear term in $`q`$ vanishes. For stability, $`\kappa `$ must be positive.
The tight-binding model (7) for two coupled condensates is related by a canonically transformation to the familiar pendulum description of a Josephson junction, which treats the relative phase as a periodic “coordinate” and the number fluctuation as its corresponding “momentum.” The present analysis instead takes the number fluctuation as a discrete coordinate, with relative phase corresponding to the conjugate (crystal) momentum. The two approaches are completely equivalent, and provide complementary insights to the behavior of coupled condensates.
A lower energy state. Within a continuum approximation, the tight-binding Hamiltonian (7) describes a particle of mass $`M_{\mathrm{eff}}`$ in a harmonic potential of spring constant $`\kappa `$. The frequency of this oscillator is given by $`M_{\mathrm{eff}}\omega _{\mathrm{eff}}^2=\kappa `$, so that $`\mathrm{}\omega _{\mathrm{eff}}=\sqrt{T^{}K}`$. The ground state has the form
$$|G=\underset{q}{}\frac{e^{q^2/2\sigma ^2}}{\pi ^{1/4}\sigma ^{1/2}}e^{i\chi q}|q,$$
(14)
with a characteristic spread in $`q`$
$$\sigma =(\mathrm{}/M_{\mathrm{eff}}\omega _{\mathrm{eff}})^{1/2}=(T^{}/K)^{1/4}N^{1/2}.$$
(15)
Since $`\sigma N^{1/2}`$, we are justified a posteriori in both the continuum approximation and our decision to neglect terms of order $`q^3`$ (and higher) in $`E(q)`$.
The energy of $`|G`$ relative to $`|0`$ is
$$E_GE_0=\frac{TN}{2}+\frac{(T^{}K)^{1/2}}{2},$$
(16)
where the first term on the right is the delocalization energy of the bottom of the tight-binding band, and the second is the zero-point energy of a harmonic oscillator with energy spacing $`\mathrm{}\omega _{\mathrm{eff}}`$. Evidently the admixture of correlated condensate fluctuations $`|G`$ is lower in energy than the fragmented state $`|q=0`$ if $`N>T^{}K/(T^2)`$, where the right hand side is of order unity. Thus even for extremely weak off-diagonal couplings, a fragmented condensate will not be the ground state for a macroscopic number of particles.
Density matrix revisited. What is the nature of the true ground state $`|G`$? It is easy to show that $`|G`$ is, in fact, a conventional Bose condensate whose density matrix has a unique macroscopic eigenvalue. Since $`qN_i`$, the diagonal matrix elements of $`\widehat{\rho }`$ are unchanged
$$[\widehat{\rho }_G]_{ii}=G|b_i^{}b_i|G=N_i.$$
(17)
The off-diagonal matrix elements, however, are now macroscopic:
$$[\widehat{\rho }_G]_{12}=G|b_2^{}b_1|G=ge^{i\chi }\sqrt{N_1N_2}$$
(18)
where
$$g=e^{1/4\sigma ^2}1\frac{A}{N},$$
(19)
where $`A`$ is a number of order unity.
Rediagonalizing the density matrix, we find a single macroscopically occupied eigenstate
$$\varphi _\mathrm{g}(𝐫)=c_1\varphi _1(𝐫)+e^{i\chi }c_2\varphi _2(𝐫),$$
(20)
where $`c_i\sqrt{N_i/[N_1+N_2]}.`$ The occupation of this state is $`N_g=N_1+N_2B`$, where $`B`$ is proportional to $`A`$, and is also a number of order unity. The orthogonal combination $`\varphi _{}(𝐫)=c_2\varphi _1(𝐫)e^{i\chi }c_1\varphi _2(𝐫)`$ has eigenvalue $`B`$ of order unity. We conclude that fragmented condensates are unstable towards the formation of conventional Bose condensate with a unique macroscopically occupied state.
Spontaneous symmetry breaking. As noted above, the matrix elements $`t_i`$ may vanish for symmetry reasons. As a specific example, consider a rotationally invariant system with two fragments $`\varphi _1`$ and $`\varphi _2`$ that have angular momentum projections $`m_1`$, $`m_2`$, respectively. Then the $`t_i`$ will be zero. This situation arises in the context of rotating Bose gases trapped in an azimuthally symmetric potential. Under these circumstances, we must ask if the fragmented condensate remains stable in the presence of a small symmetry-breaking perturbation, which introduces a nonzero $`t_1`$.
We have seen, however, that even for $`T1/N^2`$ a fragmented condensate will be unstable to the formation of a conventional condensate (20) that is a superposition of the fragments, with a relative phase $`\chi `$ that is determined by the details of the perturbing potential through eq. (8). The resulting state is not an eigenstate of angular momentum, and therefore has an asymmetric density and current distribution. The fact that an order $`1/N`$ perturbation can reduce the symmetry of the ground state indicates that fragmented condensates will spontaneously break whatever symmetry (rotation, in the case of ref. ) permitted fragmentation in the first place.
This result can be understood in another way by considering the susceptibility of a fragmented condensate to a perturbation which couples to the inter-fragment current density $`\widehat{J}`$ $`i[t_1B^{}t_1^{}B]`$. Since $`(B^{})^2=`$ $`B^2=0`$ in the $`q=0`$ state, while $`B^{}B=BB^{}=1`$ as an operator identity, the mean-square current fluctuation $`J^2`$ $`|t_1|^2`$. A weak one-body potential that allows scattering between $`\varphi _1`$ and $`\varphi _2`$, however, will typically introduce a $`t_1N`$. Fragmented condensates are thus highly susceptible to perturbations that permit coherent particle transfer between the fragments.
Gauge covariance. The phase factor $`e^{i\chi }`$ which enters into the superposition (20) is not arbitrary, but is determined by the phase of the off-diagonal density matrix element (18), which in turn is given by the phase of the Hamiltonian matrix elements $`t_1`$ and $`t_2`$ that scatter particles between the two condensates. We can easily confirm that this result is properly gauge covariant.
The single-particle states $`\varphi _i`$ are eigenstates of the (fragmented) one-body density matrix $`\widehat{\rho }_0`$, and are only defined to within arbitrary phase factors. All physical observables must therefore be unchanged if each $`\varphi _i`$ is multiplied by $`e^{i\alpha _i}`$. It is easy to follow these phases through our calculations. The hopping matrix elements transform as $`t_1t_1e^{i(\alpha _1\alpha _2)}`$ and $`t_2t_2e^{2i(\alpha _1\alpha _2)}`$, so that $`\chi \chi +\alpha _1\alpha _2`$. At the end of the day, the unique condensate (20) becomes
$`\varphi _\mathrm{g}(𝐫)`$ $``$ $`c_1[e^{i\alpha _1}\varphi _1(𝐫)]+e^{i(\chi +\alpha _1\alpha _2)}[c_2e^{i\alpha _2}\varphi _2(𝐫)]`$ (21)
$`=`$ $`e^{i\alpha _1}\varphi _\mathrm{g}(𝐫).`$ (22)
The arbitrary phases $`\alpha _i`$ are only reflected in the overall phase of $`\varphi _\mathrm{g}`$ (which itself is only defined up to an overall phase). The relative phase of $`\varphi _\mathrm{g}(𝐫)`$ between any two points $`𝐫_1`$ and $`𝐫_2`$ are invariants. The superfluid velocity, which is proportional to the gradient of the phase, is also invariant. Thus the phase of the off-diagonal one-body density matrix element (proportional to $`B`$) is a gauge-covariant representation of the relative phase between two condensates.
Relative phase of previously isolated condensates. We can use a similar approach to analyse Anderson’s gedankenexperiment in which two independently created condensates $`\varphi _L`$ and $`\varphi _R`$ are placed in contact. Before contact is established, the many-body state of the combined system is a simple product of states – i.e., a “fragmented condensate” – that we can denote $`|q=0`$. Since the off-diagonal element $`q=0|b_L^{}b_R|q=0`$ of the density matrix is zero, the relative phase between the two condensates is the phase of the complex number zero, which is ill-defined.
For gedankenpurposes, let us assume that the two independent condensates are identical, with equal chemical potentials. (Our discussion is easily generalized.) As shown in the previous section, we may without loss of generality choose both $`\varphi _L`$ and $`\varphi _R`$ to be real. The time evolution following the initiation of contact is governed by a Hamiltonian of the form (7), plus damping, with $`t_2<<t_1`$. By time-reversal symmetry, the $`t_i`$ are real, and $`\chi =0`$ or $`\pi `$; for definiteness we assume conventional tunneling contact, with $`\chi =0`$. In contrast to the case of spatially overlapping fragments, however, the Josephson coupling $`t_{\mathrm{eff}}`$ between two reservoirs in contact will be proportional to their contact area, and exponentially small in the energy barrier between them.
Except for the discreteness of $`q`$, this is a familiar problem. At long times, the system ends up in a Gaussian ground state, which we have seen is simply a conventional condensate with uniform phase. Relative to this “vacuum,” the initial state $`q=0`$ is “squeezed,” with a width $`\delta q1`$ in position and $`\delta p\pi `$ in the conjugate momentum. The establishment of phase coherence between two previously isolated condensates is therefore equivalent to the decay of a “squeezed vacuum” in quantum optics. For our purposes, the important feature is that due to the $`\pm q`$ symmetry of the initial state and the Hamiltonian, $`B^{}=B`$ for all time. The off-diagonal density matrix element is then always real, so as soon as the relative phase becomes defined, it is zero.
The experiment of Andrews et al. in which two nominally identical, initially isolated condensates are allowed to expand and overlap can also be cast in this form. If the condensates are initially converging symmetrically, then $`\varphi _R(z)=\varphi _L(z)^{}`$, and we again find that the transfer matrix element that scatters particle between the condensates is real. This result is consistent with Naraschewski et al.’s analysis of the interference fringes seen in the MIT experiment, which requires $`\chi =0`$.
Random vs. reproducible phases. Javanainen and Yoo and Castin and Dalibard have discussed protocols in which a series of measurements is performed on the states of individual particles removed from two isolated condensates. Their elegant analyses show that the interference patterns that emerge are consistent with a well-defined phase difference between the two condensates. This phase difference, however, is arbitrary, and varies if the entire series of measurements are repeated. How can our calculation be reconciled with these results?
In the schemes of refs. and , each measurement of the series removes a particle from the two-condensate system in a defined superposition of the original condensates $`\varphi _i`$. The remaining particles are therefore left in a specific entangled many-body state that depends on the results of the measurements. In other words, a series of observations of the removed particles generates a specific perturbation $`\delta (t)`$ acting on the remaining particles. This perturbation does not conserve particle number, and in effect applies a time-dependent, gauge-symmetry-breaking “$`\eta `$-field” to the condensates.
As we have seen, the susceptibility of the state $`q=0`$ to such a perturbation is divergent with $`N`$. Thus it only takes a small perturbation (i.e., a few measurements) to sculpt the state of the remaining particles into a single condensate, whose relative phase depends deterministically on the $`\delta (t)`$, and hence on the specific sequence of observations performed on the removed particles. We have shown above that when the two condensates are placed in particle-exchanging contact their coupling also introduces a perturbation, which reproducibly enforces a relative phase of zero. An interesting question is the manner in which this effect competes with the measurement scenarios of refs. and .
Acknowledgements. I thank N. Wilkin, R. Smith, J. Gunn, D. Butts, M. Mitchell, J. Ho, J.C. Davis, D. Weiss, J. Garrison, S. Kivelson, and R. Chiao for many interesting conversations.
|
no-problem/9812/cond-mat9812127.html
|
ar5iv
|
text
|
# Highly Optimized Tolerance: Robustness and Power Laws in Complex Systems
## ACKNOWLEDGMENTS
This work was supported by the David and Lucile Packard Foundation, NSF Grants No. DMR-9212396 and DMR-9813752, and a DOD MURI grant for “Mathematical Infrastructure for Robust Virtual Engineering.”
|
no-problem/9812/astro-ph9812418.html
|
ar5iv
|
text
|
# On the Particle Heating and Acceleration in Black Hole Accretion Systems
## 1 Introduction
Figure 1 shows three (roughly) contemporaneous broad band high energy emission spectra from three galactic black hole candidates (GBHCs; Grove et al. 1998). Although it is conventional to interpret the soft black-body-like component below $`10`$ keV as coming from an optically thick Shakura-Sunyaev (SS) disk, the origin of the hard X-ray continuum (and its extension into soft X-rays during the low-hard state) is a constant source of debate. Extracting a physically sensible model through a maze of high quality spectral and timing data on these systems remains a great challenge.
Recently, there seems to be a renewed interest in understanding particle heating/acceleration in accretion disks. We attribute this to the observations of: possible $`>0.5`$ MeV emissions from Cyg X-1 and GRO J0422; the powerlaw component of GRO J1655 extending to at least 800 keV without a cutoff (Tomsick et al. 1998); and relativistic radio jets from sources like GRO J1655 and GRS 1915. Furthermore, the clearly laid-out physical requirements of ADAF models (which have enjoyed much success, see Narayan et al. 1998 for a review) also prompted further discussions on particle heating.
In this review we will mostly discuss a few models for the so-called low-hard state where the spectrum ($`\nu F_\nu `$) is peaking around 100-200 keV. We apologize for not able to cover all the models (see Liang 1998 for a recent extensive review). The powerlaw tail that seems to extend beyond $`500`$ keV during the soft-high state also begs explanation, though the total energy contained in this tail is perhaps $`<10\%`$ of the total emission, so we will place less emphasis on them. We will focus on the electron energization processes of these theoretical models. We will not discuss any detailed spectral and temporal analyses (see other articles in this volume). Even so, we quickly realized that writing on this topic is a very difficult task because we find many questions and confusions with no clear and definite answers.
## 2 Some Models for the Origin of Hard X-rays and Gamma-rays
In all the models discussed here, the physics of angular momentum transport (or “$`\alpha `$” viscosity) during accretion is not well understood. As a direct consequence, unfortunately, modeling energy dissipation in accretion disks has many ad hoc elements. Quite generally, the matter (surface density $`\mathrm{\Sigma }`$) in accretion disk is evolved as (taken from Papaloizou & Lin 1995)
$$\frac{\mathrm{\Sigma }}{t}\frac{1}{r}\frac{}{r}\left[F_1+F_2+F_3\right]S_\mathrm{\Sigma }=0$$
(1)
where $`F_1(\nu \mathrm{\Sigma }r^{1/2})/r`$ is the local viscous transport with viscosity $`\nu `$ (i.e., the standard $`\alpha `$disk viscosity or from MHD turbulence by Balbus & Hawley 1991, 1998); $`F_2S_\mathrm{\Sigma }J`$ is the advective loss with $`J`$ being the angular momentum carried by the source/sink ($`S_\mathrm{\Sigma }`$) material (i.e., magnetic flux and/or winds, Blandford & Payne 1982); $`F_3\mathrm{\Lambda }`$ is the external perturbation (i.e., tidal interactions).
Three models (or their variants) are usually employed for explaining the high energy emissions, namely, the SS model, the SLE model (Shapiro et al. 1976), and the ADAF model. All of them use the local viscous transport prescription (the $`F_1`$ term) and the energy is also dissipated locally at the disk. In SS model disk is optically thick and geometrically thin, and the plasma is also highly collisional. The heat deposited from transporting angular momentum is successfully radiated away so that disk remains thin ($`HR`$). In SLE and ADAF models, however, an inner, hot ($`T_e100`$ keV), optically thin ($`\tau 1`$) and two-temperature ($`T_iT_e`$) region is postulated. This region is then cooled via various radiation processes, such as thermal Compton scattering and Synchrotron.
The arguments for the existence of this hot, optically thin region might be summarized as follows: if local viscous energy dissipation only heats protons, and if there is only Coulomb coupling between electrons and protons, then when the energy input rate is high enough, the system will become unstable if the cooling via radiation is not quick enough, so the plasma has to expand and become optically thin. Here, we want to emphasize that the accreting plasma, during this transition from an optically thick, thin disk to an optically thin, quasi-spherical state, has also changed from highly collisional to essentially collisionless. This brings up several immediate questions which are related to the above “if”.
## 3 Open Questions
### 3.1 Will local viscous energy dissipation only heat protons?
Bisnovatyi-Kogan & Lovelace (1997) first discussed this issue and argued that dissipation in such a magnetized collisionless plasma predominantly heats the electrons owing to reconnection of the random magnetic field. On the other hand, Quataert (1998) and Gruzinov (1998) have argued that conditions for ADAF could be true in the high $`\beta =P_{\mathrm{plasma}}/P_{\mathrm{magnetic}}5`$ limit by calculating the linear damping rates of short wavelength modes in a hot (but nonrelativistic) plasma. in an (implicit) almost uniform magnetic field. Note that even though MHD turbulence phenomenology was used in both papers, the damping rates are valid in the linear regime for plasma waves only (see below for further discussion). But these calculations perhaps are not answering the question of how to form the optically thin region in the first place because they are damping rates in the collisionless limit. Instead, one perhaps might first evaluate the energy dissipation processes (with an understanding of $`\alpha `$ viscosity) in the collisional limit which is the physical state initially. These collisions ensure thermal electron and proton distributions and efficient energy exchange between them, especially at the so-called transition radius in ADAF ($`10^310^4r_s`$).
If one uses Balbus-Hawley instability (see also Velikov 1959 and Chandrasekhar 1981) as the origin of the viscosity in the disk, then the gravitational energy is mostly released in large scale (longest wavelength of the magnetic field changes) and this energy will amplify the field first (instead of going into heating the particles). Once the nonlinear saturation is reached (say with magnetic energy density being $`10\%`$ of the kinetic energy density of the shear flow), we are actually faced with two possibilities, namely, whether the magnetic fields will be expelled (or escape) from the disk, or they will have to dissipate locally in the disk. Bisnovatyi-Kogan & Lovelace (1997) argued for the second possibility (but see Blackman 1998). Since we know that both the fluid and magnetic Reynolds numbers are exceedingly large in these flows, any “classical” viscous and ohmic dissipations will happen on timescales longer than the age of the universe, thus efficient magnetic reconnection has been sought as the primary candidate for energy dissipation in the disk. They further argued that current-driven instabilities in this turbulent plasma will give rise to large local $`E_{}`$ which mostly accelerate electrons. Thus, up to half of the magnetic energy input goes directly to electrons and is subsequently radiated away, and the disk will always stay thin and optically thick. The uncertainties in these arguments are nevertheless quite large since we don’t fully understand MHD turbulence, let alone its dissipation via kinetic effects. For example, it is unclear whether such reconnection sites are populated throughout the plasma so that most fluid elements encounter such regions. There has been some detailed numerical simulations with magnetic Reynolds number up to 1000 (Ambrosiano et al. 1988) in which test particles are observed to get accelerated by the induced small scale electric fields associated with reconnection sites in turbulent MHD flows. If indeed the magnetic energy dissipation is through accelerating particles by the induced electric fields (this is a big if), since electrons are the current carriers, it is hard to imagine that protons receive most of the energy.
### 3.2 Is there any collective process that could ensure efficient energy exchange between protons and electrons besides Coulomb?
Putting aside the uncertainties discussed above, if there is indeed an optically thin, hot, two-temperature plasma region, a pertinent question is how much energy electrons can get. This question is, unfortunately, ill-fated again because we do not know how to formulate the problem. Another way to look at it is how to identify the free energy, since most plasma instabilities require a good knowledge of the free energy as determined by the system configuration. For example, is there a relative drift between protons and electrons and can fast electrons be regarded as a beam to an Maxwellian proton core distribution? Is there temperature anisotropy parallel and perpendicular to background magnetic fields, etc.? Begelman & Chiueh (1988) have studied some plasma instabilities in detail and found plausible ways of transferring energy from ions to electrons, under the conditions that a substantial level of MHD turbulence will give a large enough proton density gradient (or curvature drifts) so that proton drift velocity can be large enough to drive certain modes unstable. The fluctuating electric field parallel to the magnetic field will then accelerate electrons. The applicability of this instability is again hampered by our lack of knowledge of the presumed MHD turbulence. Narayan & Yi (1995) argued that this mechanism does not work well in ADAF.
A conceptual difficulty is that the typical modes excited by protons (having most of the energy) are below the proton gyrofrequency $`\mathrm{\Omega }_{ci}`$. This makes resonance with the electrons difficult. But a possible avenue is to have protons excite (almost) perpendicular modes (i.e., high $`k_{}`$ and very small $`k_{}`$). Then the resonant conditions for electrons to resonate with these waves are easier to satisfy. More work is needed to explore these possibilities.
### 3.3 Could accretion disk have a magnetically dominated, hot corona like our Sun?
The formation of a “structured corona” was first proposed by Galeev et al. (1979). In this model a radial quadrupole field is wound up by differential rotation into an enhanced toroidal field. Then the helicity of the presumed convective “turbulence” converts a fraction of the toroidal flux back into poloidal field and hence produces an exponentiating dynamo that saturates by back reaction. This is the classical $`\alpha \mathrm{\Omega }`$ dynamo although not identified as same in the paper. Furthermore the saturation or back reaction limit of this disk dynamo is assumed to be the random loops of flux characteristic of the solar surface.
One important step in the above model is the requirement of vertical (thermal) convection in the $`\{R,z\}`$ plane. The convective motion may be driven by heat released at or near the mid-plane. Lin et al. (1993) have shown that under specific conditions of opacity and equation of state that convective instability should occur both linearly and nonlinearly, thus leading to large amplitude cells. However, the convective cells are highly constrained radially. The problems of restrictive initial conditions and the restrictive cell geometry leads one to conclude that this is not the universal mechanism needed to explain accretion disks. Colgate & Petschek (1986) showed that to drive convective cells whose displacement radially is of the order of the disk height $`h`$, (unrealistic) efficiency of the energy flow (a Carnot cycle of $`100\%`$ efficiency) is necessary to drive these convective eddies, and the cells created are also highly restrictive, tall but narrow radially (i.e., similar as Lin et al.). Thus the existence of strong convective turbulence is doubtful. The result of this lack of convective turbulence with rising plumes is to negate the origin of the helicity invoked in the structured corona model.
## 4 Electron Energization
Besides the possible role of magnetic reconnection in accelerating electrons which is observed in the solar corona (Tsuneta 1996), there are more standard processes which involve wave-particle interactions (see Kuijpers and Melrose 1996). Shock acceleration is not considered here. We give a quick review of the electron energization by plasma waves and turbulence.
### 4.1 Particle heating/acceleration – linear and quasilinear theory
Linear Vlasov equation is usually used to describe the collisionless plasma, which is a good approximation of astrophysical plasmas. Linearization of the Vlasov equation yields various dispersion relations $`\omega =\omega (𝐤)`$ which describe how the system will respond to small electrostatic and electromagnetic perturbations. Since the field energy of low frequency fluctuations (i.e., $`\omega <\mathrm{\Omega }_{\mathrm{cp}}`$) is predominantly magnetic, particles generally experience strong pitch-angle scattering before they can be energized. Of fundamental importance is the wave-particle resonance, that is, given an electromagnetic fluctuation of frequency $`\omega `$ and wavevector $`𝐤`$, a charged particle ($`q,m`$) is considered to be in resonance with this fluctuation when
$$\omega k_{}v_{}\mathrm{}\mathrm{\Omega }_0/\gamma =0,\mathrm{}=0,\pm 1,\pm 2,\mathrm{}$$
$`(1)`$
where the nonrelativistic gyrofrequency $`\mathrm{\Omega }_0=|q|B_0/mc`$, and $`v_{}`$ and $`\gamma `$ are the particle’s parallel velocity and Lorentz factor, respectively. When the harmonic number $`\mathrm{}=0`$, the resonance is referred to as the Landau or Cherenkov resonance, and implies that the particle speed along the magnetic field matches the speed of the parallel wave electric or magnetic field. If $`|\mathrm{}|>0`$, the process is called gyroresonance, and there is a matching between the wave transverse electric field and the cyclotron motion of the particle. The sign of $`\mathrm{}`$ depends upon the transverse polarization of the wave and the sign of $`q`$: if the transverse wave electric field and the particle rotate in the same sense about $`B_0`$ in the plasma frame, then $`\mathrm{}`$ is positive. In most settings, only $`\mathrm{}=\pm 1`$ is of importance.
The key quantities are the plasma $`\beta =n(T_i+T_e)/(B^2/8\pi )`$ factor and the temperature ratio $`T_e/T_i`$. Furthermore, we have:
$``$ Linear theory. The linear theory of plasma waves and instabilities is often reduced to a linear dispersion equation with a complex $`\omega `$, whose imaginary part gives the growth or damping of certain modes. Recent studies by Gruzinov (1998) and Quataert (1998) belong to this case. The usual candidates for wave-particle resonances are: for $`\mathrm{}=0`$, the transit time damping (TTD) for particle with oblique fast magnetosonic waves and Landau damping (LD) with kinetic Alfven waves; for $`|\mathrm{}|1`$, gyroresonances between the proton/Alfven wave and the electron/whistler wave.
$``$ Quasilinear theory. A detailed physical understanding of pitch-angle scattering and stochastic acceleration is beautifully presented in Karimabadi et al. (1992), using nonlinear orbit theory with the Hamiltonian formalism. In the presence of a continuum of plasma waves, the number of resonances between the particle and waves is greatly increased to a point that the trapping width associated with one particular resonance can overlap with neighboring resonances, thus allowing particles “jump” from one resonance to another. As particles sample different resonances, they gain energy in a “ladder-climbing” fashion. Hence the description stochastic acceleration. This approach has been adopted in several studies on electron acceleration by fast-mode waves and whistler waves in accretion disk (Li et al. 1996, Li & Miller 1997). We typically find that the electron distribution is hybrid with a nonthermal tail, which is responsible for the production of $`>500`$ keV emissions in several GBHCs. We have built a computer code which solves 3 coupled, time-dependent kinetic equations for particles, photons and waves, respectively. Namely,
$$\frac{N_e}{t}=\frac{}{E}\left\{\left[\frac{dE}{dt}+\left(\frac{dE}{dt}\right)_{\mathrm{loss}}\right]N_e\right\}+\frac{1}{2}\frac{^2}{E^2}\left[(D+D_c)N_e\right]$$
(2)
$$\frac{W_\mathrm{T}}{t}=\frac{}{k}\left[k^2D\frac{}{k}\left(k^2W_\mathrm{T}\right)\right]\gamma W_\mathrm{T}+Q_\mathrm{W}\delta (kk_0)$$
(3)
$$\frac{n_{\mathrm{ph}}(\epsilon )}{t}=n_{\mathrm{ph}}(\epsilon )𝑑EN_e(E)R(\epsilon ,E)+$$
(4)
$$𝑑\epsilon ^{}𝑑EP(\epsilon ;\epsilon ^{},E)n_{\mathrm{ph}}(\epsilon ^{})N_e(E)+\dot{n}_{\mathrm{ext}}(\epsilon )+\dot{n}_{\mathrm{emis}}(\epsilon )\dot{n}_{\mathrm{abs}}(\epsilon )\frac{n_{\mathrm{ph}}(ϵ)}{t_{\mathrm{esc}}}.$$
The particle distribution can be arbitrary. This allows us to determine from all the interactions whether the distribution is thermal or nonthermal. Pair production is not included so far. The Coulomb terms are also implemented for arbitrary particle distributions. Accurate Compton scattering is treated as a scattering matrix (Coppi 1992) with the full cross section. The Cyclo-Syn. process is calculated according to Robinson & Melrose (1984) which enters both as a cooling and heating term (Ghisellini et al. 1988). Syn-self absorption is also included. The radiation part of the kinetic code is tested against Monte Carlo simulations (Kusunose, Li, & Coppi 1998) and is found to be very good for $`\tau 3`$ and for both thermal and nonthermal electron distributions.
Figure 2 shows the time evolution of particles (upper panels), waves (middle) and photons (lower panels) from the start of continuous wave injection until the steady state is reached. There are 20 curves in each plot which are from $`t=010\tau _{dyn}`$, where $`\tau _{dyn}=R/c1.5\times 10^3`$ sec. These runs are made with application to optically thin environment in mind. The plasma density $`n`$ is varied from $`\tau =0.11`$. At early times, the particle distribution softens first as shown in upper panels, due to that waves have not fully cascaded (i.e., small $`k`$ as shown in middle panel), and losses dominate at high energies. As waves cascade over the inertial range, $`k`$ quickly grows to a level that acceleration overcomes all losses, electrons are then energized out of the thermal background and the nonthermal hard tail forms. The photon spectra indicates that gamma-rays can be produced when $`\tau <0.5`$. Furthermore, note that the nonthermal tails start to develop at $`E/m_ec^20.13`$ (corresponding to $`v_\mathrm{A}/c=0.46`$), this nicely confirms the fact that only particles with $`v>v_\mathrm{A}`$ can be accelerated.
### 4.2 MHD turbulence, are they an ensemble of waves?
The above described calculations, both linear and quasilinear, can be broadly regarded as “dissipation” in a general MHD turbulence theory. Finding a dynamical model that might adequately describe the evolution of magnetic fluctuations (such as equation (3) above) is at best phenomenological. In the dissipation range, the physics of the couplings that connect fluid and kinetic scales is not understood at all.
A critical assumption that is employed in all the kinetic calculations is that the magnetic fluctuations that cascade from large scales to small scales could be regarded as an ensemble of kinetic waves with a well-defined dispersion relation to describe them. This view is by no means proven, though it allows us to get an estimate of the particle heating rate since the kinetic theory is significantly more advanced (see an application of such an approach to the interplanetary magnetic field dissipation range, Leamon et al. 1998). On the other hand, the dynamics of MHD turbulence has been studied using statistical theories and simulations (e.g., Kraichnan & Montgomery 1980; Shebalin et al. 1983; Matthaeus & Lamkin 1986), and has never been convincingly presented or developed within a normal-mode, perturbation-type of framework. A further complication is that most (MHD) turbulence theory is based on the incompressible fluid model, how it will “carry-over” to compressible astrophysical flow is still an open question.
### 4.3 MHD turbulence Truncation
Recently, the assumption of a cascade to smaller scales of MHD turbulence is criticized in dynamo theory. It has been argued that both the more rapid folding of magnetic flux as well as the smaller energy density at small scale ensures rapid saturation or back reaction by the field stress, immobilizing the small scale fluid motions expected from the Kolmogorov spectrum. This will truncate the turbulent spectrum at the back reaction scale, initially the smallest and progressively reaching the largest. Since the energy input to the turbulence is assumed to be primarily at the largest scale, this leaves one with negligible power at the small scale (Cattaneo & Vainshtein 1991; Kulsrud & Anderson 1992; Gruzinov & Diamond 1994; Cattaneo 1994). The remaining largest scale is that of the disk itself.
In general all particle instabilities presumably leading to particle heating require large local gradients in some aspect of their phase space, i.e. temperature, density, and velocities, etc. Furthermore, all the free gravitational energy must flow through these gradients. This requirement, however, will not be met if the small scale turbulent motions are strongly damped by the back reaction of the field itself. We therefore look for a solution to this paradox in large scale magnetic structures.
## 5 A Sketch View
Here, we outline some plausible physical pictures about what might be happening in an accretion flow. Most these are ideas that have not been thoroughly investigated. It is also clear that there are obvious gaps which need to be filled with rigorous calculations.
### 5.1 Hydrodynamic transport and high-soft state
Many investigations have sought a linear instability deriving energy from the Keplerian flow to produce a growing mode leading, in the non-linear limit, to turbulence. The Papaloizou & Pringle instability (1984) seems to be the most studied instability but its relevance to Keplerian accretion disks has been questioned (Balbus & Hawley 1998).
Recently we have identified a linear instability in Keplerian disk leading to Rossby waves and presumably Rossby vortices in the nonlinear limit (Lovelace et al. 1998). This instability grows most effectively from a large radial gradient in entropy. It has the advantage that the nonlinear limit consists of co-planar, co-rotating vortices (Nelson et al. in preparation) that require only a radial gradient, not vertical gradient of entropy. The radial gradient, we believe, is astrophysically reasonable because all disks are presumably fed by matter at some outer radius by, for example, Roche-lobe overflow in low-mass X-ray binaries. If there is no angular momentum removal mechanism, the matter will accumulate until it builds up enough to trap heat, and variations in entropy would then render the onset of the above instability. We do not, however, expect this instability to lead to turbulence in the usual sense of convective turbulence. An ensemble of co-planar vortices does not lead to significant vertical flow as compared to the usual picture of convective turbulence where buoyant plumes would convect heat released at the the mid-plain to the disk surface.
We expect the angular momentum transport is done via nonlinear interactions of these vortices with the background flow, but this has to be addressed by extensive hydro simulations. The heat flow derived from the “viscosity” of the ensemble of Rossby vortices must be removed by radiation flow. We expect this not to be a problem because the radiation thickness of the disk, $`\tau `$, is small enough such that the effective diffusion velocity, $`v_{\mathrm{diff}}c/(3\tau )>>v_\varphi `$. Under these conditions, the disk solution will be essentially the same as the SS disk. Thus, this picture might be applied to the high-soft state of GBHCs. Relatively speaking, magnetic fields do not play a major role during this state but some nonthermal processes (such as a weak magnetic outflow) might be responsible for the powerlaw component from $`20`$ keV - 1 MeV.
### 5.2 Role of large scale magnetic fields and low-hard state
As pointed out by Blandford & Payne (1982), large scale magnetic fields can also be very effective in removing the disk angular momentum. These large scale magnetic outflows could be a hydromagnetic wind (Blandford & Payne 1982), or it might be a nearly force-free helix (Poynting flux) with very little matter as discussed by Lovelace et al. (1987, 1997).
The accreting plasma from, say a companion star, is likely to be magnetized. In the advection of this flux with the mass flow, there will necessarily be a convergence and strengthening of the field. In the region where an $`\alpha `$-viscosity prevails and the field acts as a passive marker of the flow, there will be both advection and diffusion. The diffusion radially outwards depends upon probably the same diffusion coefficient which allows the diffusion of angular momentum. Hence there will be a unique relationship between advection inwards and diffusion outwards leading to the relationship, $`B_zr^{3/2}`$ (see also Bisnovatyi-Kogan & Lovelace 1997 in which they argued $`B_rr^2`$).
If the initial field strength advected with the mass flow at the outer disk radius is large enough, then the field energy density could become comparable to the Keplerian stress at a certain radius. Bisnovatyi-Kogan & Lovelace (1997) argued that magnetic flux then has to be destroyed at the disk via reconnection. Alternatively, instead of destroying the flux, magnetic fields (presumably tied to the companion star) could be twisted such as they will remove the angular momentum of the flow and take away the released gravitational energy. So the energy dissipation (into radiation) might not be at the disk at all. The reconnection dissipation of the current supporting the torsion of the magnetic field will perhaps lead to the non-thermal emission of GBHCs. In fact, there is ample evidence in AGNs that perhaps most of the energy release is in the outflow/jet. Such a picture could also apply to GBHCs with the hard X-ray to gamma-ray emissions produced via nonthermal processes (such as Syn. or SSC) in the magnetized outflow away from the disk.
If the initial field strength advected with the mass flow at the outer disk radius is weaker, then the amplified magnetic field (by $`r^{3/2}`$) may never be greater than the Keplerian stress, thus is not the dominant channel of angular momentum transport. However, there will always be nonthermal energy release in the twisted magnetic field, which could be the powerlaw tail during the high-soft state.
###### Acknowledgements.
HL wish to thank the meeting organizers for their kind invitation and financial support. Part of the work on quasilinear wave-particle interactions was done together with J. Miller. We thank J. Finn and E. Liang for many useful discussions. HL gratefully acknowledges the support of an Oppenheimer Fellowship at LANL.
|
no-problem/9812/cond-mat9812335.html
|
ar5iv
|
text
|
# Normal-state magnetotransport in La1.905Ba0.095CuO4 single crystals
\[
## Abstract
The normal-state magnetotransport properties of La<sub>2-x</sub>Ba<sub>x</sub>CuO<sub>4</sub> single crystals with $`x`$=0.095 are measured; at this composition, a structural transition to a low-temperature tetragonal (LTT) phase occurs without suppression of superconductivity. None of the measured properties (in-plane and out-of-plane resistivity, magnetoresistance, and Hall coefficient) shows any sudden change at the LTT phase transition, indicating that the occurrence of the LTT phase does not necessarily cause an immediate change in the electronic state such as the charge-stripe stabilization.
\]
La<sub>2-x</sub>Ba<sub>x</sub>CuO<sub>4</sub> (LBCO) has been generally considered as a rather peculiar high-$`T_c`$ cuprate, not only because it is the first high-$`T_c`$ cuprate discovered by Bednorz and Müller , but also because $`T_c`$ of this compound is drastically suppressed in the composition rage near $`x`$=1/8 , known as the “1/8 anomaly”. Soon after the 1/8 anomaly was recognized, it was found that LBCO system shows a structural phase transition from a low-temperature orthorhombic (LTO) phase to a low-temperature tetragonal (LTT) phase in a rather wide range of $`x`$ around 1/8 . On the other hand, La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> (LSCO) system, which has the same crystal structure as LBCO, does not show a clear suppression of $`T_c`$ near 1/8; since there is no structural transition to the LTT phase in LSCO , it is generally believed that the occurrence of the LTT phase is responsible for the suppression of $`T_c`$ in LBCO.
There have been many experiments which tried to investigate the fundamental mechanism of the 1/8 anomaly. For example, Yoshida et al. studied the effect of partial substitution of Ba<sup>2+</sup> ion in LBCO with smaller divalent cations and found that such replacement of Ba<sup>2+</sup> leads to a suppression of the LTT structural transition and simultaneously to a recovery of the superconductivity . This result suggests that the LTT transition temperature and the strength of the $`T_c`$ suppression are closely tied to each other. Thus, Yoshida et al. concluded that the “1/8 anomaly” is caused by a Peierls-type mechanism with cooperative electronic and lattice instabilities. However, there are evidences which suggests that the occurrence of the LTT phase alone does not necessarily mean a destruction of superconductivity. Behaviors of La<sub>2-x-y</sub>Nd<sub>y</sub>Sr<sub>x</sub>CuO<sub>4</sub> (Nd-doped LSCO) system is one such example . In this system, while there is a clear structural phase transition to the LTT phase at 71 K and the superconductivity is almost completely destroyed for $`x`$=0.12, there remains a bulk superconductivity (with $`T_c`$=16 K) for $`x`$=0.20 even though the LTT phase transition temperature $`T_{LTT}`$ is higher (79 K) than that for $`x`$=0.12. Since high-quality single crystals are available for Nd-doped LSCO, the in-plane resistivity $`\rho _{ab}`$ and the out-of-plane resistivity $`\rho _c`$ have been studied in this system . For $`x`$=0.12, both $`\rho _{ab}`$ and $`\rho _c`$ show a clear jump at $`T_{LTT}`$, suggesting that the electronic state is changed upon the structural phase transition. It was found that $`\rho _c`$ shows a jump at $`T_{LTT}`$ even for $`x`$=0.20, indicating that the change in the electronic state persists to the $`x`$ value where the suppression of superconductivity is weak.
The known properties of LBCO is quite similar to that of Nd-doped LSCO; superconductivity is almost completely destroyed at $`x`$=1/8, bulk superconductivity remains for $`x`$$``$1/8, and the structural phase transition to the LTT phase occurs around 60 K which is almost independent of $`x`$. However, because of the difficulty in growing single crystals of LBCO, the anisotropic resistivity and the magnetotransport properties have not been well studied in LBCO with $`x`$ near 1/8 and thus the electronic states near $`x`$=1/8 is not well understood.
One of the composition of particular interest in LBCO is $`x`$=0.09; around this composition, the structural phase transition takes place but $`T_c`$ is not suppressed ($`T_c`$$``$30 K). In other words, the superconductivity for $`x`$=0.09 does not seem to be affected by the occurrence of the LTT phase. It is thus interesting to study whether the electronic system shows any change at the LTT phase transition for $`x`$=0.09, where the LTT phase does not affect superconductivity at all. This may clarify the importance (or unimportance) of the occurrence of the LTT phase to the electronic structure.
With the improvement in the crystal growth technique, high-quality single crystals of LBCO with $`x`$ near 1/8 have recently become available . In this paper, we report our detailed measurement of the anisotropic normal-state resistivity ($`\rho _{ab}`$ and $`\rho _c`$), in-plane magnetoresistance (MR), and the Hall coefficient $`R_H`$, of LBCO single crystals with $`x`$=0.095. As discussed above, this is the particular composition where $`T_c`$ is not suppressed despite the presence of the LTT phase. In fact, our $`x`$=0.095 crystals showed mid-point $`T_c`$ of 31 K, a very high value for LBCO. It was found that none of the measured transport properties shows any drastic change at the LTT phase transition, which strongly support the picture that the occurrence of the LTT phase does not necessarily change the electronic system.
The question whether the occurrence of the LTT phase alone can be responsible for the change in the electronic state is particularly intriguing in the light of the recently reported “stripe order” in the Nd-doped LSCO with $`x`$=0.12; using neutron diffraction techniques, Tranquada et al. observed elastic magnetic superlattice peaks of the type (1/2$`\pm ϵ`$,1/2,0) and charge-order peaks at (2$`\pm `$2$`ϵ`$,0,0), where $`ϵ`$=0.118 at low temperatures . Such an observation strongly suggests a presence of a one-dimensional charge order (“stripes”) which intervene in the antiferromagnetic spin order. Tranquada et al. proposed that the modulated antiferromagnetic order is pinned and stabilized in the LTT phase but not in the LTO phase, which is the reason why such static structure is not observed in pure LSCO. Following this picture, it can be inferred that the fundamental origin of the change in the electronic state in Nd-doped LSCO is the occurrence of the stripe phase and not the occurrence of the LTT phase itself. If so, it may be that the stripe order is not stabilized by the LTT phase transition in LBCO at $`x`$=0.095, which can be the reason for the coexistence of a “high” $`T_c`$ of 31 K with the LTT phase.
The single crystals of La<sub>1.905</sub>Ba<sub>0.095</sub>CuO<sub>4</sub> are grown using a traveling-solvent floating-zone (TSFZ) technique. Details of the crystals growth of LBCO are described elsewhere . After the crystallographic axis are determined, we cut the crystals to sufficiently small dimensions, typically 2 $`\times `$ 0.4 $`\times `$ 0.1 mm<sup>3</sup>, to ensure homogeneous Ba concentration in the crystal. The crystals are annealed in flowing-oxygen atmosphere at 650C for 24 hours to remove oxygen deficiencies. The actual Ba concentrations in the crystals are determined by the inductively-coupled plasma spectrometry (ICP) technique. A standard six-terminal method is used for the simultaneous $`\rho _{ab}`$ and $`R_H`$ measurement. Both the MR and $`R_H`$ data are taken in the sweeping magnetic field at fixed temperatures with an ac technique. The temperature is very carefully controlled and stabilized using both a capacitance sensor and a Cernox resistance sensor to avoid systematic temperature deviations with magnetic fields. The stability of the temperature during the MR and $`R_H`$ measurements is within 10 mK.
Figure 1 shows the temperature dependence of $`\rho _{ab}`$ and $`\rho _c`$. These data are measured in two different samples cut from the same rod. In both samples, the onset $`T_c`$ is 33 K and the resistivity becomes zero at 29 K. $`\rho _{ab}`$ is linear in $`T`$ down to 150 K and shows an upward deviation from the $`T`$-linear behavior at lower temperatures. A slight upturn in $`\rho _{ab}`$ is observed below 45 K, which is consistent with the data on polycrystalline samples around this composition . The extrapolated residual resistivity is negligibly small, which is similar to the behavior of high-quality LSCO crystals . In the case of Nd-doped LSCO, clear jumps in both $`\rho _{ab}`$ and $`\rho _c`$ have been observed at $`T_{LTT}`$ in underdoped samples ; however, there is no clear jump neither in $`\rho _{ab}`$ nor in $`\rho _c`$ in LBCO as shown in Fig. 1. Note that the structural phase transition to the LTT phase takes place at about 60 K for this $`x`$ value in LBCO . Therefore, contrary to the Nd-doped LSCO system, the resistivity data suggest that there is no sudden change in the electronic system in LBCO with $`x`$=0.095 at $`T_{LTT}`$. If we look at the temperature dependence of $`d\rho _{ab}/dT`$ (Fig. 1 inset, upper curve), there is a kink near $`T_{LTT}`$ ($``$60 K), which may suggest that the scattering of electrons gradually increases in the LTT phase. On the other hand, $`d\rho _c/dT`$ (Fig. 1 inset, lower curve) does not show any change at $`T_{LTT}`$, although there is a kink at lower temprature, about 52 K. It is intriguing that $`d\rho _{ab}/dT`$ and $`d\rho _c/dT`$ show kinks at different temperatures.
Figure 2 shows the temperature dependence of the in-plane $`R_H`$. Here, the magnetic field is applied along the $`c`$ axis and the current is along the $`ab`$ plane. For comparison, $`R_\mathrm{H}`$ data of LSCO ($`x`$=0.1) polycrystalline sample are also shown by a dashed line. $`R_H`$ of LBCO shows a peak at about 50 K, which is nearly the same temperature where $`\rho _{ab}`$ starts to show an upturn. There is no appreciable change in $`R_H`$ at $`T_{LTT}`$ (=60 K). The behavior and the absolute value of the $`R_H`$ of our LBCO crystal are quite similar to that of LSCO ($`x`$=0.1) system, which does not show an LTT phase transition.
One popular way of analyzing the normal-state transport properties of cuprates is to consider two scattering times, $`\tau _{tr}`$ and $`\tau _H`$ . $`\tau _{tr}(T)`$ and $`\tau _H(T)`$ are determined by the temperature dependence of $`\rho _{\mathrm{ab}}`$ and the cotangent of the Hall angle $`\theta _\mathrm{H}`$, respectively . Figure 3 shows $`\mathrm{cot}\theta _\mathrm{H}`$ (=$`\rho _{xx}`$/$`\rho _{xy}`$) at 10 T plotted against $`T^2`$. Since the Hall angle is proportional to the inverse of $`\tau _\mathrm{H}`$, it is clear from Fig. 3 that $`\tau _\mathrm{H}^1`$ obeys a $`T^2`$ law very well across $`T_{LTT}`$ down to 45 K. (The inset to Fig. 3 is a modified plot of the main panel to show directly the temperature region where the $`T^2`$ law holds.)
Figure 4 shows the result of the MR measurements of the LBCO crystal. We measured both the transverse MR ($`I`$ is within the $`ab`$ plane and $`H`$ is parallel to the $`c`$ axis) and the longitudinal MR ($`I`$ and $`H`$ are within the $`ab`$ plane and $`H`$ is parallel to $`I`$). The transverse MR consists of orbital and spin contributions, while the longitudinal MR comes only from the spin contribution. By comparing the two MRs, we can see that the spin contribution to the transverse MR is not large (about 30%). Although the longitudinal MR shows a smooth increase down to 40 K, the transverse MR shows a rather steep increase below 60 K, resulting in more than an order-of-magnitude difference between the two MRs at 40 K.
We tried to analyze whether this steep enhancement in the transverse MR can be understood by the superconducting fluctuation conductivity, whose contribution is large only for the transverse geometry. The fluctuation conductivity consists of Aslamasov-Larkin (AL) term and Maki-Thompson (MT) term; both terms comprises two contributions, the orbital contribution and the spin contribution . Kimura et al. have analyzed the MR in underdoped LSCO and concluded that the MT term is absent , which is actually expected for a $`d`$-wave superconductor . Thus we tried to estimate the fluctuation conductivity by considering only the AL term. The dashed line in the inset to Fig. 4 is the estimated AL orbital contribution, where the parameters are $`\xi _{\mathrm{ab}}(0)`$=30 Å and $`\xi _\mathrm{c}(0)`$=1 Å. (We just assumed these values as typical values.) The orbital part of MR, which is obtained by subtracting the longitudinal MR from the transverse MR, is also plotted in Fig. 4. Clearly, the increase of the transverse MR below 60 K can be accounted for by the superconducting fluctuations; therefore, it is not likely that the steep increase in the transverse MR is related to the occurrence of the LTT phase.
It has been proposed that the orbital MR in high-$`T_c`$ cuprates reflects the variance of a local Hall angle around the Fermi surface and therefore is proportional to the square of $`\theta _\mathrm{H}`$ , which is sometimes called “modified Kohler’s rule”. Figure 5 shows the temperature dependence of the orbital MR plotted together with $`a\times (\mathrm{cot}\theta _\mathrm{H})^2`$, where $`a`$ is a fitting parameter. \[Note that $`(\mathrm{cot}\theta _\mathrm{H})^2`$$``$$`\theta _\mathrm{H}^2`$ when $`\theta _\mathrm{H}`$ is small.\] The orbital MR does not scale so well to $`(\mathrm{cot}\theta _\mathrm{H})^2`$. Particularly, the orbital MR shows weaker temperature dependence above $``$200 K compared to $`(\mathrm{cot}\theta _\mathrm{H})^2`$. This might be an indication that the modified Kohler’s rule is not universally applicable to all high-$`T_c`$ cuprate. It would be interesting to study the applicability of the modified Kohler’s rule to the LBCO system in a wider carrier-concentration range.
The above results indicate altogether that the electronic system as inferred from $`\tau _\mathrm{H}`$ and $`\tau _{tr}`$ does not show any sudden change at the LTT phase transition, which seems to be different from the result of Nd-doped LSCO . In particular, the fact that $`\mathrm{cot}\theta _\mathrm{H}`$ shows a good $`T^2`$ behavior down to 45 K (Fig. 3) suggests that $`\tau _\mathrm{H}`$ is not influenced by the LTT phase. On the other hand, $`\tau _{tr}^1`$ seems to grow gradually with lowering temperature in the LTT phase, which causes a faster increase in resistivity. Based on these observations, we may conclude that the coexistence of a “high” $`T_c`$ of 31 K with the LTT phase is possible in LBCO at $`x`$=0.095 because the LTT phase transition does not immediately affect the electronic system. This, however, does not rule out the possibility that the electronic system is gradually changed in the LTT phase. The localization behavior in $`\rho _{ab}`$ below 45 K might actually be the result of some gradual change in the electronic state.
The authors would like to acknowledge Prof. S. Uchida for valuable discussions and for showing us unpublished results on polycrystalline LBCO.
|
no-problem/9812/cond-mat9812228.html
|
ar5iv
|
text
|
# Ab initio optical properties of Si(100)
## I Introduction
The optical spectroscopy of surfaces is becoming more and more popular as a non destructive and versatile tool of surface analysis. At variance with electron spectroscopies, it does not require ultra-high vacuum conditions, so that it can be used to monitor the growth of epilayers in Molecular Beam Epitaxy and Metal Organic Vapour Deposition techniques. However, the full potentiality of this kind of techniques can be exploited only by a strong interaction of experimental and theoretical work, due to the difficulty to interpret the spectra.
The dream of theorists is to compute realistic surface optical properties for useful comparison with experiments, given only the atomic species and positions. Until now most of the calculations were based on the semi-empirical tight-binding (TB) scheme, which, in spite of its successes, is not completely satisfactory from the theoretical point of view; in fact, it is not selfconsistent, and needs some experimental inputs to determine the electron bands and the optical properties. Only in the last years, due to the growing development of computational methods and computer performances, calculations of the optical properties within a first-principle approach started to appear. In principle, the correct procedure should be the calculation of the electronic states in a many body approach, including self-energy effects, and taking into account the electron-hole interaction and local-field effects to obtain the dielectric function. Actually, the state of the art is the determination of the single particle electron states using the Density Functional Theory (DFT) scheme in the Local Density Approximation.
The Si(100) surface is extensively studied both from the experimental and theoretical points of view, due to its technological importance. Now, it is generally accepted that its atomic structure is characterized by the presence of asymmetric buckled surface dimers along the $`[01\overline{1}]`$ direction, which implies at least a ($`2\times 1`$) reconstruction. Two other energetically competing (favored) reconstructions are possible, within the same top-atom bonding characteristic. The alternation of dimer buckling along the $`[011]`$ dimer rows leads to a p($`2\times 2`$) reconstruction, while the alternation of buckling angles also along the direction perpendicular to the rows leads to the c($`4\times 2`$) phase.
The actual geometry of the surface at room and low temperature is still controversial. The observed filled and empty surface states at room temperature were reasonably referred to c($`4\times 2`$) domains present in a ($`2\times 1`$) disordered structure. Moreover a phase transition from the ordered c($`4\times 2`$) to a disordered ($`2\times 1`$) reconstruction, with small c($`4\times 2`$) and p($`2\times 2`$) domains, has been observed with LEED at 200 K. Recently Yokoyama et al. showed with STM measurements the suppressive influence of steps on the phase transition leading to a long-range ordered c($`4\times 2`$) structure. From ab-initio molecular dynamics simulations Shkrebtii et al. established that the room temperature Si(100) corresponds to a mixture of the c($`4\times 2`$) and p($`2\times 2`$) geometries, together with instantaneous symmetric-like “twisted dimers” configurations. At room temperature Scanning Tunneling Microscopy (STM) images show a symmetric configuration, due to the quick flipping of dimers between their two possible orientations. Recently Shigekawa et al. showed that such flipping and the appearance of symmetric dimers in STM is connected to a migrating phase defect at domain boundaries.
In the last years different experimental studies, with Reflectance Anisotropy Spectroscopy (RAS) and Surface Differential Reflectivity (SDR) techniques, have been performed on Si(100), both on clean and adsorbate-covered surfaces. Several theoretical calculations are available, both based on tight-binding and first- principles , for the 2x1 and 2x2 reconstructions, but none is available for the most stable c(4x2) phase. For this reason, and also for the difficulty of preparing monodomain step-free surfaces, not all features in the optical spectra are well understood. The agreement between the theoretical spectra obtained with the TB scheme and those computed within the DFT-LDA is not always satisfactory, confirming that the study of the optical properties of this surface is not a simple matter. From the experimental point of view, Shioda and van der Weide and Jaloviar et al. pointed out recently the importance of a correct preparation of the clean surface (a terrace width about 1000 times larger than that obtained on vicinal surfaces, typically of 40-160 $`\AA `$, can be obtained with the techniques employed in Refs. and ) in order to minimize spurious effects due to the steps. In this way, the experiments are better suited to be compared with the theoretical results.
The purpose of this work is to produce a thourough ab-initio calculation of the optical properties of this surface, including the ground-state c(4x2) reconstruction, to be used to interpret RA spectra. We also calculate the optical properties of the hydrogen-covered surface and use them to determine SDR spectra.
The article is organized as follows: in Sec. II we review the theory for the calculation of the optical spectra. In Sec. III, the method of calculation is explained and in Sec. IV the optical spectra for the various surfaces are presented and compared with the experiments. Finally, the conclusions are given in Sec. V.
## II Theory
The surface contribution to the reflectivity is defined as the deviation of the reflectance with respect to the Fresnel formula, which is valid for an abrupt surface. The general formulas for the reflection coefficient of $`s`$ and $`p`$ radiation can be found solving the light-propagating equations at surfaces, when inhomogeneity and anisotropy at the surface are fully taken into account (see Ref. for a review). For normally incident light the correction to the Fresnel formula for the reflectivity is
$$\frac{\mathrm{\Delta }R_i(\omega )}{R_0(\omega )}=\frac{4\omega }{c}Im\frac{\mathrm{\Delta }ϵ_{ii}(\omega )}{ϵ_b(\omega )1},$$
(1)
where $`ϵ_b`$ is the bulk dielectric function, $`R_0`$ is the standard Fresnel reflectivity, and the subscript $`i`$ refers to the direction of light polarization. $`\mathrm{\Delta }ϵ_{ii}`$ is directly related to the macroscopic dielectric tensor $`ϵ_{ij}`$ of a semi-infinite solid, and is dimensionally a length. In a repeated slab geometry, introduced in order to simulate the real surface, $`\mathrm{\Delta }ϵ_{ii}`$ is given by
$$\mathrm{\Delta }ϵ_{ii}=d[1+4\pi \alpha _{ii}^{hs}ϵ_b(\omega )],$$
(2)
where $`d`$ is half of the slab thickness and $`\alpha _{ii}^{hs}`$ is the half-slab polarizability. For this geometry, the change of reflectivity with respect to the Fresnel formula reduces to :
$$\frac{\mathrm{\Delta }R_i(\omega )}{R_0(\omega )}=\frac{4\omega d}{c}Im\frac{4\pi \alpha _{ii}^{hs}(\omega )}{ϵ_b(\omega )1}.$$
(3)
The imaginary part of $`\alpha _{ii}^{hs}`$ in the single particle scheme adopted here can be related to the transition probabilities between slab eigenstates
$$\mathrm{}m[\alpha _{ii}^{hs}(\omega )]=\frac{\pi e^2}{m^2\omega ^2Ad}\underset{\stackrel{}{k},v,c}{}|p_{v,c}^i(\stackrel{}{k})|^2\delta (E_c(\stackrel{}{k})E_v(\stackrel{}{k})\mathrm{}\omega ),$$
(4)
where $`p_{v,c}^i(\stackrel{}{k})`$ denotes the matrix element of the i- component of the momentum operator between valence states ($`v`$) and conduction states ($`c`$), and $`A`$ is the area of the sample surface. The k-space integration is written as a sum over $`\stackrel{}{k}`$ vectors in the two dimensional Brillouin zone. The real parts of the half-slab polarizability and bulk dielectric function are obtained via the Kramers-Kronig transform. In the presence of a non-local potential, the momentum operator in (4), describing the coupling of electrons with the radiation, should be replaced by the velocity operator. However, Pulci et al. have shown that to neglect this effect has a negligible influence on the RA.
Using Eq. (3) and Eq. (4) it is then possible to compute RAS and SDR spectra. In particular, the RA measures the difference in the reflectivity, as function of photon energy, for light polarized in two orthogonal directions in the specimen surface, i.e.
$$\frac{\mathrm{\Delta }R}{R_0}=\frac{\mathrm{\Delta }R_y\mathrm{\Delta }R_x}{R_0}.$$
(5)
On the other hand, SDR measures the difference of the reflectance (with unpolarized or polarized light) between the clean and the covered surface:
$$\frac{\mathrm{\Delta }R}{R}=\frac{\mathrm{\Delta }R_{\text{clean}}\mathrm{\Delta }R_{\text{cov}}}{R_\text{0}}.$$
(6)
## III Method
The electronic wavefunctions and eigenvalues are obtained from a self-consistent calculation, in the standard DFT-LDA scheme, using a plane-waves basis set and the Ceperley-Alder exchange-correlation potential as parametrized by Perdew and Zunger. The electron-ion interactions are treated by norm-conserving, fully separable ab-initio pseudopotentials, and a kinetic energy cutoff of 15 Rydberg is used. The crystal with its surface is simulated by a repeated slab; 12 atomic layers and a vacuum region of 4 empty layers are needed to obtain a good convergence of the optical spectra of Si(100). For each structure considered the atomic positions correspond to the fully relaxed configuration obtained by Car-Parrinello molecular dynamics runs. Three different reconstructions, ($`2\times 1`$), p($`2\times 2`$) and c($`4\times 2`$), were considered. In Tab. 1 we report the structural parameters obtained for the various reconstructions of the clean surface. In order to obtain SDR spectra, calculations were performed also for a passivated surface (two Hydrogens per surface Si atom– yielding a bulk-like terminated (1x1) surface without any dimers), using the same ingredients as for the corresponding clean surfaces. In the calculation of the optical spectra, we verified that the use of a reduced kinetic energy cut-off (10 Ry. instead of 15) does not change the resulting spectra appreciably. The gain in computational cost was then exploited to perform very extended convergence tests with respect to the k–points sampling used. We found that for the ($`2\times 1`$) and hydrogenated surfaces a 64 k-points set is necessary and sufficient to obtain converged spectra in the range 0-5 eV, while for the c($`4\times 2`$) and p($`2\times 2`$) reconstructions, due to the smaller BZ, 32 k points are enough. For each surface, we compute equation (4) using a number of empty states within 1 Ry from the conduction band minimum.
In order to compare the theoretical spectra with the experimental curves, we apply an upward rigid shift $`\mathrm{\Delta }`$ = 0.5 eV to the conduction bands (in agreement with the results of previous GW calculations on this surface) due to the well known underestimation of the gaps in the DFT-LDA method. The momentum matrix elements are renormalized with the factor $`(E_cE_v+\mathrm{\Delta })/(E_cE_v)`$, according to the recipe of Del Sole and Girlanda.
## IV Optical spectra
We obtain band structures in good agreement with other, well converged DFT-LDA calculations appeared in the literature, in particular with those of Ref. . For the ($`2\times 1`$) reconstruction, the $`\pi `$ and $`\pi ^{}`$ surface states show a rather strong dispersion along the dimer rows ($`JK`$, $`J^{}\mathrm{\Gamma }`$ directions). This implies a relevant interaction between adjacent dimers in the same row, while the flatter dispersion along the dimer direction (i.e., perpendicularly to the row) indicates a small interaction of dimers belonging to different rows. The c($`4\times 2`$) and p($`2\times 2`$) reconstructions, in contrast with the ($`2\times 1`$) case, show two occupied and two empty dangling-bond surface bands, in agreement with ARUPS experiments. The dispersion of the unoccupied $`\pi ^{}`$ surface state of the c($`4\times 2`$) reconstruction is in good agreement with recent standing wave patterns measurements, obtained with STM at low temperature .
In Fig. 1 we show the RA spectra obtained for the ($`2\times 1`$) surface using three different geometries: the first with symmetric dimers, the second with a buckling of 0.56 $`\AA `$ (similar to Kipp’s calculation) and the last of 0.71 $`\AA `$ (which corresponds to our fully converged geometry at the cut-off of 15 Ry). The changes in the theoretical results are evident: the first low energy peak, essentially due to dimer surface-states, shifts from 1 eV to 1.4 eV (due to an opening of the gaps) and changes its sign, in going from the symmetric to the asymmetric case. The same peak increases in strength and shifts slightly to higher energy (1.6 eV) with the increase of the buckling. On the other hand, the positive peak at 4.3 eV, which is always present in the experimental spectra, reduces with the increase of the buckling.
Fig. 2 shows the RAS obtained for the c($`4\times 2`$) reconstruction. Our results for the p($`2\times 2`$) and ($`2\times 1`$) spectra are also included. The latter are consistent with the results obtained by Kipp et al., although not identical due to the different equilibrium geometries at which the optical spectra have been calculated. The c($`4\times 2`$) and p($`2\times 2`$) spectra show however a larger anisotropy on the low frequency side, due to transitions across the folded dangling-bond bands, while the ($`2\times 1`$) reconstruction has a single negative peak in this range, at about 1.4 eV. Differences are present expecially below $`\mathrm{}\omega =1.5eV`$ between the c($`4\times 2`$) and p($`2\times 2`$) reconstructions, confirming the strong dependence of the spectra upon the details of the reconstruction, in particular upon the ordering of the buckled dimers along and perpendicularly to the rows. The comparison of the overall shape with the measurements obtained by Shioda and van der Weide is generally good, showing that the various features of the experimental spectra are due essentially to the different reconstructions present in the specimens.
The interpretation of the experimental spectra is complicated by the disordered nature of the dimer buckling at room temperature, and by spurious effects due to the steps, or defects, that cannot be completely eliminated. The negative peak close to 3.7 eV can be explained by any of the three structures considered in Fig. 2, while the 3 eV structure implies the presence of (2x1) domains at room temperature. Especially for the peak at about 4.3 eV, the comparison with the experiment is not completely clear. Cole et al. found that its intensity is different in different samples, and decreases almost linearly while increasing the temperature. Shioda and van der Weide found a similar behavior by decreasing the miscut angle of the vicinal surfaces, which corresponds to reducing the majority domain coverage (or, in other words, the anisotropy of the surface) . Yasuda et al., from the analysis of their RT spectra, concluded that this peak is essentially originated by a step induced dicroism; however, the opposite conclusion has been reached by Jaloviar et al., who singled out the contribution of the steps to the RA. In our spectra a peak in the energy region around 4.3 eV is present for the p($`2\times 2`$) and c($`4x2`$) reconstructions, in agreement with the observed reduction with increasing temperature. On the other hand, as shown in Fig. 1, a peak at the same energy is present also in the ($`2\times 1`$) case, and its intensity depends on the dimer buckling. Hence, we conclude, in agreement with Refs. and , that this structure does not arise from the steps.
SDR experiments on the clean Si(100) surface for $`\mathrm{}\omega <1eV`$ were carried out in 1983 by Chabal et al. using the multiple internal total reflection arrangement. In 1980 Wierenga at al., and more recently Keim et al. , measured SDR using normal-incidence external reflection. Hydrogen, oxygen, or water were used in the experiments in order to saturate the dangling bonds (only oxygen was used in the last one). In Fig. 3 we report theoretical SDR spectra calculated using, as the clean surface contribution, the c($`4\times 2`$), the p($`2\times 2`$) and the ($`2\times 1`$) reconstructions, respectively. The ($`1\times 1`$) H-covered surface is used as the reference surface. The experimental results of Ref. and are also shown. The positive peak at about 1.5 eV, present in the experimental spectra, is confirmed by our theoretical analysis. The first peak below 1 eV found by Chabal et al. appears only when the reflectivity of the clean c($`4\times 2`$) reconstruction is used in eqs. 4 and 6, although a shoulder is also present for the p($`2\times 2`$) reconstruction. (A peak at the same energy is visible also in EELS spectra.) Experimentally, the intensity of this peak is found to be weakly dependent on the temperature from 40 K to room temperature. Different explanations of this structure have been proposed in the past: in EELS experiments it was interpreted as a direct transition at $`\mathrm{\Gamma }`$ from the bulk top valence to an empty surface state, while in SDR experiments it was explained as an indirect transition. Instead, in Ref. it was interpreted in terms of transitions across the dangling-bonds of flipping dimers, when these are instantaneously unbuckled: in this position indeed, the energy separation has the right value to produce a peak just below 1 eV (see the dotted curve in Fig. 1). Although this explanation cannot be ruled out, we may now argue that the SDR structure below 1 eV could be due to the presence of c($`4\times 2`$) and p($`2\times 2`$) domains, where it is essentially due to transitions from bulk states near the top valence to unoccupied surfaces states. The small intensity of the experimental peak may be due to the small fraction of c($`4\times 2`$) domains, caused by thermal disorder as well as by the presence of defects or steps, which can suppress the long-range ordered structures.
At higher energies, a peak at about 2.5 eV is reported in ref. while two peaks at 2.90 and 3.95 eV are present in the experimental spectra of Ref. when the surface is exposed to NO<sub>2</sub>, and a further small peak appears at 3.1 eV when molecular oxygen is used. The authors of Ref. argue that the peaks at 2.9 and 3.95 eV are related to transitions between surface states; instead, in our theoretical analysis for all the reconstructions studied, all spectral features above 2 eV are essentially due to surface-bulk, bulk-surface and bulk-bulk transitions.
The agreement between our DFT-LDA RA spectra and those obtained in the TB method is limited, even if the same geometry is used. We show in Fig. 4a the RA calculated according to the two methods for the 2x1 reconstruction, using a buckling of 0.56 Å(as for the full-line curve in Fig. 1), which is very close to that of Ref. . As previously found, the sign of the RAS at about 1.5 eV, due to transitions across surface states, is opposite to that obtained with the nearest-neighbors TB calculations. The latter coincides with that of a naive picture of the dimers considered as isolated Si<sub>2</sub> molecules. This can be explained on the basis of the omission, in the TB approach, of the direct interaction between atoms belonging to adjacent dimers. In fact, the explanation of this apparently paradoxical sign of the reflection anisotropy can be found looking at the surface band structure: the dispersion of the surface-localized $`\pi `$ and $`\pi ^{}`$ bands along the row direction (i.e., perpendicularly to the dimers) is comparable to the $`\pi `$\- $`\pi ^{}`$ separation. This means that the interaction between adjacent dimers is about as large as the interaction between the $`p_z`$ orbitals of the two atoms in a single dimer. Hence, the correct picture of these surfaces is that of interacting chains of dimers, oriented in the direction perpendicular to the dimers themselves, and separated by large valleys through which the dimers interact very weakly. Recently Gavrilenko and Pollack performed theoretical calculations both with the DFT-LDA and the TB schemes. However, while their TB results agree with our calculations, their DFT- LDA spectra, which are obtained with some kind of “ad-hoc” treatment of the self- energy effects, appear to be at variance with both our results and the calculations of ref. , and do not describe correctly the experimental spectra.
In Fig. 4 (b) we compare the unpolarized reflectivity spectra obtained in our TB and LDA calculations, using the same geometry (buckling of 0.56 $`\AA `$). In this case, our ab–initio calculation of the SDR spectra basically agree with the TB results and with previously published data , confirming that once averaged over the polarizations the spectra depend less critically on the theoretical scheme used. In particular, the occurrence of the (experimentally measured) first peak at about 1.4 eV as a consequence of dimer buckling, which was the main point of Ref. , is confirmed. In the case of unbuckled dimers, in fact, the first SDR peak occurs at 0.9 eV, where the first RAS structure in Fig.1 occurs.
## V Conclusions
We have carried out ab-initio calculations of the optical properties of the Si(100) surface within DFT-LDA. For the first time this approach has been applied to the c($`4\times 2`$) clean surface and to the ($`1\times 1`$) hydrogenated surface. Quite surprisingly, the optical properties of the c($`4\times 2`$) reconstructed surface are qualitatively different from those of the p($`2\times 2`$) phase on the low-frequency side. A qualitative understanding of RAS and SDR measurements is obtained. The structure appearing below 1 eV in the SDR spectrum, not well understood so far, is naturally explained in terms of the occurrence of c($`4\times 2`$) domains.
Our ab-initio DFT-LDA calculations allow us to check the reliability of the TB approach. SDR TB spectra are in qualitative agreement with DFT-LDA spectra, while some discrepancy appears in the case of RAS spectra, which are more sensitive to the details of the local geometry and require more accurate calculations. The omission of second-neighbor interactions in the $`sp^3s^{}`$ scheme used in our and previous TB calculations is mostly responsible for this discrepancy. Calculations using a second-neighbours parametrization scheme should correct this shortcoming of the TB approach.
## VI Acknowledgments
We are grateful to S. Goedecker for providing us an efficient code for Fast Fourier Transforms. This work was supported in part by the European Community programme “Human Capital and Mobility” (Contract No. ERB CHRX CT930337), and by the INFM Parallel Computing Initiative. Calculations were performed at CINECA (Interuniversity Consortium of the Northeastern Italy for Automatic Computing). B.S.M. thanks partial support of CONACyT-México (26651-E) and CNR-Italy through a collaboration program.
|
no-problem/9812/quant-ph9812013.html
|
ar5iv
|
text
|
# Purification via entanglement swapping and conserved entanglement
## Abstract
We investigate the purification of entangled states by local actions using a variant of entanglement swapping. We show that there exists a measure of entanglement which is conserved in this type of purification procedure.
The resource of entanglement has many useful applications in quantum information processing, such as secret key distribution , teleportation and dense coding . Polarization entangled photons have been used to demonstrate both dense coding and teleportation in the laboratory. Teleportation has also been realized using path-entangled photons and entangled electromagnetic field modes . Accompanying the practical applications of entanglement are some useful schemes for entanglement manipulation which may help in the distribution of entanglement between distant parties. One such scheme is entanglement swapping , which enables one to entangle two quantum systems that have never interacted directly with each other; we have discussed how this may be used in constructing a quantum telephone exchange . Recently, entanglement swapping has been demonstrated experimentally . There exists yet another useful manipulation of entanglement in which local actions and classical communication are used by two distant parties to distill a certain number of shared Bell states from a larger number of shared but less entangled states. When the initial shared but less entangled states are pure, this manipulation is termed as entanglement concentration , while for the more general case of the initial shared states being mixed the process is termed as entanglement purification or distillation . The importance of such a scheme in the distribution of entanglement is obvious as Bell pairs are essential for the implementation of quantum communication schemes with perfect fidelity. Curiously, such an important procedure remains to be realized in an experiment. In this paper, we will show that a simple variant of the entanglement swapping scheme can be viewed as a type of entanglement concentration procedure and has a physical realization with polarization entangled photons. Moreover, we show that there exists a certain measure of entanglement which remains conserved on average in this type of entanglement concentration. Note that, in this paper we will often use the terms entanglement concentration and entanglement purification in an interchangeble manner, though what we demonstrate is, in the strict sense, only the concentration of pure shared entanglement.
Let pairs of photons $`(1,2)`$ and $`(3,4)`$ be in the following polarization entangled states
$`|\mathrm{\Phi }(\theta )_{12}=\mathrm{cos}\theta |H_1,H_2+\mathrm{sin}\theta |V_1,V_2,`$ (2)
$`|\mathrm{\Phi }(\theta )_{34}=\mathrm{cos}\theta |H_3,H_4+\mathrm{sin}\theta |V_3,V_4,`$ (3)
where the phase angle $`\theta `$ satisfies $`0<\theta <\pi /2`$. There are a number of ways to prepare photons in such polarization entangled states. For example, one may first use type-II down conversion followed by suitable birefringent crystals to prepare two photons in the Bell state state $`|\mathrm{\Phi }^+=|HH+|VV`$ (the other Bell states being $`|\mathrm{\Phi }^{}=|HH|VV`$ and $`|\mathrm{\Psi }^\pm =|HV\pm |VH`$). This may be followed by placing a dichroic element (such as the local filters described in Ref.) which selectively absorbs any one of the polarization components (say $`H`$) along the path of one of the photons. In cases when this photon exits the element unabsorbed, the pair of photons are in the state $`e^{\gamma L}|HH\pm |VV`$ (not normalized), where $`L`$ is the length of the crystal and $`\gamma `$ is the absorption per unit length. Thereby states of the type given by Eqs.(2) and (3) with $`\mathrm{sin}\theta =\sqrt{1/(1+e^{2\gamma L})}`$ is generated. This procedure may seem inefficient because of the possibility of the photon being absorbed by the dichroic element. However, due to the absence of two qubit logic gates for polarization entangled photons, there is no way to proceed unitarily from $`|HH\pm |VV`$ to the states given by Eqs.(2) and (3) and dissipative processes are necessary. Alternatively, one can use the recently suggested tunable ultrabright source of polarization entangled photons to directly produce the states given in Eqs.(2)-(3). Photons $`2`$ and $`3`$ are brought together, while photons $`1`$ and $`4`$ are allowed to travel to separate distant locations as shown in Fig.1. If one now performs a Bell state measurement on photons $`2`$ and $`3`$, then immediately the states of the photons $`1`$ and $`4`$ become entangled. This effect, called entanglement swapping, has been tested for maximally entangled photons (when $`\theta =\pi /4`$) . We shall refer to this tested case (i.e when maximally entangled photons are used) as standard entanglement swapping. For the more general case of an arbitrary $`\theta `$, the combined state of the photons $`1`$ and $`4`$ is projected to either of the following four states, depending on the outcome of the Bell state measurement on photons $`2`$ and $`3`$:
$`|\mathrm{\Phi }^{^{}}(\theta )_{14}^+`$ $`=`$ $`{\displaystyle \frac{1}{N}}(\mathrm{cos}^2\theta |H_1,H_4+\mathrm{sin}^2\theta |V_1,V_4),`$ (5)
$`|\mathrm{\Phi }^{^{}}(\theta )_{14}^{}`$ $`=`$ $`{\displaystyle \frac{1}{N}}(\mathrm{cos}^2\theta |H_1,H_4\mathrm{sin}^2\theta |V_1,V_4),`$ (6)
$`|\mathrm{\Psi }_{14}^+`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(|H_1,V_4+|V_1,H_4),`$ (7)
$`|\mathrm{\Psi }_{14}^{}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(|H_1,V_4|V_1,H_4),`$ (8)
where, $`N=\sqrt{\mathrm{cos}^4\theta +\mathrm{sin}^4\theta }`$. The probabilities to obtain the four states are
$`P(|\mathrm{\Phi }^{^{}}(\theta )_{14}^+)`$ $`=`$ $`P(|\mathrm{\Phi }^{^{}}(\theta )_{14}^{})={\displaystyle \frac{\mathrm{cos}^4\theta +\mathrm{sin}^4\theta }{2}},`$ (10)
$`P(|\mathrm{\Psi }_{14}^+)`$ $`=`$ $`P(|\mathrm{\Psi }_{14}^{})=\mathrm{cos}^2\theta \mathrm{sin}^2\theta ,`$ (11)
where the symbol $`P(|\psi )`$ is used to denote the probability of the state $`|\psi `$. We now proceed to explain the sense in which the above variant of standard entanglement swapping is a purification procedure. Say photons $`2`$,$`3`$ and $`4`$ are held by one party (Alice) and photon $`1`$ is possessed by the other party (Bob) as shown in Fig.1. Now, Alice can change the magnitude of the entanglement she shares with Bob by doing a Bell state measurement on photons $`2`$ and $`3`$ and thereby projecting photons $`1`$ and $`4`$ to one of the states given by Eq.(5)-(8). In the cases when she projects photons $`1`$ and $`4`$ to either $`|\mathrm{\Psi }_{14}^+`$ or $`|\mathrm{\Psi }_{14}^{}`$ (which are Bell states), she actually increases the magnitude of entanglement she shares with Bob. In the other cases she reduces the magnitude of entanglement she shares with Bob even further. If she initially shared a large enough ensemble of photons in the state $`|\mathrm{\Phi }(\theta )_{12}`$ with Bob and applied the procedure described above on each shared pair separately, then she would be able to change the states of a certain fraction of the shared pairs to Bell states at the expense of degrading the entanglement of the other shared pairs even further. This qualifies as a type of purification procedure because local actions (by Alice) are used to concentrate the entanglement of a fraction of the shared states while the entanglement of the remaining fraction is being diluted, just as in other purification procedures .
We should pause here briefly to mention a fact relevant from the experimental viewpoint. It is known that a complete Bell state measurement is not possible with only linear elements . However, a complete Bell state measurement is not really necessary to implement the above purification protocol. One only needs to be able to discriminate the states $`|\mathrm{\Psi }_{23}^+`$ and $`|\mathrm{\Psi }_{23}^{}`$ from each other and from the rest of the Bell states, which can be done in existing experiments . The reason for this is that the photons $`1`$ and $`4`$ are projected to Bell states ($`|\mathrm{\Psi }_{14}^+`$ or $`|\mathrm{\Psi }_{14}^{}`$) only when the outcome of the Bell state measurement on photons $`2`$ and $`3`$ is either $`|\mathrm{\Psi }_{23}^+`$ or $`|\mathrm{\Psi }_{23}^{}`$. To verify whether the purification has indeed taken place, one simply has to pass the resultant Bell states of photons $`1`$ and $`4`$ through a Bell state analyser .
An interesting feature of the original collective entanglement concentration procedure (called the ”Schmidt projection method”) described in Ref. was that the average of the entropy of entanglement (the von Neumann entropy of the partial density matrix seen by either party) of all the shared pairs remained constant under purification (an asymptotic result). In the scheme described here, the average von Neumann entropy of entanglement of the shared pairs is, in fact, decreased. However, we shall show that there is a different measure of entanglement whose average remains conserved in this procedure of purification via entanglement swapping. This measure of entanglement is defined as the maximum probability with which two parties sharing a pure entangled state can convert it to a Bell state by classically communicating and performing local actions on their respective sides. As this is already the maximum probability, it can only remain constant or decrease under any set of local general measurements and classical communications (This is the basic criterion to be satisfied by any measure of entanglement ). Thereby, it qualifies as a measure of entanglement which can be termed as entanglement of single pair purification (This measure, is however, not an additive measure). We shall denote this measure by $`E_S`$ and refer to it henceforth simply as entanglement.
We now proceed to show that $`E_S`$ is conserved in the purification process described above. From the results of Lo and Popescu , it follows that the maximum probability with which a Bell state can be obtained by purifying a single entangled pair (in a pure state) is twice the modulus square of the Schmidt coefficient of smaller magnitude. In the case of the states given by Eqs.(2) and (3), this is simply $`2\mathrm{cos}^2\theta `$ if $`\mathrm{cos}\theta `$ is the smaller of the two Schmidt coefficients. Therefore, before the entanglement swapping, the average value of the entanglement shared between Alice (A) and Bob (B) is
$$E_S_{AB}=2\mathrm{cos}^2\theta .$$
(12)
After the entanglement swapping, when the shared states are those given by Eqs.(5)-(8) with probabilities given by Eqs.(10)-(11),
$`E_S_{AB}`$ $`=`$ $`P(|\mathrm{\Phi }^{^{}}(\theta )_{14}^+)E_S(|\mathrm{\Phi }^{^{}}(\theta )_{14}^+)`$ (13)
$`+`$ $`P(|\mathrm{\Phi }^{^{}}(\theta )_{14}^{})E_S(|\mathrm{\Phi }^{^{}}(\theta )_{14}^{})`$ (14)
$`+`$ $`P(|\mathrm{\Psi }^+)E_S(|\mathrm{\Psi }^+)+P(|\mathrm{\Psi }^{})E_S(|\mathrm{\Psi }^{})`$ (15)
$`=`$ $`2\mathrm{cos}^2\theta ,`$ (16)
where $`E_S(|\psi )`$ denotes the entanglement of the state $`|\psi `$. Therefore, the ensemble average of the entanglement of single pair purification is a conserved quantity in the process of purification by entanglement swapping. This result also indicates that the purification via entanglement swapping is an optimal protocol for single pair entanglement purification, as the average entanglement of the purified pairs is equal to the original entanglement. Here, by optimality we mean the best combination of entangled states that can be finally obtained by the purification. A positive implication of this optimality is that the purification process can be separately repeated on the final subensembles, which turn out to be less entangled, without destroying any entanglement on average (though for that, one needs to be able to do complete Bell state measurements, for which schemes have been suggested ). If one continues this process indefinitely, in the limit of an infinite sequence, the final ensemble generated will comprise of a certain fraction of Bell pairs and a certain fraction of completely disentangled pairs. This fraction of Bell pairs should be equal to $`2\mathrm{cos}^2\theta `$ as the average of the entanglement has been conserved in each step. Thus in the limit of an infinite sequence, the process of purification via entanglement swapping allows us to convert all the entanglement that can possibly be extracted by single pair purifications from the ensemble into Bell pairs.
Here, it is worthwhile to mention that a procedure for single pair purification called the ”Procrustean method” had also been suggested in the original entanglement concentration paper . The Procrustean method also conserves $`E_S`$ and its fractional yield of Bell states is optimal (i.e equal to $`E_S`$). From the point of view of efficiency, our method lies somewhere between the Schmidt projection method (when it is implemented with two entangled pairs, but one pair totally belonging to Alice) and the procrustean method. Our method has a fractional yield of Bell states equal to the Schmidt projection method (with two entangled pairs), but the non-Bell state outcomes are also entangled. It is because of the extra entanglement of the non-Bell state outcomes that $`E_S`$ is conserved on average in our method of purification. Hence the improvement over the Schmidt projection method, brought in by entanglement swapping is to make the non-Bell state outcomes entangled as well.
Next, we proceed to consider the case when photon pairs $`(1,2)`$ and $`(3,4)`$ are not in the same type of entangled state. Suppose, photons $`1`$ and $`2`$ start in the entangled state $`|\mathrm{\Phi }(\theta _1)_{12}`$ and photons $`3`$ and $`4`$ start in the entangled state $`|\mathrm{\Phi }(\theta _2)_{34}`$ (defined as in Eqs.(2) and (3)). Let the entanglement of the first photon pair be $`E_S=\alpha `$ and the entanglement of the second photon pair be $`E_S=\beta `$. Then a simple calculation shows that projecting photons $`2`$ and $`3`$ onto a Bell state basis projects the photons $`1`$ and $`4`$ to states $`(\mathrm{cos}\theta _1\mathrm{cos}\theta _2|H_1,H_4\pm \mathrm{sin}\theta _1\mathrm{sin}\theta _2|V_1,V_4)`$ and $`(\mathrm{cos}\theta _1\mathrm{sin}\theta _2|H_1,V_4\pm \mathrm{sin}\theta _1\mathrm{cos}\theta _2|V_1,H_4)`$ (not normalized) with specific probabilities. The average of the entanglement between the photons $`1`$ and $`4`$ after the projection turn out to be $`E_S=\text{min}\{\alpha ,\beta \}`$. The manipulation of entanglement described above can be visualized as a step towards the complete generalization of entanglement swapping. The original scheme involved the use of two Bell states and Bell state measurements . It has been generalized to cases when many particle Greenberger-Horne-Zeilinger (GHZ) states are used and many particle GHZ measurements are conducted . The procedure presented here is a generalization of the original scheme to the case when two particle pure states which are less entangled than Bell states are used, but measurements conducted are still Bell state measurements. This result is illustrated in Fig.2. One can get an intuitive feel of this result from the golden rule that entanglement, on average, cannot be increased under local actions and classical communications . In the situation depicted in Fig.2, we have the freedom to choose which photons belong to Alice and which to Bob. For $`\alpha <\beta `$, we choose to allot the photon $`1`$ to Bob and the rest to Alice, while for $`\alpha >\beta `$ we allot the photon $`4`$ to Bob and the rest to Alice. Then a Bell state measurement on Alice’s side cannot increase the entanglement she shares with Bob on average. As this initial entanglement is the smaller of the numbers $`\alpha `$ and $`\beta `$, the final average entanglement has to be smaller than or equal to $`\text{min}\{\alpha ,\beta \}`$. However, the fact that it is actually equal, is a peculiar feature of the entanglement swapping process.
One may envisage a situation in which Alice tries to purify the state $`|\mathrm{\Phi }(\theta _1)_{12}`$ shared with Bob with the help of the state $`|\mathrm{\Phi }(\theta _2)_{34}`$ (which we can call the purifier) in her possession. As long as $`\beta <\alpha `$, she degrades the entanglement shared with Bob on average, while when $`\beta \alpha `$, she conserves the entanglement shared with Bob on average. Thus in order not to loose any entanglement the purifier state should have at least as much entanglement as the state to be purified. Moreover, the degree to which the entanglement is concentrated (i.e inhomogeneously redistributed among the four measurement outcomes) gets better as $`\beta `$ approaches $`\alpha `$. Bell pairs are produced only when $`\beta `$ exactly equals $`\alpha `$. Thus, as Alice increases the entanglement of her purifier, the degree of entanglement concentration increases yielding Bell pairs when $`\beta `$ reaches $`\alpha `$ (a criterion which can be called entanglement matching). On increasing $`\beta `$ further, the degree of concentration starts going down. When $`\beta `$ reaches unity (maximal entanglement), there is no concentration of entanglement (all the four outcomes have an entanglement equal to that of the original state: a situation equivalent to the perfect fidelity teleportation of an entangled state).
In this paper we have shown that entanglement swapping can be used to purify single pairs of polarization entangled photons. Such a scheme may be achieved by an extension of an existing experiment . In contrast, physical implementation of the entanglement purification schemes involving collective measurements (like the Schmidt projection method of Ref.) will be difficult, as they would involve measurements on many photons at once. The absence of two qubit logic gates for polarization entangled photons also make purification procedures described in Ref. difficult to realize. From this point of view, the method described here should be a positive first step in the direction of implementation of purification procedures. In the paper we have also introduced a measure of entanglement for single pure pairs and demonstrated that this quantity is conserved in the above process. This makes purification via entanglement swapping interesting from a purely theoretical angle. The non-additive nature of the introduced measure implies that additivity is not an essential requirement for an entanglement measure to have a physical interpretation. As a natural generalization of our measure one may use the maximum possible fractional yield of Bell states from various finite collections of shared pairs as physically relevant measures of entanglement. The physical relevance stems from the fact that in a real implementation of entanglement concentration one would always have access to only a finite number of systems. All the results presented in this paper hold only for pure states. An extension to mixed states will not be trivial, as single pairs in such states cannot, in general, be purified . However, it would still be interesting to investigate whether one can generalize the measure of entanglement presented here to some quantity which is conserved in entanglement swapping with mixed states.
This work was supported in part by the Inlaks Foundation, the UK Engineering and Physical Sciences Research Council, the European Union, Elsag-Bailey and Hewlett-Packard.
|
no-problem/9812/nucl-th9812017.html
|
ar5iv
|
text
|
# WHAT CHIRAL SYMMETRY CAN TELL US ABOUT HADRON CORRELATORS IN MATTER
## I Introduction
There are little doubts that Chiral Symmetry (CS) is one of the most important principles of low-energy hadron physics. . In the limit of massless quarks the QCD Lagrangian is symmetric with respect to the $`SU(N)\times SU(N)`$ chiral group. It is generally believed that this symmetry is spontaneously broken. CS breaking manifests itself in the absence of chiral multiplets of the particles with the same masses but different parities, for example, $`\rho a_1`$ or $`\sigma \pi `$ mesons. In the language of the correlation functions the broken chiral symmetry means that the lowest pole positions of vector and axial vector correlators, describing the $`\rho `$ and $`a_1`$ mesons , are different. The restoration of the symmetry in vacuum would result in the identity of the corresponding correlators which in turn leads to the same masses of the chiral partners. The other manifestation of the spontaneously broken CS is the occurrence of the nonzero order parameters like two quark condensate $`\overline{q}q`$. In the case of of hadron interactions in vacuum $`q\overline{q}`$=0 would imply that hadron masses become much smaller compared to its observed values. The relationships between condensates, hadron masses and corresponding correlators are significantly more complicated in the presence of medium. For example, the change of the nucleon mass in medium is not completely determined by the corresponding change of the quark condensate. The other example is the CS restoration for the correlators of the chiral partners. In matter the identity of the corresponding in-medium correlators does not nesseseraly means the degeneracy of the effective masses of chiral partners. The identity of the masses of chiral partners at the point of restoration is only one of the posibilities. The dynamics of hadrons in nuclear medium is described by the corresponding correlators. Let’s consider the case of the two-point correlators. It can be written in the form
$$\mathrm{\Pi }(p)=id^4xe^{ipx}\mathrm{\Psi }|T\{J(x)J(0)\}|\mathrm{\Psi }.$$
(1)
Where $`J`$ is the interpolating current in the corresponding hadron channel and the matrix element is taken over the ground states of the system with finite density $`\rho `$. The position of the lowest pole as the function of density determines the in-medium mass of hadron. If wave function $`\mathrm{\Psi }`$ describes the infinite system of non interacting nucleons than the above correlator reflects the dynamics of the probe hadron moving in some mean field formed by nuclear matter. To calculate the corrections to this picture one needs to take into account the nuclear pions. Then the correlator can represented by the sum of two terms describing the contributions from the system of noninteracting nucleons and pionic corrections. Let’s consider the specific part of this corrections where pion is emitted and then absorbed by nuclear matter. The corresponding piece of the correlators can be represented in the form
$$\frac{d\text{k}}{4\omega _k}\frac{d\text{k}^{}}{\omega _k^{}}\mathrm{\Psi }|a_𝐤^aa_𝐤^{}^b|\mathrm{\Psi }id^4xe^{ipx}\mathrm{\Psi }\pi ^a(\text{k})|T\{J(x)J(0)\}|\mathrm{\Psi }\pi ^b(\text{k}^{})$$
(2)
The sum over the isospin indices is assumed. The terms with the different time orderings can be accounted for in a similar manner. We consider the nuclear pions in the chiral limit. Since we are interested in the properties of hadron correlators which are related to chiral symmetry treating nuclear pions in the chiral limit seems to be a reasonable assumption in our case. By using the soft-pion theorem the part of the correlator describing the pionic corrections can be written as follows
$$\mathrm{\Pi }^\pi =\frac{i}{2}\xi d^4xe^{ipx}\mathrm{\Psi }|[Q_5^a,[Q_5^a,T\{J(x),J(0)\}]]|\mathrm{\Psi },$$
(3)
where we denoted $`\xi =\frac{\rho \overline{\sigma }_{\pi N}}{f_\pi ^2m_\pi ^2}`$ and $`\overline{\sigma }_{\pi N}`$ is the leading nonanalytic part of the pion-nucleon sigma term $`\sigma _{\pi N}`$ The chiral expansion of $`\sigma _{\pi N}`$ reads as
$$\sigma _{\pi N}=Am_\pi ^2\frac{9}{16}(\frac{g_{\pi N}}{2M_N})^2m_\pi ^3+\mathrm{}\mathrm{}.$$
(4)
First, analytic term describes the short range contributions. In contrast, the nonanalytic term is due to long distance contribution of the pion cloud. Let’s consider the correlator of two nucleon interpolating current. Matter can influence the properties of the QCD vacuum and thus change the two quark condensate. It is usually believed that the reduction of $`<\overline{q}q>`$ is related to the effect of partial restoration of CS. To the first order in the density the evolution of the two-quark condensate is given by
$$\mathrm{\Psi }|\overline{q}q|\mathrm{\Psi }=0|\overline{q}q|0(1\frac{\sigma _{\pi N}}{f_\pi ^2m_\pi ^2}\rho ),$$
(5)
It turned out that not any change of $`\overline{q}q`$ is in the one-to-one correspondence with the CS restoration phenomena since the terms of order $`m_\pi `$ presenting in the condensate are not allowed in in-medium nucleon mass. Let’s show how to cancel the terms not allowed by CS in QCD sum rules. In QCD sum rules one relates the characteristics of QCD vacuum and the phenomenological in-medium nucleon spectral density. The effects of the low-momentum pions are long-ranged and stem from the phenomenological representation of the in-medium nucleon correlator. Assuming the Ioffe choice of the nucleon interpolating current, making use of the transformation property of this current one can get the following chiral expansion of the in-medium nucleon correlator
$$\mathrm{\Pi }(p)\mathrm{\Pi }^0(p)\frac{\xi }{2}\left(\mathrm{\Pi }^0(p)+\gamma _5\mathrm{\Pi }^0(p)\gamma _5\right),$$
(6)
Where $`\mathrm{\Pi }^0(p)`$ is the nucleon correlator in chiral limit. It is useful to decompose the correlator into three terms with the different Dirac structures
$$\mathrm{\Pi }(p)=\mathrm{\Pi }^{(s)}(p)+\mathrm{\Pi }^{(p)}(p)p\text{/}+\mathrm{\Pi }^{(u)}(p)u\text{/},$$
(7)
where $`u^\mu `$ is a unit four-vector defining the rest-frame of nuclear system. Only the piece $`\mathrm{\Pi }^{(s)}(p)`$ gets affected by the chiral corrections of order $`\rho m_\pi `$. Splitting the phenomenological expression of the nucleon correlator into pole and continuum parts one can get
$$\mathrm{\Pi }(p)\mathrm{\Pi }_{pole}(p)\frac{\xi }{2}\gamma _5\mathrm{\Pi }_{pole}(p)\gamma _5+\left(1\frac{\xi }{2}\right)\mathrm{\Pi }_{cont}^0(p)\frac{\xi }{2}\gamma _5\mathrm{\Pi }_{cont}^0(p)\gamma _5,$$
(8)
Where we denoted $`\mathrm{\Pi }_{pole}(p)(1\frac{\xi }{2})\mathrm{\Pi }_{pole}^0(p)`$ The explicit expression of the pole term has the following form
$$\mathrm{\Pi }_{pole}(p)=\lambda ^2\frac{p\text{/}+M^{}+V\gamma _0}{2E(\text{p})[p^0E(\text{p})]},$$
(9)
Here $`M^{}`$ is the in-medium nucleon mass including the scalar part of the self energy and $`\lambda ^{}`$ is the nucleon coupling. Then one writes the three independent sum rules, one for each Dirac structure
$$(1\frac{\xi }{2})𝑑p\frac{w(p)}{2E[p^0E]}\frac{(1\xi )}{\lambda ^2M^{}}𝑑pw(p)\left[\mathrm{\Pi }_{OPE}^{(0,s)}(p)\mathrm{\Pi }_{cont}^{(0,s)}(p)\right]$$
(10)
$$(1+\frac{\xi }{2})\lambda ^2𝑑p\frac{w(p)}{2E[p^0E]}𝑑pw(p)\left[\mathrm{\Pi }_{OPE}^{(0,p)}(p)\mathrm{\Pi }_{cont}^{(0,p)}(p)\right],$$
(11)
$$(1+\frac{\xi }{2})\lambda ^2V𝑑p\frac{w(p)}{2E[p^0E]}𝑑pw(p)\left[\mathrm{\Pi }_{OPE}^{(0,u)}(p)\mathrm{\Pi }_{cont}^{(0,u)}(p)\right].$$
(12)
Taking the ratio of these sum rules one can get the needed cancellation in the effective mass and vector self energy and bring the in-medium nucleon QCD sum rules in an agreement with the chiral symmetry constraints. Let’s consider now the in-medium correlation function of the vector currents. The lowest pole of the isovector-vector correlator corresponds to the $`\rho `$-meson contribution. Due to relatively large width it decays inside nuclear interior so the spectrum of the produced dileptons can carry the information about the modifications of the $`\rho `$-meson mass and width in matter. Such modification may be related with partial restoration of CS.One possible way to look at the phenomena of CS restoration is to study the correlators describing the in-medium dynamics of the chiral partners. The correlators of the chiral partners, in our case the correlators of the vector and axial-vector currents, should become identical in the chirally restored phase. Thus one can expect that these correlators get mixed when the symmetry is only partially restored. The effect of chiral mixing indeed takes place both at finite temperatures and densities . Making use of the standard commutation relation of current algebra $`[Q_5^a,J_\nu ^b]=iϵ^{abc}A_\nu ^c`$ and putting it in the expression for the correlator of the vector currents $`\mathrm{\Pi }_V`$ one can get
$$\mathrm{\Pi }_V=\mathrm{\Pi }_V^0+\xi (\mathrm{\Pi }_V^0\mathrm{\Pi }_A^0);\mathrm{\Pi }_A=\mathrm{\Pi }_A^0+\xi (\mathrm{\Pi }_A^0\mathrm{\Pi }_V^0)$$
(13)
The parameter $`\xi `$ in this equation is 4/3 times the one for the nucleon correlator. $`\mathrm{\Pi }_V^0`$($`\mathrm{\Pi }_A^0`$) is the correlator of the vector (axial) currents calculated in the approximation of the noninteracting nucleons. As one can see from the above equations the correlators get mixed when soft pion contributions are taken into account. We note that this statement is model independent and follows solely from CS. CS implies that the correlators of the chiral partners acquire, due to mixing, the additional singularities. These singularities may manifest themselves in the, for example, dilepton spectrum, produced in the heavy-ion collisions. It means that the spectrum may show the additional enhancement at the energy region close to the mass of the $`A_1`$ meson, besides that at the mass of $`\rho `$ meson. The phenomena of mixing suggests few possible ways of how chiral symmetry restoration occurs. First, the lowest singularities of the both correlators could indeed coincide at the point of CS restoration. Second, the correlators may exhibit two poles of the same strength corresponding to the $`\rho `$ and $`A_1`$ mesons. Third, the width of the mesons could become large enough to make no longer sensible the whole concept of individual quantum state at high densities. One notes that the $`\omega `$ meson being an isospin singlet is not affected by the chiral corrections. It follows from the fact that the commutator of the isoscalar vector current with the axial charge $`Q_a^5`$ is zero. Let’s briefly consider the mixing of the other type of chiral partners, namely the $`\sigma `$-$`\pi `$ pair. Making use of the current algebra commutation relation $`[Q_5^a,\pi ^b]=\delta ^{ab}i\sigma `$ and $`[Q_5^a,\sigma ]=i\pi ^a`$ one can get
$$\mathrm{\Pi }_\pi =(1+\xi )\mathrm{\Pi }_\pi ^0\xi \mathrm{\Pi }_\sigma ^0;\mathrm{\Pi }_\sigma =(1+\xi )\mathrm{\Pi }_\sigma ^0\xi \mathrm{\Pi }_\pi ^0$$
(14)
Here the mixing parameter is two times smaller than that for $`\rho `$-$`A_1`$ system and thus the effect of $`\sigma `$-$`\pi `$ mixing is less pronounced at the normal nuclear density than in the case of $`\rho `$-$`A_1`$ mixing. The point of the complete restoration of chiral symmetry should, of course, be the same for all kinds of the chiral partners but the “velocity” of approaching to this point may well be different. Similar to the case of the $`\rho `$-$`A_1`$ system the mixing in the pseudoscalar-scalar channel practically means that the correlators exhibit the singularities which are dictated by CS and should be taken into account regardless of the model used to describe the concrete hadronic processes. However, the effect of the $`\sigma `$-$`\pi `$ mixing can probably be observed at relatively large densities. The case worth studying is deeply bound pion states in heavy nuclei.
## Acknowledgments
Author would like to thank Mike Birse for discussions about the role of chiral symmetry in nuclei. It is also a pleasure to thank the organizers of Baryons98 for providing with the opportunity to attend the conference.
## References
|
no-problem/9812/cond-mat9812399.html
|
ar5iv
|
text
|
# Comment on “Critical behavior of the chain–generating function of self–avoiding walks on the Sierpinski gasket family: The Euclidean limit”
##
The self–avoiding walk (SAW) on a lattice is a random walk that must not contain self–intersections. The criticality of SAWs has been extensively studied as a challenging problem in statistical physics on the Euclidean lattices and on fractal lattices as well. Accordingly, the question has been posed whether the critical behavior of SAWs on a Euclidean lattice can be retrieved via a limit of infinite number of fractals whose properties gradually acquire the corresponding Euclidean values. In this Comment we scrutinize the methods used so far to answer the foregoing question.
The most frequently studied infinite family of finitely ramified fractals appears to be the Sierpinski gasket (SG) family. Each member of the SG family is labeled by an integer $`b`$ ($`2b\mathrm{}`$), and when $`b\mathrm{}`$ both the fractal $`d_f`$ and spectral $`d_s`$ dimension approach the Euclidean value 2. Concerning the study of the criticality of SAWs, these fractal lattices are perfect objects for application of the renormalization group (RG) method, due to their intrinsic dilation symmetry (the so-called self–similarity) and their finite ramification. The latter property enables one to construct a finite set of the RG transformations and therefrom an exact treatment of the problem. This treatment was first applied by Dhar , for $`b=2`$, and later it was extended up to $`b=8`$. The obtained results of the corresponding critical exponents, for the finite sequence $`2b8`$, were not sufficient to infer their relation to the relevant Euclidean values. However, these results inspired the finite–size scaling (FSS) approach to the problem , which brought about the prediction that the SAW critical exponents on fractals do not necessarily approach their Euclidean values when $`b\mathrm{}`$. Indeed, Dhar found that the critical exponent $`\nu `$, associated with the SAW end–to–end distance, tends to the Euclidean value 3/4, whereas the critical exponent $`\gamma `$, associated with the total number of distinct SAWs, approaches 133/32, being always larger than the Euclidean value $`\gamma _E=43/32`$.
The intriguing FSS results motivated endeavors to extend the exact RG results beyond $`b=8`$. However, since this extension appeared to be an arduous task, a new insight was needed. This insight came from a formulation of the Monte Carlo renormalization group (MCRG) method for fractals , which produced values for $`\nu `$ and $`\gamma `$ up to $`b=80`$. For $`2b8`$ the MCRG findings deviated, from the exact values, at most 0.03%, in the case of $`\nu `$, and 0.2% in the case of $`\gamma `$. In addition, the behavior of the entire sequence of the MCRG findings, as a function of $`b`$, supported the FSS predictions.
Recently, Riera and Chalub made a different type of endeavor to obtain results for the critical exponent $`\gamma `$ for large $`b`$, by applying an original series expansion method . However, the data of Riera and Chalub (RC) display a quite different behavior than the MCRG results (see Fig. 1). As regards comparison of the RC results with the available exact RG results , one may notice a surprising discrepancy: for $`b=7`$ the RC result deviates 19% (which should be compared with the respective MCRG deviation 0.13%), while for $`b=8`$ the RC result deviates 33% from the exact result (which is again much larger than the corresponding MCRG deviation 0.15%). On the other hand, concerning behavior of $`\gamma `$ beyond $`b=8`$, the RC results start to decrease, whereas the MCRG results monotonically increase. Furthermore, Riera and Chalub claimed that, in contrast with the FSS prediction , $`\gamma `$ should approach the Euclidean value 43/32=1.34375, in the limit of very large $`b`$. These discrepancies call for inspection of both methods, that is, of the MCRG technique and the series expansion method . We are going first to reexamine our MCRG approach, and then we shall comment on the applicability of the RC series expansion method for large $`b`$.
We have found that the best way to check the validity of the MCRG method, for large $`b`$, is to apply it in a case of a random walk model that is exactly solvable for all possible $`b`$. To this end, the so-called piecewise directed walk (PDW) turned out to be quite appropriate. The PDW model describes such a random self-avoiding walk on SG fractals in which the walker is allowed to choose randomly, but self-similarly, limited number of possible step directions . This model corresponds to the directed random walk on Euclidean lattices, in which case $`\nu _E=1`$ and $`\gamma _E=1`$. By applying the exact RG approach the critical exponent $`\nu `$ and $`\gamma `$ for PDW have been obtained for each $`b`$ ($`2b<\mathrm{}`$). Moreover, it was demonstrated exactly that $`\nu `$ approaches the Euclidean value $`\nu _E=1`$, while $`\gamma `$ tends to non-Euclidean value $`\gamma =2`$, when $`b\mathrm{}`$. Here, we apply the MCRG method (used in the case of SAW in ) to calculate $`\nu `$ and $`\gamma `$ of the PDW model for $`2b100`$. Our results, together with the exact findings, are presented in Table I and depicted in Fig. 2 and Fig. 3. One can see that, in the entire region under study, the agreement between the MCRG results and the exact data is excellent. Indeed, the deviation of the MCRG results for $`\nu `$ from the corresponding exact results is at most 0.08%, while in the case of $`\gamma `$ it is at most 0.8%. This test of the MCRG method provides novel reliability for its application in studies of random walks on SG with large $`b`$.
Because of the confirmed reliability of the MCRG method, and because Riera and Chalub have obtained quite different results, in the case of SAWs on SG, we have reason to assume that their conclusions were obtained in a wrong way. Thus, we may pose a question what was wrong in the application of the series expansion method in the work of Riera and Chalub . Let us start with mentioning that in the series expansion study of SAW the first task is to determine the number $`c_n(b)`$ of all possible SAWs for a given number $`n`$ of steps, where $`1nn_{max}`$. Of course, in practice, it is desirable to perform this enumeration for very large $`n_{max}`$, as the corresponding numbers $`c_n(b)`$ of all $`n`$-step SAWs represent coefficients of the relevant generating function $`C_b(x)=_{n=1}^{\mathrm{}}c_n(b)x^n`$ (where $`x`$ is the weight factor for each step), whose singular behavior determines critical exponents of SAWs. In order to take into account existence of the SG lacunarity, the average end-to-end distance of the set of $`n`$-step SAWs should be larger than the size of the smallest homogeneous part of the SG fractal , that is, $`n_{max}`$ should be larger than $`b^{4/3}`$. In a case when the number of steps is smaller than $`b^{4/3}`$ the corresponding SAWs percieve the underlying fractal lattice as a Euclidean substrate. In order to make it more transpicuous, we present in Fig. 4 the curve $`n=b^{4/3}`$ which divides the ($`b,n`$) plane in two regions so that one of them corresponds to the fractal behavior of $`n`$-step SAWs, while the other corresponds to the Euclidean behavior. In the same figure, for a given $`b`$, we also depict the number of coefficients (empty small triangles) that were obtained by Riera and Chalub for the corresponding SG fractal, in their series expansion approach. One can see that only for $`2b8`$ Riera and Chalub generated sufficient number of coefficients $`c_n(b)`$ so as to probe the fractal–behavior region (in which the condition $`n_{max}b^{4/3}`$ is satisfied). On the other hand, for $`b>8`$, in all cases studied, the maximum length $`n_{max}`$ of enumerated SAWs is not larger than 16, and thereby the corresponding generating functions $`C_b(x)`$ remain in the domain of the Euclidean behavior (see Fig. 4). For instance, in order to study criticality of SAWs on the SG fractal, for $`b=80`$, it is prerequisite to calculate all coefficients $`c_n(b)`$ in the interval $`1nn_{max}`$, where $`n_{max}`$ must be larger than $`80^{4/3}339`$, which is far beyond $`n_{max}=13`$ that was reached in for $`b=80`$. This explains why the RC results for the SG critical exponent $`\gamma `$ (see Fig. 1), with increasing $`b`$, wrongly become closer to the Euclidean value 1.34375.
The problems discussed above, that is, the problems with not long enough SAWs, do not appear in the MCRG study of SAWs on the SG fractals, as the RG method in general takes into account SAWs of all length scales. Incidentally, we would like to mention that in it was erroneously quoted that in the MCRG studies one Monte Carlo (MC) realization corresponds to simulation of one SAW. In fact, one MC realization implies simulations of all possible walks on the fractal generator, which appears to be the smallest homogeneous part of the SG fractal. For instance, for the $`b=80`$ SG fractal, in order to calculate the critical exponent $`\gamma `$, in one MC realization we simulated all $`n`$-step SAWs with $`n`$ ranging between 1 and 3240.
Finally, we would like to comment on the analytical argument, given in , which was assumed to support the claim $`lim_b\mathrm{}\gamma =\gamma _E`$. One can observe that the corresponding argument does not exploit particular properties of the SAWs studied. Thus, if the argument were valid, it could be applied to other types of SAWs on fractals leading to the same conclusion $`lim_b\mathrm{}\gamma =\gamma _E`$. However, the case of the PDW discussed in this comment is a definite counterexample to the foregoing conclusion, as it was rigorously demonstrated that in this case $`lim_b\mathrm{}\gamma \gamma _E`$.
In the conclusion, let us state that in this comment we have vindicated that the MCRG technique for studying the SAW critical exponents on fractals is a reliable method and a valuable tool in discussing the query whether the critical behavior of SAWs on a Euclidean lattice can be achieved through a limit of infinite number of fractals whose properties gradually acquire the corresponding Euclidean values. On the other hand, we have demonstrated that Riera and Chalub , in an attempt to answer the mentioned query by applying the series expansion method, have not provided sufficient number of numerical data for a study of criticality of SAWs on the SG fractals with finite scaling parameters $`b`$. Therefore, any inference from such a set of data about the large $`b`$ behaviour of the critical exponent $`\gamma `$ cannot be tenable.
We would like to acknowledge helpful and inspiring correspondence with D. Dhar concerning the matter discussed in this Comment.
|
no-problem/9812/cond-mat9812290.html
|
ar5iv
|
text
|
# Comment on “Is The Nonlinear Meissner Effect Unobservable?”
In a recent Letter by Li, Hirschfeld and Wölfle (LHW) on nonlocal effects in unconventional superconductors, it was suggested that these effects might explain the null result for the nonlinear Meissner effect (NLME) in our experiments on optimally doped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6.95</sub> (YBCO) single crystals for which an appreciable signal is predicted by theory .
We have no objection to the main part of the LHW letter, which deals with a detailed calculation of the nonlocal effects. However, the remarks made about our experimental results do not directly follow from these detailed calculations but are critically dependent on a qualitative argument which fails to work for YBCO.
The qualitative argument relies on treating YBCO as a “weakly 3-D” system. This leads LHW to the conclusion that nonlocal effects will wipe out the NLME (for the geometry of our experiments) for fields below about $`0.81H_{c1}`$. However, the estimate of $`H_{c1}`$ from the “weakly 3-D” argument is $`\mathrm{\Phi }_0/(2\pi \lambda _0\lambda _{0c})`$ (where $`\mathrm{\Phi }_0`$ is the flux quantum, and $`\lambda _0`$ and $`\lambda _{0c}`$ are the zero temperature penetration depths for currents flowing in the a-b plane and along the c-axis respectively), which leads to a value of $`H_{c1}`$ of twenty Gauss or less. This is over an order of magnitude below the experimental value of the field at which first flux penetration occurs which is 300 Gauss or more. In our experiments, the samples used have typical dimensions of 1.5mm x 1.5mm x 50$`\mu `$m (a x b x c). The magnetic field is applied in the a-b plane and given the sample geometry, most of the screening current flows in the a-b plane with components along the nodal directions, with the exception of the return currents that flow near the edges in the c-axis direction. Thus, we expect the field of first flux entry to be closer to $`\mathrm{\Phi }_0/(2\pi \lambda _0^2),`$ which is borne out by experiment. For currents in the a-b plane, the nonlocal effects are very small, much smaller than the NLME at fields of 300G. For the return currents, the nonlocal effects may be relevant, but these do not contribute in any way to NLME, as they have no components in the nodal directions. Thus, for our experimental geometry, the nonlocal contributions are irrelevant.
The same conclusion can be reached even more starkly by starting from the estimate of the characteristic nonlocal energy, given in LHW as $`E_{nl}=\xi _{0c}\mathrm{\Delta }_0/\lambda _0`$. Quasiparticle effects will be ineffective, due to nonlocality, for quasiparticles within an angle of less than $`\varphi _{nl}`$ from a node, where $`\varphi _{nl}`$ is determined from the condition $`\mathrm{\Delta }(\varphi _{nl})/E_{nl}1`$. This yields $`\varphi _{nl}0.001`$ implying that the NLME requires an applied field $`H>H_m`$ with $`H_m/H_00.001`$ where $`H_0`$ is the characteristic field scale of the NLME. Since $`H_0`$ is about 8000 gauss, we have that $`H_m`$ is about ten gauss, in rough agreement with the argument in the previous paragraph.
To summarize: we find it quite plausible that the nonlocal effects indeed render the NLME unobservable at fields below ten or twenty Gauss. Since the experiments are performed at fields over one order of magnitude larger, however, with the sample remaining in the Meissner state, it is obvious that the explanation for our negative result must lie elsewhere. In our opinion the presence of at least a few percent component of imaginary $`s`$ or $`d_{xy}`$ character in the gap remains the most likely explanation.
|
no-problem/9812/astro-ph9812241.html
|
ar5iv
|
text
|
# Formation of Dark Matter Halos
## 1. Radial Profiles of Dark Matter Dominated Galaxies
Navarro, Frenk, & White (1995) proposed the following simple radial density profile
$$\rho _{\mathrm{NFW}}(r)=\frac{\rho _0}{(r/r_s)(1+r/r_s)^2}$$
(1)
for dark matter halos corresponding to X-ray clusters. In very influential papers (hereafter referred to as NFW96 and NFW97) they subsequently showed that $`\rho _{\mathrm{NFW}}(r)`$ is a good fit to profiles of dark matter halos in SCDM and $`\mathrm{\Lambda }`$CDM models, and in CDM models with power law $`P(k)=Ak^n`$ fluctuation spectra with $`n=0`$ to -1.5. They characterized the halos by a concentration parameter $`cr_{200}/r_s`$, where $`r_{200}`$ refers to the radius within which the average overdensity is 200 times critical density. For $`\mathrm{\Omega }_{\mathrm{matter}}=1`$ models, $`r_{200}`$ is approximately the same as the virial radius. The mass enclosed is called $`M_{200}`$, and it is useful to express this in units of the nonlinear mass $`M_{}`$, which is defined for $`\mathrm{\Omega }_{\mathrm{matter}}=1`$ models by $`\mathrm{\Delta }_0(M_{})=\delta _0(1+z)`$, where $`\mathrm{\Delta }_0(M)`$ is the rms fluctuation of the mass in a sphere of average mass $`M`$ (calculated from the linear power spectrum) and $`\delta _0=1.69`$ for top-hat collapse. Finally, NFW96 argued that halos with lower $`M_{200}/M_{}`$ are less concentrated because they form earlier.
The $`\rho _{\mathrm{NFW}}(r)`$ profile seems consistent with data on clusters, and the $`\rho 1/r`$ behavior at small $`r`$ was consistent with earlier high-resolution N-body simulations (e.g., Dubinski & Carlberg 1991). But, as pointed out by Flores & Primack (1994) and Moore (1994), and acknowledged by NFW96, $`\rho 1/r`$ for small $`r`$ is inconsistent with data on dark matter dominated dwarf irregular galaxies. Moreover, Burkert (1995) showed that the four galaxies considered by Moore (1994) have essentially the same rotation curve shape, corresponding to<sup>1</sup><sup>1</sup>1 The analytically implied rotation curve shape is $`V_\mathrm{B}^2(r)=2V_\mathrm{B}^2(r_b){\displaystyle \frac{\mathrm{ln}[(1+r/r_b)^2(1+(r/r_b)^2)]2\mathrm{t}\mathrm{a}\mathrm{n}^1[r/r_b]}{r/r_b}},`$ where $`V(r)r`$ for small $`r`$. The peak in velocity occurs at $`r_{\mathrm{max}}3.3r_b`$, and $`V_{\mathrm{max}}=V_\mathrm{B}(r_{\mathrm{max}})2.4V_\mathrm{B}(r_b)`$.
$`\rho _\mathrm{B}(r)={\displaystyle \frac{\rho _b}{(1+r/r_b)[1+(r/r_b)^2]}}.`$ (2)
This makes it implausible that a complicated starburst process leading to non-adiabatic expulsion of much of the central baryonic content of the galaxy, such as proposed by Navarro, Eke, & Frenk (1996), could account for the apparent inconsistency between simulations and real galaxies. The implausiblilty that the resolution of the discrepancy lies in this direction was further increased when Kravtsov, Klypin, Bullock, & Primack (1998, hereafter KKBP98) showed that a larger set of ten dark matter dominated dwarf irregular galaxies, and also a set of seven dark matter dominated low surface brightness (LSB) galaxies, all have the same rotation curve shape, again corresponding to $`\rho _\mathrm{B}(r)`$. This is shown in Figure 1. In our sample we included only those galaxies in which the dark matter component was shown to constitute $`>85\%`$ of the total mass inside the last measured point of the rotation curve (in most cases with the maximum disk assumption). These dark matter dominated galaxies offer a unique opportunity for probing directly the density structure of dark matter halos which can be then compared with predictions of theoretical models. It hardly seems possible that all of these galaxies could have had the same sort of complicated conspiracy between dark matter and star formation, nor that the LSBs could have lost a significant fraction of their central baryons. A more plausible interpretation of the self-similarity of the radial mass distribution in these dark matter dominated galaxies is that it reflects an underlying similarity in the dark matter distribution.
McGaugh & de Blok (1998) emphasized that the sharp $`r^{1/2}`$ rise in central circular velocity predicted by the NFW $`r^1`$ density profile is in striking disagreement with the roughly linear rise in rotation velocity observed in LSB galaxies: “Even treating both $`c`$ and $`V_{200}`$ as completely free parameters, no fit can be obtained.”
KKBP98 fit the observed rotation curves with a more general profile,
$$\rho _{DM}(r)=\frac{\rho _0}{(r/r_0)^\gamma [1+(r/r_0)^\alpha ]^{(\beta \gamma )/\alpha }},V(r)=V_t\frac{(r/r_t)^g}{[1+(r/r_t)^a]^{(g+b)/a}},$$
(3)
in which $`\rho (rr_0)r^\gamma `$, $`\rho (rr_0)r^\beta `$, and $`\alpha `$ characterizes the sharpness of the change in logarithmic slope. This is equivalent to $`\rho _{\mathrm{NFW}}(r)`$ for $`(\alpha ,\beta ,\gamma )=(1,3,1)`$, and to the so-called isothermal profile with a core $`r_0`$ for $`(\alpha ,\beta ,\gamma )=(2,2,0)`$. Parameters $`r_t`$ and $`V_t`$ of the corresponding rotation curve are the effective “turnover” radius and velocity, and $`a`$ parameterizes the sharpness of the turnover. Because of the relatively small range of radii probed by measured rotation curves, such measurements cannot be used to constrain all five parameters of $`\rho _{DM}`$, so we fixed the outer logarithmic slope to the value suggested by the theoretical models $`\beta =3`$, $`b=0.34`$. The plausible value of the parameter $`\alpha =2`$ was determined using galaxy rotation curves that do show a turnover. We then found that $`\gamma 0.2`$ fits the observed rotation curves of our sample of 17 galaxies. The corresponding best-fit slopes of the velocity profile are $`(a,b,g)=(1.50,0.34,0.9)`$. Note that $`g=1\gamma /2`$. With parameters $`\alpha `$, $`\beta `$, and $`\gamma `$ (or $`a,b,g`$) fixed, we fitted the data for the remaining two free parameters: $`\rho _0`$ and $`r_0`$ (or $`V_t`$ and $`r_t`$, or in terms of Burkert’s profile, $`\rho _b`$ and $`r_b`$). The resulting structural relations show a decrease in the characteristic density with increasing characteristic radii, or an increase in maximum rotation velocity with increase in the radius at which the maximum occurs. Matching these observational relations is a challenge for any theory that aspires to explain the observed rotation curves.
## 2. Radial Profiles From Simulations
Does the disagreement between CDM-type simulations and the observed rotation curves of dark matter dominated galaxies at small radii mean that the dark matter in these galaxies is not mostly cold dark matter? That would be the implication if the discrepancy were real.
We have used the Adaptive Refinement Tree (ART) $`N`$-body code (Kravtsov et al., 1997), which adaptively refines spatial and temporal resolution in high density environments, to simulate the evolution of collisionless dark matter in the three cosmological structure formation models<sup>2</sup><sup>2</sup>2The simulations followed trajectories of $`128^3`$ CDM particles in a box of size of $`L_{box}=7.5h^1`$ Mpc. The CHDM simulation had $`3\times 128^3`$ particles due to the addition of two equal-mass neutrino species.: (a) the standard cold dark matter (CDM) model ($`\sigma _8=0.67`$, $`h=0.5`$); (b) a low-density CDM model with cosmological constant ($`\mathrm{\Lambda }`$CDM) with parameters favored by the high-redshift supernovae data ($`\mathrm{\Omega }_{\mathrm{matter}}=0.3`$, $`h=0.7`$, $`\sigma _8=1`$); and (c) a cold$`+`$hot dark matter model with two types of neutrino (CHDM) with favorite parameters $`\mathrm{\Omega }_\nu =0.2`$, $`h=0.5`$) (Primack et al. 1995, Gawiser & Silk 1998, but cf. Primack & Gross 1998). For the dark matter halos used in KKBP98, the spatial resolution is equal to $`0.52h^1`$ kpc, and for each of the analyzed halos we have taken into account only those regions of the density and circular velocity profiles that correspond to scales at least twice as large as the formal resolution. The reliability of the simulated density and velocity profiles was tested by comparing results of the simulations with different resolutions and time steps (Kravtsov et al. 1997). A sample of $`50`$ halos was extracted from each simulation, and rotation curves of halos were fitted with the same density distribution as the data (the same set of $`\alpha `$, $`\beta `$, and $`\gamma `$). The halo rotation curves were then renormalized to their best fit values of $`r_0`$ and $`v_0=v(r_0)`$ and these renormalized rotation curves (i.e. $`v/v_0(r/r_0)`$) were averaged over the whole sample for each model.
Figure 2 shows the average normalized dark matter velocity profiles for halos formed in each of the three cosmological structure formation models compared with the rotation curves of the dark matter dominated galaxies in our sample. We find that on average, the velocity profiles of the halos formed in the hierarchical structure formation models and observed dark matter halos are in reasonably good agreement. Why did we not see the significant discrepancy between numerical simulations and rotation curve measurements indicated by previous work? One possible explanation is that we have not used an (extrapolated) analytic model fit, but have compared the data with the average shape of the dark matter halos directly. We also find that dark matter dominated dwarf and LSB galaxies show structural correlations between their characteristic density, $`\rho _0`$, and radius, $`r_0`$, consistent with the correlations of our simulated dark matter halos: physically smaller halos are denser. We find a similar correlation between the maximum of the rotation curve, $`v_{max}`$, and the corresponding radius $`r_{max}`$ (see KKBP98 for details, and also Figure 3 below). This increases our confidence that the agreement between the simulations and the observed dark matter dominated galaxies is not a fluke.
In recent work, finished after KKBP98, we have run and analyzed a $`128^3`$ particle CDM ART simulation ($`\sigma _8=1.0`$) in a box of 2.5 $`h^1`$ Mpc, 1/3 the linear size of that used for the simulations shown in Figure 2. Figure 3 shows the well-resolved halos fit to the Burkert profile; as already mentioned, the $`\rho _{DM}`$ profile with $`\gamma 0.2`$ is essentially identical. Again we have verified that the structural relations, such as the correlation between the maximum rotation velocity and the radius at which this maximum occurs, are in reasonable agreement with our sample of dwarf and LSB galaxies — see Figure 4.
Moore et al. (1998) reports results from two very high resolution simulations of clusters, which had steep central density profiles. It is not clear that our results are in disagreement, since we consider galaxy-mass halos and find that a statistical sample of halos has a range of central profiles. But it is important to understand why different simulations give different results, and in particular to understand the effects of different resolution and different simulation techniques, of two-body relaxation (which is suppressed in the ART approach), and of the selection of halos to be simulated and analyzed.
It would appear to be better to compare a sample of galaxies with a statistical sample of galaxy-mass halos, as we have done. This has many other advantages. In work now being written up for publication, Bullock et al. have analyzed large ART simulations at many redshifts, to study the radial distribution of mass and angular momentum in dark matter halos at a given epoch, and also the evolution of these properties. We find that the dispersion of the concentration is roughly log-normal and large, and that while it can be explained at $`z=0`$ by the argument presented by NFW96, this does not work at higher redshifts. In addition, halos in higher density environments tend to be more concentrated than isolated halos, suggesting that halo selection criteria may indeed be important for interpreting the conflicting results mentioned above. Since the dispersion of the concentration at $`z=0`$ in a given cosmology (Bullock et al. 1998) is larger than the difference between different cosmologies (which comes largely from the difference between $`r_{200}`$ and the true virial radius), it is incorrect to draw conclusions about cosmology from a few galaxies (as done, for example, by Navarro 1998, who argues that a few low-concentration galaxies favor $`\mathrm{\Omega }_{\mathrm{matter}}`$ very low).
One problem that may be resolved in the light of new data is the apparently greater dispersion of the inner radial profiles of the simulations than of the dark matter dominated galaxies in our sample. The work of Swaters et al. (1997) discussed at this meeting, and also of Côté, Freeman, & Carignan (1997), appears to indicate that the dispersion of the properties of dark matter dominated galaxies may be greater than our sample suggested. However, as our sample in KKBP98 represented only the $`50`$ most massive halos in our simulations, our scatter may be artificially small. A major challenge to theorists remains, to explain why the dark matter profiles of galaxies and simulations have the radial dependences observed. Regarding the inner profiles, Syer & White (1998) argued that the closer the power spectrum approximates the asymptotic CDM slope of -3, where halos of all sizes collapse at roughly the same epoch and therefore have the same density, the shallower the resulting profile, while on cluster scales where smaller-mass halos collapse earlier at higher density and are subsequently incorporated into larger-mass halos, these will go to the center and give rise to a steeper radial density profile. This seems consistent with our results, as mentioned in KKBP98. We should then understand why the results of Huss, Jain, & Steinmetz (1998) seem inconsistent with this argument. Regarding why the outer radial dependence is roughly $`r^3`$, recent work by Henriksen & Widrow (1998) may be relevant.
## 3. Press-Schechter
The Press-Schechter (1974, hereafter PS) formula for the number density of dark matter halos as a function of their mass is based on two simple assumptions – Gaussian statistics for density fluctuations, and spherical top-hat collapse of these fluctuations. (See, e.g., Peebles (1993), pp. 630-635, or White (1996) for modern treatments.) Since both of these assumptions are known to be wrong, or at best oversimplified, the wonder is not that the PS approximation is not perfect, but rather that it works at all. In fact, the PS formula predicts the number density of virialized cluster-mass halos in N-body simulations remarkably well. However, several groups recently noticed a discrepancy: $`N_{PS}(>M)2N_{\mathrm{simulations}}(>M)`$ for $`M<M_{}/10`$. Since this has been seen in many simulations using different methods of simulating and identifying halos, it should be taken seriously. For example, Gross et al. (1998), using high-resolution particle mesh (PM) simulations and both spherical and ellipsoidal overdensity halo finders, found that the number density of galaxy-mass halos at the current epoch is overestimated by PS by about a factor of 2 in many currently popular CDM-type cosmological models, regardless of the collapse parameter $`\delta _c`$ used in the PS formula. Gross (1997), Appendix G, showed that the discrepancy of the PS number density is about this big for higher redshifts also, as long as $`M<M_{}/10`$. Kauffmann et al. (1998) found the same factor of $`2`$ discrepancy at the current epoch for both the $`\tau `$CDM and $`\mathrm{\Lambda }`$CDM models, and Somerville, Lemson, Kolatt, & Dekel (1998) generalized this to the halo merging trees of the Extended PS theory. This work is based on AP<sup>3</sup>M simulations and a friends-of-friends halo finder, as is related work on larger mass halos, the abundance of which is underpredicted by the PS approximation (Governato et al. 1998). Finally, Sigad et al. (1998) have found the same phenomenon in the ART simulations, using a version of the bound density maximum (BDM) halo finder (Klypin et al. 1997). A similar result was actually found independently in an analytic calculation based on approximations relevant to the highly nonlinear regime (Valageas & Schaeffer 1997).
The overprediction of the number density of galaxy-mass halos at low redshift in the PS approximation has several implications, including an amelioration of the overprediction by semi-analytic models of galaxy formation (based on the extended PS theory) of the luminosity function of galaxies and the star formation rate at low redshift (see, e.g., Somerville & Primack 1998ab). The underprediction by PS of the number density of massive halos, especially at high redshifts, means that strong conclusions about the density of the universe based on observations of clusters compared with PS predictions should be treated with caution (Governato et al. 1998).
## 4. Lyman Break Galaxies as Merger-Triggered Starbursts
Semi-analytic models of galaxy formation were pioneered by White & Frenk (1991), Kauffmann, White, & Gurderdoni (1993), and Cole et al. (1994); see Somerville & Primack (1998a) for a review. Such models follow the evolution of the dark and luminous contents of the universe using simple approximations to treat gas cooling, star formation, and feedback within dark halos, and Extended PS theory to predict the merger rate of halos in order to construct merger trees (e.g., Somerville & Kolatt 1998). When halos merge, their luminous contents are usually assumed to merge only as dynamical friction brings smaller galaxies to the large galaxy at the center of the new halo, and most of the star formation even at high redshift is quiescent (e.g., Baugh et al. 1998). However, both observational and theoretical arguments suggest that many of the small but very bright galaxies now being identified in very large numbers at redshifts $`z>3`$ by the Lyman break method (Steidel et al. 1996, Dickinson 1998) are low-mass starbursts rather than large galaxies that have been quiescently forming stars for a long time (Lowenthal et al. 1997; Sawicki & Yee 1998; and Somerville, Primack, & Faber 1998, hereafter SPF98). SPF98 assumed that random collisions of dark matter subhalos, as well as decay of orbits due to dynamical friction, would trigger mergers of (proto-)galaxies that could lead to starbursts, and they based their semi-analytic modeling of the number of such mergers on recent dissipationless simulations of the mergers of dark halos (Makino & Hut 1997), and of the star formation efficiency and timescale on hydrodynamic simulations of starbursts triggered by mergers (Mihos & Hernquist 1994, 1996). The result was that most of the star formation in CDM-type hierarchical models at redshifts $`z>2`$ occurs in merger-triggered starbursts, and that most of the Lyman break galaxies (LBGs) are expected to be such starbursts. This perhaps resolves the apparent paradox that the LBGs appear to cluster like massive halos (Steidel et al. 1998, Giavalisco et al. 1998, Adelberger et al. 1998; cf. Wechsler et al. 1998) while their relatively low linewidths (Pettini et al. 1998) and their spectral energy distributions (Sawicki & Yee 1998) suggest that they have relatively low mass (few$`\times 10^{10}M_{}`$) and young ages (few$`\times 10^8`$ yr). Including merger-triggered starbursts also predicts much more star formation at high redshift in CDM-type hierarchical models for structure formation than if only quiescent star formation is included (Somerville & Primack 1998b).
Do the merger rates actually grow so large at high redshift that this merger-triggered starburst scenario is plausible? With many stored timesteps from large ART simulations, we have begun to study the merger rate of dark matter halos as a function of redshift. This sort of study requires careful definitions of halos and subhalos, and of the criteria for identifying mergers, which will be given in papers now in preparation (Kolatt et al. 1998ab, Bullock 1999). But to summarize the situation briefly, the answer is yes. The collision rate of halos and subhalos grows in physical coordinates roughly as $`(1+z)^3`$ up to a redshift that depends on the mass of the halo. The comoving collision rate of $`>10^{10}h^1M_{}`$ halos peaks in the $`\mathrm{\Lambda }`$CDM simulations at $`z3`$, at a rate high enough to account for the observed number density of LBGs (Kolatt et al. 1998a). Indeed, the comoving number density of LBGs with AB magnitude brighter than 25.5 is predicted to be almost as high at $`z=4`$ as at $`z=3`$. We also find that the mergers occur mainly in and near the most massive halos, so that the high bias of the bright LBGs arises naturally.
### Acknowledgments.
We are grateful to our collaborators, including Avishai Dekel, Sandra Faber, Patrik Jonsson, Tsafrir Kolatt, Michael Gross, Yair Sigad, and Rachel Somerville, for allowing us to present preliminary reports of our new results here. We acknowledge support from NASA and NSF grants at UCSC and NMSU.
## References
Adelberger, K., et al. 1998, ApJ, 505, 18
Baugh, C., Cole, S., Frenk, C. S., & Lacey, C. 1998, ApJ, 498, 501
Bullock 1999, PhD dissertation, UCSC
Bullock et al. 1998, in preparation
Burkert, A. 1995, ApJ, 447, L25
Cole et al. 1994, MNRAS, 271, 781
Côté, S., Freeman, K., & Carignan, C. 1997, astro-ph/9704031, in Dark and Visible Matter in Galaxies and Cosmological Implications (Sesto Pusteria, Italy, 2-5 July 1996), eds M. Persic & P. Salucci, ASP Conference Series, 117, 52
Dickinson, M. 1998, in The Hubble Deep Field, ed. Livio, M. et al., STScI Symp. Ser., in press, astro-ph/9802064
Dubinski & Carlberg 1991, ApJ, 378, 496
Flores, R., & Primack, J. R. 1994, ApJ, 427, L1
Gawiser, E., & Silk, J. 1998, Science, 280, 1405
Giavalisco, M., et al. 1998, ApJ, 503, 602
Governato et al. 1998, astro-ph/9810189
Gross, M. A. K. 1997, PhD dissertation, UCSC, http://fozzie.gsfc.nasa.gov/
Gross, M. A. K., Somerville, R. S., Primack, J. R., Holtzman, J., & Klypin, A. A. 1998, MNRAS, in press, astro-ph/9712142
Henriksen, R. N., & Widrow, L. M. 1998, astro-ph/9805277; see also Widrow, L. M., this volume
Huss, A., Jain, B., & Steinmetz, M. 1998, ApJ, submitted, astro-ph/9803117
Kauffmann et al. 1998, MNRAS, 297, L23
Kauffmann, G., White, S. D. M., & Gurderdoni, B. 1993, MNRAS, 264, 201
Klypin, A. A., Gottlöber, S., Kravtsov, A., & Khokhlov, A. 1997, ApJ, submitted, astro-ph/9708191
Kolatt, T., et al. 1998a, Nature, submitted
Kolatt, T., et al. 1998b, in preparation
Kravtsov, A. V., Klypin, A. A., & Khokhlov, A. M. 1997, ApJS, 111, 73
Kravtsov, A. V., Klypin, A. A., Bullock, J. S., & Primack, J. R. 1998, ApJ, 502, 48 (KKBP98)
Kravtsov, A. V., Klypin, A. A., Bullock, J. S., & Primack, J. R. 1998, in Galactic halos: a UC Santa Cruz Workshop ed. D. Zaritsky, ASP Conference Series, 136, 421
Lowenthal et al. 1997, ApJ, 481, 673
Makino, J., & Hut, P. 1997, ApJ, 481, 83
McGaugh & de Blok 1998, ApJ, 499, 41
Mihos, J., & Hernquist, L. 1994, ApJ, 425, L13
Mihos, J., & Hernquist, L. 1996, ApJ, 431, L9
Moore, B. 1994, Nature, 370, 629
Moore et al. 1998, in Galactic halos: a UC Santa Cruz Workshop ed. D. Zaritsky, ASP Conference Series, 136, 426
Navarro, J. F. 1998, ApJ, submitted, astro-ph/9807084
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1995, MNRAS, 275, 56
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1996, ApJ, 462, 563
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1997, ApJ, 490, 493
Navarro, Eke, & Frenk 1996, MNRAS, 283, L72
Peebles, P. J., E. 1993, Principles of Physical Cosmology, Princeton: Princeton University Press
Pettini , M., et al. 1998, astro-ph/9806219
Press, W., & Schechter, P. 1974, ApJ, 187, 425
Primack, J. R., Holtzman, J., Klypin, A. A., & Caldwell, D. O. 1995, Phys. Rev. Lett., 74, 2160
Primack, J. R., & Gross, M. A. K. 1998, to appear in The Birth of Galaxies (Proc. Xth Recontre de Blois), astro-ph/9810204
Sawicki, M., & Yee, H. K. C. 1998, AJ, 115, 132
Sigad, Y., et al. 1998, in preparation
Somerville, R. S., & Kolatt, T. 1998, MNRAS, in press, astro-ph/9711080
Somerville, R. S., & Primack, J. R. 1998a, MNRAS, in press
Somerville, R. S., & Primack, J. R. 1998b, astro-ph/9811001, to appear in The Birth of Galaxies (Proc. Xth Recontre de Blois)
Somerville, R. S., Lemson, G., Kolatt, T., & Dekel, A. 1998, MNRAS, in press, astro-ph/9807277
Somerville, R. S., Primack, J. R. & Faber, S. M. 1998, MNRAS, in press, astro-ph/9806263 (SPF98)
Steidel et al. 1996, ApJ, 462, L17
Steidel et al. 1998, ApJ, 492, 428
Swaters, R. A., Sancisi, R., & van der Hulst, J. M. 1997, ApJ, 491, 140; cf. Swaters, R. A., this volume, and Sancisi, R., this volume
Syer, D., & White, S. D. M. 1998, MNRAS, 293, 377
Valageas, P., & Schaeffer, R. 1997, A&A 328, 435
Wechsler, R., Gross, M., Primack, J., Blumenthal, G., & Dekel, A. 1998, ApJ, 506, 19
White, S. D. M. 1996, in Cosmology and large scale structure (Les Houches, session LX, August 1993), ed. R. Schaeffer et al., Amsterdam: Elsevier Science
White, S. D. M., & Frenk, C. S. 1991, ApJ, 379, 52
|
no-problem/9812/astro-ph9812358.html
|
ar5iv
|
text
|
# 𝑅𝑋𝑇𝐸 OBSERVATIONS OF LMC X-1
## 1 INTRODUCTION
LMC X-1 is one of four extremely luminous ($`>10^{38}`$ ergs s<sup>-1</sup>) X-ray binaries in the Large Magellanic Cloud. The source has long been known to show a rather soft X-ray spectrum ($`kT2.7`$ keV; Markert & Clark 1975) and irregular X-ray variability by at least a factor of three (Griffiths & Seward 1977, Johnston, Bradt, & Doxsey 1979). However, its optical identification was uncertain for many years. The source lies within the bright emission nebula N159 (Henize 1956) containing many early-type stars (see finding chart of region in Cowley, Crampton, & Hutchings 1978). Based on the rather uncertain X-ray position, it was originally thought that the B5 supergiant R148 ($`V12.5`$) was the optical counterpart, although the peculiar O7 III star (Star #32 of Cowley et al.) $`6^{\prime \prime }`$ away could not be ruled out. Recent analysis of $`ROSAT`$ High-Resolution Imager data has confirmed that Star #32 ($`V14.8`$) is the most probable identification (Cowley et al. 1995). Spectroscopic studies of this star reveal an orbital period near 4 days and a probable mass for the compact star of $`M_X4M_{}`$ (Hutchings, Crampton, & Cowley 1983, Hutchings et al. 1987), making it a strong black-hole candidate.
Because LMC X-1 is so luminous, its X-ray properties were studied even with the earliest X-ray detectors. White & Marshall (1984) discussed the complex character of its X-ray spectrum, showing that it could not be described by a simple model because of the high-energy excess above 3 keV. Thus a two-component model with a soft thermal spectrum and plus a hard high-energy tail was needed. They pointed out the similarity in spectral properties of LMC X-1 to another black-hole candidate, LMC X-3, and they were the first to recognize that an unusually soft X-ray spectrum may be a reliable signature of a black-hole candidate.
Using Ginga data, Ebisawa, Mitsuda, & Inoue (1989) also found a two-component model was needed to fit the spectrum: an ultrasoft blackbody ($`kT0.8`$ keV) and a hard power law (photon index $``$2.5). Later, Schlegel et al. (1994a, 1994b) modeled the spectrum, based on Broad Band X-ray Telescope (BBXRT) data, finding results similar to those of Ebisawa et al.
Ebisawa et al. undertook a timing analysis of their Ginga data and found quasi-periodic oscillations (QPO) with a peak frequency of 0.0751 Hz in one of the observations. They concluded that the QPO came from the hard-tail component, which was unusually bright at the time in comparison to the thermal component. QPO were not found in the BBXRT data (Schlegel et al.), confirming the transient nature of these aperiodic signals. Since the presence of QPO might be related to the spectral state of the system, we undertook a series of observations with Rossi X-ray Timing Explorer ($`RXTE`$) over a period of nine months to search for QPO and to further define their origin.
## 2 OBSERVATIONS AND DATA ANALYSIS
A series of nine observations were made between 1996 February and 1996 October. Table 1 lists the details of the observational data, in reverse time order, following the numbering system of the $`RXTE`$ Science Data Center. Only data from the Proportional Counter Array (PCA) were analyzed due to the very low net count rate for events with energies $`>`$16 keV. Each observation was broken into two or three segments due to Earth occultations or passages through the South Atlantic Anomaly. In 1996 March, two proportional counter units (PCU) were found to be discharging, making it necessary to shut them down periodically. The result was that some observations were made with only three of the five PCU operating (observations #4, #6, and #7), as noted in Table 1. Observation #3 experienced a shutdown of two PCU during the second half of the exposure, so it was necessary to analyze this observation in two parts (#3A and #3B).
### 2.1 Light Curves and Hardness Ratios
Background-subtracted light curves, using 16-s time bins, were derived for the 2–5.9 keV and 5.9–15.9 keV energy ranges. The spectral hardness within each bin was found by calculating the ratio of counts, $`HR=`$ (5.9–15.9 keV)/(2–5.9 keV). Typical values for the hardness ratio are $``$0.1–0.2; mean values for each observation are listed in Table 1. The light and $`HR`$ curves for observations #2, #3B and #4 are shown in Fig. 1. The selected observations span a wide range of source intensity and hardness. In observation #3B (‘average’ count rate; lowest $`HR`$) LMC X-1 was in a quiescent mode with minimal flickering, while in both #2 (second highest count rate; highest $`HR`$) and #4 (lowest count rate; ‘average’ $`HR`$) the source varied rapidly on a time scale of minutes. Examination of all of the observations reveals no dependence of the source’s flickering characteristics on either the count rate or hardness. However, for each observation the plot of hardness ratio is similar to that of the high-energy count rate, implying the ratio is driven primarily by the number of 5.9–15.9 keV counts. $`Ginga`$ data for the black hole candidate GX 339$``$4 in its very high state (Miyamoto et al. 1991) show the same behavior (see their Fig. 2). While in this very high state, the source is 2–3$`\times `$ more intense and shows significantly enhanced variability (on timescales of minutes) compared to when it is in a (relatively quiescient) high state (Makishima et al. 1986). Although the $`RXTE`$ observations also show various degrees of short-term variability, there is no clear indication that LMC X-1 enters a very high spectral state.
To test the dependence of $`HR`$ on the source count rate, we used data from all observations to construct the hardness-intensity diagrams shown in Fig. 2. The plot of $`HR`$ versus low-energy count rate has no significant relationship between the parameters. However, the plot of $`HR`$ versus 5.9–15.9 keV count rate exhibits a very strong correlation which is also present in each of the individual observations. This dependence confirms the trend inferred from examination of the simple light and hardness-ratio curves.
We examined the data for time lags between high and low-energy photons by calculating the cross-correlation of the two light curves. A small positive lag, in the sense that high-energy photons tend to arrive after their low-energy counterparts, was found for all observations. However, the mean lag time in each case was comparable to the 0.025-s step used in the calculations, so that the results are not conclusive.
### 2.2 Power Density Spectra and Quasi-periodic Oscillations
To further investigate the temporal variations, power density spectra (PDS) were constructed using the background-subtracted data, rebinned at 0.125-s resolution. Within an observation, the Fourier transformation was calculated for each 512-s data segment (4096 points), covering the frequency range 0.002–4 Hz. Typically, each PDS represents the average of $``$11 individual transforms. Two sets of PDS were made, one covering the entire PCA energy range (2–60 keV) and the other restricted to primarily those photons in the hard spectral tail (5.9–15.9 keV). No significant differences between these were found. In Fig. 3 we present the power density spectrum (5.9–15.9 keV) for observation #7. The power density increases at low frequency (i.e., red noise) and can be modeled in terms of a power law plus a constant (i.e., power $`=\nu ^\alpha +C`$, where $`\alpha `$ is the power index). Fitted values of the power index, for both the 5.9–15.9 keV and 2–60 keV PDS, are listed in Table 2. The average indices are $`\alpha =1.06`$ and 0.87 for the restricted and full energy ranges, respectively. There is no significant dependence of the modeled power index with either the hardness ratio or the source count rate. For comparison, Ebisawa et al. (1989) measured an index of 0.81 in a $`Ginga`$ observation taken when LMC X-1 was in an X-ray bright state.
If present, quasi-periodic oscillations would appear as a broad peak in the PDS. None of our observations shows this signature, although Ebisawa et al. found strong QPO near 0.0751 Hz in some of the $`Ginga`$ data, with weaker QPO around 0.142 Hz. (Most likely, this other peak is the second harmonic.) Upper limits for the amplitude of QPO signals in the $`RXTE`$ observations were found by calculating the percent r.m.s. variation of the average excess power (after subtracting the fitted model) between 0.05 and 0.10 Hz. Typical 3$`\sigma `$ limits are $``$0.8% for the 2–60 keV PDS, as listed in Table 2. For comparison, the amplitudes of the primary and secondary $`Ginga`$ QPO peaks were 2.9% r.m.s. and 1.8% r.m.s., respectively. Hence, the power found in the $`Ginga`$ QPO was much larger than the upper limits placed on the $`RXTE`$ data. Clearly, the appearance of QPO in LMC X-1 is a transient phenomenon, as further documented by the lack measurable QPO in BBXRT observations of this source (Schlegel et al. 1994a, 1994b).
### 2.3 Spectral Fits
The X-ray energy spectrum of LMC X-1 was separately fit by two multi-component models over the energy range 3.6–15.9 keV. The soft thermal component was represented by both a blackbody and by a multicolor disk model (Mitsuda et al. 1984; Makishima et al. 1986), while the high-energy excess was always accounted for using a simple power law. Within limitations imposed by an uncertain background subtraction at low energies in the $`RXTE`$ data, both combinations yield acceptable fits. The intervening hydrogen absorption was fixed at $`N_H=0.6\times 10^{22}`$ cm<sup>-2</sup> in all cases, since the derivation of $`N_H`$ as a free parameter is very sensitive to errors in background subtraction. Mean values for the fitted parameters, which are similar to those of Ebisawa et al. (1989) and Schlegel et al. (1994a), are summarized in Table 3. Since the hardness ratio depends on source intensity, further modeling was performed on data selected by PCA count rate count (i.e., 130–140 cnts s<sup>-1</sup>, 140–150 cnts s<sup>-1</sup>, etc.). The results of this fitting for the blackbody plus power-law model are shown in Fig. 4. Within each observation, we find a trend for an increasing ‘scale’ of the power law (i.e., comparable to intensity provided the photon index is constant) with increasing source count rate (3.6–15.9 keV). The remaining parameters are nearly constant within a given observation, although a slight increase in temperature of the thermal component may be present at higher count rates. We note that the fitted parameters are highly coupled. If the temperature of the soft component were held constant, then the dependence of the power-law scale on count rate would become even more pronounced.
### 2.4 Long-term and Phase-related Variations
The All Sky Monitor (ASM) aboard $`RXTE`$ provides nearly continuous monitoring of the entire sky, producing long-term light curves for bright, persistent X-ray sources as well as discovering new transient sources. We have examined the LMC X-1 data in this public archive. They compare well with the mean count rates for our pointed observations, as presented in this paper. Overall, the source varied by about a factor of two through the nine months during which our observations were taken. Analysis of $``$3500 ASM data points taken over 76 weeks of observing does not reveal any significant periods in the range 0.5–20 days, including at the optically determined orbital period of 4.22 days. This result has been confirmed by scientists on the instrument team (Levine 1997).
In addition to searching for periods in the ASM observations, we have folded our $`RXTE`$ data on the 4.22-day orbital period. The phase coverage is poor because the observations were only obtained at nine discrete times. As it turned out, no observations were obtained between spectroscopic phases 0.9 to 1.3, leaving a large gap in the phase plots. From the folded data there is no clear indication of any trends related to the orbital period either in the count rate or the hardness ratio. Additional data obtained at more closely spaced intervals are need to determine if any small orbital modulation or other signature is present in the X-ray data.
## 3 SUMMARY
We obtained a series of $`RXTE`$ observations of the black-hole candidate LMC X-1 to search for quasi-periodic oscillations and examine their behavior with source intensity. No QPO were detected, although Ebisawa et al. had detected them in LMC X-1 when the source was in a bright state. Comparison of the source intensity during the $`RXTE`$ and Ginga observations suggests that LMC X-1 never became as bright during our $`RXTE`$ observations as when QPO were found in Ginga data. However, because of the uncertainty of the background subtraction with the $`RXTE`$ data, it is difficult to compare these data sets. In addition, no other periodic behavior was found, either on the orbital period or any other in the range of 0.5-20 days.
Fits of the X-ray spectral energy distribution require a two-component model, and the values derived are similar to those found by previous workers. A strong correlation was detected between the hardness ratio and the high-energy count rate, indicating that changes in the X-ray spectrum of LMC X-1 come from the high-energy tail.
We wish to thank the staff at the $`RXTE`$ Guest Observer Facility. We also acknowledge support from NASA.
|
no-problem/9812/astro-ph9812035.html
|
ar5iv
|
text
|
# Magnetic Field Evolution of Accreting Neutron Stars - III
## 1 Introduction
There is as yet no satisfactory theory for the generation of the neutron star magnetic field. There are two main possibilities - the field can either be a fossil remnant from the progenitor star, or be generated after the formation of the neutron star (for a review see Bhattacharya & Srinivasan 1995, Srinivasan 1995 and references therein). Whereas post-formation generation mechanisms give rise to fields supported entirely by crustal currents (Blandford, Hernquist & Applegate 1983), the fossil field resides in the core of a neutron star. In the core, the rotation of the star is supported by creation of Onsager-Feynman vortices in the neutron superfluid and the magnetic flux is sustained by Abrikosov fluxoids in the proton superconductor (Baym, Pethick & Pines 1969, Ruderman 1972, Bhattacharya & Srinivasan 1995).
Evidently no consensus regarding the theory of field evolution has been reached since this depends crucially upon the nature of the underlying current configuration. Observations, though, indicate that the magnetic field decays significantly only if the neutron star is in an interacting binary (Bailes 1989, Bhattacharya 1991, Taam & van den Heuvel 1986). There have been three major theoretical endeavours to link the field evolution with the binary history of the star - a) expulsion of the magnetic flux from the superconducting core during the phase of propeller spin-down, b) screening of the magnetic field by accreted matter and c) rapid ohmic decay of crustal magnetic field as a result of heating during accretion (see Bhattacharya 1995, 1996, Bhattacharya & Srinivasan 1995, Ruderman 1995 for detailed reviews).
Except screening the other two models of field evolution depend on ohmic decay of the underlying current loops for a permanent decrease in the field strength. Such ohmic dissipation is possible only if the current loops are situated in the crust where the electrical conductivity is finite. Models that assume an initial core-field configuration, therefore, require a phase of flux expulsion from the core. Muslimov & Tsygan (1985) and Sauls (1989) showed that there is likely to be a strong inter-pinning between the proton fluxoids and the neutron vortices. In a spinning down neutron star the neutron vortices migrate outward and by virtue of the inter-pinning drag the proton fluxoids along to the outer crust. Srinivasan et al. (1990) pointed out that neutron stars interacting with the companion’s wind would experience a major spin-down, causing the superconducting core to expel a large fraction of the magnetic flux. The nature of such flux expulsion as a result of spin-evolution has been investigated in detail for both isolated pulsars (undergoing pure dipole spin-down) and for the neutron stars that are members of binaries (Ding, Cheng & Chau 1993; Jahan Miri & Bhattacharya 1994; Jahan Miri 1996). Recently, Ruderman, Zhu & Chen (1998) has investigated the outward (inward) motion of core superfluid neutron vortices during spin-down (spin-up) of a neutron star which might alter the core’s magnetic field in detail.
Jahan Miri & Bhattacharya (1994) and Jahan Miri (1996) have investigated the link between the magnetic field evolution and the rotational history of a neutron star due to the interaction of its magnetosphere with the stellar wind of its companion. They assumed an uniform ohmic decay time-scale in the crust irrespective of the accretion rate. Later, Bhattacharya & Datta (1996) incorporated the crustal micro-physics into their calculation of the ohmic diffusion of an expelled field. However this work did not include the material movement that takes place in the crust as a result of accretion. Assuming that the field evolution starts only after the process of flux expulsion is over, in the present work we incorporate the material movement and investigate the evolution of an expelled field in the crust of an accreting neutron star.
## 2 Results and Discussions
Using the methodology developed by Konar & Bhattacharya (1997 - paper-I hereafter) we solve the induction equation for an initial flux just expelled at the core-crust boundary due to spin-down. As in previous studies we solve this equation by introducing the vector potential $`\stackrel{}{A}=(0,0,A_\varphi )`$, where $`A_\varphi =g(r,t)\mathrm{sin}\theta /r`$; $`(r,\theta ,\varphi )`$ being the spherical polar co-ordinates, assuming the field to be purely poloidal. For our calculation we have assumed an initial $`g`$-profile following the profile used by Bhattacharya & Datta (1996) plotted in figure . The sharply peaked nature of the $`g`$-profile is caused by the flux expelled from the entire core being deposited in a thin layer, thereby substantially increasing the local field strength. The evolution of the crustal currents depend on the following parameters :
* depth at which the currents are concentrated,
* width of the current distribution,
* impurity content of the crust,
* rate of accretion.
In the present work we assume the flux to be deposited at the bottom of the crust and therefore the depth of the initial current configurations is taken to be the thickness of the crust. The evolution of such flux is not very sensitive to the width of the current distribution (Bhattacharya & Datta 1996). Therefore we keep the width of current distribution fixed for all of our calculations. For details of the computation, crustal physics and binary parameters see Konar & Bhattacharya (1999 - paper-II hereafter). As in earlier papers, we denote the impurity strength by the parameter
$$Q\underset{i}{}\frac{n_i}{n}(Z_iZ)^2$$
where $`n`$ and $`Z`$ are the number density and charge of background ions in the pure lattice and $`n_i`$ and $`Z_i`$ those of the $`i`$-th impurity species. The sum extends over all species of impurities.
### 2.1 Field Evolution with Uniform Accretion
In figure we plot the evolution of the surface field for different values of the accretion rate. The field strengths go down by about only an order and a half in magnitude even for fairly large values of the impurity strength. The characteristic features of field evolution, with uniform accretion, are as follows.
1. An initial rapid decay (ignoring the early increase) is followed by a slow down and an eventual freezing.
2. The onset of ‘freezing’ is faster with higher rates of accretion.
3. Lower final ‘frozen’ fields are achieved for lower rates of accretion.
4. To achieve a significant reduction in the field strength, very large values of the impurity strength are required.
Hence, the general nature of field evolution in the case of an expelled flux is qualitatively similar to that in the case of an initial crustal flux (paper-I). However, to begin with, the expelled flux is deposited at the bottom of the crust. Therefore the initial rapid decay that is observed in case of crustal field is not as dramatic. Moreover, for higher rates of accretion the ‘freezing’ happens much earlier. For example, with an accretion rate of $`10^9`$ M yr<sup>-1</sup>the original crust is entirely assimilated into the core in about $`10^7`$ years. But the surface field levels off long before that since in this case the currents start returning to high density regions even before they have had time to spread out in the outermost regions of the crust. Due to the same reason large values of $`Q`$ do not change the final surface field much for higher rates of accretion, as for high enough temperatures and high densities the conductivity ceases to be sensitive to the impurity content of the crust (see paper-I for details).
### 2.2 Field Evolution in Binaries
In paper-II we have considered three phases of binary evolution, namely - the isolated, the wind and the Roche-contact phase. In the wind phase there are two distinct possibilities of interaction between the neutron star and its companion. If the system is in the ‘propeller phase’ then there is no mass accretion. But this phase is important because the star rapidly slows down to very long periods and as a result a significant flux-expulsion is achieved. From the point of view of flux expulsion, therefore, we assume the flux to be completely contained within the superconducting core (neglecting the small flux-expulsion caused by the dipole spin-down in the isolated phase) prior to this phase. The ohmic decay is assumed to take place only after this phase is over - that is in the phase of wind-accretion and in the phase of Roche-contact. In case of low mass X-ray binaries, it is not very clear as to how long the phase of wind-accretion lasts or whether such a phase is at all realized after the ‘propeller phase’ is over. Therefore, in our calculations we have considered cases with and without a phase of wind accretion.
#### 2.2.1 High Mass Binaries
Figure shows the evolution of the surface field in high mass X-ray binaries for different rates of accretion in the wind phase. The surface field shows an initial increase. This is followed by a sharp decay (of about an order of magnitude) only for sufficiently large rates of accretion in the wind phase. The decay in the Roche-contact phase is very small. In fact, the decay as seen in figure is interrupted by Roche-contact in which the currents are quickly pushed to the core thereby ‘freezing’ the field. This process also results in lower final field values for higher rates of accretion in the wind phase. Figure shows that the field indeed decays faster for higher rates of accretion to begin with but the ‘freezing’ happens earlier for higher rates thereby making the final saturation field lower for lower rates of accretion. Due to the short-lived nature of the wind-phase in massive binaries the ‘freezing’ takes place before this saturation field is attained, giving rise to a behaviour (lower final field strengths for higher rates of accretion) contrary to that seen in low-mass binaries. For accretion rates appropriate to high mass X-ray binaries the field decreases at most by an order of magnitude. This result is quite insensitive to the impurity strength of the crust because the time available for the evolution before the currents are pushed back into the core is rather small.
#### 2.2.2 Low Mass Binaries
Figure shows the evolution of the surface field in low mass X-ray binaries, for different values of the impurity concentration in the crust. It should be noted that a difference in the wind accretion rate does not manifest itself in either the nature of the field evolution or the final field strength. However, a difference in the accretion rate in the Roche-contact phase shows up very clearly. Comparing the different curves (for different values of the impurity parameter) we see that a large value of impurity strength gives rise to a rapid initial decay and therefore a lower value of the final surface field.
In figure we have plotted the evolution of the surface field assuming the wind accretion phase to be absent. Once again we find that for higher rates of accretion higher final field values are obtained. It should be noted here that the final field values obtained now are only about an order and a half of magnitude lower than the original surface field strengths. Even though the impurity strengths assumed now are much higher than those assumed for the field evolution in low mass X-ray binaries with a phase of wind-accretion. In absence of a prior phase of wind accretion the flux does not have enough time to diffuse out to low density regions when the Roche-contact is established. Therefore the role of accretion, in the Roche-contact phase, is mainly to push the currents towards the high density interior rather than to enhance ohmic decay rate. Evidently, much larger impurity strength is required for the final field values to decrease by three to four orders of magnitude. Unfortunately, our code is unable to handle high $`Q`$ at present because of prohibitive requirement of computer time to ensure stability. However, this figure clearly establishes a trend as to how the final field values behave with $`Q`$ and it is evident that we need $`Q`$ values much larger than those considered here to achieve millisecond pulsar field strengths in systems without a phase of wind accretion. Alternatively, systems without wind accretion might lead to high field pulsars in low-mass binaries, such as PSR 0820+02.
The most important point to note here is that similar to an initial crustal field configuration, the amount of field decay is much larger than that achieved in the case of high mass X-ray binaries. Although in low mass X-ray binaries the surface field does go down by three to four orders of magnitude from its original value for large values of impurity strength, the final field could remain fairly high if the impurity strength is small. If the wind-accretion phase is absent in these systems then to achieve large amount of field reduction even higher values of the impurity strength becomes necessary. Therefore, the ‘spin-down induced flux expulsion model’ will be consistent with the overall scenario of field evolution and in particular millisecond pulsars can be produced in low mass X-ray binaries provided the impurity strength in the crust of the neutron stars is assumed to be extremely large.
The results of our investigation (the present work and that described in paper-II) clearly indicate that both the models - assuming an initial crustal field or, alternatively a spin-down induced flux expulsion, place very stringent limits on the impurity strength. Whereas for the crustal model only $`Q\stackrel{<}{_{}}0.01`$ is allowed, the results described here show that much larger values of impurity are needed in the latter model. These requirements are quite different. If there is an independent way of estimating the impurity content of the crust then we could differentiate between these two models. However, in all of our investigation we have assumed that the impurity content of the crust does not change as a result of accretion, which may not be quite correct since accretion changes the crustal composition substantially (Ha͡ensel & Zdunik 1990).
## 3 conclusions
In this work we have investigated the consequences of ‘spin-down induced flux expulsion’. The general nature of field evolution seems to fit the overall scenario. The nature of field evolution is quite similar to that in the case of a purely crustal model of field evolution though the details differ. Most significantly, this model has the requirement of large values of the impurity strength $`Q`$ in direct contrast to the crustal model. To summarize then:
* The field values in the high mass X-ray binaries can remain fairly large for a moderate range of impurity strength.
* A reduction of three to four orders of magnitude in the field strength can be achieved in the low mass X-ray binaries provided the impurity strength is as large as 0.05.
* If the wind accretion phase is absent then to achieve millisecond pulsar field values, impurity strength in excess of unity is required.
|
no-problem/9812/astro-ph9812095.html
|
ar5iv
|
text
|
# 1 Mixing along the Red Giant Branch (RGB)
## 1 Mixing along the Red Giant Branch (RGB)
As a small mass star evolves up the RGB, the outer convective envelope expands inward and penetrates into the CN-cycle processed interior regions (first dredge-up). Some further mixing is possible in the latest phases of the RGB (see Charbonnel 1995: C95), when the advancing H-burning shell reaches the chemical discontinuity left behind by the receding convective envelope. Thereafter, no mean molecular weight gradient is present between the convective envelope and the near vicinity of the shell, and is possible that circulation currents give rise to further mixing.
Here we test this scenario presenting abundances for Fe, Li, C, N, O for about 60 metal-poor ($`2<`$\[Fe/H\]$`<1`$) field stars in different evolutionary stages. Details of the analysis will be given in a forthcoming paper. The main results of our study are presented in Figure 1, where abundances as a function of luminosity are shown.
There are two distinct mixing/dilution episodes along the RGB evolution of small mass field stars:
1) the first dredge-up follows canonical predictions (see e.g. C95). This is quite clear in the pattern shown by elements and isotopic ratios in panels a,b,c. The luminosity is about $`\mathrm{log}L/L_{}=1`$; Li is diluted by about a factor of 20, <sup>12</sup>C abundance decreases by $`0.1`$ dex.
2) a second mixing episode occurs when the star becomes brighter than the RGB bump (and the molecular weight barrier is canceled) at $`\mathrm{log}L/L_{}2`$, again in agreement with predictions by C95. Most of the remaining Li is destroyed, <sup>12</sup>C abundance further decreases by a factor $`2`$, and <sup>12</sup>C/<sup>13</sup>C ratio raises to $`6`$. These values are observed also in the following evolutionary phase (RHB and early AGB)
O and Na abundances in bright RGB and HB field stars are similar to those observed in unevolved stars. Field stars do not display any signature of the Na-O anticorrelation seen amongst globular cluster giants (see e.g. Kraft 1994).
|
no-problem/9812/nucl-th9812020.html
|
ar5iv
|
text
|
# References
THEF-NYM-98.02
Comment on “$`\pi NN`$ Coupling
from High Precision np Charge Exchange at 162 MeV”
M.C.M. Rentmeester, R.A.M. Klomp, and J.J. de Swart,
Institute for Theoretical Physics, University of Nijmegen,
Nijmegen, The Netherlands
Submitted to Phys.Rev.Letters on March 29, 1996,
Published in Phys.Rev.Letters 81, 5253 (1998)
ABSTRACT
In this updated and expanded version of our delayed Comment we show that the np backward cross section, as presented by the Uppsala group, is seriously flawed (more than 25 sd.). The main reason is the incorrect normalization of the data. We also show that their extrapolation method, used to determine the charged $`\pi NN`$ coupling constant, is a factor of about 10 less accurate than claimed by Ericson et al.. The large extrapolation error makes the determination of the coupling constant by the Uppsala group totally uninteresting.
PACS numbers: 13.75.Cs, 13.75.Gx, 21.30.-x
E-mail: swart@sci.kun.nl
In a not so recent anymore Letter, a measurement of the np differential cross-section in the backward direction at a single energy T<sub>lab</sub>=162 MeV was reported. These 31 data were then used to extract for the charged pion-nucleon coupling constant the large value $`f_c^2=0.0808`$. An incredibly small extrapolation error of only $`0.0003`$ and a normalization error of $`0.0017`$ are claimed. The systematic errors, however, have not been properly dealt with. This Uppsala value for the charged coupling constant is in agreement with the old-fashioned textbook values, but in strong disagreement with modern determinations.
First we make the observation that this coupling constant has been determined in the last decade by several groups (and not only the Nijmegen group as suggested in the Letter ) in various energy-dependent Partial Wave Analyses (PWA’s). For a review see reference . These PWA’s give very good fits to about 12,000 np , $`\pi `$N , and $`\overline{\mathrm{p}}`$p charge-exchange scattering data. The values of $`f_c^2`$ determined in these different energy-dependent PWA’s from thousands of data are all in excellent agreement with each other, and in flagrant disagreement with the determination from the merely 31 Uppsala data. A representative value (with error) for this coupling constant is $`f_c^2=0.0748(3)`$ . This charged coupling constant $`f_c`$ is via charge independence related to the pp$`\pi ^0`$ coupling constant $`f_p`$. This latter coupling constant has been determined from almost two thousand pp data with the result $`f_p^2=0.0753(5)`$. An incredibly large breaking of charge independence would therefore be implied if the Uppsala value of the charged coupling constant would be correct.
The backward np differential cross section is sensitive to $`f_c^2`$. That it is therefore a good place to determine this coupling constant is a widespread misunderstanding. This has been shown in an energy-dependent PWA of the np data. The backward np data do not show any particular sensitivity to $`f_c^2`$. In table V of one can see that in our energy-dependent PWA using all np scattering data and all types of observables, $`f_c^2`$ shows no special sensitivity to any particular type of observable. The claim in that “the Uppsala group has shown the contrary using pseudodata” is false. The Uppsala group did not study anything else but differential cross sections. Therefore they cannot make any statement about the relative importance of various observables (like differential cross sections, polarizations, spin transfer coefficients, etc.) in the determination of the coupling constant.
No particular sensitivity for any particular observable implies that all datatypes contribute with about the same weight. This means that the statistical errors in the different analyses are roughly inversely proportional to the square root of the number of datapoints. The pole extrapolations use about a factor of 100 less data than the energy dependent PWA’s, implying a statistical extrapolation error that is about 10 times larger than the error in the PWA’s. Such a large error makes the determination of $`f_c^2`$, as described in the Letter , totally uninteresting. It must be clear from statistical reasons that a rather small data set cannot be used for an accurate determination.
In the same paper it has been explicitly shown, using physical extrapolation techniques, that analyzing backward np data at a single energy, as in ref. , gives values of $`f_c^2`$ with a large spread that results in a total error of $`0.003`$, which is 10 times larger than the extrapolation error claimed in . This was confirmed by Arndt et al. , who used exactly the same techniques as used in for all the available backward data, and not for only one dataset as was done in . Their values for $`f_c^2`$ as determined at a single energy vary from 0.061 to 0.091 with an average of $`0.075`$ and an error of $`0.009`$, which is 30 times the extrapolation error quoted in .
In their Reply to our Comment the authors of the Letter imply that “the analysis of Arndt et al. is not detailed enough and their examination of the input data not critical enough”. That expresses exactly our opinion about the work of Ericson et al. as presented in and subsequent publications. From decades of experience with the work of Arndt we know that it definitely does not apply to the work of Arndt.
In the Letter the authors use a self-invented extrapolation method, which they call the Difference Method. However, they did not study properly the systematic errors in their new method. This was done in , where it was shown that the model-dependence of their method is enormous. This large model-dependence gives rise to very large systematic errors in their value for the coupling constant and in their estimate of the error.
The authors of the Letter state in the Abstract that they can extrapolate precisely and model-independently to the pion pole. That is definitely incorrect. Their extrapolation method is strongly model-dependent, with large systematic errors, and as inaccurate as any other extrapolation method. Not better, not worse. A new extrapolation method that really produces extrapolation errors a factor of 10 smaller than other extrapolation methods would have been a sensational discovery in numerical analysis and/or statistics. This Difference Method is definitely not better than the standard Chew extrapolation technique. However, it is certainly much more cumbersome. It is so cumbersome, that the Uppsala group could not properly determine their errors.
The pole-extrapolation method used by Ericson et al. relies heavily on the absolute normalization of the data. Normalizing np cross-sections is very difficult. In their determination of $`f_c^2`$ the normalization is another important source of uncertainty. In energy-dependent PWA’s, as in , one does not need normalized data to determine the coupling constant; one can use the shapes of the measured differential cross-sections.
Do not misinterpret the above statement. We do not say that we apply all our methods directly to unnormalized data. We normalize data very accurately with the help of our PWA. This has been explained extensively in most of our publications about PWA’s. This is definitely one of the successes of energy-dependent PWA’s; we determine the normalization of differential cross sections in np scattering with a typical uncertainty of about 0.5 %. This is a lot better than the 4 % normalization error used in the Letter. The remarks in the Reply about “loose normalizations” show an unfortunate lack of knowledge by the Uppsala group of the methods of modern PWA. The corresponding sentences in their Reply do not correspond to the truth, but are fabrications of the unbridled fantasies of the authors of the Reply.
The authors have applied their method for extraction of $`f_c^2`$ to data which cannot be described satisfactory by either the Nijmegen PWA or the VZ40 PWA of Arndt et al. . The Nijmegen PWA gives, after refitting, $`\chi ^2=264.0`$ for these 31 data and the VPI&SU PWA gives $`\chi ^2=236.7`$. One reason for the bad fit can be seen in the large discrepancy between the shape of the newly reported data and the shape of the older data of Bonner et al. at exactly the same energy. The authors should have reported $`f_c^2`$ from applying their extrapolation method to the Bonner data and compared the results.
In their Reply the authors claim: “the data of the present experiment are of a far better quality than those of Bonner at 162 MeV”. When reading this above quote one must realize here that people, not known for their familiarity with data analysis, are claiming that their own experiment is the best. This is definitely not the opinion of the Nijmegen data analysis group. They find in their careful, detailed, and critical analysis of the data that the old Bonner data are of better quality than the new Uppsala data. The Bonner data are included in the Nijmegen and the VPI&SU databases; the Uppsala data are not!
The new data disagree not only with the Bonner data, they disagree with the whole Nijmegen np data set, currently consisting of circa 5000 data below 500 MeV. They disagree, because of their wrong shape. The shape of the Uppsala differential cross section is more than 25 sd away from both the Nijmegen and the VPI&SU databases. More than 3 sd is already called “wrong”.
We have studied these data to see what is really wrong with them. In their experiment the Uppsala group performed 3 different measurements in 3 angular regions, which were then separately normalized. They have 49 (partially overlapping) datapoints. These 3 datasets are then combined to one dataset with 31 points. We have pinpointed two errors. Firstly, in those angular regions where these datasets overlap one clearly notices internal inconsistencies in the slopes. This discrepancy is nowhere discussed in the Uppsala papers, but just ignored. Secondly, we can improve the $`\chi ^2`$ for the total dataset dramatically by just renormalizing these 3 sets and discarding 4 datapoints . However, we cannot improve them so much that the data become acceptable. They are, after renormalization and discarding the 4 bad points, still more than 3 sd. away from the Nijmegen database. This is for statistical reasons unacceptable. But ….. , we made the Uppsala data at least almost acceptable.
In their Reply the authors refer to the Hürster et al. data , which are not used (but intensively studied) in the Nijmegen PWA’s and also not used in the Arndt et al. PWA’s because of their high $`\chi ^2`$. The authors claim that the shape of the incorrectly normalized Uppsala data agrees with the shape of these Hürster et al. data. This is almost certainly incorrect. Because then we need to assume that the Freiburg people made exactly the same mistakes as the Uppsala people in normalizing their data and that these data have the same kind of internal inconsistency. We see no reason to make such drastic assumptions. Also the $`\chi ^2`$/datapoint for the Hürster et al. data is much smaller than for the (incorrectly normalized) Uppsala data.
The authors state in their Reply that possibly the inclusion of the Bonner data in the Nijmegen and VPI&SU PWA’s is responsible for the large $`\chi ^2`$ of the Uppsala data. This also is incorrect. We have done PWA’s in which we discarded all Bonner data; the $`\chi ^2`$ of the Uppsala data was still unacceptably high. These and other studies performed in Nijmegen show that the Uppsala data are in disagreement with the whole database and not only with the Bonner data.
In their Reply to our Comment it is stated that: “Their Letter argues that a rather small, but well-controlled data set on a relevant observable can be used for an accurate determination when carefully analyzed.” We observe that the dataset is indeed rather small, only 31 points. This is insufficient for an accurate determination. According to the energy-dependent PWA’s the dataset is definitely not well-controlled. In 1993 it was already shown in that the backward differential cross section, combined with a pole-extrapolation method, is not an especially relevant observable. In the Reply one can read the unbelievable remark that our proof is not relevant to their approach. We find this an unprofessional way of discarding unwanted facts. We think that the statement “when carefully analyzed” is neither applicable to these incorrectly normalized data with internal inconsistencies, nor to their extrapolation analysis with their unnoticed, huge, systematic errors.
Our conclusions are:
i) The experimental data as presented are seriously flawed (more than 25 sd.). This is mainly caused by the way these data are normalized. Similar data at 96 MeV from the same group are not included in the Nijmegen database either because they too disagree significantly with the total dataset.
ii) Achieving an accurate determination of $`f_c^2`$ from the backward np data at one single energy is a rather unrealistic exercise. The label “dedicated” for such experiments is presumptuous and completely unwarranted. We have shown that to determine $`f_c^2`$ accurately the energy-dependent PWA’s are vastly superior over the pole-extrapolation methods.
We would like to thank Prof. N. Olsson for providing us the data, and R. Timmermans and T. Rijken for stimulating discussions.
|
no-problem/9812/cond-mat9812102.html
|
ar5iv
|
text
|
# abstract
## abstract
Nanocrystalline metals, i.e. metals with grain sizes from 5 to 50 nm, display technologically interesting properties, such as dramatically increased hardness, increasing with decreasing grain size. Due to the small grain size, direct atomic-scale simulations of plastic deformation of these materials are possible, as such a polycrystalline system can be modeled with the computational resources available today.
We present molecular dynamics simulations of nanocrystalline copper with grain sizes up to 13 nm. Two different deformation mechanisms are active, one is deformation through the motion of dislocations, the other is sliding in the grain boundaries. At the grain sizes studied here the latter dominates, leading to a softening as the grain size is reduced. This implies that there is an “optimal” grain size, where the hardness is maximal.
Since the grain boundaries participate actively in the deformation, it is interesting to study the effects of introducing impurity atoms in the grain boundaries. We study how silver atoms in the grain boundaries influence the mechanical properties of nanocrystalline copper.
## Introduction
In recent years, advanced production techniques have made it possible to create metals, alloys and ceramics with grain sizes down to 5 nm. For metals, this represents a reduction of the grain size by approximately four orders of magnitude compared to most conventionally produced metals. As can be expected, such a dramatic change in the microstructure leads to significant changes in the mechanical properties of the metals. For example, the hardness of typical nanocrystalline metals is far higher than what is seen in their coarse-grained counterparts \[1, 2, and references therein\].
Nanocrystalline metals are an attractive group of metals to model, as the small grain size provides a “cut-off” of the typical length scales, where structures appear during deformation. In coarse-grained materials structures appear on vastly different length scales, making it very difficult to model the properties of these materials. The models must include processes that occur on length scales from the sub-nanometer scale of the atomic processes in dislocation cores, to the micro- or even millimeter scale of grain and subgrain structures . Atomic-scale simulations of systems of these sizes are beyond the reach of even the most powerful of todays supercomputers. One is thus forced to split the system into sub-problems at different length-scales, and only treat sub-problems at the atomic scale with atomistic models. On coarser length scales other modeling paradigms must be used, such as Dislocation Dynamics and continuum plasticity calculations.
Dividing the problem into sub-problems at different length scales often results in a better understanding of the problem, as it draws attention to the structures which are relevant at a given length scale (for example, when studying the formation of dislocation structures it is clearly more relevant to focus on dislocations as the fundamental concept rather than on individual atoms). On the other hand, one becomes dependent on this understanding, when creating the coarse-grained models, as many assumptions about the relevant phenomena at different length-scales will by necessity be built into the multi-scale model. The low grain size of nanocrystalline metals “compresses” this range of length scales to a range, where the whole deformation problem can be modeled at the atomic scale, as many grains in the polycrystalline material can be handled in an atomic-scale simulation. This makes it possible to perform unbiased simulations of the deformation process, where no *a priori* assumptions are made about the deformation mechanisms.
In recent papers, we have presented simulations of the plastic deformation of nanocrystalline Cu and Pd . Other authors have presented simulations of the structure and elastic properties of nanocrystalline metals and semiconductors , and of the plastic deformation of Ni under constant stress loading . In this paper we review our simulations of the deformation mechanisms in nanocrystalline metals, and present simulations of the effects of impurities in the grain boundaries.
## Simulations of pure metals
### Simulation setup
We have attempted to generate three-dimensional systems with a microstructure similar to that observed experimentally in samples generated by inert gas condensation (IGC). The grains appear to be randomly oriented, approximately equiaxed, and dislocation free. The grain size distribution is close to log-normal \[1, 15, and references therein\]. We try to create systems that match this description.
The grains are produced using a three-dimensional Voronoi tesselation: random grain centers are chosen, and space is divided into regions in such a way that each region consists of the points in space closer to a given grain center than to any other grain center. Each region is then filled with a randomly oriented fcc lattice. An example of a two-dimensional Voronoi tessellation is shown in Figure 1.
The generated samples are annealed for 50 ps at 300 K to relax the grain boundary structure. We found that the annealing time and temperature are uncritical, but that the properties of the system are different if *no* annealing is done.
The interactions between the atoms are modeled using the Effective Medium Theory (EMT) . EMT is a many-body potential providing a realistic description of the metallic bonding, in particular in face-centered cubic (fcc) metals and alloys of fcc metals. Computationally, EMT is not much more demanding than pair potentials, but provide a significantly more realistic description of the metallic bonding.
The systems were deformed using a molecular dynamics (MD) procedure. A conventional MD simulation was performed, but at each timestep the atomic coordinates in the pulling direction (the $`z`$ coordinates) were rescaled in order to deform the system gradually. At the same time, the box dimensions in the transverse directions are allowed to shrink to keep the transversal components of the stress ($`\sigma _{xx}`$ and $`\sigma _{yy}`$) close to zero . The change in system dimensions is slow at the timescale of the simulation, the relative elongation is $`2.5\times 10^6`$ each timestep. With a timestep of 5 fs, this nevertheless results in a very high strain rate ($`\dot{\epsilon }=5\times 10^8s^1`$). The results presented here are not very sensitive to the strain rate, although some dependence is seen (typical stresses increase by 20% when the strain rate is changed from $`2.5\times 10^7s^1`$ to $`1\times 10^9s^1`$) .
### Results
Figure 2 shows the deformation of a typical sample with an average grain size of 5.21 nm. The atoms have been color coded according to the local crystal structure . White atoms are in local fcc order, and thus situated inside the grains. Light grey atoms are in local hexagonal close-packed (hcp) order, these atoms are at stacking faults. Atoms in all other local environments are colored dark grey. These are typically atoms at grain boundaries and in dislocation cores.
Some dislocation activity is seen in the system, as witnessed by the generation of stacking faults. The dislocation activity is not sufficient to account for the observed plastic deformation. A detailed analysis of the deformation shows that the main deformation mode is sliding in the grain boundaries .
As the volume fraction of atoms in the grain boundary increases with decreasing grain size, one would expect that increasing the grain size increases the strength of the material as long as the grain boundaries remain the carriers of the deformation. This is indeed what we see in the simulations. Figure 3 shows the stress-strain curves obtained from simulations of systems with various average grain sizes. A “reverse Hall-Petch effect”, i.e. a softening of the material with decreasing grain size, is observed.
### Discussion
The simulations of the deformation of nanocrystalline metals show a reverse Hall-Petch effect in Cu and Pd for the grain sizes studied. This softening of the material, when the grain size is reduced is caused by plastic deformation in the grain boundaries. There appears to be two different deformation mechanisms active at different grain sizes. In metals with the very small grain sizes studied here, the dominating deformation mechanism is sliding in the grain boundaries through a large number of essentially uncorrelated events, where a few atoms move with respect to each other at each event .
At much coarser grain sizes, dislocations are known to be the dominating carriers of deformation. In that regime, a hardening of the material is seen, when the grain size is reduced, as the grain boundaries act as barriers to the dislocation motion. Experimentally, this behavior is seen to continue far down into the nanocrystalline range.
As the grain size is reduced, dislocation-mediated deformation becomes more and more difficult. On the other hand, the volume fraction of the grain boundaries increases, favoring a deformation mechanism where the grain boundaries carry the deformation. Furthermore, as the grain size is approaching the grain boundary thickness, it becomes geometrically easier for slip to occur on more than one grain boundary, without large stress concentrations where the grain boundaries meet .
The emerging picture is one where two deformation mechanisms compete. One is active at very small grain sizes, another at larger grain sizes. This leads to a maximum in the yield stress and hardness of nanocrystalline metals at intermediate grain sizes, see Figure 4.
There have been many reports in the literature of a “reverse Hall-Petch effect” at sufficiently small grain sizes. However, the hardness measurements are very sensitive to sample defects, and in particular to sample porosity . In high-quality Cu samples, the Hall-Petch effect is seen to continue at least down to grain sizes around 15 nm .
There does not appear to be any *unequivocal* experimental evidence for a reverse Hall-Petch relationship in porosity free nanocrystalline metals. There are, however, indications of a break-down of the ordinary Hall-Petch relation at grain sizes below 15 nm in high quality copper samples produced by inert gas condensation . The ordinary Hall-Petch effect appears to cease, although a reverse Hall-Petch relation is not seen. It is still difficult to make direct comparisons between simulations and experiment, as there is little overlap in grain size. As the experimental techniques improve, there is hope that more experimental data will be gathered on high-quality samples with grain sizes below 10–15 nm. It should then become clear if the Hall-Petch effect does indeed break down at these grain sizes. At the same time, computer simulations of larger systems might lead to observations of the cross-over region between the reverse and the ordinary Hall-Petch effect. That cross-over region is beyond the reach of the largest simulations presented here. The grain boundaries remain the main carriers of the deformation even when the grain size is increased to 13.2 nm, but be *do* observe a slight increase in dislocation activity as the grain size is increased.
## The effects of alloying
As the main deformation mechanism is sliding in the grain boundaries, it could be expected that altering the structure and composition of the grain boundaries might have an effect on the properties of the material. One such modification is the addition of impurity atoms in the grain boundaries.
We have chosen to study the effects of silver impurities. A reason for choosing silver was that silver and copper are immiscible, low concentrations of silver impurities in nanocrystalline copper can therefore be expected to segregate to the grain boundaries. We have not studied the segregation process itself, as it is beyond the scope of this study. Segregation is a slow, diffusional process that cannot be studied directly with MD simulations due to the timescales involved. Other approaches (such as Monte Carlo or Kinetic Monte Carlo simulations) may be more appropriate.
Instead of simulating the segregation process, we have generated systems that model nanocrystalline copper after such a segregation has occurred. It is done by replacing 25% of the atoms in the grain boundaries with silver atoms, see Figure 5. The system is then annealed and deformed in the same way as the pure systems.
Figure 6 shows how the stress-strain curves have changed, when silver is introduced in the grain boundaries. The general trend seems to be a slight softening of the material, although the effect is very weak and not seen in all systems. Figure 7 shows the flow stress levels for the different simulations, i.e. the stress level at the horizontal part of the stress-strain curve (for simplicity the flow stress was defined as the average stress for $`6\%\epsilon 10\%`$). We again see the tendency for the silver-containing systems to be softer.
In the simulations of impurities in the grain boundaries, the mechanical behavior is close to what is seen in the pure systems.
As the major part of the deformation happens in the grain boundaries, one could expect that adding impurities in the grain boundaries could have a relatively large effect on the mechanical properties. The simulations presented here show that this is not the case, when silver is used as impurities in copper. We see a tendency towards a weakening of the material, but the effect is barely detectable.
The atomic bonding in copper and silver are of a similar nature, and the size of the atoms are not very different. This may account for the absence of a stronger effect of alloying in the grain boundaries.
## Conclusion
Atomic-scale simulations have been used to study the deformation mechanisms in nanocrystalline copper with and without impurities in the grain boundaries. In both cases, we find that the main deformation mode is sliding in the grain boundaries through a large number of apparently uncorrelated events, each involving only a few atoms. Some dislocation activity is seen in the grains, the dislocations are probably necessary to allow the grains to deform a little, as they glide past each other.
We observe a *reverse Hall-Petch effect*, i.e. a hardening of the material as the grain size is increased: as the amount of grain boundary atoms is decreased, the deformation becomes harder. At the same time the dislocation activity is seen to increase a little with grain size. This effect is seen in the entire range of grain sizes that we have studied (3 to 13 nm), but at some point we expect that dislocation motion will begin to dominate the behavior of the system. When that happens, the yield strength should begin to *decrease* with increasing grain size.
Adding silver to the grain boundaries has only a weak effect on the properties of the material. The strength of the material is seen to decrease marginally, when the silver is introduced. We expect that other elements, which chemically behave in a way that is more different from copper will have a larger effect, but such simulations remain to be made.
## Acknowledgments
This work was financed by The Danish Technical Research Council through Grant No. 9601119. Parallel computer time was financed by the Danish Research Councils through Grant No. 9501775. Center for Atomic-scale Materials Physics is sponsored by the Danish National Research Council.
|
no-problem/9812/cond-mat9812240.html
|
ar5iv
|
text
|
# Potential Scattering and the Kondo Effect
## Abstract
We study a generalized Kondo model in which a spin-$`\frac{1}{2}`$ impurity is coupled to a conduction band by both s-d exchange and potential interactions. A strong potential scattering is shown to screen an exchange scattering, and the Kondo temperature of the system is decreased.
It is well known that the behavior of a magnetic impurity in a metal can be described by an effective 1D Hamiltonian,
$$H=\underset{\sigma }{}\frac{dk}{2\pi }kc_\sigma ^{}(k)c_\sigma (k)+\frac{1}{2}I\underset{\sigma ,\sigma ^{}}{}\frac{dk}{2\pi }\frac{dk^{}}{2\pi }c_\sigma ^{}(k)\left(\stackrel{}{\sigma }_{\sigma \sigma ^{}}\stackrel{}{S}\right)c_\sigma ^{}(k^{})+V\underset{\sigma }{}\frac{dk}{2\pi }\frac{dϵ^{}}{2\pi }c_\sigma ^{}(k)c_\sigma (k^{}).$$
(1)
Here, $`\stackrel{}{\sigma }`$ are the Pauli matrices, $`\stackrel{}{S}`$ is the impurity spin operator, and $`I`$ is the s-d exchange coupling constant. The operators $`c_\sigma ^{}(k)`$ refer to conduction electrons with spin $`\sigma =,`$ in a $`s`$-wave state of momentum modulus $`k`$. The last term in Eq. (1) corresponds to a potential (spin independent) scattering of an electron on the impurity site, $`V`$ being the coupling constant. The electron energies and momenta in Eq. (1) are taken relative to the Fermi values, which are set to be equal to zero, while the Fermi velocity $`v_F=1`$. For simplicity, we confine ourselves to the case of $`S=1/2`$ which is of the most physical interest.
In the absence of a potential coupling, $`V=0`$, the model (1) is solved exactly by the Bethe ansatz (BA). In the BA approach to the theory of dilute magnetic alloys , the spectrum of a host is alternatively described in terms of charge and spin excitations rather than in terms of free particles with spin “up” and ‘down”. The BA equations for charge, $`k_j`$, $`j=1,\mathrm{},N`$, and spin $`\lambda _\alpha `$, $`\alpha =1,\mathrm{},M`$, rapidities describing $`N`$ particles on an interval of size $`L`$ have the form
$`\mathrm{exp}(ik_jL)\mathrm{\Phi }_{\text{ch}}`$ $`=`$ $`\left({\displaystyle \frac{\lambda _\alpha +\frac{i}{2}}{\lambda _\alpha \frac{i}{2}}}\right)^M`$ (3)
$`\left({\displaystyle \frac{\lambda _\alpha +\frac{i}{2}}{\lambda _\alpha \frac{i}{2}}}\right)^N\mathrm{\Phi }_{\text{sp}}(\lambda _\alpha )`$ $`=`$ $`{\displaystyle \underset{\beta =1}{\overset{M}{}}}{\displaystyle \frac{\lambda _\alpha \lambda _\beta +i}{\lambda _\alpha \lambda _\beta i}}`$ (4)
where $`M`$ is the number of particles with spin “down”. The eigenenergy $`E`$ and the $`z`$ component of the total spin of the system $`S^z`$ are found to be
$$E=\underset{j=1}{\overset{N}{}}k_j,S^z=\frac{1}{2}+\frac{N}{2}M.$$
(5)
In Eqs. (2) the phase factors
$`\mathrm{\Phi }_{\text{ch}}`$ $`=`$ $`{\displaystyle \frac{1\frac{i}{2}U_0}{1+\frac{i}{2}U_0}}\mathrm{exp}(iU_0)`$ (7)
$`\mathrm{\Phi }_{\text{sp}}(\lambda )`$ $`=`$ $`{\displaystyle \frac{\lambda +\frac{1}{\text{g}_0}+\frac{i}{2}}{\lambda +\frac{1}{\text{g}_0}\frac{i}{2}}}`$ (8)
describe the scattering of charge and spin excitation of the host on the impurity. Here
$$U_0=\frac{1}{4}I,\text{g}_0=\frac{1}{2}I.$$
(9)
In the absence of the exchange coupling, $`I=0`$ or $`S=0`$, multiparticle effects are also absent. The model (1) is diagonalized in terms of independent particles with spin “up” and “down” scattering on the impurity potential. Pure potential impurities are clear to change physical properties of a host only if they significantly modify the density of band states near the Fermi level. Therefore, dramatic changes in the thermodynamics of a host under a doping with magnetic impurities is associated with an s-d exchange coupling only, while the potential scattering term is assumed can be omitted.
Here, we note that not changing the structure of the BA equations (2) a potential scattering renormalizes the parameters $`U_0`$ and $`\text{g}_0`$. In the presence of potential scattering these parameters are replaced, respectively, by
$`U`$ $`=`$ $`U_0+V`$ (11)
g $`=`$ $`{\displaystyle \frac{\text{g}_0}{1+\frac{1}{4}(V+\frac{1}{4}I)(V\frac{3}{4}I)}}.`$ (12)
At $`V=0`$, Eqs. (4) reduce to expressions given in Eq. (3c), provided that, as it is assumed in deriving Eqs. (2), $`I1`$. In the domain $`I|V|1`$, the impurity term in Eq. (2a) for charge excitation is determined by potential scattering, $`UV`$, while scattering of spin excitations at the impurity is still determined by exchange coupling, $`\text{g}\frac{1}{2}I`$. Finally, at $`|V|1`$, the effective coupling constant is given by $`\text{g}=2I/V^2=4\text{g}_0/V^2`$, and hence the exchange scattering is screened by the potential one.
Thus, a strong potential scattering does not affect the qualitative behavior of the system, but it renormalizes physical parameters of the Kondo effect. In particular, the Kondo temperature is now found to be
$$T_Kϵ_F\mathrm{exp}\left(\frac{\pi }{\text{g}}\right)=ϵ_F\mathrm{exp}\left(\frac{\pi V^2}{\text{2I}}\right),$$
(13)
where $`ϵ_F`$ is the Fermi energy, while in the absence of potential scattering $`T_K^0ϵ_F\mathrm{exp}(2\pi /I)`$.
To derive Eqs. (4), it is enough to study how a potential scattering is incorporated in the particle-impurity scattering matrix. Let us look for one-particle eigenstates of the system in the form
$$|\mathrm{\Psi }_1=\underset{\sigma }{}\underset{s=0,1}{}\frac{dk}{2\pi }\psi _{\sigma ;s}(k)c_\sigma ^{}(k)\left(S^+\right)^s|0,$$
where the vacuum state $`|0`$ contains no electrons and the impurity spin is assumed to be “down”. The Schrödinger equation is then easily found to be
$`(k\omega )\psi (k)+{\displaystyle \frac{1}{2}}I\left(\stackrel{}{\sigma }\stackrel{}{S}\right)A+VA=0`$ (15)
$`A={\displaystyle \frac{dk}{2\pi }\psi (k)},`$ (16)
where $`\omega `$ is the eigenenergy, and spin indexes are omitted.
Inserting the general solution of Eq. (6a),
$$\psi (k)=2\pi \delta (k\omega )\chi \frac{V}{k\omega i0}A\frac{\frac{1}{2}I}{k\omega i0}\left(\stackrel{}{\sigma }\stackrel{}{S}\right)A,$$
with an arbitrary spinor $`\chi `$, into Eq. (6b), one obtains
$$\left\{1+\frac{i}{2}\left[V+\frac{1}{2}I\left(\stackrel{}{\sigma }\stackrel{}{S}\right)\right]\right\}A=\chi .$$
For the Fourier image of the function $`\psi (k)`$,
$$\psi (x)=_{\mathrm{}}^{\mathrm{}}\frac{dk}{2\pi }\psi (k)\mathrm{exp}(ikx),$$
a solution of Eqs. (6) is then found to be
$$\psi (x)=e^{ikx}\{\begin{array}{cc}\hfill \chi ,& x<0\hfill \\ \hfill 𝐑\chi ,& x>0\hfill \end{array}$$
where $`k=\omega `$. Here, the electron-impurity scattering matrix is given by
$$𝐑=u+2v\left(\stackrel{}{\sigma }\stackrel{}{S}\right),$$
(18)
where the parameters $`u`$ and $`v`$ are found from the equations
$`u+v`$ $`=`$ $`{\displaystyle \frac{1\frac{i}{2}(V+\frac{1}{4}I)}{1+\frac{i}{2}(V+\frac{1}{4}I)}}`$ (19)
$`u3v`$ $`=`$ $`{\displaystyle \frac{1\frac{i}{2}(V\frac{3}{4}I)}{1+\frac{i}{2}(V\frac{3}{4}I)}}.`$ (20)
Diagonalization of the system with the particle-impurity scattering matrix $`𝐑`$ results in the BA equations (2) with scattering parameters given in Eqs. (4).
A screening of the exchange scattering and a lowering the Kondo temperature by quite a strong potential scattering, $`|V|I`$, find a natural physical explanation. In the absence of the potential scattering, multiparticle effects in a spin subsystem of a metal doped with a Kondo impurities are generated due to an essential difference of electron-impurity scattering amplitudes in the triplet, $`u+v`$, and singlet, $`u3v`$, channels of scattering. As $`V`$ grows, this difference in scattering amplitudes is easily seen from Eqs. (7) to be decreased and become negligible small at $`|V|I`$. Therefore, the impurity term $`\mathrm{\Phi }_{\text{sp}}`$ disappears from Eq. (2b) describing the behavior of the spin subsystem, and the impurity is decoupled from the band states.
|
no-problem/9812/cond-mat9812272.html
|
ar5iv
|
text
|
# Single fluxon in double stacked Josephson junctions: Analytic solution.
## Abstract
We derive an approximate analytic solution for a single fluxon in a double stacked Josephson junctions (SJJ’s) for arbitrary junction parameters and coupling strengths. It is shown that the fluxon in a double SJJ’s can be characterized by two components, with different Swihart velocities and Josephson penetration depths. Using the perturbation theory we find the second order correction to the solution and analyze its accuracy. Comparison with direct numerical simulations shows a quantitative agreement between exact and approximate analytic solutions. It is shown that due to the presence of two components, the fluxon in SJJ’s may have an unusual shape with an inverted magnetic field in the second junction when the velocity of the fluxon is approaching the lower Swihart velocity.
Magnetic field, $`H`$, parallel to layers, penetrates stacked Josephson junctions (SJJ’s) in the form of fluxons. Fluxons in SJJ’s are different both from Abricosov vortices in type-II superconductors, since they don’t have a normal core, and from Josephson vortices in a single Josephson junction (JJ’s), since magnetic field is spread over many JJ’s. Despite of a large body of works concerning vortex related properties of layered superconductors, for review see e.g. Ref. , exact solution for a single fluxon in SJJ’s is still missing. Distribution of magnetic induction, $`B`$, at large distances from a fluxon was studied in Refs. and , for tightly packed stacks with identical and nonidentical layers, respectively. In Ref. a solution for weekly coupled double SJJ’s was obtained using perturbation approach. However, this solution is not valid for strongly coupled SJJ’s, which is the most interesting case. The electromagnetic coupling is strong when interlayer space periodicity of the stack, $`s`$, is (much) less than the effective London penetration depth, $`\lambda _{ab}`$, as it is the case for high-$`T_c`$ superconductors.
Recently, a simple approximate analytic solution for double SJJ’s was suggested. The approximate solution was shown to be in a good agreement with numerical simulations in the static case for arbitrary coupling strengths and parameters of the stack. However, the accuracy of the solution was not strictly proven. Another question which is still opened is how the shape of the fluxon is changed in the dynamic case. This is a crucial question since fluxon dynamics determines perpendicular ($`c`$-axis) transport properties of SJJ’s.
In this paper we rigorously derive the approximate analytic solution for a single fluxon in a double SJJ’s, which is valid for arbitrary junction parameters and coupling strengths. Using the perturbation theory we find the second order correction to the solution and analyze its accuracy. Comparison with direct numerical simulations shows a quantitative agreement between exact and approximate analytic solutions. It is shown that the fluxon in a double SJJ’s can be characterized by two components, with different Swihart velocities and Josephson penetration depths. We also studied the transformation of the fluxon shape with increasing propagation velocities. It is shown that due to the presence of two components, the fluxon in SJJ’s in the dynamic case may have unusual shape with an inverted magnetic field in the second junction at high propagation velocities.
We consider a double SJJ’s with the overlap geometry, consisting of JJ’s 1 and 2 with the following parameters: $`J_{ci}`$ -the critical current density, $`t_i`$\- the thickness of the tunnel barrier between the layers, $`d_i`$ and $`\lambda _{Si}`$ \- the thickness and London penetration depth of superconducting layers. Hereafter the subscript $`i`$ on a quantity represents its number. The strength of electromagnetic coupling of junctions is determined by the coupling parameter, $`S`$, which varies from 0 to 0.5 for the stack with identical JJ’s. More details about definitions can be found in Ref. . We consider frictionless fluxon motion with a constant velocity, $`u`$. The fluxon will be placed in junction 1.
The physical properties of SJJ’s are described by coupled sine-Gordon equation (CSGE) for gauge invariant phase differences, $`\phi _i`$. The problem with solving CSGE is the coupling of nonlinear $`sin\left(\phi _i\right)`$ terms. To decouple the variables, we take linear combination of equation in CSGE and rewrite it as:
$$\stackrel{~}{\lambda }_{1,2}^2F_{1,2\xi \xi }^{\prime \prime }sin\left(F_{1,2}\right)=Er_{1,2}\left(\xi \right),$$
(1)
where $`\xi =xut`$ is the self-coordinate of the fluxon,
$$F_{1,2}=\phi _1\kappa _{2,1}\phi _2,$$
(2)
$`Er_{1,2}`$ $`=`$ $`sin(\phi _1)\kappa _{2,1}sin(\phi _2)sin(\phi _1\kappa _{2,1}\phi _2)`$ (3)
$``$ $`\kappa _{2,1}\phi _2\left(1cos(\phi _1)\right)+O\left(\phi _2^3\right).`$ (4)
Here $`\stackrel{~}{\lambda }_{1,2}^2=\lambda _{1,2}^2(1u^2/c_{1,2}^2)`$ and coefficients $`\kappa _{1,2}`$, characteristic Josephson penetration depths, $`\lambda _{1,2}`$, and characteristic Swihart velocities, $`c_{1,2}`$ are given by Eqs.(17,20,21) from Ref.. Such choice for coefficients $`\kappa _{1,2}`$ minimizes $`Er_{1,2}`$ far from the fluxon center. The phase differences should satisfy boundary conditions:
$`\phi _1(\mathrm{})`$ $`=`$ $`0,\phi _1(0)=\pi ,\phi _1(+\mathrm{})=2\pi ;`$ (5)
$`\phi _2(\pm \mathrm{},0)`$ $`=`$ $`0.`$ (6)
Therefore, functions $`Er_{1,2}`$ have a form of ripple around zero value and will be considered as perturbation. Solutions of Eq.(1) can be easily found. Solutions of uniform Eq.(1), i.e. with zero r.h.s., are:
$$F_{1,2}=4arctan\left[exp\left(\xi /\stackrel{~}{\lambda }_{1,2}\right)\right],$$
(7)
From Eq.(2) we obtain the first approximation for phase differences:
$`\phi _1`$ $`=`$ $`{\displaystyle \frac{\kappa _1F_1\kappa _2F_2}{\kappa _1\kappa _2}},`$ ()
$`\phi _2`$ $`=`$ $`{\displaystyle \frac{F_1F_2}{\kappa _1\kappa _2}},`$ ()
which coincides with the approximate analytic solution, obtained in Ref. . The approximate solution is asymptotically correct at large distances from the fluxon center and has correct values at $`x=0`$, as follows from Eqs. (3,4).
Next, we look for a solution of nonuniform Eq.(1) in the form $`F_{1,2}=F_{1,2}+\delta F_{1,2}`$, where $`\delta F_{1,2}`$ are corrections due to perturbation terms $`Er_{1,2}`$:
$$\left(\stackrel{~}{\lambda }_{1,2}^2\frac{d^2}{d\xi ^2}1+\frac{2}{cosh^2\left(\xi /\stackrel{~}{\lambda }_{1,2}\right)}\right)\delta F_{1,2}=Er_{1,2},$$
(8)
with boundary conditions $`\delta F_{1,2}(\pm \mathrm{},0)=0`$. The solution of Eq. (7) is
$$\delta F_{1,2}=a_{1,2}(\xi )f_{1,2}+b_{1,2}(\xi )g_{1,2},$$
(9)
where
$`f_{1,2}`$ $`=`$ $`1/cosh(\xi /\stackrel{~}{\lambda }_{1,2}),`$ (10)
$`g_{1,2}`$ $`=`$ $`sinh\left({\displaystyle \frac{\xi }{\stackrel{~}{\lambda }_{1,2}}}\right)+{\displaystyle \frac{\xi }{\stackrel{~}{\lambda }_{1,2}cosh(\xi /\stackrel{~}{\lambda }_{1,2})}},`$ ()
are partial solutions of the uniform Eq.(7) and
$`a_{1,2}`$ $`=`$ $`{\displaystyle \frac{1}{2\stackrel{~}{\lambda }_{1,2}}}{\displaystyle _0^\xi }Er_{1,2}\left(x^{}\right)g_{1,2}\left(x^{}\right)𝑑x^{},`$ (11)
$`b_{1,2}`$ $`=`$ $`{\displaystyle \frac{1}{2\stackrel{~}{\lambda }_{1,2}}}{\displaystyle _{\mathrm{}}^\xi }Er_{1,2}\left(x^{}\right)f_{1,2}\left(x^{}\right)𝑑x^{}.`$ ()
The perturbation corrections to the approximate fluxon solution, Eq.(6) are
$`\delta \phi _1`$ $`=`$ $`{\displaystyle \frac{\kappa _1\delta F_1\kappa _2\delta F_2}{\kappa _1\kappa _2}},`$ ()
$`\delta \phi _2`$ $`=`$ $`{\displaystyle \frac{\delta F_1\delta F_2}{\kappa _1\kappa _2}}.`$ ()
The corrections $`\delta \phi _{1,2}`$ are used to improve the approximate solution Eq.(6) and to estimate it’s accuracy.
Lets observe that from Eq. (3), $`\kappa _1Er_1\kappa _2Er_2`$. From Eqs.(9-11) it can be seen that for $`\stackrel{~}{\lambda }_2=\stackrel{~}{\lambda }_1`$, $`\kappa _1\delta F_1=\kappa _2\delta F_2`$ and $`\delta \phi _10`$ for arbitrary $`\phi _20`$. Moreover, taking into account Eq.(6 b), it can be shown that $`d(\delta \phi _1)/d\stackrel{~}{\lambda }_{2(\stackrel{~}{\lambda }_2=\stackrel{~}{\lambda }_1)}=0`$.
In Fig. 1, central part of a fluxon is shown for a stack of two identical JJ’s with strong coupling, $`S`$=0.495, for the static case, $`u`$=0. Phase distribution in the full scale is shown in Fig. 3. Parameters of the stack are: $`d_{13}=t_{1,2}=0.01\lambda _{J1}`$, $`\lambda _{S13}=0.1\lambda _{J1}`$, where $`\lambda _{J1}`$ is the Josephson penetration depth of the single junction 1. Solid lines represent results of direct numerical integration of CSGE, Eq. (1), dashed lines show the approximate analytic solution, Eq.(6), and dotted line shows the corrected analytic solution, $`\phi _{2a}+\delta \phi _{2a}`$, Eqs.(6,11), in junction 2. Solid and dashed curves, marked as $`\delta \phi _{1,2}`$, represent the overall discrepancy between numerical and approximate analytic solutions and the perturbation correction, Eq.(11), respectively.
From Fig. 1 it is seen that correction to the fluxon image, $`\phi _2`$, in the second junction vanishes far from and in the fluxon center, while the accuracy decreases at distances $`\lambda _{J1}`$ from the center. Such behavior is expected from the shape of perturbation functions $`Er_{1,2}`$ in Eq. (1). On the other hand, for junction 1, correction $`\delta \phi _1`$ is small in the whole space region and analytic solution gives an excellent fit to the ”exact” numerical solution, in agreement with discussion above. The most crucial test for the approximate solution is the accuracy of derivative at $`x=0`$
$$\frac{\delta \phi _1^{}(0)}{\phi _1^{}(0)}=\frac{\kappa _1b_1(0)/\stackrel{~}{\lambda }_1\kappa _2b_2(0)/\stackrel{~}{\lambda }_2}{\kappa _1/\stackrel{~}{\lambda }_1\kappa _2/\stackrel{~}{\lambda }_2}.$$
(12)
An estimation for $`u=0`$ yields
$$\frac{\delta \phi _1^{}(0)}{\phi _1^{}(0)}\frac{2\alpha \kappa _1\kappa _2\left(\stackrel{~}{\lambda }_1\stackrel{~}{\lambda }_2\right)^2\left(\stackrel{~}{\lambda }_1^1+\stackrel{~}{\lambda }_2^1+\frac{\lambda _0}{\stackrel{~}{\lambda }_1\stackrel{~}{\lambda }_2}\right)}{\left(\kappa _1\kappa _2\right)^2\left(\stackrel{~}{\lambda }_1+\stackrel{~}{\lambda }_2+\frac{2\stackrel{~}{\lambda }_1\stackrel{~}{\lambda }_2}{\lambda _0}\right)\left(1+\frac{\stackrel{~}{\lambda }_1}{\lambda _0}\right)\left(1+\frac{\stackrel{~}{\lambda }_2}{\lambda _0}\right)},$$
(13)
where $`\alpha `$ is a factor of the order of unity and $`\lambda _0`$ given by Eq.(26) from Ref.. From Eq. (13) it is seen that both $`\delta \phi _1^{}(0)/\phi _1^{}(0)`$ and it’s derivative with respect to $`\stackrel{~}{\lambda }_2`$ goes to zero at $`\stackrel{~}{\lambda }_2=\stackrel{~}{\lambda }_1`$, as discussed above.
In Fig. 2, the maximum of $`\delta \phi _1\left(x\right)`$ (top panel) and the relative correction to derivative at $`x=0`$, $`\delta \phi _1^{}(0)/\phi _1^{}(0)`$, (bottom panel) for $`u`$=0 are shown as a function of $`J_{c2}/J_{c1}`$ for four different coupling parameters $`S`$=0.495 (solid lines), $`S`$=0.433 (dashed lines), $`S`$=0.312 (dotted lines) and $`S`$=0.127 (dashed-dotted lines). For $`S`$=0.495 parameters of the stack are the same as in Fig.1; for $`S`$=0.433 $`d_i=0.5\lambda _{Si}`$; for $`S`$=0.312 $`d_i=\lambda _{Si}`$; and for $`S`$=0.127 $`d_i=2\lambda _{Si}`$. From Fig. 2 it is seen that the accuracy of solution improves with decreasing $`S`$. This is naturally explained by a decrease of $`\phi _2`$, see Eq.(3) and decrease of splitting between $`\lambda _{1,2}`$, see Eq. (20) from Ref.. However, even for strongly coupled case, the analytic solution, Eq.(6), gives quantitatively good approximation not only for the value, but also for the derivative of $`\phi _1`$ for arbitrary parameters of the stack. The gray solid line in the bottom panel of Fig. 2 shows the estimation for $`\delta \phi _1^{}(0)/\phi _1^{}(0)`$, calculated from Eq.(13) for $`S`$=0.495. It is seen, that estimation gives qualitatively correct result. Namely, $`\delta \phi _1^{}(0)/\phi _1^{}(0)`$ vanishes both for $`J_{c2}/J_{c1}0`$, as $`J_{c2}/J_{c1}`$, and for $`J_{c2}/J_{c1}\mathrm{}`$, as $`\sqrt{J_{c1}/J_{c2}}`$. Note, that for $`J_{c2}/J_{c1}0`$, $`\delta \phi _1`$ vanishes even though the splitting of $`\lambda _{1,2}`$ becomes extremely large.
Sofar we have considered the static case, $`u`$=0. On the other hand, according to Eq. (6), radical changes should take place in the dynamic state, and the shape of the fluxon in SJJ’s may become qualitatively different from that in the single Josephson junction. Indeed, as the velocity approaches the lower characteristic velocity, $`uc_1`$, $`\stackrel{~}{\lambda }_10`$, i.e. the $`F_1`$ component of the fluxon contracts, while contraction of the $`F_2`$ component remain marginal. This implies, that at $`uc_1`$, the fluxon in SJJ’s consists of a contracted core and uncontracted ”tails” decaying at distances much larger than the core size. Such behavior is clearly different from that in a single Josephson junction. We note that characteristic velocities, $`c_{1,2}`$ may depend on $`u`$, therefore, contraction of each component, $`F_{1,2}`$, may be different from Lorentz contraction. Another interesting consequence of the approximate analytic solution, Eq.(6), is that with increasing $`u`$, the magnetic field in the second junction may change the sign with respect to that in junction 1. Such behavior was predicted in Ref. from the approximate analytic solution, Eq. (6), and it was suggested, that this will result in attractive fluxon interaction in SJJ’s. The magnetic induction in SJJ’s is equal to
$`B_1`$ $`=`$ $`{\displaystyle \frac{H_0\lambda _{J1}}{2\left(1S^2\right)}}\left[\phi _1^{}+S\sqrt{{\displaystyle \frac{\mathrm{\Lambda }_2}{\mathrm{\Lambda }_1}}}\phi _2^{}\right],`$ ()
$`B_2`$ $`=`$ $`{\displaystyle \frac{H_0\lambda _{J1}}{2\left(1S^2\right)}}\left[S\sqrt{{\displaystyle \frac{\mathrm{\Lambda }_1}{\mathrm{\Lambda }_2}}}\phi _1^{}+{\displaystyle \frac{\mathrm{\Lambda }_1}{\mathrm{\Lambda }_2}}\phi _2^{}\right],`$ ()
where $`H_0=\frac{\mathrm{\Phi }_0}{\pi \lambda _{J1}\mathrm{\Lambda }_1}`$ and $`\mathrm{\Lambda }_{1,2}`$ are defined in Ref. .
From Eq.(14) it is seen that estimation of magnetic induction in SJJ’s requires the accuracy of derivatives in both junctions, while Eq.(6) is not valid with the accuracy of derivative at $`x=0`$ for $`\phi _2`$. Therefore, more elaborate analysis is needed for the study of magnetic field distributions in the fluxon.
In Fig. 3, we show a) phase distributions and b) magnetic field distributions in the fluxon for different fluxon velocities. Parameters of the stack are the same as in Fig. 1. Solid lines in Fig. 3 a) represent results of direct numerical simulations of CSGE, Eq.(1), and dotted lines show the approximate analytic solution, Eq.(6). It is seen that quantitative agreement between ”exact” and approximate solutions sustain up to $`c_1`$. For the case of identical junctions, considered here, exactly one half of the fluxon belongs to each of the components, $`F_{1,2}`$. Indeed, from Fig. 3 a) it is seen that for $`uc_1`$ there is a contracted core at $`x=0`$ with a one $`\pi `$ step in $`\phi _1`$. On both sides of the core, there are two $`\pi /2`$ tails, which are slowly decaying at distances $`\stackrel{~}{\lambda }_2\stackrel{~}{\lambda }_1`$. Solid and dashed curves in Fig. 3 b) represent numerically simulated profiles, $`B_{1,2}\left(x\right)`$, in junctions 1 and 2, respectively. From Fig. 3 b) it is clearly seen, that with increasing fluxon velocity, a dip in $`B_2`$ develops in the center of the fluxon. At velocities close to $`c_1`$, $`B_2\left(0\right)`$ changes sign and finally at $`u=c_1`$, $`B_2\left(0\right)=B_1\left(0\right)`$. Such behavior is in agreement with predictions of Ref. . From numerical simulations we have found, that the fluxon shape in double SJJ’s is well described by Eq.(6) up to at least $`u0.98c_1`$ for any reasonable parameters of the stack, although the accuracy of the approximate solution may decrease with increasing $`u`$. The decrease of accuracy is caused by the increase of $`\phi _2`$ as is seen from Fig. 3. In this case, perturbation correction, Eq. (11), should be taken into account.
In conclusion, a simple approximate analytic solution for a single fluxon in a double stacked Josephson junctions for arbitrary junction parameters and coupling strengths is derived. It is shown that the fluxon in a double SJJs can be characterized by two components, with different Swihart velocities and Josephson penetration depths. Using the perturbation theory we find the second order correction to the solution and analyze its accuracy. Comparison with direct numerical simulations shows a quantitative agreement between exact and approximate analytic solutions for all studied parameters of the stack and fluxon velocities up to at least 0.98$`c_1`$. It is shown that due to the presence of two components, the fluxon in SJJ’s may have an unusual shape with an inverted magnetic field in the second junction at large propagation velocities. This may lead to attractive fluxon interaction in the dynamic state of SJJ’s
Discussions with D.Winkler are gratefully acknowledged. The work was supported by the Russian Foundation for Basic Research under Grant No. 96-02-19319.
|
no-problem/9812/cond-mat9812190.html
|
ar5iv
|
text
|
# Bulk Tunneling at Integer Quantum Hall Transitions
\[
## Abstract
The tunneling into the bulk of a 2D electron system (2DES) in strong magnetic field is studied near the integer quantum Hall transitions. We present a nonperturbative calculation of the tunneling density of states (TDOS) for both Coulomb and short-ranged electron-electron interactions. In the case of Coulomb interaction, the TDOS exhibits a 2D quantum Coulomb gap behavior, $`\nu (\epsilon )=C_Q|\epsilon |/e^4`$, with $`C_Q`$ a nonuniversal coefficient of quantum mechanical origin. For short-ranged interactions, we find that the TDOS at low bias follows $`\nu (\epsilon )/\nu (0)=1+(|\epsilon |/\epsilon _0)^\gamma `$, where $`\gamma `$ is a universal exponent determined by the scaling dimension of short-ranged interactions.
PACS numbers: 73.50.Jt, 05.30.-d, 74.20.-z
\] The integer quantum Hall transition (IQHT) refers to the continuous, zero-temperature phase transition between two consecutive integer quantum Hall states in a 2DES . It happens when the Fermi energy of the 2DES moves across a critical energy located near the center of each disorder-broadened Landau level. On both sides of the transition, while the Hall conductivity $`\sigma _{xy}`$ is integer-quantized, the dissipative conductivity $`\sigma _{xx}`$ vanishes at low temperatures. At the transition, however, both $`\sigma _{xx}`$ and $`\sigma _{xy}`$ remain finite, supporting a critical, conducting state in two dimensions.
Theoretically, the IQHT has been studied extensively in terms of its important, noninteracting analog in which electron-electron interactions are ignored . The extend to which this disordered, free electron model describes the IQHT in real materials depends on the effects of electronic interactions. Recently, the stability of the noninteracting theory has been analyzed by calculations of the scaling dimensions of the interactions . It was found that short-ranged interactions are irrelevant in the renormalization group sense, but the long-ranged $`1/r`$-Coulomb interaction is a relevant perturbation. In the latter case, the noninteracting theory becomes unstable, and the universality class (e.g. the critical exponents) of the transition is expected to change based on the standard theory of critical phenomena. The experimentally measured value of the dynamical scaling exponent, $`z=1`$, is consistent with Coulomb interaction being relevant at the transitions . However, the precise role played by Coulomb interaction and the mechanism by which it governs the scaling behavior have not been understood.
The tunneling density of states (TDOS) is a simple and useful probe of the nature and the effects of electronic interactions. Recent tunneling experiments in the integer quantum Hall regime discovered, remarkably, that the TDOS vanishes linearly on approaching the Fermi energy . Since at the transition, $`\sigma _{xx}`$ is finite and the localization length $`\xi `$ is very large, the linear suppression of the TDOS at criticality is expected to have a different origin than the 2D classical Coulomb gap behavior deep in the insulating phases. The results of numerical calculations of the TDOS using Hartree-Fock approximation (HFA) of Coulomb interaction also show a linear Coulomb gap at all Fermi energies in the lowest Landau level . Since HFA does not include screening of the exchange interaction, the correct behavior of the TDOS remained unclear, especially at the transition.
In this paper, we present a nonperturbative calculation of the TDOS at the IQHT, taking into account the dynamical screening of Coulomb interaction by the diffusive electrons. We show that the TDOS is given by,
$$\nu (\epsilon )=C_Q|\epsilon |/e^4,$$
(1)
at low energy $`\epsilon `$ (bias). We shall refer to Eq. (1) as the 2D quantum Coulomb gap behavior. The coefficient $`C_Q`$ is not a universal number as in the 2D classical Coulomb gap expression , but rather a quantity of quantum mechanical origin. In general, it depends on magnetic field and the microscopic details of the sample such as the mobility. We find, at the IQHT,
$$C_Q=\sqrt{1/\pi g}\left[1+\mathrm{\Phi }(\sqrt{g})\right]e^{g+\frac{1}{4g}\mathrm{ln}^2\mathrm{\Delta }},$$
(2)
where $`g=2\pi ^2\sigma _c`$ with $`\sigma _c`$ the critical conductivity in units of $`e^2/\mathrm{}`$ and $`\mathrm{\Delta }=(k_fa_B)^2/(2\pi ^2gk_fl)`$ with $`a_B`$ the Bohr radius and $`l`$ the zero-field mean free path. $`\mathrm{\Phi }`$ is the error function. For large $`\epsilon `$, $`\nu (\epsilon )`$ crosses over to the perturbative diagrammatic result in strong magnetic fields . We also study the case of short-ranged interactions corresponding to experimental situations when the presence of nearby ground planes or the tunneling electrodes themselves screen the long-ranged Coulomb interaction between electrons are large distances. In this case, we find that the zero-bias TDOS is suppressed from the noninteracting value but remains finite. The corrections at finite $`\epsilon `$ follow a universal power law: $`[\nu (\epsilon )\nu (0)]/\nu (0)=(|\epsilon |/\epsilon _0)^\gamma `$ with $`\gamma =x/z`$, where $`x0.65`$ is the scaling dimension of short-ranged interactions and $`z=2`$ .
We begin with the partition function,
$$Z=𝒟[\overline{\psi },\psi ]𝒟[\varphi ]𝒟[V]P[V]e^S,$$
(3)
where the imaginary time action $`S=S_0+S_I`$,
$`S_0`$ $`=`$ $`{\displaystyle _0^\beta }𝑑td^2r\overline{\psi }\left[_t{\displaystyle \frac{1}{2m}}(_\alpha ieA_\alpha )^2+V\right]\psi `$ (4)
$`S_I`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _0^\beta }𝑑td^2rd^2r^{}\varphi (r)U^1(rr^{})\varphi (r^{})`$ (5)
$`+`$ $`i{\displaystyle _0^\beta }𝑑td^2r\varphi (r)\overline{\psi }(r)\psi (r).`$ (6)
Here, $`\overline{\psi }`$ and $`\psi `$ are the electron Grassmann fields, $`\varphi `$ is a scalar potential coupled to electron density, and $`A_\alpha `$, $`\alpha =x,y`$ is the external vector potential, $`ϵ_{\alpha \beta }_\alpha A_\beta =B\widehat{z}`$. Integrating out $`\varphi `$ in Eq. (3) gives rise to the electron-electron interaction $`U(rr^{})`$ in Coulomb gauge. Quench average over the short-ranged impurity potential $`V`$ is represented in Eq. (3) by integrating over a Gaussian distribution $`P[V]`$ using the standard replica method.
The TDOS can be obtained from the impurity averaged Green’s function $`G(\tau )=\psi (r,\tau )\overline{\psi }(r,0)`$,
$$\nu (\epsilon )=\frac{1}{\pi }\mathrm{Im}𝑑\tau e^{i\omega _n\tau }G(\tau )|_{i\omega _n\epsilon +i\eta }.$$
(7)
In the imaginary time path integral, $`G(\tau )`$ is given by
$$G(\tau )=Z^1𝒟[\overline{\psi },\psi ]𝒟[\varphi ]𝒟[V]P[V]\psi (\tau )\overline{\psi }(0)e^S.$$
(8)
Consider a U(1) gauge rotation, $`\psi \psi e^{i\theta }`$, $`\overline{\psi }e^{i\theta }`$ with $`\theta =\theta (r,\tau )`$. In the rotated frame, $`G(\tau )`$ has the same form as in Eq. (8), but with a transformed action
$$S_\tau =Si_0^\beta d^2r𝑑t(j_0^\tau a_0j_\mu a_\mu ),$$
(9)
where $`a_\mu =_\mu \theta `$, $`\mu =\tau ,x,y`$, is the longitudinal U(1) gauge field coupled to the fermion 3-current $`j_\mu `$, and
$$j_0^\tau =\delta (r)[\mathrm{\Theta }(t)\mathrm{\Theta }(t\tau )]$$
(10)
is the source density (current) disturbance due to the tunneling electrons. Because of the latter, the ground state develops a nonvanishing charge and current density, and the saddle point of the tunneling action $`S_\tau `$ is shifted from that of $`S`$ in Eq. (8). Minimizing the action, $`S/\theta =0`$, one finds that the induced charge and current density, $`\rho =j_0`$ and $`𝐉=𝐣`$, satisfy the continuity equation,
$$\frac{}{t}\rho +𝐉=\mathrm{\Gamma }(r,t),$$
(11)
where $`\mathrm{\Gamma }(r,t)=_\tau j_0^\tau =\delta (r)[\delta (t)\delta (t\tau )]`$ corresponds to the process of injecting an electron at $`r=0`$ and time $`t=0`$ and removing it at $`t=\tau `$. Since the critical conductivity is finite at the IQHT, the charge spreading is expected to be described by (anomalous) diffusion, we have $`𝐉=D(\gamma _H\widehat{z}\times )\rho `$ where $`D`$ is the field-dependent diffusion coefficient and $`\gamma _H=\sigma _{xy}/\sigma _{xx}`$ is the Hall angle. Inserting this into Eq. (11), we obtain the diffusion equation,
$$\frac{}{t}\rho D^2\rho =\mathrm{\Gamma }(r,t).$$
(12)
Notice that $`\gamma _H`$ does not enter because the transverse force does not affect the charge spreading in the bulk of the sample. Solving Eq. (12) leads to
$$\rho (q,\omega _n)=\frac{\mathrm{\Gamma }(\tau ,\omega _n)}{|\omega _n|+D(q,\omega _n)q^2},$$
(13)
where $`\mathrm{\Gamma }(\tau ,\omega _n)=1\mathrm{exp}(i\omega _n\tau )`$ and $`\omega _n`$ is the boson Matsubara frequency.
Next we go back to the tunneling action in Eq. (9) and expand the currents around the expectation values: $`j_0=\rho +\delta j_0`$ and $`𝐣=𝐉+\delta 𝐣`$. The fluctuating parts now satisfy the homogeneous continuity equation: $`_t\delta j_0+\delta 𝐣=0`$. For convenience, we choose the unitary gauge by setting $`\varphi (r,t)=a_0(r,t)`$. Using Eq. (11), the tunneling action for charge spreading becomes,
$$S_\tau =S_ui_0^\beta 𝑑td^2r\rho a_0,$$
(14)
where the unitary gauge action $`S_u`$ is given by
$`S_u`$ $`=`$ $`S_0+i{\displaystyle _0^\beta }𝑑td^2r\delta j_\alpha a_\alpha `$ (15)
$`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _0^\beta }𝑑td^2rd^2r^{}a_0(r)U^1a_0(r^{}).`$ (16)
The one-particle Green’s function in Eq. (8) becomes,
$$G(\tau )=Z^1𝒟[\overline{\psi },\psi ]𝒟[\theta ]𝒟[V]P[V]\psi (\tau )\overline{\psi }(0)e^{S_\tau }.$$
(17)
We now quench average over the impurity potential and integrating out the fermions for a fixed gauge configuration . Eq. (17) becomes,
$$G(\tau )=𝒟[\theta ]G_\theta (\tau )e^{i{\displaystyle 𝑑td^2r\rho a_0}S_{\mathrm{eff}}\left[a_\mu \right]},$$
(18)
where $`G_\theta (\tau )\psi (\tau )\overline{\psi }(0)_\theta `$ is given by,
$$G_\theta (\tau )=\frac{𝒟[\psi ,\overline{\psi }]𝒟[V]P[V]\psi (\tau )\overline{\psi }(0)e^{S_u[\psi ,\overline{\psi },a_\mu ]}}{e^{S_{\mathrm{eff}}[a_\mu ]}},$$
(19)
with the gauge field effective action,
$$e^{S_{\mathrm{eff}}[a_\mu ]}=𝒟[\psi ,\overline{\psi }]𝒟[V]P[V]e^{S_u[\psi ,\overline{\psi },a_\mu ]}.$$
(20)
Thus far, all evaluations have been formal and exact, and explicitly displayed the structure of the theory in the U(1) sector.
To calculate $`G(\tau )`$ in Eq. (18), one has to make systematic approximations to obtain $`S_{\mathrm{eff}}[a_\mu ]`$ and $`G_\theta (\tau )`$. Following essentially the formalism of Finkelshtein , we obtain $`S_{\mathrm{eff}}`$ to quadratic order in $`a`$,
$$S_{\mathrm{eff}}=T\underset{n}{}d^2qa_0(q,\omega _n)\mathrm{\Pi }(q,\omega _n)a_0(q,\omega _n).$$
(21)
Here $`\mathrm{\Pi }`$ is the polarization function,
$$\mathrm{\Pi }^1(q,\omega _n)=\frac{U(q)}{1+U(q)\frac{dn}{d\mu }\frac{Dq^2}{|\omega _n|+Dq^2}},$$
(22)
which coincides with the dynamical screened Coulomb interaction . In Eq. (22), $`dn/d\mu `$ is the compressibility. The bare diffusion constant in the self-consistent Born approximation (SCBA) is $`D=\frac{1}{2}r_c^2\tau _0^1`$, where $`r_c=(2N+1)^{1/2}l_B`$ with $`l_B`$ the magnetic length and $`N`$ the Landau-level index. The effects of anomalous diffusion will be discussed later. Notice that in strong field, $`D`$ is proportional to the field dependent scattering rate $`1/\tau _0(B)\sqrt{\omega _c/\tau _0(0)}`$.
Next we turn to $`G_\theta `$ in Eq. (19). In the unitary gauge, the important interference effects between the phase of the electron wave functions have been accounted for in the tunneling action in Eq. (18). The amplitude fluctuations are small for the slowly varying gauge potentials that dominate the dynamically screened Coulomb potential in Eq. (22) . This is a unique feature of the slow diffusive dynamics of the electrons. We therefore neglect the $`\theta `$-dependence and write $`G_\theta (\tau )G_0(\tau )`$. The functional integral over the gauge potential in Eq. (18) can thus be carried out explicitly using the effective action in Eq. (21), leading to
$$G(\tau )=G_0(\tau )e^{W(\tau )}.$$
(23)
For Coulomb interaction, $`U(q)=2\pi e^2/q`$ in Eq. (22). Making use of Eq. (13), $`W(\tau )`$ has the form,
$$W(\tau )=\frac{T}{2}\underset{n}{}d^2q\frac{|\mathrm{\Gamma }(\tau ,\omega _n)|^2}{(|\omega _n|+Dq^2)^2}\frac{2\pi e^2}{q+\frac{\kappa Dq^2}{|\omega _n|+Dq^2}}.$$
(24)
where $`\kappa =2\pi e^2dn/d\mu `$ is the inverse screening length at the transition. Notice that $`W(\tau )`$ has a similar structure as the leading correction to the TDOS in the diagrammatic perturbation theory in both zero and strong magnetic field . The main contribution to the $`q`$-integral in Eq. (24) comes from the region $`|\omega _n|/D\kappa <q<(|\omega _n|/D)^{1/2}`$. In this region, the diffusion coefficient $`D`$ is a constant. The anomalous diffusion only appears in the opposite limit $`Dq^2|\omega _n|`$. As a result, $`W(\tau )`$ becomes, in the $`T=0`$ limit,
$$W(\tau )=\frac{1}{4\pi ^2}\frac{d\omega }{\omega }\frac{1}{\sigma _{xx}}\mathrm{ln}(\omega \tau _s)(1\mathrm{cos}\omega \tau ),$$
(25)
where $`\tau _s=1/D\kappa ^2`$, and the conductivity $`\sigma _{xx}=Ddn/d\mu `$ in units of $`e^2/\mathrm{}`$ following Einstein’s relation. The upper limit of the integral in Eq. (25) is $`\mathrm{}/\tau _0`$ since one can show $`\tau _s\tau _0`$ near Landau level centers. The long-time behavior of $`W`$ is therefore given by $`W(\tau )=(1/8\pi ^2\sigma _{xx})\mathrm{ln}(\tau /\tau _0)\mathrm{ln}(\tau /(\tau _s^2/\tau _0))`$ . Substituting this result into Eq. (23), we obtain,
$$G(\tau )=G_0(\tau )\mathrm{exp}\left[\frac{1}{8\pi ^2\sigma _{xx}}\mathrm{ln}\left(\frac{\tau }{\tau _0}\right)\mathrm{ln}\left(\frac{\tau \tau _0}{\tau _s^2}\right)\right].$$
(26)
The asymptotic behavior of $`G_0(\tau )`$ for large $`\tau `$ is $`G_0(\tau )\nu _0/\tau `$, where $`\nu _0`$ is the corresponding DOS . In the high-field limit of the SCBA, it is well known that $`\nu _0=(1/2\pi l_B^2)(2\tau _0)/\mathrm{}`$ and $`\sigma _{xx}=\nu _0D(2N+1)/2\pi ^2`$ in the center of the N-th Landau level. After analytical continuation to real time , we obtain the final result for the TDOS defined in Eq. (7),
$$\nu (\epsilon )=\frac{2\nu _0}{\pi }_0^{\mathrm{}}𝑑t\frac{\mathrm{sin}|\epsilon |t}{t}e^{\frac{1}{8\pi ^2\sigma _{xx}}\mathrm{ln}(t/\tau _0)\mathrm{ln}(t\tau _0/\tau _s^2)}.$$
(27)
We now discuss the behavior of $`\nu (\epsilon )`$ in different regimes, after making a few remarks. (i) The quantity in the exponential in the above equation did not follow from an expansion in $`1/\sigma _{xx}`$, but rather resulted from the leading contribution dominated by the anomalously divergent $`\mathrm{ln}^2`$-term at long-times. Next order corrections to the latter are of the order $`\{1/\sigma _{xx},1/\sigma _{xx}^2\}\mathrm{ln}t/\tau _0`$. (ii) While Eq. (27) correctly captures the double-log contributions, we have assumed that the conductivity $`\sigma _{xx}`$ does not depend on frequency in Eq. (25). However, $`\sigma _{xx}`$ can be renormalized by localization effects (of leading order $`(1/\sigma _{xx})\mathrm{ln}\omega \tau _0`$ in the unitary ensemble) and interaction effects (of leading order $`\mathrm{ln}\omega \tau _0`$ in strong magnetic field ). Thus $`\sigma _{xx}`$ assumes the frequency-independent SCBA value only if $`\mathrm{ln}(\epsilon \tau _0)\sigma _{xx}`$. In this regime, and for $`\mathrm{ln}(\epsilon \tau _0)\sqrt{\sigma _{xx}}`$, the integral in Eq. (27) gives,
$$\nu (\epsilon )=\nu _0\mathrm{exp}\left[\frac{1}{8\pi ^2\sigma _{xx}}\mathrm{ln}(|\epsilon |\tau _0)\mathrm{ln}(|\epsilon |\tau _s^2/\tau _0)\right].$$
(28)
Expanding the exponential to leading order in $`1/\sigma _{xx}`$, Eq. (28) reproduces the high-field diagrammatic perturbative result of Girvin, et. al. . In the case of $`B=0`$, such nonperturbative resummation of the perturbative double-log divergences was pointed out by Finkelshtein , and recently reexamined using different approaches . The result of Eq. (28) can be regarded as an extension of the latter to high magnetic fields.
(iii) As stated in (ii), at low-bias, i.e. for $`\mathrm{ln}(\epsilon \tau _0)\sigma _{xx}`$, the localization and interaction effects lead to, in general, a frequency dependent conductivity $`\sigma _{xx}(\omega )`$ in Eq. (25). However, at the IQHT, the critical conductivity $`\sigma _c`$ is finite and of the order of $`e^2/h`$ . Thus, the validity of our analysis, i.e. the structure of the double-log divergence at long-times, can be extended into the regime of small $`\epsilon `$, provided that $`\sigma _{xx}`$ in Eq. (27) is replaced by the critical conductivity $`\sigma _c`$. It is important to note that because of the double-log term, the exponential factor in Eq. (27) converges very fast such that the TDOS becomes analytic in small $`|\epsilon |`$. Expanding to leading order in $`|\epsilon |`$,
$$\nu (\epsilon )=\nu _0|\epsilon |\frac{2}{\pi }_0^{\mathrm{}}𝑑te^{{\displaystyle \frac{1}{8\pi ^2\sigma _c}}\mathrm{ln}\left(t/\tau _0\right)\mathrm{ln}\left(t\tau _0/\tau _s^2\right)}.$$
(29)
Performing this integral, and use the fact that the compressibility is only weakly renormalized, i.e. $`dn/d\mu \nu _0`$, we obtain the results given in Eqs. (1) and (2) with $`\mathrm{\Delta }=\tau _s/\tau _0`$. We thus conclude that for the 1/r-Coulomb interaction, the bulk TDOS at IQHT exhibits a linear Coulomb gap behavior of quantum mechanical origin, i.e. the 2D quantum Coulomb gap. As a consequence, the level spacing at the transition becomes $`\mathrm{\Delta }E1/L`$, where $`L`$ is the size of the system, leading to a dynamical scaling exponent $`z=1`$. This behavior of the TDOS is qualitatively different from those obtained in the clean case and in a weak magnetic field .
We emphasize that the linear quantum Coulomb gap behavior results from the combined effects of (i) two-dimensionality, (ii) 1/r—Coulomb potential, and (iii) quantum diffusion, i.e. a finite conductivity at $`T=0`$. It pertains therefore to other metal-insulator transitions in 2D amorphous electron systems, provided that the critical conductivity is finite . An example is the recently discovered 2D $`B=0`$ metal-insulator transition , although the asymptotic low temperature behavior of the critical conductivity extracted from the experimental data in this case is still controversial.
Finally, we turn to short-ranged interactions. It has been shown interacting potentials $`1/r^p`$ that decay faster than $`1/r^2`$ are irrelevant and scale to zero with scaling dimensions $`x=p2`$ for $`2<p<2+x_{4s}`$, $`x_{4s}0.65`$, and $`x=x_{4s}`$ for $`p>2+x_{4s}`$ at the IQHT . For simplicity, we will consider $`U(rr^{})=u\delta (rr^{})`$, thus $`U(q)=u`$ in the screened interaction in Eq. (22). Inserting the latter into Eq. (24), the $`q`$ integral no longer leads to the log-divergent term in $`\omega `$ as in Eq. (25) due to the short range nature of the interaction. Instead, we find ,
$$W_{\mathrm{sr}}(\tau )=_{1/\tau }^{1/\tau _0}\frac{\alpha }{|\omega |}𝑑\omega ,$$
(30)
where $`\alpha `$ is a nonuniversal quantity dependent on the interacting strength. It is given by
$$\alpha =\frac{1}{8\pi ^2\sigma _c}\lambda \frac{2+\lambda }{(1+\lambda )^2}\left(C+\mathrm{ln}\sqrt{1+\lambda }\right),$$
(31)
with $`\lambda =u\frac{dn}{d\mu }`$, $`C=1/2+1/(23\eta /2)`$, and $`\eta `$ the anomalous diffusion exponent . Once again in this case, the critical conductivity $`\sigma _c`$ is finite at the transition such that Eqs. (30) and (31) represent the leading contribution to $`W_{\mathrm{sr}}(\tau )`$ in the asymptotic limit. If the interaction $`u`$ were a marginal perturbation, then one would obtain, as in the Luttinger liquid case, $`W_{\mathrm{sr}}=\alpha \mathrm{ln}(\tau /\tau _0)`$, $`G_{\mathrm{sr}}(\tau )\tau ^{(1+\alpha )}`$, and $`\nu (\epsilon )\nu _0|\epsilon |^\alpha `$. However, this is not true, because $`u`$ is an irrelevant perturbation and the effective interaction scales to zero according to $`u_{\mathrm{eff}}u\omega ^{x/z}`$ with $`z=2`$ at the noninteracting fixed point . As a result, $`\alpha `$ obeys a scaling form $`\alpha (u,\omega )=𝒜(u\omega ^{x/z})`$. The fact that $`𝒜(u0,\omega )=0`$ implies, together with Eq. (31), the leading scaling behavior $`\alpha A\lambda (\omega \tau _0)^{x/z}`$ with $`A=C/4\pi ^2\sigma _c`$. Substituting this into Eq. (30), one finds that $`W_{\mathrm{sr}}(\tau )=A\lambda \gamma ^1[(\tau _0/\tau )^\gamma 1]`$ where $`\gamma =x/z0.32`$. That $`W_{\mathrm{sr}}(\tau )`$ converges in the large-$`\tau `$ limit is a consequence of the interaction being irrelevant, i.e. $`\gamma >0`$. Thus, we find for large $`\tau `$,
$$G_{\mathrm{sr}}(\tau )=\nu (0)\frac{1}{\tau }\mathrm{exp}\left[\frac{A\lambda }{\gamma }\left(\frac{\tau }{\tau _0}\right)^\gamma \right],$$
(32)
where $`\nu (0)=\nu _0\mathrm{exp}(A\lambda /\gamma )<\nu _0`$. After analytic continuation, the TDOS at small $`\epsilon `$ is given by,
$$\nu _{\mathrm{sr}}(\epsilon )=\nu (0)\left[1+\left(\frac{|\epsilon |}{\epsilon _0}\right)^\gamma \right],$$
(33)
where $`\epsilon _0=\tau _0^1(A\lambda /\gamma )^{1/\gamma }`$. This result leads to several interesting predictions: (a) For short-ranged interactions, the TDOS is finite and nonuniversal at zero bias $`\nu _{\mathrm{sr}}(\epsilon =0)=\nu (0)0`$. (b) Since $`\nu (0)\nu _0`$ for large $`\lambda `$, interactions still lead to strong density of states suppression at low bias. (c) The plus sign in Eq. (33) indicates that the TDOS increases with bias $`\epsilon `$ according to a universal power law with an initial cusp singularity for our value of $`\gamma `$. These predictions can, in principle, be tested experimentally by deliberately screening out the long-ranged Coulomb interaction using metallic gates.
The authors thank R. Ashoori, M. P. A. Fisher, S. M. Girvin, I. A. Gruzberg, A. Kamenev, Y. B. Kim, D.-H. Lee, and N. Read for useful discussions. This work was supported in part by NSF Grant No. PHY94-07194 (at ITP), DOE Grant No. DE-FG02-99ER45747, and an award from Research Corporation.
|
no-problem/9812/quant-ph9812053.html
|
ar5iv
|
text
|
# Two-photon Franson-type interference experiments are not tests of local realism
## Abstract
We report a local hidden-variable model which reproduces quantum predictions for the two-photon interferometric experiment proposed by Franson \[Phys. Rev. Lett. 62, 2205 (1989)\]. The model works for the ideal case of full visibility and perfect detection efficiency. This result changes the interpretation of a series of experiments performed in the current decade.
The ingenious two-particle interferometer introduced by Franson is an exciting tool to reveal properties of entangled states. In the current decade this device was used in many two-photon interferometric experiments , that beautifully reveal complementarity between single and two-photon interference. The results of the experiments cannot be described using standard methods involving classical electromagnetic fields .
However, in the original paper, entitled *Bell Inequality for Position and Time*, and in many papers that followed, it was claimed that the experiment constitutes a “test of local realism involving time and energy”. Some authors were more sceptical, noting that even the ideal gedanken model of the experiment involves a postselection procedure, in which $`50\%`$ of the events are discarded when computing the correlation functions . If all events are taken into account standard Bell inequalities are not violated. This does not prove the existence of a local hidden-variable (LHV) model, but merely states that such a model is not ruled out.
The situation is made even less transparent by similar claims concerning certain two-photon polarization experiments where the problem of discarded events also appears. This has earlier been treated on equal footing with the problems of the Franson-type experiments, but a recent analysis in of the entire pattern of events in the experiments in reestablishes the unconditional violation of local realism. One could be tempted to adapt the procedure of to the Franson experiment, but as will be shown below, this is not possible.
In short, no one has been able to explicitly show an unconditional incompatibility of local realistic models of Franson experiments with the quantum mechanical predictions, but on the other hand, no one has thus far shown the existence of a LHV model for such predictions . Our aim is to resolve this uncertainty about the interpretation of the Franson interferometry by constructing a LHV model which fully reproduces the quantum predictions for the ideal case, i.e for $`100\%`$ efficient detectors and $`100\%`$ visibility of the two-particle interference.
First, let us describe the main idea behind the Franson type experiments. The two-particle interferometer is presented schematically in Fig. 1. In the actual experiments the source of the photon pairs was the process of spontaneous parametric down conversion (PDC) in a nonlinear crystal pumped by a monochromatic cw laser field. Both the photon traveling to the left as well as the one traveling to the right were fed into two identical unbalanced Mach-Zehnder interferometers . The difference of the optical paths in those interferometers, $`\mathrm{\Delta }L`$, satisfies the relation $`cT_{\text{coh}}\mathrm{\Delta }L`$, where $`c`$ is the speed of light and $`T_{\text{coh}}`$ is the coherence time of the down-converted photons (it is effectively defined by the filters and the geometry of the collection of the PDC radiation). Such optical path differences prohibit any single photon interference. However, two-particle interference is observable, provided there is no way to know the actual paths of the photons of a pair which caused two spatially separated detectors to click.
The down-converted photons have the property that their detection times (barring retardation effects) are correlated to within their coherence times . Thus, the two photons either cause clicks in the two detection stations which are either coincident (i.e., within coherence times), or one of the clicks is delayed with respect to the other one by a time difference of the order $`\mathrm{\Delta }L/c`$. In the second case, one has the full “which way” information about the process that caused the two clicks (one knows exactly which photon went via the longer path, and which via the shorter one). In the first case, however, either both went via the longer arms (denoted by $`L`$) or both via the shorter arms ($`S`$) of the local interferometers. Therefore, provided the time of emission is unknown , there is no way to distinguish the two processes, and thus they interfere. The relative phase of the quantum amplitudes for the two processes can be controlled by the phase shifters within the Mach-Zehnder interferometers, and is equal to their sum.
Formally this can be described in the following simple way. Right before the exit beamsplitters of the local interferometers (Fig. 1) the photon state is
$$|\psi _{12}=\frac{1}{2}\left(|S_1+e^{i\varphi _1}|L_1\right)\left(|S_2+e^{i\varphi _2}|L_2\right).$$
(1)
The state vector $`|S_n`$ represents the $`n`$-th photon in the shorter arm, and $`L`$ denotes the longer arm. The detectors are located behind the exit 50–50 beamsplitters, and thus they hide the direct information about the path of a photon (the indirect information can still be revealed by the detection times). The part of the state vector $`|\psi _{12}`$ responsible for coincident events (up to $`T_{\text{coh}}`$) is given by
$$\frac{1}{2}\left(|S_1|S_2+e^{i(\varphi _1+\varphi _2)}|L_1|L_2\right).$$
(2)
As long as the emission times of the pairs of photons are in principle undefined, the two terms can interfere . The norm of this component is $`1/\sqrt{2}`$, and thus only half of the events will belong to this class. If the exit 50–50 beamsplitters are symmetric, the probabilities for two-particle processes to give the result $`l=\pm 1`$ for particle 1 and the result $`m=\pm 1`$ for particle 2, under specified phase settings, and in coincidence, are
$`P(l;m(\text{coincidence})|\varphi _1,\varphi _2)`$ (3)
$`=\frac{1}{8}\left(1+lm\mathrm{cos}(\varphi _1+\varphi _2)\right).`$ (4)
The other terms of $`|\psi _{12}`$ cannot lead to coincident counts, and because the paths taken by the photons is exactly known there is no interference. The probabilities are
$$P(l,\text{L};m,\text{E}|\varphi _1,\varphi _2)=P(l,\text{E};m,\text{L}|\varphi _1,\varphi _2)=\frac{1}{16},$$
(5)
where E denotes an earlier count, and L denotes a later count. No single photon interference is observed, in other words
$$P(l|\varphi _1)=P(m|\varphi _2)=\frac{1}{2}.$$
(6)
An essential property of any deterministic LHV model for the experiment is that it should contain the emission time as one of the variables describing the experiment. The reason is that the beamsplitters of, say, the right interferometer may be removed at any moment of the measurement process. In this case, the photons would be detected solely by the detector $`+1`$, and the detection time would indicate the moment of emission (of course, only up to the coherence time, but this is not essential). I.e., there exist an operational situation in which the emission time can be measured, and therefore it must be included in the LHV model. Further, under the same operational situation, the detections behind the left interferometer are either coincident with the detections on the right side, or retarded by $`\mathrm{\Delta }L/c`$. The LHV model must give predictions for the local events at one side of the experiment, independent of what measuring device is used at the other side. I.e., it must predict whether a given count on the left side would be coincident (we shall call this an early detection) or delayed (a late detection) with respect to a count at the right side, when the right interferometer is dismantled. From now on we shall assume, just like in the quantum model, that the average time between emissions is much longer than all other characteristic times of the experiment. Otherwise, we assume it to be completely random.
Let us now present a LHV model for a single emission at a specific time for the Franson experiment. The hidden variables are chosen to be an angular coordinate $`\varphi [0,2\pi ]`$ and an additional coordinate $`r[0,1]`$. The ensemble of hidden variables is chosen as that of an even distribution in this rectangle in $`(\varphi ,r)`$-space, but each pair of particles is described by a definite point $`(\varphi ,r)`$ in the rectangle, defined at the source at the moment of emission. At the left detector station, the measurement result is decided by the hidden variables $`(\varphi ,r)`$ and the local setting $`\varphi _1`$ of the apparatus. Upon arrival at the detection station, the local variable $`\varphi `$ is shifted to $`\varphi ^{}=\varphi \varphi _1`$, i.e. by the current setting of the local phase shifter.
This shifted value of the angular hidden variable, $`\varphi ^{}`$, together with $`r`$, determine the result of the local dichotomic observable $`l=\pm 1`$ and whether the particle is detected *early* E or *late* L (Fig. 2). E.g., if the shifted hidden variables $`(\varphi ^{},r)`$ end up in an area denoted $`+1_\text{E}`$, the detector $`+1`$ fires *early*, if in an area denoted $`1_\text{L}`$ the detector $`1`$ fires *late*, and so on.
At the right detector station, a similar procedure is followed. The result now depends on $`(\varphi ,r)`$ and the local setting $`\varphi _2`$ of the apparatus. In this case, the shift is to the value $`\varphi ^{\prime \prime }=\varphi +\varphi _2`$, and the result is obtained in Fig. 3 in the same manner as before .
The single-particle detection probabilities follow the quantum predictions as given by (6). The coincidence probabilities are determined by interposing Fig. 2 and Fig. 3 with the proper shifts. The probability of having $`l=+1_\text{E}`$ and $`m=1_\text{E}`$ simultaneously is the area of the set indicated in Fig. 4 divided by $`2\pi `$ (the total area is $`2\pi `$ whereas the total probability is 1). The probability of the result $`(+1,\text{E};1,\text{E})`$ is easily obtained, and because of the symmetry of the model, the probability of $`(+1,\text{L};1,\text{L})`$ is the same. Since it is not possible to distinguish these two results from each other, the probability of interest is
$`P(+1;1(\text{coincidence})|\varphi _1,\varphi _2)`$ (7)
$`=P(+1,\text{E};1,\text{E}|\varphi _1,\varphi _2)`$ (8)
$`+P(+1,\text{L};1,\text{L}|\varphi _1,\varphi _2)`$ (9)
$`={\displaystyle \frac{2}{2\pi }}{\displaystyle _0^{\varphi _1+\varphi _2}}{\displaystyle \frac{\pi }{8}}\mathrm{sin}(\varphi )𝑑\varphi `$ (10)
$`={\displaystyle \frac{1}{8}}\left(1\mathrm{cos}(\varphi _1+\varphi _2)\right),`$ (11)
which is equal to the quantum prediction. The other probabilities of simultaneous detection are obtained in the same manner. The probabilities for non-simultaneous detection are also obtained by integration, but here the symmetry of the model is such that e.g.
$`P(+1,\text{E};1,\text{L}|\varphi _1,\varphi _2)`$ (12)
$`={\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{\pi /2}}\left({\displaystyle \frac{1}{2}}{\displaystyle \frac{\pi }{8}}\mathrm{sin}(\varphi )\right)𝑑\varphi ={\displaystyle \frac{1}{16}},`$ (13)
independently of the detector settings, also in accordance with the quantum predictions. Finally, if the emission time on one side is monitored (by removing the interferometer), the counts on the other side still must follow the same local model, and as it is evident from the figures 1 and 2 the counts split evenly between early and late ones.
Let us now move into the conclusions. As the coincident events constitute only $`50\%`$ of all events, one might want to dismiss the whole problem by stating that effectively only around $`\sqrt{50\%}71\%`$ events at a single detection station enter into the Bell analysis, which is much below the usual threshold of minimum $`83\%`$ . However, this is also the case in many other interferometric Bell-type configurations, e.g. , but performing a careful analysis it is still possible to show unconditional violations of local realism for the quantum predictions describing the expected phenomena in the ideal case . The above construction shows that it is not possible to extend this analysis to Franson-type experiments.
The main conclusion of our work is that all experiments with Franson-type two-particle interferometers have to be reinterpreted. They cannot ever serve as demonstrations of violation of local realism (or, if preferred, violation of a Bell inequality), simply because there exists a LHV model for the expected quantum predictions in the ideal case. We emphasize that this model does not rely on any imperfections of the actual experiments (like the notorious detector-efficiency loophole).
One should notice here that majority of the performed quantum cryptography experiments that involve entanglement are based on the Franson two-particle interferometry. The original idea of harnessing quantum entanglement to cryptographic jobs was based on the fact that security checks can always be performed by testing whether the signals violate the Bell inequalities . We have shown that such violations are only *apparent* for the studied Franson-type processes. Therefore, the basis for the application of this type of phenomena for quantum cryptography has to be carefully re-examined (compare ).
Facing our result, one could say that currently only the polarization entanglement setups are capable to produce true long-distance EPR-Bell type phenomena. Nevertheless, the Franson two-particle interferometry remains one of the most beautiful ways to demonstrate the non-classical nature of light (as the phenomena cannot be described by any classical field theory). Our *ad hoc* model is important only in relation to the Bell theorem.
Sven Aerts and Marek Żukowski were supported by the Flemish-Polish Scientific Collaboration Program No. 007. Marek Żukowski acknowledges support of UG Program BW/5400-5-0202-8. Sven Aerts is supported by the Flemish Institute for the advancement of Scientific-Technological Research in the Industry (IWT). Jan-Åke Larsson acknowledges support from the Swedish Natural Science Research Council.
|
no-problem/9812/astro-ph9812316.html
|
ar5iv
|
text
|
# Ultra-high-energy cosmic ray acceleration by relativistic blast waves
## 1 Introduction
The observation of some hundred cosmic ray events to date with energy above the Greisen–Zatsepin–Kuz’min cutoff, $`E_{\mathrm{GZK}}10^{19.5}`$ eV, (e.g. Takeda et al. 1998, and references therein) has sparked renewed interest in their origin. Ultra-high-energy cosmic rays (UHECRs), with energies in the range $`E10^{18.5}`$$`10^{20.5}`$ eV, are generally believed to be extragalactic in origin, based on their harder spectrum, isotropic arrival directions on the sky and the fact that they are not confined by the Galactic magnetic field. One class of models involves continuous production of UHECRs at large shocks, such as those associated with powerful radio galaxies, or with ongoing large-scale structure formation in clusters of galaxies (e.g. Norman, Melrose & Achterberg 1995, and references therein). A major difficulty with these models is that the number of possible sources within the volume contributing to the flux above $`E_{\mathrm{GZK}}`$, which has a radius $`D_{\mathrm{max}}50`$$`100`$ Mpc, is too small to explain the observed number of independent events.
An alternative model (Waxman 1995a; Vietri 1995; Milgrom & Usov 1995) considers impulsive production in the sources of gamma-ray bursts (GRBs). Observations of X-ray and optical afterglows of GRBs (van Paradijs et al. 1997; Metzger et al. 1997) have confirmed the cosmological origin of the phenomenon; the energy associated with each event is then thought to be $`_{\mathrm{GRB}}10^{51}`$$`10^{53}`$ erg. Waxman (1995a) and Vietri (1995) noted that assuming comparable efficiencies for gamma-ray and UHECR production, the estimated GRB rate, $`Q_{\mathrm{GRB}}10^8`$ Mpc<sup>-3</sup> yr<sup>-1</sup>, implies a flux of UHECRs reaching Earth from within $`D_{\mathrm{max}}`$ remarkably similar to the one observed. Another, perhaps more compelling, argument is that the dispersion in UHECR arrival times due to small-angle deflections in the intergalactic magnetic field implies that at any one time, enough GRB sources contribute to the UHECR flux to account for the observed number of independent arrival directions (Miralda-Escudé & Waxman 1996; Achterberg et al., in preparation).
Relativistic fireball models for GRBs involve an ultra-relativistic blast wave with Lorentz factor $`\mathrm{\Gamma }_\mathrm{s}10^2`$$`10^3`$ bounding the fireball (Rees & Mészáros 1992), and internal mildly relativistic shocks ($`\mathrm{\Gamma }_\mathrm{s}2`$$`10`$) due to unsteady outflows (Rees & Mészáros 1994). The quenching of interstellar scintillation observed in GRB radio afterglows (Frail et al. 1997) confirms the relativistic expansion of the source. Vietri (1995) proposed that these relativistic shocks can be sites of efficient particle acceleration up to energies exceeding $`10^{20}`$ eV. In particular, he argued that an ultra-relativistic shock with Lorentz factor $`\mathrm{\Gamma }_\mathrm{s}`$ will lead to an energy gain per crossing cycle $`E_\mathrm{f}/E_\mathrm{i}\mathrm{\Gamma }_\mathrm{s}^2`$, where $`E_\mathrm{i}`$ and $`E_\mathrm{f}`$ are the initial and final particle energies. If such an energy gain could be obtained repeatedly at the ultra-relativistic blast wave, UHECR energies would be reached in only a few cycles.
In this letter, we first consider the Fermi acceleration process at ultra-relativistic shocks in some detail; apart from the recent simulations by Bednarz and Ostrowski (1998), earlier calculations of such relativistic shock acceleration have concentrated on mildly relativistic shocks ($`\mathrm{\Gamma }_\mathrm{s}10`$). We distinguish between the energy gain in the initial and subsequent shock crossing cycles, and show that for physically realistic particle deflection processes upstream, the latter yield only an energy gain $`E_\mathrm{f}/E_\mathrm{i}2`$. We estimate the time scale for the acceleration process, and examine the maximum energy attainable, as well as the global energetics, for UHECR production scenarios at relativistic blast waves, first in a typical interstellar medium, and then in the plausible environment of a pulsar wind bubble.
## 2 Ultra-Relativistic Shock Acceleration
We consider an ultra-relativistic shock of Lorentz factor $`\mathrm{\Gamma }_\mathrm{s}1`$ relative to the upstream medium. Assuming the fluid is weakly magnetised, the shock jump conditions then imply that the shock velocity relative to the downstream medium reduces to $`c/3`$, while the relative Lorentz factor of the downstream and upstream media satisfies $`\mathrm{\Gamma }_\mathrm{r}=\mathrm{\Gamma }_\mathrm{s}/\sqrt{2}`$ (e.g. Blandford & McKee 1976).
### 2.1 Energy gain: initial vs. repeated crossings
At non-relativistic shocks, the standard scenario of particle acceleration (Krymskii 1977; Bell 1978; Axford, Leer & Skadron 1978; Blandford & Ostriker 1978) assumes that both the upstream and downstream media contain magnetic fluctuations which tend to isotropise particles by elastic scattering in the respective fluid rest frames.
Assuming that the same general principle operates at a relativistic shock, the energy gain per shock crossing cycle follows from relativistic kinematics (e.g. Peacock 1981; Achterberg 1993). For ultra-relativistic particles (of Lorentz factor $`\gamma 1`$), it can be written in terms of initial and final energies $`E_\mathrm{i}`$ and $`E_\mathrm{f}`$ as:
$$\frac{E_\mathrm{f}}{E_\mathrm{i}}=\mathrm{\Gamma }_\mathrm{r}^2\left(1\beta _\mathrm{r}\mu _\mathrm{d}\right)\left(1+\beta _\mathrm{r}\mu _\mathrm{u}^{}\right)=\frac{E_\mathrm{f}^{}}{E_\mathrm{i}^{}},$$
(1)
where the first and second equalities apply to shock crossing cycles which begin and end in the upstream and downstream media, respectively. Here $`\beta _\mathrm{r}`$ is the velocity of the downstream medium (in units of the speed of light) relative to upstream, $`\mathrm{\Gamma }_\mathrm{r}`$ is the associated Lorentz factor, and the quantities $`\mu _\mathrm{d}`$ and $`\mu _\mathrm{u}^{}`$ measure the cosine of the angle between the particle velocity and the shock normal, when the particle crosses the shock into the downstream and upstream respectively. We use the convention that the shock normal points into the upstream medium, so that $`\mu _\mathrm{u}^{}>0`$. Primed and unprimed quantities are respectively measured in the downstream and upstream rest frames. The only assumption is that scattering is elastic in the local fluid frame.
Kinematics require that $`1\mu _\mathrm{u}^{}>\beta _\mathrm{s}^{}=\frac{1}{3}`$, so that the factor $`(1+\beta _\mathrm{r}\mu _\mathrm{u}^{})`$ in (1) is of order unity. If $`\mu _\mathrm{d}`$ is more or less isotropically distributed, as would be the case for a population of relativistic particles already present in the undisturbed upstream medium, the factor $`(1\beta _\mathrm{r}\mu _\mathrm{d})`$ is also of order unity, and energy gains $`E_\mathrm{f}/E_\mathrm{i}\mathrm{\Gamma }_\mathrm{s}^2`$ can be achieved, as envisioned by Vietri (1995). A similar conclusion holds for particles which are initially non-relativistic upstream.
For all but the initial shock crossing into the downstream medium, however, the distribution in $`\mu _\mathrm{d}`$ for a relativistic shock will be highly anisotropic (Peacock 1981). For an ultra-relativistic particle with Lorentz factor $`\gamma \mathrm{\Gamma }_\mathrm{s}`$, the Lorentz transformation of the incidence angle reduces to $`\mu ^{}(2\mathrm{\Gamma }_\mathrm{s}^2\theta ^2)/(2+\mathrm{\Gamma }_\mathrm{s}^2\theta ^2)`$ when $`\theta \mathrm{cos}^1\mu 1`$. The kinematic condition $`\mu _\mathrm{u}^{}>\frac{1}{3}`$ is thus equivalent to $`\theta _\mathrm{u}<1/\mathrm{\Gamma }_\mathrm{s}`$, which defines the shock ‘loss cone’ in the upstream frame. We show below that for physically realistic deflection processes upstream, the particle cannot be deflected very far beyond this loss cone before the shock overtakes it, so that the angle $`\theta _\mathrm{d}1/\mathrm{\Gamma }_\mathrm{s}`$ as well. This severely limits the energy gain attainable for all but the initial shock crossing: when $`\theta _\mathrm{d}1`$, the downstream energy gain in (1) reduces to
$$\frac{E_\mathrm{f}^{}}{E_\mathrm{i}^{}}=\frac{1+\beta _\mathrm{r}\mu _\mathrm{u}^{}}{1+\beta _\mathrm{r}\mu _\mathrm{d}^{}}\frac{2+\mathrm{\Gamma }_\mathrm{s}^2\theta _\mathrm{d}^2}{2+\mathrm{\Gamma }_\mathrm{s}^2\theta _\mathrm{u}^2},$$
(2)
which is of order unity if $`\theta _\mathrm{d}1/\mathrm{\Gamma }_\mathrm{s}`$.
### 2.2 Upstream dynamics
We consider two mechanisms for the upstream deflection of particles needed to allow repeated shock crossings: regular deflection by a large-scale magnetic field, and scattering by small-scale magnetic fluctuations.
#### 2.2.1 Regular deflection
We first examine the case where the magnetic field may be considered uniform over the region sampled by the particle in its excursion upstream. We take the shock velocity to define the $`z`$-direction, and without loss of generality assume that the magnetic field lies in the $`xz`$ plane, $`𝑩=(B_{},\mathrm{\hspace{0.17em}0},B_{})`$, with $`qB_{}0`$.
The equation of motion for the velocity $`𝜷=𝒗/c`$ of an ultra-relativistic particle of charge $`q`$ can be solved approximately by expanding $`\beta _z=\mu 1\frac{1}{2}\theta ^2`$ to lowest order in $`\beta _x`$ and $`\beta _y`$, as $`\theta ^2\beta _x^2+\beta _y^2=𝒪(1/\mathrm{\Gamma }_\mathrm{s}^2)`$ throughout the particle’s orbit upstream. Then as long as $`B_{}B_{}/\mathrm{\Gamma }_\mathrm{s}`$, which will be the case for all but a fraction $`1/\mathrm{\Gamma }_\mathrm{s}^2`$ of possible magnetic field orientations, we can ignore the terms in $`B_{}`$ in the equation of motion. If we denote by $`𝜷_\mathrm{i}`$ and $`𝜷_\mathrm{f}`$ the ingress and egress particle velocities, we find that $`\beta _{\mathrm{f}x}=\beta _{\mathrm{i}x}`$ and
$$\beta _{\mathrm{f}y}=\frac{1}{2}\beta _{\mathrm{i}y}+\sqrt{\frac{3}{\mathrm{\Gamma }_\mathrm{s}^2}3\beta _{\mathrm{i}x}^2\frac{3}{4}\beta _{\mathrm{i}y}^2}.$$
(3)
Since $`\beta _{\mathrm{i}x}^2+\beta _{\mathrm{i}y}^2<1/\mathrm{\Gamma }_\mathrm{s}^2`$, eq. (3) implies
$$1<\mathrm{\Gamma }_\mathrm{s}\theta _\mathrm{d}2\frac{1}{3}>\mu _\mathrm{d}^{}\frac{1}{3}.$$
(4)
The energy gain ratio can then be obtained by substituting $`\theta _\mathrm{d}^2\beta _{\mathrm{f}x}^2+\beta _{\mathrm{f}y}^2`$ in (2). It has its maximum at $`\beta _{\mathrm{i}x}=0`$ and $`\mathrm{\Gamma }_\mathrm{s}\beta _{\mathrm{i}y}0.27`$, where $`E_\mathrm{f}^{}/E_\mathrm{i}^{}2.62`$.
#### 2.2.2 Direction-angle scattering
We now consider the case where the magnetic field upstream is not uniform but irregular. In the simplest such model, the field consists of randomly oriented magnetic cells with field amplitude $`B`$ and radius (correlation length) $`\mathrm{}_\mathrm{c}`$. If the particle’s gyration radius, $`r_\mathrm{g}E/qB`$, satisfies $`r_\mathrm{g}/\mathrm{\Gamma }_\mathrm{s}\mathrm{}_\mathrm{c}`$, it will be deflected sufficiently to recross the shock within a single cell, and the regular deflection regime above will apply. In the opposite case, $`r_\mathrm{g}/\mathrm{\Gamma }_\mathrm{s}\mathrm{}_\mathrm{c}`$, the particle’s momentum direction will diffuse in time, with angular diffusion coefficient $`D_0=c\mathrm{}_\mathrm{c}/(3r_\mathrm{g}^2)`$ (Achterberg et al., in preparation).
In this scattering regime, a particle initially crossing the shock at an angle $`\theta _\mathrm{u}`$ will, after a time $`t`$, be at an average distance upstream relative to the shock given by
$$z(t)z_\mathrm{s}(t)ct\left(\frac{1}{2\mathrm{\Gamma }_\mathrm{s}^2}\frac{\theta _\mathrm{u}^2}{2}D_0t\right),$$
(5)
where we have used the fact that $`\theta =𝒪(1/\mathrm{\Gamma }_\mathrm{s})`$ and expanded to lowest order in $`1/\mathrm{\Gamma }_\mathrm{s}`$. Defining the typical upstream residence time $`t_\mathrm{u}`$ as the solution of $`z(t_\mathrm{u})=z_\mathrm{s}(t_\mathrm{u})`$, the typical downstream recrossing angle may be estimated as
$$\theta _\mathrm{d}^2\theta ^2(t_\mathrm{u})\frac{2}{\mathrm{\Gamma }_\mathrm{s}^2}\theta _\mathrm{u}^2.$$
(6)
The typical direction angle upon recrossing the shock, $`\theta _\mathrm{d}^2^{1/2}`$, is thus again of order $`1/\mathrm{\Gamma }_\mathrm{s}`$. Substituting (6) into the energy gain formula (2) shows that the energy gain typically reaches its largest values when $`\theta _\mathrm{u}=0`$, where $`E_\mathrm{f}^{}/E_\mathrm{i}^{}2`$. These results are approximate, rather than exact averages, in that they neglect the correlation between individual shock recrossing times and crossing angles.
### 2.3 Acceleration time scale
As the energy gain per shock crossing cycle, $`\mathrm{\Delta }EE_\mathrm{f}E_\mathrm{i}`$, is comparable to the initial energy $`E_\mathrm{i}`$, the acceleration time scale will be of order the cycle time, which is the sum of the typical upstream and downstream residence times $`t_\mathrm{u}`$ and $`t_\mathrm{d}`$. In the case of deflection by a uniform magnetic field, the upstream residence time is the time required for the particle to be deflected by an angle of order $`1/\mathrm{\Gamma }_\mathrm{s}`$ \[eq. (3)\], i.e. $`t_\mathrm{u}(\mathrm{\Gamma }_\mathrm{s}\omega _\mathrm{c})^1E/(q\mathrm{\Gamma }_\mathrm{s}B_{}c)`$.
The downstream residence time will depend on the downstream scattering process, but if we assume Bohm diffusion, it can be roughly estimated as the gyrotime, i.e. $`t_\mathrm{d}^{}\omega _\mathrm{c}^1E^{}/(qB^{}c)`$. If the downstream magnetic field is simply due to the compression resulting from the shock jump conditions, $`B^{}B_{}^{}=\sqrt{8}\mathrm{\Gamma }_\mathrm{s}B_{}`$, then taking into account the Lorentz transformation of the energy for a typical particle emerging upstream ($`E^{}E/\mathrm{\Gamma }_\mathrm{s}`$) and the transformation of the time interval between two events occurring at the shock ($`t_\mathrm{d}^{}t_\mathrm{d}/\mathrm{\Gamma }_\mathrm{s}`$), one finds that $`t_\mathrm{d}t_\mathrm{u}`$. Moreover, turbulence downstream could amplify the magnetic field well above the value resulting from simple compression; assuming it reaches equipartition with the thermal pressure downstream, one can show that in this case $`B^{}(c/v_\mathrm{A})\mathrm{\Gamma }_\mathrm{s}B`$, where $`v_\mathrm{A}`$ is the upstream Alfvén speed. This yields a correspondingly shorter downstream residence time:
$$t_\mathrm{d}\frac{E}{qc^2\mathrm{\Gamma }_\mathrm{s}\sqrt{4\pi \rho }},$$
(7)
where $`\rho `$ is the upstream mass density.
The full shock crossing cycle time is thus dominated by $`t_\mathrm{u}`$ in all cases of interest. In the case of scattering upstream, where $`r_\mathrm{g}\mathrm{}_\mathrm{c}`$, solving (5) yields a typical upstream residence time $`t_\mathrm{u}r_\mathrm{g}^2/(c\mathrm{\Gamma }_\mathrm{s}^2\mathrm{}_\mathrm{c})`$. This may be combined with the regular deflection case in the single expression
$$t_\mathrm{u}\frac{E}{q\mathrm{\Gamma }_\mathrm{s}Bc}\times \mathrm{max}(\mathrm{\hspace{0.25em}1},\frac{r_\mathrm{g}}{\mathrm{\Gamma }_\mathrm{s}\mathrm{}_\mathrm{c}}).$$
(8)
For a given upstream magnetic field amplitude $`B`$, the regular deflection regime thus constitutes an absolute lower limit on the acceleration time for repeated shock crossings.
## 3 Fireballs in the general ISM
We now review the properties of the relativistic blast wave driven into the surrounding interstellar medium in fireball models of gamma-ray bursts. After an acceleration phase, an initially radiation-dominated fireball enters a relativistic ‘free expansion’ phase, driving a blast wave with the approximately constant Lorentz factor $`\mathrm{\Gamma }_\mathrm{s}\sqrt{2}\eta `$, where $`\eta _{\mathrm{GRB}}/(Mc^2)`$, $`M`$ being the baryonic mass in which the fireball energy $`_{\mathrm{GRB}}`$ is initially deposited (Piran, Shemi & Narayan 1993; Mészáros, Laguna & Rees 1993). The parameter $`\eta `$ is generally assumed to lie in the range $`10^2`$$`10^3`$.
If cooling of the shocked material can be ignored, this is followed by an adiabatic deceleration stage in which the blast wave Lorentz factor decreases with radius $`R_\mathrm{s}`$ as $`\mathrm{\Gamma }_\mathrm{s}R_\mathrm{s}^{3/2}`$ (Blandford & McKee 1976). The transition between the free expansion and deceleration phases occurs roughly at the radius $`R_\mathrm{d}`$ at which the energy in the swept-up surrounding material becomes comparable to $`_{\mathrm{GRB}}`$:
$$R_\mathrm{d}\left(\frac{3}{4\pi }\frac{_{\mathrm{GRB}}}{\eta ^2e}\right)^{1/3},$$
(9)
where $`e`$ is the total energy density of the surrounding material; for a typical interstellar medium of number density $`n`$, this is simply the rest-mass energy $`enm_\mathrm{p}c^2`$. The subsequent evolution of $`\mathrm{\Gamma }_\mathrm{s}`$ is then given by
$$\mathrm{\Gamma }_\mathrm{s}(R_\mathrm{s})\sqrt{2}\eta \left(\frac{R_\mathrm{s}}{R_\mathrm{d}}\right)^{3/2}\left(\frac{3}{2\pi }\frac{_{\mathrm{GRB}}}{e}\right)^{1/2}R_\mathrm{s}^{3/2},$$
(10)
and thus once $`R_\mathrm{s}>R_\mathrm{d}`$ the behaviour of $`\mathrm{\Gamma }_\mathrm{s}`$ is independent of the value of $`\eta `$.
We now consider various scenarios for the acceleration of UHECRs by the relativistic blast wave of a fireball expanding into the general interstellar medium.
### 3.1 Maximum energy for Fermi acceleration
In order for particles to be accelerated beyond the initial boost by shock acceleration of the Fermi type, repeated shock crossings must occur. The acceleration time, which we argued above is dominated by the upstream residence time $`t_\mathrm{u}`$, must therefore be shorter than the age of the blast wave measured in the upstream frame, $`R_\mathrm{s}/c`$. Using (8), the maximum energy for which this can be the case is
$$EqB\mathrm{\Gamma }_\mathrm{s}\times \text{min}(R_\mathrm{s},\sqrt{\mathrm{}_\mathrm{c}R_\mathrm{s}}).$$
(11)
The most favourable regime energetically is that of regular deflection upstream, which requires $`\mathrm{}_\mathrm{c}R_\mathrm{s}`$; in what follows we will assume that this is the case. Note that the maximum energy (11) is larger by a factor $`\mathrm{\Gamma }_\mathrm{s}`$ than that resulting from a simple geometrical comparison of the gyration radius with $`R_\mathrm{s}`$. This is due to the fact that a particle typically only executes a fraction $`\mathrm{\Gamma }_\mathrm{s}^1`$ of a Larmor orbit upstream before recrossing the shock.
The behaviour of $`\mathrm{\Gamma }_\mathrm{s}`$ as a function of $`R_\mathrm{s}`$ implies that the highest energy in (11) is reached at the transition between the free expansion and deceleration phases, $`R_\mathrm{s}R_\mathrm{d}`$. Using the definition (9) of $`R_\mathrm{d}`$, we obtain:
$$E5\times 10^{15}ZB_6_{52}^{1/3}\eta _3^{1/3}n_0^{1/3}\text{eV},$$
(12)
for ions of charge $`q=Ze`$, where $`B_6B/\left(10^6\mathrm{G}\right)`$, $`_{52}_{\mathrm{GRB}}/\left(10^{52}\text{erg}\right)`$, $`\eta _3\eta /10^3`$, and $`n_0n/\left(1\text{cm}^3\right)`$. Given typical interstellar magnetic fields of a few microgauss and the weak dependence on the other parameters, this rules out the production of UHECRs by Fermi acceleration at the unmodified external blast waves of relativistic fireballs.
### 3.2 Initial boost of galactic cosmic rays
#### 3.2.1 Maximum energy
One remaining possibility for blast wave acceleration of UHECRs is that they result from the initial shock crossing cycle energy boost of a pre-existing upstream population of relativistic particles. This is, in essence, the first of the two acceleration mechanisms considered by Vietri (1995).
This initial boost requires only the time $`t_\mathrm{d}`$ for the particle to be scattered downstream, which if equipartition fields are generated can be considerably shorter than $`t_\mathrm{u}`$, yielding a correspondingly higher maximum energy. Requiring the downstream half-cycle time (7) for a particle of final energy $`E`$ to be shorter than the age of the fireball, we obtain
$$E7\times 10^{20}Z_{52}^{1/3}\eta _3^{1/3}n_0^{1/6}\text{eV}.$$
(13)
This process can thus attain UHECR energies, provided particles with sufficient initial energy to be boosted in this range are present upstream. To reach $`10^{20}\text{eV}`$, this requires relativistic particles with energy above $`10^{14}\eta _3^2\text{eV}`$.
#### 3.2.2 Global energetics
The energy invested in boosting pre-existing relativistic particles is of order $`\mathrm{\Gamma }_\mathrm{s}^2e_{\mathrm{CR}}`$ per unit volume swept up by the blast wave, where $`e_{\mathrm{CR}}`$ is the upstream energy density of relativistic particles. Meanwhile, the blast wave expends an energy $`\mathrm{\Gamma }_\mathrm{s}^2e`$ per unit volume in shock-heating the surrounding medium, which has upstream energy density $`enm_\mathrm{p}c^2`$. Thus the fraction of the fireball’s energy that can go into boosting cosmic rays is of order $`fe_{\mathrm{CR}}/e`$. Taking the interstellar medium values in our Galaxy to be typical, $`e_{\mathrm{CR}}1\text{eV}\text{cm}^3`$ and $`e10^9n_0\text{eV}\text{cm}^3`$, we see that this mechanism can only have a very low efficiency, $`f10^9`$.
The efficient yield of UHECRs required in the GRB hypothesis, $`f0.1`$ (Waxman 1995a; Vietri 1995), could only be obtained if a large fraction of the surrounding medium’s energy density was in relativistic particles. This is in fact not implausible in the neutron star binary merger scenario for GRBs, as we now argue.
## 4 Fireballs in pulsar wind bubbles
The surrounding medium in which the relativistic fireball explodes is probably one modified by the activity of the progenitor system prior to the burst event. In the neutron star binary merger scenario for GRBs, these progenitors are identified with the binary pulsar systems observed in our own Galaxy (Narayan, Paczyński & Piran 1992). The pulsar in these systems can be expected to emit a relativistic wind, which over the lifetime of the binary can fill a large surrounding volume with shocked relativistic plasma. While this plasma will probably be predominantly in the form of electron-positron pairs created in the pulsar magnetosphere, it has been argued that pulsar winds must also contain ions in order to account for the accelerated pair spectrum observed in the Crab Nebula (Hoshino et al. 1992). The Crab Nebula’s wisps provide additional observational evidence for this ion component (Gallant & Arons 1994).
### 4.1 Maximum energy for boosted ions
We will assume the known binary millisecond pulsar systems PSR 1913+16, PSR 1534+12, and PSR 2127+11C to be typical of GRB progenitors. The spiral-in times of these binaries due to gravitational radiation are $`\tau 3\times 10^7\text{yr}`$ (Narayan et al. 1992), while the periods and period derivatives of the millisecond pulsars observed in these systems yield spindown luminosities $`10^{33}\text{erg}\text{s}^1`$ and characteristic times $`10^8\text{yr}`$ (Taylor, Manchester & Lyne 1993). Thus these pulsars will continuously inject relativistic particles into the surrounding medium over the lifetime of the system, with an approximately constant wind luminosity.
The relativistic flow of the pulsar wind will be thermalised in a termination shock, and this shocked material will form a relativistic plasma bubble in the interstellar medium. At the ages involved, we expect this pulsar wind bubble to be approximately isobaric and in pressure equilibrium with the surrounding medium. Once their directed kinetic energy is randomised by the shock, the pairs may suffer energy losses by synchrotron radiation over time, but these are negligible for the ions, which can only suffer adiabatic expansion losses. In an isobaric bubble these losses are negligible, so that the ions approximately conserve their post-shock energy throughout the bubble.
The bulk Lorentz factor $`\gamma _\mathrm{w}`$ of the pulsar wind can be estimated by assuming that the entire spindown luminosity $`\dot{E}`$ goes into a wind composed of electron-positron pairs and ions, with number fluxes parameterised in terms of the Goldreich–Julian flux $`\dot{N}_{\mathrm{GJ}}`$: $`\dot{E}=(h_\mathrm{i}m_\mathrm{i}+h_\pm m_\mathrm{e})\gamma _\mathrm{w}c^2\dot{N}_{\mathrm{GJ}}`$, where $`m_\mathrm{i}`$ and $`m_\mathrm{e}`$ are the ion and electron masses and $`h_\mathrm{i}`$ and $`h_\pm `$ are the ion and pair multiplicities. The wind energy per Goldreich–Julian charge is comparable to the polar cap potential, and can be estimated as $`\dot{E}/\dot{N}_{\mathrm{GJ}}e(\dot{E}/c)^{1/2}`$ (see, e.g., Michel 1991). We expect the ions, although a minor component of the pulsar wind particles by number ($`h_\mathrm{i}h_\pm `$), to form the dominant component by mass ($`h_\mathrm{i}m_\mathrm{i}h_\pm m_\mathrm{e}`$), so that the fraction of the wind energy carried by them, $`\xi _\mathrm{i}[1+h_\pm m_\mathrm{e}/(h_\mathrm{i}m_\mathrm{i})]^1`$, will be close to unity (Hoshino et al. 1992). We will also assume that the ions carry the magnetospheric return current, so that $`h_\mathrm{i}1/Z`$ (Gallant & Arons 1994). This then determines the typical energy, $`m_\mathrm{i}\gamma _\mathrm{w}c^2`$, of the thermalised post-shock ions.
One can show that for an initially isotropic distribution of relativistic particles upstream, the average energy gain (1) for boosted particles satisfies $`\frac{16}{9}\mathrm{\Gamma }_\mathrm{r}^2<E_\mathrm{f}/E_\mathrm{i}<\frac{8}{3}\mathrm{\Gamma }_\mathrm{r}^2`$, the exact value depending on the typical angle $`\mu _\mathrm{u}^{}`$ of those particles recrossing upstream. Taking a factor $`2\mathrm{\Gamma }_\mathrm{r}^2`$ to be representative, the typical energy of the ions boosted in the free expansion phase of the blast wave is
$$E_{\mathrm{max}}2\eta ^2m_\mathrm{i}\gamma _\mathrm{w}c^210^{20}\eta _3^2Z\xi _\mathrm{i}\dot{E}_{33}^{1/2}\text{eV},$$
(14)
where $`\dot{E}_{33}\dot{E}/\left(10^{33}\text{erg}\text{s}^1\right)`$. UHECR energies are thus naturally obtained is this scenario, for typical values of the pulsar wind parameters, as long as $`\eta 10^3`$.
### 4.2 Spectrum and efficiency
Although the upstream ions have a relatively narrow energy spectrum centered around $`m_\mathrm{i}\gamma _\mathrm{w}c^2`$, the boosted ions will have a broader distribution in energies provided the blast wave decelerates within the pulsar wind bubble, so that the boosting factor $`\mathrm{\Gamma }_\mathrm{s}^2`$ is not constant. The deceleration radius $`R_\mathrm{d}`$ is given by (9), with $`e`$ the energy density of the relativistic plasma, while the pulsar wind bubble radius $`R_\mathrm{b}`$ is obtained by equating the volume integral of $`e`$ with the total energy output of the pulsar. It follows that
$$\frac{R_\mathrm{d}}{R_\mathrm{b}}0.3_{52}^{1/3}\eta _3^{2/3}\dot{E}_{33}^{1/3}\tau _7^{1/3},$$
(15)
where $`\tau \tau _7\times 10^7\text{yr}`$ is the age of the system. Thus for typical parameters, the relativistic fireball will decelerate well within the pulsar wind bubble.
Since $`E\mathrm{\Gamma }_\mathrm{s}^2R_\mathrm{s}^3`$ in the deceleration phase, and the number of ions swept up by the blast wave scales as $`\mathrm{d}NR_\mathrm{s}^2\mathrm{d}R_\mathrm{s}`$, the spectrum of the boosted ions will be
$$\frac{\mathrm{d}N}{\mathrm{d}E}E^2.$$
(16)
This is consistent with the observed UHECR spectrum (Waxman 1995b). It can be shown that the same spectral index follows more generally if the ion density is not uniform, as assumed here, but obeys a power law in radius.
The lower bound of the spectrum (16) follows from the boosting factor $`\mathrm{\Gamma }_\mathrm{s}^2`$ of the blast wave when it reaches the edge of the pulsar wind bubble at $`R_\mathrm{b}`$, and is
$$E_{\mathrm{min}}3\times 10^{18}Z\xi _\mathrm{i}_{52}\dot{E}_{33}^{1/2}\tau _7^1\text{eV}.$$
(17)
Thus not only can a large fraction of the fireball energy go into boosted ions, but these ions have a power-law spectrum extending more or less exactly over the energy range where the extragalactic UHECR component is observed.
## 5 Summary
In this letter, we examined Fermi-type acceleration at an ultra-relativistic shock, and found that particles with initially isotropic momenta upstream can increase their energy by a factor of order $`\mathrm{\Gamma }_\mathrm{s}^2`$ in the initial shock crossing cycle. In all subsequent shock crossing cycles, however, we showed that the particle energy typically only doubles. This is due to the fact that particles do not have time to re-isotropise upstream before being overtaken by the shock, which we demonstrated in the specific cases of a large-scale ordered magnetic field and of small-scale magnetic fluctuations.
We argued that the maximum energy that can be reached by repeated shock crossings at the unmodified relativistic blast wave from a GRB fireball is well below the UHECR range. Pre-existing relativistic particles of sufficient energy can nonetheless be boosted to UHECR energies in the first shock crossing cycle. For a fireball expanding into a typical interstellar medium, where galactic cosmic rays provide the seed particles, we showed, however, that this process is too inefficient to account for UHECR production.
We proposed that the blast wave instead expands into a pulsar wind bubble produced by the progenitor system, a plausible hypothesis in the neutron star binary merger scenario for GRBs. We showed that for parameters typical of the neutron star binary systems observed in our Galaxy, relativistic ions in the pulsar wind bubble can be efficiently boosted by the blast wave to energies exceeding $`10^{20}\text{eV}`$. We argued that these boosted ions would have an $`E^2`$ spectrum, extending down in energy to about $`10^{18.5}\text{eV}`$.
## Acknowledgements
This work was supported by the Netherlands Foundation for Research in Astronomy (ASTRON) project 781–71–050.
|
no-problem/9812/astro-ph9812147.html
|
ar5iv
|
text
|
# Hot Stars in Globular ClustersBased on observations obtained at the ESO La Silla Observatory, the German-Spanish Calar Alto Observatory and with the Hubble Space Telescope
## 1 Historical Background
Today we know that galactic globular clusters are old stellar systems and people are therefore often surprised by the presence of hot stars in these clusters. As the following paragraphs will show hot stars have been known to exist in globular clusters for quite some time:
Barnard (1900) reports the detection of stars in globular clusters that were much brighter on photographic plates than they appeared visually: “Of course the simple explanation of this peculiarity is that these stars, so bright photographically and so faint visually, are shining with a much bluer light than the stars which make up the main body of the clusters”.
In 1915 Harlow Shapley started a project to obtain colours and magnitudes of individual stars in globular and open clusters (Shapley 1915a) hoping that “considerable advance can be made in our understanding of the internal arrangement and physical characteristics” of these clusters. In the first globular cluster studied (M 3, Shapley 1915b) he found a double peaked distribution of colours, with a red maximum and a blue secondary peak. He noticed that - in contrast to what was known for field dwarf stars - the stars in M 3 became bluer as they became fainter. Ten Bruggencate (1927, p.130) used Shapley’s data on M 3 and other clusters to plot magnitude versus colour (replacing luminosity and spectral type in the Hertzsprung-Russell diagram) and thus produced the first colour-magnitude diagrams<sup>1</sup><sup>1</sup>1Shapley (1930, p.26, footnote) disliked the idea of plotting individual data points - he thought that the small number of measurements might lead to spurious results. (“Farbenhelligkeitsdiagramme”). From these colour-magnitude diagrams (CMD’s) ten Bruggencate noted the presence of a giant branch that became bluer towards fainter magnitudes, in agreement with Shapley (1915b). In addition, however, he saw a horizontal branch (“horizontaler Ast”) that parted from the red giant branch and extended far to the blue at constant brightness.
Greenstein (1939) produced a colour-magnitude diagram for M 4 (again noting the presence of a sequence of blue stars at constant brightness) and mentioned that “the general appearance of the colour-magnitude diagram of M4 is almost completely different from that of any galactic (i.e. open) cluster”. He also noticed that - while main-sequence B and A type stars were completely missing - there existed a group of bright stars above the horizontal branch and on the blue side of the giant branch. Similar stars appeared also in the CMD’s presented by Arp (1955). As more CMD’s of globular clusters were obtained it became apparent that the horizontal branch morphology varies quite considerably between individual clusters. The clusters observed by Arp (1955) exhibited extensions of the blue horizontal branch towards bluer colours and fainter visual magnitudes, i.e. towards hotter stars<sup>2</sup><sup>2</sup>2The change in slope of the horizontal branch is caused by the decreasing sensitivity of B$``$V to temperature on one hand and by the increasing bolometric correction for hotter stars on the other hand. (see Fig. 1). In some of Arp’s CMD’s (e.g. M 15, M 2) these blue tails are separated from the horizontal part by gaps (see also Fig. 3).
About 25 years after their discovery first ideas about the nature of the horizontal branch stars began to emerge: Hoyle & Schwarzschild (1955) were the first to identify the horizontal branch with post-red giant branch (RGB) stars that burn helium in their cores.
Sandage & Wallerstein (1960) noted a correlation between the metal abundance and the horizontal branch morphology seen in globular cluster CMD’s: the horizontal branch became bluer with decreasing metallicity. Faulkner (1966) managed for the first time to compute zero age horizontal branch (HB) models that qualitatively reproduced this trend of HB morphology with metallicity (i.e. for a constant total mass stars become bluer with decreasing metallicity) without taking any mass loss into account but assuming a rather high helium abundance of Y = 0.35. Iben & Rood (1970), however, found that “In fact for the values of Y and Z most favored (Y $``$ 0.25 $``$ 0.28, Z = $`10^310^4`$), individual tracks are the stubbiest. We can account for the observed spread in color along the horizontal branch by accepting that there is also a spread in stellar mass along this branch, bluer stars being less massive (on the average) and less luminous than redder stars. It is somewhat sobering to realize that this conclusion comes near the end of an investigation that has for several years relied heavily on aesthetic arguments against mass loss and has been guided by the expectation of obtaining, as a final result, individual tracks whose color amplitudes equal the entire spread in color along the observed horizontal branches”. In the same paper they found that “During most of the double-shell-source phase, models evolve upwards and to the red along a secondary giant branch<sup>3</sup><sup>3</sup>3This secondary giant branch is called asymptotic giant branch (AGB) later in the text and consists of stars with a hydrogen and a helium burning shell. that, for the models shown, approaches the giant branch defined by models burning hydrogen in a shell.”
Comparing HB models to observed globular cluster CMD’s Rood (1973) found that an HB that “…is made up of stars with the same core mass and slightly varying total mass, produces theoretical c-m diagrams very similar to those observed. …A mass loss of perhaps 0.2 M with a random dispersion of several hundredths of a solar mass is required somewhere along the giant branch.” The assumption of mass loss also diminished the need for very high helium abundances.
Sweigart & Gross (1974, 1976) computed HB tracks including semi-convection and found that this inclusion considerably extends the temperature range covered by the tracks. However, Sweigart (1987) noted that “For more typical globular cluster compositions, however, the track lengths are clearly too short to explain the observed effective temperature distributions along many HB’s, and thus semiconvection does not alleviate the need for a spread in mass (or some other parameter), a point first emphasized by Rood (1973)”.
Caloi (1972) investigated zero age HB locations of stars with very low envelope masses ($``$ 0.02 $`\mathrm{M}_{}`$; extended or extreme HB = EHB) and found that they can be identified with subdwarf B stars in the field (Greenstein 1971). Sweigart et al. (1974) and Gingold (1976) studied the post-HB evolution and found that – in contrast to the more massive blue HB stars – EHB models do not ascend the second (asymptotic) giant branch (AGB).
Thus our current understanding sees blue horizontal branch stars as stars that burn helium in a core of about 0.5 $`\mathrm{M}_{}`$ and hydrogen in a shell. Their hydrogen envelopes vary between $``$ 0.02 $`\mathrm{M}_{}`$ (less massive envelopes belong to EHB stars which do not have any hydrogen shell burning) and 0.1 – 0.15 $`\mathrm{M}_{}`$. Depending on the mass of their hydrogen envelopes they evolve to the asymptotic giant branch (BHB stars) or directly to the white dwarf domain (EHB stars, AGB manqué stars). For a review see Sweigart (1994).
But blue horizontal branch stars are neither the brightest nor the bluest stars in globular clusters: Already Shapley (1930, p.30) remarked that “Occasionally, there are abnormally bright blue stars, as in Messier 13, but even these are faint absolutely, compared with some of the galactic B stars”. This statement refers to stars like those mentioned by Barnard (1900) which in colour-magnitude diagrams lie above the horizontal branch and blueward of the red giant branch (see Fig. 1). This is also the region where one would expect to find central stars of planetary nebulae, which are, however, rare in globular clusters: Until recently Ps1 (Pease 1928), the planetary nebula in M 15 with its central star K 648, remained the only such object known in globular clusters (see also Jacoby et al. 1997).
The bright blue stars are clearly visible in the colour-magnitude diagrams of Arp (1955). Apart from analyses of individual stars like vZ 1128 in M 3 (Strom & Strom 1970, and references therein) and Barnard 29 in M 13 (Traving 1962, Stoeckley & Greenstein 1968) the first systematic work was done by Strom et al. (1970). All stars analysed there show close to solar helium content, contrary to the blue horizontal branch stars, which in general are depleted in helium (Heber 1987). Strom et al. identified the brightest and bluest stars with models of post-AGB stars (confirming the ideas of Schwarzschild & Härm 1970) and the remaining ones with stars evolving from the horizontal branch towards the AGB. This means that all of the stars in this study are in the double-shell burning stage. Zinn et al. (1972) performed a systematic search for such stars using the fact that they are brighter in the U band than all other cluster stars. This also resulted in the name UV Bright Stars for stars brighter than the horizontal branch and bluer than the red giant branch. Zinn (1974) found from a spectroscopic analysis of UV bright stars in 8 globular clusters “a strong correlation between the presence of supra-HB stars in a globular cluster and the presence of HB stars hotter than log $`T_{eff}`$ = 4.1”. Harris et al. (1983) extended the compilation of UV bright stars in globular clusters and de Boer (1987) gave another list of UV bright stars in globular clusters, together with estimates of effective temperatures and luminosities.
De Boer (1985) found from analyses of IUE spectra of UV bright stars in 7 globular clusters that their contribution to the total cluster intensity ranges “from, on average, over 50% at 1200 Å to a few percent at 3000 Å.” Most of the UV bright stars found in ground based searches are cooler than 30,000 K, although theory predicts stars with temperatures up to 100,000 K (e.g. Schönberner 1983) The ground based searches, however, are biased towards these cooler stars due to the large bolometric corrections for hotter stars. It is therefore not very surprising that space based searches in the UV (Ultraviolet Imaging Telescope, Stecher et al. 1997) discovered a considerable number of additional hot UV bright stars in a number of globular clusters.
Space based observatories also contributed a lot of other information about hot stars in globular clusters: UIT observations showed the unexpected presence of blue HB stars in metal-rich globular clusters like NGC 362 (Dorman et al. 1997) and 47 Tuc (O’Connell et al. 1997). At about the same time Hubble Space Telescope (HST) observations of the core regions of globular clusters showed long blue tails in metal-rich bulge globular clusters (Rich et al. 1997). Observations of the very dense core regions of globular clusters show that the colour-magnitude diagrams seen there may differ considerably from those seen in the outer regions of the same clusters (e.g. Sosin et al. 1997). The most recent addition to the family of hot stars in globular clusters are the white dwarfs seen in HST observations of M 4 (Richer et al. 1995, 1997), NGC 6752 (Renzini et al. 1996) and NGC 6397 (Cool et al. 1996), which unfortunately are at the very limit for any spectroscopic observations even with 10m class telescopes.
The interest in old hot stars like blue horizontal branch and UV bright stars has been revived and extended by the discovery of the UV excess in elliptical galaxies (Code & Welch 1979; de Boer 1982) for which they are the most likely sources (Greggio & Renzini 1990, Brown et al. 1997).
## 2 Spectroscopic Analysis Methods
Much of the discussion and findings described above are based solely on the photometric properties of hot stars in globular clusters. Much more physical information regarding their evolutionary status can be gained from spectroscopic analyses: From spectra of various resolutions in combination with photometric data we can determine their atmospheric parameters (effective temperature, surface gravity, and helium abundance) and compare those to the predictions of the stellar evolutionary theory. The disadvantage of spectroscopic observations (compared to photometric ones) is the fact that they require larger telescopes and/or more observing time: For the observations of the blue HB stars in M 15 we used the 3.5m telescope of the German-Spanish Calar Alto observatory in Spain and the targets in NGC 6752 were mostly observed with the NTT at the ESO La Silla observatory in Chile.
To determine effective temperatures and surface gravities we compare various spectroscopic and photometric observations to their theoretical counterparts. Depending on the available observational and theoretical data and the amount of software sophistication a wide variety of analysis methods is currently available. The following paragraphs attempt to give an overview that allows to judge our results – for detailed information we refer the reader to the cited papers.
### 2.1 Effective Temperature
The ideal temperature indicator is insensitive to variations of surface gravity because it then allows to derive the effective temperature independently from the surface gravity. The UV flux distribution meets this requirement for all blue HB stars and the Balmer jump fulfills it for stars with effective temperatures between about 11,000 K and 30,000 K<sup>4</sup><sup>4</sup>4Johnson UBV photometry becomes rather gravity independent for temperatures above about 20,000 K - at the same time, however, it also loses temperature sensitivity. Strömgren uvby photometry stays temperature sensitive up to higher effective temperatures but is not available for most globular clusters.. As interstellar extinction changes (reddens) the flux distribution of a star the observational data must be corrected for this effect. We dereddened the observed spectra using the extinction law of Savage & Mathis (1979) and the appropriate reddening values for the respective globular cluster. For the analysis published ATLAS9 model spectra (Kurucz 1992) for the metallicity closest to the globular cluster metallicity were used. If not mentioned otherwise the quality of the fit was judged by eye. An example is shown in Fig. 2. Temperature determinations that include UV data (e.g. IUE spectrophotometry) are in general more reliable than those relying solely on optical observations, as the UV flux distribution is more sensitive to temperature variations than the optical continuum. However, as only a very limited amount of UV spectrophotometry is available such data were used only by Heber et al. (1986), de Boer et al. (1995), Cacciari et al. (1995), and Moehler et al. (1997b).
If only optical spectrophotometry is available (most stars in NGC 6752 and M 15) the overall continuum slope and the Balmer jump should be fitted simultaneously, including - if possible - optical photometric data as well. It turns out that straylight from red (i.e. cool) neighbours can cause problems for optical spectrophotometry<sup>5</sup><sup>5</sup>5This problem does not affect the IUE data as the flux of cool stars decreases rapidly towards shorter wavelengths.: A model that fits the Balmer jump (and the BV photometry) cannot fit the spectrophotometric continuum longward of 4000 Å but instead predicts too little flux there - the straylight from the cool star causes a red excess. An attempt to quantify these effects is described in Moehler et al. (1997b).
Crocker et al. (1988) fit the continuum ($`\lambda \lambda `$ 3450–3700 Å and 4000–5100 Å) in their spectrophotometric data and employ a $`\chi ^2`$ test to find the best fit. In addition they use the star’s position along the observed HB to obtain another estimate of its temperature and finally average both values for $`\mathrm{T}_{\mathrm{eff}}`$.
### 2.2 Surface Gravity
Provided the effective temperature has been determined as described above the surface gravity can be derived by fitting the shape of the Balmer line profiles at a fixed temperature. For this purpose the spectra are normalized and corrected for Doppler shifts introduced by the radial velocities of the stars. The model spectra have to be convolved with a profile representing the instrumental resolution, which was generally determined from the FWHM of the calibration lines (for more details see Moehler et al. 1995). We computed (at fixed temperature) for H<sub>β</sub> to H<sub>δ</sub> the squared difference between the observed spectrum and the theoretical line profile and used the sum of these differences as estimator for the quality of the fit (Moehler et al. 1995, 1997b, see also Fig. 2). Crocker et al. (1988) used the same lines and employed a $`\chi ^2`$ test to determine $`\mathrm{log}\mathrm{g}`$. In addition they corrected their results for the subsolar helium abundance normally present in blue HB stars.
### 2.3 Simultaneous Determination of $`\mathrm{T}_{\mathrm{eff}}`$ and $`\mathrm{log}\mathrm{g}`$
For the cooler stars (below about 20,000 K) one can use a combination of optical photometry and Balmer line profile fits to determine effective temperature and surface gravity simultaneously: Reddening free indices (Q for Johnson UBV photometry, Moehler et al. 1995; \[c1\], \[u-b\] for Strömgren uvby photometry, de Boer et al. 1995) in comparison with theoretical values allow to determine a relation between effective temperature and surface gravity. Fits to the lower Balmer lines (H<sub>β</sub> to H<sub>δ</sub>) yield another relation between $`\mathrm{T}_{\mathrm{eff}}`$ and $`\mathrm{log}\mathrm{g}`$ and from its intersection with the photometric relation effective temperature and surface gravity can be derived (for examples see de Boer et al. 1995 and Moehler & Heber 1998).
For stars below about 8,500 K (Moehler et al. 1995, M 15) the Balmer lines depend more on $`\mathrm{T}_{\mathrm{eff}}`$ than on $`\mathrm{log}\mathrm{g}`$. In these cases the Balmer lines are used to estimate the temperature and $`\mathrm{log}\mathrm{g}`$ is derived from the Q value.
Including also the higher Balmer lines (H<sub>ϵ</sub> to H<sub>10</sub>) allows to derive effective temperature and surface gravity by fitting all Balmer lines (H<sub>β</sub> to H<sub>10</sub>) simultaneously (Bergeron et al. 1992; Saffer et al. 1994). This method has been used for the UV bright stars (Moehler et al. 1998a) and the blue HB stars in metal-rich globular clusters. We used the procedures developed by Bergeron et al. (1992) and Saffer et al. (1994), which employ a $`\chi ^2`$ test to establish the best fit. Using only the lower Balmer lines (H<sub>β</sub> to H<sub>δ</sub>) yields rather shallow minima of $`\chi ^2`$, which allow a large range of values for $`\mathrm{T}_{\mathrm{eff}}`$ and $`\mathrm{log}\mathrm{g}`$.
### 2.4 Helium Abundances
Helium abundances were either derived from the simultaneous fitting of the Balmer and He i/He ii lines (Moehler et al. 1998a) or from measured equivalent widths that are compared to theoretical curves-of-growth for the appropriate values of effective temperature and surface gravity (Moehler et al. 1997b).
### 2.5 Model atmospheres
Most of the stars discussed here are in a temperature–gravity range where LTE (local thermal equilibrium) is a valid approximation for the calculation of model atmospheres (Napiwotzki 1997). For the older data published ATLAS model spectra were used: ATLAS6 (Kurucz 1979) by Crocker et al. (1988) resp. ATLAS9 (Kurucz 1992) by de Boer et al. (1995) and Moehler et al. (1995, 1997b). The stars in NGC 6752 (Moehler et al. 1997b) required an extension of the model atmosphere grid to higher surface gravities, for which we used an updated version of the code of Heber (1983). The new fit procedures (Bergeron et al. 1992; Saffer et al. 1994) which we employed for the recent data (Moehler et al. 1998a) required line profiles for the higher Balmer lines (shortward of H<sub>δ</sub>) that are not available from Kurucz. We therefore calculated model atmospheres using ATLAS9 (Kurucz 1991, priv. comm.) and used the LINFOR program (developed originally by Holweger, Steffen, and Steenbock at Kiel university) to compute a grid of theoretical spectra that contain the Balmer lines H<sub>α</sub> to H<sub>22</sub> and He i lines. For those stars which show He ii lines in their spectra (and are thus considerably hotter than the bulk of our programme stars) it is necessary to take non-LTE effects into account (Napiwotzki 1997; Moehler et al. 1998a).
## 3 Gaps and Blue Tails
As mentioned above the blue tails seen in many CMD’s of globular clusters are often separated from the more horizontal part of the HB by gaps at varying brightness (examples are shown in Fig. 3; for a list of globular clusters with blue tails see Fusi Pecci et al. ; Catelan et al. and Ferraro et al. give comprehensive lists of clusters that show gaps and/or bimodal horizontal branches). Such gaps can be found already in Arp’s (1955) CMD’s and have caused a lot of puzzlement, since they are not predicted by any canonical HB evolution. One of the first ideas was that the gaps are created by diverging evolutionary paths that turn a unimodal distribution on the ZAHB into a bimodal one as the stars evolve away from the ZAHB (Newell 1973; Lee et al. 1994). Evolutionary calculations, however, do not show any such behaviour as horizontal branch stars spend most of their lifetime close to the ZAHB (Dorman et al. 1991; Catelan et al. 1998). Rood & Crocker (1985) suggested that the gaps separate two groups of HB stars that differ in, e.g., CNO abundance or core rotation. A more extreme version of this idea was suggested by Iben (1990): blue tail stars are produced differently from the blue HB stars, e.g. by merging of two helium white dwarfs. So far, no precursor systems of such stars have been observed. Quite recently, Caloi (1999) proposed a change in the stellar atmospheres from convection to diffusion as an explanation for the gaps around (B$``$V)<sub>0</sub> = 0, whereas Catelan et al. (1998) suggested that at least some of the gaps may be due to statistical fluctuations. More detailed descriptions of possible explanations for the gaps can be found in Crocker et al. (1988), Catelan et al. (1998), and Ferraro et al. (1998).
The need for more information on the stars along the blue tails led to our project to obtain atmospheric parameters for blue HB and blue tail stars in several globular clusters: NGC 6397 (de Boer et al. 1995), NGC 6752 (Heber et al. 1986; Moehler et al. 1997b), and M 15 (Moehler et al. 1995, 1997a). To enlarge our sample we also used the data of NGC 288, M 5, and M 92 published by Crocker et al. (1988). The CMD’s of these clusters can be found in Fig. 3.
### Evolutionary status
In Fig. 4 the physical parameters of the HB stars analysed by Crocker et al. (1988; M 5, M 92, NGC 288), de Boer et al. (1995; NGC 6397), and Moehler et al. (1995, 1997a, M 15; 1997b, NGC6752) are compared to evolutionary tracks. The zero-age HB (ZAHB) marks the position where the HB stars have settled down and started to quietly burn helium in their cores. The terminal-age HB (TAHB) is defined by helium exhaustion in the core of the HB star. The distribution of stars belonging to an individual cluster is hard to judge in this plot but it is obvious that the observed positions in the ($`\mathrm{log}\mathrm{g}`$, $`\mathrm{T}_{\mathrm{eff}}`$)-diagram fall mostly above the ZAHB and in some cases even above the TAHB<sup>6</sup><sup>6</sup>6Preliminary results of Bragaglia et al. (1999) indicate deviations from this trend. An indication of a low-temperature gap can be seen at $`\mathrm{log}\mathrm{T}_{\mathrm{eff}}`$ $``$ 4.1. The gaps seen in the CMD of NGC 6752 and in the M 15 data of Durrell & Harris 1993 (from which the two hottest stars in M 15 were selected) are visible in the ($`\mathrm{log}\mathrm{g}`$, $`\mathrm{T}_{\mathrm{eff}}`$) plane at about 24,000 K, where they separate BHB from EHB stars. In all other clusters the stars above and below the gaps are blue horizontal branch stars cooler than 20,000 K.
Independent of the occurrence of any gaps stars with temperatures between 11,000 ($`\mathrm{log}\mathrm{T}_{\mathrm{eff}}`$ = 4.04) and 20,000 K ($`\mathrm{log}\mathrm{T}_{\mathrm{eff}}`$ = 4.30) show lower gravities than expected from canonical scenarios, whereas stars outside this temperature range are well described by canonical HB and EHB evolutionary tracks. The UIT observations of M 13 (Parise et al. 1998) and the HUT spectra of M 79 (Dixon et al. 1996) also suggest lower than expected gravities for blue HB stars. Whitney et al. (1998) use UIT observations of the hot stars in $`\omega `$ Cen to claim that the extreme HB stars – which agree with theoretical expectations in our results – have lower than expected luminosities, which would mean higher than expected gravities. These deviations in $`\mathrm{log}\mathrm{g}`$ could indicate that some assumptions used for the calculations of model atmospheres may not be appropriate for the analysis of the BHB stars (see also de Boer et al. 1995, Moehler et al. 1995):
Diffusion might lead to peculiar abundance patterns, because radiative levitation might push up some metals into the atmospheres whereas other elements might be depleted due to gravitational settling. Line blanketing effects of enhanced metals may change the atmospheric structure. We found, however, that even an increase of 2 dex in \[M/H\] results in an increase of only 0.1 dex in $`\mathrm{log}\mathrm{g}`$<sup>7</sup><sup>7</sup>7This is consistent with the findings of Leone & Manfrè (1997) that Balmer-line gravities can be underestimated by 0.25 dex if a solar metal abundance is assumed for metal-rich helium weak stars.. Another effect of diffusion might be a stratification of the atmosphere, i.e. an increase of helium abundance with depth, which has been predicted for white dwarf atmospheres (Jordan & Koester, 1986). In order to affect the Balmer jump significantly the transition from low to high He abundance must take place at an optical depth intermediate between the formation depths of the Paschen and the Balmer continua. Such a fine tuning is unlikely to occur.
Rapid rotation rotation of the stars – if neglected in the model atmospheres – would lower the determined gravities. This effect, however, becomes significant only if the rotation velocity exceeds about half of the break-up velocity. As measured rotation velocities for HB stars are small (Peterson et al. 1995) this possibility can be ruled out as well.
As we did not find any systematic effects in our analysis that are large enough to explain the observed offsets in surface gravity we assume for now that the physical parameters are correct and look for a scenario that can explain them<sup>8</sup><sup>8</sup>8Scenarios like the merging of two helium white dwarfs (Iben & Tutukov 1984) or the stripping of red giant cores (Iben & Tutukov 1993, Tuchman 1985) may produce stars that deviate from the ZAHB. Such stars, however, are either too hot (merger) or too short-lived (stripped core) to reproduce our results.:
#### Deep mixing
Abundance variations (C, N, O, Na, Al) in globular cluster red giant stars (Kraft 1994, Kraft et al. 1995, Pilachowski et al. 1996) suggest that nuclearly processed material from deeper regions is mixed to the surface of the stars. Depending on the element considered this mixing has to reach down into varying depths. The enhancement of aluminium, for instance, requires the mixing to extend down into the hydrogen burning (= helium producing) shell (e.g. Cavallo et al. 1998). This means that any mixing that dredges up aluminium will also dredge up helium (helium mixing or deep mixing). Table 1 lists the evidence for deep mixing for the clusters shown in Fig. 3.
Such “helium mixed” red giant stars evolve to higher luminosities and therefore lose more mass than their canonical counterparts. The resulting HB stars then have less massive hydrogen envelopes and are thus hotter than in the canonical case. In addition the higher helium abundance in the hydrogen envelopes of the HB stars increases the efficiency of the hydrogen shell burning and thereby leads to higher luminosities at a given effective temperature. This increase in luminosity leads to lower gravities for “deep mixed” HB stars than predicted by canonical evolution. For a more detailed discussion of the effects of deep mixing on post-RGB evolution see Sweigart (1997, 1999).
From Fig. 5 it can be seen that most stars hotter than 11,000 K are well fitted by tracks that assume deep mixing<sup>9</sup><sup>9</sup>9The good fit of the helium-mixed tracks to the stars in NGC 288 is problematic as there is no evidence for deep mixing in this cluster (cf. Table 1).. The cooler stars, however, are better explained by canonical evolution. As deep mixing leads to hotter and brighter blue HB stars it is possible that cool blue HB stars result from unmixed RGB stars. Unfortunately it is not possible to determine the envelope helium abundance of a blue HB star because almost all of these stars are helium-deficient due to diffusion (Heber 1987). As it remains unclear what causes deep mixing (although rotation probably plays a role, Sweigart & Mengel 1979) we do also not know whether all RGB stars within one cluster experience the same degree of mixing.
### Masses
Knowing effective temperatures and surface gravities of the stars allows to determine the theoretical brightness at the stellar surface, which together with the absolute brightness of the star yields its radius and thus its mass (see de Boer et al. 1995, Moehler et al. 1995, 1997b). The distances to the globular clusters (necessary to determine the absolute brightnesses of the stars) were taken from the compilation of Djorgovski (1993). The results are plotted in Fig. 6 and can be summarized as follows:
While the masses of the stars in M 5 and NGC 6752 scatter around the canonical values the blue HB stars in all other clusters show masses that are significantly lower than predicted by canonical HB evolution - even for temperatures cooler than 11,000 K where we saw no deviation in surface gravity from the canonical tracks. The fact that the stars in two of the clusters show “normal” mass values makes errors in the analyses an unlikely cause for the problem (for a more detailed discussion see Moehler et al. 1995, 1997b, and de Boer et al. 1995). Also the merger models of Iben (1990) cannot explain these masses since the resulting stars are much hotter. However, if some of the distance moduli we used were too small this could cause such an effect – larger distances would result in brighter absolute magnitudes, i.e. larger radii and thus larger masses.
#### Distances to Globular Clusters
Using Hipparcos data for local subdwarfs several authors (Reid 1997, 1998; Gratton et al. 1997; Pont et al. 1998) determined distances to globular clusters by main sequence fitting. The results are given in Table 2 and show that the new distance moduli are in general larger than the old ones, in some cases by up to $`0\stackrel{\mathrm{m}}{.}4`$$`0\stackrel{\mathrm{m}}{.}6`$<sup>10</sup><sup>10</sup>10An increase of $`0\stackrel{\mathrm{m}}{.}2`$ in (m-M)<sub>V</sub> increases the mass of a cluster star by 20%..
It is interesting to note that for M 5 and NGC 6752 (where the masses almost agree with the canonical expectations) the new distances are close to the old ones, whereas for the metal poor clusters M 15, M 92, and NGC 6397 the new distance moduli are $`0\stackrel{\mathrm{m}}{.}3`$$`0\stackrel{\mathrm{m}}{.}6`$ larger than the old ones, thereby greatly reducing the mass discrepancies (see also Heber et al., 1997). The resulting new masses are plotted in Fig. 7 and in Table 3 we list the average ratio between the mass calculated for an HB star (as described in the text) and the supposed ZAHB mass for its temperature (from Dorman et al. 1993). It can be seen that in all cases (except NGC 6752) the agreement between expected and calculated mass improves with the new distance moduli, although the masses in NGC 288 remain significantly too low. From our observations we therefore favour the longer distance scale for globular clusters as suggested by most analyses of the Hipparcos data.
## 4 Blue HB Stars in Metal-Rich Globular Clusters
As mentioned in Section 1 metal-rich globular clusters tend to have red horizontal branches. This is plausible as according to canonical stellar evolutionary theory metal-rich HB stars have to have much smaller envelope masses than metal-poor HB stars to achieve the same temperature. Therefore a fine tuning of mass loss is required to produce blue HB stars in metal-rich environments in the framework of classical stellar evolution. Deep mixing or the merging of two helium white dwarfs offer other, more exotic, possibilities to produce hot stars. Yi et al. (1997, 1998) discuss possible mechanisms to produce blue HB stars in elliptical galaxies and d’Cruz et al. (1996) describe mechanisms to create extreme HB stars in metal-rich open clusters like NGC 6791, where Liebert et al. (1994) found subdwarf B stars. Despite this recent theoretical work it came as a surprise when blue HB stars really showed up in metal-rich globular clusters:
UIT images of 47 Tuc (\[Fe/H\] = $``$0.71; O’Connell et al. 1997) and of NGC 362<sup>11</sup><sup>11</sup>11While not exactly metal-rich NGC 362 has been famous as part of the second-parameter pair NGC 288/NGC 362: Both clusters have similar metallicities, but NGC 288 shows a well populated blue HB, whereas NGC 362 shows almost only red HB stars. (\[Fe/H\] = $``$1.28; Dorman et al. 1997) show the presence of blue stars. Colour-magnitude diagrams of the central regions of NGC 6388 (\[Fe/H\] = $``$0.60) and NGC 6441 (\[Fe/H\] = $``$0.53) obtained with the Hubble Space Telescope (HST) show sloped blue HB’s and long blue tails in both clusters (Rich et al. 1997). The slope means that in these clusters bluer HB stars are visually brighter than redder ones, in contrast to canonical expectations. Their brighter luminosities require the blue HB stars to have lower gravities, which can be caused by rotation, deep mixing and/or higher primordial helium abundance (Sweigart & Catelan 1998).
To find out what really causes the unexpected presence of blue HB stars in these metal-rich clusters we decided to perform a spectroscopic investigation. Unfortunately we were not too lucky with weather and technical conditions and the number of observed stars is small. In addition, some of the blue stars in 47 Tuc and NGC 362 turned out to be field HB stars or SMC main sequence stars. Those stars, that are members of the clusters, are confirmed to be blue HB stars with effective temperatures between 7,500 K and 15,000 K. Due to the weather conditions we could not observe the fainter and therefore hotter stars in these clusters, which would be especially interesting for the question of deep mixing. More details can be found in the forthcoming papers Moehler, Landsman, Dorman (47 Tuc, NGC 362) and Moehler, Catelan, Sweigart, Ortolani (NGC 6388, NGC 6441). The results of our spectroscopic analyses are plotted in Fig. 8 and show that so far there is no evidence for deep mixing or primordial helium enrichment in these stars<sup>12</sup><sup>12</sup>12The helium white dwarf merging model of Iben (1990) is unable to produce stars with so low temperatures, because available hydrogen envelope masses are small ($`<10^4`$ $`\mathrm{M}_{}`$).. More spectra, especially of fainter stars, are necessary to verify this statement.
## 5 Hot UV Bright Stars in Globular Clusters
As mentioned in Section 1 optical searches for UV bright stars in globular clusters yielded mainly stars cooler than 30,000 K (the majority of which was even cooler than 15,000 K) due to the increasing bolometric corrections for hotter stars. The vast majority of stars selected this way will evolve either from the blue HB to the asymptotic giant branch or from there to the white dwarf domain. It is rather unlikely to find post-EHB stars this way as they spend only short time in such cool regions (if they reach them at all, see Fig. 9). In addition their overall fainter magnitudes work against their detection. Searches in the ultraviolet regime, on the other hand, will favour hotter stars and thereby increase the chance to detect post-EHB stars. We therefore decided to spectroscopically analyse the many hot UV bright stars that were found in globular clusters by the Ultraviolet Imaging Telescope. Details of the observations, reduction, and analyses can be found in Moehler et al. (1998a). The main goal was to find out how the physical parameters of these stars compare to evolutionary tracks.
The derived effective temperatures and gravities of the target stars are plotted in Fig. 9 and compared to various evolutionary tracks. Three of the stars (in NGC 6121 and NGC 6723) appear to fit the post-early AGB<sup>13</sup><sup>13</sup>13Post-early AGB stars left the asymptotic giant branch before the thermally pulsing stage track, while the remaining targets (in NGC 2808 and NGC 6752<sup>14</sup><sup>14</sup>14including three stars analysed by Moehler et al. (1997b)) are consistent with post-EHB evolutionary tracks. Like the extreme HB stars themselves (Moehler et al. 1997b) the post-EHB stars show subsolar helium abundances probably caused by diffusion. In contrast the three post-early AGB stars - which are supposed to be successors to helium-deficient blue HB stars \- have approximately solar helium abundances. This agrees with the expectation that already during the early AGB stages convection is strong enough to eliminate any previous abundance patterns caused by diffusion.
As expected, the two clusters with populous EHB’s (NGC 2808 and NGC 6752) have post-EHB stars but no post-AGB stars. The number ratio of post-EHB to EHB stars in NGC 6752, however, is much lower than expected from stellar evolutionary theory: 6% instead of 15 – 20%. This discrepancy has first been noted by Landsman et al. (1996) and has been confirmed by our studies, which verified all four post-EHB candidates, but found no additional ones (Moehler et al. 1997b, 1998a). The clusters NGC 6723 and M 4, on the other hand, do not have an EHB population, although they do have stars blueward of the RR Lyrae gap (which are potential progenitors of post-early AGB stars). The lack of genuine post-AGB stars may be understood from the different lifetimes: The lifetime of Schönberner’s (1983) post-early AGB track is about 10 times longer than that of his lowest mass post-AGB track. Thus, even if only a small fraction of stars follow post-early AGB tracks, those stars may be more numerous than genuine post-AGB stars. Due to their relatively long lifetime, post-early AGB stars are unlikely to be observed as central stars of planetary nebulae since any nebulosity is probably dispersed before the central star is hot enough to ionize it. These different life times in combination with the fact that a considerable number of globular clusters stars (all post-EHB stars and some post-BHB stars) do not reach the thermally pulsing AGB stage could be an explanation for the lack of planetary nebulae in globular clusters reported by Jacoby et al. (1997).
## 6 Abundance Patterns of UV Bright Stars in Globular Clusters
Up to now detailed abundance analyses have been performed mainly for post-AGB stars in the field of the Milky Way (McCausland et al. 1992 and references therein; Conlon 1994; Napiwotzki et al. 1994; Moehler & Heber 1998), for which the population membership is difficult to establish. The summarized result of these analyses is that the abundances of N, O, and Si are roughly 1/10 of the solar values, while Fe and C are closer to 1/100 solar. McCausland et al. (1992) and Conlon (1994) interpret the observed abundances as the results of dredge-up processes on the AGB, i.e. the mixing of nuclearly processed material from the stellar interior to the surface. Standard stellar evolutionary theories (Renzini & Voli 1981; Vassiliadis & Wood 1993) do not predict any dredge-up processes for the low-mass precursors of these objects. Nevertheless the planetary nebula Ps 1 in M 15 as well as the atmosphere of its central star K 648 are both strongly enriched in carbon when compared to the cluster carbon abundance<sup>15</sup><sup>15</sup>15A preliminary analysis of ZNG1 in M 5 also shows evidence for a third dredge-up, but no trace of a nebula (Heber & Napiwotzki 1999). (Adams et al. 1984; Heber et al. 1993), pinpointing the dredge-up of triple $`\alpha `$ processed material to the stellar surface and suggesting a possible connection between dredge-up and planetary nebula ejection (Sweigart 1998). This discrepancy may be solved by newer evolutionary calculations which are able to produce a third dredge-up also in low-mass AGB stars (Herwig et al. 1997).
Napiwotzki et al. (1994) on the other hand suggest that the photospheric abundances are caused by gas-dust separation towards the end of the AGB phase: If the mass loss at the end of AGB ceases rapidly gas can fall back onto the stellar surface while the dust particles are blown away by radiative pressure. This process has been proposed by Bond (1991) to explain the extreme iron deficiencies seen in some cooler post-AGB stars and is described in more detail by Mathis & Lamers (1992).
As iron is very sensitive to depletion by gas-dust separation the iron abundance is the crucial key to the distinction between dredge-up and gas-dust separation. To verify any elemental depletion, however, one has to know the original abundance of the star, which is generally not the case for field stars. Therefore UV bright stars in globular clusters with known metallicities provide ideal test cases for this problem and we started a project to derive iron abundances from high-resolution UV spectra obtained with HST.
### Abundance analysis of Barnard 29 in M 13 and ROA 5701 in $`\omega `$ Cen
For Barnard 29 a detailed abundance analysis from optical spectra has been done by Conlon et al. (1994) and we use their results for $`\mathrm{T}_{\mathrm{eff}}`$ and $`\mathrm{log}\mathrm{g}`$ for our analysis. For ROA 5701 we determined these parameters from IUE low-resolution spectra, optical photometry, and optical spectroscopy. For the iron abundances of both stars we used GHRS spectra of 0.07 Å resolution that cover the range 1860 $``$ 1906 Å. The abundances have been derived using the classical curve-of-growth technique. We computed model atmospheres for the appropriate values of effective temperature, surface gravity, and cluster metallicity and used the LINFOR spectrum synthesis package (developed originally by Holweger, Steffen, and Steenbock at Kiel university) for the further analysis. A more detailed description of our analysis can be found in Moehler et al. (1998b).
For ROA 5701 we find an iron abundance of $`\mathrm{log}ϵ_{Fe}=4.89\pm 0.12`$ (\[Fe/H\] = $``$2.61) and for Barnard 29 we get $`\mathrm{log}ϵ_{Fe}=5.38\pm 0.14`$ (\[Fe/H\] = $``$2.12). Both stars thus show iron abundances significantly below the mean cluster abundances of \[Fe/H\] $``$ $``$1.5 …$``$1.7. To look for any abundance trends in Barnard 29 and ROA 5701 in comparison with other stars in these clusters we used the abundances of C, N, O, Si in addition to iron. For ROA 5701 we determined these abundances from optical high-resolution spectra. Abundance analyses of Barnard 29 have been performed by Conlon et al. (1994, N, O, Si) and Dixon & Hurwitz<sup>16</sup><sup>16</sup>16They also give an iron abundance of $`\mathrm{log}ϵ_{Fe}=5.30_{0.26}^{+0.22}`$, somewhat higher than ours. (1998, C).
It can be seen in Fig. 10 that N, O, and Si in our two objects show a behaviour similar to that in red giant stars. C seems to be depleted in ROA 5701, but this abundance is based on an upper limit for one line only (4267 Å) which may be affected by non-LTE effects. One should note here, however, that Gonzalez & Wallerstein (1994) find strong enhancements of CNO and s-process elements for the brightest of the cool UV bright stars in $`\omega `$ Cen which they interpret as evidence for a third dredge-up. Taking the sum of C+N+O as indicator for the original iron abundance (cf. Fig. 11) shows that ROA 5701 and Barnard 29 were not born iron depleted and – contrary to the brightest cool UV bright stars in $`\omega `$ Cen – also do not show any evidence for a third dredge up. The results of our analysis thus favour the gas-dust separation scenario as explanation for the abundance patterns of low-mass post-AGB and post-early AGB stars.
## 7 Summary
About nine years ago we began a spectroscopic study of blue horizontal branch stars to find the reason for the gaps seen in the CMD’s of many globular clusters. While we haven’t yet achieved this goal the study led to others which altogether resulted in some interesting findings about the evolutionary status of hot stars in globular clusters:
### Blue Horizontal Branch Stars
We studied stars above and below the gaps seen along the blue horizontal branch in the CMD’s of many globular clusters (cf. Fig. 3) and found that most of the stars below the gaps are physically the same as the stars above the gaps, i.e. blue horizontal branch stars with a helium burning core and a hydrogen burning shell. So far extreme horizontal branch stars have been verified spectroscopically only in two clusters (NGC 6752 and M 15).
The blue HB stars with temperatures between 11,000 K and 20,000 K show lower gravities than expected from canonical stellar evolution, which can be explained by deep mixing. The lower than expected masses that are found for most stars cooler than 20,000 K can be understood if we assume that the distance moduli to the globular clusters are larger than previously thought. Analyses of BHB and EHB stars within the same cluster will provide a crucial test of these two hypotheses.
We verified that most of the blue stars seen in the colour-magnitude diagrams of several metal-rich globular clusters are indeed blue horizontal branch stars in these clusters. We did so far not find any significant evidence for deep mixing or a higher primordial helium abundance in these metal-rich globular clusters but have currently too few data to draw any firm conclusions.
### UV Bright Stars in Globular Clusters
Analyses of hot UV bright stars in globular clusters uncovered a lack of genuine post-AGB stars – we found only post-early AGB and post-EHB stars. This may be an explanation for the lack of planetary nebulae in globular clusters seen by Jacoby et al. (1997). Abundance analyses of post-AGB stars in two globular clusters suggest that gas and dust separate during the AGB phase.
##### Acknowledgements
I want to thank M. Catelan, K.S. de Boer, U. Heber, W.B. Landsman, M. Lemke, R. Napiwotzki, S. Ortolani, and A.V. Sweigart for their collaboration on these projects. Thanks go also to the staff of the La Silla, ESO, and Calar Alto, DSAZ, and HST observatories for their support during and after observations. I gratefully acknowledge support for this work by the DFG (grants Mo 602/1,5,6), DARA (grant 50 OR 96029-ZA), Alexander von Humboldt-Foundation (Feodor Lynen fellowship) and Dr. R. Williams as director of the Space Telescope Science Institute (DDRF grant).
### References
* Adams S., Seaton M.J., Howarth I.D., Auriére M., Walsh J.R., 1984, MNRAS 207, 471
* Alcaino G., Buonanno R., Caloi V., Castellani V., Corsi C.E., Iannicola G., Liller W., 1987, AJ 94, 917
* Arp H.C., 1955, AJ 60, 317
* Barnard E.E., 1900, ApJ 12, 176
* Bell R.A., Briley M.M., Norris J.E., 1992, AJ 104, 1127
* Bergeron, P., Saffer, R.A., Liebert, J., 1992, ApJ 394, 228
* Bond H.E., 1991, in Evolution of Stars: the Photospheric Abundance Connection, eds. G. Michaud, A. Tutukov, IAU Symp. 145 (Kluwer: Dordrecht) p. 341
* Bragaglia, A., C. Cacciari, C., Carretta, E., Fusi Pecci, F., 1999, in The 3<sup>rd</sup> Conf. on Faint Blue Stars, eds. A.G.D. Philip, J. Liebert & R.A. Saffer (Cambridge:CUP), p. 447
* Brown J.A., Wallerstein G., Cunha K., Smith V.V., 1991, A&A 249, L13
* Brown J.A., Wallerstein G., 1993, AJ 106, 133
* Brown T.M., Ferguson H.C., Davidsen A.F., Dorman B., 1997, ApJ 482, 685
* Buonanno R., Corsi C.E., Fusi Pecci F., 1981, MNRAS, 196, 435
* Buonanno R., Buscema G. Corsi C., Iannicola G., Fusi Pecci F., 1983a, A&AS 51, 83
* Buonanno R., Buscema G. Corsi C., Iannicola G., Smriglio F., 1983b, A&AS 53, 1
* Buonanno R., Corsi C.E., Fusi Pecci F., Alcaino G., Liller W., 1984, A&AS 57, 75
* Buonanno R., Caloi V., Castellani V., Corsi C.E., Fusi Pecci F., Gratton R., 1986, A&AS 66, 79
* Buonanno R., Corsi C.E., Buzzoni A., Cacciari C., Ferraro F.R., Fusi Pecci F., 1994, A&A 290, 69
* Cacciari C., Fusi Pecci F., Bragaglia A., Buzzoni A., 1995, A&A 301, 684
* Caloi V., 1972, A&A 20, 357
* Caloi V., 1999, A&A in press
* Carretta E., Gratton R.G. 1997, A&AS 121, 95
* Catelan M., Borissova J., Sweigart A.V., Spassova N., 1998, ApJ 494, 265
* Cavallo R.M., Sweigart A.V., Bell R.A., 1998, ApJ 492, 575
* Code A.D., Welch G.A., 1979, ApJ 228, 95
* Conlon E.S., 1994, in Hot Stars in the Galactic Halo, eds. S. Adelman, A. Upgren, C.J. Adelman, CUP, p. 309
* Conlon E.S., Dufton, P.L., Keenan, F.P., 1994, A&A 290, 897
* Cool A.M., Piotto G., King I.R., 1996, ApJ 468, 655
* Crocker D.A., Rood R.T., O’Connell R.W., 1988, ApJ 332, 236
* d’Cruz N.L., Dorman B., Rood R.T., O’Connell R.W., 1996, ApJ 466, 359
* de Boer K.S., 1982, A&AS 50, 247
* de Boer K.S., 1985, A&A 142, 321
* de Boer K.S., 1987, in The 2<sup>nd</sup> Conference on Faint Blue Stars, eds. A.G.D. Philip, D.S. Hayes, J. Liebert, Davis Press, p. 95
* de Boer K.S., Schmidt J.H.K., Heber U., 1995, A&A 303, 95
* Dickens R.J., Croke B.F.W., Cannon R.D., Bell R.A., 1991, Nature 351, 212
* Dixon W.V., Davidsen A.F., Dorman B., Ferguson H.C., 1996, AJ 111, 1936
* Dixon W.V., Hurwitz M., 1998, ApJ 500, L29
* Djorgovski S., 1993, in Structure and Dynamics of Globular Clusters, eds. S.G. Djorgovski & G. Meylan, ASP Conf. Ser. 50, p. 373
* Dorman, B., Lee Y.-W., VandenBerg D.A., 1991, ApJ 366, 115
* Dorman, B., Rood, R.T., O’Connell, W.O., 1993, ApJ 419, 596
* Dorman B., Shah R.Y., O’Connell R.W., Landsman W.B., Rood R.T., et al., 1997, ApJ 480, L31
* Durrell P.R., Harris W.E., 1993, AJ 105, 1420
* Faulkner J., 1966, ApJ 144, 978
* Ferraro F.R., Paltrinieri B., Fusi Pecci F., Dorman B., Rood R.T., 1998, ApJ 500, 311
* Fusi Pecci F., Ferraro F.R., Bellazzini M., Djorgovski S., Piotto G., Buonanno R., 1993, AJ 105, 1145
* Gingold R.A., 1976, ApJ 204, 116
* Gonzalez G., Wallerstein G., 1994, AJ 108, 1325
* Gratton R.G., Fusi Pecci F., Carretta E., Clementini G., Corsi C.E., Lattanzi M. 1997, ApJ 491, 749
* Greenstein J.L., 1939, ApJ 90, 387
* Greenstein J.L., 1971, in White Dwarfs, ed. W.J. Luyten, IAU Symp. 42, (Reidel), p. 46
* Greggio L., Renzini A., 1990, ApJ 364, 35
* Harris H.C., Nemec J.M., Hesser J.E., 1983, PASP 95, 256
* Heber U., 1983, A&A 118, 39
* Heber U., 1987, Mitt. Astron. Ges. 70, 79
* Heber U., Kudritzki R.P., Caloi V., Castellani V., Danziger J., Gilmozzi R., 1986, A&A 162, 171
* Heber U., Kudritzki R.P., 1986, A&A 169, 244
* Heber U., Dreizler S., Werner, K., 1993, Acta Astron. 43, 337
* Heber U., Moehler S., Reid I.N., 1997, in HIPPARCOS Venice ’97, ed. B. Battrick, ESA-SP 402, p. 461
* Heber U., Napiwotzki R., 1999, in The 3<sup>rd</sup> Conf. on Faint Blue Stars, eds. A.G.D. Philip, J. Liebert & R.A. Saffer (Cambridge:CUP), p. 439 (astro-ph/9809129)
* Herwig F., Blöcker T., Schönberber D., El Eid M., 1997, A&A 324, L81
* Hoyle F., Schwarzschild M., 1955, ApJS 2, 1
* Iben I. Jr., 1990, ApJ 353, 215
* Iben I.Jr., Rood R.T., 1970, ApJ 161, 587
* Iben I. Jr., Tutukov A.V., 1984, ApJS 54, 335
* Iben I. Jr., Tutukov A.V., 1993, ApJ 418, 343
* Jacoby G.H., Morse J. A., Fullton L.K., Kwitter K.B., Henry R.B.C, 1997, AJ 114, 2611
* Jordan S., Koester D., A&AS 65, 367
* Kraft R.P., 1994, PASP 106, 553
* Kraft R.P., Sneden C., Langer G.E., Shetrone M.D., Bolte M., 1995, AJ 109, 2586
* Kraft R.P., Sneden C., Smith G.H., Shetrone M.D., Langer G.E., Pilachowski C.A., 1997, AJ 113, 279
* Kurucz R.L., 1979, ApJS 40, 1
* Kurucz R.L., 1992, in The Stellar Populations of Galaxies, eds. B. Barbuy & A. Renzini, IAU Symp. 149 (Kluwer:Dordrecht), 225
* Landsman W.B., Sweigart A.V., Bohlin R.C., Neff S.G., O’Connell R.W., et al., 1996, ApJ 472, L93
* Lee Y.-W., Demarque P., Zinn R., 1994, ApJ 423, 248
* Leone F., Manfrè M., 1997, A&A 320, 257
* Liebert J., Saffer R.A., Green E.M., 1994, AJ 107, 1408
* Mathis J.S., Lamers, H.J.G.L.M., 1992 A&A 259, L39
* McCausland R.J.H., Conlon E.S., Dufton P.L., Keenan F.P., 1992, ApJ 394, 298
* Moehler S., Heber U., de Boer K.S., 1995, A&A 294, 65
* Moehler S., Heber U., Durrell P., 1997a, A&A 317, L83
* Moehler S., Heber U., Rupprecht G., 1997b, A&A 319, 109
* Moehler S., Landsman W., Napiwotzki R., 1998a, A&A 335, 510
* Moehler S., Heber U., 1998, A&A 335, 985
* Moehler S., Heber U., Lemke M., Napiwotzki R., 1998b, A&A 339, 537
* Napiwotzki R., 1997 A&A 322, 256
* Napiwotzki R., Heber U., Köppen, J., 1994, A&A 292, 239
* Newell E.B., 1973, ApJS 26, 37
* O’Connell R.W., Dorman B., Shah R.Y., Rood R.T., Landsman W.B., et al., 1997, AJ 114, 1982
* Paczynski B., 1971, Acta Astron. 21, 1
* Paltoglou G., Norris J.E., 1989, ApJ 336, 185
* Parise R.A., Bohlin R.C., Neff S.G., O’Connell R.W., Roberts M.S., et al., 1998, ApJ 501, L67
* Pease F.G., 1928, PASP 40, 342
* Peterson R.C., Rood R.T., Crocker D.A., 1995, ApJ 453, 214
* Pilachowski C.A., Sneden C., Kraft R.P., Langer G.E., 1996, AJ 112, 545
* Pont F., Mayor M., Turon C., VandenBerg D.A. 1998, A&A 329, 87
* Reid I.N. 1997, AJ 114, 161
* Reid I.N. 1998, AJ 115, 204
* Renzini A., Voli, M., 1981, A&A 94, 175
* Renzini A., Bragaglia A., Ferraro F.R., Gilmozzi R., Ortolani S., et al., 1996, ApJ 465, L23
* Rich R.M., Sosin C., Djorgovski S.G., Piotto G., King I.R., et al., 1997, ApJ 484, L25
* Richer H.B., Fahlmann G.G., Ibata R.A., Stetson P.B., Bell R.A., et al., 1995, ApJ 451, L17
* Richer H.B., Fahlmann G.G., Ibata R.A., Pryor C., Bell R.A., et al., 1997, ApJ 484, 741
* Rood R.T., 1973, ApJ 184, 815
* Rood R.T., Crocker D.A., 1985, in Horizontal-Branch and UV-Bright Stars”, ed. A.G.D. Philip (Schenectady: L.Davis Press), p. 99
* Saffer R.A., Bergeron P., Koester D., Liebert J., 1994, ApJ 432, 351
* Sandage A.R., Wallerstein G., 1960, ApJ 131, 598
* Savage B.D., Mathis F.S., 1979, ARAA 17,73
* Schönberner D., 1983, ApJ 272, 708
* Schwarzschild M., Härm R., 1970, ApJ 160, 341
* Shapley H., 1915a, Contr. Mt. Wilson 115
* Shapley H., 1915b, Contr. Mt. Wilson 116
* Shapley H., 1930, Star Clusters, The Maple Press Company, York, PA, USA,
* Shetrone M.D., 1996, AJ 112, 1517
* Shetrone M.D., 1997, in Fundamental Stellar Properties: The Interaction between Observation and Theory, IAU Symp. 189 (poster proceedings) (Kluwer: Dordrecht), p. 158
* Smith V.V., Cunha K., Lambert D.L., 1995, AJ 110, 2827
* Smith G.H., Shetrone M.D., Bell R.A., Churchill C.W., Briley M.M., 1996, AJ 112, 1511
* Sneden C., Kraft R.P., Prosser C.F., Langer G.E., 1992, AJ 104, 2121
* Sneden C., Kraft R.P., Shetrone M.D., Smith G.H., Langer G.E., Prosser C.F., 1997, AJ 114, 1964
* Sosin C., Piotto G., Djorgovski S.G., King I.R., Rich R.M., Dorman B., Liebert J., Renzini A., 1997, in Advances in Stellar Evolution, eds. R.T. Rood & A. Renzini, CUP, p. 92
* Stecher T., Cornett R.H., Greason M.R., Landsman W.B., Hill J.K., et al., 1997, PASP 109, 584
* Stoeckley R., Grennstein J.L., 1968, ApJ 154, 909
* Strom S.E., Strom K.M., 1970, ApJ 159, 195
* Strom S.E., Strom K.M., Rood R.T., Iben I.Jr., 1970, A&A 8, 243
* Sweigart A.V., 1987, ApJS 65, 95
* Sweigart A.V., 1994, in Hot Stars in the Galactic Halo, eds. S. Adelman, A. Upgren, C.J. Adelman, CUP, p. 17
* Sweigart A.V., 1997, ApJ 474, L23
* Sweigart A.V. 1998, to appear in New Views on the Magellanic Clouds, eds. Y.-H. Chu, J. Hesser & N. Suntzeff, IAU Symp. 190 (ASPC)
* Sweigart A.V. 1999, in The 3<sup>rd</sup> Conf. on Faint Blue Stars, ed. A.G.D. Philip, J. Liebert & R.A. Saffer (Cambridge:CUP), 3 (astro-ph/9708164)
* Sweigart A.V., Mengel J.G., Demarque P., 1974, A&A 30, 13
* Sweigart A.V., Gross P.G., 1974, ApJ, 190, 101
* Sweigart A.V., Gross P.G., 1976, ApJS 32, 367
* Sweigart A.V., Mengel J.G., 1979, ApJ 229, 624
* Sweigart A.V., Catelan M., 1998, ApJ 501, L63
* ten Bruggencate P., 1927, Sternhaufen, Julius Springer Vlg., Berlin
* Traving G., 1962, ApJ 135, 439
* Tuchman Y., 1985, ApJ 288, 248
* Vassiliadis E., Wood P.R., 1993, ApJ 413, 641
* Whitney J.H., Rood R.T., O’Connell R.W., D’Cruz N.L., Dorman B., et al., 1998, ApJ 495, 284
* Yi S., Demarque P., Kim Y.-C., 1997, ApJ 482, 677
* Yi S., Demarque P., Oemler A. Jr., 1998, ApJ 492, 480
* Zinn R., 1974, ApJ, 193, 593
* Zinn R.J., Newell E.B., Gibson J.B., 1972, A&A 18, 390
* Zinn R., West M.J. 1984, ApJS 55, 45
|
no-problem/9812/cond-mat9812016.html
|
ar5iv
|
text
|
# Testing quantum correlations in a confined atomic cloud by scattering fast atoms
## Abstract
We suggest measuring one-particle density matrix of a trapped ultracold atomic cloud by scattering fast atoms in a pure momentum state off the cloud. The lowest-order probability of the inelastic process, resulting in a pair of outcoming fast atoms for each incoming one, turns out to be given by a Fourier transform of the density matrix. Accordingly, important information about quantum correlations can be deduced directly from the differential scattering cross-section. A possible design of the atomic detector is also discussed.
PACS numbers: 03.75.Fi, 05.30.Jp, 32.80.Pj, 67.90.+z
Successful advances in achieving collective quantum states in confined clouds of alkaline atoms and in atomic hydrogen make possible studying quantum coherent properties of these systems, as well as the revealing fundamental kinetic processes leading to the formation of the coherence. In Ref., a dynamics of the condensate growth has been observed by means of detecting a density increase usually associated with the condensate formation in the trap. The coherence of the condensate has been demonstrated directly by letting two released condensates form the interference pattern .
Mechanism of formation of quantum correlations is a matter of great attention and controversy. An emergence of the condensate due to either lowering temperature or reaching an equilibrium after fast quenching is associated with a formation of the off-diagonal long-range order (ODLRO) . A primary object displaying such an order is the one-particle density matrix (OPDM) $`\rho (𝐱_1,𝐱_2)`$. Typical distances over which these correlations become important are comparable with the interatomic spacing $`r_a`$. Consequently, an “early” detection of such emerging correlations is very difficult to achieve by light with the wavelength $`\lambda r_a`$ usually employed for probing the cloud density. In Ref., it has been suggested that a resonant fluorescence of an external atom identical to those forming the condensate is sensitive to the relative phase of two condensates. However, this method is not suitable for testing distances shorter than $`\lambda `$. Information about short-range density correlations (at distances $`<r_a`$) can, in principle, be obtained from the absorption of detuned resonant light . The change in local $`m`$-body density correlations (the so-called $`m!`$-effect ) can be seen by measuring recombination rates, and this has been already done experimentally for the equilibrium case . However, measurements of the correlation length $`r_c`$ of the forming ODLRO seem very unlikely to be achieved by these methods.
Thus it is tempting to have a tool which could make possible seeing the OPDM directly without limitations on the accessible distances.
Currently, great efforts are being dedicated to a creation of a controllable source of coherent atoms – atomic LASER (see Ref. and references therein). In Ref., a mechanism of accelerating neutral atoms has been proposed. Hence, it is very likely that a source of fast and coherent atoms will be available soon. In this paper, we suggest a method of detecting the OPDM which relies on inelastic scattering of such atomic beam off the atomic cloud.
We note that methods of scattering of neutrons and He atoms off liquid He are well known. In Ref. , the Impulse Approximation has been suggested to employ for interpreting the differential cross-section of fast neutrons. In such an approximation, it is possible to relate the momentum transfer distribution to some integral of the population factor in He. However, liquid He is a strongly interacting system where the gas parameter $`\xi =na^3`$ ($`n`$ is the density and $`a`$ is the scattering length) is not small. Accordingly, multiparticle excitations dominate in the final-state channel, which makes a direct measurement of the OPDM impossible. Certain assumptions about the role of the final-state effects should be made . In contrast to this, the gas parameter in the trapped atomic condensates can be as small as $`\xi 10^5`$. Accordingly, the mean free path is $`a/\xi 10^2`$ cm, which greatly exceeds the mean particle separation, and can become greater then the cloud size. This implies that a contribution of the multiparticle excitations can be safely neglected as long as a wavelength of the incoming atom is much smaller than the interparticle separation in the atomic cloud. Under these conditions, as we will discuss below, it becomes possible to measure the OPDM directly.
If the mass of the incoming fast atom is comparable to that of the atoms forming a cloud (or, in particular, the fast atom is just identical to the atoms of the cloud), the lowest-order inelastic scattering event is a production of two fast outcoming atoms (the size of atomic cloud is supposed to be small enough to neglect multiple scattering). The quantum-mechanical probability of this process turns out to be proportional to a Fourier transform of the OPDM. Another process - an elastic scattering of a single fast atom off the cloud - also exists. In this case one has a single fast atom in the final state, in a complete analogy with elastic scattering of light. In principle, the elastic processes might mask the inelastic ones. However, if the momentum of incoming atom, $`k`$, is much larger than the typical momentum of the the cloud, $`1/r_c`$, then the elastic-scattering angles are small in the parameter $`(kr_c)^11`$. This feature allows one to distinguish between the two processes. One more feature, arising at $`kr_c1`$ from the conservation laws, is the fact that the angle between the two created fast atoms is always close to $`\pi /2`$. This facilitates identification of pairs resulted from one and the same scattering event.
Apart from extremely small uncertainty of the momentum of the incident atom, the method we suggest implies that the total momentum of the outcoming pair can be detected with a sufficient precision. Thus, a reliable atomic detector is required. Therefore, we will also discuss a possible design of such a detector.
Now let us derive an expression for the inelastic cross-section in the lowest order with respect to the two-body interaction. For the sake of simplicity, we will not consider the complexity of all possible scattering channels and will concentrate on a simplest case of spin-polarized bosons, when the incoming fast atom is identical to particles forming the cloud, and its spin polarization is the same as that of the cloud. We emphasize that the validity of the suggested method relies on a possibility to have the incident momentum $`k`$ obeying the relation
$$\xi ^{1/3}ka1.$$
(1)
Then, one can represent the interaction Hamiltonian in a traditional form ($`\mathrm{}=1`$)
$$H_{int}=\frac{u_0}{2}𝑑𝐱\mathrm{\Psi }^{}(𝐱)\mathrm{\Psi }^{}(𝐱)\mathrm{\Psi }(𝐱)\mathrm{\Psi }(𝐱),u_0=\frac{4\pi a}{m},$$
(2)
with $`m`$ standing for the atomic mass. The total field $`\mathrm{\Psi }`$ can be subdivided into the low- and the high-energy parts, $`\psi `$ and $`\psi ^{}`$, respectively:
$$\mathrm{\Psi }=\psi +\psi ^{},\psi ^{}=\underset{𝐤}{}a_𝐤\mathrm{e}^{i\mathrm{𝐤𝐱}},$$
(3)
where the incident and the scattering states are described in terms of plane waves normalized to unit volume; $`a_𝐤`$ destroys the high-energy particle with the momentum $`𝐤`$. A substitution of Eq. (3) into Eq. (2) and selection of the terms that describe a process involving one incident fast atom carrying momentum $`𝐤`$ and two ejected fast atoms carrying momenta $`𝐤_1`$ and $`𝐤_2`$, as well as its reverse, yield
$$H_{int}^{}=u_0\underset{𝐤_1,𝐤_2,𝐤}{}a_{𝐤_1}^{}a_{𝐤_2}^{}a_𝐤𝑑𝐱\mathrm{e}^{i\mathrm{𝐪𝐱}}\psi +\text{H.c.},$$
(4)
where $`𝐪=𝐤_1+𝐤_2𝐤`$ is the transferred momentum. The double-differential cross-section for scattering with given energy and momentum transfers, $`\omega `$ and $`𝐪`$, in the lowest order with respect to $`H_{int}^{}`$ is given by the Golden rule formula
$$W(𝐪,\omega )=\frac{2mu_0^2}{k}\frac{d𝐤_1}{(2\pi )^3}\delta (\omega \omega _{fi})𝑑𝐱_1𝑑𝐱_2_{\mathrm{}}^{\mathrm{}}𝑑t\mathrm{e}^{i𝐪(𝐱_1𝐱_2)i\omega t}\rho (𝐱_1,t;𝐱_2,0),$$
(5)
where $`\rho (𝐱_1,t_1;𝐱_2,t_2)=\psi ^{}(𝐱_1,t_1)\psi (𝐱_2,t_2)`$ \[Note that unlike the uniformity in time, $`\rho (𝐱_1,t_1;𝐱_2,t_2)=\rho (𝐱_1,t_1t_2;𝐱_2,0)`$, the space uniformity cannot be, in general, assumed as long as a trapping potential exists.\]; $`\omega _{fi}`$ is the difference of the kinetic energies of the fast atoms in the final and initial states:
$$\omega _{fi}=\frac{1}{m}\left[\frac{q^2}{2}+\mathrm{𝐪𝐤}+k_1^2(𝐪+𝐤)𝐤_\mathrm{𝟏}\right],$$
(6)
with $`𝐤_1`$ being the momentum of one of the outcoming atoms while the momentum of the second one, $`𝐤_2`$, is fixed by the relation $`𝐤_2=𝐪+𝐤𝐤_1`$.
The integration over $`𝐤_1`$ in Eq. (5) can be carried out explicitly. However, first we notice that the requirement of large $`k`$ means that the values of $`q`$ and $`\omega `$, which are effectively selected by the correlator $`\rho (𝐱_1,t;𝐱_2,0)`$ in the right-hand side of Eq. (5), satisfy the conditions $`qk`$ and $`|\omega |k^2/m`$. This immediately leads to the approximation $`\delta (\omega \omega _{fi})m\delta (k_1^2\mathrm{𝐤𝐤}_1)`$, which yields
$$W(𝐪,\omega )=4a^2𝑑𝐱_1𝑑𝐱_2_{\mathrm{}}^{\mathrm{}}𝑑t\mathrm{e}^{i𝐪(𝐱_1𝐱_2)i\omega t}\rho (𝐱_1,t;𝐱_2,0).$$
(7)
We thus see that the double-differential cross-section $`W(𝐪,\omega )`$ is almost directly related to the dynamic correlator $`\rho (𝐱_1,t;𝐱_2,0)`$ which, as is known, contains rather rich information about the system, including, for one thing, the elementary excitation spectrum. Confining ourselves to the static correlations, described by OPDM $`\rho (𝐱_1,𝐱_2)=\rho (𝐱_1,0;𝐱_2,0)`$, we arrive at even more simple relation
$$W(𝐪)=8\pi a^2𝑑𝐱_1𝑑𝐱_2\mathrm{e}^{i𝐪(𝐱_1𝐱_2)}\rho (𝐱_1,𝐱_2)$$
(8)
in terms of the differential cross-section $`W(𝐪)=𝑑\omega W(𝐪,\omega )`$.
Consider the structure of $`W(𝐪)`$ in the most characteristic cases. In the case of pure Bose-Einstein condensate, when $`\rho (𝐱_1,𝐱_2)=\mathrm{\Phi }^{}(𝐱_1)\mathrm{\Phi }(𝐱_2)`$ \[$`\mathrm{\Phi }(𝐱)`$ is the condensate wave-function\],
$$W(𝐪)=\mathrm{\hspace{0.17em}8}\pi a^2|\mathrm{\Phi }_𝐪|^2,\mathrm{\Phi }_𝐪=𝑑𝐱\mathrm{e}^{i\mathrm{𝐪𝐱}}\mathrm{\Phi }(𝐱).$$
(9)
In particular, if $`\mathrm{\Phi }_𝐪`$ is real, Eqs. (9) can be reversed, and correspondingly $`\mathrm{\Phi }(𝐱)`$ can be restored from $`W(𝐪)`$.
Another instructive case is the axially symmetric quantum vortex. A presence of a single vortex in a center of the axially symmetric condensate drastically changes the scattering pattern. Indeed, in this situation, $`\mathrm{\Phi }=\mathrm{exp}(i\theta )\sqrt{n(r,z)}`$, where $`\theta `$ is the axial angle and $`n(r,z)`$ stands for the axially symmetric condensate density as a function of the distances $`r,z`$ perpendicular to the axis and along the axis, respectively. Accordingly, Eq. (9) shows that $`W(𝐪)=0`$ for $`𝐪`$ directed along the vortex line. The differential cross-section becomes finite as long as there is a component $`𝐪_{}`$ of $`𝐪`$ perpendicular to the axis or the vortex displaces from the condensate center. In the first case, $`W(𝐪)q_{}^2`$ for $`q_{}0`$.
Quite similar to the quantum vortex is the case of the supercurrent state of a toroidal Bose condensate. The suppression of $`W(𝐪)`$ for $`𝐪`$ perpendicular to the plane of the torus allows distinguishing the supercurrent state from the currentless genuine ground state.
In the absence of condensate, or for the above-the-condensate part of OPDM, one normally deals with the so-called quasi-homogeneous regime, when typical inverse momentum of particles is much less then the scale of density variation. In this case, it is reasonable to introduce the variables $`𝐫=𝐱_2𝐱_1`$ and $`𝐑=(𝐱_1+𝐱_2)/2`$, and to represent $`W(𝐪)`$ as
$$W(𝐪)=8\pi a^2𝑑𝐑\rho _𝐪(𝐑),\rho _𝐪(𝐑)=𝑑𝐱\mathrm{e}^{i\mathrm{𝐪𝐱}}\rho (𝐱,𝐑),$$
(10)
since in the quasi-homogeneous regime, OPDM in the Wigner representation, $`\rho _𝐪(𝐑)`$, has a semiclassical meaning of local (at the point $`𝐑`$) distribution of the particle momentum $`𝐪`$. We thus see that in this case $`W(𝐪)`$ yields spatially averaged momentum distribution. Moreover, without contradiction with the uncertainty principle, this averaging can be partially (totally) removed by collimating the incident (both the incident and outcoming) beams. Even without removing the averaging, $`W(𝐪)`$ contains valuable information about long-range correlations in the system, since the the averaging does not affect the order-of-magnitude value of the correlation radius $`r_c`$ (equal to the inverse typical momentum $`k_c`$). If the system contains both condensate and quasi-homogeneous above-the-condensate fraction, the cross-section $`W(𝐪)`$ is given by a combination of Eqs. (9) and (10).
Now let us discuss one possible method of detecting a total momentum of an outcoming pair of atoms. There is a significant requirement for such a detector: an uncertainty of detecting the momentum transfer must be less than the inverse correlation length $`r_c`$ in the cloud. This imposes a limitation on the detector size $`R_D`$. Indeed, let us suppose that cells sensitive to the arrival of the scattered atoms are located on a sphere of the radius $`R_D`$, and the atomic cloud of the size $`RR_D`$ is at the center of this sphere. Then, the uncertainty in the scattering angle is $`R/R_D`$. This produces the uncertainty in the momentum $`kR/R_D`$. This uncertainty must be much less than $`1/r_c`$, or
$$R_Dkr_cR.$$
(11)
It can be shown that the absolute values of the scattered momenta measured by the time of flight are also subjected to the same uncertainty. Thus, the condition (11) determines a precision of the time-of-flight measurements as well.
Our method implies that two outcoming atoms produced by one incoming atom can be identified. As mentioned above, this identification can be done due to the $`\pi /2`$ angle between the scattered atoms. However, if the incoming coherent beam produces too many scattered pairs, the erroneous identification is very likely. Let us derive a limitation on the number $`N_s`$ of the scattered pairs per typical time of flight, which would keep the momentum uncertainty less than $`1/r_c`$. The scattering events under consideration should occur with approximately equal probability at any angle inside the $`4\pi `$ solid angle, so that the angular area occupied by a single event is $`4\pi /N_s`$. It should be much less than $`2\pi /r_ck`$ which determines the angular area of the strip where the second atom of a pair can be found, the direction of the first being fixed. Thus,
$$N_s2kr_c.$$
(12)
If a typical speed of the fast atoms is 10cm/s, and $`r_c10^3`$ cm, $`R10^2`$ cm, the conditions (11) and (12) yield $`R_D10`$ cm and $`N_s10^3`$. Accordingly, a typical time of flight will be longer than 1 s.
Now let us discuss a mechanism of detection of a single neutral atom. We suggest employing resonant atomic fluorescence in the evanescent field of light propagating inside a waveguide. Specifically, light sensitive cells are mounted on a side of a long waveguide so that light propagating inside the waveguide does not excite these cells. If holes are made through the cells and the waveguide, the light will remain confined as long as a diameter of the hole is much smaller than the wavelength of light. However, a single atom may enter the hole and feel the resonant field inside the hole, or in the close vicinity of the entrance to it, due to the evanescent field of the light. Consequently, the atom will reemit resonantly one or several photons. These photons can then be detected by the nearest cell indicating an arrival of an atom at a specific location at the detector surface. Accordingly, a scattering angle of the coming atom can be identified. The velocity of the atom can be deduced from the time of flight.
A probability $`P_D`$ for atom to penetrate into a hole is simply the area $`A_h`$ of the holes per unit area of the detector exposed to the atomic flux. In order to detect an atom in the hole, the atom must reemit at least few photon which are captured by the cell mounted outside the waveguide. Before we estimate a number $`N_D`$ of the reemitted photons which can be detected, we note that the photons remitted by the atom, which has penetrated deeply inside the waveguide, should remain confined inside it, and therefore they will not be detected by the cell. Only those photons, which are reemitted by the coming atom while being close to the hole entrance, will be scattered almost isotropically and can be absorbed by the cell. Taking into account that the penetration length of the evanescent field is comparable with the wavelength $`\lambda `$ of light, one can find the time $`t_D\lambda /v`$ during which the atom moving at speed $`v`$ is subjected to the evanescent field and reemits light isotropically. Then, the number of photons reemitted during this time is $`N_D\gamma t_D=\gamma \lambda /v`$, where $`\gamma `$ stands for the natural width of the line. Choosing typical values $`\lambda =700`$nm, $`\gamma 10^7`$ Hz and $`v=10`$ cm/s one finds $`N_{ph}=70`$. Some geometrical factor of the order of one should reduce the number of photons reaching the cell. Quantum efficiency of photodetectors can be easily achieved to be 10%-20%. This implies that once an atom entered a hole, it will be detected with high certainty. Therefore, a probability to detect an atom at a given position is just $`P_D`$. In other words, if a diameter of each hole is 100 nm and two closest holes are 300 nm apart, the probability is $`P_D0.1`$. Thus, in order to achieve a resolution of at least .1 of the typical momentum region in the scattering intensity, one needs at least 10 scattering events for each component of the momentum. This corresponds to 10 atoms ejected from the cloud (plus 10 fast incident atoms). Accordingly, for the 3D geometry, it translates into $`10^3/P_D10^4`$ atoms in total. In a typical condensate of $`10^6`$ atoms , such a bombardment by fast atoms will result in a depletion of the cloud by only 1%. Evidently, this method is not appropriate for clouds containing less than $`10^5`$ atoms. In the effective 2D or 1D geometries, the same resolution requires much less numbers of the scattering events. Specifically, $`10^3`$ and $`10^2`$ of them for the 2D and 1D geometries, respectively, will satisfy the above criteria of resolution.
In conclusion, we have suggested a method of scattering of fast atoms in a pure enough momentum state off a trapped atomic cloud in order to test directly one-particle density matrix of this cloud. The differential cross-section of the inelastic process, when one incoming fast atom produces two fast ones, allows measuring the correlation length of the local off-diagonal order. This gives, in particular, a powerful tool for testing different scenarios of formation of the off-diagonal long-range order in the traps. This method can also be employed for detecting quantum vortices and supercurrent states, as well as the effect of quantum depletion of the condensate. The main principles and a design of the detector of scattered atoms are suggested.
|
no-problem/9812/astro-ph9812030.html
|
ar5iv
|
text
|
# The Andromeda Dwarf Spheroidal Galaxies
## 1. Introduction
Dwarf spheroidal galaxies are probably the most common type of galaxy in the Universe. Because of their ubiquity, and because the accretion of dwarf galaxies may have played an important role in the formation/evolution of large galaxies, an understanding of dwarf spheroidal (dSph) galaxies is key to many problems in astronomy. The great majority of our knowledge of dwarf spheroidal galaxies results from the Galaxy’s nine dSph companions. In order to broaden this knowledge and to investigate the effects of environment on the formation and evolution of dwarf galaxies, it is important to study dSph galaxies beyond the Milky Way’s companions. Some mechanisms by which environment can influence the evolution of dwarf galaxies include: tidal interactions, removal of a dwarf’s interstellar medium via ram pressure sweeping, confinement of a dwarf’s interstellar medium via ambient pressure, and photoionization of the interstellar medium by a nearby source of UV radiation (such as a quasar or AGN).
Detailed study of dwarf spheroidals beyond the Milky Way will help reveal how much we can generalize from the nine Galactic companions. For example, the Galactic dwarf spheroidals follow reasonably tight relations between central surface brightness and absolute magnitude and between absolute magnitude and mean metal abundance (e.g., Caldwell et al. 1992). Are these relations universal or do they change with environment? The present, limited data have not revealed changes with environment (Caldwell et al. 1998). As another example, the Galactic dwarf spheroidals show an amazing diversity in their star formation histories, ranging from a single star-formation episode of globular cluster age to ongoing star formation that peaked $``$3 Gyr ago (e.g., Da Costa 1998). Is this diversity universal or did special circumstances contribute to the varied star formation histories of the Milky Way’s dwarf spheroidals? Another puzzle is the (imperfect) correlation among the Galactic dwarfs between the duration of star formation and distance from the Galactic center, with the outer dwarfs continuing to form stars over longer periods than the inner dwarfs (van den Bergh 1994). Again, it would be valuable to understand whether this trend is seen in other systems. Finally, what is the behavior of the galaxy luminosity function at the faint end and is this dependent on environment? There is some indication that rich clusters of galaxies have a steeper slope for the faint end of the luminosity function than the Local Group, though this could be caused by incompleteness in the Local Group (Trentham 1998).
The nearest collection of dwarf spheroidals beyond the Milky Way’s halo is M31’s companions. Van den Bergh’s (1972, 1974) survey of 700 square degrees around M31 with the Palomar 48-inch Schmidt and IIIaJ plates revealed three dwarf spheroidal companions (And I, II & III). Surface brightness profiles and structural parameters have been measured for these dSphs by Caldwell et al. (1992). Color–magnitude diagrams from 4m-class ground-based telescopes are also available for all three galaxies (Mould & Kristian 1990; Konig et al. 1993; Armandroff et al. 1993). These show the brightest $``$2 magnitudes of the red giant branch and yield line-of-sight distances similar to that of M31. All these data indicate a considerable degree of similarity between these three M31 dwarf spheroidals and those of the Galaxy. The most significant developments that have occurred since the last review article devoted exclusively to the M31 dwarf spheroidals (Armandroff 1994) are: 1) color–magnitude diagrams from HST/WFPC2 that reach below the horizontal branch for And I & II; 2) searches for additional dwarf spheroidals that have revealed three new M31 companions. Consequently, this review will focus on these two areas of recent progress.
## 2. WFPC2 Studies
Although ground-based images of the M31 dSph companions can provide a large amount of useful information, such data are limited by image crowding to the brightest 2 or 3 magnitudes of the red giant branch. On the other hand, because of the Hubble Space Telescope’s superior resolution, images obtained with the WFPC2 camera aboard HST can yield accurate photometry for stars at the level of the horizontal branch, or fainter, in these galaxies. In this section we outline some of the results obtained in our team’s (other Co-I’s are N. Caldwell and P. Seitzer) program to image the M31 dSph companions with WFPC2. The results for And I have been published in Da Costa et al. (1996) while those for And II are presented here for the first time. Our observations of And III are scheduled to be executed in early 1999.
### 2.1. Observations and Photometry
Our program uses the $`F555W`$ (wide–$`V`$) and $`F450W`$ (wide–$`B`$) filters; the $`F450W`$ filter being preferred to the more common $`F814W`$ ($``$$`I`$) because it allows greater sensitivity to faint blue horizontal branch stars. For each dSph our procedure is to take a first series of deep exposures and then, a few ($``$5) days later, take a second series of exposures at the same orientation but with the field center displaced by $``$20 PC pixels. This displacement of the field center allows us to readily separate faint stars from the instrumental effects present in WFPC2 images, while the lag between the observation sets allows us to detect variable stars, particularly RR Lyrae variables. The total exposure times for each dSph are of order 2.3 hrs with the $`F555W`$ filter and 5.4 hrs with $`F450W`$.
Even when centered on the centers of these dSph galaxies, the WFPC2 frames are sufficiently uncrowded (cf. Fig. 1 of Da Costa et al. 1996) that aperture photometry techniques can be successfully employed. We then use the zeropoints and transformation equations listed in Holtzman et al. (1995) to convert the WFPC2 photometry to the standard $`V`$, $`BV`$ system, though we have a ground-based program underway to confirm the zeropoints of these data. The final color-magnitude diagram (cmd) for And I derived from our WFPC2 images is shown in Fig. 1, which is taken from Da Costa et al. (1996).
From the data presented in Fig. 1, and from the similar cmd for And II, we can draw a number of inferences. We will consider these in the following order: results that derive primarily from the morphology of the cmds; results, such as dSph line-of-sight distances, that depend on the $`F555W/V`$ zeropoint; and results, such as mean abundances and abundance dispersions, that depend on the $`F555W/V`$ zeropoint and the $`F450W/B`$ transformation and zeropoint.
### 2.2. Horizontal Branch Morphologies
It is immediately apparent from Fig. 1 that the morphology of the horizontal branch (HB) in And I is dominated by red HB stars. However, it is equally obvious that blue HB stars occur and, as Da Costa et al. (1996) have shown, RR Lyrae variables are also present. If we use a HB morphology index $`i`$ of the form $`i=b/(b+r)`$, where $`b`$ and $`r`$ are the numbers of blue and red HB stars, respectively (see Da Costa et al. 1996 for formal definition of these quantities), then $`i`$ = 0.13 $`\pm `$ 0.01 for And I. As discussed below, the mean abundance for And I is $`<`$\[Fe/H\]$`>`$ = –1.45 $`\pm `$ 0.2 dex. At this abundance, the HB morphologies of Galactic globular clusters that follow the (HB morphology, \[Fe/H\]) relation defined by the majority of such clusters, have relatively many more blue HB stars and many fewer red HB stars than does And I. For example, the standard globular cluster M5, which has a similar abundance to And I, has $`i`$ $``$ 0.75. For this reason, we can say unequivocally that And I shows the second parameter effect in the same way as do many of the Galactic dSph companions. Da Costa et al. (1996) go on to argue that in And I, the dominance of red HB stars results from a mean age for the bulk of the population that is younger than that of the Galactic globular clusters, though the presence of blue HB and RR Lyrae stars testifies to the presence of a minority older population. In other words, the And I cmd implies that star formation in this dSph has continued for an extended period. This interpretation is supported by the results of Mighell & Rich (1996) for the Galactic dSph Leo II, which has similar HB morphology and mean abundance to And I. Mighell & Rich (1996) used WFPC2 data that reaches below the main sequence turnoff to infer that the mean age of the stellar population in Leo II is $``$9 Gyr, with the majority of the star formation having occurred in a $``$4 Gyr interval about that epoch, though some older stars are also present.
Our WFPC2 cmd for And II is similar to that shown for And I in Fig. 1. In particular, once again the HB morphology is dominated by red stars though, as for And I, blue HB and candidate RR Lyrae variable stars are also present. Indeed, there are relatively more blue HB stars in the And II cmd than is the case for And I: the HB morphology index for And II is $`i`$ = 0.18 $`\pm `$ 0.02 (cf. $`i`$ = 0.13 $`\pm `$ 0.01 for And I). Given the somewhat higher mean abundance of And II ($`<`$\[Fe/H\]$`>`$ $``$ –1.3, see below) and its large abundance dispersion, this dSph is less obviously a strong second parameter candidate in the way And I is. Nevertheless, it is likely that the second parameter effect is at work in And II. For example, the standard Galactic globular cluster M4, which has an abundance similar to the And II mean, has a relatively evenly populated HB with $`i`$ = 0.45 $`\pm `$ 0.05, while NGC 362, which also has an abundance similar to the And II mean, has a dominant red HB morphology with $`i`$ = 0.04 $`\pm `$ 0.02. This latter cluster is one of the ‘classic’ Galactic halo second parameter clusters and has an age $``$2 Gyr younger than the majority of Galactic halo globular clusters. We conclude therefore, pending further analysis, that the And II cmd does show the second parameter effect and that the most likely cause is that the bulk of the stellar population in And II is somewhat younger than the age of most Galactic halo globular clusters.
Perhaps the most interesting result concerning HB morphology, however, is that in And I it varies with distance from the galaxy’s center. Inside the core radius we find no radial variation in the HB morphology index, but outside the core radius the relative number of blue HB stars increases by approximately a factor of two. In other words, the blue HB stars in And I are more widely distributed than the red HB stars that make up the bulk of the population. The origin of this radial gradient is not easily identified but the most likely explanation is that at the epoch of its initial episode of star formation, the proto-And I was more spatially extended than it was when the bulk of the stars formed. Apparently similar HB morphology gradients are also present in two (Sculptor and Leo II) of the three Galactic dSphs where there is sufficient data to consider the question. We have not yet performed the equivalent analysis on our And II data.
### 2.3. Distances
For And I, the mean magnitude of the horizontal branch is $`V`$ = 25.25 $`\pm `$ 0.04 mag. Adopting a reddening E($`BV`$) = 0.04 and the distance scale based on the horizontal branch models of Lee et al. (1990), in which $`M_V`$(HB) = 0.17\[Fe/H\] + 0.82 mag,<sup>2</sup><sup>2</sup>2This relation also underlies the $`I`$ magnitude of the red giant branch tip distance scale (cf. Da Costa & Armandroff 1990, Lee et al. 1993). the line-of-sight distance to And I is 810 $`\pm `$ 30 kpc for a mean abundance of $`<`$\[Fe/H\]$`>`$ = –1.45 $`\pm `$ 0.2 dex. This distance agrees well with that, 790 $`\pm `$ 60 kpc, derived by Mould & Kristian (1990) from the $`I`$ magnitude of the And I red giant branch tip. The most directly comparable distance determinations for M31, based on either red giant branch tip stars or RR Lyraes in the M31 halo or horizontal branch stars in M31 globular clusters, are 760 $`\pm `$ 45 kpc or 850 $`\pm `$ 20 kpc, both on the same distance scale (see the discussion in Da Costa et al. 1996). We can then conclude only that the relative M31/And I line-of-sight distance is 0 $`\pm `$ 70 kpc. Given that And I lies $``$45 kpc in projection from the center of M31, the true distance of And I from the center of M31 then lies between $``$45 and $``$85 kpc. The lower limit is smaller than the galactocentric distances of any of the Galaxy’s dSph companions with the exception of Sagittarius, while the upper limit is comparable to the galactocentric distances of the nearer Galactic dSphs such as Ursa Minor, Sculptor and Draco.
For And II, the mean magnitude of the horizontal branch is $`V`$ = 24.90 $`\pm `$ 0.04 mag from which we can immediately infer that And II lies closer to us than And I. Indeed, adopting the same distance scale and reddening as above, and assuming a mean metal abundance of $`<`$\[Fe/H\]$`>`$ $``$ –1.3 (see next section), we find that And II lies at a distance of 690 $`\pm `$ 25 kpc, or 120 $`\pm `$ 70 kpc in front of M31 along the line-of-sight. This is consistent with the ground-based results of Konig et al. (1993) who found And II to be 0.4 $`\pm `$ 0.2 mag closer than M31. On the sky And II lies closer to M33 than M31. However, our results now unambiguously associate And II with M31 since M33 lies $``$150 kpc beyond M31 along the line of sight. When combined with And II’s projected distance of $``$130 kpc from the center of M31, these estimates indicate that the true distance of this dSph from the center of M31 lies between $``$150 and $``$240 kpc. The lower limit corresponds approximately to the galactocentric distance of Fornax while the upper limit is similar to the galactocentric distances of the Leo systems, the most distant of the Galactic dSph companions. And II, however, may not be the most distant of the M31 dSph companions, since And VI (see below) lies at least $``$270 kpc (the projected distance) from M31.
### 2.4. Mean Metal Abundances and Abundance Dispersions
With the distances established, the mean metal abundances of these dSphs can be determined via the comparison of the dSph cmds with the giant branches of Galactic globular clusters whose \[Fe/H\] values are well known. An example of this is shown in Fig. 1 where the giant branches of five standard globular clusters, whose abundances range from \[Fe/H\] = –2.1 to \[Fe/H\] = –0.7, are overlaid on the And I WFPC2 observations. Da Costa et al. (1996) discuss the mean abundance determination process in some detail and their conclusion is that the mean abundance of And I is $`<`$\[Fe/H\]$`>`$ = –1.45 $`\pm `$ 0.2 dex. The uncertainty given includes uncertainty in the calibration relation, the statistical error in the observed giant branch mean color, and the effect of an adopted $`\pm `$0.03 mag systematic uncertainty in the $`BV`$ color zeropoint. This value, however, is in good accord with that, $`<`$\[Fe/H\]$`>`$ = –1.4 $`\pm `$ 0.2, determined by Mould & Kristian (1990) from the mean $`(VI)_0`$ color of And I’s upper red giant branch. Further, there does not appear to be any change in this mean abundance with distance from the center of And I.
Application of the same technique to the And II WFPC2 cmd yields a mean abundance of $`<`$\[Fe/H\]$`>`$ $``$ –1.3, with an uncertainty comparable to that derived for And I. This mean abundance is somewhat more metal-rich than previous ground-based determinations. For example, Konig et al. (1993) give $`<`$\[Fe/H\]$`>`$ = –1.59 (+0.44,–0.12) from their Gunn system $`g`$,$`r`$ cmd study while Armandroff (1994) lists $`<`$\[Fe/H\]$`>`$ = –1.63 $`\pm `$ 0.25 from the mean giant branch color in a $`(I,VI)`$ cmd. This relatively high mean abundance for And II derived from the WFPC2 data has an interesting consequence: at $`<`$\[Fe/H\]$`>`$ $``$ –1.3 and $`M_V`$ $``$ –11.6 (which is based on the new distance derived above), And II now deviates significantly from the relation between mean abundance and absolute magnitude apparently followed by most dSph and dE galaxies (cf. Fig. 18 of Caldwell et al. 1998). For example, this mean abundance for And II exceeds that of the Fornax dSph, a system which is more than two magnitudes more luminous.
The lack of image crowding on the WFPC2 images permits the photometric errors, unlike the case for ground-based images, to approach the photon statistics limit for all but the very brightest stars, where systematic effects seem to limit the precision to $``$0.015 mag (see Da Costa et al. 1996). Consequently, investigating the intrinsic color width of the giant branch in these WFPC2 cmds is relatively straightforward. The occurrence of internal ranges in abundance is one of the characteristics of the Galactic dSph galaxies and the M31 dSph companions prove to be no exceptions. For And I, the results of Da Costa et al. (1996) can be summarized as follows: on the upper part of the red giant branch, there is an intrinsic color spread that exceeds by a substantial factor the color spread expected on the basis of the photometric errors alone. Applying the same calibration as for the mean abundance, this color spread can be characterized in a number of ways. For example, $`\sigma `$(\[Fe/H\]) = 0.20 dex, the inter-quartile range is 0.30 dex, the central two-thirds of the sample abundance range is 0.45 dex, and the full range of the sample is $``$0.6 dex. These values are comparable to those seen in Galactic dSphs. For example, Suntzeff (1993) lists values of $`\sigma `$(\[Fe/H\]) for five Galactic dSphs that range from $``$0.2 to $``$0.3 dex.
On the other hand, Konig et al. (1993) have suggested that the internal abundance range in And II is unusually large, with $`\sigma `$(\[Fe/H\]) $``$ 0.43 dex, though they admit that their photometric errors, which have to be subtracted in quadrature from their observed color dispersion, are large and not well constrained. Thus the validity of their result is questionable. However, our WFPC2 data appear also to indicate that the intrinsic abundance distribution in this dSph is quite broad, much broader than could be explained by the photometric errors alone. Further, because we have essentially two sets of data resulting from the two different pointings, our photometric errors are well determined (and small). Applying a similar analysis technique to that used by Da Costa et al. (1996) for And I, we find that $`\sigma `$(\[Fe/H\]) $``$ 0.38 dex for And II, while the inter-quartile range is 0.65 dex, the central two-thirds of the sample abundance range is 0.85 dex and the full range of the sample is $``$1.4 dex. These numbers are significantly larger than those for And I. We thus have the interesting situation of two M31 companion dSphs, of similar luminosity and mean abundance, yet whose internal abundance ranges appear to be substantially different. This result reinforces the point that the internal abundance distributions of the Local Group dSphs apparently do not seem to readily correlate with luminosity in the way that the mean abundances do. Indeed the internal abundance distribution is a characteristic that has the potential to provide additional significant constraints on the star formation and enrichment histories in these galaxies. Further discussion of this point, however, is beyond the scope of this contribution.
## 3. New Dwarf Companions to M31
### 3.1. Searches for Additional M31 Dwarf Spheroidals
And I, II & III were found by van den Bergh (1972, 1974) using the Palomar Schmidt and IIIaJ plates to survey 700 square degrees around M31. A number of surveys for nearby dwarf galaxies have been undertaken in recent years. Irwin (1994) reported a survey for nearby dwarfs using automated star counts on sky survey plates covering over 20,000 square degrees of high-latitude sky. This survey revealed the Sextans dwarf spheroidal (Irwin et al. 1990). Whiting et al. (1997) visually examined survey plates of the entire southern sky, finding the Antlia dwarf. Neither of these surveys would be expected to reveal new M31 companions because, in the first instance, M31 is too distant for its companions to resolve into stars on sky survey plates and because, in the second instance, M31 is a northern object.
Because van den Bergh’s (1972, 1974) survey yielded a total of three M31 dwarf spheroidals, when the Galaxy has nine, it has been recognized that additional M31 companions may be awaiting new surveys. Two other factors suggest that a new survey may be profitable: 1) the absolute magnitudes of And I, II & III are significantly brighter than the faintest Galactic dwarf spheroidals (see Fig. 2 of Armandroff 1994); 2) the radial extent of van den Bergh’s (1972, 1974) survey does not reach the galactocentric distances of the most distant Galactic dwarfs (see Fig. 6 of Armandroff 1994). Two surveys have recently been undertaken to find new M31 dwarfs.
The first survey, by Armandroff, Davies & Jacoby (1998a; see also Armandroff et al. 1999a,b), uses digitized images from the Second Palomar Sky Survey (POSS-II; Reid et al. 1991; Lasker & Postman 1993). The POSS-II has better resolution and depth than the POSS-I. Armandroff et al. (1998a) found that cleaning the POSS-II images of stars and bright galaxies, then applying a matched filter, easily reveals nearby low surface brightness dwarf galaxies. They optimized their filter and detection procedure for the known M31 dwarf spheroidals. To date, their detection procedure has been applied to 1550 square degrees around M31 (see Fig. 2). Once an object that resembles the known M31 dSphs on the processed and raw POSS-II images is found, small telescope CCD imaging is undertaken. This imaging eliminates most “contaminants” (e.g., distant galaxy clusters, distant low surface brightness spirals, Galactic cirrus clouds) and highlights true nearby dwarf galaxies by their incipient resolution into stars.
The second survey, by Karachentsev & Karachentseva (1999), also uses the POSS-II and is part of their larger program to search the POSS-II films for nearby dwarf galaxies (see Karachentseva & Karachentsev 1998). They visually searched a circular area of 22 radius around M31, using morphological criteria to identify nearby low surface brightness dwarf candidates.
Armandroff et al. (1998a,b, 1999a,b) discovered two M31 dwarf spheroidal candidates: And V & VI. Karachentsev & Karachentseva (1998, 1999) found two strong candidate M31 dwarf galaxies: Pegasus Dwarf and Cas Dwarf. One candidate was found independently by both groups: And VI = Pegasus Dwarf. The Cas Dwarf is located in a region that lacked POSS-II digital data; therefore it was not found by Armandroff et al. (see Fig. 2).
### 3.2. Properties of the Newly Discovered Dwarf Galaxies
Three of the candidate M31 dwarf spheroidals found by Armandroff et al. and Karachentsev & Karachentseva have been resolved into stars, indicating that they are indeed nearby dwarf galaxies. In this section, we summarize what is known about these three galaxies as of November 1998.
Armandroff et al. (1998a) presented follow-up observations of And V from the KPNO 4-m telescope and prime-focus CCD imager in the $`V`$, $`R`$, $`I`$ and H$`\alpha `$ narrow-band filters. In the broad-band filters, And V resolves nicely into stars and exhibits a smooth stellar distribution, resembling the other M31 dwarf spheroidals. And V does not exhibit the features of classical dwarf irregulars, such as obvious regions of star formation or substantial asymmetries in its stellar distribution. In the And V continuum-subtracted H$`\alpha `$ image, no diffuse H$`\alpha `$ emission or H ii regions are detected. The lack of H$`\alpha `$ emission in And V reinforces the conclusion, based on And V’s appearance on the broad-band images, that it is a dwarf spheroidal galaxy rather than a dwarf irregular. And V is not detected in any of the IRAS far-infrared bands either. Because far-infrared emission traces warm dust, and because some Local Group dwarf irregular galaxies are detected by IRAS, And V’s lack of far-infrared emission serves as additional, weaker evidence that it is a dSph.
And V’s apparent central surface brightness was measured by Armandroff et al. (1998a): 25.7 mag/arcsec<sup>2</sup> in $`V`$. And V has a fainter apparent central surface brightness than And I, II & III (24.9, 24.8, and 25.3 mag/arcsec<sup>2</sup> in $`V`$, respectively; Caldwell et al. 1992). And V probably eluded detection until now due to its very dim apparent surface brightness. Armandroff et al. (1998a) also constructed a color–magnitude diagram for And V stars, in order to determine its distance and stellar populations characteristics. Color–magnitude diagrams for the parts of the images dominated by And V stars reveal a red giant branch, which is absent in the outer regions of the images. The tip of the red giant branch is well defined in the cmd and in the luminosity function. A distance was derived for And V based on the $`I`$ magnitude of the tip of the red giant branch (Da Costa & Armandroff 1990, Lee et al. 1993). On the distance scale of Lee et al. (1990), the resulting And V distance is 810 $`\pm `$ 45 kpc. As discussed in Sec. 2.3, the best and most comparable estimates of the M31 distance are 760 $`\pm `$ 45 kpc or 850 $`\pm `$ 20 kpc. This implies that And V is located at the same distance along the line of sight as M31 to within the uncertainties. And V’s projected distance from the center of M31 is 112 kpc; And I, II & III have projected M31-centric distances of 46, 144 and 69 kpc, respectively. The above line-of-sight and projected distances strongly suggest that And V is indeed associated with M31.
Armandroff et al. (1998a) compared And V’s color–magnitude diagram with fiducials representing the red giant branches of Galactic globular clusters that span a range of metal abundance (Da Costa & Armandroff 1990). Based on the position of the And V giant branch relative to the fiducials, the mean metal abundance of And V is $``$ –1.5, which is normal for a dSph (e.g., Fig. 9 of Armandroff et al. 1993). No bright blue stars are present in the And V cmd. Interpreting via isochrones, this lack of blue stars rules out any stars younger than 200 Myr in And V and is further evidence that And V is a dwarf spheroidal and not a dwarf irregular. From the luminosities and numbers of upper asymptotic giant branch stars in a metal-poor stellar system, one can infer the age and strength of its intermediate age component (Renzini & Buzzoni 1986). Using the And V field-subtracted luminosity function, there is no evidence for upper asymptotic giant branch stars that are more luminous than and redward of the red giant branch tip. Therefore, And V does not have a prominent intermediate age population; in this sense, it is similar to And I & III.
And VI = Pegasus Dwarf has been resolved into stars and studied by two groups. Armandroff et al. (1998b) imaged And VI with the KPNO 4-m telescope prime-focus CCD on 1998 January 23 in $`V`$ for 300 seconds. And VI resolved nicely into stars in this short $`V`$ image, suggesting that it is indeed a nearby dwarf galaxy. On 1998 July 15, Armandroff et al. (1999b) obtained deeper And VI images with the KPNO 4-m telescope through the $`R`$ and H$`\alpha `$ filters. The $`R`$ image of And VI exhibits a smooth stellar distribution and shows a resemblance with the other M31 dwarf spheroidals. And VI does not look lumpy or show obvious regions of star formation, suggesting that it is a dwarf spheroidal as opposed to a dwarf irregular. In the And VI continuum-subtracted H$`\alpha `$ image, no diffuse H$`\alpha `$ emission or H ii regions are detected. The lack of H$`\alpha `$ emission rules out current high-mass star formation in And VI and serves as further evidence that And VI is a dwarf spheroidal. Like And I, II, III & V, And VI is not detected in any of the IRAS far-infrared bands. Armandroff et al. (1999b) are constructing a cmd for And VI based on images obtained with the WIYN 3.5-m telescope during excellent seeing conditions ($`B`$, $`V`$ & $`I`$).
Grebel & Guhathakurta (1999) have presented a cmd for And VI based on images from the Keck-II telescope in $`V`$ and $`I`$, which clearly shows the red giant branch. Based on the $`I`$ magnitude of the red giant branch tip, they determined a line-of-sight distance of 780 $`\pm `$ 40 kpc, associating And VI with M31. From the position of the red giant branch relative to standard globular cluster fiducials, Grebel & Guhathakurta found $`<`$\[Fe/H\]$`>`$ = –1.2 $`\pm `$ 0.3. They also measured an intrinsic central surface brightness of 24.5 mag/arcsec<sup>2</sup> in $`V`$.
Grebel & Guhathakurta (1999) also obtained similar data for the Cas Dwarf from Keck-II. As for And VI, their cmd clearly shows the red giant branch, and from it they derived a distance and mean metallicity. Their line-of-sight distance of 710 $`\pm `$ 35 kpc places the Cas Dwarf in the extended M31 satellite system. The resulting metallicity is $`<`$\[Fe/H\]$`>`$ = –1.3 $`\pm `$ 0.3. Grebel & Guhathakurta also measured an extinction-corrected central surface brightness of 23.6 mag/arcsec<sup>2</sup> in $`V`$. Both And VI and the Cas Dwarf have apparent central surface brightnesses that are much brighter than And V, and somewhat brighter than And I, II & III. Both And VI and the Cas Dwarf are outside of van den Bergh’s (1972, 1974) survey area, which likely accounts for their anonymity until now.
Once accurate $`M_V`$ values are measured for And V, And VI and Cas, and once their $`<`$\[Fe/H\]$`>`$ values are established with definity, it will be important to re-evaluate whether the M31 and Galactic dSphs follow the same relations between absolute magnitude, central surface brightness, and mean metal abundance. The sample will be twice as large and the data of higher quality than the previous determinations (e.g., Fig. 7 of Armandroff 1994).
### 3.3. Impact
The discovery of the M31 dwarf spheroidals And V, And VI = Pegasus Dwarf, and Cas Dwarf doubles the number of known M31 dSphs. This changes the properties of M31’s satellite system, as discussed below. The most obvious change is that M31 is not as poor in dwarf spheroidals as previously thought.
Karachentsev (1996) discussed the spatial distribution of the companions to M31. The discovery of And V, And VI and Cas changes the spatial distribution of the M31 satellites. Curiously, And I, II & III are all located south of M31, while the three more luminous dwarf elliptical companions NGC 147, 185 & 205 are all positioned north of M31. Also, Karachentsev (1996) noted that there are more M31 companions overall south of M31 than north of M31. The locations of And V and Cas north of M31 lessen both of these asymmetries (though And VI is south of M31; see Fig. 2). With projected radii from the center of M31 of 112, 271 and 224 kpc, respectively, And V, And VI and Cas increase the mean projected radius of the M31 dwarf spheroidals from 86 kpc to 144 kpc.
The discovery of nearby dwarf galaxies like And V, And VI and Cas augments the faint end of the luminosity function of the Local Group. We do not yet have reliable $`M_V`$ values for And V, And VI or Cas, but they appear to be in the range $`12<M_V10`$. This increases the number of galaxies in the Local Group in this $`M_V`$ range by 27%. From a survey of nine clusters of galaxies, Trentham (1998) derived a composite luminosity function that is steeper at the faint end than that of the Local Group (see his Fig. 2). He attributed the difference to poor counting statistics and/or incompleteness among the Local Group sample. The discovery of And V, And VI and Cas reduces somewhat the discrepancy between the Local Group luminosity function and the extrapolation of Trentham’s (1998) function.
## 4. Future Directions
Several opportunities for advancing our understanding of M31’s dwarf spheroidal companions are apparent:
* The search for dSph companions to M31 is not yet complete. The ongoing searches based on the POSS-II may reveal additional companions. The Sloan Digital Sky Survey should allow a significantly deeper search for faint, low surface brightness M31 companions.
* It will be valuable to obtain HST/WFPC2 cmds for all six M31 dwarf spheroidals. These cmds will yield precise line-of-sight distances, hopefully facilitating a 3-dimensional map of the distribution of M31 satellites. Whether this distribution is random or exhibits “streams” will provide an interesting comparison with the Galactic dSph distribution. Also, having HB morphologies for all six known M31 dSphs will allow a more comprehensive evaluation of the second parameter effect in the halo of M31.
* No information is currently available about the H i content of And II, And V, And VI or Cas Dwarf via 21 cm observations. Either H i detections or strict upper limits would be valuable. The upper limits for And I & III should be refined (see Thuan & Martin 1979).
* Measurements of radial velocity are possible for red giants at the M31 distance with 8–10 meter telescopes. One can constrain the total mass of M31 via the velocity dispersion of its dwarf satellite system (as, for example, Zaritsky et al. 1989 did for the Galaxy). The impressive projected radii of And VI and Cas will result in mass limits to very large radii.
* Radial velocities will also allow mass-to-light ratios (M/L) to be derived for the M31 dSphs. Measurement of a high M/L for an M31 dSph would be a step toward universalizing the large velocity dispersions and consequent large M/L values observed for Galactic dSphs. Also, the large galactocentric radii of And II, And VI and Cas should render tidal effects negligible.
#### Acknowledgments.
We are grateful to our collaborators, Nelson Caldwell and Pat Seitzer on the And II WFPC2 cmd, and James Davies and George Jacoby on the M31 dSph survey, for allowing us to present results in advance of publication and for helpful discussions.
## References
Armandroff, T.E. 1994, in ESO/OHP Workshop on Dwarf Galaxies, ed. G. Meylan & P. Prugniel (Garching: ESO), 211
Armandroff, T.E., Da Costa, G.S., Caldwell, N., & Seitzer, P. 1993, AJ, 106, 986
Armandroff, T.E., Davies, J.E., & Jacoby, G.H. 1998a, AJ, 116, 2287
Armandroff, T.E., Davies, J.E., & Jacoby, G.H. 1998b, in “Dwarf Tales,” ed. E. Brinks & E.K. Grebel, Vol. 3, p. 2
Armandroff, T.E., Davies, J.E., & Jacoby, G.H. 1999a, in The Low Surface Brightness Universe, IAU Colloquium No. 171, ed. J.I. Davies, C. Impey, & S. Phillipps (San Francisco: ASP), in press (astro-ph/9810299)
Armandroff, T.E., Jacoby, G.H., & Davies, J.E. 1999b, AJ, to be submitted
Caldwell, N., Armandroff, T.E., Da Costa, G.S., & Seitzer, P. 1998, AJ, 115, 535
Caldwell, N., Armandroff, T.E., Seitzer, P., & Da Costa, G.S. 1992, AJ, 103, 840
Da Costa, G.S. 1998, in Stellar Astrophysics for the Local Group, ed. A. Aparicio, A. Herrero, & F. Sanchez (Cambridge: Cambridge Univ. Press), 351
Da Costa, G.S., & Armandroff, T.E. 1990, AJ, 100, 162
Da Costa, G.S., Armandroff, T.E., Caldwell, N., & Seitzer, P. 1996, AJ, 112, 2576
Grebel, E.K., & Guhathakurta, P. 1999, ApJ, submitted
Holtzman, J., et al. 1995, PASP, 107, 1065
Irwin, M.J. 1994, in ESO/OHP Workshop on Dwarf Galaxies, ed. G. Meylan & P. Prugniel (Garching: ESO), 27
Irwin, M.J., Bunclark, P.S., Bridgeland, M.T., & McMahon, R.G. 1990, MNRAS, 244, 16p
Karachentsev, I. 1996, A&A, 305, 33
Karachentsev, I., & Karachentseva, V. 1998, in “Dwarf Tales,” ed. E. Brinks & E.K. Grebel, Vol. 3, p. 1
Karachentsev, I.D., & Karachentseva, V.E. 1999, A&A, in press
Karachentseva, V.E., & Karachentsev, I.D. 1998, A&AS, 127, 409
Konig, C.H.B., Nemec, J.M., Mould, J.R., & Fahlman, G.G. 1993, AJ, 106, 1819
Lasker, B.M., & Postman, M. 1993, in ASP Conf. Ser. 43, Sky Surveys: Protostars to Protogalaxies, ed. B.T. Soifer (San Francisco: ASP), 131
Lee, M.G., Freedman, W.L., & Madore, B.F. 1993, ApJ, 417, 553
Lee, Y.-W., Demarque, P., & Zinn, R. 1990, ApJ, 350, 155
Mighell, K.J., & Rich, R.M. 1996, AJ, 111, 777
Mould, J., & Kristian, J. 1990, ApJ, 354, 438
Reid, I.N., et al. 1991, PASP, 103, 661
Renzini, A., & Buzzoni, A. 1986, in Spectral Evolution of Galaxies, ed. C. Chiosi & A. Renzini (Dordrecht: Reidel), 195
Suntzeff, N.B. 1993, in ASP Conf. Ser. 48, The Globular Cluster – Galaxy Connection, ed. G.H. Smith & J.P. Brodie (San Francisco: ASP), 167
Thuan, T.X., & Martin, G.E. 1979, ApJ, 232, L11
Trentham, N. 1998, MNRAS, 294, 193
van den Bergh, S. 1972, ApJ, 171, L31
van den Bergh, S. 1974, ApJ, 191, 271
van den Bergh, S. 1994, ApJ, 428, 617
Whiting, A.B., Irwin, M.J., & Hau, G.K.T. 1997, AJ, 114, 996
Zaritsky, D., Olszewski, E.W., Schommer, R.A., Peterson, R.C., & Aaronson, M. 1989, ApJ, 345, 759
|
no-problem/9812/astro-ph9812162.html
|
ar5iv
|
text
|
# Multiwavelength Observations of GX 339–4 in 1996. III. Keck Spectroscopy
## 1 Introduction
Most Galactic black hole candidates exhibit at least two distinct spectral states (see Liang & Narayan 1997, Liang 1998, Poutanen 1998 for reviews). In the hard state (= soft X-ray low state) the spectrum from $``$ keV to a few hundred keV is a hard power law (photon index $`1.5\pm 0.5`$) with an exponential cutoff. This can be interpreted as inverse Comptonization of soft photons. In the soft state (often, but not always, accompanied by the soft X-ray high state), the spectrum above $`10`$ keV is a steep power law (photon index $`>2.2`$) with no detectable cutoff out to $``$ MeV. This multi-state behavior is seen in both persistent sources (e.g. Cygnus X-1) and transient black hole X-ray novae (BHXRN, e.g. GRS 1009–45). GX 339–4 is unusual in that it is a persistent source, being detected by X-ray telescopes most of the time, but it also has nova-like flaring states.
In 1996, we performed a series of multiwavelength observations of GX 339–4 when it was in a very low state. This paper is one of a series that describe the results of this campaign. In Paper I (Smith et al. (1999)) we discuss the radio, X-ray, and gamma-ray daily light curves and spectra obtained in 1996 July. In Paper II (Smith & Liang (1999)) we discuss the X-ray timing analysis from 1996 July. In Böttcher, Liang, & Smith (1998) we use the GX 339–4 spectral data to test our detailed self-consistent accretion disk corona models. These papers expand significantly on our preliminary analyses (Smith et al. 1997a ; Smith et al. 1997b ).
In this paper we focus on Keck spectroscopy performed on 1996 May 12 UT. Beginning with the discovery observations in the 1970s (Penston et al. (1975); Doxsey et al. (1979); Grindlay (1979)) it was clear that the optical counterpart to GX 339–4 is highly variable (e.g. Motch, Ilovaisky, & Chevalier (1982); Motch et al. (1983, 1985); Callanan et al. (1992)). For example, one month prior to our Keck observations, when the source was in a similar low state, it was found to have 16 second optical quasi-periodic oscillations (Steiman-Cameron et al. (1997)).
In previous spectroscopic observations, the optical emission has always been dominated by the highly variable accretion disk, preventing a direct observation of the spectrum of the companion star. This is also true of our Keck observations. For the first time we are able to resolve the $`\mathrm{H}\alpha `$ emission line, showing it is double peaked. This is similar to the $`\mathrm{H}\alpha `$ emission lines from the quiescent optical counterparts of many BHXRN. However, we show that the peak separation in GX 339–4 is much smaller than in the BHXRN. Instead, the $`\mathrm{H}\alpha `$ emission line may be more akin to the one in Cygnus X-1.
## 2 Observations and Reductions
Our Keck observations were performed on 1996 May 12 UT (MJD 50215). At this time, the ASM on the Rossi X-Ray Timing Explorer did not detect the source, with count rates of $`0.636\pm 0.395`$ cps in the 1.3 – 3.0 keV band, $`0.027\pm 0.331`$ cps in the 3.0 – 5.0 keV band, and $`0.420\pm 0.403`$ cps in the 5.0 – 12.1 keV band. Similarly, using the Earth occultation method, BATSE on the Compton Gamma-Ray Observatory also did not detect it with flux $`0.0024\pm 0.0061\mathrm{photons}\mathrm{cm}^2\mathrm{s}^1`$ in the 20 – 100 keV band. The source was very faint or not detected for weeks on either side of this time: see Figure 1 of Smith et al. (1997a) for the RXTE light curve. This is as close to “quiescence” as the high-energy emission comes in GX 339–4.
Our observations of GX 339–4 were made during one of the same runs in which we studied the BHXRN Nova Ophiuchi 1977 in quiescence and found its orbital period and mass function; see Filippenko et al. (1997) and Harlaftis et al. (1997) for full instrumental and reduction details. The Low-Resolution Imaging Spectrometer was used at the Cassegrain focus of the Keck I telescope. Two consecutive 400 second exposures were taken, starting at 11:34 and 11:41 UT. A 1″ slit was used, but the seeing was poor (2.2″) and variable, probably due to the high airmass (2.8) and possibly affected by thin clouds.
## 3 Results
### 3.1 Whole Spectrum
The original goal of our observations was to study the source in optical quiescence to determine the nature of the companion star; it was hoped that the lack of detectable high-energy emission would make this possible. However, we found that the optical emission was still dominated by the accretion disk. Figure 1 shows the two consecutive 400 s spectra. The data are not fully photometric but give $`\mathrm{V}17`$ mag. This is typical for the source in the low state, $`\mathrm{V}15.420.2`$: here we include both the results reported for the traditional “soft X-ray low” state ($`\mathrm{V}15.417`$) and the sometimes discussed “off” state ($`\mathrm{V}17.720.2`$), since it is now generally believed that the off state is a lower luminosity low hard state (Van der Klis (1995)).
There is no artificial offset in the two curves in Figure 1. The ratio of the spectra is consistent with being constant across the whole range, including the emission lines. Such a variability and constant spectral shape agrees with previous observations. For example, in the high state in 1986 June/July it was found that there were large changes in $`V`$, but no accompanying variations in the (B V) index (Corbet et al. (1987)). However, since the slit was narrow relative to the seeing conditions at the time of the observation, we can make no accurate statements regarding the true variability of the source.
The Keck continua can be approximately fit by a linear relationship. Extrapolating this, we find that the flux density goes to zero at $`4000`$ Å, easily consistent with the lack of observed X-ray emission (5 keV corresponds to 2.45 Å). However, given the relatively narrow wavelength range of the Keck spectrum, and a misalignment between the slit position angle (60°) and the parallactic angle (175°) which can distort the continuum shape at the edges of the spectrum (Filippenko (1982)), we cannot make conclusive statements about the spectrum at higher energies.
### 3.2 $`\mathrm{H}\alpha `$ Lines
The dominant emission line in our spectra is $`\mathrm{H}\alpha `$. Its equivalent width (EW) of $`6.5`$ Å is similar to the lines seen in previous GX 339–4 observations (Grindlay (1979); Cowley, Crampton, & Hutchings (1987)). However, this is small compared to other BHXRN; for example, the $`\mathrm{H}\alpha `$ line in Nova Oph 1977 had EW $`=85`$ Å during the same observing run (Filippenko et al. (1997)).
An expanded view of the $`\mathrm{H}\alpha `$ line in GX 339–4 is shown in Figure 2. For the first time we are able to resolve a double peaked line about the rest wavelength of 6562.80 Å in both spectra.
We used a linear fit to the continuum around the line to determine the rms noise. Adding a single Gaussian line (with unconstrained centroid) does not give a good fit in either of our spectra. For the upper spectrum, the reduced $`\chi _\nu ^2`$= 1.57 for $`\nu =51`$ degrees of freedom. The probability that a random set of data points would give a value of $`\chi _\nu ^2`$ as large or larger than this is $`Q=5.6\times 10^3`$. The lower spectrum gives very similar results: $`\chi _\nu ^2`$= 1.54, $`\nu =73`$, $`Q=2.2\times 10^3`$.
Using two unconstrained Gaussian lines greatly improves the fits. Figure 2 shows the best fit results. The fit to the upper spectrum now has $`\chi _\nu ^2`$= 1.23, $`\nu =48`$, $`Q=0.13`$, while the lower spectrum now has $`\chi _\nu ^2`$= 0.95, $`\nu =70`$, $`Q=0.59`$. It therefore appears that the $`\mathrm{H}\alpha `$ line in GX 339–4 is double peaked, as is the case for the quiescent optical counterparts of many BHXRN.
The best two Gaussian fits have different fit parameters for the two spectra. However, if we multiply the fit to the upper spectrum by a constant 1/1.062 given by the ratio of the continua shown in Figure 1, we find that this gives an adequate fit to the lower spectrum: $`\chi _\nu ^2`$= 1.11, $`Q=0.25`$. Thus we are unable to claim there is significant variability in the $`\mathrm{H}\alpha `$ lines in these data. Searching for this variability is the subject of an ongoing project, at which point it will also be possible to use more realistic emission-line profiles (e.g. Orosz et al. 1994, and references therein).
The separation of the peaks is $`\mathrm{\Delta }\lambda =8.0\pm 0.8`$ Å, implying a $`\mathrm{\Delta }v=370\pm 40\mathrm{km}\mathrm{s}^1`$. If we assume this peak emission comes from a circular Keplerian orbit, it would be at a distance $`4\times 10^{11}(M/M_{\mathrm{}})`$ cm, where $`M`$ is the mass of the black hole. This would be in the outer regions of the accretion disk. This distance can be contrasted with the suggestion by Motch et al. (1982) that the 20 second optical QPO they detected came from a ring orbiting with a Keplerian period of 20 seconds at a radius of $`10^9`$ cm.
### 3.3 He I Lines
The only other emission lines in the spectra are from He I. For $`\lambda `$5875, the combined equivalent width is $`1.3\pm 0.3`$ Å, although this is difficult to fit accurately because of the adjacent NaD interstellar absorption lines. For $`\lambda `$6678, the combined equivalent width is $`1.0\pm 0.1`$ Å.
### 3.4 Li I Line
We do not see any evidence for an absorption line from Li I at $`\lambda `$6708 Å. This result agrees with previous observations of GX 339–4 (Sood, James, & Durouchoux (1997)).
## 4 Comparison with other Galactic Black Hole Candidates
A double peaked $`\mathrm{H}\alpha `$ emission line has been seen in several BHXRN in quiescence, showing that there is optical emission from their accretion disks long after the X-ray nova event. Table 1 lists representative values for their peak separations. The remarkable feature that has already been noted in previous studies is that the peak separations in these different sources are all surprisingly similar.
The $`\mathrm{H}\alpha `$ peak separation in GX 339–4 clearly shows that this source is different from the usual BHXRN. This may not be too surprising. Although GX 339–4 had no detectable high-energy emission around the time of our Keck observation, the source did not exhibit a very long period of “quiescence” in the same way as the other BHXRN. The narrower $`\mathrm{H}\alpha `$ peak separation in GX 339–4 implies that the optical emission comes from a larger Keplerian radius than in the BHXRN, which may be a clue to its different behavior.
GX 339–4 is most often compared to Cygnus X-1. Unfortunately, for Cygnus X-1 the absorption and emission lines from the companion star dominate, and only the wings of the $`\mathrm{H}\alpha `$ emission line are detected (Canalizo et al. (1995); Sowers et al. (1998)). It is therefore very difficult to determine if the $`\mathrm{H}\alpha `$ emission line in Cygnus X-1 is double peaked. However, the width of the base of the $`\mathrm{H}\alpha `$ emission line in Cygnus X-1 is quite similar to that in GX 339–4, suggesting that the optical emission from their accretion disks may be similar.
It is also interesting to note that the peak separation was $`350550\mathrm{km}\mathrm{s}^1`$ in the $`\mathrm{H}\alpha `$ emission lines from GRO J1655–40 during outbursts (Soria et al. (1998)). This is a superluminal jet source, and may indicate a connection to the suggested radio jet in GX 339–4 (Fender et al. (1997)).
We thank Philippe Durouchoux for suggesting we check for a Li I line. We also thank the referee for carefully reading the manuscript and providing useful suggestions and clarifications. This work was supported by NASA grants NAG 5-1547 and 5-3824 at Rice University, and NSF grant AST-9417213 at UC Berkeley. This work made use of the RXTE ASM data products provided by the ASM/RXTE teams at MIT and at the RXTE SOF and GOF at NASA’s Goddard Space Flight Center. The BATSE daily light curves were provided by the Compton Observatory Science Support Center at NASA’s Goddard Space Flight Center.
|
no-problem/9812/hep-ph9812460.html
|
ar5iv
|
text
|
# Introduction
## Introduction
Since the discovery of charm in 1974 numerous experiments on charm production at fixed target have been performed . In recent years the precision has increased significantly , which enables a detailed comparison with theory. Perturbative QCD can describe some of the phenomenology of charm production, but not all . Most notably the asymmetry between leading and non-leading particles, which is negligible in NLO QCD, has been shown to increase with $`x_\mathrm{F}`$ . Also the momentum spectra of produced D mesons are harder than the NLO QCD c quark predictions, especially for leading particles, see e.g. . These facts imply that nonperturbative effects are important in the production of charmed hadrons. In the string fragmentation approach both these aspects are included in the ’drag’ effect, whereby a charm quark can gain momentum when it is connected to a beam remnant. The extreme case of this effect is the collapse of a small string into a single hadron, which gives rise to a dependence on the flavour contents of the beam.
## Basics of string fragmentation
The Lund String Fragmentation Model is best explored in $`\mathrm{e}^+\mathrm{e}^{}`$ annihilation, where the produced $`\mathrm{q}`$ and $`\overline{\mathrm{q}}`$ are connected by a linear force field with a string-like topology. The $`\mathrm{q}\overline{\mathrm{q}}`$ production process is described by perturbative QCD, with a parton-shower approximation to higher orders. Radiated gluons are interpreted as ’kinks’ on the string. The nonperturbative hadronization of a string proceeds via the production of $`\mathrm{q}\overline{\mathrm{q}}`$ pairs in the colour force field, which arrange themselves to produce the observed hadrons. A strongly constrained fragmentation function can be derived from very general and physically intuitive assumptions about the fragmentation process . Because of the non-negligible mass of the charm quark the fragmentation function has to be modified for heavy-flavour production . This model has been implemented in the Monte Carlo program Pythia, which has been tuned to $`\mathrm{e}^+\mathrm{e}^{}`$ experiments to give a good description of available data.
### Hadron-hadron collisions
When we carry this model over to hadron-hadron physics, we again divide the process into a perturbative and a nonperturbative part and assume factorization between the two. In addition we assume that the fragmentation process is universal, i.e. the hadronization of a colour singlet is independent of how it was produced. In a hadron-hadron collision, such as $`\pi ^{}\mathrm{p}`$, several ambiguities not present in $`\mathrm{e}^+\mathrm{e}^{}`$ annihilation are introduced. The main ones are presented in the following.
Structure of the incoming hadrons. The particles that participate in the collision are not point-like but have an internal structure. The longitudinal structure is parameterized by parton distribution functions (PDFs) which have been determined from other experiments such as deep inelastic scattering (DIS). These will not be discussed in the following, but in principle they give rise to some ambiguity, especially for small momentum fractions.
Structure of the beam remnant. When a parton has been picked out of a hadron, what is left continues in the direction of the beam and is called a beam remnant. If the beam remnant can be viewed as consisting of two or more objects its structure must be described. How this should be done is not known from first principles and has not been studied much. This aspect is therefore parameterized in beam remnant distribution functions (BRDFs) and several variants are considered.
Primordial transverse momentum. The partons inside the hadron are confined to a transverse dimension less than 1 fm; therefore by the uncertainty principle the spread of the transverse momentum should be of the order .2 GeV. This is modeled by adding a primordial $`k_{}`$ to the partons going into the hard scattering process. In the default version of Pythia, $`k_{}`$ is assumed distributed as a Gaussian with a width of 0.44 GeV. Several studies imply that this value is too small and a value of 1 GeV or more is needed to describe the data. This remains somewhat of a mystery and will not be resolved here.
Small strings in hadronization. Quark masses. When the colour topology of an event has been determined, every colour singlet is hadronized as a string would in $`\mathrm{e}^+\mathrm{e}^{}`$. This works for strings with a mass of a few GeV or more. For strings with masses near the two-particle threshold, the standard string fragmentation approach can not be used and some other scheme is needed. As we will see later this introduces a large dependence on the quark masses.
### String topologies and the ’beam drag’ effect
In a $`\pi ^{}\mathrm{p}`$ collision we include charm production via the leading-order production mechanisms of quark and gluon fusion ($`\mathrm{q}\overline{\mathrm{q}}\mathrm{c}\overline{\mathrm{c}}`$ and $`\mathrm{gg}\mathrm{c}\overline{\mathrm{c}}`$ respectively). The partons of the hard interaction and of the beam remnants are connected by strings, representing the confining colour field . Each string contains a colour triplet endpoint, a number (possibly zero) of intermediate gluons and a colour anti-triplet end. The string topology can usually be derived from the colour flow of the hard process. For instance, consider the process $`\mathrm{u}\overline{\mathrm{u}}\mathrm{c}\overline{\mathrm{c}}`$ in a $`\pi ^{}\mathrm{p}`$ collision. The colour of the incoming $`\mathrm{u}`$ is inherited by the outgoing $`\mathrm{c}`$, so it will form a colour-singlet together with the proton remnant, here represented by a colour anti-triplet $`\mathrm{ud}`$ diquark. In total, the event will thus contain two strings, one $`\mathrm{c}`$$`\mathrm{ud}`$ and one $`\overline{\mathrm{c}}`$$`\mathrm{d}`$ (Fig. 1a). In $`\mathrm{gg}\mathrm{c}\overline{\mathrm{c}}`$ a similar inspection shows that two distinct colour topologies are possible. Representing the proton remnant by a $`\mathrm{u}`$ quark and a $`\mathrm{ud}`$ diquark (alternatively $`\mathrm{d}`$ plus $`\mathrm{uu}`$), one possibility is to have three strings $`\mathrm{c}`$$`\overline{\mathrm{u}}`$, $`\overline{\mathrm{c}}`$$`\mathrm{u}`$ and $`\mathrm{d}`$$`\mathrm{ud}`$ (Fig. 1b), and the other is the three strings $`\mathrm{c}`$$`\mathrm{ud}`$, $`\overline{\mathrm{c}}`$$`\mathrm{d}`$ and $`\mathrm{u}`$$`\overline{\mathrm{u}}`$ (Fig. 1c).
Other production mechanisms such as charm excitation and charm from parton showers (i.e. higher order effects) are not included in this study. Because of the relatively low virtuality of the hard process at fixed target energies the second mechanism is negligible, but it will become increasingly important at higher energies; at the LHC, e.g., this mechanism will dominate over the fusion mechanisms. Charm excited from the sea could give a non-negligible contribution, but we will not include it here.
Consider a colour singlet in Fig. 1 containing a charm endpoint. The hadronization of this string is performed in the CM system of the string. In that frame hadronization always results in a deceleration of the quark. After the rotation and boost to the hadron-hadron CM system, on the other hand, the net result of hadronization can be either an acceleration or a deceleration of the charm quark. This is interpreted as the beam remnant dragging the charm quark in the direction of the beam. This effect alone does not account for the asymmetry because of the cancellation between the diagrams in Fig 1b and c. We must therefore consider the flavour contents of the beam.
### Cluster collapse of small strings
In $`\mathrm{e}^+\mathrm{e}^{}`$ annihilation the string mass is fixed by the CM energy of the process. In a hadron-hadron collision, on the other hand, charmed strings can have any mass ranging from $`m_\mathrm{q}`$+$`m_\mathrm{c}`$ to $`\sqrt{s}`$. For string masses larger than some cut-off (here taken as $`m_\mathrm{q}`$+$`m_\mathrm{c}`$+$`m_0`$, with $`m_0`$$``$1 GeV and q a light quark) the Lund string fragmentation approach can be used. The model assumes a continuous final-state phase space, and an iterative scheme is used which demands that at least two particles are produced from a string. A string with a mass smaller than the cut-off we call a cluster and it is hadronized in the following way:
Cluster decay. A $`\mathrm{q}\overline{\mathrm{q}}`$ pair is created, using standard flavour selection, from the force-field connecting the cluster endpoints, and two hadrons are produced. If kinematically possible the cluster will decay isotropically into these two hadrons.
Cluster collapse. If it is not kinematically possible for the cluster to decay into two hadrons, it will be forced to collapse into a single hadron under conservation of the flavour quantum numbers. Since the mass of the cluster most likely will not correspond to any physical state (e.g. D or $`\mathrm{D}^{}`$), the energy and momentum of the cluster must be slightly modified in order to put it on the hadronic mass shell.
There are two main ambiguities in this scheme. The first is that there is no clear separation between the two hadronization mechanisms, and the second is energy/momentum conservation in the cluster collapse.
To justify the cluster collapse approach we use an argument based on local duality, which has been shown to hold in $`\mathrm{e}^+\mathrm{e}^{}`$ annihilation, DIS and $`\tau `$ decay . In $`\mathrm{e}^+\mathrm{e}^{}`$ annihilation into hadrons around the $`\mathrm{c}\overline{\mathrm{c}}`$ threshold the observed cross section consists of peaks at $`\mathrm{J}/\psi `$ and $`\psi ^{}`$ and a continuum above the $`\mathrm{D}\overline{\mathrm{D}}`$ threshold. The perturbatively calculated cross section, on the other hand, is continuous from $`\sqrt{s}=2m_\mathrm{c}`$ onwards. However, if the experimental cross section is suitably smeared, it approximately agrees with the perturbative one. Another way of stating this is that the integrated cross section (over $`\sqrt{s}`$) should be the same provided that the integration interval is suitably large. We use the same argument in the present case, by replacing $`\sqrt{s}`$ with $`M_{\mathrm{string}}`$ and the $`\mathrm{J}/\psi `$ and $`\psi ^{}`$ peaks with $`\mathrm{D}`$ and $`\mathrm{D}^{}`$. The duality argument could then be stated in the following way:
$$_{m_1}^{m_2}\frac{\mathrm{d}\sigma _{\mathrm{Partons}}}{\mathrm{d}M_{\mathrm{String}}}dM_{m_1}^{m_2}\frac{\mathrm{d}\sigma _{\mathrm{Hadrons}}}{\mathrm{d}M_{\mathrm{String}}}dM$$
(1)
Fig. 2a shows how this looks using Pythia with standard parameters. The solid line is the mass distribution of produced clusters at the parton level and the dashed one is the produced hadrons. The clusters in the gray area have collapsed into single hadrons. The parton level distribution depends on many of the parameters such as the BRDF, primordial $`k_{}`$, and quark masses. This is shown in Fig 2b. Consider e.g. an increase of the charm mass. The threshold will be shifted towards higher values and fewer clusters will be forced to collapse (the gray area is decreased). It is also possible to decrease the number of collapses from above by increasing the probability for a cluster above the $`\mathrm{D}\pi `$ threshold to decay.
Fig. 3 shows the $`x_\mathrm{F}`$ distributions for different production channels and different parameter sets. These parameters will be discussed in the following. The explanation of the leading particle asymmetry in this model is that $`\mathrm{D}^+`$ cannot be produced from cluster collapse because it has no quark in common with the beam.
### Dependence on parameters
In this section the different parameters that have already been introduced will be discussed in more detail. We start with the BRDFs. How the energy and momentum in the beam remnant should be split between the constituents is not known from first principles, so we consider two extreme cases. In the first (default) scenario, we use BRDFs similar to those in PDFs, where one object takes a small fraction of the available energy. In the other extreme, we use naive counting rules where the energy is shared democratically between the constituents. How this effects the distributions is shown in Fig. 3. Most notably the dip around $`x_\mathrm{F}0.5`$ in the cluster collapse distribution is smeared out when an even sharing is used. Also note that the asymmetry is much more sensitive to the BRDF in the proton region, where one of the constituents of the beam remnant is a diquark. In the tuned version we will use an intermediate energy sharing.
We now come to the problem of energy/momentum conservation in the cluster collapse. To understand the dependence on this we consider two mechanisms that are of opposite nature:
Old method: ’far away’. In the first scheme, energy and momentum is shuffled to the parton in the event that is farthest from the cluster, in order to minimize the recoil. In this approach the D meson is often harder than the cluster.
New method: ’local’. In this new scheme we conserve energy locally by exchanging ’gluons’ between the cluster and the string in the event that is closest to the cluster.
The details are presented in and the conclusion is that the observables are not sensitive to the details of the energy/momentum conservation scheme, except for $`x_\mathrm{F}>0.5`$, where cluster collapse dominates but data is scarce.
Looking at Fig. 3 we see that the reason for the large asymmetry using the default parameters is that they allow many clusters to collapse for $`x_\mathrm{F}>0`$. The following set of parameters are inspired by the E791 collaboration and data from WA82 and they aim at decreasing the asymmetry by decreasing the number of clusters that collapse into one particle .
* Quark masses. The charm mass used in Pythia is by default set to the current algebra one (1.35 GeV). This is the value used in e.g. Higgs physics but it is far from obvious that this is the relevant mass in the present case. We therefore use constituent quark masses: $`m_\mathrm{c}=`$ 1.5 GeV; $`m_\mathrm{d}=m_\mathrm{u}=`$ 0.33 GeV; $`m_\mathrm{s}=0.5`$ GeV. This will shift the threshold of the distribution of cluster masses and thus decrease the number of cluster collapses.
* Cluster decay. Increase the probability for a cluster to decay.
* Cluster collapse. Use new ’local’ collapse mechanism.
* BRDF. We use an intermediate energy sharing that is more democratic, but not completely.
* Primordial $`k_{}`$. The width of the Gaussian primordial $`k_{}`$ is increased from 0.44 to 1.0 GeV. This allows cluster collapses between a charm quark and a beam remnant to occur also at fairly large values of $`p_{}`$, thus leading to an essentially $`p_{}`$-independent asymmetry. In addition, the $`p_{}`$ kick added to charm quarks and beam remnants tends to increase the average invariant mass of the produced clusters, thereby reducing the number of cluster collapses.
## Comparison with experiment
In this section we compare the model (with the new parameter set ) to some data. Fig. 4 shows the asymmetry as a function of $`x_\mathrm{F}`$ and $`p_{}^2`$ as well as single-charm spectra from WA82 for $`\mathrm{D}^+`$ and $`\mathrm{D}^{}`$ individually. In this case the data is nicely described by the new parameter set.
Next we compare the model to the new correlation data from E791 . Several correlations in events where a pair of $`\mathrm{D}\overline{\mathrm{D}}`$ mesons with rapidity in $`0.5<y<2.5`$ is fully reconstructed are studied. The distributions in Fig. 5 show that the longitudinal correlations in the string model are generally different from the data and are better described by NLO QCD . The transverse correlations on the other hand are better described by the model, mostly because of the increased primordial $`k_{}`$.
It is clear that the correlation between charm quark pairs should be modified by hadronization, but the string model seems to produce D mesons that are too far from each other in momentum (rapidity). To attempt a description of the data we use the independent fragmentation approach where a charmed hadron is simply a scaled-down version of the charm quark. In this way the beam remnants do not affect the charm quarks at all. We use the fragmentation function of Peterson et. al. with $`ϵ_\mathrm{c}=0.05`$. Surprisingly the longitudinal correlations are quite nicely described by this scheme. Especially the rapidity distribution is interesting because NLO QCD also fails to describe the strong peaking at small rapidities seen in the E791 data, see Fig. 6. However, independent fragmentation of course fails to describe the single-charm spectra from WA82, falling way below the data and lacking any asymmetry at large $`x_\mathrm{F}`$. Thus there seems to be a contradiction between the data on single-charm and correlations that we cannot explain. There are several possibilities that deserve to be investigated. There could be some cut in the data that we don’t understand, or problems with acceptance corrections at large $`x_\mathrm{F}`$. Alternatively, some other mechanisms that we have not yet included (sea, intrinsic charm, parton showers) could give a sizable contribution.
## Summary
To summarize we have described the string fragmentation approach to charm production in hadronic collisions. A number of uncertainties have been identified and studied in detail, in particular the transition from a continuous string-mass distribution to a discrete hadron-mass one. The conclusion is that the model can describe asymmetries, single-charm spectra and transverse correlations but not longitudinal correlations. Also, these data do not fully constrain the choice of model parameters. Further data on charm production in $`\pi ^{}\mathrm{p}`$ collisions may provide further information, as may charm production e.g. in $`\mathrm{ep}`$ collisions.
|
no-problem/9812/cs9812018.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
In order to support the efficient development of NL generation systems, two orthogonal methods are currently pursued with emphasis: (1) reusable, general, and linguistically motivated surface realization components, and (2) simple, task-oriented template-based techniques. Surface realization components impose a layer of intermediate representations that has become fairly standard, such as the Sentence Plan Language (SPL) \[Kasper and Whitney (1989)\]. This layer allows for the use of existing software with well-defined interfaces, often reducing the development effort for surface realization considerably. Template-based techniques recently had some sort of revival through several application-oriented projects such as idas \[Reiter et al. (1995)\], that combine pre-defined surface expressions with freely generated text in one or another way. However, the benefits of both surface realization components and template-based techniques are still limited from an application-oriented perspective. Surface realization components are difficult to use because of the differences between domain-oriented and linguistically motivated ontologies (as in SPL), and existing template-based techniques are too inflexible.
In this paper we suggest and evaluate flexible shallow methods for report generation applications requiring limited linguistic resources that are adaptable with little effort. We advise a close connection between domain-motivated and linguistic ontologies, and we suggest a layer of intermediate representation that is oriented towards the domain and the given task. This layer may contain representations of different granularity, some highly implicit, others very elaborate. We show how this is used by the processing components in a beneficial way.
The approach suggested does not only change the modularization generally assumed for NLG systems drastically, it also renders the system much more application-dependent. At first glance, however, such an approach seems to abandon generality and reusability completely, but, as we will demonstrate, this is not necessarily the case.
The rest of this paper is organized as follows: Section 2 identifies deficits with current approaches to surface realization that may occur for particular applications. In Section 3 we propose alternative methods implemented into our sample application, the generation of air-quality reports from current environmental data. In Section 4 we discuss the pros and cons of our approach, and we summarize the conditions for successful use.
## 2 In-Depth and Shallow Generation
### 2.1 Shallow generation
Recently, the distinction between in-depth and shallow approaches to language processing has emerged from the need to build sensible applications. In language understanding deep analysis attempts to “understand” every part of the input, while shallow analysis tries to identify only parts of interest for a particular application. Shallow analysis is a key concept for information extraction from huge text bases and many other real-world application types.
In language generation a corresponding distinction which we term in-depth vs. shallow generation<sup>2</sup><sup>2</sup>2We thus avoid confusion with the common distinction between deep and surface generation. is becoming prominent. While in-depth generation is inherently knowledge-based and theoretically motivated, shallow generation quite opportunistically models only the parts of interest for the application in hand. Often such models will turn out to be extremely shallow and simple, but in other cases much more detail is required. Thus, developing techniques for varying modeling granularity according to the requirements posed by the application is a prerequisite for more custom-tailored systems.
According to Reiter and Mellish, shallow techniques (which they call “intermediate”) are appropriate as long as corresponding in-depth approaches are poorly understood, less efficient, or more costly to develop \[Reiter and Mellish (1993)\]. While our motivation for shallow techniques is in essence based on the cost factor, our assessment is even more pronounced than Reiter’s and Mellish’s in that we claim that shallow approaches combining different granularity in a flexible way are better suited for small applications. We are convinced that shallow generation systems will have a similar impact on the development of feasible applications as shallow analyzers.
### 2.2 Potential shortcomings of approaches to surface realization
Current approaches to surface realization are mostly in-depth, based on general, linguistically motivated, and widely reusable realization components, such as Penman \[Penman (1989)\], KPML \[Bateman (1997)\], and SURGE \[Elhadad and Robin (1996)\]. These components are domain-independent and based on sound linguistic principles. KPML and SURGE also exhibit a broad coverage of English, while several other language models are also available or under development. Despite their being reusable in general, the fact that the modularization of grammatical knowledge follows linguistic criteria rather than the needs of different types of applications may cause a number of problems for an efficient development of concrete applications:
* The substantial differences between domain- and linguistically motivated ontologies may render the mapping between them difficult; for instance, the use of case relations such as “agent” or “objective” requires compatible models of deep case semantics.
* The need to encapsulate grammar knowledge within the surface realizer may require details in the intermediate representation to be spelled out that are irrelevant to the intended application, even for rather small systems.
* The fixed granularity of grammatical modeling requires a realizer to cover many more languages, language fragments, or stylistic variations than would be needed for one particular application, which can lead to a considerable inefficiency of the realizer.
In addition, there may be linguistic constructs needed for some applications that are still outside the scope of the general tool. Their inclusion may require the intermediate representation layer to be modified.
### 2.3 Potential shortcomings of shallow generation methods
A prominent example for an early shallow generation system is Ana \[Kukich (1983)\], which reports about stock market developments. While the kind of texts it produces can still be considered valuable today, Ana is implemented as a widely unstructured rule-based system, which does not seem to be easily extendable and portable. Since then, various shallow methods including canned text parts and some template-based techniques have been utilized, e.g. in CogentHelp \[White and Caldwell (1997)\], in the system described in \[Cawsey et al. (1995)\], and in idas \[Reiter et al. (1995)\]. They feature simplicity where the intended application does not require fine-grained distinctions, such as the following techniques used in idas:
* canned text with embedded KB references (“Carefully slide \[x\] out along its guide”),
* case frames with textual slot fillers, (“gently” in (manner: "gently")).
Although these techniques seem to be able to provide the necessary distinctions for many practical applications in a much simpler way than in-depth surface realization components can do, a serious limitation lies in their inflexibility. The first example above requires the realization of \[x\] to agree in number with the canned part; as this is not explicitly treated, the system seems to implicitly “know” that only singular descriptions will be inserted. Moreover, canned texts as case role fillers may bear contextual influence, too, such as pronominals, or word order phenomena. Thus, the flexibility of shallow generation techniques should be increased significantly.
## 3 Shallow Generation in TEMSIS
In order to tailor the design of a generation system towards an application, we must account for different levels of granularity. We need a formalism capable of adapting to the expressivity of the domain-oriented information. Parts of the texts to be generated may be canned, some require templates, others require a more elaborate grammatical model.
In this section we first introduce an instance of the kind of applications we have in mind. We then proceed by discussing aspects of different granularity from the point of view of the intermediate representation (IR) layer and the components it interfaces. These include text organization and text realization. The text organizer is also responsible for content selection. It retrieves the relevant data from the TEMSIS database. It combines fixed text blocks with the results of the realizer in a language-neutral way. IR expressions are consumed by the text realizer, which is a version of the production system TG/2 described in \[Busemann (1996)\].
### 3.1 The TEMSIS application
With TEMSIS a Transnational Environmental Management Support and Information System was created as part of a transnational cooperation between the communities in the French-German urban agglomeration, Moselle Est and Stadtverband Saarbrücken. Networked information kiosks are being installed in a number of communities to provide public and expert environmental information. The timely availability of relevant environmental information will improve the planning and reactive capabilities of the administration considerably. Current measurement data are made available on the TEMSIS web server. The data include the pollutant, the measurement values, the location and the time the measurements were taken, and a variety of thresholds. Besides such data, the server provides metadata that allow for descriptions of the measuring locations, of the pollutants measured and of regulations or laws according to which a comparison between measurements and thresholds can be performed. This information can be accessed via the internet through a hyperlink navigation interface (http://www-temsis.dfki.uni-sb.de/).
The verbalization of NL air quality information in German and French is an additional service reducing the need to look up multiple heterogeneous data. The generated texts can be complemented with diagrams of time series. The material can be edited and further processed by the administrations to fit additional needs.
In order to request a report, a user specifies his demand by choosing from a hierarchy of options presented to him within the hyperlink navigation interface. He selects a report type by indicating whether he is interested in average values, maximum values, or situations where thresholds are exceeded. Further choices include the language, the country the environmental legislation of which should apply, the measurement location, the pollutant, the period of time for which measurements should be retrieved, and in some cases comparison parameters. In addition, descriptions of pollutants and measurement stations can be requested. They are stored as canned texts in the TEMSIS database. Not all choices are needed in every case, and the TEMSIS navigator restricts the combination of choices to the meaningful ones.
Let us assume that the user wants a French text comparing thresholds for sulfur dioxide with measurements taken in the winter period of 1996/97 at Völklingen City, and the applicable legislation should be from Germany. He also wants a confirmation of some of his choices. The user receives the following text on his browser<sup>3</sup><sup>3</sup>3A demo version of the system is available at http://www.dfki.de/service/nlg-demo/. (translated into English for the reader’s convenience):
> You would like information about the concentration of sulfur dioxide in the air during the winter season 1996/97. At the measurement station of Völklingen City, the early warning threshold for sulfur dioxide at an exposition of three hours (600 $`\mu `$g/m<sup>3</sup> according to the German decree “Smogverordnung”) was not exceeded. In winter 1995/96, the early warning threshold was not exceeded either.
Reports are organized into one or several paragraphs. Their length may range from a few lines to a page.
### 3.2 The intermediate representation
The main purpose of the IR layer for the report generation system consists in ensuring that all facets of the domain with their different degrees of specificity can be verbally expressed, and in keeping the realization task simple when no or little variety in language is needed. While SPL and similar languages interfacing to in-depth surface realization are either linguistic in nature or largely constrain the surface form of an utterance, the IR specifies domain information to be conveyed to the user and logical predicates about it. Abstracting away from language-specific information in the IR like this has the additional advantage that multi-lingual aspects can be kept internal to the realizer. They depend on the LANGUAGE feature in an IR expression.
The IR in Figure 1 roughly corresponds to the key statement of the sample report in the previous section (the second sentence), which also appears at the end of each report as a summary. It constitutes a threshold comparison, as stated by the value of the COOP<sup>4</sup><sup>4</sup>4The COOP value can correspond to the report type, as in the example, to confirmations of user choices, or to meta comments such as an introductory statement to a diagram, generated by a dedicated component. slot. There is only little indication as to how IR expressions should be expressed linguistically. Many semantic relations between the elements of an IR expression are left implicit. For instance, the value of DURATION relates to the time of exposure according to the threshold’s definition and not to the period of time the user is interested in (TIME). Another example is the relation between EXCEEDS and THRESHOLD-VALUE, which leads to the message that the early warning threshold was not exceeded at all. Wordings are not prescribed. For instance, our sample IR does not contain a basis for the generation of “exposure” or “measurement station”.
IR expressions contain specifications at different degrees of granularity. For coarse-grained specifications, it is up to the text realizer to make missing or underspecified parts explicit on the surface so that, in a sense, shallow text realization determines parts of the contents. For more fine-grained specifications, such as time expressions, text realization behaves like a general surface generator with a fully-detailed interface. Ensuring an appropriate textual realization from IR expressions is left to the language template design within the realizer.
The syntax of IR expressions is defined by a standard Backus-Naur form. All syntactically correct expressions have a compositional semantic interpretation and can be realized as a surface text provided corresponding realization rules are defined. Sharing the IR definitions between the text organization and the realization component thus avoids problems of realizability described in \[Meteer (1992)\].
### 3.3 Text organization
The goal of text organization in our context is to retrieve and express, in terms suitable for the definition of the IR, (1) report specifications provided by the user, (2) the relevant domain data accessed from the database according to these specifications, including e.g. explicit comparisons between measurements and threshold values, and (3) implicitly associated meta-information from the database, such as the duration of exposure, the decree and the value of the threshold. This task is accomplished by a staged process that is application-oriented rather than based on linguistically motivated principles.
The process starts with building some sort of a representation sketch, by instantiating a report skeleton that consists of a sequence of assertion statement specifications. Assertion statements consist of a top level predicate that represents the assertion’s type (e.g. threshold-exceeding) and encapsulates the entire meaning of the associated assertion, except to attached specifications and domain data, to make local parameters and data dependencies explicit.
In order to transform this initial representation to meet the application-oriented requirements of the IR, it is necessary to recast the information, which comprises augmenting, restructuring, and aggregating its components.
Augmenting statement specifications means making information implicitly contained or available elsewhere explicitly at the place it is needed. This concerns reestablishing report-wide information, as well as making locally entailed information accessible. An example for the former is the number of diagrams copied into the introductory statement to these diagrams. This treatment is much simpler than using a reference generation algorithm, but it relies on knowing the number of diagrams in advance. An example for the latter is the unit in which the value of a measurement is expressed.
Restructuring information imposes some manipulations on the specifications obtained so far to rearrange the pieces of information contained so that they meet the definition of the IR. The associated operations include reifying an attribute as a structured value and raising an embedded partial description. These operations are realized by mapping schemata similar to those elaborated for linguistically motivated lexicalization \[Horacek (1996)\]. However, some of our schemata are purely application-oriented and tailored to the domain, which manifests itself in the larger size of the structures covered.
Aggregation, the last part of information recasting, comprises removing partial descriptions or adding simple structures. These operations are driven by a small set of declaratively represented rules that access a discourse memory. Most of the rules aim at avoiding repetitions of optional constituents (e.g., temporal and locative information) over adjacent statements. For example, the TIME specification is elided in the second sentence of our sample text, since the time specification in the first sentence still applies. An example for adding a simple structure to an IR expression is the insertion of a marker indicating a strong correspondence between adjacent assertions, which gives rise to inserting “either” in the sample text. Altogether, the underlying rules are formulated to meet application particularities, such as impacts of certain combinations of a value, a status, and a threshold comparison outcome, rather than to capture linguistic principles.
### 3.4 Text realization with TG/2
TG/2 is a flexible and reusable application-oriented text realization system that can be smoothly combined with deep generation processes. It integrates canned text, templates, and context-free rules into a single production-rule formalism and is thus extremely well suited for coping with IR subexpressions of different granularity.
TG/2 is based on production system techniques \[Davis and King (1977)\] that preserve the modularity of processing and linguistic knowledge. Productions are applied through the familiar three-step processing cycle: (i) identify the applicable rules, (ii) select a rule on the basis of some conflict resolution mechanism, and (iii) apply that rule. Productions are used to encode grammar rules in the language TGL \[Busemann (1996)\]. A rule is applicable if its preconditions are met. The TGL rule in Figure 2 is applicable to input material as shown in Figure 1, because the COOP slot matches, and there is information about the THRESHOLD-VALUE available (otherwise a different sentence pattern, and hence a different rule, would be required).
TGL rules contain categories as in a context-free grammar, which are used for rule selection (see below). The rule’s actions are carried out in a top-down, depth-first and left-to-right manner. They include the activation of other rules (:RULE, :OPTRULE), the execution of a function, or the return of an ASCII string as a (partial) result. When selecting other rules by virtue of a category, the relevant portion of the input structure for which a candidate rule must pass its associated tests must be identified. The function get-param in Figure 2 yields the substructure of the current input depicted by the argument. The first action selects all rules with category PPtime; the relevant substructure is the TIME slot of an IR.
TGL rules are defined according to the IR syntax definitions. This includes optional IR elements, many of which can simply be omitted without disturbing fluency. In these cases, optional rules (OPTRULE) are defined in TGL. Optional actions are ignored if the input structure does not contain relevant information. In certain cases, the omission of an IR element would suggest a different sentence structure, which is accounted for by defining alternative TGL rules with appropriate tests for the presence of some IR element. Agreement relations are encoded into TGL by virtue of a PATR style feature percolation mechanism \[Shieber et al. (1983)\]. The rules can be annotated by equations that either assert equality of a feature’s value at two or more constituents, or introduce a feature value at a constituent. The constraint in Figure 2 requires the categories THTYPE and EXCEEDS to agree in gender, thus implementing a subject-participle agreement relation in French. This general mechanism provides a considerable amount of flexibility and goes beyond simple template filling techniques.
A TGL rule is successfully applied if all actions are carried out. The rule returns the concatenation of the substrings produced by the “template” actions. If an action fails, backtracking can be invoked flexibly and efficiently using memoization techniques (see \[Busemann (1996)\]).
## 4 Costs and Benefits
As Reiter and Mellish note, the use of shallow techniques needs to be justified through a cost-benefit analysis \[Reiter and Mellish (1993)\]. We specify the range of possible applications our approach is useful for, exemplified by the report generator developed for the TEMSIS project.
This application took an effort of about eight person months, part of which were spent implementing interfaces to the TEMSIS server and to the database, and for making ourselves acquainted with details of the domain. The remaining time was spent on (1) the elicitation of user requirements and the definition of a small text corpus, (2) the design of IR according to the domain distinctions required for the corpus texts, and (3) text organization, adaptation of TG/2 and grammar development.
The grammars comprise 105 rules for the German and 122 for the French version. There are about twenty test predicates and IR access functions, most of which are needed for both languages. The French version was designed on the basis of the German one and took little more than a week to implement. The system covers a total of 384 different report structures that differ in at least one linguistic aspect.
### 4.1 Benefits
Altogether, the development effort was very low. We believe that reusing an in-depth surface generator for this task would not have scored better. Our method has a number of advantages:
(1) Partial reusability. Despite its domain-dependence, parts of the system are reusable. The TG/2 interpreter has been adopted without modifications. Moreover, a sub-grammar for time expressions in the domain of appointment scheduling was reused with only minor extensions.
(2) Modeling flexibility. Realization techniques of different granularity (canned text, templates, context-free grammars) allow the grammar writer to model general, linguistic knowledge as well as more specific task and domain-oriented wordings.
(3) Processing speed. Shallow processing is fast. In our system, the average generation time of less than a second can almost be neglected (the overall run-time is longer due to database access).
(4) Multi-lingual extensions. Additional languages can be included with little effort because the IR is neutral towards particular languages.
(5) Variations in wording. Alternative formulations are easily integrated by defining conflicting rules in TGL. These are ordered according to a set of criteria that cause the system to prefer certain formulations to others (cf. \[Busemann (1996)\]). Grammar rules leading to preferred formulations are selected first from a conflict set of concurring rules. The preference mechanisms will be used in a future version to tailor the texts for administrative and public uses.
### 4.2 Costs
As argued above, the orientation towards the application task and domain yields some important benefits. On the other hand, there are limitations in reusability and flexibility:
(1) IR cannot be reused for other applications. The consequences for the modules interfaced by IR, the text organizer and the text realizer, are a loss in generality. Since both modules keep a generic interpreter apart from partly domain-specific knowledge, the effort of transporting the components to new applications is, however, restricted to modifying the knowledge sources.
(2) By associating canned text with domain acts, TG/2 behaves in a domain and task specific way. This keeps the flexibility in the wording, which can only partly be influenced by the text organizer, inherently lower than with in-depth approaches.
### 4.3 When does it pay off?
We take it for granted that the TEMSIS generation application stands for a class of comparable tasks that can be characterized as follows. The generated texts are information-conveying reports in a technical domain. The sublanguage allows for a rather straight-forward mapping onto IR expressions, and IR expressions can be realized in a context-independent way. For these kinds of applications, our methods provide sufficient flexibility by omitting unnecessary or known information from both the schemes and its IR expressions, and by including particles to increase coherency. The reports could be generated in multiple languages. We recommend the opportunistic use of shallow techniques for this type of application.
Our approach is not suitable for tasks involving deliberate sentence planning, the careful choice of lexemes, or a sophisticated distribution of information onto linguistic units. Such tasks would not be compatible with the loose coupling of our components via IR. In addition, they would require complex tests to be formulated in TGL rules, rendering the grammar rather obscure. Finally, if the intended coverage of content is to be kept extensible or is not known precisely enough at an early phase of development, the eventual redesign of the intermediate structure and associated mapping rules for text organization may severely limit the usefulness of our approach.
## 5 Conclusion
We have suggested shallow approaches to NL generation that are suited for small applications requiring limited linguistic resources. While these approaches ignore many theoretical insights gained through years of NLG research and instead revive old techniques once criticized for their lack of flexibility, they nevertheless allow for the quick development of running systems. By integrating techniques of different granularity into one formalism, we have shown that lack of flexibility is not an inherent property of shallow approaches. Within the air quality report generation in TEMSIS, a non-trivial application was described. We also gave a qualitative evaluation of the domain characteristics to be met for our approach to work successfully. Further experience will show whether shallow techniques transpose to more complex tasks.
We consider it a scientific challenge to combine shallow and in-depth approaches to analysis and generation in such a way that more theoretically motivated research finds its way into real applications.
|
no-problem/9812/cond-mat9812040.html
|
ar5iv
|
text
|
# VORTEX LATTICE MELTING AND THE DAMPING OF THE DHVA OSCILLATIONS IN THE MIXED STATE
## Abstract
Phase fluctuations in the superconducting order parameter, which are responsible for the melting of the Abrikosov vortex lattice below the mean field $`H_{c2}`$ , are shown to dramatically enhance the scattering of quasi-particles by the fluctuating pair potential , thus leading to enhanced damping of the dHvA oscillations in the liquid mixed state. This effect is shown to quantitatively account for the detailed field dependence of the dHvA amplitude observed recently in the mixed state of a Quasi 2D organic SC.
PACS numbers: 74.70.Kn, 74.60.–w, 74.40.+k
Much effort has been recently invested in elucidating the fundamental mechanisms in which the Abrikosov vortex lattice is involved in the damping process of the de-Haas van-Alphen (dHvA) oscillations in the mixed state of pure, type-II superconductors. Several theoretical attempts have been made to account quantitatively for this effect (see discussions in Refs. , ). It seems , however, that non of the proposed theories is capable of fully achieving this goal, in particular, since there has been growing evidence recently to the significant sensitivity of the dHvA amplitude to extrinsic factors such as vortex lattice disorder and flux lines pinning, . Furthermore, it has been shown theoretically that any mechanism destroying the phase coherence in the vortex lattice, such as disorder or vortex lines fluctuations, should dramatically enhance the inhomogeneous Landau level broadening of quasi particles.
The effect of vortex lattice melting is of special interest since it is associated with phase fluctuations of the superconducting order parameter, which can be implemented selfconsistently into the Gorkov-Ginzburg-Landau (GGL) formalism used in ,.
In a type-II superconductor under a magnetic field a soft shear Goldstone mode, which can be described by long wavelength phase fluctuations is connected to the Abrikosov lattice melting. Unfortunately rigorous analytical approaches to this problem have encountered fundamental difficulties: large order high temperature perturbation expansion with Borel-Pade approximants to the low temperature behavior , has no indication of an ordered vortex lattice even at zero temperature. The existing non-perturbative approaches have not completely clarified the situation: Renormalization group studies have predicted no crystal vortex state in a pure 2D superconductor(SC) at finite temperature, while the novel functional integral formalism suggested in Ref. ,, has led to some kind of a vortex liquid freezing transition without breaking the $`U(1)`$ symmetry. Several Monte Carlo simulations, have recently shown <sup>-</sup> that even in a 2D SC a true vortex lattice melting phase transition takes place at finite temperature and that the transition is of the first order.
In the present letter we make the first attempt to take into account the effect of vortex lattice melting on the magnetization (dHvA) oscillations in the mixed state. A simple, non-perturbative analytical model of melting is developed for this purpose within the GGL theory for a pure 2D extremely type II SC. Our model is based on the observation that in the vortex lattice state the main correction to the mean field free energy arises from fluctuating ”Bragg chains”, that is, from fluctuations which preserve long range periodic order along some crystallographic direction in the vortex lattice. This simplification reduces our 2D fluctuation problem to a 1D one, which is then solved exactly.
Our starting point is the microscopic BCS Hamiltonian for electrons interacting via an effective two-body attractive potential. We then write down the formal functional integral expression for the partition function of this system, eliminating the electronic field by introducing bosonic Hubbard-Stratonovich complex field $`\mathrm{\Delta }\left(\stackrel{}{r}\right)`$ , which describes all possible configurations of Cooper-pairs. By expanding the resulting log determinant in the small $`\mathrm{\Delta }\left(\stackrel{}{r}\right)`$, up to fourth order, we then recover a rather general expression for the partition function in terms of the fully nonlocal Gorkov free energy functional $`F_G\left[\{\mathrm{\Delta }\left(\stackrel{}{r}\right),\mathrm{\Delta }^{}\left(\stackrel{}{r}\right)\}\right]`$ , near $`H_{c2}`$:
$$Z=D\mathrm{\Delta }\left(\stackrel{}{r}\right)D\mathrm{\Delta }^{}\left(\stackrel{}{r}\right)\mathrm{exp}\left[F_G/k_BT\right]$$
(1)
The stationary phase approximation for this functional integral yields the mean field (GGL) equation for $`\mathrm{\Delta }\left(\stackrel{}{r}\right)`$. In the lowest Landau level approximation this equation can be solved by expanding in a set of $`\sqrt{N}`$ Landau functions ($`N`$ being the total number of vortices) $`\varphi _n(x,y)=\mathrm{exp}\left[iq_nx(y+q_n/2)^2\right]`$ , with $`q_n=\frac{2\pi }{a_x}n`$ a wave number along the $`x`$-axis, which is also proportional to the $`y`$-axis position of a given $`x`$ -axis chain. Thus in the symmetric gauge
$$\mathrm{\Delta }(x,y)=e^{ixy}\underset{n=\sqrt{N}/2}{\overset{\sqrt{N}/21}{}}a_n\varphi _n(x,y)$$
(2)
where for an arbitrary $`2D`$ periodic lattice the coefficients $`a_n=\frac{\left(2\pi \right)^{1/4}}{a_x^{1/2}}\mathrm{\Delta }_0\mathrm{exp}\left(i\gamma n^2\right)`$, . All spatial variables are expressed in units of the magnetic length. The parameter $`\left|\mathrm{\Delta }_0\right|^2`$ stands for the spatial average of the order parameter square, and $`a_x`$ is the period along the $`x`$-axis. The summation in Eq.2 is performed over $`\sqrt{N}`$ wavenumbers $`q_n`$ only. The reduction of the number of degrees of freedom results from the assumed periodicity along the $`x`$ -axis. Obviously $`a_x`$ is not unique and depends on the choice of the coordinate system; any $`2D`$ lattice can be described by two different values of $`a_x`$, corresponding to the principal diagonals of the elementary cell (Fig.1).
In Refs., it was shown that the free energy functional near stationary solutions can be approximated by a local expression. This important property , resulting from gross cancellations of nonlocal terms in the free energy due to destructive interference between phase factors of the order parameter , simplifies the problem considerably , reducing it to the Ginsburg-Landau model with the well known local free energy functional
$$F_{GL}=d^2r\left[\alpha \left|\mathrm{\Delta }\left(\stackrel{}{r}\right)\right|^2+\frac{1}{2}\beta \left|\mathrm{\Delta }\left(\stackrel{}{r}\right)\right|^4\right]$$
(3)
Here $`\alpha `$ and $`\beta `$ are known function of temperature and magnetic field (see below). It should be stressed, however, that the locality of the free energy functional can be very sensitive to phase fluctuations, especially in the vortex liquid state , since these fluctuations tend to destroy the phase coherence of the order parameter. For the sake of simplicity we ignore in this paper possible nonlocality in the Boltzman factor of Eq.(1). These assumption, though inconsistent with the nonlocal form of the free energy in the vortex liquid state, can be shown by a more careful analysis to have no significant influence on the final result.
It is convenient to express the GL free energy through the Abrikosov parameter $`\beta _a`$: $`F_{GL}=N\epsilon _0/\beta _a`$, where $`\epsilon _0=\pi \alpha ^2/2\beta `$. In the lattice state $`\beta _a`$ is a function of $`a_x`$ and $`\gamma `$. The dependence of $`\beta _a`$ on $`z\pi ^2/a_x^2`$ at $`\gamma =\pi /2`$ is plotted in Fig.(2a). Both minima with $`z_1=\pi /2\sqrt{3}`$ and $`z_2=\sqrt{3}\pi /2`$ correspond to the triangular lattice with $`\beta _a1,1596`$. The maximum at $`z_3=\pi /2`$ is obtained for the square lattice.
In discussing the effect of fluctuations we first note that the order parameter in Eq.2 with arbitrary coefficients $`a_n`$ includes only those fluctuations which preserve periodic order along the $`x`$-direction. The number of independent coefficients $`a_n`$ is $`\sqrt{N}`$ whereas the total number of orbital centers is $`N`$. Therefore each $`a_n=\left|a_n\right|e^{i\phi _n}`$ describes a set of $`\sqrt{N}`$ orbital centers periodically arranged within a certain chain along the $`x`$-axis (”vortex Bragg chain”). The phase $`\phi _n`$ determines the relative position $`x_n=`$ $`\phi _n/q_n`$ of the $`n`$-th chain.
The scale of the melting temperature is determined by the value of the phase dependent terms in Eq.3. They can be readily calculated if we note that in the GL Hamiltonian there is a small parameter $`\lambda =\mathrm{exp}\left(z\right)`$, with $`z1`$ in the important regions near the minima of $`\beta _a`$ (see Fig.(2a)). Thus the quartic term can be written as an expansion in $`\lambda `$, i.e. $`\underset{n,s,t}{}\lambda ^{s^2+t^2}a_{n+s+t}^{}a_n^{}a_{n+s}a_{n+t}`$, while the leading phase dependent term is of the order $`\lambda ^2`$. Integrating over amplitude fluctuation $`\left|a_n\right|`$ by using the stationary phase approximation we find that the free energy can be written in the mean field like form $`F_{GL}N\epsilon _0/\beta _a`$,where $`\beta _a=\sqrt{\frac{z}{\pi }}\left[1+4\lambda 4\lambda ^2\mathrm{cos}\left(\chi _n\right)_{phase}\right]`$, $`\chi _n`$ $`=\phi _{n+1}+\phi _{n1}2\phi _n`$ , and $`\mathrm{}_{phase}`$ means average over phase fluctuations. Note that $`\mu =\mathrm{cos}\left(\chi _n\right)_{phase}=\frac{I_1\left(T_{cm}/T\right)}{I_0\left(T_{cm}/T\right)}`$, where $`\mu `$ is the shear modulus of the Abrikosov lattice normalized by its mean field value. Here $`I_k\left(x\right)`$ is a modified Bessel function of the $`k`$-order, and $`T_{cm}=\frac{4\lambda ^2\epsilon _0}{\left(1+4\lambda \right)\beta _m}`$ is a characteristic crossover temperature from the crystal to the liquid state, with $`\beta _m=\sqrt{\frac{z}{\pi }}\left(1+4\lambda \right)`$.Well below $`T_{cm}`$, $`\mu 1`$, so that thermal fluctuations do not distort significantly the triangular lattice and the Abrikosov parameter is very close to its minimal (mean field) value,$`\beta _A`$, i.e. $`\beta _a\beta _A\sqrt{\frac{z}{\pi }}\left(1+4\lambda 4\lambda ^2\right)`$. At $`TT_{cm}`$ , on the other hand , $`\mu 0`$ and the vortex system transforms to a new (”liquid”) state for which $`\beta _a\beta _m`$. The crossover temperature $`T_{cm}`$ and the corresponding free energy depend on the choice of vortex chains (Fig.1). At zero temperature (i.e. in the vortex crystal state) the two minima at points $`z_1`$ and $`z_2`$ have the same energy (Fig.(2a)). At a finite temperature this degeneracy is removed by the phase dependent terms, which introduce different interchain coupling within different families of Bragg chains (corresponding to $`z_1`$ and $`z_2`$). Since $`z_1<z_2`$, we have $`T_{cm}\left(z_1\right)>T_{cm}\left(z_2\right)`$. In Fig.(2b) we plot local minimal values of the average free energy $`F_{GL}`$ as functions of the parameter $`t=\sqrt{\frac{4\epsilon _0}{T}}`$, which has been used in Ref.. It is clear that at low temperatures the first state ( $`z_1`$ ) is more stable than the second one ( $`z_2`$ ). Since $`T_{cm}\left(z_1\right)>T_{cm}\left(z_2\right)`$ , the free energy of the first state increases faster with increasing temperature than the second one, and so there is an intersection point $`T_m`$ at which energies of both states are equal but the corresponding entropies are a little different. Therefore we conclude that at this point there is a weak first order transition characterized by a small jump of the lattice entropy. The position of the crossing point $`t_m16`$ and the jump of entropy $`S=T\frac{F}{T}`$ at $`T=T_m`$: $`T\mathrm{\Delta }S/F_{MF}7.5\times 10^3`$ , are in good agreement with the Monte-Carlo simulations .
At higher temperatures $`\beta _a`$ does not describe correctly the free energy. A straightforward calculation in this case yields for the partition function per unit vortex: $`Z_v\sqrt{\frac{2}{\beta }}e^{\beta _mx^2}erf\left(x\right)\mathrm{exp}\left[\lambda \left(\frac{\mathrm{ln}erf\left(x\right)}{x}\right)^2\right]`$ ,where $`x=sign\left(\alpha \right)\frac{\sqrt{\epsilon _0}}{\beta _m\sqrt{T}}`$. In contrast to mean field theory , in which there is a second order phase transition,in our case there is only a crossover to the normal metal state, with significant fluctuations of the superfluid density appearing at magnetic fields far above the mean field $`H_{c2}`$.
All current theories of the dHvA effect in the mixed state are restricted to the mean field approximation. The situation in the vortex liquid state is in some sense similar to that described by Stephen in his derivation of Maki’s damping formula. However,in this approach a random distribution of vortex lines was assumed, while the (nonzero) mean field value of the Abrikosov lattice order parameter was used in the calculation. The averaging over all realizations of the random vortex lattice in such a model yields for the quartic term of the free energy: $`a_{q+s+t}^{}a_q^{}a_{q+s}a_{q+t}\left(\delta _{s,0}+\delta _{t,0}\right)`$. A similar expression, which reflects the complete destruction of phase coherence in the vortex lattice, can be naturally and consistently derived in the model discussed here at $`TT_m`$ by averaging over phase fluctuations, while at $`TT_m`$ it is approximately valid.
Thus, similarly to the MS theory, our fluctuation theory predicts an exponential damping: $`M_{sc}=M_n\mathrm{exp}\left[\frac{\pi ^{3/2}}{n_F^{1/2}}\left|\stackrel{~}{\mathrm{\Delta }}_0\right|^2\right]`$ , where $`M_{sc}`$, $`M_n`$ are amplitudes of the magnetization oscillations in the superconducting and normal states respectively, $`\stackrel{~}{\mathrm{\Delta }}_0=\mathrm{\Delta }_0/\mathrm{}\omega _c`$, and $`n_F=E_F/\mathrm{}\omega _c`$. A simple analytic expression for the damping parameter can be derived in the liquid state well above the melting point by integrating over amplitude fluctuations ( after neglecting small linear corrections in $`\lambda `$):
$$\left|\mathrm{\Delta }_0\right|^2=\frac{1}{\pi }\frac{\mathrm{ln}Z_v}{\alpha }\frac{\alpha }{\beta }\left(1+\frac{\mathrm{ln}erf\left(x\right)}{2xx}\right)$$
(4)
This expression has no singularity at the mean field transition,where both $`\alpha 0`$ and $`x0`$ , and smoothly interpolates between the high field ($`x\mathrm{}`$) value, $`T/\pi \left|\alpha \right|`$ , and the low field ($`x\mathrm{}`$) mean field value, $`\alpha /\beta `$. We have compared this damping with the experimental results on the organic quasi 2D SC: $`\left(ET\right)_2Cu\left(SCN\right)_2`$. In this computation the coefficients $`\alpha `$ and $`\beta `$ derived previously in Refs., for the GGL expansion have been used, namely $`\alpha =.5\pi \mathrm{}\omega _c\mathrm{ln}\sqrt{H_{c2}/H}`$, and $`\beta =1.38\pi \mathrm{}\omega _c/n_F`$ , together with the well documented Fermi surface and SC parameters of the studied material. The only adjustable parameter in this scheme is the mean field $`H_{c2}`$. We have found that a single value of about $`4.7T`$ (see Fig.(3)) fits well the two different sets of available experimental data, which were taken at quite different temperatures , .
It should be stressed that in the fluctuation theory the magnetic oscillations are smoothly damped well above $`H_{c2}`$ , and disappear below the mean field $`H_{c2}`$ in remarkable quantitative agreement with experiment. This contrasts the result of the mean field MS theory , where the additional damping begins abruptly at $`H_{c2}`$ and then increases below $`H_{c2}`$ with significantly stronger rate than what is observed experimentally,
The interesting point in our derivation of the four-particle correlation function concerns the fact that its factorization in the high temperature (or high field) limit into a product of two-particle correlation functions is solely due to integration over phase fluctuations, which leads to the limiting behavior:$`a_{q+s+t}^{}a_q^{}a_{q+s}a_{q+t}_{phase}`$
$`\mathrm{exp}i\left(\phi _{q+s}+\phi _{q+t}\phi _q\phi _{q+s+t}\right)_{phase}\left(\delta _{s,0}+\delta _{t,0}\right)`$ .
This is a generalized form of the condition for the vanishing of the shear modulus $`\mu `$ in the liquid state. The apparent correlation between the vanishing of the shear modulus of the Abrikosov lattice and the factorization of the average phase factor of the quartic term in the SC free energy is a general result, independent of the nature of the melting transition. It is thus conceivable that at a first order melting transition, the vanishing of $`\mu `$ at some finite value of the parameter $`T/T_m`$ , as predicted by the Monte-Carlo simulation reported in ,will be accompanied by a transition from ’coherent’ (i.e. relatively weak) damping in the crystal state to ’incoherent’ MS-like damping in the liquid state at a certain magnetic field below $`H_{c2}`$.
Such a transition will be difficult to observe in the quasi $`2D`$ organic SC investigated due to the very small value of the melting temperature $`T_m`$ and to the strength of the $`2D`$ fluctuations. In $`3D`$ superconductors where the role of fluctuations is less important and the melting transition is shifted significantly closer to $`H_{c2}`$ , one could expect such a transition to be observable. It is thus interesting to note that in the torque dHvA measurement performed on the borocarbide SC $`YNi_2B_2C`$ , where vortex lines pinning seemed to be weak (no peak effect near $`H_{c2\text{ }}`$) , the Dingle plot exhibited a rather sharp upward turn of its (negative) slope at a certain field below $`H_{c2}`$ , which could indicate a freezing transition into an ordered vortex lattice state .
In conclusion, we have developed a simple analytical model of the vortex lattice melting in a $`2D`$ type-II SC, which is in good quantitative agreement with the state of the art numerical simulations in this field. We have shown that fluctuations in the SC order parameter, which are responsible for this melting, destroy the phase coherence in the quasi particle scattering and thus lead to enhanced damping of the dHvA oscillations in the mixed state, in remarkable quantitative agreement with the experiment.
Acknowledgment: We would like to thank Stephen Hayden and Mike Springford for stimulating discussions of the experimental aspects of this work. This research was supported by a grant from the US-Israel BSF, grant no. 94-00243, and by the fund from the promotion of research at the Technion.
Figure Captions
Fig.(1): Two different choices of the vortex chains in the triangular lattice: dash lines - $`z_1=\pi /2\sqrt{3}`$, solid lines - $`z_2=\sqrt{3}\pi /2`$.
Fig.(2): (a) Mean field Abrikosov parameter,$`\beta _a`$, as a function of $`z=(\pi /a_x)^2`$. (b) Dependence of the local minimal values of the free energy $`F/F_{MF}`$,on the parameter $`t=2\sqrt{\epsilon _0/T}`$.
Fig.(3): The magnetization damping factor $`R_s`$ $`=M_{sc}/M_n`$, as a function of magnetic field $`H`$ in the quasi $`2D`$ SC:$`\left(ET\right)_2Cu\left(SCN\right)_2`$. Solid lines are our theoretical curves for $`T=20mK,`$and for the dHvA frequency $`F=690T`$ (1), and $`T=120mK,F=600T`$ (2);stars and circles are the corresponding experimental data taken from Refs., , respectively. The other parameters used in the calculations are:$`T_c=10.4K`$ (from which $`\mathrm{\Delta }_0`$ is calculated using BCS weak coupling formula), and $`m^{}=3.5m_e`$.
|
no-problem/9812/astro-ph9812332.html
|
ar5iv
|
text
|
# GRB Redshift Distribution is Consistent with GRB Origin in Evolved Galactic Nuclei
## Introduction
In our previous work we2 , we have considered the generation of GRBs due to the radiative collisions of CSRs (for simplicity, taken to be solely NSs) in the central stellar clusters of evolved galactic nuclei. Depending on its initial radius and mass, the characteristic dynamical evolution time of a cluster, $`t_e`$, can either exceed the age of the Universe, $`t_0`$, (we call such clusters as ‘slowly evolving clusters’ hereinafter) or be less than $`t_0`$ (‘fast evolving clusters’). The GRB rate from slowly evolving clusters does not depend on redshift of the host galactic nuclei. Following an approach developed in qs87 , the mean rate of radiative collisions of NSs in the fiducial galactic nucleus with mass of a central cluster $`M`$, radius $`R`$, and stellar velocity dispersion $`v=(GM/2R)^{1/2}`$ is given by we2 :
$$\dot{N}_c9\sqrt{2}\left(\frac{v}{c}\right)^{17/7}\frac{c}{R}5.810^6\left(\frac{M}{10^7\mathrm{M}_{}}\right)^{17/14}\left(\frac{R}{0.1\mathrm{pc}}\right)^{31/14}\frac{\text{events}}{\text{yr galaxy}},$$
(1)
where we use for normalization the likely values of $`M`$ and $`R`$ to fit the inferred rate of GRBs. The GRB rate from a unit comoving volume is $`\dot{n}_0=\dot{N}_cn_g`$, where the number density of galaxies $`n_g10^2`$ Mpc<sup>-3</sup> and $`\mathrm{}`$ is an averaging throughout all types of slowly evolving galactic nuclei.
A galactic nucleus with the gravitationally dominating SMBH of mass $`M_h>M`$ might belong to slowly evolving galactic nuclei as well. The corresponding GRB rate from such a nucleus is given by we2
$$\dot{N}_{c,h}\frac{9}{2\sqrt{2}}\left(\frac{Nm}{M_h}\right)^2\left(\frac{v}{c}\right)^{17/7}\frac{c}{R},$$
(2)
which differs from that given by Eq. (1) by a factor $`\left(M/M_h\right)^21`$.
One more possible contribution to GRB rate from a slowly evolving system is due to the coalescence of tight NS binaries in the galactic discs (which is a commonly believed scheme of cosmological GRB origin). In this paper, we assume that there was no strong evolution with cosmological time of the number of tight NS binaries both in the galactic disks and in the slowly evolving nuclei so that all these contributions result in the redshift-independent GRB rate $`\dot{n}_010^8`$ yrs<sup>-1</sup> Mpc<sup>-3</sup>.
Meanwhile the actual parameters of galactic nuclei vary in a wide range and some central clusters may, in fact, be fast evolving ($`t_e<t_0`$). The GRB rate from the fast evolving clusters depends on redshift. Below, we find the redshift distribution of the GRB rate from both slow and fast evolving clusters by (i) modelling the dynamical evolution of a fast evolving cluster in the galactic nucleus, which leads to the production of a SMBH; (ii) connecting the parameters of a fast evolving cluster with the mass of its central SMBH, and (iii) using the observed distributions in luminosity and redshift for quasars, as well as in mass for SMBHs in the galactic nuclei.
## Fast Evolving Galactic Nuclei
We introduce a dimensionless time variable $`yt_0/t_e=(1+z_e)^{3/2}`$, where $`t_e`$ is the time necessary for a SMBH formation, $`z_e`$ is the redshift at that instant. Evidently, $`y>1`$ for clusters collapsed to the present epoch $`t=t_0`$ (fast evolving clusters) and $`y<1`$ for clusters which do not collapse by $`t_0`$ (slowly evolving clusters). As we show elsewhere deo99 , the dynamical evolution of a stellar cluster, in the framework of a homologous Fokker-Plank approximation qs87 , leads to the following relationship between the initial mass of the cluster and the mass of the central SMBH formed there:
$$M_f=210^6y^{2.22}\left[\frac{M_i}{2.810^8M_{}}\right]^{5.44}M_{}.$$
(3)
There are serious observational indications 9712076 , 0006 , 0003 that a fraction of galaxies with a central SMBH might be as high as $`\epsilon 0.1`$ and that there is a correlation between the central SMBH mass, $`M_h`$, and the luminosity of the host galaxy bulge, $`L_s`$, which in a simplified form can be expressed as $`M_h10^{2.5}(L_s/L_{})M_{}`$. This correlation, taken together with the Schechter luminosity function of galaxies sheh , gives the mass distribution of SMBHs, $`\varphi _1(M_h)dM_h`$ with $`\varphi _1(M_h)\epsilon `$. Then we use the observed boyle91 distribution of quasars in absolute magnitude $`M_B`$ and redshift $`z`$, $`\varphi _2(M_B,z)dM_Bdz`$, for $`z3`$. Finally, we assume the existence of an ‘Eddington luminosity phase’ in the evolution of quasars, when they shine with the Eddington luminosity (or with its fraction $`\lambda 1`$) during a certain finite time interval, which duration only depends upon $`\lambda `$ and the initial SMBH mass. In our calculations, the ‘final’ (i.e., by the end of the dynamical evolution) mass of a galactic nucleus is the initial mass of its new-born SMBH. The latter then experiences the ‘Eddington luminosity phase’, during which its mass grows up to the SMBH mass in the actually observed quasar. As a result, we find the mass distribution of active SMBHs, $`P(M_f,y)dM_fdy`$, in terms of the variables $`y`$ and mass $`M_f`$ at the instant of the SMBH formation (in the limit $`M_f<310^6M_{}`$). Thus, these active SMBHs arise due to the dynamical evolution of galactic nuclei with fast evolving central clusters.
## Distribution of GRB Rate in Redshift
In fast evolving galactic nuclei, the resulting rate of GRB generation per unit of comoving volume is given as a function of redshift $`z`$ by
$$\dot{n}_{ev}(z)=𝑑M_f\underset{1}{\overset{x(z)^3}{}}P(M_f,y)\dot{N}_{ci}(M_i(M_f),y)\frac{dyx^{3\xi }}{(x^3y)^\xi },$$
(4)
where $`x=(1+z)^{1/2}`$, $`\xi 0.6`$, $`\dot{N}_{ci}`$ is found from Eq. (1) with $`M=M_i`$ and $`R=R_i`$, and function $`M_i(M_f)`$ is the inverse function given by Eq. (3). Numerical integration over the variable $`M_f`$ in Eq. (4) proceeds in the range $`M_f10^5310^6M_{}`$, which approximately correspond to the range of $`M_i10^7210^8M_{}`$.
The total rate of GRBs from all kinds of galactic nuclei, $`\dot{n}(z)`$, includes also the contribution $`\dot{n}_010^8`$ Mpc<sup>-3</sup> yr<sup>-1</sup> from all slowly evolving galactic nuclei and disks with $`t_e>t_0`$, i.e. $`\dot{n}(z)=\dot{n}_0+\dot{n}_{ev}(z)`$. The total GRB rate as a function of $`z`$ can be fitted by a one-parametric function: $`\dot{n}(z)=\dot{n}_0(1+z)^\beta `$, where the range of $`\beta `$, $`1.5<\beta <2`$, accounts for the statistical uncertainties horack . This function is shown in Fig. 1. The consistency between the calculated and inferred GRB rates is achieved at $`\epsilon /\lambda 0.16`$. Since quasars shine at $`L(0.11)L_{\mathrm{Ed}}`$ (e.g., wandel ), it implies that $`\epsilon 0.0160.16`$, which is consistent with available data -.
|
no-problem/9812/astro-ph9812314.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The proper motion studies of stars in the Galactic Center (Eckart et al. 1995, Genzel et al. 1996, 1997, Ghez et al. 1998) have shown the astonishing accuracy of astrometric observations in the near infrared $`K`$ band. Their studies have proved the existence of a very compact dark mass, most likely a black hole of $`2.6\times 10^6\mathrm{M}_{}`$. The closest investigated star is at the projected distance $`100\mathrm{mas}`$, which corresponds to $`850\mathrm{AU}`$. The gravitational radius for the quoted mass, $`GM/c^2=0.025\mathrm{AU}`$ and corresponds to $`3\mu \mathrm{as}`$. Thus the motion at $`3\times 10^4`$ gravitational radii from the mass is already observed.
Jaroszyński & Paczyński (1998, hereafter JP98) have investigated the possibility of finding orbital parameters for stars near the Galactic Center. According to their work the systematic observations of the closest star included in the proper motion studies can be sufficient to define its orbit in $`10\mathrm{y}`$ with the accuracy allowing for the determination of the central mass with an error smaller then the present estimates. Increased accuracy of astrometric measurements and ability to observe fainter stars which may be found closer to the center, will allow for still more accurate determinations in shorter time.
The speckle interferometry with the Keck telescope (Ghez et al. 1998) reaches the resolution $`50\mathrm{mas}`$ and the accuracy of $`2\mathrm{mas}`$ in position measurements. The Keck Interferometer, which is now under construction (van Belle & Vasisht 1998) will have $`5\mathrm{mas}`$ resolution and $`20\mu \mathrm{as}`$ ($`0.17\mathrm{AU}`$) astrometric accuracy. It will be able to measure the position of a faint point object with $`K22^m`$ if a bright $`K14^m`$ star can be found within a circle of the radius $`20\mathrm{arcsec}`$ around it. Such measurement accuracy, up to few gravitational radii, suggests the possibility of investigating relativistic corrections to the motion of stars sufficiently close to the central mass. We include the periastron motion, the precession of the orbit, the gravitational radiation from the star-central mass binary, and the bending of rays in our study.
In this paper we assume that there is a $`2.6\times 10^6\mathrm{M}_{}`$ black hole in the Galactic Center and we investigate the measureability of the relativistic effects in its vicinity. We consider orbits of stars of sizes $`50`$ to $`1000\mathrm{AU}`$, which correspond to periods $`0.219`$ to $`19.6\mathrm{y}`$. Following JP98 we use Monte Carlo method to simulate the observations of the stars. We use minimization algorithms to fit the model trajectories to the synthetic data sets. Using models with different level of sophistication and comparing the results we are able to find which relativistic effects must be included in the interpretation of observations and which are below the current and near future detection limits.
In the next Section we describe the motion of particles and photons in the weak gravitational field of a rotating body. In Section 3 we present the methods of simulating the observations and procedures of parameter fitting. We also present the main results of this paper - the range of orbit sizes for which the study of relativistic effects is possible with given astrometric accuracy. We also calculate the accuracy of the distance estimate based on the measured periastron motion. In Section 4 we estimate the chance that a star (stars) bright enough to be followed by the Keck Interferometer may be found close enough to the Galactic Center, so the relativistic effects are measurable. The conclusions follow in the last Section.
## 2 Equations of motion in the weak field approximation
We describe the gravitational field far from a rotating black hole in the isotropic coordinates $`t`$, $`x`$, $`y`$, and $`z`$. We use the geometrical units for mass ($`M`$) and angular momentum ($`J`$) of the hole:
$$m\frac{GM}{c^2}j\frac{J}{Mc}$$
(1)
where $`G`$ is the gravity constant and $`c`$ is the velocity of light. The black hole angular momentum is directed along $`𝐞_𝐳𝐤`$. The observer is in the $`(𝐞_𝐱,𝐞_𝐳)`$ plane. In the approximation preserving terms of the order $`r^2`$, the lowest including effects of the black hole angular momentum, the metric takes the form (Landau & Lifshitz 1973):
$$ds^2=\left(12\frac{m}{r}+2\frac{m^2}{r^2}\right)dt^2+4\frac{jmy}{r^3}dtdx4\frac{jmx}{r^3}dtdy$$
$$+\left(1+2\frac{m}{r}+\frac{3}{2}\frac{m^2}{r^2}\right)\left(dx^2+dy^2+dz^2\right)$$
(2)
### 2.1 Orbits of Stars
In the weak field approximation the orbit of the star can be described as a classical ellipse of semimajor axis $`a`$ and eccentricity $`e`$ in a non-inertial frame of reference with coordinates $`(x_1,y_1,z_1)`$. We assume that the black hole is at the origin of this coordinate system, the orbit is at the $`(𝐞_{𝐱_\mathrm{𝟏}},𝐞_{𝐲_\mathrm{𝟏}})`$ plane, and $`𝐞_{𝐱_\mathrm{𝟏}}`$ points toward the orbit periastron. The orientation of the frame is such that the orbital angular momentum is directed along the positive direction of $`𝐞_{𝐳_\mathrm{𝟏}}𝐧`$. The inclination $`i`$ of the orbital plane relative to the equatorial plane is given by the condition $`\mathrm{cos}i=(\mathrm{𝐤𝐧})`$. The ascending node of the orbit is at the position angle $`\mathrm{\Omega }`$ measured from the $`x`$ axis. The periastron is at the angle $`\omega `$ from the ascending node. Finally the position of the star on the orbit, as measured by the eccentric anomaly $`u`$, is given by the Kepler equation:
$$\frac{2\pi }{P}(tt_0)=ue\mathrm{sin}ux_1=a(\mathrm{cos}ue)y_1=a\sqrt{1e^2}\mathrm{sin}u$$
(3)
where $`P`$ is the orbital period and $`t_0`$ is the time of the passage through the periastron.
The Lense-Thirring effect causes the precession of the orbital plane and the change of the periastron location. As a result the line of nodes rotates with the angular velocity:
$$\dot{\mathrm{\Omega }}_{\mathrm{prec}}=\frac{2jmc}{a^3(1e^2)^{3/2}}\dot{\mathrm{\Omega }}=\dot{\mathrm{\Omega }}_{\mathrm{prec}}\mathrm{sin}i$$
(4)
and the periastron position changes with the rate:
$$\dot{\omega }_{\mathrm{prec}}=3\dot{\mathrm{\Omega }}_{\mathrm{prec}}\mathrm{cos}i$$
(5)
Independently of the black hole angular momentum, the periastron of the orbit advances with the angular velocity:
$$\dot{\omega }_{\mathrm{per}}=\frac{3m^{3/2}c}{a^{5/2}(1e^2)}$$
(6)
The resulting evolution of the orbit orientation is given as:
$$\mathrm{\Omega }=\mathrm{\Omega }_0+\dot{\mathrm{\Omega }}(tt_0)\omega =\omega _0+(\dot{\omega }_{\mathrm{per}}3\dot{\mathrm{\Omega }}_{\mathrm{prec}}\mathrm{cos}i)(tt_0)$$
(7)
where $`\mathrm{\Omega }_0`$ and $`\omega _0`$ are the initial values of the angles. The star location at any reference frame can be calculated after a straightforward coordinate transformation based on the equation:
$$𝐫(t)=x_1(t)𝐞_{𝐱_\mathrm{𝟏}}(t)+y_1(t)𝐞_{𝐲_\mathrm{𝟏}}(t)$$
(8)
Suppose an observer is located in the $`(x,z)`$ plane at the angle $`\mathrm{\Theta }`$ measured from the hole rotation axis. The observed position of the star on the sky is given as:
$$Y(t_{\mathrm{obs}})=y(t)Z(t_{\mathrm{obs}})=z(t)\mathrm{sin}\mathrm{\Theta }x(t)\mathrm{cos}\mathrm{\Theta }$$
(9)
where we have chosen coordinates in the sky with $`Z`$ axis along the projection of the black hole angular momentum and $`Y`$ axis along the $`y`$ axis of our frame. The position angle $`\psi `$, which gives the orientation of the $`Z`$ axis on the sky is another parameter of the problem. The time of the observation $`t_{\mathrm{obs}}`$ depends on the source position and is given (up to an additive constant) as:
$$ct_{\mathrm{obs}}=ctx(t)\mathrm{sin}\mathrm{\Theta }z(t)\mathrm{cos}\mathrm{\Theta }$$
(10)
### 2.2 The Influence of Gravitational Lensing
In the typical lensing situation the source and the observer are far from the lens, but the rays pass through its close vicinity. Such case of lensing near the Galactic Center has been considered by Wardle & Yusef-Zadech (1992), Jaroszyński (1998), and Alexander & Sternberg (1998). In such a case one can describe a ray as two segments of a straight line deflected in the lens plane. In our case the distance of the source from the lens is of the same order as the encounter parameter for rays and the standard gravitational lensing formalism can not be employed. Instead we consider the null geodesics in the metric of Eq. 2.
In the zeroth order approximation a ray is a straight line. We assume that the point of the closest approach to the hole is at $`𝐛_\mathrm{𝟎}`$. Suppose the ray is propagating along the unit vector $`𝐥`$ and let $`l`$ be the length measured along the ray from the point of the closest approach. Since we are using the weak field approximation, we have to limit ourselves to the case $`b_0m`$. For a unit energy photon propagating along the ray the components of the four momentum are: $`p_t=1`$, $`p^t=1`$, $`p_l=1`$, $`p^l=1`$, and all other vanish.
In the first order approximation we include the terms proportional to $`m/r`$ in the metric, which preserves its spherical symmetry. Due to this symmetry the ray remains in the plane $`(𝐛_\mathrm{𝟎},𝐥)`$ and is deflected toward the hole:
$$\frac{d}{dl}p_b=\frac{1}{2}g_{tt,b}p^tp^t+\frac{1}{2}g_{ll,b}p^lp^l=\frac{2mb_0}{r^3}$$
(11)
where $`r=\sqrt{b_0^2+l^2}`$. Integrating the above equation with the boundary condition at the observer position: $`p_b=0`$ for $`l=\mathrm{}`$, one gets
$$p_b(l)=\frac{2m}{b_0}\left(1\frac{l}{\sqrt{b_0^2+l^2}}\right)$$
(12)
For sources far behind the lens the photon momentum changes its direction by an angle $`(p_b(+\mathrm{})p_b(\mathrm{}))/p_l=4m/b_0`$, which is a standard result for a ray deflection. The photon momentum perpendicular to the ray is a first order quantity, so $`p^b=p_b`$. Integrating again and assuming $`b(+\mathrm{})=b`$ we have:
$$\frac{d}{dl}b(l)=p^bb(l)=b\frac{2m}{b_0}\left(\sqrt{b_0^2+l^2}l\right)$$
(13)
where the second relation is the analog of the lens equation. For $`l=0`$ one has $`b_0=b2m`$ which is the relation between the observed source position ($`b`$) and the encounter parameter ($`b_0`$). In general the source is at the distance $`b_s`$ from the optical axis, at the position $`l_s`$ along the ray. The value of the encounter parameter $`b_0`$ can be obtained as a solution to the equation $`b(l_s)=b_s`$. In a typical case, when $`b_sm`$, the same is true of $`b`$ and $`b_0`$ and one has approximately:
$$b=b_s+\frac{2m}{b_s}\left(\sqrt{b_s^2+l^2}l\right)$$
(14)
The formula is valid for $`|l|b_s^2/m`$. In the language of gravitational lensing it means, that the source should be at a distance much larger than Einstein radius from the optical axis.
The second order terms ($`m^2/r^2`$) in the diagonal metric components introduce only quantitative corrections to the lens equation. The off diagonal terms introduce the dependence of geodesics on the black hole angular momentum. The deflection of rays is not necessarily toward the hole. Since the metric components of interest are of the second order, the influence of the hole angular momentum can be calculated along the zeroth order rays. Only the deflection perpendicular to the line of sight is interesting. In the observer’s coordinates the geodesic equations are:
$$\frac{d}{dl}p_Y=\frac{2jm\mathrm{sin}\mathrm{\Theta }}{r^3}\frac{6jmY_0^2\mathrm{sin}\mathrm{\Theta }}{r^5}$$
(15)
$$\frac{d}{dl}p_Z=\frac{6jmY_0Z_0\mathrm{sin}\mathrm{\Theta }}{r^5}$$
(16)
where the encounter vector is given as $`𝐛_\mathrm{𝟎}=(Y_0,Z_0)`$ and the distance from the hole as $`r=\sqrt{b_0^2+l^2}`$. The integration gives:
$$\delta Y=\frac{2jm\mathrm{sin}\mathrm{\Theta }}{b_0^4}(Z_0^2Y_0^2)\left(\sqrt{b_0^2+l^2}l\right)+\frac{2jm\mathrm{sin}\mathrm{\Theta }Y_0^2}{b_0^2\sqrt{b_0^2+l^2}}$$
(17)
$$\delta Z=\frac{4jmY_0Z_0\mathrm{sin}\mathrm{\Theta }}{b_0^4}\left(\sqrt{b_0^2+l^2}l\right)+\frac{2jmY_0Z_0\mathrm{sin}\mathrm{\Theta }}{b_0^2\sqrt{b_0^2+l^2}}$$
(18)
where $`\delta Y`$, $`\delta Z`$ denote the angular momentum induced shift of the ray from a trajectory neglecting these effects. Even for the maximally rotating black hole ($`j=m`$), the shifts are of the order $`mm/b_0m`$ and we neglect them in further calculations.
### 2.3 The Gravitational Radiation
The gravitational radiation lowers the energy of a binary system of masses leading to the orbit narrowing and shortening of the orbital period. For two masses $`M_1`$ and $`M_2`$ at a distance $`r`$ from each other, moving on circular orbits, one has (Landau & Lifshitz 1973):
$$\dot{r}=\frac{64G^3M_1M_2(M_1+M_2)}{5c^5r^3}$$
(19)
We estimate the relative change of the orbit size for a star with mass $`M_2=2.6M_{}`$ revolving around a black hole of the mass $`M_1=2.6\times 10^6M_{}`$. We get
$$\frac{|\dot{r}P|}{r}10^{13}\left(\frac{100\mathrm{A}\mathrm{U}}{r}\right)^{2.5}$$
(20)
which means that the gravitational radiation can be neglected for a wide range of star masses and orbit sizes.
## 3 Observability of the Relativistic Effects
We simulate the astrometric observations of stars on elliptic orbits around the Galaxy Center black hole. We check for which range of orbital periods the measurement of the star proper motion with given positional accuracy is sufficient to measure the rate of the periastron motion. We also check, whether the precession of the orbital plane, which may be present if the black hole is rotating, can cause any measurable effects.
The simulations consist of two parts. First we choose the physical parameters of the orbits ($`a`$, $`e`$, $`P`$), their orientation in space ($`i`$, $`\mathrm{\Omega }_0`$, $`\omega _0`$) and time ($`t_0`$). The value of the semimajor axis is the main parameter of this study and it covers a range of values. We use the linear measure of the orbit semimajor axis $`a`$, while the quantity more directly related to observations is the corresponding angle $`\alpha `$ ($`a\alpha R`$, where $`R`$ is the distance to the Galactic Centre, for which we use $`R=8.5\mathrm{kpc}`$). The orbital period is related to the orbit size, since we assume that the black hole has the mass $`M=2.6\times 10^6M_{}`$ (Ghez et al. 1998). Other parameters are chosen at random for each orbit. We assume that any orientation of the orbit in space and any initial phase of the orbital motion have equal probabilities. The initial time of the measurement seems to be unimportant, but due to the fact that the conditions for astronomical ground observations of Sgr A do change through the year, we treat $`t_0`$ as another random variable. For simplicity we assume, that the observations are possible from March through September or from the 10-th to the 40-th week of the year. We do not introduce any seasonal or random dependence of the accuracy of the astrometric measurements on time. The rate of the periastron motion $`\dot{\omega }`$ and the precession angular velocity $`\dot{\mathrm{\Omega }}`$ can be calculated, when other orbital parameters are known.
The equations of Sec. 2 can be used to obtain the “true” trajectory of the star projected into the plane of the sky $`𝐗(t;t_0,a,e,p,i,\mathrm{\Omega }_0,\omega _0,\dot{\mathrm{\Omega }},\dot{\omega },\mathrm{\Theta },\psi )`$, or $`𝐗(t)`$ in short notation. Observations of the star in the instants of time $`\{t_j\}`$ give the measured values of its positions in the sky $`\{𝐗_𝐣\}`$. We simulate the process of observation assuming that
$$𝐗_𝐣=𝐗(t_j)+\delta 𝐗_𝐣$$
(21)
where $`\delta 𝐗_𝐣`$ are the errors introduced by observations. We assume that each component of the position vector is measured with errors which are normally distributed with dispersion $`\sigma `$ and vanishing mean value. Equivalently the distribution of $`\delta 𝐗_𝐣`$ is given as:
$$P\{|\delta 𝐗_𝐣|>s\}=\mathrm{exp}\left(\frac{s^2}{2\sigma ^2}\right)$$
(22)
and the direction of $`\delta 𝐗_𝐣`$ has a uniform distribution. Using Monte Carlo method we obtain synthetic data sets $`\{𝐗_𝐣\}`$ representing the sequences of astrometric observations with noise.
The next step is to fit a model to each synthetic data set. We fit orbital parameters to the observations by minimizing the expression:
$$\chi ^2=\underset{j=1}{\overset{N}{}}\frac{(𝐗_𝐣𝐗(t_j;t_0,a,e,P,i,\mathrm{\Omega }_0,\omega _0,\dot{\mathrm{\Omega }},\dot{\omega },\mathrm{\Theta },\psi ))^2}{\sigma ^2}$$
(23)
The parameters fitted to a synthetic data set are different from the original “true” parameters of the orbit. Many simulations of synthetic data sets for the same orbit give the scatter in estimated parameters. This “bootstrap” method (Press et al. 1988) is a practical way to estimate the accuracy of parameter fitting in the case of real observations.
The synthetic data sets we obtain using Monte Carlo method are always based on calculations including the effects of the periastron motion, the precession of the orbit and the bending of rays by gravitational lensing. To check whether these effects are observable we compare the quality of the fits obtained with models including or neglecting them. We start from the simplest model which neglects both the precession and periastron motion of the orbit. Since we are limiting ourselves to the first order effects in gravitational lensing, the angular momentum of the black hole has no influence on the visual orbit and the reference frame defined by the hole rotation axis looses its observational basis. The inclination of the orbit should now be defined relative to the plane of the sky. The position of the line of nodes can be measured from the declination circle. We obtain the fits by minimizing $`\chi ^2`$ again, but using a simplified model $`\widehat{𝐗}(t;t_0,a,e,P,i,\mathrm{\Omega }_0,\omega _0)`$ instead of the full model depending on the higher number of parameters. This can also be achieved by fixing the values of some parameters ($`\dot{\mathrm{\Omega }}0`$, $`\dot{\omega }0`$, $`\mathrm{\Theta }0`$, and $`\psi 0`$) in the full model. As we show in Fig. 1 the model neglecting the periastron motion and the precession of the orbit can be rejected for sufficient accuracy of position measurements and short orbits.
The precession of the orbit is a weaker effect than the periastron motion. If the observations span only a few orbital periods, which is the case of our simulations, the two effects can be difficult to distinguish. To clarify this point we try a model which includes periastron motion but neglects the precession, $`\widehat{𝐗}(t;t_0,a,e,P,i,\mathrm{\Omega }_0,\omega _0,\dot{\omega })`$. Our calculations show, that in this case the successful fits are possible. By successful we mean the fits with sufficiently low value of $`\chi ^2`$:
$$\chi ^2\chi _{0.95}^2(m)$$
(24)
where $`m=2N8`$ is the number of the degrees of freedom for $`N`$ observed positions and $`8`$ parameters fitted. The subscript denotes the confidence level. (See also the dotted line on Fig. 1.)
Using the fitted orbital parameters one can calculate the mass of the central body with the help of the Third Kepler Law:
$$m=\frac{4\pi ^2a^3}{c^2P^2}$$
(25)
Assuming that the central body is not rotating, one has the expected periastron motion per one revolution:
$$\mathrm{\Delta }\omega _{\mathrm{exp}}=\dot{\omega }_{\mathrm{per}}P=\frac{6\pi m}{a(1e^2)}$$
(26)
We introduce another variables, independent of orbit eccentricity, which also measure the periastron motion:
$$Q_{\mathrm{exp}}=\mathrm{\Delta }\omega _{\mathrm{exp}}(1e^2)Q_{\mathrm{fit}}=\dot{\omega }_{\mathrm{fit}}P_{\mathrm{fit}}(1e^2)$$
(27)
The first variable is based on the theory and fitted values of mass and semimajor axis of the orbit. The second is based on the fitted values of the periastron motion and the orbital period. For the nonrotating central body the variables should be equal. If the central body rotates, they should be different. Since both variables are based on the fitting procedure and ”observations” with errors we can only compare their averaged values. We define:
$$D=\frac{|Q_{\mathrm{fit}}Q_{\mathrm{exp}}|}{Q_{\mathrm{exp}}}$$
(28)
$$S=\frac{\sqrt{\left(Q_{\mathrm{exp}}Q_{\mathrm{exp}}\right)^2}}{Q_{\mathrm{exp}}}$$
(29)
where the averages are taken for orbits of the same “true” semimajor axis $`a_0`$ and the same accuracy of astrometric observations $`\sigma `$. In Fig. 2 we compare the fitted and expected rates of periastron motion. We consider the case of stars moving in the gravitational field of a nonrotating black hole (left panel) and the case of maximally rotating hole (right panel). The left panel shows that the difference between the fitted and expected values scales linearly with the error in astrometric measurements. The same is true of the dispersion in the expected rate of periastron motion. This shows, that the differences are statistical in nature. The plots also show that the “observations” which cover a given number of rotational periods (5-10 in most cases) allow for a more accurate fit to the orbit semimajor axis and period than for a fit to the rate of the periastron motion. The right panel shows a nonlinear behavior. Comparing Eqs. 4, 5, and 6 one can get a rough estimate:
$$\frac{\mathrm{\Delta }\omega _{\mathrm{prec}}}{\mathrm{\Delta }\omega _{\mathrm{per}}}\sqrt{\frac{j^2}{ma}}=\sqrt{\frac{m}{a}}$$
(30)
where we have assumed $`|\mathrm{cos}i|=0.5`$ and neglected the factor dependent on the eccentricity. The second equality is valid for the maximally rotating hole ($`j=m`$). The presence of the precession in the “true” motion, which is not accounted for in the model causes a systematic difference between the rates of the periastron motion estimated in two ways. This difference remains finite when the error in measurements becomes very small. Its value is in agreement with the above formula for the orbits we consider. (Orbits with $`a=50`$, $`150`$, and $`250\mathrm{AU}`$ or $`P=0.22`$, $`1.14`$, and $`2.45\mathrm{y}`$ are considered here.)
The expected contribution to the periastron motion from the precession is smaller than $`1`$ % for sufficiently wide orbits ($`a250\mathrm{AU}`$), even for the maximally rotating central black hole. If we neglect this effect, Eqs. 25, 26 can be used as two independent methods of estimating the mass. After substitution we get the relation between the fitted variables of our model:
$$\mathrm{\Delta }\omega =\frac{24\pi ^3}{c^2}\frac{a^2}{(1e^2)P^2}$$
(31)
In our approach we generate the synthetic data sets using a fixed distance to the Galactic Center, which makes the linear ($`a`$) and angular ($`\alpha `$) measures of the semimajor axis indistinguishable. In reality the angular size of the orbit $`\alpha `$ is fitted directly, while its physical dimension can be calculated for the known distance to the object. Substituting $`a=\alpha R`$ to the above equation we get:
$$R=\sqrt{\frac{(1e^2)\mathrm{\Delta }\omega }{24\pi ^3}}\frac{cP}{\alpha }$$
(32)
where all the variables in the RHS can be obtained from a fit to the observations. Thus we have a method of estimating the distance to the Galactic Center independent of any “standard candles”.
We have estimated the error in the distance found by the above method for a limited number of simulations. We have assumed that the black hole is maximally rotating, so the errors resulting from the precession (which is unaccounted for in the fitting procedure) are contained in our analysis. The most important contribution to the error comes from the measurement of the periastron motion despite the square root dependence of the distance on this variable. The distances calculated for orbits of given size are scattered. We find the median value of fitted distances $`R_{\mathrm{med}}`$ and such $`\mathrm{\Delta }R`$, that 68% of the results belongs to $`[R_{\mathrm{med}}\mathrm{\Delta }R,R_{\mathrm{med}}+\mathrm{\Delta }R]`$. The plots of the relative error in the distance measurement are given in Fig. 3. Only the position measurements with errors smaller than $`100\mu \mathrm{as}`$ can give the distance estimate with the accuracy better than $`10\%`$.
## 4 Stars Very Close to the Galactic Center
The presence of stars at distances of few hundreds astronomical units from the Galactic Center is necessary to measure the effects we consider. A more detailed consideration of this subject can be found in JP98. The observations of Eckart & Genzel (1997), Genzel et al. (1996, 1997) and Ghez et al. (1998) show the presence of the dense central star cluster. The subset of stars with measured proper motions, which is seen close to the center in projection, is also close to it in 3D, since the proper velocities of stars are related to the projected distance - a fact hard to understand for background or foreground objects. These stars are relatively bright ($`K17^m`$), but their sample is not complete (Ghez et al. 1998). This makes the following reasoning a bit risky. Assuming that the stars with measured proper motions are otherwise typical we can postulate that their luminosity function is the same as that of the central star cluster. The integral luminosity function for the Galactic Center in $`K`$ can be obtained from papers of Blum et al. (1996) and Holtzman et al. (1998). It has a shape $`N(L_K)L_K^\beta `$, with $`\beta =0.875`$, and flattens at $`K21^m`$. Thus going from $`K=17^m`$ to $`K=21^M`$ makes the star volume density $`25`$ times higher and the typical distances between the stars in 3D become $`3`$ times smaller. Since the closest to the center observed star is at the distance $`850\mathrm{AU}`$, one can expect few fainter stars to be even closer. Thus the presence of observable stars at required distance from the center is likely.
## 5 Conclusions
We have investigated the observability of relativistic effects in motion of stars in the vicinity of the Galactic Center assuming the presence of massive black hole there and the accuracy of astrometric position measurements up to $`20\mu \mathrm{as}`$. We have shown that the gravitational radiation from the star-black hole binary and the second order effects in the deflection of rays related to the angular momentum of the black hole are completely negligible. The only robust effect is the motion of the orbit periastron. Systematic observations of a star orbit, covering $`25\mathrm{y}`$ or $`5`$ orbital periods (whichever takes shorter) are sufficient to measure the rate of the periastron motion. With the accuracy of position measurements $`0.1\mathrm{mas}`$ or better, it is possible to reject models neglecting the periastron motion for orbits of semimajor axis up to $`500\mathrm{AU}`$. With the highest accuracy expected for the Keck Interferometer ($`20\mu \mathrm{as}`$) it will be possible to measure the periastron motion for orbits up to $`a10^3\mathrm{AU}`$.
If the black hole in the Galactic Center is rotating it should cause the precession of the orbits of stars. We investigate this effect for orbits of the size $`a=50\mathrm{AU}`$ to $`1000\mathrm{AU}`$. The effect is weak. The models of orbits neglecting the precession can not be rejected on the basis of the $`\chi ^2`$ value. This is probably due to the fact that precession of the orbit, when observed through a small number of revolution periods is hard to distinguish from the periastron motion. (Both effects change the direction of the ellipse axes in space. Precession can also change the visual shape of the orbit, since it changes the angle between the line of sight and the orbital plane, but this is a very slow process.) The effects of precession can be seen indirectly as a discrepancy between the theoretical rate of periastron motion for a nonrotating black hole of estimated mass and the rate actually measured.
The measurements of the orbit elements (size, eccentricity, and period) and of the rate of the periastron motion can be used to estimate the distance to the source. This method is independent of any standard candles. The systematic observations of a star moving at distances $`10^3\mathrm{AU}`$ from the Galactic Center with the maximal positional accuracy of the Keck Interferometer would give its distance up to few percent.
Acknowledgements. I thank Bohdan Paczyński for many useful discussions and the anonymous referee for his suggestions. This work was supported in part by the Polish State Committee for Scientific Research grant 2-P03D-012-12.
## REFERENCES
* Alexander, T. & Sternberg, A. 1998, , , astro-ph/9811038.
* Blum, R.D., Sellgren, K., & DePoy, D.L. 1996, ApJ, 470, 864.
* Eckart, A., & Genzel, R. 1997, MNRAS, 284, 576.
* Eckart, A., Genzel, R., Hofmann, B., Sams, B.J., & Tacconi-Garman, L.E. 1995, ApJ, 445, L23.
* Genzel, R., Eckart, A., Ott, T., & Eisenhauer, F. 1997, MNRAS, 291, 219.
* Genzel, R., Thatte, N., Krabbe, A., Kroker H. & Tacconi-Garman, L.E. 1996, ApJ, 472, 153.
* Ghez, A.M., Klein, B.L., Morris, M., & Becklin, E.E. 1998, ApJ, , in press, astro-ph/9807210.
* Holtzman, J.A., Watson, A.M., Baum, W.A., Grillmair, C.J., Groth, E.J., Light, R.M., Lynds, R. & O’Neil, E.J. 1998, AJ, 115, 1946.
* Jaroszyński, M. 1998, Acta Astron., 48, 413.
* Jaroszyński, M. & Paczyński, B. 1998, , , in preparation.
* Landau, L.D. & Lifshitz, E.M. 1973, Teoria Polia, , Nauka (Moscow) .
* Press, W.H., Teukolsky, S.A., Vettering, W.T., & Flannery, B.P. 1988, Numerical Recipes: the Art of Scientiofic Computing (Cambridge University Press, Cambridge), , .
* van Belle, G.T. & Vasisht, G. 1998, The Keck Interferometer: Science Requirements Document, JPL D-15477, , http://huey.jpl.nasa.gov/keck/).
* Wardle, M. & Yusef-Zadeh, F. 1992, ApJ, 387, L65.
## 7 Note Added in Proof
The highest astrometric accuracy of the Keck Interferometer ($`20\mu \mathrm{as}`$) will in fact be limited to relatively bright objects with $`K17.6`$. Thus the measurement of the periastron motion will probably be possible only for stars already discovered. The best candidate seems to be the star S0-1 from the Ghez et al. (1998) catalog. For the fainter stars the accuracy of position measurements will be much worse, $`3\mathrm{mas}`$, not adequate for the following of the periastron motion. The results of our calculations can still be applied to observations done with other instruments providing high astrometric accuracy for faint objects, which may become operational in the future. I am grateful to Dr. Gerard T. van Belle for pointing to me my wrong interpretation of the Keck Interferometer technical data.
|
no-problem/9812/astro-ph9812265.html
|
ar5iv
|
text
|
# Intrinsic Absorption Lines in Seyfert 1 Galaxies. I. Ultraviolet Spectra from the Hubble Space TelescopeBased on observations made with the NASA/ESA Hubble Space Telescope, obtained from the data archive at the Space Telescope Science Institute. STScI is operated by the Association of Universities for Research in Astronomy, Inc. under the NASA contract NAS5-26555.
## 1 Introduction
Seyfert galaxies were originally recognized as a special class of objects from their strong emission lines (Seyfert 1943), and were later subdivided into two groups based on the widths of their emission lines in optical spectra (Khachikian & Weedman 1971). Seyfert 1 galaxies have broad permitted lines with widths $``$ 1000 km s<sup>-1</sup> (FWHM), and narrow permitted and forbidden lines with widths $``$ 500 km s<sup>-1</sup> (FWHM), whereas Seyfert 2 galaxies show only the narrow emission lines. In turn, these objects are recognized as belonging to the category of objects known as active galaxies (Osterbrock 1984). It has been known for some time that Seyfert galaxies are strong emitters over the entire electromagnetic spectrum (Weedman 1977). The optical continua of Seyfert 1 galaxies are dominated by nonstellar emission that can be characterized by a power-law, as first demonstrated by Oke & Sargent (1968), whereas this component appears to be much weaker in Seyfert 2 galaxies (Koski 1978). These trends extend to the lines and continua observed in the ultraviolet (Wu, Boggess, & Gull 1983) by the International Ultraviolet Explorer (IUE).
Seyfert (1943) recognized the presence of absorption lines in several of his objects, which he attributed to a “G-type spectrum”. In fact, most of the absorption features in the optical spectra of both types of Seyfert can be attributed to stellar features from the host galaxy (Koski 1978; Crenshaw & Peterson 1985). Oke & Sargent (1968) first reported a possible nonstellar absorption feature in the spectrum of NGC 4151; they explain that this feature was first discovered and attributed to He I $`\lambda `$3889 self-absorption by O.C. Wilson. Anderson & Kraft (1969) resolved the metastable He I absorption into three components, and discovered hydrogen Balmer (H$`\beta `$, H$`\gamma `$) absorption corresponding to these components. The components are all blue-shifted relative to the emission lines by up to $``$970 km s<sup>-1</sup>, and the authors attributed the blue-shifts to ejection of matter from the nucleus of NGC 4151. Cromwell and Weymann (1970) showed that the strengths of the Balmer absorption lines are variable, and can occasionally disappear altogether. Anderson (1974) found evidence that the Balmer absorption lines in NGC 4151 can vary on time scales as small as $``$30 days, and claimed that the absorption and continuum variations are correlated. Anderson also suggested that the absorption region lies outside of the broad emission-line region, because the depth of the H$`\alpha `$ absorption was greater than the flux in the nearby continuum.
In the ultraviolet, early observations of NGC 4151, primarily by IUE, revealed a rich assortment of intrinsic absorption lines spanning a wide range in ionization from Mg II to N V, as well as fine-structure and metastable absorption lines such as C III $`\lambda `$1175 (Davidsen & Hartig 1978; Boksenberg et al. 1978; Penston et al. 1981; Bromage et al. 1985). Bromage et al. suggested that the absorption-line variations in NGC 4151 are due to changes in the column densities, rather than the velocity spread of the clouds. In addition, they suggested that the absorbing region is likely to be composed of optically thin gas located outside of the broad emission-line region. Ulrich (1988) described other Seyfert galaxies that show obvious intrinsic absorption in IUE spectra. Most of these objects show C IV and N V absorption (along with L$`\alpha `$ in most cases), indicating high ionization. Only two Seyferts showed strong evidence for intrinsic Mg II absorption (NGC 4151 and MCG 8-11-11). The absorption features are blue-shifted by as much as $``$2500 km s<sup>-1</sup> with respect to the emission lines.
From a large set of IUE observations, Ulrich (1988) estimated that only 3 – 10% of all Seyfert 1 galaxies show intrinsic UV absorption. This would suggest that while intrinsic UV absorption is an interesting curiosity, it might not have a strong impact on our understanding of active galaxies. However, only very strong absorption lines (with equivalent widths $`>`$ 1 Å) are detectable at the low spectral resolution ($`\lambda `$/$`\mathrm{\Delta }\lambda `$ $`=`$ 200 – 400) and low signal-to-noise ratio (SNR $`<`$ 10) of the IUE spectra. From an examination of a small sample of HST spectra, we found that more than half (5/8) of the Seyfert 1 galaxies showed intrinsic C IV absorption (Crenshaw 1997). This indication that intrinsic UV absorption is much more common than previously believed was a principal motivation for expanding our sample of HST spectra, as described in this paper.
Another motivation for this study was to investigate the relationship between the UV absorber and the X-ray “warm absorber”, which is characterized by O VII and O VIII absorption edges and is present in about half of the Seyfert 1 galaxies observed in X-rays (Reynolds 1997; George et al. 1998b). Evidence for this highly ionized gas in the line-of-sight to an active galaxy was first reported by Halpern (1984), using the Einstein Observatory. Subsequent X-ray satellites provided supporting evidence for warm absorbers, and detailed analysis and modeling of many individual objects were made possible by ASCA. Warm absorbers are typically characterized by ionization parameters in the range U $`=`$ 0.1 – 10, temperatures on the order of 10<sup>5</sup> K, and ionized hydrogen column densities in the range 10<sup>21</sup> – 10<sup>23</sup> cm<sup>-2</sup> (Reynolds & Fabian 1995; George et al. 1998b).
A connection between the UV and X-ray absorbers was made by Mathur et al. (1994), based on quasi-simultaneous ROSAT and HST observations of the quasar 3C 351. Using photoionization models, these authors found that a single component of ionized gas could produce both the observed strengths of the O VII and O VIII absorption edges and the equivalent widths of the intrinsic UV absorption lines (O VI $`\lambda \lambda `$1031.9, 1037.6; N V $`\lambda \lambda `$1238.8, 1242.8; C IV $`\lambda \lambda `$1548.2, 1550.8). In these models, most of the carbon, nitrogen, and oxygen are in higher ionization states than those represented by the UV lines. These authors claim that the UV and X-ray absorbers are also the same in 3C 212 (Mathur 1994) and the Seyfert 1 galaxy NGC 5548 (Mathur et al. 1995). However, in some objects, it appears that multiple components characterized by a wide range in ionization paramater and hydrogen column density are needed to explain the wide range in ionization species; such is the case for NGC 3516 (Kriss et al. 1996a; Crenshaw et al. 1998) and NGC 4151 (Kriss et al. 1995). To further investigate the proposition that a single component is responsible for both the UV and X-ray absorption, we felt that it was important to study the absorption in other sources, particularly those with both HST and ASCA spectra.
Finally, we were motivated by the fact that the intrinsic UV absorption in Seyfert galaxies is qualitatively similar to the “associated” absorption seen in QSOs, in that the lines are relatively narrow ($``$ 300 km s<sup>-1</sup>), usually close to the emission-line redshift, and in some cases, variable (Hamann 1997; Hamann et al. 1997). The relationship between these relatively narrow absorption systems and the broad absorption lines (BALS), which can exhibit troughs reaching blueshifts as high as $``$0.1c, and are found in approximately 10% of all QSOs (Weymann et al. 1981; Weymann et al. 1991), is not known. The fact that these systems represent high-ionization gas flowing outward from the nucleus suggests that these phenomena may be related. Thus, detailed studies of the intrinsic absorption in Seyfert 1 galaxies should lead to a better understanding of outflows in active galaxies as a function of luminosity. Eventually, we hope that these studies will provide clues to the physical structure of active galaxies, and the relationship between the various absorption, emission, and continuum components in these objects.
To gain a better understanding of the intrinsic absorption in Seyfert 1 galaxies, we have obtained all of the UV spectra of these objects from the HST data archive. In this paper, we present new results on the basic properties of the intrinsic UV absorber, its connection to the X-ray warm absorber, its frequency of occurrence in Seyfert 1 galaxies, its global covering factor (relative to the continuum source and broad emission-line region), and its variability. We give a brief history of previous ultraviolet observations of each object in Appendix A, and discuss a few interesting objects that did not satisfy our selection criteria in Appendix B.
## 2 Observations and Data Reduction
Our original sample was a collection of eight Seyfert 1 galaxies observed under the GHRS GTO (Guaranteed Time Observations) programs of Boggess and Maran; preliminary results on detection of intrinsic absorption were reported in Crenshaw (1997). Results on a few observations from our sample were previously reported for Mrk 509 (Crenshaw, Boggess, & Wu 1993), NGC 3783 (Maran et al. 1996), and NGC 3516 (Crenshaw, Maran, & Mushotzky 1998).
To increase the sample, we searched the HST Data Archive to find all of the objects that were identified as Seyfert galaxies by the original proposers and have ultraviolet spectra obtained by the Faint Object Spectrograph (FOS) or the Goddard High Resolution Spectrograph (GHRS). Our search was conducted one year after the removal of these two instruments from HST (in 1997 February), so all of the data were nonproprietary and available for retrieval. We did not include the few spectra obtained with the Faint Object Camera (FOC), as they do not satisfy the criteria given below. We included only those objects in this group that have been identified as Seyfert 1, Seyfert 1.5, or narrow-line Seyfert 1 galaxies (Osterbrock & Pogge 1985) in the literature or in the NASA/IPAC Extragalactic Database (NED); hereafter, we refer to this group in a general way as Seyfert 1 galaxies. We excluded Seyfert 2 galaxies or those objects identified as Seyfert 1.8 or 1.9 galaxies (cf., Osterbrock 1981). The reason for excluding these objects is a practical one: it is difficult to detect the absorption lines if they are not seen against strong and relatively broad emission lines and/or strong UV continua. We identified $``$40 Seyfert 1 galaxies with UV spectra in the archives.
After retrieving and inspecting the UV spectra, we used the following additional criteria for including objects in our survey:
1) There must be at least one spectral region in which an intrinsic absorption line might be detected. In practice, we included objects that had an observation of the redshifted N V, Si IV, C IV, or Mg II emission-line profiles, since the strongest absorption lines come from these species and are embedded in these profiles (Ulrich 1988). We did not include objects that were only observed in the L$`\alpha `$ region, since this absorption line can originate in the low-redshift L$`\alpha `$ forest (cf., Stocke et al. 1995; Shull et al. 1996).
2) The resolution must be sufficient to detect and, in most cases, separate the absorption doublets. In practice, we required that $`\lambda `$/$`\mathrm{\Delta }\lambda `$ must be $``$ 1000, where $`\lambda `$ is the observed central wavelength and $`\mathrm{\Delta }\lambda `$ is a resolution element. Henceforth, we refer to FOS spectra with $`\lambda `$/$`\mathrm{\Delta }\lambda `$ $``$ 1000 and GHRS spectra with $`\lambda `$/$`\mathrm{\Delta }\lambda `$ $``$ 2000 as the “low resolution spectra”, and GHRS spectra with $`\lambda `$/$`\mathrm{\Delta }\lambda `$ $``$ 20,000 as the “high resolution spectra”.
3) The signal-to-noise ratio (SNR) per resolution element near the expected location of absorption must be $``$ 20 for the low resolution spectra, and $``$ 8 for the high resolution spectra. The detection limit for absorption lines varies according to the SNR; a typical 3$`\sigma `$ detection limit (in equivalent width) is $``$0.1 Å for the low resolution spectra, and $``$0.03 Å for the high resolution spectra.
We found that 17 Seyfert 1 galaxies in the HST data archive satisfied the above criteria. Only a few of the rejected spectra did not satisfy criteria 1 or 2; most of them were rejected on the basis of poor SNR. We emphasize that the observations in our sample were acquired by various investigators for a variety of purposes, and that we did not include QSOs at low redshifts. Thus, our sample is obviously not a complete one. In Appendix B, we list those Seyferts that did not satisfy our criteria but show possible evidence for intrinsic absorption.
Table 1 lists the low and high resolution observations for the 17 Seyfert 1 galaxies in our sample. Note that each object has coverage of the redshifted L$`\alpha `$, N V, Si IV, and C IV lines, whereas 15 objects have coverage of the redshifted Mg II region (NGC 3783 and II Zw 136 do not have coverage of this region). Spectra that were taken within days of each other have been averaged to increase the SNR. There are multiple epochs of observation for a number of these objects, although each observation does not necessarily contain every wavelength interval. High resolution observations are available for five Seyfert 1 galaxies observed at low resolution. For these observations, we list the redshifted emission-line profiles that are present in each wavelength setting, where we expect to find the intrinsic absorption lines.
We used the following data reduction procedures for the spectra in Table 1. For the FOS spectra, we retrieved the calibrated files from the HST data archive to obtain the wavelength, absolute flux, error flag, and photon noise vectors for each spectrum. For the GHRS spectra, we reduced the raw data using the IDL procedures written for the GHRS Instrument Definition Team (Blackwell et al. 1993); in the process, we determined an average background across the diode array and subtracted it from the gross spectrum for each readout, as described in Crenshaw et al. (1998). We interpolated all of the spectra to a linear wavelength scale, retaining the original approximate wavelength intervals (in Å per bin), and, for each object, we averaged all of the spectra obtained at a particular wavelength setting and within a few days of each other. We then combined spectra that were obtained at different wavelength settings and observed within days of each other at a single point in the wavelength region of overlap. In the few instances where the fluxes did not agree to within 5% in the region of overlap, the longer wavelength spectrum was scaled in flux to match the shorter wavelength spectrum. Thus, we have a final UV spectrum for each entry in Table 1.
In each spectrum, we identified lines that arise in the interstellar medium or halo of our Galaxy using the lists of UV resonance lines in Morton, York, & Jenkins (1988). Those lines that could not be identified as Galactic in origin were considered to be candidates for intrinsic absorption. Specifically, absorption lines with heliocentric velocities greater than a few hundred km s<sup>-1</sup>, at the proper wavelength separations for the doublets, and with equivalent widths $``$ 3 times the measurement errors (see below) were considered to be intrinsic. In addition, only L$`\alpha `$ lines that occured at radial velocites close to those of other detected lines were considered to be intrinsic, since those at other velocities could be due to the low-redshift L$`\alpha `$ forest.
Figures 1 and 2 present the far-UV (1200 – 1700 Å) portion of the low resolution spectra. The locations of the strongest Galactic absorption lines (Morton et al. 1988) are also shown. Figure 1 shows the seven Seyfert 1 galaxies that did not show detectable intrinsic absorption. Figure 2 shows the ten Seyfert 1 galaxies with intrinsic L$`\alpha `$, N V $`\lambda \lambda `$1238.8, 1242.8, Si IV $`\lambda \lambda `$1393.8, 1402.9, and/or C IV $`\lambda \lambda `$1548.2, 1550.8 absorption lines . Note that the intrinsic absorption lines span a large range in equivalent width and velocity width, and tend to occur in the cores of the emission-line profiles or are blueshifted with respect to the emission lines. In a few cases, multiple velocity components can be seen, even at this relatively low resolution.
Figures 3 and 4 give the high-resolution spectra of Seyfert 1 galaxies with intrinsic absorption. Figure 3 shows the C IV region in four Seyferts, and demonstrates the necessity of high spectral resolution for resolving absorption components. Figure 4 shows the corresponding L$`\alpha `$ and N V components, when available.
## 3 Direct measurements
We used the following procedures to measure the intrinsic absorption lines that we detected. To determine the shape of the underlying emission, we fit a cubic spline to regions on either side of the absorption. For those cases in which the absorption occurs in the very core of the line, this procedure may underestimate the underlying emission if there is a significant narrow-line contribution that has not been properly taken into account. Based on visual inspection of the unabsorbed profiles in the same spectrum (e.g., C III\] $`\lambda `$1909), we do not expect this to be a serious problem. To normalize the absorption profiles, we divided the observed spectra by the cubic spline fits. We then made direct measurements of the centroid, equivalent width (EW), and full-width at half-maximum (FWHM) of each visually distinct absorption component. Obviously, the number of visible components may depend on spectral resolution, so we consider the low and high resolution spectra separately. We note that our distinction between components and structure within components is somewhat subjective.
We determined the uncertainites in the centroids and FWHMs of each feature from different reasonable fits to the emission on either side of the absorption lines. We determined uncertainties in EWs from the sum in quadrature of the error in fitting the underlying emission and the uncertainty due to photon noise. Finally, we excluded from further consideration any absorption feature with an EW less than three times its uncertainty.
### 3.1 Detection of intrinsic absorption
Table 2 provides information on the redshifts of Seyfert 1 galaxies in our sample; 4 of these 17 objects are generally regarded as narrow-line Seyfert 1 galaxies, as noted. Redshifts from H I observations (when available) and from the narrow optical emission lines (primarily \[O III\] $`\lambda \lambda `$4959, 5007) were obtained from NED. The H I redshifts are probably more indicative of the systemic velocities of the host galaxies, but we have values for only eight galaxies. The optical redshifts tend to be slightly lower, on average, but within $``$100 km s<sup>-1</sup> of the neutral hydrogen redshifts. To be consistent, we adopt the optical redshifts when determining the radial velocities of the absorption lines relative to the Seyfert galaxy, with the understanding that the absorption velocities relative to the systemic redshift are likely to be within 100 km s<sup>-1</sup> of these values. We note that in a few cases, the absorption lines in Figure 2 appear to be slightly redshifted with respect to the cores of the broad emission lines (e.g., in Mrk 509), due to a slight blueshift of the high ionization emission lines with respect to the lower ionization lines (see Gaskell 1982).
In Table 2, we specify whether or not each Seyfert 1 galaxy shows intrinsic UV absorption, based on our criteria and according to the detection limits that we gave in the previous section. The Seyferts with intrinsic C IV absorption always show N V and L$`\alpha `$ absorption at the same approximate velocities, as we discuss in more detail later. We also specify whether or not each Seyfert 1 galaxy shows evidence for an X-ray warm absorber (Reynolds 1997; George et al. 1998b). This table confirms our earlier finding that intrinsic UV absorption is very common: more than half (10/17) of the Seyfert 1 galaxies show intrinsic absorption lines.
### 3.2 Results from the direct measurements
Tables 3 and 4 present direct measurements of the intrinsic absorption lines in the low and high resolution spectra, respectively. We give the observed centroid ($`\lambda _{obs}`$), the equivalent width (EW), our identification, and the radial velocity, relative to the optical redshift, for each absorption feature. Where there is evidence for more than one kinematic component, we number each component beginning with the one at the shortest observed wavelength. In a few instances, an intrinsic absorption component is blended with another Galactic or intrinsic line; in these cases, we identify the contributors and give the centroid for the entire feature, but do not give other direct measurements. For the high resolution spectra in Table 4, we also give the full-width at half-maximum (FWHM) of each component, when possible, since the absorption features are resolved (the instrumental FWHM is $``$15 km s<sup>-1</sup>). Some components are blended enough to prevent an accurate measure of their widths.
Measurement errors for $`\lambda _{obs}`$ are typically 0.05 – 0.1 Å (10 – 20 km s<sup>-1</sup>) for the low resolution spectra and 0.02 Å ($``$4 km s<sup>-1</sup>) for the high resolution spectra; measurement errors for FWHM are typically $``$10 km s<sup>-1</sup>. We provide more realistic uncertainties in v<sub>r</sub> and FWHM below, by comparing values for all of the ions associated with each kinematic component.
We summarize the kinematic components that we have detected in Tables 5 and 6. The radial velocity centroids and FWHM are averages from all of the lines that we think are associated with one kinematic component, and the uncertainties are standard deviations of the averages. As we noted earlier, the number of components is highly dependent on the spectral resolution. For objects that have both high and low resolution spectra, it is clear that the low resolution values are just “averages” for blends of the strongest lines. In Mrk 509, there is an offset in v<sub>r</sub> between the low and high resolution spectra of $``$80 kms<sup>-1</sup>, which is probably due to a zero-point error in the low resolution wavelength scale of $``$0.3 Å. Similar errors in v<sub>r</sub> are possible for the other low resolution spectra.
### 3.3 Conclusions based on direct measurements
We can draw some important conclusions from Tables 2, 5, and 6 and Figures 2 – 4, to support and extend previous results on the intrinsic UV absorption lines (which were based on a few Seyfert 1 galaxies, see section 1):
1. From the low resolution sample, the frequency of occurence of intrinsic absorption is f $`=`$ 10/17 $`=`$ 0.59. When intrinsic absorption is detected, high ionization lines (C IV, N V) and L$`\alpha `$ are always seen. The intermediate ionization line of Si IV is seen in only 4/17 of the Seyfert 1 galaxies. Of the 15 objects with near-UV coverage, only NGC 4151 shows evidence for low-ionization (Mg II) absorption,
2. The centroids of the absorption lines are blueshifted by up to $``$ 2100 km s<sup>-1</sup>, or occasionally at rest, with respect to the narrow emission lines (and relative to neutral hydrogen, when it is detected). Thus, the UV absorbers are, for the most part, undergoing radial outflow with respect to the nuclei of these Seyfert 1 galaxies.
3. At high resolution, the absorption lines often split into distinct kinematic components. The components are resolved, and exhibit a range of widths (20 – 400 km s<sup>-1</sup> FWHM). Most of the components have widths much larger than that expected for thermally stable photoionized gas, even if they arise in the warm absorber (FWHM of C IV $``$ 20 km s<sup>-1</sup> at 10<sup>5</sup> K). This indicates macroscopic motions within a component, such as turbulence or the superposition of additional radial velocity components.
4. Lines that are narrow and within a few hundred km s<sup>-1</sup> of the systemic radial velocity could potentially come from the interstellar medium or halo of the host galaxy. However, most of the intrinsic absorption lines arise close to the nucleus, because they are broad, have large blueshifts, and/or they are variable (see section 5).
5. In the high resolution spectra, the cores of the strongest absorption components are deeper than the heights (in flux) of the continua. This indicates that both the central continuum source and at least a portion of the BLR is occulted by the UV absorber. When we know the size of the BLR (e.g., from reverberation mapping experiments), we can determine lower limits to the size of the absorber in the plane of the sky and its distance from the continuum source (see section 4.3).
6. From Table 3, we can see that there is a one-to-one correspondence between those Seyferts that show UV and X-ray absorption in this sample; six objects show both UV and X-ray absorption, and two objects show neither (although the case for a warm absorber in NGC 7469 is not strong). Thus, although the sample is small, it appears that there is a connection between the UV and X-ray warm absorbers; however, this does not require that they arise from the same zone, characterized by a single ionization parameter.
## 4 Absorption profiles
Additional information can be derived from the resolved absorption lines in the high resolution spectra. We have omitted the low resolution spectra from this analysis for the following reasons. 1) The absorption lines seen at low resolution are often comprised of multiple kinematic components, and these components are likely to be characterized by different physical conditions. 2) It is more difficult to determine the intrinsic shapes of the underlying profiles in the low resolution spectra, since the absorption components are broadened and blended. 3) Many of the absorption lines are saturated in the low resolution spectra, which typically leads to underestimates of the column densities <sup>1</sup><sup>1</sup>1Lower limits to the column densities of the absorption lines in the low resolution spectra can be calculated directly from the EWs in Table 3, by assuming the case of unsaturated lines (Spitzer 1978). Other techniques for estimating the column densities from unresolved or marginally resolved absorption lines include the curve-of-growth method (Spitzer 1978), and the apparent optical depth method (Savage & Sembach 1991). These techniques are based on simplifying assumptions, which are discussed in these references..
### 4.1 Covering factors
The methods we used to analyze the absorption profiles follow those of Hamann et al. (1997). We consider the case where the region responsible for an absorption component does not completely cover the emission source(s) behind it. If this effect is present and not corrected for, the column densities will be underestimated. We define the covering factor, C<sub>los</sub>, to be the fraction of continuum plus BLR that is occulted by the absorber in our line of sight. We will consider the case where the covering factor could differ from one kinematic component to the next (but not within a component, as discussed below). If we normalize the absorption lines by dividing by the underlying emission, and I<sub>r</sub> is the residual flux in the core of the absorption line at a particular radial velocity, then we can determine a lower limit to the covering factor in the line of sight from the residual flux:
$$C_{los}1I_r.$$
(1)
We can also determine a lower limit to the fraction of BLR flux that is occulted by the absorber:
$$C_{los}^{BLR}\frac{1I_cI_r}{1I_c},$$
(2)
where I<sub>c</sub> is the fraction of the underlying emission at a particular radial velocity due to the central continuum source.
The actual value of C<sub>los</sub> for a component can be determined from a doublet, if both lines of the doublet are unblended with other lines (Hamann et al. 1997). If the expected ratio of the optical depths of the doublet is 2 (as for C IV and N V), then
$$C_{los}=\frac{I_1^22I_1+1}{I_22I_1+1},$$
(3)
where I<sub>1</sub> and I<sub>2</sub> are the residual fluxes in the cores of the weaker line (e.g., C IV $`\lambda `$1550.8) and stronger line (e.g., C IV $`\lambda `$1548.2), respectively.
A lower limit to C<sub>los</sub> is easily obtained, particularly if the absorption components are resolved. To determine a well-constrained value of C<sub>los</sub>, however, we need both members of the doublet to be clean (unblended with other lines), reasonably strong lines, high SNR, and an accurate estimate of the underlying emission. In addition, since even the strong components are affected by blending with other components in their wings, we are constrained to determining C<sub>los</sub> in the cores of these components.
In practice, we determined the covering factor for a few (2 – 5) resolution elements across the core of each component, and then calculated the average and uncertainty. Actual values of the covering factors, as opposed to lower limits, were determined only for a few strong, unblended components (see Figures 3 and 4). For the lower limits, we chose those lines that yield the highest values for that component (implicitly assuming that for a given kinematic component, the covering factor is the same for L$`\alpha `$, C IV, and N V).
Table 6, gives C<sub>los</sub> and C$`{}_{}{}^{BLR}{}_{los}{}^{}`$ for each kinematic component in the high resolution spectra, along with the line used to determine these values. We note that when actual values are determined, they are close to unity (except for NGC 3783).
### 4.2 Column densities
Since the absorption components are resolved, we can determine the column density of each component from its optical depth as a function of radial velocity (v<sub>r</sub>). If the covering factor C<sub>los</sub> $`=`$ 1 for a component, then the optical depth at a particular v<sub>r</sub> is given by:
$$\tau =ln\left(\frac{1}{I_r}\right),$$
(4)
If C<sub>los</sub> $``$ 1, then the optical depth at a particular radial velocity is:
$$\tau =ln\left(\frac{C_{los}}{I_r+C_{los}1}\right)$$
(5)
(Hamann et al. 1997). In principal, C<sub>los</sub> can be a function of radial velocity (v<sub>r</sub>), but we have assumed that it is constant for a given component.
The column density is then obtained by integrating the optical depth across the profile:
$$N=\frac{m_ec}{\pi e^2f\lambda }\tau (v_r)𝑑v_r$$
(6)
(Savage & Sembach 1991), where f is the oscillator strength and $`\lambda `$ is the laboratory wavelength (Morton et al. 1988).
In Table 7, we give the column densities for each component in the high resolution spectra, for the case of C<sub>los</sub> $`=`$ 1. The values are averages of those from each member of the doublet, if unblended with other components; otherwise they are from the line that is least affected by blending. Horizontal lines in the table indicate that the component was not detected at that epoch. The column densities in Table 7 are larger than those derived by assuming unsaturated lines by as much as factor of $``$3, due to the substantial optical depths in many of the components (e.g., see Crenshaw et al. 1998).
To illustrate the effects of partial covering on the measurements, Table 7 also gives column densities for the few instances that we could actually determine reliable C<sub>los</sub> values. The greatest effect is seen for NGC 3783, where a covering factor of $``$0.5 leads to an increase in the column density by a factor of $``$2. We also note that if the optical depths are high, even a covering factor close to one can, if not corrected for, have a significant effect on the column densities. For example, C<sub>los</sub> = 0.99 and 0.98 for components 1 and 2 in NGC 3516, respectively, but the effects on the measured column densities are still significant due to the large optical depths. In this particular case, we have suggested that the measured covering factors are not unity because of grating-scattered light, which occurs at the 1 – 2% level (Crenshaw et al. 1998). Grating-scattered light is almost certainly present at this level in all of the GHRS spectra presented here.
We have only a few column densities for the case in which C<sub>los</sub> has been derived. Thus, for the purpose of discussion, we will refer to the column densities derived from C<sub>los</sub> = 1, with the understanding that the true column densities may be underestimated by a factor of $``$2 or less.
### 4.3 Conclusions from the absorption profiles
In addition to the effects that the covering factors have on column density measurements, we can use them to place important constraints on the absorption regions. Table 6 shows that C$`{}_{}{}^{BLR}{}_{los}{}^{}`$ is always positive, indicating that at least a portion of the BLR is occulted by each component. This result alone does not provide a tight constraint, because the absorber could be closer to the continuum source than the BLR, and still intercept a fraction of the emission from the far side of the BLR. However, we note that there is at least one component in each object with C$`{}_{}{}^{BLR}{}_{los}{}^{}`$ $``$ 0.9, except for NGC 3783. Thus, in four of the five Seyfert 1 galaxies in Table 6, there is at least one component of the absorbing gas that essentially covers the BLR (or at least the high-ionization portion characterized by C IV and N V emission). Since the sizes of the C IV emitting regions in these objects are a few light days (Wanders et al. 1997, and references therein), the absorption regions responsible for these components must be greater than a few light days from the continuum source, and greater than a few light days in extent (in the plane of the sky).
We define C<sub>global</sub> to be the fraction of the underlying emission that the intrinsic UV absorber (as a whole) covers, averaged over all lines of sight for all Seyfert 1 galaxies. Thus,
$$C_{global}=<C_{los}>f,$$
(7)
where f is the fraction of Seyfert 1 galaxies that show intrinsic absorption, which we have determined to be 0.59 for our sample. For any Seyfert 1 galaxy, C<sub>los</sub> for the ensemble of absorption components must be greater than or equal to the largest value for any single component. We take the average of the largest values or lower limits of C<sub>los</sub> from each Seyfert, and find $`<`$C<sub>los</sub>$`>`$ $``$ 0.86, which leads to C<sub>global</sub> $``$ 0.51. We note that these values are estimates, and may be affected by the small sample size, as well as biases on the part of the observers who selected these Seyfert galaxies.
Thus, as an ensemble, the regions responsible for the UV absorption typically cover at least half of the sky as seen by the central continuum source. To illustrate this result, we take two extremes: 1) the covering factor could be close to one-half in all Seyfert 1 galaxies, or 2) the covering factor could be one in the Seyferts with observed absorption and the absorbing gas is not present in the other Seyferts. Obviously, the true situation could be somewhere in between. In either case, C<sub>global</sub> provides an important geometric constraint for the intrinsic absorber. For example, geometries such as spherical shells, large pieces of shells or sheets, or cones with large opening angles are favored.
Finally, we discuss the column densities in Table 7. We concentrate on the values determined for C<sub>los</sub> $`=`$ 1 (with the understanding that they may be slight underestimates of the true column densities). There is a substantial range in the column densities of the components, from 0.1 to 14 x 10<sup>14</sup> cm<sup>-2</sup> (the former value is at our detection limit). By comparison, the column density of C IV is much higher in a typical BLR cloud: N(C IV) $``$ 10<sup>18</sup> cm<sup>-2</sup> (Ferland & Mushotzky 1982; Ferland et al. 1992).
As we mentioned previously, it is difficult to accomodate the larger values of the C IV column density ($``$10<sup>15</sup> cm<sup>-2</sup>) in X-ray “warm absorber” models that are characterized by a single ionization parameter. In addition, we note that in the case of NGC 5548, where multiple components of N V absorption have been measured, the ratio of N(N V)/N(C IV) varies substantially, ranging from 2.1 to 18.2. This is strong evidence that the ionization parameter varies significantly from one kinematic component to the next. Combined with the evidence that Si IV absorption is present in some objects and absent in others, and the detection of Mg II absorption and other low ionization lines in only one object (NGC 4151), these results indicate that there is a wide range in physical properties, particularly ionization parameter, among UV absorbers in different Seyferts and among different components in the same Seyfert.
## 5 Variability
To investigate the variability of the intrinsic UV absorption in our sample, we concentrate on the high resolution spectra, since it is difficult to determine accurate column densities in the low resolution data. As in the past, when absorption variations are detected, thay are found in the column densities of the lines and not in the radial velocities. We are limited to the three Seyfert 1 galaxies with multiple epoch observations at high resolution: NGC 3516, NGC 3783, and NGC 4151. (Variability studies of other objects with other satellites are described in Appendix A.)
NGC 3783 shows extreme variability in its C IV absorption lines (Figure 3). There is no evidence for absorption in the GHRS spectrum on 1993 February 5, but a C IV absorption doublet is present 11 months later (on 1994 January 6) at $``$560 km s<sup>-1</sup> (relative to the optical emission). After a subsequent interval of 15 months (on 1995 April 15), an additional C IV doublet is evident at a radial velocity of $``$1420 km s<sup>-1</sup> (there was no significant change in the column density of the other component). We reported the first variation in Maran et al. (1996), and the second in Crenshaw et al. (1997); in Table 7, we give the magnitude of these variations in terms of the derived column densities. We also note that the single spectrum of the N V region on 1993 February 21 (Figure 4) was obtained just 16 days after the first C IV spectrum, which showed no absorption. Since C IV and N V are always present together in our low-resolution spectra, this suggests (but does not prove) that the absorption varied rapidly over this 16 day interval.
As reported in Crenshaw et al. (1998), we detected no variability in the C IV absorption components of NGC 3516 over a 6-month period, although changes in the column density at the $``$25% level cannot be ruled out, due to the fact that the lines are saturated. In this case, we know that previous variations of the C IV absorption were due to a broad, blue-shifted component that disappeared between 1989 October and 1993 February (Koratkar et al. 1996). For NGC 4151, Weymann et al. (1997) found that the C IV absorption was constant over a 4-year period, with the exception of one transient event. We note that over the time span of the HST observations, the lines of C IV and N V in these two Seyferts were saturated, and therefore modest ($``$25%) variations in the column densities of these high-ionization lines would not be detected. In NGC 4151, variations in weaker (and lower ionization) absorption lines have been detected with HUT and ORFEUS, as described in Appendix A.
There are two likely sources of absorption variations: 1) changes in the ionization fractions due to continuum variations, and/or 2) changes in the total column of gas (e.g., due to bulk motion across the line-of-sight). The former could be identified from a correlation between continuum and absorption variations, but obviously we do not have enough observations of NGC 3783, closely spaced in time, to establish such a correlation. The best case for a correlation between continuum and absorption variations has been established for the neutral hydrogen absorption and low-ionization lines of NGC 4151 (Kriss et al. 1996c; Espey et al. 1998).
The best case for bulk motion as a source of variability is the aforementioned disappearance of the broad, blueshifted absorption component of C IV in NGC 3516. This component has not reappeared in recent observations, despite substantial continuum variations, which suggests that the absorbing gas has moved out of the line-of-sight. Assuming tangential velocities comparable to the largest observed radial velocities, an absorber could move across the BLR on a time scale of months to years (Maran et al. 1996). It is possible that both continuum variability and bulk motion are responsible for absorption variability, and that they could very well be occuring on different time scales.
If variations in column density can be attributed to changes in the ionizing continuum, then monitoring observations can be used to probe the physical conditions in the gas. The time lag between continuum and absorption variations will provide the recombination and/or ionization time scales, and thus the electron density (n<sub>e</sub>) and/or radial location (r) of the gas, respectively (Shields & Hamann 1997; Krolik & Kriss 1997). With only one of these quantities (n<sub>e</sub> or r), the other can be determined with the assistance of photoionization models, which provide the ionization parameter and ionic ratios from the observed column densities of different ions. For example, Shields & Hamann (1997) used our first detection of variability in NGC 3783 to derive lower limits to the electron density (n<sub>e</sub> $``$ 300 cm<sup>-3</sup>) and upper limits to the distance of the absorbing gas from the central continuum source (r $``$ 10 pc). Other examples are given in Appendix A, but in every case, only an upper limit has been determined for the time lag between continuum and absorption variations, which leads to a lower limit on the density and an upper limit on the radial distance.
## 6 Summary
We have presented the first systematic study of intrinsic absorption in active galaxies that is based on ultraviolet spectra from HST. Although the sample is not complete and subject to biases on the part of the observers that selected these objects, we find that intrinsic absorption is much more common than previously believed (Ulrich 1988), occuring in more than half (10/17) of the Seyfert 1 galaxies in this study. The absorption is due to highly ionized gas that is flowing outward from the nucleus. At high resolution ($`\lambda `$/$`\mathrm{\Delta }\lambda `$ $``$ 20,000) the absorption often breaks up into multiple kinematic components, which are likely characterized by different physical conditions, and probably different distances from the nucleus.
We have shown that at least some of the absorption components arise from regions that are completely outside of the BLR, and that these regions must be at least a few light days in size (in the plane of the sky) to occult the BLR. As an ensemble, the intrinsic absorbers cover a large part of the sky ($``$ 50%) as seen from the continuum source. We note that these values
The absorption components in our sample remain stable in radial velocity over at least several years, but, in one case, show extreme variations in column density over one year. Previous studies have suffered from a lack of spectral or temporal resolution, but indicate that the UV absorber lies between the BLR and NLR (except for narrow components near the systemic velocity of the host galaxy that may arise from that galaxy’s halo.) Intensive monitoring observations (e.g, with HST) are needed produce well-sampled continuum and absorption light curves, in order to explore the range of physical characteristics of the absorbing gas, determine its origin, and explore its relation to the X-ray warm absorber.
Eight Seyfert 1 galaxies in this study have high-quality HST and ASCA spectra, and there is a one-to-one correspondence between those objects that show intrinsic UV absorption and those that show X-ray warm absorbers. Thus these two phenomena are related, but additional studies are needed to determine the exact nature of their relationship. Specifically, detailed photoionization calculations are needed to match the observed column densities, and thereby test the need for multi-zone models. We will address this issue in a companion paper.
This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. This research has also made use of NASA’s Astrophysics Data System Abstract Service. D.M.C. and S.B.K. acknowledge support from NASA grant NAG 5-4103. Support for this work was provided by NASA through grant number AR-08011.01-96A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555.
## Appendix A Notes on individual Seyfert 1 galaxies in the sample
### A.1 WPVS 007 (WPV85 007)
This relatively unknown narrow-line Seyfert 1 shows variable and ultrasoft X-ray emission (Grupe et al. 1995); no information is available on the possibility of an X-ray warm absorber. In addition to deep L$`\alpha `$, N V, and C IV absorption, we have found intrinsic Si IV absorption at the same approximate radial velocity. The FOS observations are also described in Goodrich et al. (1998).
### A.2 I Zw 1
Weak intrinsic absorption was discovered in the FOS spectra of this narrow-line Seyfert 1 galaxy by Laor et al. (1997). Although there are several observations, there is only one G130H spectrum, which contains the high-ionization absorption lines, and therefore no information is available on absorption variability.
### A.3 NGC 3516
Intrinsic absorption lines of C IV, N V, and Si IV were originally detected in IUE spectra by Ulrich & Boisson (1983). NGC 3516 also exhibits a strong and variable X-ray warm absorber (Kolman et al. 1993; Nandra & Pounds 1994; Kriss et al. 1996b; Mathur, Wilkes, & Aldcroft 1997; George et al. 1998b). Intrinsic O VI $`\lambda \lambda `$ 1031.9, 1037.6 absorption has been detected with the Hopkins Ultraviolet Telescope (HUT) by Kriss et al. (1996a). Kriss et al. have also detected intrinsic neutral hydrogen absorption (in the Lyman series) at a blueshift of $``$400 km s<sup>-1</sup>, relative to the systemic velocity.
IUE monitoring found evidence for variability in the C IV absorption on time scales as small as weeks (Voit, Shull, & Begelman 1987, Walter et al. 1990; Kolman, Halpern, & Martin 1993). Walter et al. characterized the C IV absorption as a blend of a narrow stable component in the core of the emission-line, and a variable and broad blueshifted component. Kolman et al. (1993) found that n<sub>e</sub> $``$ 10<sup>5</sup> cm<sup>-3</sup> and r $``$ 9 pc for this variable component. Koratkar et al. (1996) show that the variable component disappeared between 1989 October and 1993 February, when there were no IUE observations, and has not reappeared.
Additional information on the FOS observations can be found in Goad et al. (1998), and a more detailed description of the GHRS observations is presented in Crenshaw et al. (1998). The variable blueshifted component of C IV has not reappeared in the HST spectra, and there is no evidence for variations in the remaining C IV absorption over the span of HST observations (1995 April - 1996 August).
### A.4 NGC 3783
Intrinsic L$`\alpha `$ and C IV absorption were discovered in the FOS spectrum by Reichert et al. (1994), and intrinsic N V absorption was found by Lu, Savage, & Sembach (1994) in the GHRS spectra of this region. George et al. (1998a) found evidence for variability in the X-ray “warm absorber” between ASCA observations on 1993 December and 1996 July.
We presented the first two GHRS spectra of the C IV region in Maran et al. (1996), and discovered that the C IV absorption is extremely variable. We present the third GHRS spectrum of C IV in the current paper, and find that another C IV component has appeared at a higher blueshift (section 5). The C IV absorption detected in the GHRS spectrum on 1994 January 16 at $``$560 km s<sup>-1</sup> could be the same component as that seen in the FOS spectrum on 1992 July 27 (at $``$ 660 km s<sup>-1</sup>), given the weakness of the feature and the wavelength uncertainties in the FOS spectra. This component completely disappeared on 1993 February 5, but is present in N V just 16 days later, on 1993 February 21, suggesting: 1) rapid variability, or 2) unusually high ionization during this time period (because in every Seyfert with simultaneous observations, C IV and N V absorption are always present together). As we discussed earlier, Shields & Hamann (1997) find n<sub>e</sub> $``$ 300 cm<sup>-3</sup> and r $``$ 10 pc for this component.
### A.5 NGC 4151
A history of the early observations of intrinsic UV absorption is given in the Introduction. Previous observations of the complex X-ray absorption are described by George et al. (1998b). Observations with HUT and ORFEUS have extended the UV coverage down to 912 Å, and have resulted in the detection of a number of new absorption lines, including the strong lines of C III $`\lambda `$977.0, N III $`\lambda `$989.8, O VI $`\lambda \lambda `$1031.9, 1037.6, and the hydrogen Lyman series (Kriss et al. 1992, 1995; Espey et al. 1998). Espey et al. show that the C III $`\lambda `$1175 (metastable) absorption line varies on a time scale of $``$1 day, and thus r $``$ 25 pc, similar to the value obtained from H I absorption variations (Krolik & Kriss 1997). No variations were found in the high-ionization absorption lines (C IV, N V, O VI).
The FOS and GHRS spectra were obtained primarily by Weymann et al. (1997). Due to the large columns, widths, and number of components, the deconvolution of the C IV absorption is much more difficult than for any other object in this sample, even at high resolution. Weymann et al. present their deconvolution and measurement of individual components, and Kriss (1998) presents a slightly different deconvolution; we do not attempt to repeat their analyses here. Weymann et al. find no evidence for variations in the C IV absorption in five separate observations obtained over $``$4 years, except for the appearance of a broad absorption feature in the blue wing of C IV on one occasion. They also present evidence for variations in the S IV absorption in the low resolution spectra.
The only other line observed at high resolution is Mg II (Weymann et al. 1997). The Mg II/C IV ratio varies greatly from one component to the next, indicating a wide range in ionization parameter (Kriss 1998). NGC 4151 is the only object that shows Mg II absorption in our sample, and it would be interesting to test the uniqueness of this object, by looking for low-ionization absorption in other Seyfert 1 galaxies.
### A.6 NGC 5548
The intrinsic C IV absorption was first studied in detail by Shull & Sachs (1993), using data from the 1988 - 1989 IUE monitoring campaign (Clavel et al. 1991). Shull & Sachs claim that C IV EW varies on a time scale of 4 days or less, and is anticorrelated with the continuum flux at 1570 Å, indicating that n<sub>e</sub> $``$ 5 x 10<sup>5</sup> cm <sup>-3</sup> and r $``$ 14 pc. Properties of the X-ray warm absorber are given by George et al. (1998b, and references therein).
The FOS observations that we used are from the 1993 HST monitoring campaign (Korista et al. 1995). Measurements of the absorption in these data are given by Mathur et al. (1995). We found no evidence for strong variations in these lines (given the difficulties imposed by low resolution), and therefore used the average spectrum from this campaign. The GHRS spectra of C IV (at low and high resolution) were obtained by Mathur et al. (1998), and the GHRS spectra of the N V region are also shown in Savage et al. (1997).
### A.7 Mrk 509
York et al. (1984) found two kinematic components of intrinsic L$`\alpha `$ and C IV absorption in high-dispersion IUE spectra. Evidence for an X-ray warm absorber is given by Reynolds (1997) and George et al. (1998b).
Our FOS observations found the same two kinematic components as those identified by York et al. (1984) $``$ 12 years earlier, and we found N V at these velocities as well (Crenshaw et al. 1995). We originally agreed with the suggestion of York et al. (1984) that these components arise from extended regions of ionized gas in the host galaxy. However the high resolution GHRS spectra of L$`\alpha `$ (also shown in Savage et al. 1997) demonstrates that these two components are broad, and are likely to arise close to the nucleus. A determination of the location of the absorbers will require variability monitoring.
### A.8 II Zw 136
We find intrinsic L$`\alpha `$, N V, and C IV absorption at two widely separated radial velocities in the GHRS low resolution spectra.
### A.9 Akn 564
The absorption in this narrow-line Seyfert 1 galaxy is similar to that in WPVS 007, in that Si IV absorption is present in addition to strong L$`\alpha `$, N V, and C IV. The FOS observations are also presented by Goodrich et al. (1998).
### A.10 NGC 7469
The FOS observations are also shown in Kriss et al. (1998). We find intrinsic L$`\alpha `$, N V, and C IV absorption at a large blueshift: $``$1770 km s<sup>-1</sup> relative to the emission lines. George et al. (1998b) found marginal evidence for a warm absorber in ASCA spectra, whereas Reynolds (1997) found no absorption that satisfied his criteria.
## Appendix B Additional Seyfert 1 galaxies with possible intrinsic absorption
Although these Seyferts did not satisfy our selection criteria, they show broad L$`\alpha `$ absorption, indicating possible intrinsic absorption close to the nucleus.
### B.1 III Zw 2
A noisy FOS G130H spectrum obtained on 1992 January 18 shows moderately broad L$`\alpha `$ absorption in the core of the emission profile, and possible intrinsic N V absorption, near the emission-line redshift of z $`=`$ 0.0894.
### B.2 Mrk 279
A high resolution GHRS spectra obtained on 1997 January 16 shows two broad, deep L$`\alpha `$ absorption components, centered at $``$490 km s<sup>-1</sup> and $`+`$60 km s<sup>-1</sup> with respect to the emission-line redshift (z $`=`$ 0.0306).
### B.3 Mrk 290
A high resolution GHRS spectra obtained on 1997 January 9 shows a broad, deep L$`\alpha `$ absorption component, centered at $``$240 km s<sup>-1</sup> relative to the emission-line redshift (z $`=`$ 0.0296).
|
no-problem/9812/hep-ph9812255.html
|
ar5iv
|
text
|
# The intercept of the BFKL pomeron from Forward Jets at HERA
## 1 Introduction
Using a physical gauge, deep inelastic scattering processes can be represented within perturbative QCD with ladder diagrams like the one shown in figure 1. In this picture the interaction between an electron and a proton is mediated through the exchange of a virtual photon of four momentum squared $`q_\mu q^\mu =Q^2`$, which couples to a quark–antiquark box at the top of the parton ladder inside the proton.
The DGLAP and BFKL schemes select different leading logarithmic regions of the phase space to describe the partonic evolution along this ladder. The DGLAP approach takes into account the leading terms in $`\mathrm{ln}Q^2`$, but neglects those in $`\mathrm{ln}(1/x)`$. In this case the square of the transverse momentum of the partons along the ladder are strongly order, $`k_i^2k_{i1}^2`$ (cf. fig. 1), whereas the longitudinal momenta obey $`x_i<x_{i1}`$. On the other hand, in the BFKL formalism the $`\mathrm{ln}(1/x)`$ are resummed in the region of $`k_i^2k_{i1}^2`$ and $`x_ix_{i1}`$. Due to the inherent approximations involved the domain of applicability of the DGLAP equations is medium to high $`x`$–Bjorken, but the BFKL approach should only be used at small $`x`$–Bjorken.
The electron–proton collider HERA opened up the study of DIS to new domains of small values of $`x`$–Bjorken and still sizeable virtuality of the photon. It was thus expected that experiments at HERA would be able to measure the transition from the region of applicability of DGLAP to that of BFKL. Among the different proposed observables to test the BFKL approach to DIS, the measurement of forward jets is considered to be one of the most promising.
Higher order corrections to the BFKL formalism have been recently calculated . They turned out to be sizable and pointed out the need of still higher order corrections. This reduced the quantitative predictive power of the leading logarithmic approximation to BFKL calculations, but the qualitative behavior at small $`x`$ may still remain the same. Taking this into account a fit to HERA data, which of course resumes all orders, may shed light into the problem of the numerical stability of the BFKL kernel under higher order corrections. In this spirit the intercept of the BFKL pomeron is extracted from a fit to the forward jet data of the H1 and ZEUS collaborations.
## 2 Forward jets at HERA
Back in 1990 A. H. Mueller proposed to look for DIS events with one jet –other than the jet originating from the struck quark– fulfilling the following characteristics (cf. fig. 1):
* $`𝐱`$ small. The selection of the smallest $`x`$–Bjorken experimentally possible implies going away from the domain of validity of DGLAP and into a phase space area governed by BFKL.
* $`𝐱_𝐉`$ large. This, along with the previous item, provides the phase space for parton evolution and has the extra advantage to enable the use of parton density functions (PDF) of the proton in a region where they have already been measured, so no extrapolation is needed.
* $`𝐤_𝐉^\mathrm{𝟐}𝐐^\mathrm{𝟐}`$. This requirement suppresses DGLAP evolution without affecting the BFKL dynamics. It also provides a big scale at both ends of the ladder, avoiding thus the dangerous infrared regions of phase space and increasing the solidity of the analytic predictions.
The process so selected provides a well defined topology for the experimentalist. One has to look for a DIS event with a jet at high rapidities and enforce that the virtuality of the proton is of the same scale as the transverse momentum squared of the jet. As the direction of the proton is termed by the HERA experiments the forward direction, this kind of events are known as forward jets events. Recently both the H1 and the ZEUS collaborations have presented cross sections for this observable. The H1 collaboration selects DIS events requiring the energy ($`E_e`$) and polar angle ($`\theta _e`$) of the scattered electron to be $`E_e>11`$ GeV and $`160^{}<\theta _e<173^{}`$. ZEUS only requires $`E_e>10`$ GeV and the polar angle $`\theta _e`$ is restricted by the rest of the kinematic constrains. Both of them select events with $`y_{Bj}>0.1`$ to avoid the jet from the struck quark going forward. In the jet side, selected with a cone algorithm of radius 1, the cuts used are for H1: $`k_J>3.5`$ GeV, $`x_JE_J/E_{beam}=E_J/820`$GeV$`>0.035`$ and $`7^{}<\theta _J<20^{}`$. ZEUS ask for jets with $`k_J>5.0`$ GeV, $`x_J>0.036`$ and $`8.5^{}<\theta _J`$. Both collaborations require that $`0.5<k_J^2/Q^2<2.0`$. The luminosity used is 2.8 and 6.4 pb<sup>-1</sup> for the H1 and ZEUS collaborations respectively. The H1 collaboration also reports a similar search, where all cuts were kept the same, except that $`k_J>5.0`$ GeV was required. The measured cross sections for different bins in $`x`$–Bjorken are presented in the first column of tables 1 and 2. Note that both collaborations presented slightly asymmetrical systematic errors. The errors presented in tables 1 and 2 are an average of the systematic errors added in quadrature to the statistical error of the measurement. Note also that the measurements at the lowest $`x`$–Bjorken have not been included. This is because, due to the $`0.5<k_J^2/Q^2<2.0`$ cut, the experiments ran out of phase space. Thus these points mix dynamic effects with phase space restrictions and are not useful for the analysis presented here.
## 3 BFKL fits to the forward jet data
The cross section for DIS events containing a forward jet has been calculated at leading logarithmic approximation within the BFKL formalism in references . There the following form has been found:
$$k_J^2x_J\frac{d^4\sigma }{dxdQ^2dx_Jdk_J^2}=C\alpha _s(Q^2)F(x_J,Q^2)\sqrt{\frac{Q^2}{k_J^2}}\frac{\mathrm{exp}[(\alpha _p1)\mathrm{ln}x_J/x]}{(\mathrm{ln}x_J/x)^{1/2}}$$
(1)
where $`\alpha _s(Q^2)`$ is the QCD coupling constant at the scale $`Q^2`$ and $`F`$ is a generic parton distribution in the proton given by
$$F(x_J,Q^2)=x_JG(x_J,Q^2)+\frac{4}{9}[x_Jq(x_J,Q^2)+x_J\overline{q}(x_J,Q^2)].$$
(2)
In an explicit form for the parameters $`C`$ and $`\alpha _p`$ can be found. Recently the NLO corrections to the BFKL kernel have been presented . These corrections turn out to be very large implying that NNLO calculations are needed. In particular the corrections affected the intercept of the BFKL pomeron $`\alpha _p`$, reducing the predictive power of the explicit forms given in at LO and in at NLO. This means that the exact power of the leading behavior of equation (1),
$$(\frac{x_J}{x})^{\alpha _p1},$$
(3)
is not numerically known from BFKL calculations. Nonetheless given that the first experimental data for this process are available, it is tempting to test if the form of equation (1) does indeed describe the measurement, and if so, which intercept of the BFKL pomeron is favored by the data.
To perform the fit there are some other ingredients needed. Standard parton density functions are required to evaluate $`F`$. Also as equation (1) is a four differential expression, values for $`x`$, $`Q^2`$, $`x_J`$ and $`k_J^2`$ are needed in each $`x`$–Bjorken bin. Both collaborations, H1 and ZEUS, report that the Monte Carlo generator Ariadne describes very well, not only the $`x`$–Bjorken dependence, but all distributions involved in the analysis. So a similar analysis to that reported by both collaborations has been performed at the hadron level of the Ariadne Monte Carlo. It has been checked that the cross section obtained with this procedure agrees with both, the reported data and also with the expectations from the Ariadne Monte Carlo as given by the H1 and ZEUS collaborations. Using this Monte Carlo data set for forward jet events the mean values of the variables $`x`$, $`Q^2`$, $`x_J`$ and $`k_J`$ have been estimated for each measured point of both collaborations. The results are presented in tables 1 and 2 along with the measured data points and their errors.
Using the input of tables 1 and 2, the PDFs given by GRV-LO , GRV-HO , CTEQ-4M and MRS-R1 –last three calculated in the $`\overline{\text{MS}}`$ scheme–, and an $`\alpha _s`$ value consistent with each of the different parton density functions, a 2 parameter fit was performed separately to the H1 and the ZEUS points using formula (1). The $`\chi ^2`$ and the value of $`\alpha _p`$ obtained are shown in table 3.
## 4 Discussion
The results of the fits can be summarized as follows:
1. All data points could be successfully fit to the form of equation (1). This is quite encouraging because that was the main motivation for this kind of measurements.
2. The fits are insensitive to the PDF used. This result is also as expected. As a matter of fact this was one of the main advantages of the forward jet proposal.
3. Different values of the exponent $`\alpha _p`$ are found when using LO or NLO PDF. On the one hand, being the formula (1) a LO approximation to BFKL, one is tempted to consider only the use of LO PDFs. On the other hand, the whole idea of performing the fits is to have a data driven estimation of the effects of higher order corrections to the BFKL kernel. The actual variation of the value of $`\alpha _p`$ is expected in the basis of it being proportional to $`\alpha _s`$ in LO. As the value of $`\alpha _s`$ decreases in going from LO to NLO, the value of $`\alpha _p`$ must compensate this trend and increase from one case to the other.
4. The values of $`\alpha _p`$ found using H1 data are different to those found using ZEUS measurements. On the one hand ZEUS data is more precise having a factor of 3 more luminosity. This allowed the ZEUS collaboration to measure the cross section in a region definitely dominated by DGLAP, having thus a clean transition from the box diagram at the top of figure 1 to the case when the box is complemented with a ladder. This greatly constrains the fit to equation (1). On the other hand H1 points reach smaller $`x`$, which is the interesting region for BFKL studies, although still with huge errors. One possible explanation for the difference in the values obtained using the data of both collaborations, could be again the $`\alpha _s`$ dependence of $`\alpha _p`$. Note that the average $`Q^2`$ is a lot bigger in the ZEUS measurement than in the H1 case. In the LO approximation this would naively produce a 15 to 20% increase of $`\alpha _p`$ from H1 data with respect to that from ZEUS cross sections.
5. The formula (1) is at parton level, i.e. it does not includes hadronization effects. As reported by both collaborations these are uncertain. Different models yield not only different normalization, but may also yield a $`x`$ dependence (see for example figure 8 on reference ) of the correction from hadron to parton level.
6. Using the dipole approach to BFKL, it has been shown that the measurements of the structure function $`F_2`$ can be described using an intercept of $`\alpha _p=1.28`$ . Some care is necessary when comparing this value with the one obtained here. The question of infrared divergencies due to the random walk generated by the BFKL kernel is quite sensitive for the $`F_2`$ case, but it does not appear in the case of forward jets. Nevertheless is quite comforting that two so different observables yield results compatible with the BFKL formalism.
7. The possibility to experimentally reach lower $`x`$ at still sizable $`Q^2`$ is very important. The fits presented here, although they could not used the lowest $`x`$ points due to phase space constrains, have shown that the BFKL dynamics enforces a steep rise of the cross section for a process governed by a hard pomeron inspite the NLO corrections to the BFKL kernel. Reaching smaller $`x`$ will allow to access the saturation region of hot spots in the proton, which was one of the primary motivations of the forward jet proposal. This goal may still be reachable at HERA.
## 5 Conclusion
A fit was performed of a BFKL prediction to forward jet production as measured by the H1 and ZEUS collaborations. All data were consistent with the assumption of using the BFKL formula for this process. Difference in the intercept of the pomeron obtained with different sets of data, may be assigned to its $`\alpha _s`$ dependence. The fits support the idea of a hard pomeron and point in the direction that when all higher order corrections are taken into account the forward jet cross section will still be rising quite rapidly.
|
no-problem/9812/astro-ph9812018.html
|
ar5iv
|
text
|
# The “angular size - redshift” relation for compact radio structures in quasars and radio galaxies Table 1 is available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/Abstract.html
## 1 Introduction
Classical tests of cosmological world models using the observed dependence of the angular size of galaxies or kiloparsec-scale radio sources have been inconclusive. At optical wavelengths, observational uncertainties at large redshift are large due to the small size of a galactic disk, seeing, the difficulty in defining a true metric rod, and possible evolutionary effects (e.g. Sandage 1988). At radio wavelengths, the separation of the lobes of extended double radio sources may be determined with great accuracy even at large redshift, but the interpretation of the “angular size – redshift” ($`\theta z`$) relation for double radio sources has been obscured by possible selection and evolutionary effects. The observed $`\theta z`$ relation for double radio sources appears to follow a simple $`1/z`$ law even at high redshift, in apparent contradiction to any simple Friedmann-Robertson-Walker (FRW) model without evolution (e.g. Kapahi 1989). Most researchers interpret the observed $`\theta z`$ diagram for double radio sources as evidence for a decrease in linear size with redshift (Kapahi 1987, Barthel and Miley 1988, Neeser et al. 1995). However, Singal (1993) and Nilsson et al. (1993) consider that the observed departure from the FRW curves is due to an inverse “linear size – luminosity” correlation which preferentially selects the smaller (high luminosity) sources at high redshifts. It is curious, however, that these selection or evolutionary effects apparently combine with cosmological effects to give the simple observed $`1/z`$ relation.
More recently, Buchalter et al. (1998) have studied a sample of 103 double lobed quasars with $`z>0.3`$ using the VLA at 20 cm in its B-configuration. In contrast to the $`1/z`$ “angular size – redshift” relation found for double lobed radio sources by other workers, Buchalter et al. find no change in apparent angular size in the range of $`1.0z2.7`$, consistent with FRW models without significant evolution. But, it is not clear to what extent their results are affected by the limited range of angular size, between 12 and 120 arcseconds, which can be observed with the VLA in the B-configuration at 20 cm.
The size of extended double lobe source whose linear extent is typically hundreds of kiloparsecs, may depend on the systematic changes in the properties of the intergalactic medium with $`z`$. Moreover, high redshift extended sources have ages which are comparable to the age of the Universe, and so evolutionary effects are not unexpected. Compact radio jets associated with quasars and AGN, by contrast, are typically less than a hundred parsec in extent. Their morphology and kinematics probably depends more on the nature of the “central engine” than on the surrounding intergalactic medium. The “central engine” itself is thought to be controlled by a limited number of physical parameters, such as the mass of central black hole, the strength of magnetic field, the accretion rate, and, possibly, the angular momentum. This central region may be “standard” for sources in which these parameters are confined within restricted ranges. Also, because the compact radio jets have typical ages of only some tens of years, they are young compared to the age of the Universe, at any reasonable redshift. Therefore, compact radio sources may offer an evolution free sample to test world models over a wide range of redshift.
However, the size of compact radio jets is not unambiguously defined and depends on the frequency of observation as well as on resolution. Moreover, differences in the spectral index between the core and jet components may introduce a K-like correction which can be important for high redshift sources, as this may introduce an apparent “linear size – redshift” dependence even in the absence of evolution (Kellermann 1993). Frey et al. (1997) have shown that this is likely to be a weak dependence, however more detailed images at various frequencies with matched resolution be needed to verify the importance of any K-like correction.
In several previous studies we have reported on the observed $`\theta z`$ relation for compact radio sources. Kellermann (1993) studied a sample of 79 quasars and AGN’s which had been observed with VLBI at 5 GHz and which have a 5 GHz luminosity greater than $`10^{24}`$ W/Hz. There are only a few sources at low redshift which meet the luminosity restriction, but these are consistent with a $`1/z`$ relation, characteristic of the Euclidean geometry which describes the local Universe. The Kellermann (1993) sample already includes all sources with luminosity greater than $`10^{24}`$ W/Hz at redshifts less than a few tenths, and further surveys down to fainter flux density limits will not find any additional sources which satisfy the luminosity criteria. The important point of the Kellermann (1993) paper was that for redshifts in the range $`0.5<z<3`$, the angular size appears to be essentially independent of redshift, in contrast to the $`\theta z`$ relation for powerful extended sources which continues its apparent Euclidean form out to large redshifts. Kellermann noted that the observed form of the $`\theta z`$ relation for the compact source sample was qualitatively consistent with a standard FRW cosmology with $`\mathrm{\Omega }=1`$ without the need to appeal to arguments based on size evolution or a “linear size – luminosity” dependence. A more rigorous quantitative analysis of this data by Stepanas and Saha (1995) find a best fit of $`q_{}=2.6_{2.2}^{+2.1}`$ with the 90% confidence, and they exclude the simple $`\theta 1/z`$ relation at the 99% confidence level. Using the same data, Kayser (1995) has applied a Kolmogorov-Smirnov test to compare the linear sizes of high ($`z>0.75`$) and low ($`z<0.75`$) redshift compact radio sources for different cosmological models and also concludes that the available data allow models with a wide range of $`\mathrm{\Omega }`$ and the cosmological constant, $`\mathrm{\Lambda }`$.
In a separate investigation, Gurvits (1993) used two point VLBI visibility data obtained at 13 cm (Preston et al. 1985) for 337 sources in order to show qualitatively that the observed data suggests $`q_{}0.5`$. A four-parameter regression analysis of the same sample gave a value of $`q_{}=0.16\pm 0.71`$ (Gurvits 1994). The same analysis also gave estimates of the dependence of the apparent angular sizes of compact sources on their luminosity and emitting frequency.
More recently, Wilkinson et al. (1998) have reported on the $`\theta z`$ relation for sources taken from the Caltech-Jodrell Bank VLBI CJF sample of 160 flat spectrum radio sources (Taylor et al. 1996 and references therein). As in the studies of Kellermann (1993) and Gurvits (1994), Wilkinson et al. find no dependence of angular size on redshift for sources with $`0.5<z<3`$, but conclude that uncertainties in defining the angular size of complex jets, in the K-like correction, in a possible “size – luminosity” dependence, in the effects of orientation, as well as in possible size or luminosity evolution restrict the usefulness of compact sources to accurately constrain the value of $`q_{}`$. Dabrowski, Lasenby, and Saunders (1995), as well, have pointed out the difficulty in obtaining a meaningful constraint on $`\mathrm{\Omega }`$ due to the effects of relativistic beaming in limited source samples.
Krauss and Schramm (1993) and Stelmach (1994) have pointed out that if evolutionary effects can be ruled out, then the form of the “angular size – redshift” relation can put significant limits on the value of the cosmological constant, $`\mathrm{\Lambda }`$, as well as on $`\mathrm{\Omega }`$. Jackson and Dodgson (1996) have shown that while the data presented by Kellermann (1993) are consistent with $`\mathrm{\Omega }_{}1`$ and $`\mathrm{\Lambda }_{}=0`$, since there is not a well defined minimum in the $`\theta z`$ dependence, equally good fits to the data are obtained with smaller values of matter density described by $`\mathrm{\Omega }_{}<1`$, and negative values of $`6\mathrm{\Lambda }_{}2`$.
With the aim of better restricting the allowable range of cosmological parameters, we have compiled a new larger sample of sources than used by Kellermann (1993) or by Wilkinson et al. (1998) but with more complete structural data than used by Gurvits (1993, 1994). We note, that the sample discussed here is inhomogeneous as it is based on VLBI images published by various authors using a variety of antenna configurations and different techniques for image reconstruction. In Section 2 we discuss the definition of our sample, and in the following sections we discuss the apparent “angular size – redshift” dependence.
## 2 The source sample
The new list contains all sources found in the literature which were imaged with VLBI at 5 GHz with a nominal resolution of about 1.5 mas in the east-west direction and with a dynamic range of at least 100. The list includes the all-sky set of 79 sources discussed by Kellermann (1993), but enhanced by more recently published work, mostly by the Caltech–Jodrell Bank group (Xu et al. 1995, Henstock et al. 1995, Taylor et al. 1994, 1996) of sources at declination greater than 35 degrees. We also included in the sample a number of sources published by other authors and our own recent observations of quasars with measured redshifts greater than 3 (Frey et al. 1997, Paragi et al. 1998).
Our sample differs from the compilation of Wilkinson et al. (1998) primarily in that it includes a number of relatively strong sources at declinations south of +35 degrees, and other sources not presented in the CJF sample. By including sources outside the range of CJF declinations we are able to better sample the sparsely populated low redshift (Euclidean) part of the $`\theta z`$ diagram not included in the CJF sample. The observations of Gurvits et al. (1992, 1994), Frey et al. (1997) and Paragi et al. (1998) were made in an attempt to better sample the high redshift part of the $`\theta z`$ diagram which is particularly sensitive to the value of $`q_{}`$. In particular, the unambiguous detection of an increase in angular size at the highest redshifts would indicate a value of $`q_{}>0`$. The increase in the size of our present sample comes at the expense of homogeneity and the need to use published VLBI contour maps instead of the primary data. We have attempted to minimize the effect of these inhomogeneities by using the following criteria.
As in Kellermann (1993), we define the characteristic angular size of each source as the distance between the strongest component, which we refer to as the core, and the most distant component which has a peak brightness greater than or equal to 2% of the peak brightness of the core. For sources which are slightly resolved or unresolved, we adopted the following procedure. We assume that sources are one dimensional. For sources which appear resolved in at least one direction, we estimated the distance of a secondary component from the core, or its upper limit, from the published contours. If the source was not resolved, we took the size of the major axis of the synthesized beam (FWHM) as an upper limit to the size, unless there was additional information which indicated that the source structure axis lies along a specific direction different from the direction of major axis. The latter applies to those sources which show extensions in a particular direction, including an extension along the minor axis of the synthesized beam. In this case we estimated the upper limit of the size as the size of the beam along the direction of extension. Thus, our approach to measuring source size allows for four different cases:
Case C: the distance between the core and a 2%-component;
Case J: an upper limit of the size measured as the size of the synthesized beam along the direction of apparent extension, most likely – a jet;
Case L: an upper limit of the size measured along the major axis of the synthesized beam;
Case S: an upper limit of the size measured along the minor axis of the synthesized beam.
For those sources where multi epoch VLBI images are available, we have used the most recent epoch that meets our criteria of sensitivity and dynamic range. Finally, we excluded from our analysis all unresolved sources if the major axis of the primary beam exceeded 2.2 mas (i.e. all cases L with too large a synthesized beam) and sources which are known to be gravitationally lensed.
The resulting sample of 330 sources is presented in Table 1<sup>1</sup><sup>1</sup>1Table 1 is available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/Abstract.html, where we show the IAU source designation and alternative name in columns 1 and 2. The redshift and optical counterpart are given in columns 3 and 4. Columns 5, 6 and 7 give the flux density at 6 and 20 cm (or a footnote for alternative wavelength) and the two-point spectral index, $`\alpha `$ ($`S\nu ^\alpha `$), respectively. Columns 8 and 9 give the angular size (or its upper limit) and the one-letter structure code as explained above. In columns 10–13, we list references for redshift, flux densities at 6 and 20 cm, and for the VLBI image respectively.
## 3 Properties of the sample
The distribution of redshifts for the sources in our sample is shown in Fig. 1. The 79 sources used in the analysis of Kellermann (1993) are shown shaded.
The histogram of the spectral index distribution is shown in Fig. 2. We note, that we have used the value of spectral index as calculated, in most cases, from measurements of total flux density on arcsecond or larger angular scales although our discussion of angular dimensions is based on milliarcsecond-scale structures, which account for only part of the total flux. For most of the sources, this distinction is not important as nearly all of the flux density in sources of interest is contained in the compact component. Where relevant, such as for Cyg A, we specifically used the flux density of the core component. In a few special cases marked in column 6, when flux density at 20 cm was not available, the spectral index was calculated between 6 cm and another longer wavelength as explained in the footnotes.
Fig. 3 shows the luminosity of all the sources in our sample as a function of their redshift. (Throughout this paper we use $`H_{}=100h\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$ and a deceleration parameter $`q_{}=0.5`$ to calculate the luminosity). The shape of the luminosity – redshift diagram and the narrow dispersion simply reflects the fact that our sample, although compiled on an ad-hoc basis from the literature and based upon various selection criteria, is basically a flux-limited sample.
## 4 Properties of angular size
In Fig. 4, we plot the measured angular size against redshift for all 330 sources in our sample. For the well resolved sources, the procedure of measuring $`\theta `$ gives an unambiguous estimate of a metric size. But, for sources with maximum dimensions comparable to or smaller than the beam size, there are large uncertainties. For this reason, and to minimize the influence on our analysis of the few sources with extremely large dimensions, we have chosen to bin the data and to examine the change in median angular size with redshift. This allows us to treat equally true metric sizes of resolved sources and upper limits of sizes for slightly resolved or unresolved ones. Fig. 5 shows the binned data of median angular size plotted against redshift for the same data. (Here and throughout this paper we use nearly equally populated bins, which number is close to $`\sqrt{N}`$, where $`N`$ is the size of the sample.) As found in previous studies, for $`z0.5`$, the median angular size is nearly independent of redshift. In this figure, as an example we show a family of curves for a standard rod in various world models. We note, that none of these curves represent the best fit discussed below.
### 4.1 “Angular size – luminosity” and <br>“angular size – spectral index” relations
As it is clear from Fig. 3, our sample contains sources with luminosity ranging over more than 4 orders of magnitude. Fig. 6 shows the relation between median angular size and luminosity (the same binning in redshift space as in Fig. 5).
Fig. 7 shows the dependence of the angular size on spectral index. As expected from simple consideration of self absorption arguments, sources with flat and inverted spectra should, on average, have smaller sizes. Fig. 7 qualitatively confirms this expectation. It may also indicate a presence of one or several selection effects. However, as illustrated by Fig. 8, we do not find an evidence on systematic correlation between $`\alpha `$ and $`z`$, which might be responsible for the appearance of the “$`\theta \alpha `$” dependence, shown in Fig. 7. The only possible exception could correspond to the lowest redshift bin. However, this bin represents sources of considerably lower luminosity (cf. Fig. 3), which could differ intrinsically from their higher redshift counterparts.
### 4.2 Toward estimating cosmological parameters from the <br>$`\theta z`$ relation
The $`\theta z`$ relation based on the data described here is qualitatively consistent with $`0q_{}1`$ and $`\mathrm{\Lambda }=0`$ without the need to introduce evolutionary effects. The new data, in agreement with presented earlier by Kellermann (1993), Gurvits (1994) and Wilkinson et al. (1998), do not show clear evidence for an angular size minimum near $`z=1.25`$ as expected for models with $`\mathrm{\Omega }=1`$ and $`\mathrm{\Lambda }=0`$. The near asymptotic slope of the $`\theta z`$ relation is more characteristic of models with $`\mathrm{\Omega }<1`$ and allows values of $`\mathrm{\Lambda }0`$.
These results are, however, based on very inhomogeneous data obtained by many different observers using different instruments and imaging techniques. New VLBI observations now in progress will improve the accuracy of the observed $`\theta z`$ relation as it will provide a uniform data set for analysis using the $`uv`$-data and images rather than published contour maps.
With all the reservations discussed above, as an example of a cosmological test with the $`\theta z`$ relation on milliarcsecond scale, we consider a multi-parameter regression analysis as described by Gurvits (1994) with modifications made by Frey (1998). It is based on the following phenomenological expression
$$\theta l_mD^1(z)lhL^\beta (1+z)^nD^1(z),$$
(1)
where $`l_m`$ is the metric linear size, $`D`$ is the angular size distance, $`lh`$ is the linear size scaling factor, $`L`$ is the source luminosity. Parameters $`\beta `$ and $`n`$ represent the dependence of the linear size on the source luminosity and redshift, respectively. For a homogeneous, isotropic Universe ($`q_{}>0`$) with the cosmological constant, $`\mathrm{\Lambda }=0`$, $`D(z)`$ is given by the usual expression
$$D(z)=\frac{q_oz+(q_o1)\left(\sqrt{1+2q_oz}1\right)}{q_o^2(1+z)^2}.$$
(2)
The regression model (Gurvits 1994, Frey 1998) allows us to fit the $`\theta z`$ relation with four free parameters, the linear size scaling factor $`lh`$, the deceleration parameter $`q_0`$ and two parameters related to the physics of compact radio emitting regions, $`\beta `$ and $`n`$. The value of $`n`$, in turn, could in principle represent three different physical dependences: (i) a cosmological evolution of the linear size; (ii) a dependence of the linear size on the emitted frequency; and (iii) an impact of sources broadening due to scattering in the propagation medium. The latter effect is not important for our sample with the lowest emitted frequency of 5 GHz (which corresponds to $`z=0`$). The distinction between the former two effects is beyond the scope of this paper and will require multifrequency $`\theta z`$ tests.
To minimize any possible dependence of linear size on luminosity, we restrict the regression analysis to sources with $`Lh^210^{26}`$ W/Hz. Furthermore, as is evident from Fig. 7, there is an obvious dependence of angular size on spectral index. In order to minimize this effect on the regression, we choose only those sources which form a flat segment of the $`\theta \alpha `$ diagram $`0.38\alpha 0.18`$ (Fig. 7). This selection criterion also partially excludes from the analysis the lowest redshift bin which represents the highest deviation on the $`\alpha z`$ diagram (Fig. 8). By restricting the range of spectral indices, we are able to further restrict the dispersion in intrinsic size in our analysis. Specifically, we exclude many of the relatively large compact steep spectrum sources and most compact inverted spectrum sources. There are 145 sources which meet these criteria, their distribution in redshift space is shown in Fig. 9, and the median angular sizes versus redshift is shown in Fig. 10.
As an example, we apply the four parameter regression model for median values of this sub-sample grouped into 12 redshift bins. The best fit values and corresponding 1$`\sigma `$ errors are: $`lh=23.8\pm 17.0`$ pc, $`\beta =0.37\pm 0.27`$, $`n=0.58\pm 1.0`$, and $`q_0=0.33\pm 0.11`$. This result is in qualitative agreement with similar estimates obtained for an independent sample of sources and different technique of measuring their angular sizes (Gurvits 1994).
In Table 2, we show the results of regression modeling of the same binned sub-sample for the two parameters, $`lh`$ and $`q_0`$, for different fixed values of $`\beta `$ and $`n`$. The ranges for $`\beta `$ and $`n`$ shown do not require a substantial evolution of linear sizes with redshift and luminosity. We note, that the range of parameter $`\beta `$ used covers the estimate obtained for kiloparsec-scale structures in FRII sources by Buchalter et al. (1998; $`\beta 0.13\pm 0.06`$) and is close to the estimate obtained earlier for kiloparsec-scale structures in quasars by Singal (1993; $`\beta 0.23\pm 0.12`$). Similarly, our range of the parameter $`n`$ is close to the estimates obtained in both papers for kiloparsec-scale structures (Singal 1993, Buchalter et al. 1998). However, one must keep in mind that closeness of these values for kiloparsec-scale structures in double radio sources and in our parsec-scale structures could be superficial since the radio emission on these scales, differed by several orders of magnitude, is governed by different physical processes.
As a test of our method of using median values for binned data, we repeated the same procedure for the sub-sample of 145 sources in which the upper limits of angular size (shown in Table 1 with the sign “$`<`$” in column 8) are replaced with an arbitrary value of 0.1 mas. The difference between estimates of $`lh`$ and $`q_0`$ for this test case and values presented in Table 2 does not exceed 4.1% within the range of $`\beta `$ and $`n`$ shown in Table 2. We therefore conclude that the use of median values is justified.
Values of $`q_0`$ shown in Table 2 should be treated with caution due to the deficiencies of the sample and the method described above. For the simple case with no dependence of the source linear size on the source luminosity and redshift (“true” standard rod, $`\beta =0`$ and $`n=0`$) $`q_0=0.21\pm 0.30`$. This result does not contradict to the estimate of $`q_{}`$ in $`\mathrm{\Lambda }=0`$, ”no-evolution” ($`\beta =n=0`$) by Buchalter et al. (1998). Solutions, which allow evolution of source size with redshift ($`n0`$) or a dependence on luminosity ($`\beta 0`$), favor values of $`q_00.5`$ for $`\beta +n0.15`$.
## 5 Summary
The 5 GHz VLBI data are consistent with standard FRW cosmologies with $`0q_{}0.5`$ and $`\mathrm{\Lambda }=0`$ without the need to introduce evolution of the population or to appeal to selection effects caused by a possible “luminosity – linear size” dependence. This conclusion is based on the “angular size – redshift” test using an inhomogeneous sample of 330 VLBI images, with the 1.5 mas nominal angular resolution and the dynamic range at least 100. A two-parameter regression model applied for a plausible range of dependence of linear size on luminosity and redshift is used to separate the “$`\beta n`$”parameter space and gives a deceleration parameter somewhat lower than the critical, $`q_{}0.5`$, for $`\beta +n0.15`$. Such an approach might be useful to further restrict the deceleration parameter using a better understanding of physics of the compact radio structures, represented by parameters $`\beta `$ and $`n`$.
We also find a dependence of angular size with spectral index which, if not considered, increases the dispersion in linear size. Elimination of extreme values of spectral indices with $`\alpha <0.38`$ and $`\alpha >0.18`$ better defines compact sources as standard rods.
In view of the size and selection bias in the currently available sample, we have chosen not to consider more general models with $`\mathrm{\Lambda }0`$. However, we present our full data set for those may wish to use these data to further investigate constraints on the cosmological parameters.
###### Acknowledgements.
We are grateful to the referee for helpful comments. LIG acknowledges partial support from the European Union under contract No. CHGECT 920011, the Netherlands Organization for Scientific Research (NWO) programme on the Early Universe, and the European Commission, TMR Programme, Research Network Contract ERBFMRXCT 96–0034 “CERES”. SF acknowledges financial support received from the NWO and the Hungarian Space Office, and hospitality of JIVE and NFRA during his fellowship in Dwingeloo. KIK acknowledges the hospitality of JIVE during two visits. We acknowledge the use of the NASA/IPAC Extragalactic Database (NED), which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The National Radio Astronomy Observatory is operated by Associated Universities, Inc. under a Cooperative Agreement with the National Science Foundation.
|
no-problem/9812/patt-sol9812003.html
|
ar5iv
|
text
|
# The exact solutions of differential equation with delay
\[
## Abstract
The exact solutions of the first order differential equation with delay are derived. The equation has been introduced as a model of traffic flow. The solution describes the traveling cluster of jam, which is characterized by Jacobi’s elliptic function. We also obtain the family of solutions of such type of equations.
PACS numbers: 05.45.Yv, 02.30.Ks, 45.70.Vn, 05.45.-a
patt-sol/9812003
\]
In this paper, we investigate following differential-difference equation,
$$\frac{d}{dt}x_n(t+\tau )=f(x_{n+1}(t)x_n(t)),$$
(1)
where $`\tau `$ is a real positive constant called “delay”. The index $`n`$ takes integer. The set of equations of this type has been introduced for a car-following model of traffic flow . In that case $`x_n`$ is the position of the $`n`$th car. Equation (1) has been popular in many physical phenomena of relaxation towards an optimal equilibrium state, such as relaxation effect in gases, chemical reactions and synchronization problem.
As a model for traffic flow, $`f(x)=\mathrm{tanh}(x)`$ is the reasonable choice , which was first introduced in the model ,
$$\tau \frac{d^2}{dt^2}x_n(t)=f(x_{n+1}(t)x_n(t))\frac{d}{dt}x_n(t).$$
(2)
This model may have some relation to Eq.(1) for relatively small $`\tau `$ . Many studies have been made to the model of Eq.(2) with $`f(x)=\mathrm{tanh}(x)`$, which has the stable traveling cluster solution as traffic jam . The traveling cluster solution is characterized as a soliton of modified Korteweg-de Vries (MKdV) equation in the vicinity of the critical point . One interesting question is whether Eq.(1) has such traveling cluster solution or not. Our simulation result suggest the existence of such solution, which describes a traveling cluster moving backward with the velocity $`v=1/(2\tau )`$ . Recently, we have an information that Igarashi, Itoh and Nakanishi have found the exact solution of Eq.(1), which characterized by the theta function . In this paper we also derive a series of analytic solutions of traveling cluster in the context of our work, and confirm their stability by the numerical simulations.
It is convenient to introduce new variable $`h_n(t)=x_{n+1}(t)x_n(t)`$ and rewrite Eq.(1),
$$\frac{d}{dt}h_n(t+\tau )=f(h_{n+1}(t))f(h_n(t)).$$
(3)
We start at the linear theory assuming that the amplitude $`h_n(t)`$ is infinitesimal and $`f(h_n)=f(0)+f^{}(0)h_n`$. Equation (3) becomes
$$\frac{d}{dt}h_n(t+\tau )=h_{n+1}(t)h_n(t).$$
(4)
Here we set $`f^{}(0)=1`$ without loss of generality. We investigate the solution of the form
$$h_n(t)=\mathrm{exp}(i\alpha n+i\omega t)$$
(5)
for real $`\alpha `$ and $`\omega `$. Inserting Eq. (5) to Eq. (4) we obtain
$$\frac{\mathrm{sin}\alpha /2}{\alpha /2}=\frac{1}{2\tau },$$
(6)
and
$$\alpha =2\omega \tau .$$
(7)
Equation (6) has a solution if $`2\tau `$ is larger than $`1`$, which means $`\tau =1/2`$ is critical. If $`\tau `$ is a little bit larger than $`1/2`$, Eq.(6) has only two solutions $`\alpha =\pm \lambda `$. Then, we obtain the solution of Eq.(4) as
$$h_n(t)=\mathrm{exp}\pm i\lambda (n+\frac{t}{2\tau }).$$
(8)
This represents a traveling wave solution with the velocity $`1/(2\tau )`$ in the space of index $`n`$, which is treated as a continuous variable. The wave moves backward against the numbering direction, which appears as the traveling wave in the real space moving backward in the flow of $`x_n`$. The above analysis is first given by Whitham .
Now let us investigate the exact traveling wave solution of Eq.(3). We treat the index $`n`$ as a continuous variable, and change the notation $`h_n(t)`$ to $`h(n,t)`$. We introduce new variables on the moving flame of traveling wave as $`u=n+vt`$, where $`v`$ is the velocity of the traveling wave. We search the solution which does not change its form on this flame. We define the amplitude of traveling wave
$$H(u)=H(n+vt)h(n,t).$$
(9)
Eq.(3) for the amplitude $`H(u)`$ is expressed as
$$v\frac{d}{du}H(u+v\tau )=f(H(u+1))f(H(u)).$$
(10)
Replacing $`u`$ by $`u1/2`$, we get more symmetric form
$$v\frac{d}{du}H(u+\sigma )=f(H(u+\frac{1}{2}))f(H(u\frac{1}{2})),$$
(11)
where
$$\sigma =v\tau \frac{1}{2}.$$
(12)
We investigate the solution under the condition $`\sigma =0`$. This means that the traveling wave of such solutions propagates backward with just the same velocity as the linear theory,
$$v=\frac{1}{2\tau }.$$
(13)
Now we present the definite form of exact solutions giving concrete examples of $`f(x)`$. First we take $`f(x)=\mathrm{tanh}(x)`$, which is a suitable choice for a model of traffic flow. Introducing a new amplitude
$$G=f(H),$$
(14)
Eq.(11) with $`\sigma =0`$ is rewritten as
$$v\frac{dG(u)/du}{1G(u)^2}=G(u+\frac{1}{2})G(u\frac{1}{2}).$$
(15)
We can easily find a solution of Eq.(15) in the form
$$G(u)=\beta \mathrm{sn}(\alpha u,k),$$
(16)
where sn is Jacobi’s elliptic function with modulus $`k`$. The parameter $`\alpha `$ is determined by
$$\frac{\mathrm{sn}(\alpha /2,k)}{\alpha /2}=\frac{1}{2\tau },$$
(17)
and $`\beta `$ is given by
$$\beta =\pm k\frac{\alpha }{4\tau }.$$
(18)
Eq.(17) has a real solution only if $`1/(2\tau )<1`$. In the case of $`k=0`$, Eq.(17) reduces to the result of linear theory, Eq.(6). The modulus $`k`$ is a free parameter of the solution, which indicates the existence of many solutions for the same traveling velocity $`v=1/(2\tau )`$. The relation between the modulus and solutions is discussed later. Next we take $`f(x)=\mathrm{tan}(x)`$. In this case Eq.(15) is replaced by
$$v\frac{dG(u)/du}{1+G(u)^2}=G(u+\frac{1}{2})G(u\frac{1}{2}).$$
(19)
It is also easy to see that Eq.(19) has a solution
$$G(u)=\pm k\frac{\alpha }{4\tau }\mathrm{cn}(\alpha u,k),$$
(20)
where cn is another Jacobi’s elliptic function. The parameter $`\alpha `$ is determined by
$$\frac{\mathrm{sd}(\alpha /2,k)}{\alpha /2}=\frac{1}{2\tau },$$
(21)
which also reduces to the result of linear theory, if we take $`k=0`$.
The above two solutions suggest us to expect the existence of another solutions using elliptic functions. Actually, we have found the family of solutions of Eqs. (15) or (19). The solutions of this family are denoted in the form
$$G(u)=\pm C_k\frac{\alpha }{4\tau }\mathrm{g}(\alpha u,k),$$
(22)
where $`\alpha `$ is determined by
$$\frac{A(\alpha /2,k)}{\alpha /2}=\frac{1}{2\tau }.$$
(23)
In the above, g and $`A`$ are appropriate elliptic functions. The solution is given by each set of ($`\mathrm{g}`$,$`A`$,$`C_k`$) in Tables I and II.
Thus, all Jacobi’s elliptic functions are the solutions for either Eq. (15) or (19). This result is well understood from the following fact. Jacobi’s elliptic functions $`\mathrm{g}(\alpha u,k)`$ are connected each other by Jacobi’s imaginary transformation: $`\alpha i\alpha `$, the imaginary transformation of modulus: $`kik`$ and the translation of $`u`$: $`uu+K(k)/\alpha `$, where $`K(k)`$ is a quarter of the period of elliptic functions. These transformations preserve the form of Eq. (15) or (19), or exchange each other. Each pair of two elliptic functions connected by the translation gives an equivalent solution. The relation of solutions is shown in Fig.1.
The explicit form for the solution of Eq.(3) is given by $`H=f^1(G)`$. In this step some solutions for $`G`$ needs appropriate interpretation for the realistic meaning. The solution $`G`$ cs for $`G=\mathrm{tanh}(H)`$ is the case, which has the region of $`u`$ where $`|G|>1`$. On the other hand, the solution $`G`$ sn has a realistic meaning in the whole range of $`u`$.
Equation (15) or (19) includes the solutions for another choice of $`f(x)`$ besides $`\mathrm{tanh}(x)`$ and $`\mathrm{tan}(x)`$. Actually, $`f(x)=\mathrm{coth}(x)`$ leads to Eq.(15), and $`f(x)=\mathrm{cot}(x)`$ leads to Eq.(19) with $`vv`$. Jacobi’s elliptic functions also provide the solutions for these systems. This result presents some interesting picture. Let us compare the systems controlled by $`f(x)=\mathrm{tanh}(x)`$ and $`\mathrm{coth}(x)`$. At first sight, the behavior of these two systems seems different from each other. We note sn and ns are the solutions for both systems, which obey Eq.(15). The solution $`H\mathrm{arccoth}(\mathrm{ns})`$ in the system of $`f(x)=\mathrm{coth}(x)`$ has essentially the same form as the solution $`H\mathrm{arctanh}(\mathrm{sn})`$ in the system of $`f(x)=\mathrm{tanh}(x)`$. As the result, these two systems may have the same global phenomenon in spite of the different type of local interaction. This analysis is not simply applicable to the solution nc in the system of $`f(x)=\mathrm{coth}(x)`$, because cn is the solution of the system of $`f(x)=\mathrm{tan}(x)`$, which obeys the other Eq.(19).
We discuss the meaning of the modulus in our solution. The modulus $`k`$ determines the period of elliptic functions, which is related to the number of traveling cluster and the boundary condition. For example, $`k=1`$ gives kink like solution corresponding to the boundary condition $`G(\mathrm{})=G(\mathrm{})`$ ,
$$G(z)=\mathrm{tanh}(\alpha /2)\mathrm{tanh}(\alpha u)$$
(24)
with $`\alpha `$ satisfying
$$\frac{\mathrm{tanh}(\alpha /2)}{\alpha /2}=\frac{1}{2\tau }.$$
(25)
We perform the simulation to check the stability of our analytic solution. Fig.2 shows an elliptic solution for $`f(x)=\mathrm{tanh}(x)`$ given by sn
$$H(u,k)=\mathrm{arc}\mathrm{tanh}(k\frac{\alpha }{4\tau }\mathrm{sn}(\alpha u,k)),$$
(26)
with $`\tau =0.501`$, $`k=0.9965`$ and $`\alpha =0.15522`$ together with the result of simulation for Eq.(1) performed in the periodic boundary condition with a suitable initial condition.
We are convinced the set of Eqs. (15) and (19) has much more solutions constructed with elliptic functions. As a fact, the solution found by Igarashi, Itoh and Nakanishi is rewritten in our formulation as
$$G(u)=\beta \frac{\mathrm{sn}(\alpha (u+a))+\mathrm{sn}(\alpha (ua))}{\mathrm{sn}(\alpha u)}+\gamma ,$$
(27)
where $`a`$ is a free parameter. Actually this satisfies Eq.(15). Their solution includes Eq.(16) as a special case if we take $`\alpha a=K(k)/2`$ . The parameter $`\alpha `$ is determined by
$$\frac{\alpha }{4\tau }\frac{\mathrm{cn}(\alpha a)\mathrm{dn}(\alpha a)}{\mathrm{sn}(\alpha a)}+\frac{x_+x_{}}{x_+x_{}}=0,$$
(28)
where
$$x_\pm =1\frac{\mathrm{sn}^2(\alpha a)}{\mathrm{sn}^2(\alpha (a\pm \frac{1}{2}))}.$$
(29)
Then $`\beta `$ and $`\gamma `$ are given by
$$\beta =\pm \frac{\alpha }{4\tau \mathrm{sn}(\alpha a)},$$
(30)
and
$$\gamma =\pm \frac{x_++x_{}}{x_+x_{}}.$$
(31)
This solution represents the asymmetry of the widths of upper and lower plateaus for traveling cluster in contrast of the solution given by a single sn shown in Fig.2. Some combinations of elliptic functions in the numerator and denominator of Eq.(27) are solutions, those are constructed by the three transformations represented in Fig. 1. This type of solutions has asymmetry in contrast of the solutions given by a single elliptic function in Tables I and II. Probably, the set of Eqs. (15) and (19) has some algebraic structure in the space of solutions constructed with elliptic functions.
We remark all these solutions have the common velocity $`v=1/(2\tau )`$. Numerical simulation shows that the traveling cluster solutions of Eq.(1) preserve their velocity as $`v=1/(2\tau )`$ in the deformation of $`f(x)`$ beyond $`\mathrm{tanh}(x)`$. This fact suggests $`v`$ is some invariant quantity of the structure in the set of the solutions.
The set of differential-difference equations for $`G`$: Eqs. (15) and (19) offers rich contents of the system with traveling cluster solutions which characterized by elliptic functions. Eqs. (15) and (19) are related to some soliton systems. Our equations can be derived as the traveling wave equations for such soliton systems. Eqs. (15) and (19) correspond to one of the evolution equations discussed by Ablowitz and Ladik . The soliton systems related to Eq.(19) were widely discussed in the self dual network equations of nonlinear inductors and capacitors by Wadati , Hirota and Satsuma . Wadati showed the corresponding soliton system to Eq.(19) was derived from Lotka-Volterra system by Bäcklund transformation . These soliton systems reduce to MKdV equation. Our system of difference equation with delay may be related to soliton systems.
This work was partly supported by a Grant-in-Aid (No. 10650066) of the Japanese Ministry of Education, Science, Sports and Culture.
|
no-problem/9812/cond-mat9812107.html
|
ar5iv
|
text
|
# Cluster formulation of spin glasses and the frustrated percolation model: statics and dynamics
## I Introduction
Cluster concepts have been extremely useful in critical phenomena to elucidate the mechanism underlying a thermodynamical transition, by providing a geometrical interpretation of thermodynamic correlations. In the Ising model defined by the Hamiltonian
$$=J\underset{ij}{}S_iS_j,$$
(1)
a correct definition of clusters was given by Kasteleyn and Fortuin and by Coniglio and Klein . In the Coniglio-Klein approach, one throws a bond between pairs of parallel nearest neighbor spins with a probability $`p=1e^{2\beta J}`$, where $`\beta =1/k_BT`$. By summing over the spin configurations with the Boltzmann weight, the partition function of the model defined by Eq. (1) can be written as a sum over bond configurations $`C`$,
$$Z=\underset{C}{}e^{\mu b(C)}q^{N(C)},$$
(2)
where $`q=2`$, $`b(C)`$ is the number of bonds in the configuration $`C`$, $`N(C)`$ the number of connected clusters, and $`\mu =\mathrm{log}(\frac{p}{1p})=\mathrm{log}(e^{2\beta J}1)`$ is the chemical potential of the bonds. Thermodynamic averages can be related in this approach to corresponding percolative quantities. One finds that the clusters represent spin fluctuations, and percolate at the Ising critical temperature with Ising critical exponents.
This approach can be extended to the Potts model, in which spins can have $`q2`$ different states. In this case the parameter $`q`$ in Eq. (2) can assume a value $`q2`$, and for every value of $`q`$ the percolation model has the same critical temperature and exponents of the corresponding Potts model. For $`q=1`$ one recovers the random bond percolation. Sweeny studied the weighted percolation problem defined by Eq. (2) by Monte Carlo techniques on a two-dimensional square lattice. He showed that a simulation based on this approach does not suffer from critical slowing down for $`q<4`$. In a few Monte Carlo steps one can equilibrate even a very large lattice at the critical temperature. Thus he extracted informations about the critical point of the Potts model by measuring geometric quantities like the mean cluster size at the transition point. Later, the cluster approach was further elaborated by Swendsen and Wang and Wolff , by implementing an efficient cluster dynamics to simulate directly the spin model (1), which drastically reduces the critical slowing down of conventional spin flip simulations.
The cluster approach has been extended also to frustrated systems like spin glasses . With respect to the ferromagnetic case, new features appear, due to the phenomenon of frustration. In fact the percolation model that one obtains has the same complexity of the original spin model, and is not useful to define fast Monte Carlo dynamics, as in the unfrustrated case. Nevertheless, it represents an interesting tool to investigate the properties of frustrated spin systems from the geometrical point of view. Moreover, it can be considered on its own as a model of percolation in a frustrated medium, that makes it of relevance in the study of systems in which frustration and connectivity play a central role, as the structural glasses.
To illustrate the cluster approach to frustrated spin models, let us consider the Ising spin glass, defined by the Hamiltonian
$$=J\underset{ij}{}ϵ_{ij}S_iS_j,$$
(3)
where $`ϵ_{ij}`$ are quenched random variables that can have the values $`ϵ_{ij}=\pm 1`$. As in the ferromagnetic case, one throws a bond between nearest neighbor spins that satisfy the interaction (in this case if $`ϵ_{ij}S_iS_j=1`$) with probability $`p=1e^{2\beta J}`$. The crucial difference with the ferromagnetic case is that in general not all the interactions can be satisfied simultaneously. Indeed a closed loop such that the product of $`ϵ_{ij}`$ along the loop is negative does not admit any spin configuration satisfying all the interactions, and is called frustrated. Since bonds can be put only between spins that satisfy the interaction, one cannot put bonds on the lattice that close a frustrated loop. In terms of bond configurations, the partition function of the Ising spin glass turns out to be
$$Z=\underset{C}{}^{}e^{\mu b(C)}q^{N(C)},$$
(4)
where $`q=2`$, $`\mu =\mathrm{log}(\frac{p}{1p})=\mathrm{log}(e^{2\beta J}1)`$, $`b(C)`$ is the number of bonds, and $`N(C)`$ the number of clusters in the configuration $`C`$. Here the sum $`_C^{}`$ is extended to all the bond configurations that do not contain a frustrated loop. Therefore, in this cluster formalism, the only difference between the Ising spin glass and the ferromagnetic model is that, in the ferromagnetic case (2) the sum is over all the bond configurations, while in the spin glass (4) due to the geometrical constraint the sum is restricted to a subset of the bond configurations. In particular the ground state at $`T=0`$ ($`\mu =\mathrm{}`$) is obtained by maximizing the number of bonds, under the constraint that the bond configuration does not contain a frustrated loop.
Due to frustration, clusters defined in the spin glass model no longer correspond to thermodynamical fluctuations. In fact the correlations between spins can be either positive (if they propagate along a path that contains an even number of negative interactions) or negative (if the path contains an odd number of negative interactions), so they interphere and are canceled out at least in part . On the other hand connectivity is always positive and, as a result, clusters in the spin glass model percolate at a higher temperature $`T_p`$ respect to the critical temperature $`T_{SG}`$. In the three dimensional Ising spin glass simulations show that $`T_p3.95J/k_B`$ , while $`T_{SG}1.11J/k_B`$ .
As in the ferromagnetic case, the model defined by Eq. (4) can be extended to values $`q2`$ of the spin multiplicity. We call this model the “$`q`$-state frustrated percolation model”. For $`q`$ integer and even, Eq. (4) is the partition function of the model defined by the Hamiltonian
$$=J\underset{ij}{}\delta _{\sigma _i\sigma _j}(ϵ_{ij}S_iS_j+1),$$
(5)
where $`\sigma _i=1\mathrm{}q/2`$ are Potts spins, and $`\delta _{\sigma _i\sigma _j}`$ is the Kroneker’s delta. In analogy to the case $`q=2`$, the model is expected to have two transitions, one at a temperature $`T_{SG}(q)`$ corresponding to freezing of Ising spins, and the other at a temperature $`T_p(q)>T_{SG}(q)`$ corresponding to the percolation of the clusters. A renormalization group calculation carried over a hierarchical lattice has confirmed this expectation, and has shown that the transition at $`T_{SG}`$ should be in the universality class of the Ising spin glass, no matter what is the value of $`q`$, while the percolation transition at $`T_p`$ should be in the universality class of the $`q/2`$ state Potts model. For $`q=1`$ the model assume a very simple form, as the factor $`q^{N(C)}`$ disappears. The resulting model has been called “frustrated percolation”, and can be viewed as a simple model of percolation in a frustrated medium. Despite its simplicity, its dynamical properties exhibit a complex behavior, with features in common with both structural glasses and spin glasses.
In this paper, we perform a Monte Carlo study of the $`q`$-state frustrated percolation model on a two-dimensional square lattice, using a “bond flip” dynamics, in which bonds are added and removed from the lattice with appropriate probabilities, in order to satisfy the principle of detailed balance. To do this, we have realized an algorithm that allows to determine the connectedness of two given sites, and the presence of frustrated loops, with a time that in the worst case (at the percolation temperature) scales only with the logarithm of the lattice size, as we describe in Sect. II. The algorithm is a modification of the algorithm used by Sweeny in the ferromagnetic case . In Sect. III we report our results on the percolation transition of the model for $`q=1,2,4`$. Note that while for $`q`$ multiple of $`2`$ the percolation transition can be studied also by conventional spin flip , for other values of $`q`$ the “bond flip” dynamics is the only way to simulate the model. In Sect. IV and V we study the dynamical properties of the model, analyzing the relaxation functions of the number of bonds.
## II Monte Carlo algorithm
We have implemented a Monte Carlo algorithm to simulate the bond percolation model defined by Eq. (4) on a two-dimensional square lattice, which can be applied for any value of the parameter $`q[0,\mathrm{})`$. The interactions $`ϵ_{ij}`$ between pairs $`ij`$ of nearest neighbor sites are set at the beginning to a value $`+1`$ or $`1`$ randomly, with equal probability. These variables are quenched, and their state is not changed by the dynamics.
Each edge of the lattice, that is each pair of nearest neighbor sites, can be in two possible states: connected by a bond or not. At each step of the dynamics, we try to flip the state of an edge chosen randomly, with a probability determined in such a way to satisfy the principle of detailed balance. If we try to remove a bond from the system, or to add a bond that do not close a frustrated loop, then the probability of carrying out the “bond flip” will be
$$P_{\text{flip}}=\mathrm{min}(1,e^{\mu \delta b}q^{\delta N}),$$
(6)
where $`\delta b`$ is the change in the number of bonds, $`\delta N`$ is the change in the number of connected clusters. If we are trying to add a bond that closes a frustrated loop, then we have simply $`P_{\text{flip}}=0`$. A Monte Carlo Step (MCS) is defined as $`2V`$ single bond flip trials, where $`V=L^2`$ is the total number of sites, and $`2V`$ the total number of edges of the lattice.
The nontrivial point here is to determine the change in number of connected clusters, and to verify if a bond added between two given sites closes or not a frustrated loop. To do this, we have used the algorithm used by Sweeny in the ferromagnetic case , suitably modified to treat the frustration occurrence. Consider a two-dimensional square lattice, together with its dual lattice. If a bond is present on the lattice, then its dual bond is absent, and viceversa. The boundaries between connected clusters on the lattice and on its dual will form a collection of closed loops, as shown in Fig. 1(a). These loops are represented in the computer as chains of pointers. Each site on the lattice has four pointers adjacent to it, as shown in Fig. 1(b). At the beginning of the simulation, pointers are organized in a hierarchical way, by giving them a defined “level”, that do not change in the following. A fraction $`(4^n4^{(n+1)})`$ of the pointers are at level $`n=0,1,\mathrm{},n_{\text{max}}1`$, and $`4^{n_{\text{max}}}`$ at level $`n_{\text{max}}`$, where level $`n_{\text{max}}`$ is chosen so that it counts not more than four pointers. Chains then are formed by making each pointer point to other two pointers of the chain, one in the direction of the arrows, called the ALONG pointer, and one in the opposite direction, called the UP pointer. The ALONG pointer must be at least at the same level of the one pointing to it, and the UP pointer at a higher level (except if there are no higher level pointers in the chain), so they in general do not correspond to the nearest pointers in the chain.
When we add a bond to the lattice, (and remove its dual), two things can happen: the bond links two sites already belonging to the same connected cluster, and therefore $`\delta N=0`$; the bond links two previously disconnected clusters, and $`\delta N=1`$. In the first case the bond will cut a single chain into two distinct chains, see Fig. 2(a), while in the second case it will join two distinct chains into a single chain, see Fig. 2(b). When we remove a bond from the lattice, (and add its dual), it happens the other way round. Using this auxiliary data structure, one can determine if two given chain pointers belong to the same chain or not, and cut and rejoin chain segments, in a CPU time that grows only with the logarithm of the chain length.
To simulate the frustrated percolation model, we must also determine if a bond added to the lattice closes a frustrated loop or not. To do this, we must be able to count the number of antiferromagnetic bonds encountered along a path that joins two given sites A and B. This can be done if every chain pointer contains information about the number of antiferromagnetic bonds “skirted” when one traverses the chain to its UP pointer. Call this number the “phase” of the pointer respect to its UP pointer. We then go on jumping from each pointer to its UP pointer, and adding the relative phases, until we reach a reference pointer R in the chain. Then the number we seek for is found as the difference between the phases of two pointers adjacent to the sites A and B, respect to the reference pointer R, as shown in Fig. 3. This reference pointer is chosen between the highest level pointers in the chain, and is the UP pointer of all the pointers belonging to the highest level in the chain. When we cut and rejoin chain segments, we must update coherently the relative phases of the chain pointers involved, and assure that there is one and only one reference pointer per chain.
## III The percolation transition
We have studied the percolation transition of the model defined by Eq. (4), for $`q=1,2,4`$. For each value of $`q`$ we have simulated the model for lattice sizes $`L=32,64,128`$. The histogram method was used to analyze the data. In this Section, we use the probability $`p`$ as the independent variable. It is connected to the temperature via the simple relation $`p=1e^{2\beta J}`$. For each value of $`q`$ and $`L`$, sixteen probabilities were simulated around the percolation transition point, taking $`10^3`$ MCS for thermalization, and between $`10^4`$ MCS (for $`L=128`$) and $`10^5`$ MCS (for $`L=32`$) for the acquisition of the histograms. Histograms were taken of the number of bonds; of the mean cluster size, defined as $`{\displaystyle \frac{1}{V}}{\displaystyle \underset{s}{}}s^2n_s`$, where $`n_s`$ is the number of clusters of size $`s`$; and of the occurrence of a spanning cluster, defined as a cluster that spans from the bottom to the top of the lattice. These histograms were used to calculate the average number of bonds $`b`$, the fluctuation in the number of bonds $`b^2b^2`$, the average mean cluster size $`\chi `$, and the spanning probability $`P_{\mathrm{}}`$, in a whole interval of probabilities around the percolation transition probability.
The spanning probability $`P_{\mathrm{}}(p)`$ and the mean cluster size $`\chi (p)`$, as a function of the lattice size $`L`$ and of the probability $`p=1e^{2\beta J}`$, around the percolation probability $`p_c`$ should obey the scaling laws
$$P_{\mathrm{}}(p)\stackrel{~}{P}_{\mathrm{}}[L^{1/\nu }(pp_c)],$$
(8)
$$\chi (p)L^{\gamma /\nu }\stackrel{~}{\chi }[L^{1/\nu }(pp_c)],$$
(9)
where $`\gamma `$ and $`\nu `$ are the critical exponents of mean cluster size and connectivity length, $`\stackrel{~}{P}_{\mathrm{}}`$ and $`\stackrel{~}{\chi }`$ are universal functions. Given Eq. (8), the value of the transition probability $`p_c`$ can be evaluated from the point at which the curves $`P_{\mathrm{}}(p)`$ for different values of $`L`$ intersect. In Fig. 4(a) are plotted the measured curves $`P_{\mathrm{}}(p)`$ for $`q=1`$ and $`L=32,64,128`$. One must extrapolate the value at which curves for $`L,L^{}\mathrm{}`$ intersect. Then we have evaluated the critical exponent $`1/\nu `$, by choosing the value that gives the best data collapse of the curves, when one plots $`P_{\mathrm{}}(p)`$ in function of $`L^{1/\nu }(pp_c)`$, see Fig. 4(b). The errors on $`p_c`$ and $`1/\nu `$ were evaluated as the amplitudes of the intervals within which a good data collapse was obtained. Given Eq. (9), the mean cluster size $`\chi (p)`$ has a maximum that scales as $`L^{\gamma /\nu }`$, see Fig. 5(a). From a log-log plot of $`\chi ^{\text{max}}`$ in function of $`L`$, we can extract the exponent $`\gamma /\nu `$, making a linear fit, as shown in Fig. 5(b). Results are summarized in Tab. I. These results are in good agreement with the prediction that the percolation transition of the model falls in the universality class of the $`q/2`$-state ferromagnetic Potts model , whose critical exponents are reported in Tab. II .
Renormalization group calculations show that the model should also have a singularity in the free energy density $`F(p)`$ at $`p_c(q)`$, with a singular part $`F_{\text{sing}}(p)A(q)(pp_c)^{2\alpha (q/2)}`$, where $`\alpha (q/2)`$ is the specific heat exponent of the $`q/2`$-state ferromagnetic Potts model. In particular the model with $`q=4`$ should have a singularity corresponding to the ferromagnetic Ising model, that is a logarithmic divergence of the second derivative of the free energy (specific heat). From Eq. (4) it is possible to show, that the derivatives of $`\beta F=V^1\mathrm{log}Z`$ respect to $`\mu `$ are equal to the cumulants of the distribution of the number of bonds, divided by the total number of sites $`V`$. As $`\mu `$ is a regular function of the probability $`p`$ for $`0<p<1`$, we conclude that the model with $`q=4`$ should have a divergence in the second cumulant, that is the fluctuation, of the number of bonds divided by $`V`$. For finite size systems, we expect to see a peak whose maximum scales as $`\mathrm{log}(L)`$. In Fig. 6 we show our results for $`q=4`$ and $`L=32,64,128`$. It is evident that there is a divergence, but statistical errors do not allow to distinguish between logarithmic and a weak power law divergence.
For $`q=2`$ one would expect a singularity in the free energy, at the percolation transition, characterized by an exponent $`2\alpha (1)=8/3`$, that is a divergence in the third cumulant of the number of bonds. For the Ising spin glass ($`q=2`$) this would imply a divergence in the third cumulant of the distribution of the energy. However several arguments, including a renormalization group calculation , predict that the singularity could be canceled out by the vanishing of the prefactor $`A(q)`$ for $`q2`$. This prediction could be checked by showing that the third cumulant does not diverge, but much more extensive simulations are needed to verify this prediction.
## IV Onset of stretched exponentials
We have studied the dynamical properties of the $`q`$-state frustrated percolation model for $`q=1`$, 2, and 4, by calculating the autocorrelation function of the number of bonds. This is defined as
$$F(t)=\frac{b(0)b(t)b^2}{b^2b^2},$$
(10)
and is normalized so that $`F(0)=1`$, while for $`t\mathrm{}`$ it relaxes to zero. We simulated the model on a two-dimensional square lattice of size $`32\times 32`$, and took $`10^5`$ MCS for thermalization, and between $`10^5`$ and $`2\times 10^6`$ MCS for acquisition. All the functions were then averaged over 16 different configurations of the interactions $`ϵ_{ij}`$. Errors are evaluated as mean standard deviation of this last averaging.
In Fig. 7(a), 8(a), and 9(a), we show respectively the results for $`q=1`$, 2, and 4. For all the three values of the multiplicity $`q`$ considered, we observe the same behavior. For high temperatures, the functions show a single exponential decay, while at lower temperatures they show instead a two step decay, reminiscent of what one observes in glass forming liquids at low temperature. This behavior can be explained by the presence, at temperatures below the percolation transition, of a rough landscape of the free energy in configuration space, with many minima separated by high barriers. The short time decay corresponds to the relaxation inside the single valley, while the long time tail is due to the tunneling through barriers and final decay to equilibrium ($`\alpha `$ relaxation). Furthermore, the second step of relaxation functions is well fitted by a stretched exponential function
$$F(t)\mathrm{exp}\left(\frac{t}{\tau }\right)^\beta ,$$
(11)
where $`\beta `$ is an exponent lower than one.
In disordered and frustrated spin systems like spin glasses, simulated by conventional spin flip, the appearing of non exponential relaxation is believed to be caused by the existence of unfrustrated ferromagnetic clusters of interactions, see Randeria et al. . Below the ferromagnetic transition temperature $`T_c`$ of the pure model, each unfrustrated cluster relaxes with a time that depends from its size. Due to the disorder of the interactions, the sizes of the unfrustrated clusters are distributed in a wide range, giving rise to a wide distribution of relaxation times in the model. Therefore, according to this picture, the temperature $`T_c`$ of the ferromagnetic transition of the pure model marks the onset of non exponential relaxation.
The $`q`$-state frustrated percolation model is equivalent, for what concerns static properties, to a disordered spin system. In particular there are unfrustrated clusters of interactions with different sizes, due to the disorder of the variables $`ϵ_{ij}`$. On the other hand, the bond flip dynamics does not suffer from critical slowing down near the ferromagnetic critical point , so we expect that the Randeria mechanism does not apply in this case. We have verified this point plotting the stretching exponent $`\beta (T)`$ as a function of the temperature $`T`$, for $`q=1`$, 2, and 4, as shown in Fig. 7(b), 8(b), and 9(b), respectively. In the case of $`q=2`$ and 4, it is quite evident from the data that the transition point of the pure model $`T_c`$ does not mark any change in the behavior of relaxation functions. Instead the temperature $`T_p`$, at which clusters of bond percolate, appears as the point that marks the onset of stretched exponential relaxation. The case $`q=1`$ was not so evident from our results. For this reason we made a single simulation for a much larger system ($`L=100`$), at a temperature $`T=2.3`$ slightly higher than the percolation threshold ($`T_p=2.25`$). By fitting the function for times greater than $`1.5`$ MCS, where $`F(t)<0.05`$, we obtain a stretching exponent $`\beta =0.97`$, definitely higher than that obtained at the same temperature for $`L=32`$ ($`\beta =0.86`$). Thus it seems that for $`q=1`$ finite size effects at the percolation transition are more important, but nevertheless the relaxation is asyntotically purely exponential for $`T>T_p`$. In conclusion our results show that, for all the values of $`q`$ studied, in this model the relevant mechanism for the appearing of non exponential relaxation is the percolation transition.
## V Dynamical properties at low temperature
We have evaluated the relaxation functions, in the model with $`q=1`$ and 2, for very low temperatures. In Fig. 10(a) and (b) we show the results for $`L=40`$, $`q=1`$ and 2, and for different temperatures, averaged over 16 different configurations of interactions. For the lowest temperatures, the functions do not relax smoothly to zero. This is clearly an effect due to the relaxation time being greater than the total time of the run, that was between $`2\times 10^6`$ and $`5\times 10^6`$ MCS. For such very low temperatures, the relaxation functions show a behavior very similar to what can be observed in glass forming liquids near or below the mode coupling theory transition temperature . A first short time decay is followed by a very long plateau, and eventually there is a final relaxation to equilibrium, for very long times.
We compare this behavior to what one observes in the Ising spin glass simulated by conventional spin flip. In Fig. 11 the relaxation functions of the energy, for the Ising spin glass on a two-dimensional square lattice with $`L=40`$, and for the same temperatures of Fig. 10(b), are shown. Note that in this case there is not a clear separation between the first short time decay and the long time tail of the functions, and the functions do not show any plateau.
## VI Conclusions
We have studied the $`q`$-state frustrated percolation model, by means of a “bond flip” Monte Carlo dynamics. The model is equivalent from the thermodynamic point of view to the Potts spin glass model (5), which for $`q=2`$ coincides with the Ising spin glass. We have studied the percolation transition, that happens at a temperature $`T_p`$ greater than the spin glass temperature $`T_{SG}`$, and found that it belongs to the universality class of the $`q/2`$-state ferromagnetic Potts model.
We have then studied the dynamical properties of the model. Differently from what happens in spin glass systems, simulated by conventional spin flip, the transition temperature $`T_c`$ of the pure model does not play here any role in determining the dynamical behavior. Instead, the percolation temperature $`T_p`$ appears to mark the onset of two step decay, and stretched exponentials in autocorrelation functions.
At very low temperatures the autocorrelation functions develop a long plateau, as observed in glass forming liquids, and predicted by the mode coupling theory. This is a feature of the bond flip dynamics we have performed, while the spin flip dynamics of the spin glass model, that is thermodynamically equivalent to our model for $`q=2`$, is very different, and does not show any plateau.
This results show that the frustrated percolation model bridges the spin glass to the glass forming liquids. On one hand it is equivalent thermodynamically (for $`q=2`$) to the Ising spin glass. On the other hand the model describes a bond packing problem, where the bond are subjected to the constraint that no frustrated loop can be closed. This makes the dynamics similar to that of glass forming liquids, where irregular molecules (or groups of molecules) move under the constraint of some kind of steric hindrance.
To make more contact with glass forming liquids, the site version of the frustrated percolation model has been developed . In this model particles are allowed to diffuse on the lattice, under the constraint that no frustrated loop can be fully occupied. Consequently it is possible to calculate the mean square displacement and the diffusion coefficient of the particles. Numerically it is found that also these quantities reproduce qualitatively the corresponding quantities measured in glass forming liquids.
## Acknowledgments
This work was supported in part by the European TMR Network-Fractals c. n. FMRXCT980183, and by MURST (PRIN-97). Part of the Monte Carlo calculations were performed on the Cray T3D and T3E at CINECA.
|
no-problem/9812/quant-ph9812031.html
|
ar5iv
|
text
|
# Modified “delta kick cooling” for studies of atomic tunneling
## I Introduction
Advances in laser cooling and atom optics have led both to the ability to directly observe new physical effects (e.g., in connection with weakly interacting Bose gases) and to develop new, potentially applicable technologies (such as atom-wave gravimetry). There has been a great deal of interest in paving the way for future developments by studying atom optical components such as mirrors, lenses, and beam-splitters . Often, these components themselves rely on new and interesting physical effects, which is one of the explanations for the self-sustaining excitement in this field.
The long de Broglie wavelengths of ultracold atoms make them ideally suited for studies of quantum mechanics, as evidenced for example by the interference between two Bose condensates, or by tests of quantum chaos using cold atoms. We plan to make use of the wave nature of laser-cooled Rubidium atoms to study the tunneling of atoms through optical dipole-force barriers; by focusing light into a thin sheet, one can create strong repulsive potentials for atoms, with spatial widths several times the optical wavelength. By cooling atoms to near or below the recoil temperature (where the atomic de Broglie wavelength is equal to the optical wavelength), we can enter a regime where there is significant tunneling probability. Although tunneling has been observed indirectly in several atom-optics experiments, this would be to our knowledge the first experiment where spatially resolvable tunneling is observed. That is, we plan to directly image the incident and transmitted atomic clouds, using the internal degrees of freedom of the atoms to address questions about the “history” of transmitted particles. This unique system should make it possible to answer long-controversial questions related to tunneling and to quantum measurement theory. As a simple example, the question of what you see if you attempt to image atoms while they are within a forbidden region is already nontrivial.
In addition to the intrinsic interest of atomic tunneling, focussed dipole-force barriers may prove useful as coherent beam-splitters, velocity-selective elements, and in related roles. In this paper, we present the status of our experimental project to observe atomic tunneling, including some simulations of planned velocity-selection experiments. As a first step towards this experiment, we have used time-dependent magnetic forces to cool atoms in one dimension, in a scheme based loosely on Ammann and Christensen’s “delta-kick cooling” proposal. We have achieved temperatures below 700 nK, and much lower temperatures are possible in principle. The technique is also generalizable to three dimensions. We also see some interesting effects when these magnetic kicks are applied to atoms from an optical molasses. We are still working towards a full understanding of these effects, but they appear to contain the signature of spatial spin correlations within the atom cloud. We believe that in addition to the usefulness of magnetic kicks for cooling, this Stern-Gerlach-like system will prove to be a novel probe of the atomic spin in laser-cooled atom clouds.
As alluded to above, the practical observation of atomic tunneling will require atoms with de Broglie wavelengths on the order of the width of the tunnel barrier. Our barrier is produced by using the dipole force from a 500 mW laser focused into a sheet of light which is about 10 $`\mu `$m wide and 2 mm tall. The atoms leaving our laser cooling and trapping system are at temperatures of 6 $`\mu `$K, implying a wavelength of about 0.2 $`\mu `$m, too short to observe any significant tunneling through an optical-scale barrier. Therefore we need to further cool our atoms to sub-recoil temperatures. Our first stage of cooling following optical molasses is to perform an improved variant of the “delta kick cooling” proposed by Ammann and Christensen, also similar to an independent proposal by Chu et al. . Following this stage, we will have atoms with 1-D temperatures on the order of the recoil temperature of Rb and de Broglie wavelengths on the order of the optical wavelength of 780 nm. We shall perform a final step of velocity selection to select only the least energetic atoms and separate them from the remaining higher-energy atoms. By using 1-D velocity selection we can retain about 7% of our atoms at 1/200th of the initial temperature. Given the fact that we are concerned with 1D temperatures this is more efficient than comparable evaporative cooling in 3D. Moreover, the velocity selection will suppress the thermal tails which could otherwise lead to ambiguity between tunneling and thermal activation.
In the original proposal of “delta-kick cooling” Ammann and Christensen suggested a form of cooling in which an atomic wavepacket prepared in a minimum-uncertainty state within an optical lattice is first allowed to freely expand for a short period to allow position to become correlated with momentum. Application of a position-dependent force from a standing wave can then be used to reduce the mean momentum of the atoms. Near the bottom of the potential well the atoms essentially experience a harmonic potential. As the atoms expand beyond this harmonic region the proposed cooling process breaks down. Instead of a sinusoidal potential, we consider a true harmonic potential, easily generated with magnetic field coils. This variant is not only easier to understand and to implement; it is also immune to the original proposal’s cooling limit.
Consider an ensemble of laser-cooled atoms initially confined within a small region of size $`r_i`$ with mean thermal velocity $`v_o`$. At time $`t=0`$ the trap is turned off and atoms are allowed to freely expand away from the trap center. After a time $`t_\mathrm{f}`$, long compared to $`r_i/v_o`$ the atoms are located at positions essentially given by $`x_a=v_at_\mathrm{f}`$, where $`v_a`$ is the velocity of a particular atom. Application of a harmonic potential $`U=\frac{1}{2}m\omega ^2x^2`$ for a short time will apply an impulse to the atoms proportional to position $`\mathrm{\Delta }\stackrel{}{p}=m\omega ^2\stackrel{}{x}t_\mathrm{k}`$ where $`t_\mathrm{k}`$ is the duration of the kick. If this impulse is chosen to be equal to $`mx_a/t_fmv_a`$, all atoms will essentially be brought to rest. The condition for such an optimal kick is $`t_\mathrm{f}t_\mathrm{k}=1/\omega ^2`$.
In reality several factors place limits on the achievable temperature. At time $`t=0`$ the atoms are not all located at the center of the trap, but instead have some distribution of initial positions. As the ratio between the final size to initial size increases, the correlation between position and momentum improves, allowing greater degrees of cooling. At any finite time, the correlations will be imperfect, preventing the cooling from being optimal. Practical considerations may limit the time for which the atoms can be allowed to expand.
After free expansion, the typical atom is at a position on the order of $`r_f=\sqrt{r_0^2+(v_0t)^2}`$, and has a velocity on the order of $`v_0`$. Each individual velocity class has a spread in position given by the initial size $`r_0`$ of the cloud. A transferred impulse proportional to distance, and designed to cancel out the mean velocity of $`v_0`$ at a typical distance of $`r_f`$, will therefore transfer a random velocity on the order of $`v_0r_0/r_f`$ to each velocity group of atoms. Indeed, a more careful phase-space treatment shows clearly that the rms velocity of the kicked atoms decreases by a factor of $`r_0/r_f`$, leading to a temperature reduction of $`(r_0/r_f)^2`$.
In other words, this technique is the moral equivalent of adiabatic expansion. At a practical level, however, it has the advantage that there is no adiabatic criterion to meet. To cool a cloud by a factor of $`N`$ would require a time of $`\sqrt{N}\tau _0`$ where $`\tau _0`$ is a secular period in the harmonic oscillator. A similar adiabatic expansion would need to be accomplished slowly relative to the instantaneous secular period($`N\tau _0`$ by the end of the expansion).
The cooling process can be readily understood in a phase-space picture. Figure 1 shows the evolution of an atomic cloud in phase space from the initial state at the start of expansion, to after free expansion and finally, the distribution after application of a kick. We start with a phase-space distribution characterized by the widths of the distribution in momentum and position. Free expansion stretches the distribution in position, but has no effect on momentum. The effect of a harmonic kick is to rotate the distribution in phase space. If the duration of the kick is chosen correctly, this rotates the distribution back onto the position axis, thereby lowering the temperature. Note that this does not require a true “delta-kick” in the sense of a very short duration. By Liouville’s theorem, the area in phase space is conserved both during free expansion and during the kick, yielding $`v_f=v_0x_0/x_f`$. The longer the cloud is allowed to undergo free expansion before the kick, the narrower the final distribution is in momentum, resulting in lower temperatures.
In addition to cooling via harmonic potentials, it is also possible to cool atoms using a pulsed quadrupole potential. Although this technique cannot match the ideal harmonic kicks, it is substantially easier to generate strong field gradients than higher-derivative terms. In 1D, a quadrupole field exerts a fixed impulse toward the center of the potential. If the atoms have been allowed to expand significantly, most are moving away from the center, and a kick chosen to be equal to the mean thermal velocity concentrates the atoms near v=0 (in a highly non-thermal, nearly flat-topped distribution). In 1D, the mean kinetic energy may be reduced by up to a factor of 6 in this way. Simulations for a true 3D quadrupole potential show even greater cooling(predominantly along the symmetry axis), due to the “rounding” of the potential energy surface when transverse excursions are taken into account. The cooling is predominantly along the quadrupole axis, but some cooling does occur in the other directions as well.
The data presented in this paper all concern one-dimensional cooling. However, the harmonic kicks are easily generalizable to 3D by application of successive kicks along the three Cartesian axes. With quadrupole kicks, perfect spherical symmetry cannot be achieved, but may be approximated using a similar approach.
## II Experiment
Our MOT coils are located horizontally around the cuvette and define the z-axis of the trap. The coils are made of 200 turns of wire with a radius of 4 cm and separated by 8 cm. This is twice the separation required for a Helmholtz configuration, leading to a nonvanishing $`d^2B/dz^2`$. With the currents in opposing directions as used in a MOT or a quadrupole kick, the coils can produce gradients up to 180 G/cm given our present maximum current of 18 A. These same coils are also used for generation of a harmonic potential. By reversing the direction of current in one of the coils we can achieve harmonic fields with a second derivative of about 60 G/cm<sup>2</sup>, corresponding to a trap frequency of about 60 Hz. The coils can be switched on and off in 200 $`\mu `$s, residual fields falling to 1% in 1 ms.
We start by cooling and trapping <sup>85</sup>Rb atoms within our MOT using a field gradient of 20 G/cm, and 40 mW of total power in trapping beams with 2 cm diameter. The trapping beam is detuned 10 MHz to the red of the D2 ($`F=3F=4`$) transition at 780 nm. We cool and trap about $`10^8`$ atoms in our MOT with a diameter of 0.5 mm. We further cool the atoms to 6 $`\mu `$K in 1 millisecond of $`\sigma ^+\sigma ^{}`$ optical molasses detuned $`34`$ MHz to the red of the $`F=3F=4`$ transition.
We then turn off all light and magnetic fields to allow the atoms to undergo free expansion. After a time of 9 to 15 ms we apply a short (3 ms) pulse of the quadrupole field. The resulting force is directed towards the origin, mostly along the symmetry axis of the coils, and applies an impulse to the atoms opposing their direction of motion.
The temperature of the resulting cloud is determined by time-of-flight imaging. A series of images is taken as a function of time after the kick, and the temperature is extracted from the expansion curve.
The existence of multiple spin levels in atoms released from an optical molasses implies that different atoms experience different magnetic potentials, making distributions difficult to interpret. For this reason we select the atoms in a doubly polarized state $`F=3,m_F=3`$ by capturing the atoms within a weak magnetic trap which is unable to hold atoms with $`m_F<3`$ against gravity. We release the doubly polarized atoms from the magnetic trap after 200 ms and allow the $`m_F=3`$ atoms to expand freely before application of the quadrupole magnetic field.
Figure 2 shows fluorescence images of the atoms under free expansion and after application of a delta kick. The atoms released from the magnetic trap have a temperature of 7.5 $`\mu `$K with an rms velocity of 2.7 cm/s. The atoms are first allowed to undergo free expansion for a time of 11 ms and then have a kick applied for 3 ms, imparting a change in velocity of 3 cm/s. After application of the kick, the temperature of the cloud along the coil axis has decreased by a factor of about 6, from 7.5 $`\mu `$K down to 1.2 $`\mu `$K. Expansion curves for a cloud for several kick strengths are observed in Figure 3. For a kick of 2.4 cm/s, near the original rms velocity of the atoms, we observe optimal cooling. Over 20 ms, essentially no expansion of the atom cloud is seen, and fits indicate a temperature below 700 nK. For stronger kicks, the atoms are impelled back to the center of the cloud, where they come to a focus. As the strength of the kick increases the atoms come to a smaller focus (consequently heating the atom cloud), until the strongest kick strength results in a cloud hotter than the cloud from the original magnetic trap.
At short times such that the radius of the cloud has not increased by very much, the positions and momenta of the atoms are not very correlated and the kick does little to cool the atoms. In Figure 4 we show results (solid circles) and simulations (smooth curve) on the cooling ratio versus the ratio of final size to inital size.
The temperatures obtained experimentally were lower than the initial temperature by as much as a factor of 10, even at times when the simulations predicted only a factor of 5. Additionally, Figure 5 suggests that optimal cooling occurs for a somewhat stronger kick strength than originally expected theoretically. Both these effects can be qualitatively understood by considering the effect of gravity. When the atoms are allowed to freely expand for a long time, they also begin to fall under the influence of gravity. If the cloud falls in the conical potential by a distance larger than the spatial extent of the cloud, then the horizontal potential seen by the atoms looks like a conic section, i.e. a parabola. Thus transverse cooling is accomplished by a harmonic kick, much more efficient than the quadrupole, yet much stronger than the true harmonic potential achievable by reversing the current in one of our coils.
When we include gravity into our simulations, both the extra cooling and increased potential strength requirements are observed. Figure 6 shows the achievable temperature with the inclusion of gravity. As the atoms move down from the center of the potential, the strength of the horizontal force decreases. To optimize cooling requires the use of a stronger potential than expected for a 1D quadrupole. Simulations run with various parameters have shown cooling by as much as a factor of 30.
Kicks have also been performed upon atoms coming out of an optical molasses. Since there are 7 different spin states in the F=3 ground state of Rb, we expect the different states to undergo different amounts of cooling or heating. We therefore expect to see a multimodal distribution, in which from the initial cloud radius, each spin component expands at a different rate. Figure 7 shows the results of a kick on an optical molasses. The upper set of expansion images shown free expansion and the lower set displays the atom cloud after a kick. After the kick, the cloud separates as expected into a bi-modal distribution consisting of a very cold central stripe and a broader, hotter background. The unusual feature comes in when we realize that the central stripe is significantly smaller than the size of the atom cloud at the time of kick as seen clearly in Figure 8. We believe this may indicate a pre-existing correlation between position in the optical molasses and spin state. We are continuing to study these effects, for which the Stern-Gerlach-like analysis possible using pulsed field gradients appears to be a very promising new probe.
Through the use of these kicks we have achieved temperatures as low as 700 nK both on atoms released from a magnetic trap and on a subset of the atoms released from an optical molasses. Cooling ratios as large as a factor of 10 have been observed. The effect of gravity may be used to enhance cooling, but also limits this technique for certain applications. Since the atoms begin to fall due to gravity, they are no longer located at the origin of the potential, and cannot be recaptured without heating. The gain in cooling may still be useful for experiments in which the atoms are allowed to fall for some distance, such as atomic mirrors, lensing and atom deposition systems. Along these lines Maréchal et. al. have studied a system where inhomogeneous magnetic potentials separate the spin components of a cloud and provide cooling as the cloud falls a distance of about a meter. The addition of a 1D optical lattice or anti-gravity field will allow us to reach greater ratios in the size of the atoms without having the atoms fall from the center of the trap, to obtain cooling greater than a factor of 10.
## III Tunneling and Velocity Selection
Having achieved recoil-velocity atoms we are moving on to study properties of atoms while they are tunneling. We shall use a line-focused beam of intense light detuned far to the blue of the D2 line to create a dipole-force potential for the atoms. Using a 500 mW laser at 770 nm focused into a sheet 10 $`\mu `$m wide and 2 mm tall, we will be able to make repulsive potentials with maxima as large as 50 $`\mu `$K, without significant scattering rates. The potential will be modulated using an acousto-optic deflector allowing us to rapidly shift the focus of the beam. Since we may shift the focus of the beam more quickly than the atomic center-of-mass motion can follow, the atoms see a time-averaged potential. This will allows us to make nearly arbitrary potential profiles.
Following the application of a delta-kick to obtain sub-recoil 1-D velocities, our atoms will have a thermal de Broglie wavelength of about 1 micron, still too small to observe significant tunneling through a 10 $`\mu `$m dipole-force barrier. We will therefore follow the delta kick with a velocity-selection phase. We will use the same beam which is to form the tunnel barrier but dither the focus of the laser quickly. The atoms shall see a time-averaged potential many times larger than the de Broglie wavelength of the atoms, thereby appearing as an essentialy classical barrier. By sliding this “classical” barrier through our atomic cloud, it will be possible to adiabatically sweep the lowest energy atoms away from the center of the magnetic trap, leaving the hot atoms behind. The coldest atoms will thus be in a spatially separated local minimum from which tunneling may be observed when the width of the barrier is decreased. Figure 9 shows the results of quantum-mechanical simulations for atoms at an initial temperature of 1.3 $`\mu `$K, trapped in a 5 G/cm field (300 $`\mu `$ K/cm). We superpose a 20 $`\mu `$m Gaussian beam with a peak height of 600 nK onto the V-shaped potential. This creates a potential minimum of 16 nK which is swept through the magnetic trap at a rate of 0.5 mm/s. The minimum supports a quasi-bound state with energy 5 nK, and we see 7% of the atoms transferred into this minimum. Classically, we would expect the number of atoms to be transferred equivalent to $`\sqrt{T_f/T_i}`$. In Figure 10 we observe steps in the probability of transfer as a function of the barrier height, indicating the number of quasi-bound states supported.
After velocity-selection we will narrow the beam to 10 $`\mu `$m to allow the cold atoms to tunnel. The velocity-selected sub-sample of atoms will have a de Broglie wavelength of 6 $`\mu `$m, leading to a significant tunneling probability through a 10-micron barrier. We expect rates of order 5% per secular period, causing the auxiliary trap to decay via tunneling on a timescale of the order of 600 ms.
Progressing towards this tunneling experiment we have achieved temperatures of 700 nK for atoms released both from magnetic traps and optical molasses by application of a quadrupole delta-kick. Given longer expansion times, this temperature should drop even further. Once these cooling and velocity selection techniques are perfected, we shall have a unique system in which to study tunneling. By using optical probes such as absorption, optical pumping and stimulated Raman transitions, we will be able to go beyond studies of simple wavepacket tunneling to investigate the interactions of tunneling atoms while in the forbidden region. Studies of measurement of the tunneling time are planned as well as investigations into quantum-mechanical nonlocality.
|
no-problem/9812/math9812073.html
|
ar5iv
|
text
|
# An Infinite Antichain of Permutations
## 1 Introduction
When considering a partially ordered set with infinitely many elements, one should wonder whether it contains an infinite antichain (that is, a subset in which each pair of elements are incomparable). It is well known that all antichains of $`N^k`$ (where $`(x_1,x_2,\mathrm{},x_k)(y_1,y_2,\mathrm{},y_k)`$ if and only if $`x_iy_i`$ for $`1ik`$ ) are finite. (See ). Another basic result is that all antichains of the partially ordered set of the finite words of a finite alphabet are finite, where $`x<y`$ if one can delete some letters from $`y`$ to get $`x`$. (This result is due to Higman and can be found in ).
In this paper we examine this question for the partially ordered set $`P`$ of finite permutations with the following $`<`$ relation: if $`m`$ is less than $`n`$, and $`p_1`$ is a permutation of the set $`\{1,2,\mathrm{},m\}`$ and $`p_2`$ is a permutation of the set $`\{1,2,\mathrm{},n\}`$, then $`p_1<p_2`$ if and only if we can delete $`nm`$ elements from $`p_2`$ so that when we re-name the remaining elements according to their rank, we obtain $`p_1`$. For example, $`\mathrm{1\hspace{0.25em}3\hspace{0.25em}2}<\mathrm{2\hspace{0.25em}4\hspace{0.25em}5\hspace{0.25em}3\hspace{0.25em}1}`$ as we can delete 4 and 1 from the latter to get $`\mathrm{2\hspace{0.25em}5\hspace{0.25em}3}`$, which becomes $`\mathrm{1\hspace{0.25em}3\hspace{0.25em}2}`$ after re-naming. Another way to view this relation is that $`p_1<p_2`$ if there are $`nm`$ elements of $`p_2`$ that we can delete so that the $`i`$-th smallest of the remaining elements preceeds exactly $`b_i`$ elements, where $`b_i`$ is the number of elements preceeded by $`i`$ in $`p_1`$. In other words, the $`i`$-th smallest remaining entry of $`p_2`$ preceeds the $`j`$-th one if and only if $`i`$ preceeds $`j`$ in $`p_1`$. In short, $`p_1<p_2`$ if $`p_1`$ is “contained” in $`p_2`$, that is, there is a subsequence in $`p_2`$ in which any two entries relate to each other as the corresponding entries in $`p_1`$.
We would like to point out that any answer to this question would be somewhat surprising. If there were no infinite antichains in this partially ordered set, that would be surprising because, unlike the two partially ordered sets we mentioned in the first paragraph, $`P`$ is defined over an infinite alphabet and the “size” of its elements can be arbitrarily large. On the other hand, if there is an infinite antichain, and we will find one, then it shows that this poset is more complex in this sense than the poset of graphs ordered by the operations of edge contraction and vertex deletion. (That this poset of graphs does not contain an infinite antichain is a famous theorem of Robertson and Seymour ). This is surprising too, as graphs are usually much more complex than permutations.
## 2 The infinite antichain
We are going to construct an infinite antichain, $`\{a_i\}`$. The elements of this antichain will be very much alike; in fact, they will be identical at the beginning and at the end. Their middle parts will be very similar, too. These properties will help ensure that no element is contained in another one.
Let $`a_1=13,12,10,14,8,11,6,9,4,7,3,2,1,5`$. We view $`a_1`$ as having three parts: a decreasing sequence of length three at its beginning, a long alternating permutation starting with the maximal element of the permutation and ending with the entry 7 at the fifth position from the right (In this alternating part odd entries have only even neighbors and vice cersa. Moreover, the odd entries and the even entries form two decreasing subsequences so that $`2i`$ is between $`2i+5`$ and $`2i+3`$), and a terminating subsequence 3 2 1 5.
To get $`a_{i+1}`$ from $`a_i`$, simply insert two consecutive elements right after the maximum element $`m`$ of $`a_i`$, and give them the values $`(m4)`$ and $`(m1)`$. Then make the necessary corrections to the rest of the elements, that is, increment all old entries on the left of $`m`$ ($`m`$ included) by two and leave the rest unchanged (see Figure 1).
Thus the structure of any $`a_{i+1}`$ is very similar to that of $`a_i`$—only the middle part becomes two entries longer.
We claim that the $`a_i`$ form an infinite antichain. Assume by way of contradiction that there are indices $`i,j`$ so that $`a_i<a_j`$. How could that possibly happen? First, note that the rightmost element of $`a_j`$ must map to the rightmost element of $`a_i`$, since this is the only element in $`a_j`$ preceeded by four elements less than itself. Similarly, the maximal element of $`a_j`$ must map to the maximal element of $`a_i`$, since, excluding the rightmost element, this is the only element preceeded by three smaller elements. This implies that the first four and the last six elements of $`a_j`$ must be mapped to the first four and last six elements of $`a_i`$, thus none of them can be deleted.
Therefore, when deleting elements of $`a_j`$ in order to get $`a_i`$, we can only delete elements from the middle part, $`M_j`$. We have already seen that the maximum element cannot be deleted. Suppose we can delete a set $`D`$ of entries from $`M_j`$ so that the remaining pattern is $`a_i`$. First note that $`D`$ cannot contain three consecutive elements, otherwise every element before those three elements would be larger than every element after them, and $`a_i`$ cannot be divided in two parts with this property. Similarly, $`D`$ cannot contain two consecutive elements in which the first is even. Thus $`D`$ can only consist of separate single elements (elements whose neighbors are not in $`D`$) and consecutive pairs in which the first element is odd. Clearly, $`D`$ cannot contain a separate single element as in that case the middle part of resulting permutation would contain a decreasing 3-subsequnce, but the middle part, $`M_i`$, of $`a_i`$ does not. On the other hand, if $`D`$ contained two consecutive elements $`x`$ and $`y`$ so that $`x`$ is odd, then the odd element $`z`$ on the right of $`y`$ would not be in $`D`$ as we cannot have three consecutive elements in $`D`$, therefore $`z`$ would be in the remaining copy of $`a_i`$ and $`z`$ wouldn’t be preceeded by two entries smaller than itself. This is a contradiction as all odd entries of $`M_i`$ have this property.
This shows that $`D`$ is necessarily empty, thus we cannot delete any elements from $`a_j`$ to obtain some $`a_i`$ where $`i<j`$.
We have shown that no two elements in $`\{a_i\}`$ are comparable, so $`\{a_i\}`$ is an infinite antichain. $`\mathrm{}`$
|
no-problem/9812/hep-ph9812379.html
|
ar5iv
|
text
|
# References
Does Weak CP Phase Originate from a Certain Geometry ?
As is well known, CP violation is one of the most important problem in particle physics \[1-6\]. People have been concerning it for more than thirty years. However, we still know little about it. With the runing of the $`B`$factories, we expect to know more about it in near future.
According to the standard CKM mechanism, CP violation originates from a phase presenting in the three by three quark mixing matrix \[7-8\]. Last winter, we have found that, the weak CP phase and the other three mixing angles in CKM matrix satisfy a certain geometry relation \[9-10\], so we postulated that, weak CP phase is a geometry phase. Although it is a ad hoc postulation, all the conclusions extracted from it are consistent with the present experimental results.
Our postulation in the standard parametrization can be expressed as
$$\mathrm{sin}\delta _{13}=\frac{(1+s_{12}+s_{23}+s_{13})\sqrt{1s_{12}^2s_{23}^2s_{13}^2+2s_{12}s_{23}s_{13}}}{(1+s_{12})(1+s_{23})(1+s_{13})}$$
(1)
where $`s_{ij}`$ and $`\delta _{13}`$ are the parameters in the standard parametrization \[11-12\]
$$V_{KM}=\left(\begin{array}{ccc}c_{12}c_{13}& s_{12}c_{13}& s_{13}e^{i\delta _{13}}\\ s_{12}c_{23}c_{12}s_{23}s_{13}e^{i\delta _{13}}& c_{12}c_{23}s_{12}s_{23}s_{13}e^{i\delta _{13}}& s_{23}c_{13}\\ s_{12}s_{23}c_{12}c_{23}s_{13}e^{i\delta _{13}}& c_{12}s_{23}s_{12}c_{23}s_{13}e^{i\delta _{13}}& c_{23}c_{13}\end{array}\right)$$
(2)
with $`c_{ij}=\mathrm{cos}\theta _{ij}`$ and $`s_{ij}=\mathrm{sin}\theta _{ij}`$ for the ”generation” labels $`i,j=1,2,3`$. Here, the real angles $`\theta _{12},\theta _{23}`$ and $`\theta _{13}`$ can all be made to lie in the first quadrant. The phase $`\delta _{13}`$ lies in the range $`0<\delta _{13}<2\pi `$. In following, we will make the three angles $`\theta _{ij}`$ lie in the first quadrant.
The central purpose of this work is to further investigate our postulation carefully. Firstly, we investigate the whole CKM matrix, secondly, we investigate the three angles in unitarity triangle $`\mathrm{𝐃𝐁}`$, both with our postulation being used. In fact, if we notice that, the moduli of the elements in CKM matrix are squared moduli invariants, and the three angles in unitarity triangle are composed of squared and quartic invariants , we can see that is meaningful to do so.
1. About the Moduli of the Elements in CKM Matrix
We investigate the whole CKM matrix firstly, we hope to extract some useful information from our postulation. The programme is
$`a`$. For each group of given $`V_{ud},V_{ub}`$ and $`V_{tb}`$, solve $`s_{12},s_{23},s_{13}`$ from the following equation
$$V_{ud}=c_{12}c_{13}V_{ub}=s_{13}V_{tb}=c_{23}c_{13}.$$
(3)
$`b`$. Substituting Eq.(1) into CKM matrix Eq.(2). Then, solve the moduli of all the elements.
$`c`$. Let $`V_{ud},V_{ub}`$ and $`V_{tb}`$ vary in certain ranges. Repeat the steps $`a`$ and $`b`$.
When we let $`V_{ud},V_{ub}`$ and $`V_{tb}`$ vary in the ranges
$$0.9745V_{ud}\mathrm{0.9760\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0.0018}V_{ub}\mathrm{0.0045\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0.9991}V_{tb}0.9993$$
(4)
$$0.9745V_{ud}\mathrm{0.9760\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0.00045}V_{ub}\mathrm{0.00585\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0.9991}V_{tb}0.9993$$
(5)
and
$$0.97375V_{ud}\mathrm{0.97675\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0.00045}V_{ub}\mathrm{0.00585\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0.99895}V_{tb}0.99955$$
(6)
we get the magnitudes of the elements of the complete matrix as
$$\left(\begin{array}{ccc}0.97450.9760& 0.21820.2244& 0.00180.0045\\ 0.21800.2242& 0.97360.9752& 0.03710.0424\\ 0.00790.0093& 0.03650.0415& 0.99910.9993\end{array}\right)$$
(7)
$$\left(\begin{array}{ccc}0.97450.9760& 0.21820.2244& 0.000450.00585\\ 0.21800.2242& 0.97360.9757& 0.02960.0424\\ 0.00790.0094& 0.03640.0415& 0.99910.9993\end{array}\right)$$
(8)
and
$$\left(\begin{array}{ccc}0.973750.97675& 0.21430.2276& 0.000450.00585\\ 0.21420.2275& 0.97270.9762& 0.02960.0458\\ 0.00720.0102& 0.03370.0448& 0.998950.99955\end{array}\right)$$
(9)
respectively.
Compare with that given by
$$\left(\begin{array}{ccc}0.97450.9760& 0.2170.224& 0.00180.0045\\ 0.2170.224& 0.97370.9753& 0.0360.042\\ 0.0040.013& 0.0350.042& 0.99910.9993\end{array}\right)$$
(10)
we find that, the predicted results are well in agreement with that given by data book. In the meantime, $`|V_{td}|`$ is not sensitive to the variations of the inputs. It lies in a very narrow window with the central value about ($`0.0086`$), even if we take a little more large error ranges for the inputs.
On the other hand, the relevant result extracted from the experiment on $`B_d^0\overline{B_d^0}`$ mixing is
$$|V_{tb}^{}V_{td}|=0.0084\pm 0.0018.$$
(11)
we find that, the prediction about $`|V_{td}|`$ based on our postulation, not only coincide with the experimental result very well, but also gives a more strict constraint.
2. About the Three Angles of the Unitarity Triangle
The three angles $`\alpha ,\beta `$ and $`\gamma `$ in the unitarity triangle $`\mathrm{𝐃𝐁}`$ defined as
$$\alpha arg(\frac{V_{td}V_{tb}^{}}{V_{ud}V_{ub}^{}})\beta arg(\frac{V_{cd}V_{cb}^{}}{V_{td}V_{tb}^{}})\gamma arg(\frac{V_{ud}V_{ub}^{}}{V_{cd}V_{cb}^{}})$$
(12)
If a small change being made, it is easy to see that, the definited angles are composed of squared and quartic invariants. For example,
$$\alpha arg(\frac{V_{td}V_{tb}^{}}{V_{ud}V_{ub}^{}})=arg(\frac{V_{td}V_{ud}^{}V_{tb}^{}V_{ub}}{V_{ud}V_{ud}^{}V_{ub}V_{ub}^{}})$$
the numerator in the definition is a quartic invariant and the denominator is a product of two squared invariants.
Now, we investigate the three angles. The programme is similar to that in section 1.
$`a`$. For each group of given $`V_{ud},V_{ub}`$ and $`V_{tb}`$, solve $`s_{12},s_{23},s_{13}`$ from Eq.(3).
$`b`$. Substituting Eq.(1) into CKM matrix Eq.(2). Then, solve all of the elements with the results of $`a`$ being used.
$`c`$. Solve $`\alpha ,\beta `$ and $`\gamma `$ according to the definition Eq.(12).
$`d`$. Let $`V_{ud},V_{ub}`$ and $`V_{tb}`$ vary in certain ranges. Repeat the steps $`a`$, $`b`$ and $`c`$.
We still let $`V_{ud},V_{ub}`$ and $`V_{tb}`$ vary in the ranges given by Eq.(4), Eq.(5) and Eq.(6). The corresponding results are
$$73.3^0\alpha 94.4^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}10.6}^0\beta 31.3^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}74.9}^0\gamma 75.6^0$$
(13)
$$70.3^0\alpha 98.9^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}6.1}^0\beta 34.2^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}74.8}^0\gamma 75.6^0$$
(14)
and
$$66.9^0\alpha 99.8^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}5.6}^0\beta 34.2^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}74.5}^0\gamma 76.0^0$$
(15)
respectively.
The recent analysis by Buras gives
$$35^0\alpha 115^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}11}^0\beta 27^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}41}^0\gamma 134^0$$
(16)
or more strictly
$$70^0\alpha 93^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}19}^0\beta 22^0\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}65}^0\gamma 90^0$$
(17)
It is easy to find, similar to that on $`V_{td}`$ in above, starting from our postulation, we obtain a more strict constraint on $`\gamma `$. We predict a very narrow window for $`\gamma `$ with the central value about ($`75.3^0`$). Furthermore, all the predictions about $`\alpha ,\beta `$ and $`\gamma `$ coincide with the relevant analysis .
3. Conclusions and Discussions
In conclusion, we have further investigated the postulation that, weak CP phase originates in a certain geometry. Based on this postulation, we obtian two strict constraints on the magnitude of CKM matrix element $`V_{td}`$ and angle $`\gamma `$ respectively. These can be put to the more precisely test on our postulation in $`B`$factory in near future.
The significance of Eq.(1) is evident. If it can be further verified in the future, we can at least remove a uncertainty coming from the weak interaction. Through the study on heavy flavors, with Eq.(1) being considered, we can then extract more imformations on strong interaction. For instance, it becomes possible to extract the CP phase related to the part of strong interaction of final states. The further work along this direction is under way.
|
no-problem/9812/astro-ph9812057.html
|
ar5iv
|
text
|
# 1 Three Scenarios for Bulge Formation
## 1 Three Scenarios for Bulge Formation
We consider the following possibilities for bulge formation: bulges form before, contemporaneously with, or after disks.
Bulges are old: Monolithic collapse (Eggen, Lynden-Bell and Sandage 1962) described the formation of population II prior to disk formation. Evidence on the age of the inner Milky Way bulge stars generally supports an old population that formed before the disk. Galaxies with massive bulges would have necessarily formed by primordial collapse, major mergers at high redshifts, or infall of satellite galaxies (Pfenniger 1992).
The chemical evidence is less clear: the predictions of a simple monolithic formation model cannot be easily reconciled with the observed abundance spread in bulge stars. Indeed theory has largely supplanted a monolithic collapse picture with a clumpy collapse model. The theory of galaxy formation has had considerable success in predicting various properties of large-scale structure. Hierarchical formation is closer in spirit to the Searle-Zinn view of halo and bulge formation in which many globular cluster mass clumps merge together. We regard bulge and luminous halo formation as closely related phenomena, the bulge simply being the core of the field star halo. Hierarchical galaxy formation involves a sequence of successive mergers of larger and larger subgalactic scale clumps. At any given stage, gas dissipation and settling produces disks that are destroyed in subsequent mergers. Not all disk structure is erased: dwarf satellites and even globular clusters may be substructure relics. The disk only forms after the last massive merger via gas infall. In environments such as rich clusters disk infall is largely suppressed and the cluster cores are dominated by spheroidal galaxies. Bulges are inevitably older than disks, and formed on a dynamical time scale. Their formation is characterized by a series of intense formation episodes or starbursts, produced by each merger. Most stars that are now in the bulge formed during the process of bulge accumulation.
Bulges are of intermediate age: One can also envisage the following prescription for bulge formation. Merging of dwarf irregular galaxies with a massive disk galaxy will result in the dwarf being stripped of gas. The angular momentum of the gas guarantees that it will eventually dissipate to provide infall into the disk. The stellar component, however, as it interacts with the disk is partially tidally disrupted, to form the thick disk, but the dense cores undergo dynamical friction, spiraling into the center to form the bulge. While this may not be appropriate to the inner Milky Way bulge, such a picture may be relevant to the outer spheroid. There is evidence for tidal streams that are continuously generated by disruption of satellites such as the LMC. The age spread of the globular star clusters is consistent with a model in which bulges and disks would be of similar age.
Bulges are young: Bulges may also form slowly by dynamical instabilities of disks. Secular evolution of disks has occurred in at least some galaxies, particularly in late-type galaxies (Kormendy 1992; Courteau 1996). The secular evolution of a cold disk inevitably results in gravitational instability. On galaxy scales this is dominated by the non-axisymmetric modes that induce formation of a bar. Tidal torques are expected by the bar on the disk gas, which consequently suffers angular momentum transfer and forms a massive central concentration. This in turn eventually tidally disrupts the bar, as well as undergoes a central starburst and forms the bulge. This process can repeat as infall of gas continues and the disk becomes sufficiently massive to again be gravitationally unstable. Bar disruption takes up to 100 bar dynamical timescales.
## 2 Signatures of Bulge Formation
The observed properties of bulges provide a fossil testament to their formation. The various signatures do not lead to any unique conclusion, however. Consider first the Milky way as a prototype for bulge formation, bearing in mind that our local neighborhood is necessarily limited in scope. The following remarks are largely summarized from an excellent review of the Milky Way bulge by Wyse, Gilmore and Franx (1997).
Ages: This should be the cleanest signature. The Milky Way bulge appears to be indistinguishable in age from the inner globular clusters that form a flattened subsystem and appear to be $`2`$ Gyr younger than the oldest globular cluster systems. However the colors of other bulges,especially in late-type disks, show a broad dispersion which may reflect age differences.
Abundances: The observed abundances of Milky Way bulge stars show a broad dispersion, suggestive of inhomogeneous enrichment, and the mean metallicity is lower than that of the disk, but higher than that of the halo. This supports the inference of an old bulge from observed ages in localized regions. In particular the outer bulge appears to be more metal-poor than the old disk. If age traces metallicity, one infers that the bulge precedes the disk. However dynamical processes could delay bulge star formation without inducing chemical evolution.
Angular momentum: The angular momentum distribution of the inner bulge resembles that of the halo (or outer bulge), rather than that of the disk. This is suggestive of a sequential formation process.
Other bulges provide a broader basis with which to search for clues to bulge formation.
Profiles: Core radii of bulges and disks are well correlated. This suggests that the formation of the two components is closely coupled.
Colors: Bulge colors show a broad dispersion, but generally track disk colors. Both metallicity and age must therefore be correlated.
Dynamics: Low luminosity bulges are rotationally supported, as are disks, but luminous bulges generally are supported by anisotropic velocity dispersion. Bulges overlap with, but generally have lower anisotropic velocity dispersion support (i. e. $`\sigma /v_{circ}`$) than do ellipticals. This suggests that massive bulges are distinct from disks and closer to ellipticals in dynamical origin, whereas low luminosity bulges are more closely associated with disk star formation.
Fundamental plane: Bulges lie in the identical fundamental plane as ellipticals, although there is a slight offset of the zero point. This is suggestive of a similar early formation phase for bulges to that for ellipticals.
The signatures provide mixed signals on the epoch of bulge formation. It is probably true that many bulges, especially if massive, form early, while some, especially if associated with later-type spirals, form late. The age differences provide an interesting environment with which to probe bulge formation models.
## 3 Bulges at High Redshift
The high redshift universe potentially provides a unique discriminant. The differences between the bulge models described above are magnified at high redshift.
We may broadly classify these bulge formation scenarios into three types: secular evolution in which bulges form relatively late by a series of bar-induced starbursts, one in which bulges form simultaneously with disks, and an early bulge formation model in which bulges form earlier than disks. Adjusting the three models to produce optimal agreement with $`z=0`$ observations, we compare their high-redshift predictions with present-day observations, in particular, with data compiled in various studies based on the CFRS (Schade et al. 1996; Lilly et al. 1998) and the HDF (Abraham et al. 1998). We have developed a simple scheme for simulating the rival models and comparing bulge colors and sizes with observations. Hubble Space Telescope photometry and color information is available for galaxy samples that extend up to redshift unity and beyond. At this epoch one may hope to detect the difference between old and young bulge models.
For the purposes of normalizing our models, we examine two local $`z=0`$ samples (de Jong & van der Kruit 1994; hereinafter, DJ; Peletier & Balcells 1996, hereinafter, PB). Both are diameter-limited samples but differ in orientation selection, so that edge-on disks in the PB sample are simply redder and less prominent relative to the bulges.
Starting with the local properties of disks and a reasonable distribution of formation times, we construct a fiducial disk evolution model, to which we add three different models for bulge formation, the principle difference being simply the time the bulges form relative to that of their associated disks. We do not attempt to model the internal dynamics or structure of spirals (e.g., Friedli & Benz 1995). We adopt the usual Sabc and Sdm luminosity functions (LFs) for disk galaxies (Binggeli, Sandage, and Tammann 1988). We evolve these galaxies backwards in time and in luminosity according to their individual star formation histories without number evolution. We take halo formation time to equal the time over which 0.25 of the final halo mass is assembled. We normalize by assuming a constant mass-to-light ratio where $`M_{b_J}=21.1`$ corresponds to $`4\times 10^{12}\mathrm{M}_{}`$ and we adopt the usual CDM matter power spectrum. We take star formation in the disk to commence at the halo formation time with an e-folding time that depends on the $`z=0`$ galaxy luminosity, to roughly fit the $`z=0`$ colour-magnitude relationship. We assume exponential profiles for the disks with a $`b_J`$ central surface brightness modified to account for the observed correlation between surface brightness and luminosity (e.g., de Jong 1996; McGaugh & de Blok 1997). We compute bulge spectra for the purposes of determining colours and magnitudes using the Bruzual & Charlot instantaneous-burst metallicity-dependent spectral synthesis tables (Leitherer et al. 1996).
To calibrate our fiducial disk evolution models, we compare the model predictions to both the colour-magnitude relationship of disks in spirals and the cosmic history of luminosity density. There is good agreement with the colour-magnitude relationship. All models, for which bulge, disk, and E/S0 contributions have been considered, produce fair agreement with the luminosity density of the universe at all redshifts for which observable constraints are available (Lilly et al. 1996; Madau et al. 1996; Connolly et al. 1997).
If bulges form through the merging of disk galaxies, the formation of the stars found in bulges is expected to precede the formation of stars in the disks which form out of gas that accretes around the spheroid (e.g., Kauffmann & White 1993; Frenk et al. 1996). For simplicity, we commence star formation in the bulge 4 Gyr prior to the formation of disks in our fiducial model and suppose that it continues for $`\tau _{burst}=0.1`$ Gyr.
Suppose next that star formation in the bulge commences at the formation time of disks, for example because high angular momentum gas forms the disk while low angular momentum gas simultaneously forms the bulge which undergoes a mild starburst. One can imagine gas-rich satellite infall. The gas is tidally stripped and accreted onto the disk at large radii, whereas the dense cores lose angular momentum by dynamical friction and are incorporated into the bulge. A refinement of this model would allow for a sequence of early mergers that formed the bulge. However the final merger dominates the luminosity and therefore the spectral evolution, since the star formation efficiency is greatest for the most massive systems.
Finally, in secular evolution, bulges form after disks, with gas accretion onto the disk triggering the formation of a bar that drives gas inflow into the center followed by star formation (Friedli & Benz 1995). The build-up of a central mass destroys the bar and inhibits gas inflow (Norman, Sellwood, & Hasan 1996), consequently stopping star formation in the bulge until enough gas accretes onto the galaxy to trigger the formation of a second bar, followed by a second central starburst. Somewhat arbitrarily, we suppose that the first central starbursts occur some 2 Gyr after disk formation in our fiducial model, that central starbursts last $`\tau _{burst}`$ = 0.1 Gyr, as in the simulations by Friedli & Benz (1995), and that 2.4 Gyr separate central starbursts, in order to illustrate the general effect of a late secular evolution model for the bulge. We assume that the star formation rate follows an envelope with an e-folding time equivalent to the history of disk star formation. We thereby force star formation in the disk and the bulge to follow very similar time scales, given the extent to which they are both driven by gas infall processes.
We add a simple model for E/S0 galaxies to aid with the interpretation of observed high redshift, high $`B/T`$ systems, somewhat arbitrarily assuming that the distribution of formation redshifts for the $`E/S0`$ population is scaled to be at exactly twice the distribution of formation redshifts.
We perform all our calculations using a galaxy evolution software package written by one of the authors for calculating how the gas, metallicity, star formation, luminosity, and colours vary as a function of time for a wide variety of morphological types, formation times, and star formation histories. With this software package, we present representative HDF simulations for our three bulge formation models in Figure 1 for comparison with the observations.
Clearly the secular evolution model, with late bulge formation, has a paucity of large B/T objects relative to the other models (Figure 2). The simultaneous bulge formation model has a large number of such galaxies simply because a large number of bulges were forming at this time, while the early bulge formation model has a slightly lower value due to the fact that bulges in this model had long been in place within their spiral hosts.
As expected, in all redshift bins, bulges are slightly bluer in the late bulge formation models than are the disks (Figure 3). A blue tail may be marginally detectable in the Schade et al. data in the highest redshift bins. Unfortunately, given the extremely limited amount of data and uncertainties therein, little can be said about the comparison of the models in all three redshift bins, except that the range of bulge and relative bulge-to-disk colours found in the data appears to be consistent with that found in the models.
While consistent with currently available data, our models for bulge formation are schematic and are intended to illustrate the observable predictions that will eventually be made when improved data sets are available in the near future. Our models are still quite crude, assuming among other things that the effects of number evolution on the present population of disks can be ignored to $`z1`$ as suggested, for example, in Lilly et al. 1998. In contrast, one recent analysis (Mao, Mo, & White 1998) has argued that observations favor the interpretation that a non-negligible amount of merging has taken place in the disk population from $`z=0`$ to $`z=1`$. For this particular interpretation, it remains to be seen how all the present stellar mass in disks could have built up if disks were continually destroyed by merging to low $`z`$ given the constraints on the cosmic star formation history. Infall of metal-poor gas provides a non-destructive alternative that is supported by chemical evolution modeling of the old disk and even by observations of a reservoir of high velocity outer halo clouds.
## Acknowledgements
We thank our collaborator Laura Cayon for many inspiring discussions of bulge issues. This research has been supported in part by NSF.
## 4 References
|
no-problem/9812/cond-mat9812305.html
|
ar5iv
|
text
|
# Positive magnetoresistance and orbital ordering in La1-xSrxMnO3
## Abstract
We report on detailed transprort measurements of single crystalline La<sub>1-x</sub>Sr<sub>x</sub>MnO<sub>3</sub> ($`x0.2`$). We have found a giant positive magnetoresistance in the compositions range between $`0.1x0.125`$ and give an explaination in terms of orbital ordering due to the interplay between superexchange interactions and Jahn-Teller distortions.
In the last years an overwhelming interest in the manganite perovskites La<sub>1-x</sub>A<sub>x</sub>MnO<sub>3</sub> arose primarily due to the observation of a collosal negative magnetoresistance (CMR) close to $`x=0.3`$ . These CMR effects at the ferromagnetic (FM) phase transition were explained within extended double-exchange (DE) models . Since then it became clear that this compounds reveal many interesting and puzzling phenomena which can not be accounted for by double exchange alone. It was Millis et al. who first related dynamic Jahn-Teller (JT) distortions to the CMR effect. In a recent paper we have presented a detailed phase diagram of La<sub>1-x</sub>Sr<sub>x</sub>MnO<sub>3</sub> at low Sr concentrations. In the concentration regime $`0.1x0.15`$ we detected a ferromagnetic (FM) and insulating (I) ground state which is followed by a canted antiferromagnetic (CA) or mixed phase at elevated temperatures . This insulating FM-phase results from superexchange (SE) interactions in a charge-ordered (CO) phase which probably also reveals orbital order.
In this paper we will present detailed transport measurements on La<sub>1-x</sub>Sr<sub>x</sub>MnO<sub>3</sub> single crystals. It will be shown that the double degeneracy of the $`e_g`$ orbitals and their ordering in real space must be considered in order to understand the ferromagnetic and insulating ground state in this doping regime. Recently theoretical models considered the orbital degrees of freedom, but have mainly focused on the properties of the LaMnO<sub>3</sub> and La<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub> compounds . Although Ahn and Millis discussed the interplay of orbital and charge ordering for La<sub>0.875</sub>Sr<sub>0.125</sub>MnO<sub>3</sub> they did not considered the effect of this types of ordering on the mangetic properties. Using a mean field approximation Maezono $`et`$ $`al.`$ have obtained an overall phase diagram of doped manganites based on the interplay of SE- and DE-interactions.
Here we provide strong experimental evidence that the competition between JT-, SE- and DE-interactions is responsible for the rich variety of magnetic and structural phase transitions in the low doping phase diagram ($`x0.2`$). At the phase boundary of the insulating FM ground state, a strong (collossal) positive MR appears. We will show that the close coupling of structural and magnetic phase transitions and especially the magnetic field induced structural transitions can only be explained assuming orbital order in the O´´-phase.
Single crystals of La<sub>1-x</sub>Sr<sub>x</sub>MnO<sub>3</sub>, with concentrations $`0.1x0.2`$ were grown by a floating zone method with radiation heating in air atmosphere. X-ray diffraction of crushed single crystals revealed high-quality single-phase materials. X-ray topography indicated twinning of the crystals. Transport measurements in the temperature range 1.5-300 K were performed with the standard four-probe method in fields up to 14 T.
The relevant part of the phase diagram around $`x=0.125`$ is reproduced in Fig. 1. Close to this concentration a rather unusual sequence of phase transitions has been detected. At room temperature La<sub>0.875</sub>Sr<sub>0.125</sub>MnO<sub>3</sub> is orthorhombic. A long range JT distortion (and hence orbital order) appears at T$`_{\text{OO´}}`$=265 K, both orthorhombic phases beeing insulating . At T<sub>CA</sub>=180 K a CA structure is established where the resistivity is reduced almost by a factor of 10. It is this regime which has been treated as a FM and metallic phase during the last years. This however was not based on experimental results but only due to the fact that $`d\rho /dT>0`$ in a limited temperature regime and that a FM regime which is followed by a canted spin structure at lower temperatures nicely fits into de Gennes phase diagram . We have shown that this phase reveals a canted structure . Of course, we also can not exclude a mixed phase in this regime, but it is definitely not metallic. The decrease in the resistivity probably indicates an increase of the charge transfer matrix elements due to the increasing importance of the DE mechanism combined with the freezing out of spin-disorder. Finally at T$`_{\text{O´O´´}}`$=150 K and T<sub>C</sub>=140 K a structural phase transition is immediately followed by the onset of FM order. The O´´-phase reveals charge order in an almost pseudocubic lattice and the ground state is a FM insulator.
The field dependence of the magnetoresistance ($`\mathrm{\Delta }\rho `$/$`\rho `$(0)=$`\rho `$(H)-$`\rho `$(0)/$`\rho `$(0)) for $`x=0.1`$ and $`x=0.125`$ is shown in Fig. 2 for various temperatures. For temperatures T$`>`$T$`_{\text{O´O´´}}`$ we find negative MR effects. However we would like to point out that here the negative MR appears at the transition from a paramagnetic (PM) insulator into a CA or mixed phase. Small negative MR effects again appear for T$`<`$T<sub>C</sub>=105 K. However, between T$`_{\text{O´O´´}}`$ and T<sub>C</sub> large or even colossal positive MR effects appear, due to the fact that the FM ground state is indeed more insulating than the CA phase. A closer inspection of Fig. 2 reveals that two subsequent jumps appear as a function of increasing field. Both jumps induce higher resistivity states and correspond to those ones observed in the magnetization curves (see Fig. 3 in ). The first jump is due to a field induced structural phase transition from the JT-distorted O´- to the pseudocubic O´´- orthorhombic phase and obviously is accompanied by a real space charge ordering of the Mn<sup>3+</sup> and Mn<sup>4+</sup> ions . At the subsequent second jump the sample undergoes a transition into a FM state with the magnetoresistance showing a saturated behavior at higher fields. For $`x=0.125`$ these jumps are not clearly separated, a fact that has also been observed in the magnetization curves. In contradiction to the $`x=0.1`$ compound, the magnetoresistance of the $`x=0.125`$ sample decreases with increasing field below the field induced transition to the charge ordered FM state. Finally for the $`x=0.15`$ sample (not shown) only negative CMR effects have been observed in agreement with previous published results . It is remarkable that in the temperature range between T<sub>CA</sub> and T<sub>C</sub> the magnetoresistance at a given field changes sign and becomes positive.
To show these large positive MR effects, the temperature dependence of the magnetoresistance for various fields and doping levels is plotted in Fig 3. For $`x=0.1`$ two pronounced peaks yielding an increase in resistivity up to +400% when the magnetic field is raised to 5 T are clearly visible. These peaks corresponds to the transitions into the O´´- and FM-state respectively. Around T<sub>CA</sub> the typical negative MR effect can be seen. The same features are also present for $`x=0.125`$, though the two positive enhachments at T$`_{\text{O´O´´}}`$ and T<sub>C</sub> obviously merge resulting into a single peak. The positive MR is maximal for $`x=0.1`$, becomes significant smaller for $`x=0.125`$ and finally dissappears for $`x0.15`$. Summarizing the magnetoresistance measurements indicate a large positive MR at the transitions O´/CA$``$O´´/CA and O´´/CA$``$O´´/FM, while negative MR effects appear close to the O´/PM$``$O´/CA and O/PM$``$O/FM phase boundaries (see Fig. 1). The negative MR obviously result from the onset of spin order below T<sub>CA</sub> (i.e. T<sub>C</sub> for $`x>0.15`$) and can be explained within a double-exchange picture as proposed by . The electronic properties in the O´´-phase are more complicated and can not be explained taking only double-exchange interactions in account.
Finally in Fig. 4 we show a H-T-phase diagram for La<sub>0.9</sub>Sr<sub>0.1</sub>MnO<sub>3</sub>. We choose this concentration as for this sample the sequence of two strongly coupled phase transitions (O´$``$O´´, CA$``$FM) can be easily documented (see e.g. Figs. 2 and 3). The insulating region of positive MR is enbedded in insulating phases which reveal negative MR effects. Both transition temperatures (T$`_{\text{O´O´´}}`$, T<sub>C</sub>) are shifted to higher temperatures as the external field is increased. At the two closely related phase boundaries T$`_{\text{O´O´´}}`$ and T<sub>C</sub> three degrees of freedom, namely the charges (Mn<sup>3+</sup>/Mn<sup>4+</sup>), the spins and the orbitals undergo an ordering process. The structural phase transition T$`_{\text{O´O´´}}`$ indicates charge ordering and most probably also orbital ordering.
We start our discussion summarizing the most important experimental results presented here. It is obvious that the transition from the canted-AFM and JT-distorted O´-phase to the pseudocubic charge-ordered O´´-phase is intimately coupled to a transition into an insulating FM state. The positive CMR effect observed in the vicinity of the commensurate doping value $`x=0.125`$ counts for a very different picture as it can be given by the interplay of JT-distortion and DE-interactions alone.
It is well know that the double degeneracy of the $`e_g`$ levels of $`3d`$ ions in an octahedral enviroment is lifted by an JT-distortion of the lattice , accompanied by an ordering of the orbitals in real space. The possibility of orbital ordering in transition-metal ions due to exchange interactions different than that resulting from JT-distortions was first pointed out by Roth and has been extensively studied by Kugel and Khomskii . It has been shown that two transitions take place, one into an orbitally ordered state and a second into a spin-ordered state, both driven by SE-mechanisms. This SE differs from the ordinary one due to the fact that each electron has four degrees of freedom, two orbital states ($`d_{z^2}`$, $`d_{x^2y^2}`$) and two spin states (spin-up, spin-down). The presence of intra-atomic exchange in case of orbital degeneracy produce ferromagnetism below the orbital-ordering phase transition . It is important to notice that above the orbital-ordering transition temperature the effective spin-spin interaction is AFM, which is modified by the appearance of the orbital ordered state and finally goes over into a FM coupling. A modern view of this problem has been recently considered by Held and Vollhardt .
For undoped LaMnO<sub>3</sub> the magnetic properties can be well explained taking only the predominant JT-distortion of the MnO<sub>6</sub> octaherda into account. The double degeneracy of the $`e_g`$ orbitals is lifted by a long range cooperative Jahn-Teller distortion resulting in an ordering of the $`d_{z^2}`$ orbitals as has been argued by Solovey $`et`$ $`al`$. and has recently been confirmed experimentaly by Murakami $`et`$ $`al.`$ . As a result of this JT driven orbital ordering the A-type AFM state is established below T<sub>N</sub>.
Upon doping with holes the long-range JT-distortion become suppressed and concomitantly double-exchange interactions have to be taken into acount. At room temperature the lattice is distorted as in LaMnO<sub>3</sub> (O´-phase) due to the JT effect removing the double degeneracy of the $`e_g`$ levels. Since mobile holes are present, the double exchange mechanism plays a fundamental role. The magnetic structure observed in undoped LaMnO<sub>3</sub> is modified by the appearance of three-dimensional ferromagnetic DE-interactions, competing with the JT driven A-type AFM yielding a canted AFM state (or mixed phase). Thus canted antiferromagnetism is established below T<sub>CA</sub>=150 K, accompanied by a drop in the resistivity, since a gain in kinetic energy of the carriers due to DE-interactions can be achieved. Further lowering of the temperature favours SE-interactions between Mn<sup>3+</sup> ions in a manner discussed above, yielding the ordering of orbitals and subsequently the evolution of ferromagnetism. At T$`_{\text{O´O´´}}`$ the SE-interactions become dominant, suppress the JT-distortion and enforces the structural transition into the pseudocubic O´´-phase, followed by the onset of ferromagnetism at T<sub>C</sub>. At the same time an insulating behavior appears due to charge and orbital ordering, which decreases the hole mobility and explains the positive jumps in the MR curves. The transition to the orbital ordered FM state, is stabilized by the application of an external magnetic field as can be seen in Fig. 1 and Fig. 2. We remind that by application of an external magnetic field jumps in the magnetization curves has been observed, leading the sample to higher magnetization states . This field induced transitions can not be ascribed to magnetocrystalline anisotropy since no dependence was found for different orientations of the crystal axes in respect to the applied external field. The possibility of magnetic field induced transitions due to the interplay of the JT-effect and SE-interactions has been first pointed out by Kugel and Khomskii . It was shown that when orbital ordering is enforced by the SE mechanism such transitions are possible. On the other hand if orbital ordering is established due to electron-lattice interactions (JT-effects) as it is believed to be in LaMnO<sub>3</sub>, then such transitions may not appear.
At the moment only predictions about how the orbitals are ordered in the low temperature state can be made, but an alternation of occupied $`d_{z^2}`$ and $`d_{x^2y^2}`$ orbitals on neighbouring Mn<sup>3+</sup>-sites appear most probably to us (Fig. 5). This would cause a displacement of the O<sup>2-</sup> ions due to secondary electrostatic interactions , yielding alterning long and short Mn-O bonds. Together with the real space ordering of the Mn<sup>3+</sup>/Mn<sup>4+</sup> ions this would result in additional superstructure reflections in the diffraction patterns. Finally for doping levels higher than $`x0.175`$ the double exchange mechanism becomes dominant and a metallic ferromagnetic state is established. The orbital liquid picture recently proposed by Ishihara $`et`$ $`al`$. may be appropriate to describe the physics in the metallic phase.
The authors would like to thank D. I. Khomskii for helpfull discussions. This work has in part be supported by the BMBF under the contract number 13N6917.
FIGURE CAPTIONS:
Fig. 1: Phase diagram of La<sub>1-x</sub>Sr<sub>x</sub>MnO<sub>3</sub> at low doping concentrations. The shaded area represents an orbitally and charge ordered insulating ferromagnet.
Fig. 2: Isothermal magnetoresistance curves for the $`x=0.1`$ and $`x=0.125`$ samples.
Fig. 3: Temperature dependence of the MR for $`x=0.1`$ and $`x=0.125`$ at various applied magnetic fields.
Fig. 4: H-T-phase diagram for $`x=0.1`$. The shaded area corresponds to the region of positive MR. A strong negative MR appears close to the OO´-phase boundary.
Fig. 5: Orbital ordering in La<sub>0.875</sub>Sr<sub>0.125</sub>MnO<sub>3</sub> which is in accord with the charge order proposed by Yamada $`et`$ $`al`$..
|
no-problem/9812/nucl-ex9812003.html
|
ar5iv
|
text
|
# Energy dependence of the isomeric cross section ratio in the 58Ni(𝑛,𝑝)⁵⁸Com,g reactions
## ACKNOWLEDGEMENTS
This work was supported in part by the Hungarian Research Found (Contract No. T 025024), the International Atomic Energy Agency, Vienna (Contract Nos. 7687/R0 and 8205), the International Science and Technology Center (Project No. 176), and the Romanian Ministry of Research and Technology Grant No. 3028GR/B10.
TABLE I. Measured cross sections for the <sup>58</sup>Ni$`(n,p)^{58}`$Co and <sup>58</sup>Ni$`(n,p)^{58}`$Co<sup>m</sup> reactions, and measured and deduced isomeric cross section ratios for the former reaction.
| $`Neutron`$ | Measured | Measured | Measured | Deduced |
| --- | --- | --- | --- | --- |
| $`energy`$ | $`\sigma (n,p)`$ | $`\sigma _m`$ | $`\sigma _m/(\sigma _g+\sigma _m)`$ | |
| $`(MeV)`$ | (mb) | (mb) | | |
| $`2.14`$ | | 19.2$`\pm `$1.3 | 0.260$`\pm `$0.014 | |
| $`2.21`$ | | | 0.259$`\pm `$0.011 | |
| $`2.23`$ | | | 0.274$`\pm `$0.026 | |
| $`2.30`$ | | 20.6$`\pm `$1.4 | 0.234$`\pm `$0.010 | |
| $`2.43`$ | | | 0.254$`\pm `$0.012 | |
| $`2.59`$ | | | 0.277$`\pm `$0.011 | |
| $`2.60`$ | | 29.1$`\pm `$2.0 | | |
| $`2.74`$ | | 41.0$`\pm `$3.0 | 0.289$`\pm `$0.012 | |
| $`2.83`$ | | | 0.269$`\pm `$0.008 | |
| $`2.84`$ | | | 0.278$`\pm `$0.009 | |
| $`2.94`$ | | 56.3$`\pm `$3.5 | 0.276$`\pm `$0.011 | |
| $`10.3`$ | | | 0.445$`\pm `$0.019 | |
| $`12.3`$ | | | 0.452$`\pm `$0.020 | |
| $`13.4`$ | | | 0.545$`\pm `$0.004 | |
| $`13.56`$ | 413.6$`\pm `$13.2 | 234$`\pm `$9 | | 0.566$`\pm `$0.028 |
| $`13.6`$ | | | 0.536$`\pm `$0.004 | |
| $`13.74`$ | 383.4$`\pm `$16.1 | 218$`\pm `$9 | | 0.569$`\pm `$0.034 |
| $`13.96`$ | 359.1$`\pm `$17.2 | 197$`\pm `$8 | | 0.549$`\pm `$0.035 |
| $`14.03`$ | | | 0.545$`\pm `$0.007 | |
| $`14.05`$ | | | 0.552$`\pm `$0.005 | |
| $`14.19`$ | 329.9$`\pm `$12.8 | 182$`\pm `$8 | | 0.552$`\pm `$0.032 |
| $`14.42`$ | 313.8$`\pm `$10.8 | 170$`\pm `$7 | | 0.542$`\pm `$0.029 |
| $`14.48`$ | | | 0.573$`\pm `$0.012 | |
| $`14.61`$ | 292.3$`\pm `$14.0 | 173$`\pm `$11 | | 0.592$`\pm `$0.047 |
| $`14.68`$ | | | 0.556$`\pm `$0.006 | |
| $`14.78`$ | 275.9$`\pm `$11.8 | 150$`\pm `$6 | | 0.544$`\pm `$0.032 |
| $`14.88`$ | | | 0.573$`\pm `$0.005 | |
TABLE II. The number of discrete levels $`N_d`$ up to excitation energy $`E_d`$ used in Hauser–Feshbach calculations, taken from the corresponding references, and the low–lying levels as well as s–wave nucleon–resonance spacings $`\overline{D}_{exp}`$ in the nucleon energy range $`\mathrm{\Delta }`$E above the corresponding binding energy B<sub>n</sub> used to obtain the BSFG parameters, i.e. the level–density parameter a, the ratio of the nuclear moment of inertia $`I/I_r`$, and the ground–state shift $`\mathrm{\Delta }`$.
| Nucleus | $`N_d`$ | $`E_d`$ | Ref. | Fitted level and resonance data | | | | $`a`$ | $`I/I_r`$ | $`\mathrm{\Delta }`$ |
| --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- |
| | | (MeV) | | $`N_d`$ | $`E_d`$ | $`B_n+\frac{\mathrm{\Delta }E}{2}`$ | $`\overline{D}_{exp}`$ | $`(MeV^1)`$ | | $`(MeV)`$ |
| | | | | | (MeV) | $`(MeV)`$ | (keV) | | | |
| <sup>59</sup>Ni | $`13`$ | 1.948 | | $`20`$ | 2.48 | $`9.33`$ | 12.5$`\pm 0.9^a`$ | $`6.25`$ | 1.0 | $`1.20`$ |
| <sup>58</sup>Ni | $`28`$ | 4.475 | | $`32`$ | 4.58 | | | $`6.00`$ | 1.0 | $`0.28`$ |
| <sup>58</sup>Co | $`28`$ | 1.555 | | $`28`$ | 1.56 | | | $`6.60`$ | 1.0 | $`2.37`$ |
| <sup>58</sup>Co | $`28`$ | 1.555 | | $`28`$ | 1.56 | | | $`6.11`$ | 0.5 | $`2.40`$ |
| <sup>55</sup>Fe | $`16`$ | 2.600 | | $`16`$ | 2.60 | $`9.55`$ | 18.0$`\pm 2.4^a`$ | $`5.60`$ | 1.0 | $`1.30`$ |
<sup>a</sup>Reference .
## Figure Captions
FIG. 1. Comparison of the measured and calculated excitation functions and isomeric cross section ratios for the <sup>58</sup>Ni$`(n,p)^{58}`$Co<sup>m,g</sup> and <sup>59</sup>Co$`(n,2n)^{58}`$Co<sup>m,g</sup> reactions. The calculated cross-section curves (c),(d) were obtained by using the fitted value of the branching ratio for the 52.8 keV$``$24.9 keV transition (solid curves), the isomeric cross-section ratios (a),(b) were found by using also the evaluated branching-ratio (dotted curves), using only the two levels (dashed-dotted curves), as well as the fitted branching-ratio but considering two assigned 4<sup>+</sup> levels populating either the g.s. (lower dashed curves) or the isomeric state (higher dashed curves). For the experimental data see Refs. .
FIG. 2. (a) Partial cross sections for the compound nucleus formation versus the corresponding total angular momentum $`J_{CN}`$, at the given incident energies of neutrons on the target nucleus <sup>58</sup>Ni. (b) The yrast plot for the residual nucleus <sup>58</sup>Co, of the adopted discrete levels including the spin assignment ($`+`$), while in the opposite case the spin values considered in the present calculations are additionally marked if the corresponding level has an adopted decay scheme ($`\times `$) or only the excitation energy ($``$); the yrast lines showed only for orientation correspond to the effective excitation energies obtained by using the BSFG parameters in Table II, and the nuclear moment of inertia for a rigid body $`I_r`$ (dashed curve) and respectively one-half of $`I_r`$ (dotted curve) with a reduced radius $`r_0`$=1.25 fm.
FIG. 3. The same as Fig. 1(a), except the calculated values are obtained by using (a) 28 discrete levels up to the excitation energy $`E^{}`$=4.475 MeV of the nucleus <sup>58</sup>Ni, while for the nucleus <sup>58</sup>Co are used either 28 discrete levels (solid curve) or 29 levels (dashed curve) up to $`E^{}`$=1.555 MeV, 3 discrete levels up to $`E^{}`$=0.053 MeV (dotted curve), or 2 discrete levels up to $`E^{}`$=0.025 MeV (dashed-dotted curve), and (b) only 2 discrete levels up to $`E^{}`$=0.025 MeV for the nucleus <sup>58</sup>Co while for the nucleus <sup>58</sup>Ni are used either 28 levels up to $`E^{}`$=4.475 MeV (solid curve), only g.s. (dashed curve), 2 levels up to $`E^{}`$=1.454 MeV (dashed-dotted curve), or 4 levels up to $`E^{}`$=2.776 MeV (dotted curve).
|
no-problem/9812/astro-ph9812312.html
|
ar5iv
|
text
|
# M84 - A Warp Caused By Jet Induced Pressure Gradients?Based on observations with the NASA/ESA Hubble Space Telescope obtained at the Space Telescope Science Institute which is operated by the Association of University for Research in Astronomy, Inc. (AURA), under NASA contract NAS5-26555.
## 1 Introduction
Recent imaging from the Hubble Space Telescope (HST) and spectroscopic studies of active elliptical galaxies have established that there are gas disks in the central few hundred pc of these galaxies which could be feeding the massive $`10^9M_{}`$ black holes at their centers (e.g M84; Bower et al. 1997a , Bower et al. 1997b , M87; Harms et al. (1994) and NGC 4261; Ferrarese, Ford, & Jaffe (1996)). These gas disks are often observed to be nearly perpendicular to the jets, establishing a link between the angular momentum of the disk and the direction of the jet. Surveys of the observed angular difference, $`\mathrm{\Psi }`$, between radio and disk axes in radio ellipticals and find a statistically significant peak in the distribution at $`\mathrm{\Psi }=90^{}`$ (Kotanyi & Ekers (1979), Möllenhoff, Hummel & Bender (1992)). This has established that radio jets are generally perpendicular to gas disks on large (kpc) scales in nearby radio galaxies, even though no correlation between the radio axis and the galaxy isophotal major axis exists at low redshift (e.g., Sansom et al. (1987)). The strong peak at $`\mathrm{\Psi }=90^{}`$ has been confirmed at smaller (100 pc) scales by van Dokkum & Franx (1995) using HST/WFPC images.
Strangely, the jet direction is not necessarily expected to be correlated with the angular momentum of the gas feeding the black hole (and so the orientation of the disk) because of Lense–Thirring precession, otherwise known as ‘dragging of inertial frames’ (Rees (1978)). A spinning black hole causes a gas disk near the gravitational radius, $`r_g=GM/c^210^{14}`$ cm, to precess, and so the disk or torus near the black hole is expected to fill a volume that is axisymmetric and aligned with the spin axis of the black hole itself (Rees (1984), Bardeen & Petterson (1975)). This inner torus is the proposed site for jet collimation and acceleration (Rees et al. (1982)) so we expect that the jet should be aligned with the spin axis of the black hole but not necessarily the angular momentum of the disk well outside of $`r_g`$. A massive spinning black hole can be described as an angular momentum reservoir that does vary in momentum but only very slowly with the influx of fuel. The rate of change of angular momentum is expected to be particularly slow ($`10^9`$ years) in radio galaxies where the black holes are massive ($`10^9M_{}`$) and the fueling rates are probably low ($`10^4M_{}`$/yr; Rees (1978)), though accretion of a second black hole could change the spin axis of the AGN on a very short timescale. If, however, the black hole does not spin significantly, then the jet axis would be determined by the angular momentum of the inner disk. This disk, because of its lower mass, could vary in orientation on shorter timescales than the black hole axis (for a spinning system). However, the observed alignment of jet and disk axes would then require that the orientation of this inner disk be coupled to that of the outer disk (at $`100`$pc). Possible mechanisms causing alignment between the jets and a disk at 100 pc scales (well beyond $`r_g`$) have not been explored.
To explore alignment mechanisms, we consider the geometry of the warped dust features and ionized emission in M84 (NGC 4374, 3C 272.1), a radio elliptical in the Virgo cluster. M84 has a dusty disk seen in extinction in optical images (see Fig. 1) and in emission in H$`\alpha `$ +\[NII\] that although nearly perpendicular to the jet, is misaligned with the galaxy isophotes (Hansen, Norgaard-Nielsen, & Jorgensen (1985), van Dokkum & Franx (1995), Jaffe et al. (1994), Bower et al. 1997a , Baum et al. (1988)). van Dokkum & Franx (1995) noted that the timescale for the disk to settle into the galaxy plane of symmetry is probably short; $`10^8`$ years. These authors therefore suggested that the misalignment could be explained by a rapid inflow of gas into the central region of the galaxy (to radii smaller than $`r600`$ pc). However if this is the case, it is difficult to explain the S-shape of the H$`\alpha `$ \+ \[NII\] emission (Baum et al. (1988)) and the detailed morphology of the dusty disk itself. This emission is elongated and lies within $`20^{}`$ of PA $`70^{}`$ for $`r7^{\prime \prime }`$ but at this radius the disk twists to PA $`115^{}`$ on either side of the nucleus. It is difficult to imagine how a cooling flow could generate such a morphology. As an alternative to this possibility, we also consider an interaction between the gas disk and a hot ambient interstellar medium (ISM) as a possible mechanism for affecting the alignment of the disk. By integrating the light of the galaxy through a warped gas and dust disk we find that the geometry of gas disk in M84 is likely to differ that predicted from a simple precession model. We search for a simple physical model that can account for the observed disk morphology.
Throughout this paper we adopt a distance to M84 of 17Mpc (Mould et al. (1995); $`H_0=75`$km s<sup>-1</sup> Mpc<sup>-1</sup>). At this distance $`1^{\prime \prime }`$ corresponds to 82 pc.
## 2 Misalignment of the Dust Features with the Galaxy
That the dust features are misaligned with the galaxy isophotes in M84 for $`r600`$ pc was emphasized by van Dokkum & Franx (1995). The galaxy isophotes have a major axis PA $`=129^{}`$ (van Dokkum & Franx (1995); see Fig. 1) whereas the dust features are elongated along PA$`=8065^{}`$ (for $`r<7^{\prime \prime }`$) depending upon the region used to estimate the angle (Bower et al. 1997a , van Dokkum & Franx (1995)). This is a misalignment of $`5065^{}`$ of the dust features major axis with respect to the galaxy major axis.
Misalignment between gas and stellar isophotes is not necessarily rare, (e.g. see van Dokkum & Franx (1995)). When there are kpc scale gas/galaxy misalignments usually the misalignment is the result of a merger (e.g. Tubbs (1980) and Steiman–Cameron, Kormendy, & Durisen (1992)). A gas disk in a non-spherical but axisymmetric galaxy will precess about the galaxy axis of symmetry. At a radius of a few kpc the time for a gas disk to precess about the galaxy axis of symmetry is long, $`10^8`$ years, requiring a merger event to have occurred within this time if a merger is responsible for the gas/galaxy misalignment. Because the precession is faster in the center of the galaxy in the central regions the warp can be multiply folded and in the outer regions where the precession is slower the angle of the disk is more directly related to the orbital angular momentum of the merger event (Quillen, Graham, & Frogel (1993)). A multiply warped disk will form a band of absorption that is roughly perpendicular to the galaxy axis of symmetry. The morphology of the dust lanes in M84 resembles those of Centaurus A and NGC 4753 which are caused by multiply folded thin dusty surfaces (Quillen et al. (1993), Steiman–Cameron et al. (1992)). However precession in an axisymmetric galaxy potential is an unlikely explanation for the warp in M84 since the band of absorption in the multiply folded region is not aligned with the galaxy isophotes.
van Dokkum & Franx (1995) emphasized that the timescale for the gas disks to settle into the plane of the galaxy is short in the central few hundred pc. The timescale for a gas ring to precess about the galaxy axis of symmetry, is shorter still. The angular precession frequency $`d\alpha /dt`$ about the galaxy axis of symmetry for a gas ring of radius $`r`$ and inclination $`\theta _g`$ with respect to the galaxy axis of symmetry is
$$\frac{d\alpha }{dt}=ϵ_\mathrm{\Phi }\mathrm{\Omega }\mathrm{cos}\theta _g$$
(1)
(e.g. Gunn (1979)) where $`ϵ_\mathrm{\Phi }`$ is the ellipticity of the gravitational potential which we assume is axisymmetric and $`\mathrm{\Omega }v_c/r`$ for $`v_c`$ the velocity of a particle which has only a gravitational force on it in a circular orbit. Except in the case of a polar ring ($`\theta _g90^{}`$), the dependence on inclination is weak. In a time
$$t_p\frac{3}{ϵ_\mathrm{\Phi }\mathrm{\Omega }}$$
(2)
which we refer to as a precession time, rings that differ in radius by a factor of 2 should have an angular difference of $`\pi /2`$.
To calculate the precession timescale we estimate the circular velocity, $`v_c1.1\sigma _{}`$ (for a mildly anisotropic velocity distribution) where $`\sigma _{}=303\pm 5`$km/s for $`r5^{\prime \prime }`$ (Davies & Birkinshaw (1988)) is observed stellar velocity dispersion. The ellipticity of the gravitational potential, $`ϵ_\mathrm{\Phi }`$, which is directly related to the ellipticity of the isophotes (see Fig. 1) $`ϵ_\mathrm{\Phi }ϵ/3`$ (see Binney & Tremaine (1987), Figure 2-13). Here the isophote ellipticity, $`ϵ1b/a`$, where $`b/a`$ is the axis ratio of the isophotes. In the near-infrared NICMOS 2.2$`\mu `$m and $`1.6\mu `$m images the isophotes of M84 are symmetrical, do not vary in position angle, and show little extinction from dust. The ellipticity from these NICMOS images shows that an $`ϵ>0.1`$ persists to well within the core break radius ($`r_b2^{\prime \prime }`$) to $`r0.3^{\prime \prime }`$ (Quillen, Bower & Stritzinger (1999)). This results in
$$t_p=2\times 10^7\mathrm{yr}\left(\frac{0.05}{ϵ_\mathrm{\Phi }}\right)\left(\frac{r}{100\mathrm{p}\mathrm{c}}\right)\left(\frac{330\mathrm{k}\mathrm{m}/\mathrm{s}}{v_c}\right).$$
(3)
Since in M84 the dust is misaligned with the galaxy for $`r600`$ pc, a mechanism operating at a timescale faster than the above timescale must be causing the disk to remain at an angle differing from the galaxy isophotes.
### 2.1 Pressure Required to Overcome the Galaxy Torque
Here we consider the possibility that the misalignment of the disk with the galaxy is caused locally. Radio galaxies put a substantial fraction of their luminosity into the kinetic energy of their jets (e.g. Rees et al. (1982)). Emission in H$`\alpha `$ \+ \[NII\] commonly shows high velocity dispersions of the order of a few hundred km/s (Baum, Heckman, & van Breugel (1992), Tadhunter, Fosbury, & Quinn (1989), Axon et al. (1989)). In radio galaxies there is strong evidence that the jets themselves impart significant motions in the surrounding ISM, implying that a significant fraction of the jet mechanical energy is dissipated locally in the surrounding ISM (e.g. De Young (1981); Begelman (1982)). If the jets are responsible for motions in the surrounding ISM then there should be a differential in this medium, with the largest pressures nearest the jets, and the lowest pressures in the plane perpendicular to the jets.
If the disk orientation is affected by the local medium then there must be a pressure on the gas disk greater than that of the gravitational potential which would be causing it to precess. The torque per unit mass on a ring of radius $`r`$ is
$$\tau =ϵ_\mathrm{\Phi }\mathrm{cos}(\theta _g)\mathrm{sin}(\theta _g)v_c^2$$
(4)
and so the pressure required to keep the disk from precessing is
$$P_\tau >ϵ_\mathrm{\Phi }\mathrm{cos}(\theta _g)\mathrm{sin}(\theta _g)\mathrm{\Sigma }v_c^2/r$$
(5)
where $`\mathrm{\Sigma }`$ is the mass per unit area in the gas disk. This pressure is
$$P_\tau 2\times 10^{11}\mathrm{dynes}\mathrm{cm}^2\left(\frac{ϵ_\mathrm{\Phi }}{0.05}\right)\left(\frac{\mathrm{\Sigma }}{1M_{}/\mathrm{pc}^2}\right)\left(\frac{v_c}{330\mathrm{k}\mathrm{m}/\mathrm{s}}\right)^2\left(\frac{100\mathrm{p}\mathrm{c}}{r}\right)\mathrm{cos}(\theta _g)\mathrm{sin}(\theta _g)$$
(6)
and can be compared to that estimated from the hot medium based on the X-ray emission. From the X-ray emission, Thomas et al. (1986) find an electron density that is $`n_e0.5(100\mathrm{p}\mathrm{c}/r)\mathrm{cm}^3`$ (where we have extrapolated from their profile which ranges from 0.5 – 20 kpc). This extrapolation is somewhat justified by comparison of pressures estimated from the \[SII\] 6717, 6731Å lines in cooling flow galaxies (including M87 within $`r<200`$ pc) which are not inconsistent with high central densities extrapolated from the lower resolution X-ray estimated pressures (Heckman et al. (1989)). The above density can be converted to a thermal pressure in the hot medium of
$$P_{therm}(r)10^9\mathrm{dynes}\mathrm{cm}^2\left(\frac{100\mathrm{p}\mathrm{c}}{r}\right)$$
(7)
(using a temperature of 0.67 kev; Matsumoto et al. (1997)). This pressure could be larger than that required to to overcome the galaxy torque. This suggests that hydrostatic forces could dominate the gravitational torque. If there is a pressure gradient aligned with the jet axis in the hot medium then it is likely that the disk orientation could be strongly affected by a torque from the hot ambient medium.
## 3 Modeling the Warped Disk in M84
We first discuss constructing geometrical models for the dust features in M84 where the light of the galaxy is integrated taking into account the opacity from a warped disk. We then consider two physical models for the disk geometry which are based on a disk made up of rings of gas which are precessing differentially. Precession is either caused by torques from a triaxial galaxy or a combination of torques from an oblate galaxy and due to a pressure gradient in the ambient X-ray emitting medium (as proposed above).
### 3.1 A geometrical model for the warp
We can begin with a model for the geometry of the gas disk which assumes the disk is made up of rings of material which precess about a given symmetry axis (e.g. used by Tubbs (1980), Quillen et al. (1993), and Steiman–Cameron, Kormendy, & Durisen (1992)). Here we here refer to this axis as $`\stackrel{}{m}`$ and orient it such that it is near the angular momentum axis of the gas disk. Since the disk rotates in a direction such that the eastern side is blueshifted with respect to the systemic velocity (Baum, Heckman, & van Breugel (1990); Bower et al. 1997b ), $`\stackrel{}{m}`$ should be pointing near the north. We describe $`\stackrel{}{m}`$ by a position angle on the sky $`\beta `$ and an inclination angle $`\theta `$ which is the angle between this axis and the line of sight ($`\theta =0`$ refers to a symmetry axis pointing towards us).
We assume that a gas ring with radius $`r`$, rotation velocity $`v_c`$ and angular momentum $`\stackrel{}{L}=rv_c\widehat{n}`$ precesses about $`\stackrel{}{m}`$. We define the inclination angle, $`\omega `$, of a gas ring with respect to the symmetry axis as the angle between $`\widehat{n}`$ and $`\stackrel{}{m}`$. The azimuthal or precession angle, $`\alpha `$, is defined as the angle in the equatorial plane (plane perpendicular to $`\stackrel{}{m}`$) of the projection of $`\widehat{n}`$ into this plane minus the angle of the projection of the line of sight. $`\alpha `$ increases in the direction opposite the rotation (e.g., see Fig. 6 of Quillen et al. (1993)). We can then describe the geometry of the disk with inclination and precession angles as a function of $`r`$;
$$\alpha (r)=\alpha _0(r)=\frac{d\alpha }{dt}\mathrm{\Delta }T+\mathrm{constant}$$
(8)
$$\omega (r)=\omega _0.$$
(9)
Here we have assumed a planar initial state for the gas disk, and $`\mathrm{\Delta }T`$ is the timescale since this initial condition. When the magnitude of the torque is known an estimate for the radial dependence of the azimuthal angle then yields an estimate for $`\mathrm{\Delta }T`$. When the torque is $`r^1`$ then $`\frac{d\alpha }{dt}r^1`$; this true for the torque caused by an oblate or prolate galaxy when the rotation curve is nearly flat. We can then write
$$\alpha _0(r)=B_\alpha \left(\frac{100pc}{r}\right)+\alpha _c.$$
(10)
To produce a model we integrate the stellar light of the galaxy taking into account the opacity caused by the warped disk. The surface brightness profile at $`1.6\mu `$m can be well fit in the inner $`20^{\prime \prime }`$ with a blend of two power laws (e.g. from Ferrarese et al. (1994))
$$I(r)=\sqrt{2}I_c\left(\frac{r_b}{r}\right)^{\beta _1}\left[1+\left(\frac{r}{r_b}\right)^{2(\beta _2\beta _1)}\right]^{1/2}$$
(11)
and fit as a function of semi-major axis. The break radius, $`r_b`$, is where the profile changes shape or can also be described as near a point of maximum curvature in log log coordinates. For our model we assume a light density in 3 dimensions which is consistent with the above profile,
$$\rho (r)s^{\gamma _1}\left(1+s^{2(\gamma _2\gamma _1)}\right)^{1/2}$$
(12)
where $`s\frac{1}{r_b}\sqrt{x^2+y^2+z^2/q^2}`$ and $`q`$ is the observed isophotal axis ratio. Here we found that $`\gamma _1=0.5`$ and $`\gamma _2=2`$ and the break radius measured at $`1.6\mu `$m ($`r_b=1.97\pm 0.15^{\prime \prime }160`$ pc, Quillen, Bower & Stritzinger (1999)) yields a pretty good fit to the observed surface brightness profile (with $`\beta _1=0.12\pm 0.04`$, $`\beta _2=1.30\pm 0.06`$). For the disk itself we assume a powerlaw form for the opacity of the disk (seen face on) $`\tau (r)=\tau _0\left(\frac{r}{1\mathrm{p}\mathrm{c}}\right)^{\tau _p}`$. We then compared the resulting integration qualitatively with the morphology of the disk in an optical band (WFPC2/PC image). For the models shown in Fig. 2 we used $`\tau _0=1`$ at F547M (0.55$`\mu `$m) and $`\tau _p=0.2`$.
Fig. 2 shows an example of such a model (referred to as Model 1) compared to the F547M band image. Qualitatively this model succeeds in having two self similar dust features which roughly correspond to the two broad features observed, one at large scales and one at smaller scales. The model, however does not reproduce the strong asymmetries. For example the galaxy dust features are wider on the eastern side than on the western side of the nucleus, but reach a minimum just north of the nucleus. The model on the other hand has a dust feature that is nearly constant width running from east of the nucleus to the north west. The observed dust features also extend more linearly to the west on the western side than the model does. A model which ‘wobbles’ would be required to exhibit these traits. By wobbling we mean that the ring angular momentum vector $`\stackrel{}{n}`$ could follow an elliptical path (corresponding to variations in $`\omega `$) which has an azimuthally varying precession rate about this path (corresponding to azimuthal variations in $`d\alpha /dt`$).
#### 3.1.1 A Geometrical description for wobble
We can increase the complexity of our geometric model with additional terms
$$\alpha (r)=\alpha _0(r)+A_{\alpha 1}\mathrm{cos}(\alpha _0\alpha _d)+A_{\alpha 2}\mathrm{sin}2(\alpha _0\alpha _d)$$
(13)
$$\omega (r)=\omega _0+A_\omega \mathrm{cos}2(\alpha _0\alpha _d)$$
(14)
where $`\alpha _0(r)`$ comes from our more simplistic model described above. $`\alpha _d`$ refers to the orientation angle of the maxima of the $`\alpha `$ azimuthal variations projected onto plane perpendicular to $`\stackrel{}{m}`$. The inclination term only contains a $`\mathrm{cos}2\alpha _0`$ dependence because an $`m=1`$ dependence can be removed with a redefinition for the symmetry axis $`\stackrel{}{m}`$. A model with these additional parameters (listed in Table 1, and referred to as Model 2) is also shown in Fig. 2. This model succeeds at reproducing the asymmetries of the observed dust features. Most of the features observed in the model are present in the galaxy. We infer that the dust disk is probably precessing in a way that is more complicated than described by the simple precession model (Eqns. 8-10) and which matches observations in galaxies such as Cen A (Quillen et al. (1993)) and NGC 4753 (Steiman–Cameron et al. (1992)). We now examine what kind of physical models can produce significant extra ‘wobble’ terms in Eqns. 13 and 14.
#### 3.1.2 Triaxial models
In the above discussion we have assumed that M84 is axisymmetric, and not triaxial. To reproduce the alignment of dusty features we require a model with symmetry axis that is not aligned with the galaxy isophotal major axis. When the galaxy is triaxial, the axes of symmetry do not coincide with the projected isophotal axes (e.g. de Zeeuw & Franx (1989)). Using equations for the projected angles listed in Appendix A of de Zeeuw & Franx (1989) and assuming a galaxy major axis inclination angle consistent with our fit to the warp geometry we find that either extreme axes ratios are required (one galaxy axis ratio lower than 0.7) or only a very small region (less than $`10^{}`$ in the azimuthal projection angle $`\varphi `$ of de Zeeuw & Franx (1989)) of possible galaxy projection angles is allowed, making it unlikely. The lack of galaxy rotation ($`<8`$km/s; Davies & Birkinshaw (1988)) makes it unlikely that the galaxy is tumbling (as explored in van Albada, Kotanyi, & Schwarzschild (1982)).
Additional dynamical concerns also imply that the central region of M84 should not be triaxial. Stochasticity induced by the black hole (Merritt & Valuri (1996)) should cause triaxiality to be short lived (of order the local dynamical time) in the central regions of the galaxy. There is little twist in the isophotes of M84 (van den Bosch et al. (1994), Peletier et al. (1990)) which should be observed if the galaxy is strongly triaxial and not oriented with an axis coincident with the line of sight. We also note that the warped disk appearance and its misalignment with the galaxy major axis persists to well within the break radius $`2^{\prime \prime }160`$ pc of the nucleus where the ellipticity is slightly reduced and the slight boxiness disappears (as seen in the NICMOS $`1.6\mu `$m isophotes; Quillen, Bower & Stritzinger (1999)). Mechanisms that reduce ellipticity and boxiness (such as scattering) are likely to destroy triaxiality as well. In addition van Dokkum & Franx (1995) found that triaxiality could not explain the observed disk/galaxy misalignments in a sample of elliptical galaxies. Whereas precession times are similar, disk settling times in triaxial galaxies are substantially faster than in axisymmetric systems (Habe & Ikeuchi (1985)).
We can also determine if a triaxial model is viable based on the size and form of the ‘wobble’ terms introduced above, which are probably needed to describe the disk morphology. We associate the symmetry axis of the disk $`\stackrel{}{m}`$ with one symmetry axis of the triaxial galaxy. The amplitude of variation in inclination angle described by $`A_\omega `$ can be estimated from the azimuthal variation in the torque on a gas ring as the ring precesses. This results in a torque that causes the ring to change inclination ($`\omega `$) rather than precession rate. This amplitude should be roughly equivalent to the size of the ellipticity of the gravitational potential in the plane perpendicular to $`\stackrel{}{m}`$ times the angle $`\omega `$ (for $`\omega `$ small). Since this angle is small in M84 and the ellipticity of the potential should be $`1/3`$ the axis ratio of the density in this plane, $`A_\omega `$ should be quite small (we estimate less than a degree). This is far less than exhibited in our Model 2 described above. In a triaxial galaxy it makes sense that this amplitude should be small because the gravitational potential is always smoother than the actual density distribution. For a triaxial galaxy we also expect no $`\mathrm{sin}(\alpha _0)`$ variation in $`d\alpha /dt`$ (or $`A_{\alpha 1}`$ term) which were a part of Model 2 discussed above. Dynamical models which neglect triaxiality for the disk kinematics and morphology in Centaurus A (e.g. Sparke (1996), Quillen et al. (1993)) are successful despite the evidence that this galaxy probably is triaxial (Hui et al. (1995)). This supports our statement that triaxial galaxies are unlikely to result in disk precession models with large ‘wobble’ terms.
### 3.2 Warp model including a pressure torque
Above we proposed that the ambient X-ray emitting ISM could exert sufficient torque on the gas disk to affect the orientation of the gas disk. Here we elaborate on this possibility and search for a model that can accurately match the inferred geometry of the gas disk.
The disk rotates in a direction such that the eastern side is blueshifted with respect to the systemic velocity (Baum et al. (1990); Bower et al. 1997b ). The gas disk, if warped, has orientation such that with increasing radius the azimuthal angle of greatest disk inclination moves counter (retrograde) to the direction of rotation. This orientation would be consistent with an oblate galaxy potential (if the warp were indeed caused by the torque from the galaxy; e.g., Tubbs (1980)). For an oblate galaxy potential the shape of the potential causes a force towards the galactic plane of symmetry, resulting in a ‘retrograde’ warp. If higher pressures exist in the ISM along the jet axis then the pressure gradient in this gas would exert a force on the gas disk towards the plane perpendicular to the jets. When the precession period decreases with increasing radius this again would cause a ‘retrograde’ warp but about a symmetry axis aligned with the jet rather than a galactic axis of symmetry.
We can estimate the torque resulting from the hot ISM as follows: As typically assumed for a non-axisymmetric galaxy, we can describe the pressure in the volume filling X-ray emitting ISM as having an the axis of symmetry, $`\widehat{p}`$, of the isobars in the hot gas. We chose to orient $`\widehat{p}`$ with a sign such that it is nearest the average angular momentum vector of the disk. The pressure gradient across a gas disk with mass surface density $`\mathrm{\Sigma }`$ and thickness $`h`$ (corresponding to the vertical scale height) results in an average torque per unit mass (around a ring of radius $`r`$)
$$\left|\tau _p\right|=\frac{h}{2\mathrm{\Sigma }}\frac{P}{\theta }$$
(15)
where the direction of the torque on the ring is given by $`\widehat{p}\times \widehat{n}`$.
The average torque per unit mass from a non-axisymmetric gravitational potential can be determined similarly with
$$\stackrel{}{\tau _g}=ϵ_\mathrm{\Phi }v_c^2\mathrm{cos}\theta _g(\widehat{n}\times \widehat{g})$$
(16)
where we have assumed a $`\mathrm{cos}(2\theta _g)`$ form for the non-axisymmetric part of the gravitational potential. Here $`\theta _g`$ refers to the ring inclination with respect to the axis of symmetry of the galaxy, or the angle between $`\widehat{n}`$ and $`\widehat{g}`$, for $`\widehat{g}`$ the galaxy symmetry axis and $`ϵ_\mathrm{\Phi }`$ is the ellipticity of the gravitational potential. The total average torque per unit mass on a gas ring would then be given by the sum of the two torques $`\stackrel{}{\tau }=\stackrel{}{\tau _g}+\stackrel{}{\tau _p}`$.
Precession of the gas disk takes place approximately about a vector
$$\stackrel{}{m}=\frac{P}{\theta }\frac{h}{2\mathrm{\Sigma }}\widehat{p}+ϵ_\mathrm{\Phi }v_c^2\widehat{g}$$
(17)
where $`\frac{P}{\theta }`$ is estimated at the average angle between $`\widehat{n}`$ and $`\widehat{p}`$. We can define orientation angles for the gas disk as a function of radius with respect to this vector $`\stackrel{}{m}`$. As in §3.1 we define the inclination angle, $`\omega `$, and $`\alpha `$ for a ring with angular momentum axis $`\widehat{n}`$ with respect to the precession axis $`\stackrel{}{m}`$. The total torque is
$$\stackrel{}{\tau }\stackrel{}{m}\times \widehat{n}$$
(18)
and the angular precession rate about $`\stackrel{}{m}`$ is
$$\frac{d\alpha }{dt}\frac{\stackrel{}{m}\widehat{n}}{rv_c}$$
(19)
When the galaxy is oblate the gas disk will precess about an axis intermediate between the galaxy axis of symmetry ($`\widehat{g}`$) and the pressure axis of symmetry ($`\widehat{p}`$, which we associate with the jet axis). However when the galaxy is prolate the sign of the galactic torque term changes (equivalent to setting $`ϵ_\mathrm{\Phi }`$ to be negative) and precession could occur about an axis which is not obviously between the two symmetry axes.
Because of the sense of the warp, in M84 we expect that the galaxy is nearly oblate and so that the gas disk should precess about an axis intermediate between the two symmetry axes. In fact we can measure the orientation of this axis and from it estimate the size of the pressure torque compared to the the galactic term (assuming that the galaxy is not strongly triaxial).
#### 3.2.1 Ratio of pressure to galactic torque
In M84 we observe that the dust features are aligned at an angle (PA$`=6580^{}`$) intermediate between the jet axis (PA$`=0`$; Jones et al. (1981)) and the galaxy major axis (PA$`=129^{}`$, van Dokkum & Franx (1995)). Because the north and south jets are have similar intensities we assume that the jet axis is perpendicular to the line of sight. We find a rough fit to the warp with a symmetry axis with PA between $`10`$ and $`15^{}`$. So the angle between the jet and $`\stackrel{}{m}`$ is about $`1015^{}`$ and the angle between $`\stackrel{}{m}`$ and the galaxy axis is about $`3540^{}`$. This suggests that the torque from the pressure is larger than the galactic torque so that
$$\frac{\tau _g}{\tau _p}24$$
(20)
This lets us estimate the size of the azimuthal pressure component. For $`\frac{P}{\theta }ϵ_pP_0(r)`$
$`P_0(r)`$ $``$ $`3{\displaystyle \frac{ϵ_\mathrm{\Phi }v_c^2\mathrm{\Sigma }}{ϵ_ph}}`$
$``$ $`3\times 10^{10}\mathrm{dynes}\mathrm{cm}^2\left({\displaystyle \frac{ϵ_\mathrm{\Phi }}{0.05}}\right)\left({\displaystyle \frac{ϵ_p}{1.00}}\right)^1\left({\displaystyle \frac{v_c}{300\mathrm{k}\mathrm{m}\mathrm{s}^1}}\right)^2\left({\displaystyle \frac{\mathrm{\Sigma }}{2M_{}\mathrm{pc}^2}}\right)\left({\displaystyle \frac{h/r}{0.1}}\right)^1\left({\displaystyle \frac{r}{100\mathrm{p}\mathrm{c}}}\right)^1.`$
Here $`ϵ_p`$ represents the ellipticity of the isobars. We have adopted a range for the surface density $`\mathrm{\Sigma }`$ based on the extinction in the dust features, van Dokkum & Franx (1995) and Bower et al. (1997a) estimate a total disk mass of $`10^6M_{}`$ and $`9\times 10^6M_{}`$ respectively. An estimate for the disk mass based on the NICMOS images is consistent with the low end $`10^6M_{}`$. For a disk of constant surface density (consistent with a small $`\tau _p`$, see §3.1) truncated at $`r=400`$ pc the lower of these mass estimates give $`\mathrm{\Sigma }2M_{}\mathrm{pc}^2`$.
Although there are uncertainties in the parameters, the pressure we estimate above is similar to the thermal pressure we estimated from the ambient volume filling X-ray emission (see Eqn. 7). The alignment and precession of the disk therefore could be consistent with the model proposed here where a pressure exerted by the X-ray medium (aligned with the jet) and the non-axisymmetric galaxy both exert torques on the gas disk.
#### 3.2.2 Warp timescale
As mentioned in §3.1, once the size and radial form of the torque is known, the radial dependence of $`\alpha `$ can be used to estimate a timescale. Here we have assumed a planar initial state for the gas disk, and $`\mathrm{\Delta }T`$ is the timescale since this initial condition (see Eqn. 2). For a warp caused purely by torque from the galaxy $`d\alpha /dtϵ_\mathrm{\Phi }\mathrm{\Omega }r^1`$ when the rotation curve is nearly flat. From our modeling we estimate $`B_\alpha =1.2\pm 0.3\times 10^3`$ radians pc yielding an estimate of
$$\mathrm{\Delta }T=1.8\times 10^7\mathrm{yr}\left(\frac{B_\alpha }{1.2\times 10^3\mathrm{radians}\mathrm{pc}}\right)\left(\frac{ϵ_\mathrm{\Phi }}{0.05}\right)^1\left(\frac{v_c}{300\mathrm{k}\mathrm{m}\mathrm{s}^1}\right)^1\left(\frac{\tau _p/\tau _g}{3}\right)^1$$
(22)
This suggests that precession has occurred only about $`10^7`$ years since an event responsible for distributing the gas and dust, or causing the present orientation of the jets. Our model places time limits on the stability of the jet orientation. Here we link the morphology of the disk to the history of pressure asymmetries caused by the jets themselves. If the jets were at a different angle previously (as suggested from the large scale radio emission) then the large scale morphology of the disk could be linked to the past jet orientation. An alternate mechanism is then required to change the jet orientation in a way not related to the disk orientation (for example as proposed by Roos, Kaastra, & Hummel (1993) involving binary black holes).
#### 3.2.3 Wobble
In general the sum of the two torques will lead to an elliptical path (or ‘wobble’) for the angular momentum vector of a gas ring, $`\widehat{n}`$, about an axis $`\stackrel{}{m}`$ which also varies in angular precession rate. This wobble corresponds to the variation in torque that a ring feels as it precesses. From a measurement of this ‘wobble’ we can determine if the pressure gradient is strongly dependent on $`\theta _p`$, the angle between $`\widehat{n}`$ and $`\widehat{p}`$. As in the case of a triaxial galaxy, for a $`\mathrm{cos}2\theta _p`$ dependence we would expect that the wobble about simple precession should be small (less than a degree). However we achieve a better correspondence to the observed morphology with larger azimuthal variations in $`\omega `$ and $`d\alpha /dt`$ suggesting that the pressure is strongly dependent on the inclination angle $`\theta _p`$ from the jet axis.
We now estimate the difference in the torque as the angular momentum axis precessed about $`\stackrel{}{m}`$ As the ring precesses variation in the torque reach extremes nearest and furthest from the jet axis of the size
$$\mathrm{\Delta }\tau \left(\frac{\tau _p}{\theta }\frac{\tau _g}{\theta }\right)\omega .$$
(23)
This results in a variation in $`\frac{d\alpha }{dt}`$ that is proportional to $`\mathrm{cos}(\alpha _0)`$ (described in our model by $`A_{\alpha 1}`$). The component of $`\mathrm{\Delta }\tau `$ in the plane perpendicular to $`\stackrel{}{m}`$ causes the ring to change inclination and is about the same size as $`\mathrm{\Delta }\tau `$ times an additional factor of $`\omega `$ and results in a $`\mathrm{cos}(2\alpha _0)`$ dependence of $`\omega `$ which we described in our model by $`A_\omega `$. (The $`\mathrm{cos}(\alpha _0)`$ dependence can be removed by a redefinition of $`\stackrel{}{m}`$). Here our estimates have assumed that $`\omega `$ is small.
If both the pressure and gravitational torques show the same angular dependence (e.g. $`\mathrm{cos}(2\theta )`$) then the dependence of $`\alpha `$ cancels to first order in the angles $`\theta _g`$ and $`\theta _p`$ and $`\frac{d\alpha }{dt}`$ varies only to second order in $`\omega `$. The dependence on $`\alpha `$ can be approximated as
$$A_{\alpha 1}\omega _0(m^24)\mathrm{cos}(\alpha _0)$$
(24)
where $`m`$ refers to the angular dependence of $`\frac{P}{\theta }`$ on $`\theta _p`$ and the variation in $`\omega `$ is $`A_\omega A_{\alpha 1}\omega `$. $`\alpha (r)`$ is then described by Eqns. 8, 10 and 13 and $`\omega (r)`$ is described by Eqn. 14. The precession rate is expected to be fastest when the ring angular momentum vector is furthest from the pressure or jet axis, $`\widehat{p}`$, and nearest $`\widehat{g}`$, the galaxy symmetry axis. From this we can see that large azimuthal variations in the angular precession rate and inclination could be symptoms of a strong angular dependence of the isobars. The signs of $`A_\omega `$ and $`A_{\alpha 1}`$ in our Model 2 (see Table 1) suggests that the galaxy axis (to the north-west on the sky) is pointing somewhat towards us with the angle between it (projected on to the planet perpendicular to $`\stackrel{}{m}`$) and the line of sight $`|\alpha _d|50^{}`$.
#### 3.2.4 A note
We have ignored the possibility that the isobars might actually be affected by the galaxy potential. This situation would arise naturally to some extent because as energy is lost from the jet into the ISM energy would be transfered in the direction of least gravitational force, i.e. towards the galactic major axis. However if hydrostatic equilibrium applies, the sound speed of the gas sets a particular scale length in the hot gas whereby $`\frac{h}{r}\frac{c_s}{v_c}`$. Because the sound speed in the hot gas $`c_sv_c300`$ km/s hydrostatic equilibrium would set $`h/r1`$ and so the isobars are unlikely to be significantly affected by the gravitational potential, and are more likely to be determined by the nature of the dissipation from the jet into the ISM.
## 4 Summary and Discussion
In this paper we have estimated the timescale of a gas disk to precess in the non-spherical gravitational potential of M84. This timescale is a few times $`10^7`$ years at 100 pc where the dust features are misaligned with the galaxy isophotes. For the disk to remain misaligned with the galaxy potential some mechanism must operate faster than this. While a cooling flow could replenish the disk on this timescale it is difficult to explain why the disk is at a roughly constant angle within $`r<7^{\prime \prime }`$ and yet twists at this radius forming an overall S-shape in the H$`\alpha `$ \+ \[NII\] emission (Baum et al. (1988)). Extremely fast accretion through the disk itself would require a substantial gas reservoir at large radii or a very short disk lifetime. It would also require a place to put the excess accreted gas mass, such as a wind, an inner disk or advection-dominated accretion into the black hole. A combination of fast accretion and replenishment by a cooling flow could possibly result in inflow faster than the precession rate, but it is not clear whether this combination could account for the disk morphology. The AGN is not luminous enough for the radiative induced warp mechanism of Pringle (1996) to operate. None of these possibilities would provide a good explanation for the observed morphology of the gas and dust disk.
As an alternative to these external mechanisms we consider the possibility of a local force on the disk. We estimate the pressure required to overcome the torque from the galaxy and find that it is small compared to the thermal pressure inferred from X-ray observations. Pressure gradients in this ambient hot ISM could therefore overcome the galaxy torque. We therefore propose that pressure gradients in an energetic low density medium in M84 could strongly affect the orientation of the gas disk on the scale of a few hundred pc.
By integrating the light of the galaxy through a dusty warped disk we find that the gas disk in M84 is likely to differ from a simple precession model where the precession rate is constant with azimuthal angle and the angular momentum axis of a gas ring traces a circular path. A triaxial model for the galaxy, though it would explain the misalignment of the dust features with the galaxy isophotal major axis, is not a good explanation for a variety of reasons. Because of the expected weak dependence of the gravitational torque on the position angle wobble of the angular momentum axis of a gas ring is likely to be smaller than that we infer from the disk geometry. The misalignment also persists to small radii where the galaxy is expected to be oblate and not triaxial. We propose instead that the morphology of the gas disk in M84 is consistent with a warped geometry where precession is caused by a combination of a galactic torque and a torque due to pressure gradients in the ambient X-ray emitting gas. The alignment of the disk can be used to estimate the ratio of these torques. Precession occurs at an axis between the jet and galaxy major axis, but nearer to the jet axis. This angle implies that the pressure torque is 2-4 times larger than the galactic torque. A rough model to the morphology of gas disk also allows us to estimate the degree of precession that has taken place. Assuming the initial condition of gas in a plane we estimate the timescale since then to be a few times $`10^7`$ years. A better model to the morphology of the disk is achieved when precession takes place about an elliptical rather than circular path and precesses fastest when the angular momentum vector is furthest from the jet axis. This suggests that the isobars are strongly dependent on angle from the jet axis.
Recent investigations find that almost all ellipticals have dust (van den Bosch et al. (1994)). In cases of non-active elliptical galaxies having dusty disks misaligned with the galaxy isophotes, the misalignment must be caused by another mechanism (perhaps a cooling flow or a galaxy merger). van Dokkum & Franx’s (1995) sample contains only five galaxies with radio power $`\mathrm{log}P_\nu `$ (6 cm)$`<20`$ (W/Hz) and misaligned dust features. Further investigation is needed with a larger sample to determine how common these non-AGN misaligned cases are and if the models proposed here are applicable. One consequence of our proposed mechanism is that on very long timescales we expect the disk to become multiply warped or rippled. On a timescale of roughly 10 times the precession time we would then expect the disk to settle into a quasi-stationary surface nearly perpendicular to the jet axis. This might account for the observed alignments between jets and dust features. Comparison of high resolution X-ray morphology (such as will be possible with AXAF), dust morphologies and ionized gas kinematics should determine if the type of models introduced here are appropriate.
We acknowledge helpful discussions and correspondence with E. Emsellem, A. Eckart, R. Green, G. Rieke, M. Rieke, G. Schmidt, D. Hines, P. Pinto, D. DeYoung and F. Melia. Support for this work was provided by NASA through grant number GO-07868.01-96A from the Space Telescope Institute, which is operated by the Association of Universities for Research in Astronomy, Incorporated, under NASA contract NAS5-26555. ACQ also acknowledges support from NSF grant AST-9529190 to M. and G. Rieke and NASA project no. NAG-53359. GB acknowledges support from the STIS Investigation Definition Team.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.