cond-mat0604552/cm2.tex
1: \documentclass[prb,tighten]{revtex4}
2: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3: %\documentclass[prb,eqsecnum,preprint]{revtex4}
4: %\documentclass[prb,twocolumn,eqsecnum,showpacs]{revtex4}
5: %\documentclass[pre,preprint,draft,showpacs,tighten]{revtex4}
6: %\usepackage[spanish]{babel}
7: %\usepackage[cp850]{inputenc}
8: %\usepackage[latin1]{inputenc}
9: 
10: \usepackage{amssymb,amsmath}
11: 
12: \usepackage{epsfig}
13: %\usepackage{showkeys}
14: % Muestra etiquetas de formulas, figuras, referencias,...
15: 
16: \usepackage{graphicx}% Include figure files
17: 
18: %\usepackage{dcolumn}% Align table columns on decimal point
19: %\usepackage{bm}% bold math
20: %\psdraft
21: 
22: \begin{document}
23: \title{Navier-Stokes transport coefficients of $d$-dimensional granular binary mixtures
24: at low density}
25: \author{Vicente Garz\'{o}\footnote[1]{Electronic address: vicenteg@unex.es;
26: URL: http://www.unex.es/eweb/fisteor/vicente/}}
27: \affiliation{Departamento de F\'{\i}sica, Universidad de
28: Extremadura, E-06071 Badajoz, Spain}
29: \author{Jos\'e Mar\'{\i}a
30: Montanero\footnote[2] {Electronic address: jmm@unex.es}}
31: \affiliation{Departamento de Electr\'onica e Ingenier\'{\i}a
32: Electromec\'anica, Universidad de Extremadura, E-06071 Badajoz,
33: Spain}
34: 
35: \begin{abstract}
36: The Navier-Stokes transport coefficients for binary mixtures of smooth inelastic hard disks or spheres under
37: gravity are determined from the Boltzmann kinetic theory by application of the Chapman-Enskog method for states
38: near the local homogeneous cooling state. It is shown that the Navier-Stokes transport coefficients are not
39: affected by the presence of gravity. As in the elastic case, the transport coefficients of the mixture verify a
40: set of coupled linear integral equations that are approximately solved by using the leading terms in a Sonine
41: polynomial expansion. The results reported here extend previous calculations [V. Garz\'o and J. W. Dufty, Phys.
42: Fluids {\bf 14}:1476--1490 (2002)] to an arbitrary number of dimensions and provide explicit expressions for the
43: seven Navier-Stokes transport coefficients in terms of the coefficients of restitution and the masses,
44: composition, and sizes of the constituents of the mixture. In addition, to check the accuracy of our theory, the
45: inelastic Boltzmann equation is also numerically solved by means of the direct simulation Monte Carlo method to
46: evaluate the diffusion and shear viscosity coefficients for hard disks. The comparison shows a good agreement
47: over a wide range of values of the coefficients of restitution and the parameters of the mixture (masses and
48: sizes).
49: 
50: {\bf KEY WORDS}: Granular binary mixtures; inelastic Boltzmann
51: equation; Navier-Stokes transport coefficients; DSMC method.
52: 
53:  Running title: Navier-Stokes transport coefficients
54: 
55: \end{abstract}
56: 
57: 
58: %\pacs{05.20.Dd, 45.70.Mg, 51.10.+y}
59: \draft
60: \date{\today}
61: \maketitle
62: 
63: \section{Introduction}
64: \label{sec1}
65: 
66: 
67: The simplest model for a granular fluid is a system composed by
68: smooth hard spheres or disks with inelastic collisions. The only
69: difference from the corresponding model for normal fluids is the
70: loss of energy in each binary collision, characterized by a
71: (constant) coefficient of normal restitution. For a low density
72: gas, the Boltzmann kinetic equation conveniently modified to
73: account for inelastic collisions \cite{GS95,BDS97,BP04} has been
74: used in recent years as the starting point to derive the
75: hydrodynamic-like equations of the system. Thus, assuming the
76: existence of a {\em normal} (hydrodynamic) solution for
77: sufficiently long space and time scales, the Chapman-Enskog (CE)
78: method \cite{CC70} has been applied to solve the Boltzmann
79: equation to Navier-Stokes (NS) order and get explicit expressions
80: for the transport coefficients. While this goal has been widely
81: covered in the case of a monocomponent gas, \cite{simple} much
82: less is known for systems composed by grains of different masses,
83: sizes, and concentrations (granular mixtures).
84: 
85: 
86: Needless to say, the determination of the NS transport
87: coefficients for a multicomponent granular fluid is much more
88: complicated than for a single granular system. Many attempts to
89: derive these coefficients \cite{mixture} have been carried out by
90: means of the CE expansion around Maxwellians at the {\em same}
91: temperature $T$ for each species. The use of this distribution can
92: only be considered as acceptable for nearly elastic systems where
93: the assumption of the equipartition of energy still holds. In
94: addition, according to this level of approximation, the
95: inelasticity is only accounted for by the presence of a sink term
96: in the energy balance equation and so the expressions of the NS
97: transport coefficients are the same as those obtained for elastic
98: collisions. However, as the dissipation increases, different
99: species of a granular mixture have different partial temperatures
100: $T_i$ and consequently, the energy equipartition is seriously
101: broken ($T_i\neq T$). The failure of energy equipartition in
102: granular fluids \cite{GD99b,MP99} has also been confirmed by
103: computer simulations \cite{computer} and even observed in real
104: experiments \cite{exp1} of agitated mixtures. All the results show
105: that deviations from equipartition depend on the mechanical
106: differences between the particles of each species and the
107: coefficients of restitution of the system. Given that the
108: inclusion of nonequipartition effects increases the level of
109: complexity of the problem, it is interesting from a practical
110: point of view to assess the influence of this effect on the
111: transport properties of the system. If the NS transport
112: coefficients turned out to be quite sensitive to nonequipartition,
113: the predictions made from previous theories \cite{mixture} should
114: be reexamined by theories that take into account the
115: nonequipartition of energy. For this reason, although the
116: possibility of nonequipartition was already pointed out many years
117: ago, \cite{JM87} a careful study of its influence on transport has
118: only been carried out recently. In this context, Garz\'o and Dufty
119: \cite{GD02} have developed a kinetic theory for a binary granular
120: mixture of inelastic hard spheres at low density which accounts
121: for nonequipartition effects. Their results show that in general
122: the consequences of the temperature differences for the transport
123: coefficients are quite significant, especially for strong
124: dissipation.\cite{GA05,MGD06} It is important to remark that the
125: expressions derived in Ref.\ \onlinecite{GD02} for the
126: Navier-Stokes transport coefficients do not limit their
127: application to weak inelasticity. In fact, the results reported in
128: this paper include a domain of both weak and strong inelasticity,
129: $0.5 \leq \alpha \leq 1$, where $\alpha$ is the (common)
130: coefficient of restitution considered. The accuracy of these
131: theoretical predictions (based on a Sonine polynomial expansion)
132: has been confirmed by Monte Carlo simulations of the inelastic
133: Boltzmann equation in the cases of the diffusion coefficient
134: \cite{GM04} and the shear viscosity coefficient of a mixture
135: heated by an external thermostat. \cite{MG03} Exceptions to this
136: good agreement are extreme mass or size ratios and strong
137: dissipation, although these discrepancies are mitigated in part if
138: one retains more terms in the Sonine polynomial expansion.
139: \cite{GM04} For small dissipation, the results derived by Garz\'o
140: and Dufty \cite{GD02} agree with those recently obtained by Serero
141: {\em et al.}\cite{SGNT06} in the first order of the order
142: parameter $\epsilon_{ij}=1-\alpha_{ij}^2$.
143: 
144: The CE method solves the Boltzmann equation by expanding the
145: distribution function of each species $f_i({\bf r}, {\bf v},t)$
146: around the {\em local} homogeneous cooling state (HCS).
147: \cite{GD99b} This state plays the same role for granular gases as
148: the local equilibrium distribution for a gas with elastic
149: collisions. Given that the form of the distribution function
150: $f_i^{(0)}$ of the HCS is not exactly known, one usually considers
151: the first correction to a Maxwellian at the temperature for that
152: species, namely, a polynomial in velocity of degree four (leading
153: Sonine correction). However, the results derived for hard spheres
154: clearly show that the influence of these non-Gaussian
155: contributions to the transport coefficients are in general
156: negligible, except in the case of the heat flux for quite large
157: values of dissipation. \cite{MGD06} Accordingly, a theory
158: incorporating the contributions coming from the deviations of the
159: HCS from its Gaussian form does not seem necessary in practice for
160: computing the NS transport coefficients of the mixture.
161: 
162: 
163: The objective of this paper is twofold. First, given that the
164: results reported in Ref.\ \onlinecite{GD02} are limited to hard
165: spheres, we extend here this derivation to an arbitrary number of
166: dimensions $d$. This goal is not only academic since, from a
167: practical standpoint, many of the experiments reported for flowing
168: granular materials have created (quasi) {\em two-dimensional}
169: systems by confining grains between vertical or on a horizontal or
170: tilted surface, enabling data collection by high-speed video.
171: \cite{exp} Regarding computer simulations, most of them consider
172: hard disks to save computer time and memory. For these reasons, it
173: would be desirable to provide experimentalists and simulators with
174: theoretical tools to work when studying problems both in two and
175: three dimensions. In addition, apart from its practical interest,
176: it is also interesting from a fundamental view to explore what is
177: the influence of dimensionality on the dependence of the transport
178: coefficients on dissipation. As a second target, we want also to
179: present a simplified theory with explicit expressions for the
180: transport coefficients. As the algebra involved in the
181: calculations of Ref.\ \onlinecite{GD02} is complex, the
182: constitutive relations for the fluxes were not explicitly
183: displayed in this paper. Although the work carried out here
184: involves complex algebra as well, the use of Maxwellians at
185: different temperatures for the distribution functions of each
186: species in the reference state allows us to explicitly obtain
187: expressions for the {\em seven} relevant transport coefficients of
188: the mixture in terms of the mechanical parameters of the system:
189: masses, sizes, composition and coefficients of restitution. To
190: assess the degree of accuracy of our (approximated) expressions,
191: we have also performed Monte Carlo simulations for the diffusion
192: and the shear viscosity coefficients for hard disks ($d=2$). As
193: shown below, the good agreement found between the results derived
194: in this paper with computer simulations justifies this
195: simplification and allows one to obtain more simplified forms of
196: the transport coefficients.
197: 
198: 
199: 
200: The plan of the paper is as follows. In Sec.\ \ref{sec2}, the
201: inelastic Boltzmann equation and the corresponding hydrodynamic
202: equations are recalled. The CE expansion adapted to the inelastic
203: binary mixtures is formulated in Sec.\ \ref{sec2bis}. Assuming
204: that gradients and dissipation are independent parameters, the
205: Boltzmann equation is solved by means of an expansion in powers of
206: the spatial gradients around the local HCS distribution
207: $f_i^{(0)}$. It is shown that the use of the local HCS as the
208: reference state is not an assumption of the CE method but a
209: consequence of the exact solution to the Boltzmann equation in the
210: zeroth-order approximation. Section \ref{sec3} deals with the
211: expressions for the NS transport coefficients. As in the case of
212: elastic collisions, these coefficients are the solutions of a set
213: of coupled linear integral equations which involves the (unknown)
214: distributions $f_i^{(0)}$. The integral equations are solved by
215: considering two approximations: First, $f_i^{(0)}$ is replaced by
216: its Maxwellian form at the temperature $T_i$, and second, only the
217: leading terms in a Sonine polynomial expansion of the first-order
218: distribution $f_i^{(1)}$ are retained. Technical details of the
219: calculations carried out here are given in Appendices \ref{appA},
220: \ref{appB}, and \ref{appC}. A comparison with previous results
221: \cite{SGNT06} based on the use of Maxwellians at the same
222: temperature $T$ as a ground state is also illustrated in Sec.\
223: \ref{sec3}, showing significant discrepancies between both
224: descriptions at moderate dissipation. Section \ref{sec4} is
225: devoted to the numerical solutions of the Boltzmann equation by
226: using the direct simulation Monte Carlo (DSMC) method \cite{Bird}
227: in the cases of the diffusion $D$ and shear viscosity $\eta$
228: coefficients for hard disks. To the best of our knowledge, this is
229: the first time that the NS shear viscosity of a granular binary
230: mixture at low density has been numerically obtained from the DSMC
231: method. The paper is closed in Sec.\ \ref{sec5} with a brief discussion of the
232: results presented in this paper.
233: 
234: 
235: 
236: \section{Boltzmann equation and conservation laws}
237: \label{sec2}
238: 
239: Consider a binary mixture composed by smooth inelastic disks
240: ($d=2$) or spheres ($d=3$) of masses $m_{1}$ and $ m_{2} $, and
241: diameters $\sigma _{1}$ and $\sigma _{2}$. The inelasticity of
242: collisions among all pairs is characterized by three independent
243: constant coefficients of restitution $\alpha _{11}$, $\alpha
244: _{22}$, and $\alpha _{12}=\alpha _{21}$, where $\alpha _{ij}\leq
245: 1$ is the coefficient of restitution for collisions between
246: particles of species $i$ and $j$. The mixture is in presence of
247: the gravitational field so that each particle feels the action of
248: the force ${\bf F}_i=m_i{\bf g}$, where ${\bf g}$ is the gravity
249: acceleration. In the low density regime, the distribution
250: functions $f_{i}({\bf r},{\bf v};t)$ $(i=1,2)$ for the two species
251: are determined from the set of nonlinear Boltzmann equations
252: \cite{BDS97}
253: \begin{equation}
254: \left( \partial _{t}+{\bf v}\cdot\nabla+{\bf g} \cdot
255: \frac{\partial}{\partial {\bf v}} \right) f_{i}({\bf r},{\bf
256: v},t)=\sum_{j=1}^2 J_{ij}\left[ {\bf v}|f_{i}(t),f_{j}(t)\right]
257: \;, \label{2.1}
258: \end{equation}
259: where the Boltzmann collision operator $J_{ij}\left[ {\bf
260: v}|f_{i},f_{j}\right] $ is
261: \begin{eqnarray}
262: J_{ij}\left[ {\bf v}_{1}|f_{i},f_{j}\right] &=&\sigma
263: _{ij}^{d-1}\int d{\bf v} _{2}\int d\widehat{\boldsymbol {\sigma
264: }}\,\Theta (\widehat{{\boldsymbol {\sigma }}} \cdot {\bf
265: g}_{12})(\widehat{\boldsymbol {\sigma }}\cdot {\bf g}_{12})
266: \nonumber
267: \\
268: &&\times \left[ \alpha _{ij}^{-2}f_{i}({\bf r},{\bf v}_{1}^{\prime
269: },t)f_{j}( {\bf r},{\bf v}_{2}^{\prime },t)-f_{i}({\bf r},{\bf v}
270: _{1},t)f_{j}({\bf r}, {\bf v}_{2},t)\right] \;. \label{2.2}
271: \end{eqnarray}
272: In Eq.\ (\ref{2.2}), $d$ is the dimensionality of the system,
273: $\sigma _{ij}=\left( \sigma _{i}+\sigma _{j}\right) /2$,
274: $\widehat{\boldsymbol {\sigma}}$ is an unit vector along the line
275: of centers, $\Theta $ is the Heaviside step function, and ${\bf
276: g}_{12}={\bf v}_{1}-{\bf v}_{2}$ is the relative velocity. The
277: primes on the velocities denote the initial values $\{{\bf
278: v}_{1}^{\prime }, {\bf v}_{2}^{\prime }\}$ that lead to $\{{\bf
279: v}_{1},{\bf v}_{2}\}$ following a binary (restituting) collision:
280: \begin{equation}
281: {\bf v}_{1}^{\prime }={\bf v}_{1}-\mu _{ji}\left( 1+\alpha
282: _{ij}^{-1}\right) (\widehat{{\boldsymbol {\sigma }}}\cdot {\bf
283: g}_{12})\widehat{{\boldsymbol {\sigma }}} ,\nonumber\\
284: \end{equation}
285: \begin{equation}
286:  {\bf v}_{2}^{\prime }={\bf v}_{2}+\mu _{ij}\left( 1+\alpha
287: _{ij}^{-1}\right) (\widehat{{\boldsymbol {\sigma }}}\cdot {\bf
288: g}_{12})\widehat{ \boldsymbol {\sigma}} ,  \label{2.3}
289: \end{equation}
290: where $\mu _{ij}\equiv m_{i}/\left( m_{i}+m_{j}\right) $. The
291: relevant hydrodynamic fields are the number densities $n_{i}$, the
292: flow velocity $ {\bf u}$, and the temperature $T$. They are
293: defined in terms of moments of the distributions $f_{i}$ as
294: \begin{equation}
295: n_{i}=\int d{\bf v}f_{i}({\bf v})\;,\quad \rho {\bf u}=\sum_{i=1}^2m_{i}
296: \int d {\bf v}{\bf v}f_{i}({\bf v})\;,
297: \label{2.4}
298: \end{equation}
299: \begin{equation}
300: nT=p=\sum_{i=1}^2 n_i T_i=\sum_{i=1}^2\frac{m_{i}}{d}\int d{\bf
301: v}V^{2}f_{i}({\bf v})\;, \label{2.5}
302: \end{equation}
303: where ${\bf V}={\bf v}-{\bf u}$ is the peculiar velocity, $
304: n=n_{1}+n_{2}$ is the total number density, $\rho
305: =m_{1}n_{1}+m_{2}n_{2}$ is the total mass density, and $p$ is the
306: pressure. Furthermore, the third equality of Eq.\ (\ref{2.5})
307: defines the kinetic temperatures $T_i$ for each species, which
308: measure their mean kinetic energies.
309: 
310: The collision operators conserve the particle number of each
311: species and the total momentum but the total energy is not
312: conserved:
313: \begin{equation}
314: \int d{\bf v}J_{ij}[{\bf v}|f_{i},f_{j}]=0\;,  \label{2.6}
315: \end{equation}
316: \begin{equation}
317: \sum_{i=1}^2\sum_{j=1}^2m_i\int d{\bf v}{\bf v}J_{ij}[{\bf
318: v}|f_{i},f_{j}]=0 \;, \label{2.7}
319: \end{equation}
320: \begin{equation}
321: \sum_{i=1}^2\sum_{j=1}^2m_i\int d{\bf v}V^{2}J_{ij}[{\bf v}
322: |f_{i},f_{j}]=-d nT\zeta \;,  \label{2.8}
323: \end{equation}
324: where $\zeta$ is identified as the total ``cooling rate'' due to
325: inelastic collisions among all species. At a kinetic level, it is
326: also convenient to introduce the ``cooling rates'' $\zeta_i$ for
327: the partial temperatures $T_i$. They are defined as
328: \begin{equation}
329: \label{2.7.1} \zeta_i=\sum_{j=1}^2\zeta_{ij}=-
330: \frac{m_i}{dn_iT_i}\sum_{j=1}^2\int d{\bf v}V^{2}J_{ij}[{\bf
331: v}|f_{i},f_{j}],
332: \end{equation}
333: where the second equality defines the quantities $\zeta_{ij}$. The
334: total cooling rate $\zeta$ can be written in terms of the partial
335: cooling rates $\zeta_i$ as
336: \begin{equation}
337: \label{2.7.2} \zeta=T^{-1}\sum_{i=1}^2\;x_iT_i\zeta_i,
338: \end{equation}
339: where $x_i=n_i/n$ is the mole fraction of species $i$.
340: 
341: 
342: From Eqs.\ (\ref{2.4})--(\ref{2.8}), the macroscopic balance
343: equations for the mixture can be obtained. They are given by
344: \begin{equation}
345: D_{t}n_{i}+n_{i}\nabla \cdot {\bf u}+\frac{\nabla \cdot {\bf
346: j}_{i}}{m_{i}} =0\;,  \label{2.9}
347: \end{equation}
348: \begin{equation}
349: D_{t}{\bf u}+\rho ^{-1}\nabla \cdot {\sf P}={\bf g}\;,
350: \label{2.10}
351: \end{equation}
352: \begin{equation}
353: D_{t}T-\frac{T}{n}\sum_{i=1}^2\frac{\nabla \cdot {\bf
354: j}_{i}}{m_{i}}+\frac{2}{dn} \left( \nabla \cdot {\bf q}+{\sf
355: P}:\nabla {\bf u}\right) =-\zeta \,T\;. \label{2.11}
356: \end{equation}
357: In the above equations, $D_{t}=\partial _{t}+{\bf u}\cdot \nabla $
358: is the material derivative,
359: \begin{equation}
360: {\bf j}_{i}=m_{i}\int d{\bf v}\,{\bf V}\,f_{i}({\bf v})
361: \label{2.11b}
362: \end{equation}
363: is the mass flux for species $i$ relative to the local flow,
364: \begin{equation}
365: {\sf P}=\sum_{i=1}^2\,m_i\,\int d{\bf v}\,{\bf V}{\bf
366: V}\,f_{i}({\bf v})  \label{2.12}
367: \end{equation}
368: is the total pressure tensor, and
369: \begin{equation}
370: {\bf q}=\sum_{i=1}^2\,\frac{m_i}{2}\int d{\bf v}\,V^{2}{\bf
371: V}\,f_{i}({\bf v})  \label{2.13}
372: \end{equation}
373: is the total heat flux.
374: 
375: 
376: The macroscopic balance equations (\ref{2.9})--(\ref{2.11}) are
377: not entirely expressed in terms of the hydrodynamic fields, due to
378: the presence of the cooling rate $\zeta$, the mass flux ${\bf
379: j}_i$, the heat flux ${\bf q}$, and the pressure tensor ${\sf P}$
380: which are given as functionals of the distributions $f_i$.
381: However, it these distributions can be expressed as functionals of
382: the hydrodynamic fields, then the cooling rate and the fluxes also
383: will become functional of the hydrodynamic fields through Eqs.\
384: (\ref{2.7.1}) and (\ref{2.11b})--(\ref{2.13}). Such expressions
385: are called {\em constitutive} relations and they provide a link
386: between the exact balance equations and a closed set of equations
387: for the hydrodynamic fields. This hydrodynamic description can be
388: derived by looking for a {\em normal} solution to the Boltzmann
389: kinetic equation. A normal solution is one whose all space and
390: time dependence of the distribution function $f_i$ occurs through
391: a functional dependence on the hydrodynamic fields,
392: \begin{equation}
393: f_{i}({\bf r},{\bf v},t)=f_{i}\left[{\bf v}|x_{1} ({\bf r}, t),
394: p({\bf r}, t), T({\bf r}, t), {\bf u}({\bf r}, t) \right] \;.
395: \label{2.14}
396: \end{equation}
397: As in previous works, \cite{GD02,MGD06} we have taken the set
398: $\{x_1, p, T, {\bf u}\}$ as the $d+3$ independent fields of the
399: two-component mixture. These are the most accessible fields from
400: an experimental point of view. The determination of this normal
401: solution from the Boltzmann equation (\ref{2.1}) is a very
402: difficult task in general, unless the spatial gradients are small.
403: In this case, the CE method gives an approximate solution.
404: 
405: 
406: \section{Chapman-Enskog solution}
407: \label{sec2bis}
408: 
409: 
410: The CE method is a procedure to construct an approximate normal
411: solution. It is perturbative, using the spatial gradients as the
412: small expansion parameter. More specifically, one assumes that the
413: spatial variations of the hydrodynamic fields $n_i$, ${\bf u}$,
414: $p$, and $T$ are small on the scale of the mean free path. For
415: ordinary gases this can be controlled by the initial or boundary
416: conditions. It is more complicated for granular gases, since in
417: some cases (e.g., steady states such as the simple shear flow
418: problem\cite{SGD04}) the boundary conditions imply a relationship
419: between dissipation and some hydrodynamic gradient. As a
420: consequence, there are examples for which the NS approximation is
421: restricted to the quasi-elastic limit.\cite{SGD04} Here, we also
422: assume that the spatial gradients are independent of the
423: coefficients of restitution so that, the corresponding NS order
424: hydrodynamic equations apply for small gradients but they are not
425: limited {\em a priori} to weak inelasticity. It must be emphasized
426: that our perturbation scheme differs from the one recently carried
427: out by Serero {\em et al.} \cite{SGNT06} where the CE solution is
428: given in powers of both the hydrodynamic gradients (or
429: equivalently, the Knudsen number) and the degree of dissipation
430: $\epsilon_{ij}=1-\alpha_{ij}^2$. The results provided in Ref.\
431: \onlinecite{SGNT06} only agree with our results in the
432: quasielastic domain (small $\epsilon_{ij}$). Moreover, in the
433: presence of an external force it is necessary to characterize the
434: magnitude of the force relative to gradients as well. As in the
435: elastic case, \cite{CC70} it is assumed here that the magnitude of
436: the gravity field is at least of first order in perturbation
437: expansion.
438: 
439: For small spatial variations, the functional dependence
440: (\ref{2.14}) can be made local in space through an expansion in
441: gradients of the hydrodynamic fields. To generate it, $f_{i}$ is
442: written as a series expansion in a formal parameter $\delta$
443: measuring the nonuniformity of the system,
444: \begin{equation}
445: f_{i}=f_{i}^{(0)}+\delta \,f_{i}^{(1)}+\delta^2
446: \,f_{i}^{(2)}+\cdots \;, \label{2.15}
447: \end{equation}
448: where each factor of $\delta$ means an implicit gradient of a
449: hydrodynamic field. The local reference states $f_{i}^{(0)}$ are
450: chosen such that they verify Eqs.\ (\ref{2.4}) and (\ref{2.5}), or
451: equivalently, the remainder of the expansion must obey the
452: orthogonality conditions
453: \begin{equation}
454: \label{2.16bis}
455:  \int d{\bf v}\left[ f_{i}({\bf
456: v})-f_{i}^{(0)}({\bf v})\right] =0\;,\
457: \end{equation}
458: \begin{equation}
459: \sum_{i=1}^2\,m_i\,\int d{\bf v}\,{\bf v}\left[ f_{i}({\bf
460: v})-f_{i}^{(0)}({\bf v})\right] ={\bf 0}\;,  \label{2.16}
461: \end{equation}
462: \begin{equation}
463: \sum_{i=1}^2\,\frac{m_{i}}{2}\,\int d{\bf v}\,V^{2}\left[
464: f_{i}({\bf v})-f_{i}^{(0)}({\bf v})\right] =0\;.  \label{2.17}
465: \end{equation}
466: The time derivatives of the fields are also expanded as $\partial
467: _{t}=\partial _{t}^{(0)}+\epsilon \partial _{t}^{(1)}+\cdots $.
468: The action of the operators $\partial _{t}^{(k)}$ can be obtained
469: from the balance equations (\ref{2.9})--(\ref{2.11}) when one
470: takes into account the corresponding expansions for the fluxes and
471: the cooling rate. This is the usual CE method \cite{CC70} for
472: solving kinetic equations. The main difference in the case of
473: inelastic collisions is that the reference state has a time
474: dependence associated with the cooling that is not proportional to
475: the gradients. As a consequence, terms from the time derivative
476: $\partial_t^{(0)}$ are not zero.  In addition, the different
477: approximations $f_i^{(k)}$ are well-defined functions of the
478: coefficients of restitution $\alpha_{ij}$, regardless of the
479: applicability of the corresponding hydrodynamic equations
480: truncated at that order.
481: 
482: 
483: \subsection{Zeroth-order approximation}
484: 
485: 
486: To zeroth order in the gradients, Eq.\ (\ref{2.1}) becomes
487: \begin{equation}
488: \label{2.17.1}
489: \partial_t^{(0)}f_i^{(0)}=\sum_{j=1}^2\;
490: J_{ij}[f_i^{(0)},f_j^{(0)}],
491: \end{equation}
492: where use has been made of the fact that gravity is assumed to be
493: of first order in the uniformity parameter $\delta$. The balance
494: equations to this order give
495: \begin{equation}
496: \label{2.17.2}
497: \partial_t^{(0)}x_1=\partial_t^{(0)}u_\ell=0,\quad
498: T^{-1}\partial_t^{(0)}T=p^{-1}\partial_t^{(0)}p=-\zeta^{(0)},
499: \end{equation}
500: where $\zeta^{(0)}$ is determined by Eqs.\ (\ref{2.7.1}) and
501: (\ref{2.7.2}) to zeroth order in the gradients. Since $f_i^{(0)}$
502: is a normal solution, then the time derivative in Eq.\
503: (\ref{2.17.1}) can be written as
504: \begin{equation}
505: \label{2.17.3}
506: \partial_t^{(0)}f_i^{(0)}=-\zeta^{(0)}\left(T\partial_T+p\partial_p\right)f_i^{(0)}=
507: \frac{1}{2}\zeta^{(0)}\frac{\partial}{\partial {\bf V}}\cdot
508: \left({\bf V}f_i^{(0)}\right).
509: \end{equation}
510: The second equality in Eq.\ (\ref{2.17.3}) follows from
511: dimensional analysis which requires that the dependence of
512: $f_i^{(0)}$ on $p$ and $T$ is of the form
513: \begin{equation}
514: \label{2.17.4} f_i^{(0)}(
515: V)=x_i\frac{p}{T}v_0^{-d}\Phi_i\left(V/v_0\right),
516: \end{equation}
517: where $v_0(t)=\sqrt{2T(m_1+m_2)/m_1m_2}$ is a thermal velocity
518: defined in terms of the temperature $T(t)$ of the mixture and
519: $\Phi_i$ is a dimensionless function of the reduced velocity
520: $V/v_0$. The dependence of $f_i^{(0)}$ on the magnitude of ${\bf
521: V}$ follows from the isotropy of the zeroth-order distribution
522: with respect to the peculiar velocity. Thus, the Boltzmann
523: equation at this order reads
524: \begin{equation}
525: \label{2.17.5} \frac{1}{2}\zeta^{(0)}\frac{\partial}{\partial {\bf
526: V}}\cdot \left({\bf V}f_i^{(0)}\right) \sum_{j=1}^2\;
527: J_{ij}[f_i^{(0)},f_j^{(0)}].
528: \end{equation}
529: Equation (\ref{2.17.5}) has the same form as the Boltzman equation
530: for a strictly {\em homogeneous} state. The latter is called the
531: homogeneous cooling state (HCS). \cite{GD99b} Here, however, the
532: state is not homogeneous because of the requirements
533: (\ref{2.16bis})--(\ref{2.17}). Instead it is a {\em local} HCS. It
534: must be emphasized that the presence of this local HCS as the
535: ground or reference state is not an assumption of the CE expansion
536: but rather a consequence of the kinetic equations at zeroth order
537: in the gradient expansion.
538: 
539: 
540: The local HCS distribution is the solution of the Boltzmann
541: equation (\ref{2.17.5}). However, its explicit form is not exactly
542: known even in the one-component case.\cite{NE98} An accurate
543: approximation for the zeroth-order solution $f_i^{(0)}$ can be
544: obtained by using low order truncation of a Sonine polynomial
545: expansion. The results show that in general, $f_i^{(0)}$ is close
546: to a Maxwellian at the temperature for that species. Further
547: details of this solution for hard spheres ($d=3$) can be found in
548: Ref.\ \onlinecite{GD99b}. An important consequence is that the
549: kinetic temperatures of each species are different for inelastic
550: collisions and, consequently the total energy is not equally
551: distributed between both species (breakdown of energy
552: equipartition). This violation of energy equipartition has been
553: confirmed by computer simulation studies \cite{computer} as well
554: as by real experiments. \cite{exp1} The condition that $f_i^{(0)}$
555: is {\em normal} in the sense of Eq.\ (\ref{2.14}) (namely, it
556: depends on time only through its functional dependence on $T$ and
557: $p$) implies that the ratio $T_i/T\equiv \gamma_i(x_1)$ depends on
558: the hydrodynamic state through the concentration $x_1$.
559: 
560: 
561: 
562: The dependence of the temperature ratio $\gamma\equiv
563: \gamma_1/\gamma_2=T_1/T_2$ on the parameters of the mixture is
564: obtained by requiring that the partial cooling rates
565: $\zeta_i^{(0)}$ must be equal,\cite{GD99b} i.e.,
566: \begin{equation}
567: \label{2.17bis} \zeta_1^{(0)}=\zeta_2^{(0)}=\zeta^{(0)}.
568: \end{equation}
569: These partial cooling rates are nonlinear functionals of the
570: distributions $f_i^{(0)}$, which are not exactly known. However,
571: to get the temperature ratio, they can be well estimated by using
572: Maxwellians at different temperatures:
573: \begin{equation}
574: \label{2.18} f_i^{(0)}({\bf V})\to f_{i,M}({\bf
575: V})=n_i\left(\frac{m_i}{2\pi T_i}\right)^{d/2}\exp\left(-
576: \frac{m_i V^2}{2T_i}\right).
577: \end{equation}
578: In this approximation, one gets \cite{G02}
579: \begin{eqnarray}
580: \label{2.19}
581: \zeta_i^{(0)}&=&\sum_{j=1}^2\zeta_{ij}^{(0)}=\frac{4\pi^{(d-1)/2}}{d\Gamma\left(\frac{d}{2}\right)}
582: v_0\sum_{j=1}^2
583: n_j\mu_{ji}\sigma_{ij}^{d-1}\left(\frac{\theta_i+\theta_j}
584: {\theta_i\theta_j}\right)^{1/2}\nonumber\\
585: & &\times (1+\alpha_{ij})
586: \left[1-\frac{\mu_{ji}}{2}(1+\alpha_{ij})
587: \frac{\theta_i+\theta_j}{\theta_j}\right],
588: \end{eqnarray}
589: where
590: \begin{equation}
591: \label{2.20} \theta_i=\frac{m_i}{\gamma_i}\sum_{j=1}^2\,m_j^{-1}.
592: \end{equation}
593: It must be remarked that the fact that $T_1(t)\neq T_2(t)$ does
594: not mean that there are additional hydrodynamic degrees of freedom
595: since the partial temperatures $T_i$ can be expressed in terms of
596: the granular temperature $T$ as
597: \begin{equation}
598: \label{2.20.0} T_1(t)=\frac{\gamma}{1+x_1(\gamma-1)}T(t),\quad
599: T_2(t)=\frac{1}{1+x_1(\gamma-1)}T(t).
600: \end{equation}
601: Note that the reference Maxwellians (\ref{2.18}) for the two
602: species can be quite different due to the temperature differences.
603: This contrasts with the ground state considered in more standard
604: derivations \cite{mixture,SGNT06} where $f_i^{(0)}$ is replaced by
605: a Maxwellian defined at the same temperature $T$, i.e.,
606: \begin{equation}
607: \label{2.20.1} f_i^{(0)}({\bf V})\to n_i\left(\frac{m_i}{2\pi
608: T}\right)^{d/2}\exp\left(- \frac{m_i V^2}{2T}\right).
609: \end{equation}
610: As will show later, the approaches (\ref{2.18}) and (\ref{2.20.1})
611: lead to different results for the NS transport coefficients.
612: \begin{figure}[htbp]
613: \begin{center}
614: \resizebox{7cm}{!}{\includegraphics{fig1.eps}}
615: \end{center}
616: \caption{Temperature ratio $\gamma\equiv T_1/T_2$ versus the
617: (common) coefficient of restitution $\alpha$ for an equimolar
618: mixture ($x_1=1/2$) of hard disks ($d=2$) with $\omega=\mu^{1/2}$.
619: Two different values of the mass ratio are considered: $\mu=4$ and
620: $\mu=8$.  The symbols refer to DSMC results while the lines
621: represent the theoretical results obtained from the condition
622: (\ref{2.17bis}).} \label{fig1}
623: \end{figure}
624: 
625: 
626: The solution to Eq.\ (\ref{2.17bis}) gives the temperature ratio
627: $T_1/T_2$ for any dimension $d$ as a function of the mole fraction
628: $x_1$, the mass ratio $\mu\equiv m_1/m_2$, the size ratio
629: $\omega\equiv \sigma_1/\sigma_2$, and the coefficients of
630: restitution $\alpha_{ij}$. To illustrate the violation of
631: equipartition theorem, in Fig.\ \ref{fig1} we plot the temperature
632: ratio versus the coefficient of restitution $\alpha$ for hard
633: disks ($d=2$) in the case of an equimolar mixture
634: $x_1=\frac{1}{2}$ for two different mixtures composed by particles
635: of the same material: $\mu=4$, $\omega=2$, and $\mu=8$,
636: $\omega=\sqrt{8}$. For the sake of simplicity, we have taken a
637: common coefficient of restitution $\alpha\equiv
638: \alpha_{11}=\alpha_{22}=\alpha_{12}$. We also include the
639: simulation data obtained by solving  numerically the Boltzmann
640: equation by means of the DSMC method. \cite{Bird} The excellent
641: agreement between theory and simulation shows the accuracy of the
642: estimate (\ref{2.19}) to compute the temperature ratio from the
643: equality of cooling rates (\ref{2.17bis}). We also observe that
644: the deviations from the energy equipartition increase as the
645: mechanical differences between the particles of each species
646: increase.
647: 
648: 
649: \section{Navier-Stokes transport coefficients}
650: \label{sec3}
651: 
652: 
653: The CE procedure allows one to determine the NS transport
654: coefficients of the mixture in the first order of the expansion.
655: The analysis to first order in gradients is similar to the one
656: worked out in Ref.\ \onlinecite{GD02} for $d=3$. Here, we only
657: present the final results with some technical details being given
658: in Appendix \ref{appA}. The mass, momentum, and heat fluxes are
659: given, respectively, by
660: \begin{equation}
661: {\bf j}_{1}^{(1)}=-\frac{m_{1}m_{2}n}{\rho } D\nabla x_{1}-\frac{
662: \rho }{p}D_{p}\nabla p-\frac{\rho }{T}D^{\prime }\nabla
663: T,\hspace{0.3in}{\bf j}_{2}^{(1)}=-{\bf j}_{1}^{(1)},  \label{3.1}
664: \end{equation}
665: \begin{equation}
666: P_{k \ell}^{(1)}=p\;\delta _{k\ell}-\eta \left( \nabla _{\ell
667: }u_{k}+\nabla _{k}u_{\ell}-\frac{2}{d}\delta _{k\ell} {\bf \nabla
668: \cdot u}\right),  \label{3.2}
669: \end{equation}
670: \begin{equation}
671: {\bf q}^{(1)}=-T^{2}D^{\prime \prime }\nabla x_{1}-L\nabla
672: p-\lambda \nabla T. \label{3.3}
673: \end{equation}
674: The transport coefficients in these equations are the diffusion
675: coefficient $D$, the pressure diffusion coefficient $D_p$, the
676: thermal diffusion coefficient $D'$, the shear viscosity $\eta$,
677: the Dufour coefficient $D''$, the pressure energy coefficient $L$,
678: and the thermal conductivity $\lambda$. These coefficients are
679: defined as
680: \begin{equation}
681: D=-\frac{\rho }{dm_{2}n}\int d{\bf v}\,{\bf V}\cdot {\boldsymbol
682: {\cal A}}_{1}, \label{3.4}
683: \end{equation}
684: \begin{equation}
685: D_{p}=-\frac{m_{1}p}{d\rho}\int d{\bf v}\,{\bf V}\cdot
686: {\boldsymbol {\cal B}}_{1}, \label{3.5}
687: \end{equation}
688: \begin{equation}
689: D^{\prime }=-\frac{m_{1}T}{d\rho }\int d{\bf v}\,{\bf V}\cdot
690: {\boldsymbol {\cal C}}_{1}, \label{3.6}
691: \end{equation}
692: \begin{equation}
693: \eta =-\frac{1}{(d-1)(d+2)}\sum_{i=1}^2\,m_i\,\int d{\bf v}\, {\bf
694: V}{\bf V}:{\boldsymbol {\cal D}}_{i}, \label{3.7}
695: \end{equation}
696: \begin{equation}
697: D^{\prime \prime
698: }=-\frac{1}{dT^{2}}\sum_{i=1}^2\,\frac{m_i}{2}\,\int d{\bf
699: v}\,V^{2}{\bf V}\cdot {\boldsymbol {\cal A}}_{i}, \label{3.8}
700: \end{equation}
701: \begin{equation}
702: L=-\frac{1}{d}\sum_{i=1}^2\,\frac{m_i}{2}\,\int d{\bf
703: v}\,V^{2}{\bf V} \cdot \,{\boldsymbol {\cal B}}_{i}, \label{3.9}
704: \end{equation}
705: \begin{equation}
706: \lambda =-\frac{1}{d}\sum_{i=1}^2\,\frac{m_i}{2}\,\int d{\bf
707: v}\,V^{2} {\bf V}\cdot {\boldsymbol {\cal C}}_{i}. \label{3.9bis}
708: \end{equation}
709: As for ordinary gases, \cite{CC70} the unknowns ${\boldsymbol
710: {\cal A}}_{i}({\bf V})$, ${\boldsymbol {\cal B}}_{i}({\bf V})$,
711: ${\boldsymbol {\cal C}}_{i}({\bf V})$, and ${\boldsymbol {\cal
712: D}}_{i}({\bf V})$ are the solutions of the following set of
713: coupled linear integral equations:
714: \begin{subequations}
715: \begin{equation}
716: \left[ -\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
717: +{\cal L}_{1} \right] {\boldsymbol {\cal A}}_{1}+{\cal
718: M}_{1}{\boldsymbol {\cal A}}_{2}={\bf A}_{1}+\left( \frac{\partial
719: \zeta ^{(0)}}{\partial x_{1}}\right) _{p,T}\left( p{\boldsymbol
720: {\cal
721:  B}}_{1}+T{\boldsymbol {\cal C}}_{1}\right) ,  \label{3.10a}
722: \end{equation}
723: \begin{equation}
724: \left[ -\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
725: +{\cal L}_{2} \right] {\boldsymbol {\cal A}}_{2}+{\cal
726: M}_{2}{\boldsymbol {\cal A}}_{1}={\bf A}_{2}+\left( \frac{\partial
727: \zeta ^{(0)}}{\partial x_{1}}\right) _{p,T}\left( p{\boldsymbol
728: {\cal
729:  B}}_{2}+T{\boldsymbol {\cal C}}_{2}\right) ,  \label{3.10b}
730: \end{equation}
731: \label{3.10}
732: \end{subequations}
733: \begin{subequations}
734: \begin{equation}
735: \left[ -\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
736: +{\cal L}_{1}-2\zeta ^{(0)}\right] {\boldsymbol {\cal
737: B}}_{1}+{\cal M}_{1}{\boldsymbol {\cal B}}_{2}= {\bf
738: B}_{1}+\frac{T\zeta ^{(0)}}{p}{\boldsymbol {\cal C}}_{1},
739: \label{3.11a}
740: \end{equation}
741: \begin{equation}
742: \left[ -\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
743: +{\cal L} _{2}-2\zeta ^{(0)}\right] {\boldsymbol {\cal
744: B}}_{2}+{\cal M}_{2}{\boldsymbol {\cal B}}_{1}= {\bf
745: B}_{2}+\frac{T\zeta ^{(0)}}{p}{\boldsymbol {\cal C}}_{2},
746: \label{3.11b}
747: \end{equation}
748: \label{3.11}
749: \end{subequations}
750: \begin{subequations}
751: \begin{equation}
752: \left[ -\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
753: +{\cal L}_{1}-\frac{1}{2}\zeta ^{(0)}\right] {\boldsymbol {\cal
754: C}}_{1}+{\cal M}_{1} {\boldsymbol {\cal C}}_{2}= {\bf
755: C}_{1}-\frac{p\zeta ^{(0)}}{2T}{\boldsymbol {\cal B}}_{1},
756: \label{3.12a}
757: \end{equation}
758: \begin{equation}
759: \left[ -\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
760: +{\cal L}_{2}-\frac{1}{2}\zeta ^{(0)}\right] {\boldsymbol {\cal
761: C}}_{2}+{\cal M}_{2} {\boldsymbol {\cal C}}_{1}= {\bf
762: C}_{2}-\frac{p\zeta ^{(0)}}{2T}{\boldsymbol {\cal B}}_{2},
763: \label{3.12b}
764: \end{equation}
765: \label{3.12}
766: \end{subequations}
767: \begin{subequations}
768: \begin{equation}
769: \label{3.12.0a} \left[ -\zeta ^{(0)}\left( T\partial
770: _{T}+p\partial _{p}\right) +{\cal L}_{1}\right] {\boldsymbol {\cal
771: D}}_{1} +{\cal M}_{1}{\boldsymbol {\cal D}}_{2}={\sf D}_1,
772: \end{equation}
773: \begin{equation}
774: \label{3.12.0b} \left[ -\zeta ^{(0)}\left( T\partial
775: _{T}+p\partial _{p}\right) +{\cal L}_{2}\right] {\boldsymbol {\cal
776: D}}_{2} +{\cal M}_{2}{\boldsymbol {\cal D}}_{1}={\sf D}_2.
777: \end{equation}
778: \label{3.12.0}
779: \end{subequations}
780: In the above equations, the quantities ${\bf A}_{i}$, ${\bf
781: B}_{i}$, ${\bf C}_{i}$, and ${\sf D}_{i}$ are given by Eqs.\
782: (\ref{a7})--(\ref{a10}), respectively. They depend on the local
783: HCS distribution $f_i^{(0)}$. In addition, we have introduced the
784: linearized Boltzmann collision operators
785: \begin{equation}
786: {\cal L}_{1}X=-\left( J_{11}[f_{1}^{(0)},X]+J_{11}[X,f_{1}^{(0)}]+
787: J_{12}[X,f_{2}^{(0)}]\right) \;, \label{3.12.1}
788: \end{equation}
789: \begin{equation}
790: {\cal M}_{1}X=-J_{12}[f_{1}^{(0)},X].  \label{3.12.2}
791: \end{equation}
792: The corresponding expressions for the operators ${\cal L}_{2}$ and
793: ${\cal M}_{2}$ can be easily obtained from Eqs.\ (\ref{3.12.1})
794: and (\ref{3.12.2}) by just making the changes $1\leftrightarrow
795: 2$. Note that in Eq.\ (\ref{3.10}) the cooling rate $\zeta^{(0)}$
796: depends on $x_1$ explicitly and through its dependence on $\gamma
797: (x_1)$. This dependence gives rise to significant new
798: contributions to the integral equations for the transport
799: coefficients. Furthermore, the external field does not occur in
800: Eqs.\ (\ref{3.10})--(\ref{3.12.0}). This is because the particular
801: form of the gravitational force.
802: 
803: \subsection{Sonine polynomial approximation}
804: 
805: So far, all the results are exact. However, explicit expressions
806: for the NS transport coefficients requires to solve Eqs.\
807: (\ref{3.10})--(\ref{3.12.0}) as well as the integral equations
808: (\ref{2.17.5}) for the reference distributions $f_i^{(0)}$. As
809: said before, the results obtained in the HCS \cite{GD99b} have
810: shown that $f_i^{(0)}$ is well represented by its Maxwellian form
811: (\ref{2.18}) in the region of thermal velocities. For this reason
812: and to provide simple but accurate expressions for the transport
813: coefficients, non-Gaussian corrections to $f_i^{(0)}$ will be
814: neglected in our theory. The full expressions for the transport
815: coefficients in the case $d=3$ (including non-Gaussian
816: corrections) can be found in Ref.\ \onlinecite{GD02}. It must
817: remarked that while the effect of these non-Gaussian corrections
818: on the transport coefficients is not important in the case of the
819: mass flux and the pressure tensor, the same does not happen for
820: the heat flux, where the influence of them is not negligible at
821: high inelasticity. \cite{MGD06} With respect to the functions
822: $\left( {\boldsymbol {\cal A}}_{i}, {\boldsymbol {\cal B}}_{i},
823: {\boldsymbol {\cal C}}_{i}, {\boldsymbol {\cal D}}_{i} \right)$,
824: we will expand them in a series expansion of Sonine polynomials
825: and will consider only the leading terms. The procedure is
826: described in Appendix \ref{appB} and only the final expressions
827: will be provided here.
828: 
829: 
830: 
831: \subsection{Mass flux}
832: 
833: 
834: In dimensionless form, the transport coefficients associated with
835: the mass flux, $D$, $D_p$, and $D'$ can be written as
836: \begin{equation}
837: D=\frac{\rho T}{m_{1}m_{2}\nu _{0}}D^{\ast },\quad
838: D_{p}=\frac{nT}{\rho \nu _{0}}D_{p}^{\ast },\quad D^{\prime
839: }=\frac{nT}{\rho \nu _{0}}D^{\prime}{}^{\ast },  \label{3.n1}
840: \end{equation}
841: where $\nu _{0}=n\sigma _{12}^{d-1}v_{0}$ is an effective
842: collision frequency. The explicit forms are
843: \begin{equation}
844: D^{\ast}=\left( \nu ^{\ast }-\frac{1}{2}\zeta ^{\ast
845: }\right)^{-1}\left[ \left( \frac{\partial }{\partial
846: x_{1}}x_{1}\gamma _{1}\right) _{p,T}+\left( \frac{\partial \zeta
847: ^{\ast }}{\partial x_{1}} \right) _{p,T}\left( 1-\frac{\zeta
848: ^{\ast}}{2\nu ^{\ast }}\right) D_{p}^{\ast }\right] , \label{3.13}
849: \end{equation}
850: \begin{equation}
851: D_{p}^{\ast }=x_{1}\left( \gamma _{1}-\frac{\mu}{x_2+\mu x_1}
852: \right) \left( \nu ^{\ast }-\frac{3}{2}\zeta ^{\ast }+\frac{\zeta
853: ^{\ast 2}}{ 2\nu ^{\ast }}\right) ^{-1}, \label{3.14}
854: \end{equation}
855: \begin{equation}
856: D^{\prime\ast }=-\frac{\zeta ^{\ast }}{2\nu ^{\ast }}D_{p}^{\ast
857: }. \label{3.15}
858: \end{equation}
859: Here, $\zeta^*=\zeta^{(0)}/\nu_0$, and $\nu^*$ is given by
860: \begin{equation}
861: \label{3.16}
862: \nu^*=\frac{2\pi^{(d-1)/2}}{d\Gamma\left(\frac{d}{2}\right)}
863: (1+\alpha_{12})
864: \left(\frac{\theta_1+\theta_2}{\theta_1\theta_2}\right)^{1/2}\left(x_2\mu_{21}+
865: x_1\mu_{12}\right).
866: \end{equation}
867: Since ${\bf j}_{1}^{(1)}=-{\bf j}_{2}^{(1)}$ and $\nabla
868: x_{1}=-\nabla x_{2}$, $D^*$ must be symmetric while $D_{p}^*$ and
869: $D^{\prime \ast}$ must be antisymmetric with respect to the
870: exchange $1\leftrightarrow 2$ . This can be easily verified by
871: noting that $x_{1}\gamma_{1}+x_{2}\gamma_{2}=1$. The expressions
872: for $D^*$, $D_{p}^*$ and $D^{\prime \ast}$ reduce to those
873: recently obtained \cite{BRM05} in the tracer limit ($x_1\to 0$).
874: 
875: \subsection{Pressure tensor}
876: 
877: The shear viscosity coefficient $\eta$ can be written as
878: \begin{equation}
879: \label{3.17}
880: \eta=\frac{p}{\nu_0}\left(x_1\gamma_1^2\eta_1^*+x_2\gamma_2^2\eta_2^*\right),
881: \end{equation}
882: where the expression of the (dimensionless) partial contribution
883: $\eta_i^*$ is
884: \begin{equation}
885: \label{3.18}
886: \eta_1^*=2\frac{\gamma_2(2\tau_{22}-\zeta^{*})-2\gamma_1\tau_{12}}
887: {\gamma_1\gamma_2[\zeta^*-2\zeta^{*}
888: (\tau_{11}+\tau_{22})+4(\tau_{11}\tau_{22}-\tau_{12}\tau_{21})]}.
889: \end{equation}
890: Here, we have introduced the (reduced) collision frequencies
891: $\tau_{11}$ and $\tau_{12}$ given by
892: \begin{widetext}
893: \begin{eqnarray}
894: \label{3.19}
895: \tau_{11}&=&\frac{2\pi^{(d-1)/2}}{d(d+2)\Gamma\left(\frac{d}{2}\right)}\left\{
896: x_1\left(\frac{\sigma_{1}}{\sigma_{12}}\right)^{d-1}(2\theta_1)^{-1/2}(3+2d-3\alpha_{11})
897: (1+\alpha_{11})\right.\nonumber\\
898: & &+2x_2 \mu_{21}(1+\alpha_{12}) \theta_1^{3/2}\theta_2^{-1/2}
899: \left[
900: (d+3)(\mu_{12}\theta_2-\mu_{21}\theta_1)\theta_1^{-2}(\theta_1+\theta_2)^{-1/2}\right.\nonumber\\
901: & &
902: \left.\left.+\frac{3+2d-3\alpha_{12}}{2}\mu_{21}\theta_1^{-2}(\theta_1+\theta_2)^{1/2}
903: +\frac{2d(d+1)-4}{2(d-1)}\theta_1^{-1}(\theta_1+\theta_2)^{-1/2}\right]\right\},
904: \end{eqnarray}
905: \begin{eqnarray}
906: \label{3.20}
907: \tau_{12}&=&\frac{4\pi^{(d-1)/2}}{d(d+2)\Gamma\left(\frac{d}{2}\right)}
908: x_2\frac{\mu_{21}^2}{\mu_{12}}\theta_1^{3/2}\theta_2^{-1/2}
909: (1+\alpha_{12})\nonumber\\
910: & \times&\left[
911: (d+3)(\mu_{12}\theta_2-\mu_{21}\theta_1)\theta_2^{-2}(\theta_1+\theta_2)^{-1/2}\right.\nonumber\\
912: & &
913: \left.+\frac{3+2d-3\alpha_{12}}{2}\mu_{21}\theta_2^{-2}(\theta_1+\theta_2)^{1/2}
914: -\frac{2d(d+1)-4}{2(d-1)}\theta_2^{-1}(\theta_1+\theta_2)^{-1/2}\right].
915: \end{eqnarray}
916: \end{widetext}
917: A similar expression can be obtained for $\eta_2^*$ by just making
918: the changes $1 \leftrightarrow 2$.
919: 
920: \subsection{Heat flux}
921: 
922: The case of the heat flux is more involved since it requires to
923: consider the second Sonine approximation. The transport
924: coefficients appearing in the heat flux, $D''$, $L$, and $\lambda$
925: can be written as
926: \begin{equation}
927: D^{\prime \prime
928: }=-\frac{d+2}{2}\frac{n}{(m_{1}+m_{2})\nu_0}\left[ \frac{
929: x_{1}\gamma _{1}^{3}}{\mu _{12}}d_{1}^*+\frac{x_{2}\gamma
930: _{2}^{3}}{\mu _{21}}d_{2}^*-\left( \frac{\gamma _{1}}{\mu
931: _{12}}-\frac{\gamma _{2}}{\mu _{21}}\right) D^{\ast }\right] ,
932: \label{4.7}
933: \end{equation}
934: \begin{equation}
935: L=-\frac{d+2}{2}\frac{T}{(m_{1}+m_{2})\nu_0}\left[
936: \frac{x_{1}\gamma _{1}^{3}}{\mu _{12}}\ell
937: _{1}^*+\frac{x_{2}\gamma _{2}^{3}}{\mu _{21}}\ell _{2}^*-\left(
938: \frac{\gamma _{1}}{\mu _{12}}-\frac{\gamma _{2}}{\mu _{21}}
939: \right) D_{p}^{\ast }\right] ,  \label{4.8}
940: \end{equation}
941: \begin{equation}
942: \lambda =-\frac{d+2}{2}\frac{nT}{(m_{1}+m_{2})\nu _{0}}\left[
943: \frac{ x_{1}\gamma _{1}^{3}}{\mu _{12}}\lambda
944: _{1}^*+\frac{x_{2}\gamma _{2}^{3}}{\mu_{21}}\lambda _{2}^*-\left(
945: \frac{\gamma _{1}}{\mu _{12}}-\frac{\gamma _{2}}{ \mu
946: _{21}}\right) D^{\prime }{}^{\ast }\right] , \label{4.9}
947: \end{equation}
948: where the coefficients $D^{\ast}$, $D_{p}^{\ast }$, and
949:  $D^{\prime \ast }$ are given by
950:  Eqs.\ (\ref{3.13})--(\ref{3.15}), respectively. The expressions
951:  of the (dimensionless) Sonine coefficients $d_{i}^{*}$, $\ell
952: _{i}^*$, and $\lambda _{i}^*$ are
953: \begin{eqnarray}
954: \label{4.n2} d_1^{*}&=&\frac{1}{\Delta}\left\{2\left[2
955: \nu_{12}Y_2-Y_1(2\nu_{22}-3\zeta^*)\right]\left[\nu_{12}\nu_{21}-\nu_{11}\nu_{22}
956: +2(\nu_{11}+\nu_{22})\zeta^*-4\zeta^{*2}\right]\right.\nonumber\\
957: & &+2\left( \frac{\partial \zeta ^{\ast }}{\partial x_{1}}\right)
958: _{p,T}(Y_3+Y_5)\left[2\nu_{12}\nu_{21}+2\nu_{22}^2-\zeta^*(
959: 7\nu_{22}-6\zeta^{*})\right]\nonumber\\
960: & & \left. -2\nu_{12}\left( \frac{\partial \zeta ^{\ast
961: }}{\partial
962: x_{1}}\right)_{p,T}(Y_4+Y_6)\left(2\nu_{11}+2\nu_{22}-7\zeta^*\right)\right\},
963: \end{eqnarray}
964: \begin{eqnarray}
965: \label{4.n3} \ell_1^*&=&\frac{1}{\Delta}\left\{-2Y_3\left[2
966: (\nu_{12}\nu_{21}-\nu_{11}\nu_{22})\nu_{22}+\zeta^*(7\nu_{11}\nu_{22}-5\nu_{12}\nu_{21}+2\nu_{22}^2
967: -6\nu_{11}\zeta^*-7\nu_{22}\zeta^*+6\zeta^{*2})\right]\right.\nonumber\\
968: &&
969: +2Y_4\nu_{12}\left[2\nu_{12}\nu_{21}-2\nu_{11}\nu_{22}+2\zeta^*(\nu_{11}+\nu_{22})
970: -\zeta^{*2}\right]\nonumber\\
971: & +&
972: \left.2Y_5\zeta^*\left[2\nu_{12}\nu_{21}+\nu_{22}(2\nu_{22}-7\zeta^*)+6\zeta^{*2}\right]
973: -2\nu_{12}\zeta^*Y_6\left[2(\nu_{11}+\nu_{22})-7\zeta^*\right]
974: \right\},
975: \end{eqnarray}
976: \begin{eqnarray}
977: \label{4.n4}
978: \lambda_1^*&=&\frac{1}{\Delta}\left\{-Y_3\zeta^*\left[2
979: \nu_{12}\nu_{21}+\nu_{22}(2\nu_{22}-7\zeta^*)+6\zeta^{*2}\right]+\nu_{12}\zeta^*
980: Y_4\left[2(\nu_{11}+\nu_{22})-7\zeta^*\right] \right. \nonumber\\
981: &
982: &-Y_5\left[4\nu_{12}\nu_{21}(\nu_{22}-\zeta^*)+2\nu_{22}^2(5\zeta^*-2\nu_{11})+2\nu_{11}
983: (7\nu_{22}\zeta^*-6\zeta^{*2})+5\zeta^{*2}(6\zeta^*-7\nu_{22})\right]\nonumber\\
984: & & \left.
985: +\nu_{12}Y_6\left[4\nu_{12}\nu_{21}+2\nu_{11}(5\zeta^*-2\nu_{22})+\zeta^*(10\nu_{22}-
986: 23\zeta^*)\right]\right\},
987: \end{eqnarray}
988: where
989: \begin{equation}
990: \label{4.n5}
991: \Delta=\left[4(\nu_{12}\nu_{21}-\nu_{11}\nu_{22})+6\zeta^*(\nu_{11}+\nu_{22})-9\zeta^{*2}\right]
992: \left[\nu_{12}\nu_{21}-\nu_{11}\nu_{22}+2\zeta^*(\nu_{11}+\nu_{22})-4\zeta^{*2}\right].
993: \end{equation}
994: In the above equations, the Y's are defined by Eqs.\
995: (\ref{4.14})--(\ref{4.16}), while the (reduced) collision
996: frequencies $\nu_{11}$ and $\nu_{12}$ are given by Eqs.\
997: (\ref{c18}) and (\ref{c19}), respectively. The expressions for
998: $d_2^{*}$, $\ell_2^*$, and $\lambda_2^*$ can be obtained from
999: Eqs.\ (\ref{4.n2})--(\ref{4.n5}) by setting $1\leftrightarrow 2$.
1000: As expected, our results for the heat flux show that $D^{\prime
1001: \prime }$ is antisymmetric with respect to the change
1002: $1\leftrightarrow 2$, while $L$ and $ \lambda $ are symmetric.
1003: Consequently, in the case of mechanically equivalent particles
1004: ($m_{1}=m_{2}\equiv m$, $\sigma _{1}=\sigma _{2}\equiv \sigma$,
1005: $\alpha _{ij}\equiv \alpha $), the coefficient $D^{\prime \prime}$
1006: vanishes.
1007: 
1008: In the three-dimensional case ($d=3$), all the above expressions
1009: for the transport coefficients reduce to those previously derived
1010: for hard spheres \cite{GD02,MGD06} when one takes Maxwellian
1011: distributions (\ref{2.18}) for the zeroth-order approximations
1012: $f_i^{(0)}$. For mechanically equivalent particles, the results
1013: obtained by Brey and Cubero \cite{BC01} for a $d$-dimensional
1014: monocomponent gas are also recovered. This confirms the
1015: self-consistency of the results derived here.
1016: 
1017: 
1018: \begin{figure}[htbp]
1019: \begin{center}
1020: \resizebox{8cm}{!}{\includegraphics{equi1.eps}}
1021: \end{center}
1022: \caption{Plot of the reduced pressure diffusion coefficient
1023: $D_p(\alpha)/D_p(1)$ as a function of the (common) coefficient of
1024: restitution $\alpha$ for binary mixtures with $x_1=0.2$,
1025: $\omega=1$ in the case of a three-dimensional system ($d=3$) and
1026: two values of the mass ratio $\mu$: $\mu=0.5$ (a) and  $\mu=4$
1027: (b). The solid lines refer to the results derived here and the
1028: dashed lines correspond to the results assuming the equality of
1029: the partial temperatures.} \label{fig4}
1030: \end{figure}
1031: \begin{figure}[htbp]
1032: \begin{center}
1033: \resizebox{8cm}{!}{\includegraphics{equi2bis.eps}}
1034: \end{center}
1035: \caption{Plot of the reduced thermal conductivity coefficient
1036: $\lambda(\alpha)/\lambda(1)$ as a function of the (common)
1037: coefficient of restitution $\alpha$ for binary mixtures with
1038: $x_1=0.2$, $\omega=1$ in the case of a three-dimensional system
1039: ($d=3$) and two values of the mass ratio $\mu$: $\mu=0.5$ (a) and
1040: $\mu=4$ (b). The solid lines refer to the results derived here and
1041: the dashed lines correspond to the results assuming the equality
1042: of the partial temperatures.} \label{fig5}
1043: \end{figure}
1044: 
1045: 
1046: 
1047: 
1048: 
1049: 
1050: \subsection{Comparison with other theories}
1051: 
1052: Before checking the accuracy of our expressions by comparing them
1053: with computer simulations, it is instructive first to make some
1054: comparison with previous results. \cite{mixture,SGNT06} These
1055: results assume energy equipartition ($T_i=T$) and so, they are
1056: based on a standard CE expansion around the Maxwellian
1057: (\ref{2.20.1}) instead of the local HCS distribution. Figures
1058: \ref{fig4} and \ref{fig5} show the dependence of the reduced
1059: pressure diffusion coefficient $D_p(\alpha)/D_p(1)$ and the
1060: reduced thermal conductivity coefficient
1061: $\lambda(\alpha)/\lambda(1)$, respectively, as a function of the
1062: (common) coefficient of restitution $\alpha_{ij}\equiv \alpha$ for
1063: $d=3$, $\omega=1$, $x_1=0.2$, and two different mass ratios $\mu$:
1064: $\mu=0.5$ (a) and  $\mu=4$ (b). Here, $D_p(1)$ and $\lambda(1)$
1065: are the values of $D_p$ and $\lambda$ for elastic collisions. We
1066: see that the deviation from the functional form for elastic
1067: collisions is quite important in both theories, even for moderate
1068: dissipation. It is apparent that the dependence of the transport
1069: coefficients on dissipation is quantitatively different in both
1070: models, especially at strong dissipation (say for instance,
1071: $\alpha=0.5$). This clearly shows the real quantitative effect of
1072: two different species temperatures on transport in granular
1073: mixtures.
1074: 
1075: 
1076: 
1077: 
1078: 
1079: 
1080: 
1081: 
1082: 
1083: \section{Comparison with Monte Carlo simulations}
1084: \label{sec4}
1085: 
1086: 
1087: A said before, the expressions derived in Sec.\ \ref{sec3} for the
1088: NS transport coefficients have been obtained by considering two
1089: different approximations. First, since the deviation of
1090: $f_i^{(0)}$ from its Maxwellian form (\ref{2.18}) is quite small
1091: in the region of thermal velocities, we have used the Maxwellian
1092: distribution (\ref{2.18}) as a trial function for $f_i^{(0)}$.
1093: Second, we have only considered the leading terms of an expansion
1094: of the distribution $f_i^{(1)}$ in Sonine polynomials. Both
1095: approximations allow one to offer a simplified kinetic theory for
1096: a $d$-dimensional granular binary mixture. To check the accuracy
1097: of the above predictions, in this Section we numerically solve the
1098: Boltzmann equation by means of the DSMC method \cite{Bird} and
1099: compare theory and simulation in the cases of the diffusion
1100: coefficient $D$ (in the tracer limit) and the shear viscosity
1101: coefficient $\eta$. Previous comparisons carried out for hard
1102: spheres \cite{GM04,MG03} when one takes into account the
1103: deviations of $f_i^{(0)}$ from their Maxwellian forms have shown
1104: good agreement between theory and simulation, even for strong
1105: dissipation (say $\alpha \gtrsim  0.5$). Here, we expect that such
1106: good agreement is also maintained in the case of hard disks when
1107: one replaces $f_i^{(0)}\to f_{i,M}$. Let us study each coefficient
1108: separately.
1109: 
1110: \subsection{Tracer diffusion coefficient}
1111: 
1112: We consider the special case in which one of the components of the
1113: mixture (say, for instance, species $1$) is present in tracer
1114: concentration ($x_1\to 0$). In this situation, $\partial
1115: \zeta^*/\partial x_1 \to 0$ and so, the expression (\ref{3.13})
1116: for the reduced diffusion coefficient $D^*$ becomes
1117: \begin{equation}
1118: \label{4.1} D^*=\frac{\gamma}{\nu^*-\frac{1}{2}\zeta^{*}},
1119: \end{equation}
1120: where now
1121: \begin{equation}
1122: \label{4.2}
1123: \zeta^{*}=\frac{\pi^{(d-1)/2}}{d\Gamma\left(d/2\right)}
1124: \left(\frac{\sigma_2}{\sigma_{12}}\right)^{d-1}\sqrt{2\mu_{12}}
1125: (1-\alpha_{22}^2),
1126: \end{equation}
1127: \begin{equation}
1128: \label{4.3} \nu^*=\frac{2\pi^{(d-1)/2}}{d\Gamma\left(d/2\right)}
1129: \mu_{21} \sqrt{\mu_{12}+\mu_{21}\gamma}
1130:  (1+\alpha_{12}).
1131: \end{equation}
1132: 
1133: 
1134: The diffusion coefficient of impurities in a granular gas
1135: undergoing homogeneous cooling state can be measured in simulation
1136: from the mean square displacement of
1137: the tracer particle after a time interval $t$: \cite{M89,GM04}
1138: \begin{equation}
1139: \label{4.4} \frac{\partial}{\partial t}\langle |{\bf r}(t)-{\bf
1140: r}(0)|^2 \rangle =\frac{2dD}{n_2}.
1141: \end{equation}
1142: Equation (\ref{4.4}) is the Einstein form. This relation
1143: (written in appropriate dimensionless variables to eliminate the
1144: time dependence of $D(t)$) was used in Ref.\ \onlinecite {GM04} to
1145: measure the diffusion coefficient for hard spheres. More details
1146: on this procedure can be found in Ref.\ \onlinecite {GM04}.
1147: 
1148: If the hydrodynamic description (or normal solution in the context
1149: of the CE method) applies, then the diffusion coefficient $D(t)$
1150: depends on time only through its dependence on the temperature
1151: $T(t)$. In this case, after a transient regime, the reduced
1152: diffusion coefficient $D^*=(m_1m_2/\rho)D(t)\nu_0(t)/T(t)$
1153: achieves a time-independent value. Here, we compare the steady
1154: state values of $D^*$ obtained from Monte Carlo simulations with
1155: the theoretical predictions given by the first Sonine
1156: approximation (\ref{4.1}). The dependence of $D^*$ on the common
1157: coefficient of restitution $\alpha_{ij}\equiv \alpha$ is shown in
1158: Fig.\ \ref{fig2} in the case of hard disks for three different
1159: systems. The symbols refer to computer simulations while the lines
1160: correspond to the kinetic theory results given by Eq.\
1161: (\ref{4.1}). Molecular dynamics (MD) results reported in Ref.\
1162: \onlinecite{BRCG00} when impurities and particles of the gas are
1163: mechanically equivalent have also been included. We observe that
1164: MD and DSMC results for $\mu=\omega=1$ are consistent among
1165: themselves in the range of values of $\alpha$ explored. This good
1166: agreement gives support to the applicability of the inelastic
1167: Boltzmann equation beyond the quasielastic limit. It is apparent
1168: that the agreement between the first Sonine approximation and
1169: simulation results is excellent when impurities and particles of
1170: the gas are mechanically equivalent and when impurities are much
1171: heavier and/or much larger than the particles of the gas (Brownian
1172: limit). However, some discrepancies between simulation an theory
1173: are found with decreasing values of the mass ratio $m_1/m_2$ and
1174: the size ratio $\sigma_1/\sigma_2$. These discrepancies are not
1175: easily observed in Fig.\ \ref{fig2} because of the small magnitude
1176: of $D^*$ for $\mu=1/4$. The above findings agree with those
1177: previously reported for hard spheres, \cite{GM04} where it was
1178: shown that the second Sonine approximation improves the
1179: qualitative predictions from the first Sonine approximation for
1180: the cases in which the gas particles are heavier and/or larger
1181: than impurities. The comparison carried out here for disks
1182: confirms the above expectations and shows that the Sonine
1183: polynomial expansion exhibits a slow convergence for sufficiently
1184: small values of the mass ratio $\mu$ and/or the size ratio
1185: $\omega$.
1186: 
1187: \begin{figure}[htbp]
1188: \begin{center}
1189: \resizebox{7cm}{!}{\includegraphics{fig2bis.eps}}
1190: \end{center}
1191: \caption{Plot of the reduced diffusion coefficient $D^*$ as a
1192: function of the (common) coefficient of restitution $\alpha$ for
1193: binary mixtures with $\omega=\mu$ in the case of a two-dimensional
1194: system ($d=2$). The symbols are computer simulation results
1195: obtained from the mean square displacement and the lines are the
1196: theoretical results obtained in the first Sonine approximation.
1197: The DSMC results correspond to $\mu=1/4$ ({\Large $\bullet$}),
1198: $\mu=4$ ({\Large $\circ$}) and $\mu=1$ ({\Large $\diamond$}).
1199: Molecular dynamics results reported in Ref.\ \onlinecite{BRCG00}
1200: for $\mu=1$ ($\triangle$) have also been included.} \label{fig2}
1201: \end{figure}
1202: 
1203: \subsection{Shear viscosity coefficient}
1204: 
1205: The shear viscosity $\eta$ is perhaps the most widely studied
1206: transport coefficient in granular fluids. In the case of granular
1207: mixtures, this coefficient has been measured \cite{MG03} when the
1208: system is {\em heated} by the action of an external driving force
1209: (thermostat) that exactly compensates for cooling effects
1210: associated with dissipation of collisions. The corresponding shear
1211: viscosity of the mixture (which slightly differs from the one
1212: obtained in the free cooling case) has been determined by means of
1213: the CE method in the low-density regime \cite{MG03} as well as for
1214: a moderate dense mixture. \cite{GM03} The theoretical predictions
1215: compare reasonably well with the corresponding numerical solutions
1216: of the Boltzmann and Enskog kinetic equations.
1217: 
1218: More recently, a new alternative method has been proposed to
1219: measure the (true) NS shear viscosity coefficient. \cite{MSG05}
1220: The method is based on the simple shear flow state modified by the
1221: introduction of a deterministic non-conservative force (which
1222: compensates for the collisional cooling) along with a stochastic
1223: process. While the external force is introduced to allow the
1224: granular fluid to approach a Newtonian regime, the stochastic
1225: process is introduced to mimic the conditions appearing in the CE
1226: method to NS order. Although the method was originally devised to
1227: a single granular gas, its extension to multicomponent systems is
1228: straightforward. Here, we use this procedure to measure the shear
1229: viscosity of the mixture by means of the DSMC method. More
1230: technical details on this procedure and its application to dense
1231: gases can be found in Ref.\ \onlinecite{MSG05}.
1232: \begin{figure}[htbp]
1233: \begin{center}
1234: \resizebox{7cm}{!}{\includegraphics{fig3.eps}}
1235: \end{center}
1236: \caption{Plot of the reduced shear viscosity
1237: $\eta^*(\alpha)=\eta(\alpha)/\eta(1)$ as a function of the
1238: (common) coefficient of restitution $\alpha$ for binary mixtures
1239: constituted by particles of the same mass density
1240: ($\omega=\mu^{1/2}$) in the case of a two-dimensional system
1241: ($d=2$). The symbols are computer simulation results and the lines
1242: are the theoretical results obtained in the first Sonine
1243: approximation. The DSMC results correspond to $\mu=1$ ({\Large
1244: $\circ$}), $\mu=4$ ({\Large $\bullet$}) and $\mu=8$ ({\Large
1245: $\diamond$}). We have also included DSMC results obtained in Ref.\
1246: \onlinecite{BRM04} for $\mu=1$ ($\triangle$) from the Green-Kubo
1247: relation.} \label{fig3}
1248: \end{figure}
1249: 
1250: Comparison between the first Sonine approximation and computer
1251: simulations for $\eta^*(\alpha)=\eta(\alpha)/\eta(1)$ is shown in
1252: Fig.\ \ref{fig3} for three different mixtures constituted by
1253: particles of the same mass density (i.e., $\mu=\omega^d$) in the
1254: case of a two-dimensional system. Here, $\eta(1)$ refers to the
1255: elastic value for the shear viscosity coefficient and we have
1256: assumed again a common value of the coefficient of restitution
1257: $\alpha$. The symbols represent the simulation data and the lines
1258: correspond to the theoretical results. We have also included
1259: recent simulation results \cite{BRM04} for $\eta^*$ obtained from
1260: the Green-Kubo relation in the one-component case. Good agreement
1261: among the data presented here and those reported in Ref.\
1262: \onlinecite{BRM04} for $\mu=1$ is observed. In addition, as
1263: happens for hard spheres, \cite{MG03} we see that in general the
1264: agreement between the first Sonine approximation and simulation is
1265: quite good. At a quantitative level, the theory slightly
1266: overestimates the simulation data, especially for strong
1267: dissipation and for mixtures of particles of different masses
1268: and/or sizes. However, such discrepancies are quite small since
1269: for instance, they are smaller than 3\% at $\alpha=0.5$ for
1270: $\mu=8$ and $\omega=\sqrt{8}$. This shows again the reliability of
1271: the first Sonine approximation for the shear viscosity
1272: coefficient. It must be noted that this conclusion cannot in
1273: principle be extended to the transport coefficients associated with
1274: the heat flux since recent comparisons for a single gas
1275: \cite{BRM04,BRMMG05,MSG06} have shown significant discrepancies between
1276: the first Sonine approximation and computer simulations for high
1277: inelasticity (say $\alpha \lesssim 0.7$). In this case, the
1278: agreement between theory and simulation can be significantly
1279: improved by the use of a modified first Sonine approximation.
1280: \cite{GSM06}
1281: 
1282: 
1283: \section{Discussion}
1284: \label{sec5}
1285: 
1286: The main objective of this work has been to obtain the NS
1287: transport coefficients of a granular binary mixture at low
1288: density. In contrast to previous works, \cite{mixture,SGNT06} the
1289: present study is based on a modified CE solution of the inelastic
1290: Boltzmann equation that takes into account non-equipartition of
1291: energy. There is no phenomenology involved as the equations and
1292: the transport coefficients have been derived systematically from
1293: the inelastic Boltzmann equation by the CE expansion around the
1294: local HCS. Since the spatial gradients are assumed to be
1295: independent of the coefficients of restitution, although the NS
1296: equations restrict their applicability to first order in gradients
1297: the corresponding transport coefficients hold {\em a priori} to
1298: arbitrary degree of inelasticity. All the calculations have been
1299: performed in an arbitrary number $d$ of dimensions, previous
1300: results \cite{GD02} being recovered for $d=3$.
1301: 
1302: The constitutive equations to NS order for the mass flux, the
1303: stress tensor, and the heat flux are given  by Eqs.\
1304: (\ref{3.1})--(\ref{3.3}), respectively. The associated transport
1305: coefficients are the mutual diffusion coefficient $D$, the
1306: pressure diffusion coefficient $D_p$, and the thermal diffusion
1307: coefficient $D'$ in the case of the mass flux, the shear viscosity
1308: coefficient $\eta$ for the pressure tensor, and the Dufour
1309: coefficient $D''$, the pressure energy coefficient $L$, and the
1310: thermal conductivity $\lambda$ in the case of the heat flux. These
1311: coefficients are determined from the solutions of the set of
1312: coupled linear integral equations (\ref{3.10})--(\ref{3.12.0}). In
1313: addition, the NS transport coefficients also depend on the
1314: reference distributions $f_i^{(0)}$, which are not Maxwellians
1315: because they obey the integral equations (\ref{2.17.5}). To solve
1316: the above integral equations and provide good estimates for the
1317: transport coefficients, we have considered two approximations: (i)
1318: the distributions $f_i^{(0)}$ have been replaced by their
1319: Maxwellian forms (\ref{2.18}) at the temperature $T_i$ for that
1320: species and, (ii) we have only considered the leading terms in a
1321: series of Sonine polynomials for the first-order distribution
1322: $f_i^{(1)}$. By using both approximations, explicit expressions of
1323: the {\em seven} NS transport coefficients have been obtained as
1324: functions of the coefficients of restitution and the concentration
1325: and the ratios of mass and diameters. In dimensionless forms, the
1326: coefficients $D$, $D_p$, and $D'$ are given by Eqs.\
1327: (\ref{3.13})--(\ref{3.15}), respectively, the shear viscosity
1328: $\eta$ is given by Eqs.\ (\ref{3.17}) and (\ref{3.18}), while the
1329: expressions of the coefficients $D''$, $L$, and $\lambda$ are
1330: provided by Eqs.\ (\ref{4.7})--(\ref{4.n5}).
1331: 
1332: 
1333: 
1334: \begin{figure}[htbp]
1335: \begin{center}
1336: \resizebox{7cm}{!}{\includegraphics{fig6.eps}}
1337: \end{center}
1338: \caption{Plot of the reduced shear viscosity coefficient
1339: $\eta^*=\eta(\alpha)/\eta(1)$ as a function of the (common)
1340: coefficient of restitution $\alpha$ for binary mixtures with
1341: $x_1=0.2$, $\omega=1$ and two values of the mass ratio $\mu$:
1342: $\mu=0.5$ (a) and $\mu=4$ (b). The solid lines refer to spheres
1343: ($d=3$) while the dashed lines correspond to disks ($d=2$).}
1344: \label{fig6}
1345: \end{figure}
1346: 
1347: 
1348: 
1349: 
1350: Previous results \cite{mixture} derived from the CE method have
1351: typically introduced additional assumptions for convenience that
1352: are not internally consistent with constructing a solution to the
1353: Boltzmann equation. Thus, in most of the cases the reference state
1354: $f_i^{(0)}$ has been chosen to be a Maxwellian at the same
1355: temperature [see Eq.\ (\ref{2.20.1})]. This assumption is presumed
1356: to give accurate results at weak dissipation where energy
1357: equipartition can be still considered as a good approximation.
1358: However, as shown in Fig.\ \ref{fig1}, the temperature ratio
1359: $T_1/T_2$ clearly differs from 1 as dissipation increases. Here,
1360: we have replaced $f_i^{(0)}\to f_{i,M}$ so that, the influence of
1361: the fourth-cumulants $c_i$ of $f_i^{(0)}$ has been
1362: ignored.\cite{GD99b} Comparison between the expressions derived in
1363: this paper by taking the approximation (\ref{2.18}) with those
1364: obtained by assuming energy equipartition shows important
1365: discrepancies as the coefficient of restitution decreases [see
1366: Figs.\ \ref{fig4} and \ref{fig5}]. Moreover, as an added value of
1367: our theory, the use of the Maxwellian approximation (\ref{2.18})
1368: for $f_i^{(0)}$ allows one to provide simple and explicit
1369: expressions for all the transport coefficients in terms of the
1370: parameters of the mixture. This contrasts with the relatively
1371: recent study for hard spheres \cite{GD02} where the constitutive
1372: relations for the fluxes were not explicitly displayed.
1373: 
1374: 
1375: 
1376: As a complementary route and to check the reliability of our
1377: theory, the analytical results derived for the diffusion
1378: coefficient $D$ and the shear viscosity $\eta$ in the first Sonine
1379: approximation have been compared with those obtained from
1380: numerical solutions of the Boltzmann equation by means of the DSMC
1381: method for a two-dimensional system. For the sake of simplicity,
1382: all the simulations have considered a common coefficient of
1383: restitution $\alpha\equiv \alpha_{ij}$. As expected, theory and
1384: simulation clearly show that the influence of dissipation on mass
1385: and momentum transport is quite important since there is a
1386: relevant dependence of the diffusion $D$ and viscosity $\eta$
1387: coefficients on $\alpha$. With respect to the accuracy of the
1388: theoretical predictions, we see that in general the CE results in
1389: the first Sonine approximation exhibit a good agreement with the
1390: simulation data. Exceptions to this agreement are extreme mass or
1391: size ratios and strong dissipation. These discrepancies are
1392: basically due to the use of the first Sonine approximation and can
1393: be partially mitigated by considering the second and third Sonine
1394: approximations \cite{GM04} or the use of a modified first Sonine
1395: approximation. \cite{GSM06}
1396: 
1397: 
1398: As said in the Introduction, the results obtained in this paper
1399: are of great practical interest since most of the experiments and
1400: simulations are performed in two dimensions. On the other hand,
1401: apart from this practical interest, the knowledge of the NS
1402: transport coefficients of a $d$-dimensional mixture allows one to
1403: investigate the influence of dimensionality on the transport
1404: properties of the system. To illustrate this effect, in Fig.\
1405: \ref{fig6} we plot the reduced shear viscosity
1406: $\eta^*\equiv\eta(\alpha)/\eta(1)$ versus the coefficient of
1407: restitution $\alpha$ for $\omega=1$, $x_1=0.2$, and two different
1408: mass ratios $\mu$: $\mu=0.5$ (a) and $\mu=4$ (b). We have
1409: considered the physical cases of hard spheres (solid lines) and
1410: hard disks (dashed lines). Although the qualitative dependence of
1411: $\eta^*$ on $\alpha$ is quite similar in both systems, we observe
1412: that the influence of dissipation on momentum transport is
1413: stronger for $d=3$ than for $d=2$. This trend is also observed in
1414: general in the remaining transport coefficients.
1415: 
1416: 
1417: 
1418: 
1419: One of the main limitations of our theory is its restriction to
1420: dilute gases. In this situation, the collisional transfer
1421: contributions to the fluxes are neglected and only their kinetic
1422: contributions are considered. Possible extension of the present
1423: kinetic theory to higher densities can be done in the context of
1424: the revised Enskog theory. Preliminary results \cite{GM03} have
1425: been focused on the uniform shear flow state to get directly the
1426: shear viscosity coefficient. The extension of this study
1427: \cite{GM03} to states with gradients of concentration, pressure,
1428: and temperature is somewhat intricate due to subtleties associated
1429: with the spatial dependence of the pair correlations functions
1430: considered in the revised Enskog theory. On the other hand, it
1431: must be remarked that many of the collision integrals appearing in
1432: the Enskog description are the same as those appearing in the
1433: Boltzmann limit so that one can take advantage of the results
1434: reported in this paper. We plan to extend the results derived for
1435: moderately dense mixtures of smooth {\em elastic} hard spheres
1436: \cite{MCK83} to inelastic collisions in the near future.
1437: 
1438: 
1439: 
1440: 
1441: \acknowledgments
1442: 
1443: Partial support of the Ministerio de Ciencia y Tecnolog\'{\i}a
1444: (Spain) through Grant No.  FIS2004-01399 (partially financed by
1445: FEDER funds) in the case of V.G. and ESP2003-02859 (partially
1446: financed by FEDER funds) in the case of J.M.M. is acknowledged. V.
1447: G. also acknowledges support from the European Community's Human
1448: Potential Programme HPRN-CT-2002-00307 (DYGLAGEMEM).
1449: 
1450: 
1451: 
1452: \appendix
1453: \section{Chapman-Enskog method}
1454: \label{appA}
1455: 
1456: 
1457: 
1458: The velocity distribution function $f_1^{(1)}$ obeys the equation
1459: \begin{equation}
1460: \left( \partial _{t}^{(0)}+{\cal L}_{1}\right) f_{1}^{(1)}+{\cal
1461: M} _{1}f_{2}^{(1)}=-\left( \partial _{t}^{(1)}+{\bf v}\cdot
1462: \nabla+ {\bf g}\cdot \frac{\partial}{\partial {\bf
1463: v}} \right)f_{1}^{(0)}\;, \label{a1}
1464: \end{equation}
1465: where the linear operators ${\cal L}_{1}$ and ${\cal M}_{1}$ are
1466: defined by Eqs.\ (\ref{3.12.1}) and (\ref{3.12.2}), respectively.
1467: A similar equation can be obtained for $f_2^{(1)}$ by
1468: interchanging $1\leftrightarrow 2$. The action of the time
1469: derivatives $\partial _{t}^{(1)}$ on the hydrodynamic fields is
1470: \begin{equation}
1471: D_{t}^{(1)}x_{1}=0,  \label{a2}
1472: \end{equation}
1473: \begin{equation}
1474: D_{t}^{(1)}p=-\frac{d+2}{d}p\nabla \cdot {\bf u},  \label{a3}
1475: \end{equation}
1476: \begin{equation}
1477: D_{t}^{(1)}T=-\frac{2T}{d}\nabla \cdot {\bf u},  \label{a4}
1478: \end{equation}
1479: \begin{equation}
1480: D_{t}^{(1)}{\bf u}=-\rho ^{-1}{\bf \nabla }p+{\bf g},  \label{a5}
1481: \end{equation}
1482: where $D_{t}^{(1)}=\partial _{t}^{(1)}+{\bf u\cdot \nabla }$ and
1483: use has been made of the results ${\bf j}_{i}^{(0)}={\bf
1484: q}^{(0)}=\zeta ^{(1)}=0$. The last equality follows from the fact
1485: that the cooling rate is a scalar, and corrections to first order
1486: in the gradients can arise only from the divergence of a vector
1487: field. However, as is demonstrated  below, there is no
1488: contribution to the distribution function proportional to this
1489: divergence. We note that this is special to the low density
1490: Boltzmann equation and such terms do occur at higher densities.
1491: \cite{GD99a} Use of Eqs.\ (\ref{a2})--(\ref{a5}) yields
1492: \begin{eqnarray}
1493: -\left( \partial _{t}^{(1)}+{\bf v}\cdot \nabla + {\bf g}\cdot
1494: \frac{\partial}{\partial {\bf v}} \right)
1495: f_{1}^{(0)} &=&-\left( \frac{\partial }{\partial
1496: x_{1}}f_{1}^{(0)}\right) _{p,T}{\bf V} \cdot \nabla x_{1}-\left[
1497: f_{1}^{(0)}{\bf V}+\frac{n T}{\rho}\left( \frac{
1498: \partial }{\partial {\bf V}}f_{1}^{(0)}\right) \right] \cdot \nabla \ln p
1499: \nonumber \\
1500: &&+\left[ f_{1}^{(0)}+\frac{1}{2}\frac{\partial}{\partial {\bf
1501: V}}\cdot \left({\bf V}f_{1}^{(0)}\right) \right] {\bf V}\cdot
1502: \nabla \ln T
1503: \nonumber \\
1504: &&+\left( V_{k}\frac{\partial}{\partial V_{\ell
1505: }}f_{1}^{(0)}-\frac{1}{d}\delta _{k\ell}{\bf V}\cdot
1506: \frac{\partial}{\partial {\bf V}} f_{1}^{(0)}\right)
1507: \nabla _{k}u_{\ell}. \label{a5bis}
1508: \end{eqnarray}
1509: Note that the external field does not appear in the right-hand
1510: side of Eq.\ (\ref{a5bis}). This is due to the particular form of
1511: the gravitational force. Using Eq.\ (\ref{a5bis}), Eq.\ (\ref{a1})
1512: can be written as
1513: \begin{equation}
1514: \left( \partial _{t}^{(0)}+{\cal L}_{1}\right) f_{1}^{(1)}+{\cal
1515: M} _{1}f_{2}^{(1)}={\bf A}_{1}\cdot \nabla x_{1}+{\bf B}_{1}\cdot
1516: \nabla p+{\bf C}_{1}\cdot \nabla T +{\sf D}_{1}:
1517: \nabla {\bf u}, \label{a6}
1518: \end{equation}
1519: where
1520: \begin{equation}
1521: {\bf A}_{i}({\bf V})=-\left(\frac{\partial}{\partial x_{1}}
1522: f_{i}^{(0)}\right)_{p,T}{\bf V}, \label{a7}
1523: \end{equation}
1524: \begin{equation}
1525: {\bf B}_{i}({\bf V})=-\frac{1}{p}\left[ f_{i}^{(0)}{\bf
1526: V}+\frac{nT}{\rho } \left(\frac{\partial}{\partial {\bf
1527: V}}f_{i}^{(0)}\right) \right] , \label{a8}
1528: \end{equation}
1529: \begin{equation}
1530: {\bf C}_{i}({\bf V})=\frac{1}{T}\left[
1531: f_{i}^{(0)}+\frac{1}{2}\frac{\partial }{\partial {\bf V}}\cdot
1532: \left( {\bf V}f_{i}^{(0)}\right) \right] {\bf V}, \label{a9}
1533: \end{equation}
1534: \begin{equation}
1535: {\sf D}_{i}({\bf V})={\bf V}\frac{\partial }{\partial {\bf V}}
1536: f_{i}^{(0)}-\frac{1}{d}\openone {\bf V}\cdot \frac{\partial
1537: }{\partial {\bf V}}f_{i}^{(0)}.  \label{a10}
1538: \end{equation}
1539: In Eqs.\ (\ref{a7})--(\ref{a10}) it is understood that $i=1,2$ and $\openone$ is the unit
1540: tensor in $d$ dimensions. Note that the trace of ${\sf D}_{i}$ vanishes, confirming that the
1541: distribution function does not have contribution from the divergence of
1542: the flow field.  The solutions to Eqs.\ (\ref{a6}) are of the form
1543: \begin{equation}
1544: f_{i}^{(1)}={\boldsymbol {\cal A}}_{i}\cdot \nabla
1545: x_{1}+{\boldsymbol {\cal B}}_{i}\cdot \nabla p+{\boldsymbol {\cal
1546: C}}_{i}\cdot \nabla T+{\cal D}_{i,k\ell }\nabla _{k}u_{\ell}\;.
1547: \label{a11}
1548: \end{equation}
1549: The coefficients ${\boldsymbol {\cal A}}_{i}$, ${\boldsymbol {\cal
1550: B}}_{i}$, ${\boldsymbol {\cal C}}_{i}$, and ${\boldsymbol {\cal D}}_{i}$
1551: are functions of the peculiar velocity ${\bf V}$ and the
1552: hydrodynamic fields. The cooling rate depends on space through its
1553: dependence on $x_{1}$, $p$, and $T$. The time derivative $\partial
1554: _{t}^{(0)}$ acting on these quantities can be evaluated by the
1555: replacement $\partial _{t}^{(0)}\rightarrow -\zeta ^{(0)}\left(
1556: T\partial _{T}+p\partial _{p}\right) $. In addition, there are
1557: contributions from $\partial _{t}^{(0)}$ acting on the temperature
1558: and pressure gradients given by
1559: \begin{eqnarray}
1560: \partial _{t}^{(0)}\nabla T &=&-\nabla \left( T\zeta ^{(0)}\right)
1561: =-\zeta
1562: ^{(0)}\nabla T-T\nabla \zeta ^{(0)}  \nonumber \\
1563: &=&-\frac{\zeta ^{(0)}}{2}\nabla T -T\left[ \left( \frac{\partial
1564: \zeta ^{(0)}}{\partial x_{1}}\right) _{p,T}\nabla x_{1}
1565: +\frac{\zeta ^{(0)}}{p} \nabla p\right], \label{a12}
1566: \end{eqnarray}
1567: \begin{eqnarray}
1568: \partial _{t}^{(0)}\nabla p &=&-\nabla \left( p\zeta ^{(0)}\right) =-\zeta
1569: ^{(0)}\nabla p-p\nabla \zeta ^{(0)}  \nonumber \\
1570: &=&-2\zeta ^{(0)}\nabla p-p\left[ \left( \frac{\partial \zeta
1571: ^{(0)}}{
1572: \partial x_{1}}\right) _{p,T}\nabla x_{1}-\frac{\zeta ^{(0)}}{2T}\nabla T
1573: \right] . \label{a13}
1574: \end{eqnarray}
1575: The corresponding integral equations for the unknowns
1576: ${\boldsymbol {\cal A}}_{i}$, ${\boldsymbol {\cal B}}_{i}$,
1577: ${\boldsymbol {\cal C}}_{i}$, and ${\boldsymbol {\cal D}}_{i}$ are
1578: identified as the coefficients of the independent gradients in
1579: (\ref{a11}). This leads to Eqs.\ (\ref{3.10})--(\ref{3.12.0}).
1580: 
1581: 
1582: \section{Leading Sonine approximations}
1583: \label{appB}
1584: 
1585: In this Appendix, we get the explicit expressions of the mass,
1586: momentum, and heat fluxes in the first Sonine approximation and
1587: neglecting the non-Gaussian corrections to the reference
1588: distributions $f_i^{(0)}$ (i.e., the cumulants $c_i=0$). The
1589: procedure to get the leading order contributions in the Sonine
1590: polynomial expansion to the transport coefficients is quite
1591: similar to the one previously used in the three-dimensional case.
1592: \cite{GD02} Only some partial results will be presented here.
1593: 
1594: 
1595: \subsection{Leading Sonine approximation to Mass Flux}
1596: 
1597: In the case of the mass flux, the leading Sonine approximations
1598: (lowest degree polynomial) of the quantities ${\boldsymbol {\cal
1599: A}}_{i}$, ${\boldsymbol {\cal B}}_{i}$, ${\boldsymbol {\cal
1600: C}}_{i}$ are
1601: \begin{equation}
1602: \label{b1} {\boldsymbol {\cal A}}_{1}({\bf V})\to -f_{1,M}{\bf
1603: V}\frac{m_1m_2n}{\rho n_1T_1}D,\quad {\boldsymbol {\cal
1604: A}}_{2}({\bf V})\to f_{2,M}{\bf V}\frac{m_1m_2n}{\rho n_2T_2}D
1605: \end{equation}
1606: \begin{equation}
1607: \label{b2} {\boldsymbol {\cal B}}_{1}({\bf V})\to -f_{1,M}{\bf
1608: V}\frac{\rho}{p n_1T_1}D_p ,\quad {\boldsymbol {\cal B}}_{2}({\bf
1609: V})\to f_{2,M}{\bf V}\frac{\rho}{p n_2T_2}D_p
1610: \end{equation}
1611: \begin{equation}
1612: \label{b3} {\boldsymbol {\cal C}}_{1}({\bf V})\to -f_{1,M}{\bf
1613: V}\frac{\rho}{T n_1T_1}D',\quad {\boldsymbol {\cal C}}_{2}({\bf
1614: V})\to f_{2,M}{\bf V}\frac{\rho}{T n_2T_2}D',
1615: \end{equation}
1616: where $f_{i,M}$ are the Maxwellian distributions (\ref{2.18}).
1617: Multiplication of Eqs.\ (\ref{3.10})--(\ref{3.12}) by $m_1 {\bf
1618: V}$ and integrating over the velocity yields
1619: \begin{equation}
1620: \left[ -\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
1621: +\nu \right] \left(-\frac{m_1m_2n}{\rho}D\right)=-\left(\frac{
1622: \partial }{\partial x_{1}}n_{1}T_{1}\right)_{p,T}
1623: -\rho\left( \frac{\partial \zeta ^{(0)}}{\partial x_{1}}\right)
1624: _{p,T}\left(D_p+D'\right) ,  \label{b4}
1625: \end{equation}
1626: \begin{equation}
1627: \left[ -\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
1628: -2\zeta ^{(0)}+\nu \right]\left(-\frac{\rho}{p}D_p\right) =-\frac{
1629: n_{1}T_{1}}{p}\left( 1-\frac{m_{1}nT}{\rho T_{1}}\right)
1630: -\frac{\rho\zeta ^{(0)}}{p}D',  \label{b5}
1631: \end{equation}
1632: \begin{equation}
1633: \left[-\zeta ^{(0)}\left( T\partial _{T}+p\partial _{p}\right)
1634: -\frac{1}{2} \zeta ^{(0)}+\nu \right]
1635: \left(-\frac{\rho}{T}D'\right)= \frac{\rho\zeta ^{(0)}}{2T}D_p.
1636: \label{b6}
1637: \end{equation}
1638: Here, $\nu $ is the collision frequency defined by
1639: \begin{eqnarray}
1640: \nu  &=&\frac{1}{dn_{1}T_{1}}\int d{\bf V}_{1}m_{1}{\bf
1641: V}_{1}\cdot \left[ {\cal L}_{1}(f_{1,M}{\bf V}_{1})-\delta \gamma
1642: {\cal M}_{1}(f_{2,M}{\bf V}_{2})
1643: \right]   \nonumber \\
1644: &=&-\frac{1}{dn_{1}T_{1}}\int d{\bf V}_{1}m_{1}{\bf V}_{1}\cdot
1645: \left( J_{12}[{\bf v}_{1}|f_{1,M}{\bf V}_{1},f_{2}^{(0)}]-\delta
1646: \gamma J_{12}[{\bf v}_{1}|f_{1}^{(0)},f_{2,M}{\bf V}_{2}]\right),
1647: \label{b7}
1648: \end{eqnarray}
1649: where $\delta\equiv x_1/x_2$. The evaluation of the collision integral
1650: (\ref{b7}) is made in Appendix \ref{appC}. The self-collision
1651: terms of ${\cal L}_{i}$ arising from $J_{11}$ do not occur in Eq.\
1652: (\ref{b7}) since they conserve momentum for species $1$. From
1653: dimensional analysis, $D\sim T^{1/2}$, $ D_p\sim p T^{-1/2}$, and
1654: $D'\sim p T^{-1/2}$ so the temperature and pressure derivatives
1655: can be performed in Eqs.\ (\ref{b4})--(\ref{b6}). After performing
1656: them, one gets the expressions (\ref{3.13}), (\ref{3.14}), and
1657: (\ref{3.15}) for the (reduced) coefficients $D^*$, $D_p^*$, and
1658: $D^{\prime\ast}$, respectively.
1659: 
1660: \subsection{Leading Sonine approximation to Pressure Tensor}
1661: 
1662: 
1663: In the case of the pressure tensor, the leading Sonine
1664: approximation for the function ${\cal D}_{i,k\ell}$ is
1665: \begin{equation}
1666: \label{b8} {\cal D}_{i,k\ell}({\bf V})\to -f_{i,M}({\bf V}) \frac{\eta_{i}}{T}
1667: R_{i,k\ell}({\bf V}),\quad i=1,2
1668: \end{equation}
1669: where
1670: \begin{equation}
1671: \label{b9} R_{i,k\ell}({\bf V})=m_i\left( V_{k}V_{\ell}-
1672: \frac{1}{d}V^2\delta_{k\ell}\right),
1673: \end{equation}
1674: and
1675: \begin{equation}
1676: \label{b10} \eta_i=-\frac{1}{(d-1)(d+2)}\frac{T}{n_iT_i^2}\int
1677: d{\bf v} R_{i,k\ell}({\bf V}){\cal D}_{i,k\ell}({\bf V}).
1678: \end{equation}
1679: The shear viscosity $\eta$ in this approximation can be written as
1680: \begin{equation}
1681: \label{b11} \eta=\frac{p}{\nu_0}\left(x_1\gamma_1^2
1682: \eta_1^*+x_2\gamma_2^2 \eta_2^*\right),
1683: \end{equation}
1684: where $\eta_i^*=\nu_0\eta_i$. The integral equations for the
1685: (reduced) coefficients $\eta_i^*$ are decoupled from the remaining
1686: transport coefficients. The two coefficients $\eta_{i}^*$ are
1687: obtained by multiplying Eqs.\ (\ref{3.12.0}) with $R_{i,k\ell}$
1688: and integrating over the velocity to get the coupled set of
1689: equations
1690: \begin{equation}
1691: \label{b12} \left(
1692: \begin{array}{cc}
1693: \tau_{11}-\frac{1}{2}\zeta^{*}& \tau_{12}\\
1694: \tau_{21}&\tau_{22}-\frac{1}{2}\zeta^{*}
1695: \end{array}
1696: \right) \left(
1697: \begin{array}{c}
1698: \eta_{1}^*\\
1699: \eta_{2}^*
1700: \end{array}
1701: \right) =\left(
1702: \begin{array}{c}
1703: \gamma_1^{-1}\\
1704: \gamma_2^{-1}
1705: \end{array}
1706: \right).
1707: \end{equation}
1708: The (reduced) collision frequencies $\tau_{ij}$ are given in terms
1709: of the linear collision operator by
1710: \begin{equation}
1711: \label{b13}
1712: \tau_{ii}=\frac{1}{(d-1)(d+2)}\frac{1}{n_iT_i^2\nu_0}\int d{\bf
1713: v}_1 R_{i,k\ell} {\cal L}_i\left(f_{i,M}R_{i,k\ell}\right),
1714: \end{equation}
1715: \begin{equation}
1716: \label{b14}
1717: \tau_{ij}=\frac{1}{(d-1)(d+2)}\frac{1}{n_iT_i^2\nu_0}\int d{\bf
1718: v}_1 R_{i,k\ell} {\cal M}_i\left(f_{j,M}R_{j,k\ell}\right) , \quad
1719: i\neq j.
1720: \end{equation}
1721: The evaluation of these collision integrals is also given in
1722: Appendix \ref{appC}. The solution of Eq.\ (\ref{b12})
1723: is elementary and yields Eq.\ (\ref{3.18}).
1724: 
1725: 
1726: \subsection{Leading Sonine approximation to Heat Flux}
1727: 
1728: The heat flux requires going up to the second Sonine
1729: approximation. In this case, the quantities ${\boldsymbol {\cal
1730: A}}_{i}$, ${\boldsymbol {\cal B}}_{i}$, ${\boldsymbol {\cal
1731: C}}_{i}$ are taken to be
1732: \begin{equation}
1733: \label{b15} {\boldsymbol {\cal A}}_{1}({\bf V})\to
1734: f_{1,M}\left[-\frac{m_1m_2n}{\rho n_1T_1}D{\bf V}+d_1''{\bf
1735: S}_1({\bf V}) \right] ,\quad {\boldsymbol {\cal A}}_{2}({\bf
1736: V})\to f_{2,M}\left[\frac{m_1m_2n}{\rho n_2T_2}D{\bf V}+d_2''{\bf
1737: S}_2({\bf V})\right]
1738: \end{equation}
1739: \begin{equation}
1740: \label{b16} {\boldsymbol {\cal B}}_{1}({\bf V})\to
1741: f_{1,M}\left[-\frac{\rho}{p n_1T_1}D_p{\bf V}+\ell_1{\bf S}_1({\bf
1742: V}) \right] ,\quad {\boldsymbol {\cal B}}_{2}({\bf V})\to
1743: f_{2,M}\left[\frac{\rho}{p n_2T_2}D_p{\bf V}+\ell_2{\bf S}_2({\bf
1744: V})\right]
1745: \end{equation}
1746: \begin{equation}
1747: \label{b17} {\boldsymbol {\cal C}}_{1}({\bf V})\to
1748: f_{1,M}\left[-\frac{\rho}{T n_1T_1}D'{\bf V}+\lambda_1{\bf
1749: S}_1({\bf V}) \right] ,\quad {\boldsymbol {\cal C}}_{2}({\bf
1750: V})\to f_{2,M}\left[\frac{\rho}{T n_2T_2}D'{\bf V}+\lambda_2{\bf
1751: S}_2({\bf V})\right],
1752: \end{equation}
1753: where
1754: \begin{equation}
1755: \label{b18} {\bf S}_i({\bf
1756: V})=\left(\frac{1}{2}m_iV^2-\frac{d+2}{2}T_i\right){\bf V}.
1757: \end{equation}
1758: In these equations, it is understood that $D$, $D_p$ and $D'$ are
1759: given by Eqs.\ (\ref{3.13}), (\ref{3.14}), and (\ref{3.15}),
1760: respectively. The coefficients $d_i''$, $\ell_i$ and $\lambda_i$
1761: are defined as
1762: \begin{equation}
1763: \left(
1764: \begin{array}{c}
1765: d_i'' \\
1766: \ell_i \\
1767: \lambda_i
1768: \end{array}
1769: \right) =\frac{2}{d(d+2)}\frac{m_i}{n_iT_i^3} \int d{\bf v}\,{\bf
1770: S}_i({\bf V})\cdot\left(
1771: \begin{array}{c}
1772: {\boldsymbol {\cal A}}_{i} \\
1773: {\boldsymbol {\cal B}}_{i} \\
1774: {\boldsymbol {\cal C}}_{i}
1775: \end{array}
1776: \right). \label{b19}
1777: \end{equation}
1778: These coefficients can be determined by multiplying Eqs.\
1779: (\ref{3.10})--(\ref{3.12}) (and their counterparts for the species
1780: 2) by ${\bf S}_i({\bf V})$ and integrating over the velocity. The
1781: final expressions can be obtained by taking into account that
1782: $d_1''\sim T^{-3/2}$, $\ell_1\sim T^{-3/2}/p$, and $\lambda_1\sim
1783: T^{-5/2}$ and the results
1784: \begin{equation}
1785: \int d{\bf v}\,m_{1}{\bf S}_1({\bf V})\cdot {\bf A}_{1}=-
1786: \frac{d(d+2)}{4}\frac{n_1T^2}{m_1} \left( \frac{
1787: \partial }{\partial x_{1}}\gamma_1^2\right) _{p,T},  \label{b20}
1788: \end{equation}
1789: \begin{equation}
1790: \int d{\bf v}\,m_{1}{\bf S}_{1}({\bf V})\cdot {\bf B}_{1}=0,
1791: \label{b21}
1792: \end{equation}
1793: \begin{equation}
1794: \int d{\bf v}\,m_{1}{\bf S}_{1}({\bf V})\cdot {\bf
1795: C}_{1}=-\frac{d(d+2)}{2} \frac{n_1T_1^2}{m_1T}. \label{b22}
1796: \end{equation}
1797: 
1798: 
1799: 
1800: By using matrix notation, the coupled set of six equations for the
1801: quantities
1802: \begin{equation}
1803: \label{4.10} \{d_1^*, d_2^*, \ell_1^*, \ell_2^*, \lambda_1^*,
1804: \lambda_2^*\}
1805: \end{equation}
1806: can be written as
1807: \begin{equation}
1808: \label{4.11} \Lambda_{\sigma \sigma'}X_{\sigma'}=Y_{\sigma}.
1809: \end{equation}
1810: Here, $d_i^*\equiv T\nu_0 d_i^{\prime\prime}$, $\ell_i^*\equiv
1811: pT\nu_0 \ell_i$, $\lambda_i^*\equiv T^2\nu_0 \lambda_i$,
1812: $X_{\sigma'}$ is the column matrix defined by the set (\ref{4.10})
1813: and $\Lambda_{\sigma \sigma'}$ is the square matrix
1814: \begin{equation}
1815: \label{4.12} \Lambda=\left(
1816: \begin{array} {cccccc}
1817: \nu_{11}-\frac{3}{2}\zeta^{*}& \nu_{12}& -\left(\frac{\partial
1818: \zeta ^{*}}{\partial x_{1}}\right)_{p,T}&0&
1819: -\left( \frac{\partial \zeta ^{*}}{\partial x_{1}}\right)_{p,T}&0 \\
1820: \nu_{21}&\nu_{22}-\frac{3}{2}\zeta^{*}&0& -\left( \frac{\partial
1821: \zeta ^{*}}{\partial x_{1}}\right)_{p,T}&0&
1822: -\left( \frac{\partial \zeta ^{*}}{\partial x_{1}}\right)_{p,T}\\
1823: 0& 0& \nu_{11}-\frac{5}{2}\zeta^{*}& \nu_{12}&
1824: -\zeta^{*}&0\\
1825: 0& 0 & \nu_{21} & \nu_{22}-\frac{5}{2}\zeta^{*} & 0 &
1826: -\zeta^{*}\\
1827: 0& 0& \zeta^{*}/2&0&\nu_{11}-\zeta^{*}& \nu_{12}\\
1828: 0& 0& 0&\zeta^{*}/2&\nu_{21}&\nu_{22}-\zeta^{*}
1829: \end{array}
1830: \right).
1831: \end{equation}
1832: The column matrix ${\bf Y}$ is
1833: \begin{equation}
1834: \label{4.13} {\bf Y}=\left(
1835: \begin{array}{c}
1836: Y_1\\
1837: Y_2\\
1838: Y_3\\
1839: Y_4\\
1840: Y_5\\
1841: Y_6
1842: \end{array}
1843: \right),
1844: \end{equation}
1845: where \cite{note}
1846: \begin{equation}
1847: \label{4.14} Y_1= \frac{D^*}{ x_1\gamma_1^2}\left(\omega_{12}-\zeta^{*}\right)-\frac{1}{ \gamma_1^2}
1848: \left(\frac{\partial \gamma_1}{\partial x_1} \right)_{p,T},\quad Y_2= -\frac{D^*}{
1849: x_2\gamma_2^2}\left(\omega_{21}-\zeta^{*}\right)-\frac{1}{ \gamma_2^2} \left(\frac{\partial \gamma_2}{\partial
1850: x_1} \right)_{p,T},
1851: \end{equation}
1852: \begin{equation}
1853: \label{4.15} Y_3=\frac{
1854: D_p^*}{x_1\gamma_1^2}\left(\omega_{12}-\zeta^{*}\right),\quad
1855: Y_4=-\frac{
1856: D_p^*}{x_2\gamma_2^2}\left(\omega_{21}-\zeta^{*}\right),
1857: \end{equation}
1858: \begin{equation}
1859: \label{4.16} Y_5= -\frac{1}{\gamma_1}+\frac{D^{\prime\ast}}{
1860: x_1\gamma_1^2}\left(\omega_{12}-\zeta^{*}\right),\quad Y_6=
1861: -\frac{1}{\gamma_2}-\frac{D^{\prime\ast}}{
1862: x_2\gamma_2^2}\left(\omega_{21}-\zeta^{*}\right).
1863: \end{equation}
1864: Here, we have introduced the (reduced) collision frequencies
1865: \begin{equation}
1866: \label{4.17}
1867: \nu_{ii}=\frac{2}{d(d+2)}\frac{m_i}{n_iT_i^3\nu_0}\int d{\bf v}_1
1868: {\bf S}_i \cdot {\cal L}_i\left(f_{i,M}{\bf S}_i\right),
1869: \end{equation}
1870: \begin{equation}
1871: \label{4.18}
1872: \nu_{ij}=\frac{2}{d(d+2)}\frac{m_i}{n_iT_i^3\nu_0}\int d{\bf v}_1
1873: {\bf S}_i \cdot {\cal M}_i\left(f_{j,M}{\bf S}_j\right), \quad
1874: i\neq j,
1875: \end{equation}
1876: \begin{equation}
1877: \label{4.19} \omega_{12}=\frac{2}{d(d+2)}\frac{m_1}{n_1T_1^2\nu_0}
1878: \left[\int d{\bf v}_1 {\bf S}_1\cdot {\cal L}_1(f_{1,M}{\bf
1879: V}_1)-\delta \gamma \int d{\bf v}_1 {\bf S}_1\cdot {\cal
1880: M}_1(f_{2,M}{\bf V}_2)\right],
1881: \end {equation}
1882: \begin{equation}
1883: \label{4.19.1}
1884: \omega_{21}=\frac{2}{d(d+2)}\frac{m_2}{n_2T_2^2\nu_0} \left[\int
1885: d{\bf v}_1 {\bf S}_2\cdot {\cal L}_2(f_{2,M}{\bf
1886: V}_1)-\frac{1}{\delta \gamma} \int d{\bf v}_1 {\bf S}_2\cdot {\cal
1887: M}_2(f_{1,M}{\bf V}_2)\right].
1888: \end {equation}
1889: The expressions of the collision integrals (\ref{4.17}),
1890: (\ref{4.18}), and (\ref{4.19}) are given in Appendix \ref{appC}.
1891: The solution to Eq.\ (\ref{4.11}) is
1892: \begin{equation}
1893: \label{4.20} X_{\sigma}=\left(\Lambda^{-1}\right)_{\sigma
1894: \sigma'}Y_{\sigma'}.
1895: \end{equation}
1896: From this relation one gets the expressions (\ref{4.n2}),
1897: (\ref{4.n3}), and (\ref{4.n4}) for the coefficients $d_i^*$,
1898: $\ell_i^*$ and $\lambda_i^*$, respectively.
1899: 
1900: 
1901: 
1902: 
1903: 
1904: \section{Collision integrals}
1905: \label{appC}
1906: 
1907: In this Appendix we compute the different collision integrals
1908: appearing in the expressions of the transport coefficients. To
1909: simplify all the integrals, we use the property
1910: \begin{equation}
1911: \int d{\bf v}_{1}h({\bf V}_{1})J_{ij}\left[ {\bf
1912: v}_{1}|f_{i},f_{j}\right] =\sigma _{ij}^{d-1}\int \,d{\bf
1913: v}_{1}\,\int \,d{\bf v}_{2}f_{i}( {\bf V}_{1})f_{j}({\bf V}_{2})
1914: \int d\widehat{\boldsymbol {\sigma }}\,\Theta
1915: (\widehat{\boldsymbol {\sigma}} \cdot {\bf
1916: g}_{12})(\widehat{\boldsymbol {\sigma }}\cdot {\bf
1917: g}_{12})\,\left[ h( {\bf V}_{1}^{^{\prime \prime }})-h({\bf
1918: V}_{1})\right] \;,  \label{c1}
1919: \end{equation}
1920: with
1921: \begin{equation}
1922: {\bf V}_{1}^{^{\prime \prime }}={\bf V}_{1}-\mu _{ji}(1+\alpha
1923: _{ij})( \widehat{\boldsymbol {\sigma }}\cdot {\bf
1924: g}_{12})\widehat{\boldsymbol {\sigma}}\;. \label{c2}
1925: \end{equation}
1926: This result applies for both $i=j$ and $i\neq j$.
1927: 
1928: Let us start with the collision frequency $\nu$ defined by Eq.\
1929: (\ref{b7}). Use of the identity (\ref{c2}) in (\ref{b7}) gives
1930: \begin{eqnarray}
1931: \nu &=&\frac{m_1}{dn_1T_1}B_3\sigma _{12}^{d-1}\mu _{21}(1+\alpha
1932: _{12}) \int \,d{\bf V}_{1}\,\int \,d{\bf V}_{2}\,g_{12}\left[
1933: f_{1,M}( {\bf V}_{1})f_{2}^{(0)}({\bf V}_{2})({\bf V}_{1}\cdot
1934: {\bf g}_{12})\right.
1935: \nonumber \\
1936: &&\left. -\delta \gamma f_{1}^{(0)}({\bf V}_{1})f_{2,M}({\bf
1937: V}_{2})({\bf V} _{2}\cdot {\bf g}_{12})\right],  \label{c3}
1938: \end{eqnarray}
1939: where use has been made of the result
1940: \begin{equation}
1941: \label{c4} \int d\widehat{\boldsymbol{\sigma}}\, \Theta
1942: (\widehat{{\boldsymbol {\sigma }}} \cdot {\bf g}_{12})\,
1943: (\widehat{\boldsymbol{\sigma}}\cdot {\bf g}_{12})^k
1944: \widehat{\boldsymbol{\sigma}}=B_{k+1} g_{12}^{k-1}{\bf g}_{12},
1945: \end{equation}
1946: whith \cite{EB02}
1947: \begin{equation}
1948: \label{c5} B_k\equiv \int d\widehat{\boldsymbol{\sigma}}\, \Theta
1949: (\widehat{{\boldsymbol {\sigma }}} \cdot {\bf g}_{12})\,
1950: (\widehat{\boldsymbol{\sigma}}\cdot {\widehat{\bf
1951: g}}_{12})^k=\pi^{(d-1)/2} \frac{
1952: \Gamma\left(\frac{k+1}{2}\right)}{\Gamma\left(\frac{k+d}{2}\right)}.
1953: \end{equation}
1954: Substitution of the Maxwellian approximation (\ref{2.18}) for
1955: $f_i^{(0)}$ gives
1956: \begin{equation}
1957: \nu =\frac{2}{d}\frac{\pi^{(d-1)/2}}
1958: {\Gamma\left(\frac{d+3}{2}\right)} \nu_0 (1+\alpha _{12}) \pi^{-d}
1959: \left(\theta_1\theta_2\right)^{d/2} \int \,d{\bf c}_{1}\,\int
1960: \,d{\bf c}_{2}\,y\, e^{-(\theta
1961: _{1}c_{1}^{2}+\theta_{2}c_{2}^{2})}\left[ x_2\gamma_1^{-1}({\bf
1962: c}_{1}\cdot {\bf y}) -x_1\gamma_2^{-1}({\bf c}_{2}\cdot {\bf
1963: y})\right], \label{c6}
1964: \end{equation}
1965: where ${\bf c}_{i}\equiv {\bf V}_{i}/v_{0}$ and ${\bf y}\equiv {\bf c}_1-{\bf
1966: c}_2$. The integral can be performed by the change of variables
1967: $\{{\bf c}_1,{\bf c}_2\}\to \{{\bf y},{\bf z}\}$, where ${\bf
1968: z}\equiv \theta _{1} {\bf c}_{1}+\theta _{2}{\bf c}_{2}$ and the
1969: Jacobian is $\left( \theta _{1}+\theta _{2}\right) ^{-d}$. With
1970: this change, Eq.\ (\ref{c6}) becomes
1971: \begin{equation}
1972: \label{c6.1} \nu=\frac{2}{d}\frac{\pi^{(d-1)/2}}
1973: {\Gamma\left(\frac{d+3}{2}\right)} \nu_0 (1+\alpha _{12}) \pi^{-d}
1974: \left(\theta_1\theta_2\right)^{(d+1)/2}\left( \theta _{1}+\theta
1975: _{2}\right) ^{-(1+d)}(x_2\mu_{21}+x_1\mu_{12})\int \,d{\bf
1976: y}\,\int \,d{\bf z}\,y^3 e^{-(a y^{2}+b z^{2})},
1977: \end{equation}
1978: where $a\equiv \theta_1\theta_2\left( \theta _{1}+\theta
1979: _{2}\right) ^{-1}$ and $b\equiv \left( \theta _{1}+\theta
1980: _{2}\right) ^{-1}$. The integral (\ref{c6.1}) can be easily
1981: computed and one directly gets the result (\ref{3.16}) given in
1982: the text for the reduced collision frequency $\nu^*=\nu/\nu_0$.
1983: 
1984: 
1985: The collision frequencies $\tau_{ij}$ defined by Eqs.\
1986: (\ref{4.17}) and (\ref{4.18}) involve collision integrals of the
1987: form
1988: \begin{equation}
1989: \label{c8} \int d{\bf v}_1 {\bf V}_{1}{\bf V}_{1}
1990:  J_{ij}[f_i,f_j]=\sigma_{ij}^{d-1}\int\, d{\bf
1991: v}_1\int\, d{\bf v}_2\, f_i({\bf V}_1) f_j({\bf V}_2)  \int
1992: d\widehat{\boldsymbol {\sigma}}\,\Theta (\widehat{\boldsymbol
1993: {\sigma}} \cdot {\bf g}_{12})(\widehat{\boldsymbol {\sigma}}\cdot
1994: {\bf g}_{12})\,\left[ {\bf V}_1''{\bf V}_{1}''-{\bf V}_{1}{\bf
1995: V}_{1}\right],
1996: \end{equation}
1997: where the identity (\ref{c1}) has been used. The scattering rule
1998: (\ref{c2}) gives
1999: \begin{equation}
2000: \label{c9} {\bf V}_{1}''{\bf V}_{1}''-{\bf V}_{1}{\bf
2001: V}_{1}=-\mu_{ji}(1+ \alpha_{ij})(\widehat{\boldsymbol
2002: {\sigma}}\cdot {\bf g}_{12})\left[ {\bf
2003: G}_{ij}\widehat{\boldsymbol
2004: {\sigma}}+\widehat{\boldsymbol{\sigma}}{\bf G}_{ij} +\mu_{ji}({\bf
2005: g}_{12}\widehat{\boldsymbol{\sigma}}+\widehat{\boldsymbol{\sigma}}{\bf
2006: g}_{12}) -\mu_{ji}(1+\alpha_{ij})(\widehat{\boldsymbol
2007: {\sigma}}\cdot {\bf g}_{12})
2008: \widehat{\boldsymbol{\sigma}}\widehat{\boldsymbol{\sigma}}\right],
2009: \end{equation}
2010: where ${\bf G}_{ij}=\mu_{ij}{\bf V}_1+\mu_{ji}{\bf V}_2$.
2011: Substitution of Eq.\ (\ref{c9}) into Eq.\ (\ref{c8}) allows the
2012: angular integral to be performed with the result
2013: \begin{eqnarray}
2014: \label{c10} \int d\widehat{\boldsymbol {\sigma}}\,\Theta
2015: (\widehat{\boldsymbol {\sigma}}\cdot {\bf g
2016: }_{12})(\widehat{\boldsymbol {\sigma}}\cdot {\bf g}_{12})&&\left[
2017: {\bf V}_{1}''{\bf V}_{1}''-{\bf V}_{1}{\bf V}_{1}\right]= -B_3 m_i
2018: \mu_{ji}(1+\alpha _{ij})\left[g_{12}({\bf G}_{ij}{\bf g}_{12}+
2019: {\bf g}_{12}{\bf G}_{ij})\right. \nonumber\\
2020: & &\left. +3\frac{\mu_{ji}}{d+3}
2021: (1+\frac{2d}{3}-\alpha_{ij})g_{12}{\bf g}_{12}{\bf g}_{12}
2022: -\frac{\mu_{ji}}{d+3}(1+\alpha_{ij})g^3\openone\right].
2023: \end{eqnarray}
2024: Using (\ref{c10}) the integrals defining $\tau_{ij}$ can be calculated by the same
2025: mathematical steps as those made before for $\nu$. After a lengthy calculation, one gets
2026: \begin{eqnarray}
2027: \label{c11} \int d{\bf v}_1 R_{1,k\ell}J_{12}[f_1^{(0)},
2028: f_{2,M}R_{2,k\ell}] &=&
2029: -\frac{\pi^{(d-1)/2}}{2d\Gamma\left(\frac{d}{2}\right)}m_1
2030: m_2n_1n_2\mu_{21}(1+\alpha_{12}) \sigma_{12}^{d-1}
2031: v_0^5 \left(\theta_1\theta_2\right)^{-1/2}\nonumber\\
2032: & \times &
2033: \left\{2(d+3)(d-1)(\mu_{12}\theta_2-\mu_{21}\theta_1)\theta_2^{-2}
2034: \left(\theta_1+\theta_2\right)^{-1/2}\right.\nonumber\\
2035: & & +3(d-1)\mu_{21}\left(1+\frac{2d}{3}-\alpha_{12}\right)
2036: \theta_2^{-2}\left(\theta_1+\theta_2\right)^{1/2}\nonumber\\
2037: & & \left.
2038: -\left[2d(d+1)-4\right]\theta_2^{-1}\left(\theta_1+\theta_2\right)^{-1/2}\right\},
2039: \end{eqnarray}
2040: \begin{eqnarray}
2041: \label{c12} \int d{\bf v}_1 R_{1,k\ell}J_{12}[f_{1,M}R_{1,k\ell},
2042: f_2^{(0)}] &=&
2043: -\frac{\pi^{(d-1)/2}}{2d\Gamma\left(\frac{d}{2}\right)}m_1
2044: m_2n_1n_2\mu_{21}(1+\alpha_{12}) \sigma_{12}^{d-1}
2045: v_0^5 \left(\theta_1\theta_2\right)^{-1/2}\nonumber\\
2046: & \times &
2047: \left\{2(d+3)(d-1)(\mu_{12}\theta_2-\mu_{21}\theta_1)\theta_1^{-2}
2048: \left(\theta_1+\theta_2\right)^{-1/2}\right.\nonumber\\
2049: & & +3(d-1)\mu_{21}\left(1+\frac{2d}{3}-\alpha_{12}\right)
2050: \theta_1^{-2}\left(\theta_1+\theta_2\right)^{1/2}\nonumber\\
2051: & & \left.
2052: +\left[2d(d+1)-4\right]\theta_1^{-1}\left(\theta_1+\theta_2\right)^{-1/2}\right\},
2053: \end{eqnarray}
2054: \begin{eqnarray}
2055: \label{c13} \int d{\bf v}_1
2056: R_{1,k\ell}\left\{J_{11}[f_1^{(0)},f_{1,M}R_{1,k\ell}]
2057: +J_{11}[f_{1,M}R_{1,k\ell},f_1^{(0)}]\right\}
2058: &=&-\frac{\pi^{(d-1)/2}}{\Gamma\left(\frac{d}{2}\right)}m_1^2
2059: n_1^2(1+\alpha_{11})\sigma_1^{d-1} (T_1/m_1)^{5/2}\nonumber\\
2060: & & \times\frac{6(d-1)}{d}
2061: \left(1+\frac{2d}{3}-\alpha_{11}\right).
2062: \end{eqnarray}
2063: The corresponding expressions for $\tau_{ij}$ can be easily
2064: inferred from Eqs. \ (\ref{c11})--(\ref{c13}).
2065: 
2066: 
2067: The collision frequencies $\nu_{ij}$ and $\omega_{ij}$ that
2068: determine the heat flux are defined by Eqs.\ (\ref{4.17}),
2069: (\ref{4.18}), and (\ref{4.19}), respectively. To evaluate these
2070: collision integrals, one needs the partial results
2071: \begin{eqnarray}
2072: \label{c15} {\bf S}_{i}({\bf V}_{1}^{^{\prime \prime}})-{\bf
2073: S}_{i}({\bf V}_{1}) &=& \frac{m_{i}}{2}(1+\alpha
2074: _{ij})\mu_{ji}(\widehat{\boldsymbol {\sigma}}\cdot {\bf g
2075: }_{12})\left\{ \left[
2076: (1-\alpha_{ij}^{2})\mu_{ji}^{2}(\widehat{\boldsymbol
2077: {\sigma}}\cdot {\bf
2078: g}_{12})^{2}-G_{ij}^{2}-\mu_{ji}^{2}g_{12}^{2}\right.
2079: \right.   \nonumber \\
2080: &&\left. -2\mu_{ji}({\bf g}_{12}\cdot {\bf
2081: G}_{ij})+2(1+\alpha_{ij})\mu _{ji}(\widehat{\boldsymbol
2082: {\sigma}}\cdot {\bf g}_{12})(\widehat{\boldsymbol {\sigma }} \cdot
2083: {\bf G}_{ij})+(d+2)\frac{T_{i}}{m_{i}}\right] \widehat{\boldsymbol
2084: {\sigma}}
2085: \nonumber \\
2086: &&-\left[ (1-\alpha_{ij})\mu_{ji}(\widehat{\boldsymbol {\sigma
2087: }}\cdot {\bf g}_{12} )+2(\widehat{\boldsymbol {\sigma}}\cdot {\bf
2088: G}_{ij})\right] {\bf G}_{ij}
2089: \nonumber \\
2090: &&\left.
2091: -\mu_{ji}\left[(1-\alpha_{ij})\mu_{ji}(\widehat{\boldsymbol
2092: {\sigma}} \cdot {\bf g}_{12})+2(\widehat{\boldsymbol{\sigma}}\cdot
2093: {\bf G}_{ij}) \right] {\bf g}_{12}\right\},
2094: \end{eqnarray}
2095: \begin{eqnarray}
2096: \label{c16} \int d\widehat{\boldsymbol {\sigma}}\,\Theta
2097: (\widehat{\boldsymbol {\sigma}}\cdot {\bf g
2098: }_{12})(\widehat{\boldsymbol {\sigma}}\cdot {\bf g}_{12}) &&\left[
2099: {\bf S}_{i}( {\bf V}_{1}^{^{\prime \prime }})-{\bf S}_{i}({\bf
2100: V}_{1})\right] =-\frac{
2101: m_{i}}{2}\frac{\pi^{(d-1)/2}}{\Gamma\left(\frac{d+3}{2}\right)}
2102: (1+\alpha_{ij})\mu_{ji}\nonumber\\
2103: & &\times \left\{ \left[
2104: g_{12}G_{ij}^{2}+\mu_{ji}^{2}\frac{4\alpha_{ij}^{2}-(d+3)\alpha
2105: _{ij}+2(d+1)}{d+3} g_{12}^{3}\right.\right.\nonumber\\
2106: & & \left. -2\mu_{ji}\frac{3\alpha_{ij}-2d-3}{d+3} g_{12}\left(
2107: {\bf g}_{12}\cdot {\bf G}_{ij}\right) -(d+2)\frac{T_{i}}{m_{i}}
2108: g_{12}\right] {\bf g}_{12}  \nonumber \\
2109: &&\left. +\left[ 2 g_{12}\left( {\bf g}_{12}\cdot {\bf
2110: G}_{ij}\right) -\mu_{ji}\frac{(d+5)\alpha_{ij}-d-1}{d+3}
2111: g_{12}^{3}\right] {\bf G}_{ij}\right\} \;.
2112: \end{eqnarray}
2113: The integrals $\omega_{ij}$ and $\nu_{ij}$ can be explicitly
2114: evaluated by using (\ref{c16}) and the same mathematical steps as
2115: before. After a lengthy algebra, one gets
2116: \begin{eqnarray}
2117: \label{c17} \omega_{12}&=&\frac{\pi^{(d-1)/2}}
2118: {\Gamma\left(\frac{d}{2}\right)}\frac{2}{d\sqrt{2}}
2119: \left(\frac{\sigma_1}{\sigma_{12}}\right)^{d-1}x_1
2120: \theta_1^{-1/2}(1-\alpha_{11}^2)\nonumber\\
2121: & & +\frac{\pi^{(d-1)/2}}
2122: {\Gamma\left(\frac{d}{2}\right)}\frac{2}{d(d+2)}
2123: x_1\mu_{21}(1+\alpha_{12})(\theta_1+\theta_2)^{-1/2}\theta_1^{1/2}
2124: \theta_2^{-3/2}\left(\frac{x_2}{x_1}A-\gamma B\right),
2125: \end{eqnarray}
2126: \begin{eqnarray}
2127: \label{c18}
2128: \nu_{11}&=&\frac{\pi^{(d-1)/2}}{\Gamma\left(\frac{d}{2}\right)}
2129: \frac{8}{d(d+2)}\left(\frac{\sigma_1}{\sigma_{12}}\right)^{d-1}
2130: x_1 (2\theta_1)^{-1/2}
2131: (1+\alpha_{11})\left[\frac{d-1}{2}+\frac{3}{16}(d+8)(1-\alpha_{11})\right]\nonumber\\
2132: & & +\frac{\pi^{(d-1)/2}}{\Gamma\left(\frac{d}{2}\right)}
2133: \frac{1}{d(d+2)}x_2\mu_{21}(1+\alpha_{12})\left(\frac{\theta_1}
2134: {\theta_2(\theta_1+\theta_2)}\right)^{3/2}\left[E-(d+2)\frac{\theta_1+\theta_2}{\theta_1}A\right],
2135: \end{eqnarray}
2136: \begin{equation}
2137: \label{c19}
2138: \nu_{12}=-\frac{\pi^{(d-1)/2}}{\Gamma\left(\frac{d}{2}\right)}
2139: \frac{1}{d(d+2)}x_2\frac{\mu_{21}^2}{\mu_{12}}(1+\alpha_{12})\left(\frac{\theta_1}
2140: {\theta_2(\theta_1+\theta_2)}\right)^{3/2}\left[F+(d+2)\frac{\theta_1+\theta_2}{\theta_2}B\right].
2141: \end{equation}
2142: In the above equations we have introduced the quantities
2143: \cite{note}
2144: \begin{eqnarray}
2145: \label{c20}
2146: A&=&(d+2)(2\beta_{12}+\theta_2)+\mu_{21}(\theta_1+\theta_2)\left\{(d+2)(1-\alpha_{12})
2147: -[(11+d)\alpha_{12}-5d-7]\beta_{12}\theta_1^{-1}\right\}\nonumber\\
2148: & &
2149: +3(d+3)\beta_{12}^2\theta_1^{-1}+2\mu_{21}^2\left(2\alpha_{12}^{2}-\frac{d+3}{2}\alpha
2150: _{12}+d+1\right)\theta_1^{-1}(\theta_1+\theta_2)^2-
2151: (d+2)\theta_2\theta_1^{-1}(\theta_1+\theta_2), \nonumber\\
2152: \end{eqnarray}
2153: \begin{eqnarray}
2154: \label{c21} B&=&
2155: (d+2)(2\beta_{12}-\theta_1)+\mu_{21}(\theta_1+\theta_2)\left\{(d+2)(1-\alpha_{12})
2156: +[(11+d)\alpha_{12}-5d-7]\beta_{12}\theta_2^{-1}\right\}\nonumber\\
2157: & &
2158: -3(d+3)\beta_{12}^2\theta_2^{-1}-2\mu_{21}^2\left(2\alpha_{12}^{2}-\frac{d+3}{2}\alpha
2159: _{12}+d+1\right)\theta_2^{-1}(\theta_1+\theta_2)^2+
2160: (d+2)(\theta_1+\theta_2), \nonumber\\
2161: \end{eqnarray}
2162: \begin{eqnarray}
2163: \label{c22} E&=&
2164:  2\mu_{21}^2\theta_1^{-2}(\theta_1+\theta_2)^2
2165: \left(2\alpha_{12}^{2}-\frac{d+3}{2}\alpha_{12}+d+1\right)
2166: \left[(d+2)\theta_1+(d+5)\theta_2\right]\nonumber\\
2167: & & -\mu_{21}(\theta_1+\theta_2)
2168: \left\{\beta_{12}\theta_1^{-2}[(d+2)\theta_1+(d+5)\theta_2][(11+d)\alpha_{12}
2169: -5d-7]\right.\nonumber\\
2170: & & \left.
2171: -\theta_2\theta_1^{-1}[20+d(15-7\alpha_{12})+d^2(1-\alpha_{12})-28\alpha_{12}]
2172: -(d+2)^2(1-\alpha_{12})\right\}
2173: \nonumber\\
2174: & & +3(d+3)\beta_{12}^2\theta_1^{-2}[(d+2)\theta_1+(d+5)\theta_2]+
2175: 2\beta_{12}\theta_1^{-1}[(d+2)^2\theta_1+(24+11d+d^2)\theta_2]
2176: \nonumber\\
2177: & & +(d+2)\theta_2\theta_1^{-1}
2178: [(d+8)\theta_1+(d+3)\theta_2]-(d+2)(\theta_1+\theta_2)\theta_1^{-2}\theta_2
2179: [(d+2)\theta_1+(d+3)\theta_2],\nonumber\\
2180: \end{eqnarray}
2181: \begin{eqnarray}
2182: \label{c23} F&=&
2183:  2\mu_{21}^2\theta_2^{-2}(\theta_1+\theta_2)^2
2184: \left(2\alpha_{12}^{2}-\frac{d+3}{2}\alpha_{12}+d+1\right)
2185: \left[(d+5)\theta_1+(d+2)\theta_2\right]\nonumber\\
2186: & & -\mu_{21}(\theta_1+\theta_2)
2187: \left\{\beta_{12}\theta_2^{-2}[(d+5)\theta_1+(d+2)\theta_2][(11+d)\alpha_{12}
2188: -5d-7]\right.\nonumber\\
2189: & & \left.
2190: +\theta_1\theta_2^{-1}[20+d(15-7\alpha_{12})+d^2(1-\alpha_{12})-28\alpha_{12}]
2191: +(d+2)^2(1-\alpha_{12})\right\}
2192: \nonumber\\
2193: & & +3(d+3)\beta_{12}^2\theta_2^{-2}[(d+5)\theta_1+(d+2)\theta_2]-
2194: 2\beta_{12}\theta_2^{-1}[(24+11d+d^2)\theta_1+(d+2)^2\theta_2]
2195: \nonumber\\
2196: & & +(d+2)\theta_1\theta_2^{-1}
2197: [(d+3)\theta_1+(d+8)\theta_2]-(d+2)(\theta_1+\theta_2)\theta_2^{-1}
2198: [(d+3)\theta_1+(d+2)\theta_2]. \nonumber\\
2199: \end{eqnarray}
2200: Here, $\beta_{12}=\mu_{12}\theta_2-\mu_{21}\theta_1$. From Eqs.\
2201: (\ref{c17})--(\ref{c23}), one easily gets the expressions for
2202: $\omega_{21}$, $\nu_{22}$ and $\nu_{21}$ by interchanging
2203: $1\leftrightarrow 2$.
2204: 
2205: 
2206: In the case of a three-dimensional system ($d=3$), all the above
2207: results reduce to those previously obtained for hard spheres when
2208: one takes Maxwellian distributions for the reference homogeneous
2209: cooling state. \cite{GD02,MGD06}
2210: 
2211: 
2212: 
2213: \begin{thebibliography} {99}
2214: 
2215: \bibitem{GS95}A. Goldshtein and M. Shapiro, Mechanics of collisional motion of granular materials.
2216: Part 1. General hydrodynamic equations, {\em J. Fluid Mech.} {\bf
2217: 282}:75--114 (1995).
2218: 
2219: \bibitem{BDS97}J. J. Brey, J. W. Dufty, and A. Santos, Dissipative
2220: dynamics for hard spheres, {\em J. Stat. Phys.} {\bf
2221: 87}:1051--1066 (1997).
2222: 
2223: \bibitem{BP04}N. V. Brilliantov and T. P\"oschel, {\em Kinetic Theory
2224: of Granular Gases} (Oxford University Press, Oxford, 2004).
2225: 
2226: \bibitem{CC70}S. Chapman and T. G. Cowling, {\em The Mathematical Theory of Nonuniform Gases}
2227: (Cambridge University Press, Cambridge, 1970).
2228: 
2229: \bibitem{simple}J. J. Brey, J. W. Dufty, C. S. Kim, and A. Santos,
2230: Hydrodynamics for granular flow at low-density, {\em Phys. Rev. E}
2231: {\bf 58}:4638--4653 (1998); J. J. Brey and D. Cubero, Hydrodynamic
2232: transport coefficients of granular gases, in {\em Granular Gases},
2233: edited by T. P\"oschel and S. Luding, Lecture Notes in Physics
2234: (Springer Verlag, Berlin, 2001), pp. 59--78; V. Garz\'o and J. M.
2235: Montanero, Transport coefficients of a heated granular gas, {\em
2236: Physica A} {\bf 313}:336--356 (2002).
2237: 
2238: \bibitem{mixture}J. T. Jenkins and F. Mancini, Kinetic theory for binary
2239: mixtures of smooth, nearly elastic spheres, {\em Phys. Fluids A}
2240: {\bf 1}:2050--2057 (1989); P. Zamankhan, Kinetic theory for
2241: multicomponent dense mixtures of slightly inelastic spherical
2242: particles, {\em Phys. Rev. E} {\bf 52}:4877--4891 (1995); B.
2243: Arnarson and J. T. Willits, Thermal diffusion in binary mixtures
2244: of smooth, nearly elastic spheres with and without gravity, {\em
2245: Phys. Fluids} {\bf 10}:1324--1328 (1998); J. T. Willits and B.
2246: Arnarson, Kinetic theory of a binary mixture of nearly elastic
2247: disks, {\em Phys. Fluids} {\bf 11}:3116--3122 (1999).
2248: 
2249: 
2250: 
2251: \bibitem{GD99b}  V. Garz\'{o} and J. W. Dufty, Homogeneous cooling
2252: state for a granular mixture, {\em Phys. Rev. E} {\bf
2253: 60}:5706--5713 (1999).
2254: 
2255: \bibitem{MP99}P. A. Martin and J. Piasecki, Thermalization of a particle by
2256: dissipative collisions, {\em Europhys. Lett.} {\bf 46}:613--616
2257: (1999).
2258: 
2259: \bibitem{computer}See for instance, J. M. Montanero and V. Garz\'o,
2260: Monte Carlo simulation of the homogeneous cooling state for a granular mixture, {\em Gran. Matt.} {\bf 4}:17--24
2261: (2002); A. Barrat and E. Trizac, Lack of energy equipartition in homogeneous heated binary granular mixtures,
2262: {\em ibid.} {\bf 4}:57--63 (2002); S. R. Dahl, C. M. Hrenya, V. Garz\'o, and J. W. Dufty, Kinetic temperatures
2263: for a granular mixture, {\em Phys. Rev. E} {\bf 66}:041301 (2002); R. Pagnani, U. M. B. Marconi, and A. Puglisi,
2264: Driven low density granular mixtures, {\em ibid.} {\bf 66}:051304 (2002); D. Paolotti, C. Cattuto, U. M. B.
2265: Marconi, and A. Puglisi, Dynamical properties of vibrofluidized granular mixtures, {\em Gran. Matt.} {\bf
2266: 5}:75--83 (2003); P. Krouskop and J. Talbot, Mass and size effects in three-dimensional vibrofluidized granular
2267: mixtures, {\em Phys. Rev. E} {\bf 68}: 021304 (2003); H. Wang, G. Jin, and Y. Ma, Simulation study on kinetic
2268: temperatures of vibrated binary granular mixtures, {\em ibid.} {\bf 68}:031301 (2003); J. J. Brey, M. J.
2269: Ruiz-Montero, and F. Moreno, Energy partition and segregation for an intruder in a vibrated granular system
2270: under gravity, {\em Phys. Rev. Lett.} {\bf 95}:098001 (2005); M. Schr\"oter, S. Ulrich , J. Kreft , J. B. Swift,
2271: and H. L. Swinney, Mechanisms in the size segregation of a binary granular mixture, {\em Phys. Rev. E} {\bf
2272: 74}:011307 (2006).
2273: 
2274: 
2275: \bibitem{exp1}R. D. Wildman and D. J. Parker, Coexistence of two granular
2276: temperatures in binary vibrofluidized beds, {\em Phys. Rev. Lett.} {\bf 88}:064301 (2002); K. Feitosa and N.
2277: Menon, Breakdown of energy equipartition in a 2D binary vibrated granular gas, {\em ibid.} {\bf 88}:198301
2278: (2002).
2279: 
2280: \bibitem{JM87}J. Jenkins and F. Mancini, Balance laws and constitutive relations
2281: for plane flows of a dense binary mixture of smooth, nearly elastic, circular disks, {\em J. Appl. Mech.} {\bf
2282: 54}:27--34 (1987).
2283: 
2284: 
2285: \bibitem{GD02}V. Garz\'o and J. W. Dufty, Hydrodynamics for a granular binary mixture
2286: at low density, {\em Phys. Fluids} {\bf 14}:1476--1490 (2002).
2287: 
2288: \bibitem{GA05}V. Garz\'o  and A. Astillero, Transport coefficients
2289: for inelastic Maxwell mixtures, {\em J. Stat. Phys.} {\bf
2290: 118}:935--971 (2005).
2291: 
2292: 
2293: 
2294: \bibitem{MGD06}V. Garz\'o, J. M. Montanero, and J. W. Dufty, Mass
2295: and heat fluxes for a binary granular mixture at low-density, {\em
2296: Phys. Fluids} {\bf 18}:083305 (2006).
2297: 
2298: \bibitem{GM04}V. Garz\'o and J. M. Montanero, Diffusion of impurities in a granular gas,
2299: {\em Phys. Rev. E} {\bf 69}: 021301 (2004).
2300: 
2301: \bibitem{MG03}J. M. Montanero and V. Garz\'o, Shear viscosity for a heated granular binary mixture
2302: at low-density, {\em Phys. Rev. E} {\bf 67}:021308 (2003).
2303: 
2304: \bibitem{SGNT06}D. Serero, I. Goldhirsch, S. H. Noskowicz, and M.-L. Tan,
2305: Hydrodynamics of granular gases and granular gas mixtures, {\em J. Fluid Mech.} {\bf 554}:237--258 (2006).
2306: 
2307: \bibitem{exp}See for instance,
2308: S. Warr, J. M. Huntley, and G. T. H. Jacques, Fluidization of a
2309: two-dimensional granular system: Experimental study and scaling
2310: behavior, {\em Phys. Rev. E} {\bf 52}:5583--5595 (1995); J. S.
2311: Olafsen and J. S. Urbach, Velocity distribution and density
2312: fluctuations in a granular gas, {\em ibid.} {\bf 60}:R2468--R2471
2313: (1999); R. D. Wildman, J. M. Huntley, and J. P. Hansen,
2314: Self-diffusion of grains in two-dimensional vibrofluidized bed,
2315: {\em ibid.} {\bf 60}:7066--7075 (1999); F. Rouyer and N. Menon,
2316: Velocity fluctuations in a homogeneous 2D granular gas in steady
2317: state, {\em Phys. Rev. Lett.} {\bf 85}:3676--3679 (2000); D. L.
2318: Blair and A. Kudrolli, Velocity correlations in dense granular
2319: flows, {\em Phys. Rev. E} {\bf 64}:050301 (2001); S. Horluck, M.
2320: van Hecke, and P. Dimon, Shock waves in two-dimensional granular
2321: flow: Effects of rough walls and polydispersity, {\em ibid.} {\bf
2322: 67}:021304 (2003).
2323: 
2324: \bibitem{Bird}G. A. Bird, {\em Molecular Gas Dynamics and the Direct
2325: Simulation Monte Carlo of Gas Flows} (Clarendon, Oxford, 1994).
2326: 
2327: \bibitem{SGD04}A. Santos, V. Garz\'o, and J. W. Dufty, Inherent
2328: rheology of a granular fluid in uniform shear flow, {\em Phys.
2329: Rev. E} {\bf 69}:061303 (2004).
2330: 
2331: \bibitem{NE98}T. P. C. van Noije and M. H. Ernst, Velocity
2332: distributions in inhomogeneous granular fluids: the free and the heated case,
2333: {\em Gran. Matt.} {\bf 1}:57--64(1998).
2334: 
2335: \bibitem{G02}V. Garz\'o, Tracer diffusion in granular shear flows, {\em Phys. Rev.
2336: E} {\bf 66}:021308 (2002).
2337: 
2338: 
2339: \bibitem{BRM05}J. J. Brey, M. J. Ruiz-Montero, and F. Moreno, Energy partition
2340: and segregation for an intruder in a vibrated granular system
2341: under gravity, {\em Phys. Rev. Lett.} {\bf 95}:098001 (2005);
2342: Hydrodynamic profiles for an impurity in an open vibrated granular
2343: gas, {\em Phys. Rev. E} {\bf 73}:031301 (2006).
2344: 
2345: \bibitem{BC01}J. J. Brey and D. Cubero, Hydrodynamic
2346: transport coefficients of granular gases, in {\em Granular Gases},
2347: edited by T. P\"oschel and S. Luding, Lecture Notes in Physics
2348: (Springer Verlag, Berlin, 2001), pp. 59--78.
2349: 
2350: 
2351: \bibitem{M89}J. A. McLennan, {\em Introduction to Nonequilibrium Statistical Mechanics}
2352: (Prentice-Hall, New Jersey, 1989).
2353: 
2354: 
2355: \bibitem{BRCG00} J. J. Brey, M. J. Ruiz-Montero, D. Cubero, and R. Garc\'{\i}a-Rojo, Self-diffusion in
2356: freely evolving granular gases, {\em Phys. Fluids} {\bf
2357: 12}:876--883 (2000).
2358: 
2359: 
2360: \bibitem{GM03}V. Garz\'o and J. M. Montanero, Shear viscosity for a moderately dense granular
2361: binary mixture, {\em Phys. Rev. E} {\bf 68}:041302 (2003).
2362: 
2363: 
2364: \bibitem{MSG05}J. M. Montanero, A. Santos, and V. Garz\'o, DSMC evaluation of the Navier-Stokes
2365: shear viscosity of a granular fluid, in {\em Rarefied Gas Dynamics
2366: 24}, edited by M. Capitelli (American Institute of Physics, Vol.
2367: 762, 2005), pp. 797--802; preprint arXiv: cond-mat/0411219.
2368: 
2369: \bibitem{BRM04}J. J. Brey and M. J. Ruiz-Montero, Simulation
2370: study of the Green-Kubo relations for dilute granular gases, {\em
2371: Phys. Rev. E} {\bf 70}:051301 (2004).
2372: 
2373: \bibitem{BRMMG05}J. J. Brey, M. J. Ruiz-Montero, P. Maynar, and M.
2374: I. Garc\'{\i}a de Soria, Hydrodynamic modes, Green-Kubo relations,
2375: and velocity correlations in dilute granular gases, {\em J. Phys.:
2376: Condens. Matter} {\bf 17}:S2489--S2502 (2005).
2377: 
2378: \bibitem{MSG06}J. M. Montanero,  A. Santos, and V. Garz\'o, First-order
2379: Chapman-Enskog velocity distribution function in a granular gas,
2380: {\em Physica A} {\bf 376}:75--93 (2007).
2381: 
2382: \bibitem{GSM06}V. Garz\'o, A. Santos, and J. M. Montanero,
2383: Modified Sonine approximation for the Navier-Stokes transport
2384: coefficients of a granular gas, {\em Physica A} {\bf 376}:94--107 (2007).
2385: 
2386: \bibitem{MCK83}M. L\'opez de Haro, E. G. D. Cohen, and J. M.
2387: Kincaid, The Enskog theory for multicomponent mixtures. I. Linear
2388: transport theory, {\em J. Chem. Phys.} {\bf 78}:2746--2759 (1983).
2389: 
2390: \bibitem{GD99a}V. Garz\'{o} and J. W. Dufty, Dense fluid transport for inelastic hard spheres,
2391: {\em Phys. Rev. E} {\bf 59}:5895--5911 (1999).
2392: 
2393: 
2394: \bibitem{EB02}M. H. Ernst and R. Brito, Scaling solutions of inelastic Boltzmann
2395: equations with over-populated high energy tails, {\em J. Stat.
2396: Phys.} {\bf 109}:407--432 (2002).
2397: 
2398: \bibitem{note}Some misprints occur in the expressions given in
2399: Ref.\ \onlinecite{GD02} for the heat flux. The results displayed
2400: here correct and extend such results to $d$ dimensions.
2401: 
2402: 
2403: \end{thebibliography}
2404: 
2405: 
2406: \end{document}
2407: