1: \documentclass{elsart}
2: \usepackage{ifpdf}
3: \usepackage{graphicx,natbib,amssymb,lineno,subfigure}
4: \ifpdf
5: \usepackage[%
6: pdftitle={},%
7: pdfauthor={},%
8: pdfsubject={},%
9: pdfkeywords={},%
10: pdfstartview=FitH,%
11: bookmarks=true,%
12: bookmarksopen=true,%
13: breaklinks=true,%
14: colorlinks=true,%
15: linkcolor=blue,anchorcolor=blue,%
16: citecolor=blue,filecolor=blue,%
17: menucolor=blue,pagecolor=blue,%
18: urlcolor=blue]{hyperref}
19: \else
20: \usepackage[%
21: breaklinks=true,%
22: colorlinks=true,%
23: linkcolor=blue,anchorcolor=blue,%
24: citecolor=blue,filecolor=blue,%
25: menucolor=blue,pagecolor=blue,%
26: urlcolor=blue]{hyperref}
27: \fi
28:
29: %\renewcommand\floatpagefraction{.2}
30: \makeatletter
31: \def\elsartstyle{%
32: \def\normalsize{\@setfontsize\normalsize\@xiipt{14.5}}
33: \def\small{\@setfontsize\small\@xipt{13.6}}
34: \let\footnotesize=\small
35: \def\large{\@setfontsize\large\@xivpt{18}}
36: \def\Large{\@setfontsize\Large\@xviipt{22}}
37: \skip\@mpfootins = 18\p@ \@plus 2\p@
38: \normalsize
39: }
40: \@ifundefined{square}{}{\let\Box\square}
41: \makeatother
42:
43: \def\file#1{\texttt{#1}}
44:
45: \renewcommand{\vec}[1]{{\bf #1}}
46: \newcommand{\pd}[2]{\frac{\partial #1}{\partial #2}}
47: \newcommand{\pdsq} [2]{\frac{\partial^2 #1}{\partial #2^2}}
48: \newcommand{\bs} [1]{\mbox{\boldmath $#1$}}
49: \newcommand{\tr}{\mathop{\mathrm{tr}}}
50: \newcommand{\unitvec}[1]{{\hat{\bm{{#1}}}}}
51:
52: \pagestyle{plain}
53: \begin{document}
54:
55: \begin{frontmatter}
56: \title{The classical granular temperature\\ and slightly beyond}
57:
58: \author[TAU]{D. Serero},
59: \ead{serero@eng.tau.ac.il}
60: \author[PMMH]{C. Goldenberg},
61: \ead{chayg@pmmh.espci.fr}
62: \thanks{I.~G. gratefully acknowledges support from the Israel Science Foundation
63: (ISF), grant. no. 689/04, the German Israel Science Foundation
64: (GIF), grant no. 795/2003 and the US-Israel Binational Science Foundation
65: (BSF), grant no. 2004391. CG gratefully acknowledge support from a European Community FP6 Marie Curie Action
66: (MEIF-CT2006-024970).}
67: \author[TAU]{S. H. Noskowicz},
68: \ead{henri@eng.tau.ac.il}
69: \author[TAU]{I. Goldhirsch}
70: \ead{isaac@eng.tau.ac.il}
71:
72: \address[TAU]{Department of Fluid Mechanics and Heat Transfer,
73: Faculty of Engineering,\\ Tel Aviv University,
74: Ramat-Aviv, Tel Aviv 69978, Israel.}
75: \address[PMMH]{Laboratoire de Physique et M\'ecanique des Milieux
76: H\'et\'erog\`enes\\ (CNRS UMR 7636), ESPCI, 10 rue Vauquelin, 75231 Paris Cedex
77: 05, France.}
78:
79: \begin{abstract}
80: One goal of this paper is to discuss the classical definition of granular
81: temperature as an extension of its thermodynamic equivalent and a
82: useful
83: concept which provides an important characterization of fluidized
84: granular matter. Following a review of some basic
85: concepts and techniques (with emphasis on fundamental issues) we present
86: new results for a system that can exhibit strong violations of equipartition,
87: yet is amenable to description by classical granular hydrodynamics,
88: namely a binary granular gas mixture. A second goal of this article is
89: to present a result that pertains to dense granular and molecular solids alike,
90: namely the existence of
91: a correction to the elastic
92: energy which is related to the heat flux in the equations of continuum
93: mechanics. The latter is of the same (second) order in the strain as the
94: elastic energy. Although recent definitions of temperatures for
95: granular matter, glasses and other disordered many-body systems are not
96: within the scope of this article we do make several general comments on this
97: subject in the closing section.
98: \end{abstract}
99: \begin{keyword}
100: Temperature, granular matter, elasticity, kinetic theory
101: \PACS 05.20.Dd, 47.45.Ab, 45.70.Mg ,83.80.Fg, 62.20.Dc
102: \end{keyword}
103: \end{frontmatter}
104: \section{Introduction}
105: \label{intro1}
106:
107: The axiomatic formulation of thermodynamics provides a precise definition of
108: temperature. Nearly any textbook on thermodynamics shows this definition
109: to be compatible with common experience.
110: Equilibrium statistical mechanics endows temperature with a microscopic
111: meaning but invokes a few assumptions (e.g., ergodicity).
112: The physical relevance and usefulness
113: of both of these theories is beyond debate.
114:
115:
116: As there are practically no equilibrium systems in nature there is clearly a
117: need to describe non-equilibrium systems. The most successful theories that
118: treat such systems are those devoted to near-equilibrium states, namely states
119: in which the macroscopic fields vary sufficiently slowly in space and time so
120: that (near) local equilibrium distributions can establish themselves
121: everywhere while the state of the system hardly changes. That this should
122: occur is not entirely obvious. It helps to note that e.g., in gases local
123: equilibration occurs within a few collisions per particle, i.e. on time scales
124: that are much smaller than macroscopic times. In other words, these
125: approaches are effective when there is significant scale separation between
126: the spatial and temporal micro-and-macroscales. In the latter case one can
127: justify hydrodynamics, elasticity and other macroscopic theories, and the
128: transport coefficients are given by Green-Kubo expressions derived using
129: linear (and non-linear) response theories \cite{IO} or projection operator
130: approaches \cite{Mori1,Mori2}, a similar statement holding for fluctuation
131: dissipation `theorems' e.g., for Brownian motion \cite{vanKampen86}.
132:
133: The macroscopic fields that represent the states of systems that possess scale
134: separation are usually taken to be the densities of the conserved entities,
135: namely the (mass or number) density, the momentum density and the energy
136: density. These fields are ``slow'' since any change in their local values
137: necessitates motion to (non-locally) transport an amount of the entity
138: represented by the field. It is usually assumed that on macroscopic time
139: scales the system `remembers' only the slow fields, the other degrees of
140: freedom rapidly accommodating themselves to those fields and becoming
141: ``noise'' whose statistics is determined by the slow fields. The description
142: of far-from-equilibrium systems, such as turbulent fluids and glasses, is not
143: as intuitively founded as that of near-equilibrium systems; in particular,
144: since many of these systems lack scale separation it is a-priori unclear which
145: fields are sufficient to provide a closed description for them (nor is the
146: nature of a relevant statistical ensemble for the unresolved degrees of
147: freedom obvious).
148:
149:
150: The BBGKY hierarchy (see e.g., \cite{HarrisX}) is an exact formulation for
151: many-body systems that does not require them to be reversible
152: \cite{Goldhirsch07a} on the
153: microscopic (particle scale) level or characterized by `slow' or
154: hydrodynamic fields. However, without approximations that
155: truncate it to a manageable size it is nearly useless. A truncation
156: which is valid at low densities (gases) begets the Boltzmann equation
157: \cite{HarrisX,Chapman70}. Like the BBGKY hierarchy, the Boltzmann equation
158: is well defined without the introduction of the concept of temperature
159: or other macroscopic fields and it resolves near-microscopic scales.
160: However it is a nontrivial integrodifferential
161: equation that can be directly solved in very few cases.
162: One of the important uses of the Boltzmann equation is the
163: derivation of hydrodynamic constitutive relations. The form of the
164: hydrodynamic equations (i.e., the equations of continuum mechanics) is
165: very general and holds for single realizations (see Appendix A). They also follow from
166: the Boltzmann equation (whose unknown is a distribution function).
167: A gradient (Chapman-Enskog) expansion of the
168: solution of the Boltzmann equation begets
169: explicit expressions for the hydrodynamic fluxes in terms of
170: gradients of the hydrodynamic fields (constitutive relations).
171: This expansion is formally valid when there is scale separation
172: (see more below) although the resulting hydrodynamic equations have been
173: successfully employed when scale separation is weak or non-existent, such as
174: in the case of weak shocks or granular matter \cite{Goldhirsch03}.
175:
176:
177:
178: Granular systems are different from molecular
179: systems not only in the respective sizes of their constituents
180: but (more importantly) the fact that the grains interact in a dissipative way
181: (e.g., collide inelastically in the granular gas). Therefore all granular
182: states are of non-equilibrium nature. An important consequence of this fact \cite{Tan98} is that granular systems do not possess strong scale
183: separation except in the near-elastic
184: case. In addition, typical granular systems comprise much fewer than
185: (say) an Avogadro
186: number of grains, therefore they are very small by molecular standards.
187: Therefore granular systems are basically
188: {\it mesoscopic} \cite{Tan98}: they are expected to exhibit strong
189: fluctuations within any reasonably defined ensemble and
190: realizations may not be representative of these ensembles, see e.g.,
191: \cite{Goldenberg02,Goldenberg06b}, even in the fluidized phase.
192:
193: Imagine a granular gas which so close to being elastic that the percentage of
194: the energy lost in a collision is minute. Since local equilibration is
195: established in a matter of a few collisions per particle the main effect of the
196: weak inelasticity is to cause the decay of the kinetic energy on long time
197: scales (of course, if the system is forced this `lost' energy can be
198: replenished). Therefore near-elastic granular gases are expected to behave in a
199: way that is close to that of elastic gases, as detailed calculations show
200: \cite{Goldhirsch03}. In other words, the elastic limit is not singular (for
201: finite times) and one expects the temperature to be a relevant field with
202: properties that are similar to those encountered in elastic systems. It is
203: a-priori unclear whether a similar statement holds for strongly inelastic
204: granular gases. It is interesting to note that even in the realm of
205: near-elastic granular gases there is no known equivalent of Boltzmann's
206: H-theorem.
207:
208: Granular solids are jammed in static positions with no possibility to explore
209: phase space. They are athermal. Furthermore, their states are typically
210: metastable (the ground state of a sand-pile is one in which all grains reside
211: on the ground). Although the concept of temperature seems to be a-priori
212: irrelevant to these systems, there have been interesting suggestions to define
213: effective (configurational) temperatures for them as properties of appropriate
214: ensembles, e.g., \cite{Mehta89,Edwards90}. Slow or quasistatic flows of dense
215: granular matter are clearly athermal as well. Here possible dynamic
216: applications of the above mentioned effective temperatures, such as
217: fluctuation-dissipation (FD) relations have been proposed, and tested, see
218: e.g., \cite{Makse02,Mayor05}, indicating a possible dynamical relevance of
219: these temperatures. Unlike the FD relations for granular gases
220: \cite{Goldhirsch00,Brilliantov05,Dufty06} (which are not trivial but can be
221: derived by means borrowed from near-equilibrium statistical mechanics), the FD
222: relations for dense and quasistatic granular flows (or glasses) have been
223: conjectured (see, though \cite{Puglisi02}). Still the above mentioned as well
224: as other findings are interesting and suggestive of a possible underlying
225: theoretical framework. Such a framework is probably not going to enjoy the
226: generality of the near-equilibrium formulations and may necessitate an increase
227: in the number of variables that characterize the considered systems (as is
228: often the case in non-equilibrium situations). A proposal for defining
229: non-equilibrium temperatures which has attracted considerable attention is due
230: to Tsallis \cite{Tsallis88}. Whether this definition is relevant to granular
231: matter is unclear as yet, see though \cite{Beck06}.
232:
233: In old theories of granular matter (gases included) the
234: granular temperature was not
235: accounted for (as it does not correspond to a conserved entity) \cite{Hutter94}. Even in
236: some modern studies the granular temperature is (used but)
237: eliminated from
238: the equations of motion {[}e.g., for length
239: scales for which heat conduction is subdominant to heat production
240: and dissipation \cite{Kumaran06}{]}; in these cases the temperature is
241: claimed to be enslaved to the other fields (but it still plays a useful
242: role as a state characterizing field).
243: Systematic derivations of hydrodynamic equations for granular gases
244: \cite{Sela98,Brey98} invariably account for the granular temperature as a
245: relevant field. These equations have been validated
246: against simulations and experiments \cite{Goldhirsch03}, even when scale
247: separation is weak.
248:
249:
250:
251:
252: The present paper has three major goals. The first is to show that the
253: temperature field is needed for the description of granular gases as a
254: characterizing field, if not a quasi-thermodynamic entity. This
255: is done in Section 2, where the concept of
256: granular temperature is explained in the context
257: of the CE expansion.
258: The goal of Section 3 is to provide an example of
259: a kinetic theory (for binary granular gases)
260: in which the temperature field is
261: one of the hydrodynamic fields and which does produce sensible results even when
262: equipartition is strongly violated (here a novel
263: method for carrying out the
264: CE expansion is presented and applied). The
265: third goal, taken up in Section 4,
266: is to show the existence of a correction to the classical formula
267: for the elastic energy of granular and molecular matter alike; this
268: correction is shown to be related to the heat flux in the equations of
269: continuum mechanics.
270: Section 5 concludes this paper and briefly discusses some issues concerning
271: non-thermodynamic temperatures. Some technical details
272: are relegated to Appendices.
273:
274:
275: \section{Granular Gases and the Chapman-Enskog expansion}
276: \label{section2}
277: As mentioned in the Introduction, the equations of continuum mechanics
278: are very general and
279: in particular, they hold for practically all single
280: realizations of many body systems (see Appendix A). These equations comprise the equation of continuity (for the density, ${\bf \rho}$), Eq.~(\ref{rhodot1}),
281: the momentum density, ${\bf p}$, equation, Eq.~(\ref{velvel}),
282: and the equation for the energy density, $e$,
283: Eq.~(\ref{energy}), in which
284: (in the dissipative/inelastic case)
285: an energy sink term, ${\bf \rho \Gamma}$, arises. An equation for the
286: internal specific energy easily follows from the above,
287: see Eq.~(\ref{internal}). Note that the derivation in Appendix A
288: produces also explicit
289: expressions for the relevant fields and fluxes in terms of the microscopic
290: degrees of freedom. In particular, note that the expression for the stress
291: tensor is not precisely the same as the commonly used Voigt or Born or Kirkwood
292: expression; the latter are limits of Eq.~(\ref{stress}) when the coarse-graining
293: volume (or width of the coarse-graining function) is very large and do not
294: satisfy the equations of continuum mechanics unless there is scale separation
295: or one considers homogeneous deformations of lattice configurations.
296:
297: In the realm of molecular or granular gases, when the internal potential
298: energy is negligible (for dilute systems) or when a model of hard spheres
299: is employed one can replace the specific internal energy by
300: the (granular) temperature, $ T$.
301: In this case Eq.~(\ref{internal}) can be rewritten as:
302: \begin{equation}
303: \rho({\bf r},t) \frac{DT({\bf r},t)}{Dt}
304: = \frac{\partial V_\beta({\bf r},t)}{\partial r_\alpha} \sigma_{\beta\alpha}
305: ({\bf r},t) - \mathrm{div}\, {\bf Q}({\bf r},t) - \rho({\bf r},t)\Gamma({\bf r},t) \label{Temp}
306: \end{equation}
307: with notation that is defined in Appendix A.
308: Following Eq.~(\ref{defT}) the (granular)
309: temperature field is given by the following intuitive expression:
310: \begin{equation}
311: T({\bf r},t)= \frac{\frac{1}{2}\sum_i m_i v^{\prime 2}_i(t) \phi\left[ {\bf r} - {\bf r}_i(t) \right]}{\sum_j m_j \phi\left[ {\bf r} - {\bf r}_j(t)\right]}
312: \end{equation}
313: Note that when scale separation does not exist, the
314: temperature, much like other fields, may be resolution dependent
315: (i.e. it
316: depends on the width of the coarse graining function $\phi$)\cite{Glasser01}.
317:
318: %\subsection{On granular hydrodynamics}
319:
320: The hydrodynamic description of a fluid, granular or otherwise,
321: is complete when constitutive relations express $\mbox{\boldmath$\sigma$}$, ${\bf Q}$ and
322: $\Gamma$ in terms of the hydrodynamic fields.
323: It is {\it a-priori}
324: unclear whether such a closure exists for granular fluids. Extended
325: Chapman-Enskog (CE) expansions provide explicit
326: constitutive relations for granular gases, as further detailed below.
327:
328:
329: The Boltzmann equation is a
330: non-trivial integrodifferential equation for the single-particle velocity
331: distribution, $f({\bf v},{\bf r},t)$, whose integral over the (particle)
332: velocities, ${\bf v}$, is the number density, $n({\bf r},t)$. The ratio
333: $f({\bf v},{\bf r},t)/n({\bf r},t)$ is the probability density for a particle
334: to have velocity ${\bf v}$ when it is found (or more accurately its center
335: of mass is found) at point ${\bf r}$ at time $t$. In the relatively simple
336: case of a dilute monodisperse collection of smooth hard
337: spheres of unit mass and diameter
338: $d$ whose collisions are characterized by a fixed coefficient of normal
339: restitution, $\alpha$, the Boltzmann equation reads
340: \cite{Sela98,Brey98,Goldshtein95}:
341: \begin{displaymath}
342: \frac{\partial f({\bf v}_1)}{\partial t}+{\bf v_{1}}\cdot\mbox{\boldmath$\nabla$}f ({\bf v}_1)
343: \end{displaymath}
344: \begin{equation}
345: =
346: d^2 \int_{{\bf \hat{k}}\cdot{\bf v_{12}}>0}d{\bf v_{2}}
347: d{\bf \hat{k}}({\bf \hat{k}}\cdot{\bf v_{12}})
348: \left(\frac{1}{{\alpha}^2}f({\bf v'_{1}})f({\bf v'_{2}})
349: -f({\bf v_{1}})f({\bf v_{2}})\right)\label{Boltzmann101}
350: \end{equation}
351: where $\mbox{\boldmath$\nabla$}$ is a spatial gradient. The unit vector $\hat{k}$ points from the center of
352: particle `1' to the center of particle `2'. The relation between
353: the precollisional (primed) and postcollisional
354: velocities is given by:
355: ${\bf v_{i}}={\bf v'_{i}}-\frac{1+\alpha }{2}({\bf \hat{k}}\cdot{\bf v'_{ij}})
356: {\bf \hat{k}}$
357: where ${\bf v'_{ij}}\equiv{\bf v'_{i}}-{\bf v'_{j}}$.
358: The dependence of $f$
359: on the spatial coordinates and time is not explicitly spelled out in
360: Eq. (\ref{Boltzmann101}).
361: Notice that in addition to the explicit dependence of Eq. (\ref{Boltzmann101}) on $\alpha$, it also
362: implicitly depends on $\alpha$ through the
363: relation between the postcollisional and precollisional velocities. The
364: condition $\hat{k}\cdot {\bf v}_{12}> 0$ represents the fact that only particles whose relative velocity is such that they approach each other, can collide.
365:
366: In the
367: derivation of the Boltzmann equation \cite{HarrisX,Chapman70} it is assumed
368: that the velocities and positions of
369: colliding particles are not correlated. This assumption of ``molecular
370: chaos'' is
371: justified for dilute gases but not for dense gases or
372: strongly inelastic granular gases \cite{Tan98,Goldhirsch93a,Soto01,Luding00}. This fact is related to the reduction in the normal component
373: of the relative velocity of colliding particles. In relatively dense gases
374: particles heading towards
375: each other may not collide as they are `stopped' by collisions with
376: other particles. This phenomenon is modeled in the
377: Enskog-corrected Boltzmann equation \cite{Chapman70}.
378:
379: The Chapman-Enskog expansion is a method to perturbatively solve the
380: Boltzmann equation. Scale separation is formally
381: necessary for its implementation.
382: When all gradients vanish, a local Maxwellian distribution, corresponding to the values of the hydrodynamic fields at each point, solves
383: the Boltzmann equation in the elastic case. A gradient expansion (formally, an expansion in the
384: Knudsen number) around this solution automatically yields the result
385: that the
386: distribution function depends on space and time only through its dependence
387: on the fields (which is quasi-local since $f({\bf r},t)$ depends
388: on the values of the fields and their spatial derivatives at
389: $({\bf r},t)$). The often made statement that within the CE expansion
390: $f$ is assumed to depend on space and time only through its
391: dependence on the hydrodynamic fields is not precise: this assumption is made only for the zeroth order of the expansion.
392:
393:
394:
395: In the case of granular gases, the
396: (kinetic) energy density is not strictly conserved, and this led
397: to questions whether it should be included in the set of hydrodynamic
398: fields \cite{Goldhirsch03,Kumaran06,Noije00,Wakou02}.
399: Following the discussion in the introduction, when the degree
400: of inelasticity is sufficiently small, it is justified to include the
401: granular temperature in the set of
402: hydrodynamic fields. An extension of the Chapman Enskog expansion
403: based on this approach has been proposed and implemented in \cite{Sela98}.
404: There, the small parameters used to expand the Boltzmann equation were
405: the Knudsen number,
406: $K\equiv\frac{\ell}{L}$, where $\ell$ is the mean free
407: path given by $\ell=\frac{1}{\pi n d^2}$, and $L$ is a macroscopic length
408: scale (i.e., the length scale which is resolved by hydrodynamics, not
409: necessarily the system size) {\it and} the degree of inelasticity, defined
410: by $\epsilon\equiv 1-{\alpha}^2$. In another approach \cite{Brey98} to the problem
411: only the Knudsen number is the small parameter and the temperature is assumed
412: to comprise a relevant hydrodynamic field for all values of the
413: degree of inelasticity. This method does not seem to lead to contradictions
414: for any value of $\alpha$; as a matter of fact it
415: has recently been used to obtain accurate values for the linear transport
416: coefficients for all physically allowed values of $\alpha$
417: \cite{Noskowicz06}. Whether the results are indeed
418: relevant to strongly inelastic granular gases, where scale separation
419: does not hold and precollisional correlations may be prominent, remains to
420: be seen.
421:
422:
423: \section{Binary granular gases: extreme violations of equipartition}
424:
425:
426: Consider a mixture of smooth hard spheres, composed of species $A$
427: and $B$, of masses $m_{A}$ and $m_{B}$, and diameters
428: $\sigma_{A}$ and $\sigma_{B}$, respectively.
429: The coefficient of normal restitution (assumed to be fixed) for a collision of a particle of
430: species $\alpha \in \{ A,B \}$ with a particle of species $\beta \in
431: \{ A,B\}$ is denoted by $e_{\alpha\beta}$ (hence, $e_{\alpha\beta}
432: \in \{ e_{AA}, \ e_{BB}, \ e_{AB} \} $).
433: The transformation of velocities due to a collision of a sphere of species
434: $\alpha$ with a sphere of species $\beta$ is given by:
435: \begin{eqnarray}
436: \mathbf{v}_{1} &=&\mathbf{v}_{1}^{\prime }-\left( 1+e_{\alpha \beta }\right)
437: M^{\alpha\beta}\left( \mathbf{v}_{12}^{\prime }\cdot \hat{k}\right) \hat{k}
438: \label{vtransfo} \\
439: \mathbf{v}_{2} &=&\mathbf{v}_{2}^{\prime }+\left( 1+e_{\alpha \beta }\right)
440: M^{\alpha\beta}\left( \mathbf{v}_{12}^{\prime }\cdot \hat{k}\right) \hat{k}
441: \end{eqnarray}
442: where $\{ \mathbf{v}^\prime_1, \mathbf{v}^\prime_2 \}$ denote the precollisional
443: velocities of the spheres (the index `$1$' refers here to species $\alpha$),
444: and $\{ \mathbf{v}_1, \mathbf{v}_2 \}$ are
445: the corresponding postcollisional velocities;
446: $\hat{k}$ is a unit vector
447: pointing from the center of sphere $\alpha $ to that of sphere $\beta $, and
448: $M^{\alpha\beta}\equiv \frac{m_{\alpha }}{m_\alpha+m_\beta}$,
449: $\mathbf{v}_{12} \equiv
450: \mathbf{v}_{1}-\mathbf{v}_{2}$, a similar definition holding for the
451: primed (precollisional) velocities. Obvious kinematic constraints require
452: that $\mathbf{v}^\prime_{12}\cdot \hat{k} \geq 0$.
453: Define the degrees of inelasticity as:
454: $\varepsilon_{\alpha\beta}\equiv 1-e_{\alpha\beta}^{2}$.
455:
456:
457: The kinetic description of a binary granular gas mixture involves two
458: Boltzmann equations, one for each
459: species \cite{Garzo02,Serero06}:
460: \begin {equation}
461: Df_{\alpha} \equiv
462: \frac{\partial f_{\alpha }}{\partial t}+\mathbf{v}_{1}\cdot
463: \mathbf{\nabla }f_{\alpha }=\mathcal{B}(f_{\alpha},f_{\alpha},e_{\alpha \alpha})+\mathcal{B}(f_{\alpha},f_{\beta},e_{\alpha \beta})\ \ \ \ \ \alpha \ne \beta,
464: \label{Boltzmann}
465: \end{equation}
466: where $f_{\alpha }\left( \mathbf{v}\right) $ is the
467: distribution function for particles of species $\alpha$, and
468: \begin{displaymath}
469: \mathcal{B}(f_{\alpha},f_{\beta},e_{\alpha \beta})
470: \end{displaymath}
471: \begin{equation}
472: =\sigma _{\alpha \beta }^{2}\int \int_{\mathbf{v}_{12}\cdot
473: \mathbf{k}>0}\left[ \frac{f_{\alpha }(\mathbf{v}_{1}^{\prime
474: })f_{\beta }(\mathbf{v}_{2}^{\prime })}{e_{\alpha \beta }^{2}}
475: -f_{\alpha }(\mathbf{v}_{1})f_{\beta }(\mathbf{v}_{2})\right]
476: \bigg( \mathbf{v}_{12}\cdot\mathbf{k}\bigg) \ d\mathbf{
477: v}_{2}\ d\mathbf{k},
478: \label{Boltzmannop}
479: \end{equation}
480: where $\sigma _{\alpha \beta }\equiv \frac{\sigma _{\alpha }+\sigma _{\beta }}{2}$. Here $\mathbf{v}_{1}$ and $\mathbf{v}_{1}^{\prime}$ pertain to species ${\alpha}$, and $\mathbf{v}_{2}$ and $\mathbf{v}_{2}^{\prime}$ pertain to $\beta$. Notice that $\mathcal{B}(f_{\alpha},f_{\beta},e_{\alpha \beta})$ depends on the coefficients of normal restitution both explicitly, as
481: shown in Eq. (\ref{Boltzmannop}), and implicitly through the
482: collision law.
483:
484: The hydrodynamic fields
485: in the present case comprise the
486: two number densities, $n_A$ and $n_B$ (or the mass densities $\rho_A=n_Am_A$ and $\rho_B=n_Bm_B$), the mixture's
487: velocity field, $\mathbf{V}$, and the temperature field, $T$, defined
488: (differently from the monodisperse case) as twice the
489: mean fluctuating kinetic energy of a particle (these are the only slow fields in the elastic limit), as explicitly defined below.
490: The continuum equations of motion
491: follow from the pertinent Boltzmann equation directly
492: (using the standard
493: procedure of computing velocity moments of the equation). Their validity
494: is general since they are based on the underlying
495: conservation laws.
496: The equation of motion for the number density, $n_\alpha$,
497: with $\alpha \in \{ A,B\}$, is:
498: \begin{equation}
499: \frac{Dn_{\alpha}}{Dt}=-{\mbox{div\ }} \mathbf{J}_{\alpha} -n_{\alpha}~{\mbox{div\ }}%
500: \mathbf{V} \label{dynameq_na}
501: \end{equation}
502: where $\frac{D}{Dt}$ is the material derivative and
503: %\begin{equation}
504: $\mathbf{J}_{\alpha}=n_{\alpha}\left( \mathbf{V}_{\alpha}-\mathbf{V}%
505: \right) \label{field_ja}$
506: %\end{equation}
507: is the particle flux density of species $\alpha$. As ${\mathbf{V}}_\alpha$,
508: the velocity field of
509: species $\alpha$ (or the flux ${\mathbf{J}}_{\alpha}$) is not
510: a hydrodynamic field, it must be given by an appropriate constitutive
511: relation.
512: The velocity field obeys (as expected):
513: %\begin{equation}
514: $\rho \frac{DV_{i}}{Dt}=-\frac{\partial \sigma_{ij}}{\partial x_{j}}+\rho \mathbf{g}$,
515: %\label{dynameanveloc}
516: %\end{equation}
517: where the summation convention is used,
518: and $\sigma_{ij}$ is the stress tensor.
519: The granular temperature field obeys:
520: %\begin{equation}
521: $n\frac{DT}{Dt}=T~{\mbox{div\ }}~\mathbf{J}-{\mbox{div\ }}~\mathbf{Q}
522: -2\sigma_{ij}\frac{\partial V_i}{{\partial x_j}}-\Gamma$,
523: % \label{dynatemp}
524: %\end{equation}
525: where ${\mathbf{J}} \equiv {\mathbf{J}}_A + {\mathbf{J}}_B$ is the total
526: particle flux, and $\mathbf{Q}$ is the heat flux.
527:
528: The form of the constitutive relations can be easily determined
529: from tensorial (and symmetry) considerations, the result being (to
530: linear order in the gradients, or Navier-Stokes order):
531: $
532: \sigma_{ij}=p\delta _{ij}-2\mu~D_{ij} -\delta_{ij} \ \eta_{\mbox{\tiny{B}}}~{\mbox{div\ }} \mathbf{V}$,
533: where
534: %\begin{equation}
535: $D_{ij}=\frac{1}{2}\ \bigg( \frac{\partial V_{i}}{\partial x_{j}}+\frac{
536: \partial V_{j}}{\partial x_{i}} -\frac{2}{3}\ \delta_{ij} \ \mbox{div}{\bf V}\bigg)$
537: is the traceless rate of strain tensor, $\mu$ is the shear viscosity,
538: $\eta_{\mbox{\tiny{B}}}$ is the
539: bulk viscosity (which vanishes in the dilute limit) and $p$ is the
540: pressure. The diffusion flux can be rewritten as
541: \begin{equation}
542: \mathbf{J}_{\alpha}=\frac{n_{\alpha}}{n}\frac{1}{\sigma_{AB}^2}\sqrt{\frac{T}{m_{\alpha}}}
543: \bigg(-\kappa_{\alpha}^{T}\mathbf{\nabla}\ln T -\kappa_{\alpha}^{n}\mathbf{\nabla}\ln n-\kappa_{\alpha}^{c}\mathbf{\nabla}\ln c \bigg),
544: \label{cr_jalpha2}
545: \end{equation}
546: where $\kappa_\alpha^T$, $\kappa_\alpha^{n}$ and $\kappa_\alpha^{c}$
547: are non-trivial functions of the parameters
548: $${\mathcal{S}} \equiv \{ \{ \varepsilon_{\alpha\beta} \}, \
549: M_A\equiv \frac{m_A}{m_A+m_B},\ \frac{\sigma_A}{\sigma_A + \sigma_B},\ c \}, $$
550: where the concentration field, $c$, is defined as $c\equiv \frac{n_A}{n}
551: =\frac{n_A}{n_A+n_B}$.
552: Similarly, the heat flux can be rewritten as
553: \begin{equation}
554: \mathbf{Q}=\frac{1}{\sigma_{AB}^2}\frac{T^{3/2}}{\sqrt{m_{0}}}\bigg( -\lambda^{T} \mathbf{\nabla}\ln T
555: -\lambda^{n}\mathbf{\nabla}\ln n-\lambda^{c}\mathbf{\nabla}\ln c \bigg)
556: \label {cr_q4}
557: \end{equation}
558: where $m_0 \equiv m_A + m_B$. In the dilute limit the
559: equation of state is the same as that for an
560: ideal gas: $p = \frac{nT}{3}$.
561: Much like in the monodisperse case
562: the number density for species $\alpha$ is given by:
563: %\begin{equation}
564: $n_{\alpha}=\int f_{\alpha} ( \mathbf{v}) \ d\mathbf{v} $,
565: %\label{density}
566: %\end{equation}
567: the corresponding mass density being
568: $\rho _{\alpha}=m_{\alpha}n_{\alpha}$; the overall number density is $n\equiv n_{A}+n_{B}$, and the overall mass density is $\rho\equiv \rho_{A}+\rho_{B}$.
569: The velocity field of species $\alpha$ is given by:
570: %\begin{equation}
571: $\mathbf{V}_{\alpha}=\frac{1}{n_{\alpha}}\int f_{\alpha}
572: (\mathbf{v})\ \mathbf{v}
573: \ d\mathbf{v} $.
574: %\label{speciesvelocity}
575: %\end{equation}
576: As mentioned, $\mathbf{V}_\alpha$ is not a hydrodynamic field, and needs
577: to be expressed as a functional of the hydrodynamic fields.
578: The mixture's velocity field is:
579: $\mathbf{V}=\frac{1}{\rho }\left( \rho _{A}\mathbf{V}_{A}+\rho _{B}\mathbf{V}
580: _{B}\right)$. The granular temperature of species $\alpha$ is defined by:
581: %\begin{equation}
582: $T_{\alpha}=\frac{1}{n_{\alpha}}\int f_{\alpha}(
583: \mathbf{v})\ m_{\alpha}\ (\mathbf{v}-\mathbf{V})^{2}\ d
584: \mathbf{v}\ $
585: % \label{specitemp}
586: %\end{equation}
587: (not a hydrodynamic field).
588: The velocity fluctuations of each of the species
589: are measured with respect to the (hydrodynamic)
590: mixture's velocity field, not the species' velocity fields.
591: The mixture's granular temperature
592: is defined as:
593: $T=\frac{1}{n}\left( n_{A}T_{A}+n_{B}T_{B}\right)$.
594: The kinetic expression for the stress tensor \cite{Chapman70} is:
595: %\begin{equation}
596: $\sigma_{ij}=m_{A}\int f_{A}\left( \mathbf{v}\right) u_{i}u_{j}~d{\bf v}+m_{B}\int f_{B}\left( \mathbf{v}\right) u_{i}u_{j}~d{\bf v}$,
597: %\label{stresstensor}
598: %\end{equation}
599: where ${\mathbf{u}} \equiv \mathbf{v} - {\mathbf{V}}$ is the peculiar
600: (fluctuating) velocity of a particle (irrespective of the species).
601: Similarly, the heat flux is composed of two contributions, for obvious reasons:
602: %\begin{equation}
603: $\mathbf{Q}=\int f_{A}\left( \mathbf{v}\right) m_{A}u^{2}
604: \mathbf{u}\ d{\bf v} +\int f_{B}\left( \mathbf{v}\right) m_{B}u^{2}
605: \mathbf{u}\ d{\bf v}$.
606: %\label{heatflux}
607: %\end{equation}
608: Our definition of the heat flux
609: may differ by a factor of $2$ from some other definitions.
610: The sink term is given by $\Gamma = \Gamma_A + \Gamma_B + \Gamma_{AB}$,
611: where:
612: \begin{equation}
613: \Gamma _{\alpha} \equiv \varepsilon_{\alpha\alpha} \frac{ m_{\alpha}
614: \pi\sigma_{\alpha}^{2}}{8}
615: \int \int f_{\alpha}\left( \mathbf{v}_{1}\right) f_{\alpha}\left(
616: \mathbf{v}_{2}\right) \left| v_{12}\right| ^{3}d{\bf v}_{1}\ d{\bf v}_{2},
617: \label {GammaA}
618: \end{equation}
619: and
620: \begin{equation}
621: \Gamma _{AB}\equiv \varepsilon _{AB}\frac{m_{AB}\pi \sigma_{AB}^{2}}{2}\int \int
622: f_{A}\left( \mathbf{v}_{1}\right) f_{B}\left (\mathbf{v}_{2}\right) \left| \mathbf{v}%
623: _{12}\right| ^{3}d{\bf v}_{1}\ d{\bf v}_{2}. \label{GammaAB}
624: \end{equation}
625:
626:
627: The HCS distributions for binary granular gas mixtures (much
628: like in the monodisperse case) are
629: scaling solutions, in which the distribution functions are rendered time independent
630: when the velocities are
631: scaled by the respective (decaying) thermal speeds of the species comprising the
632: mixture. It serves as a zeroth order in the extended Chapman-Enskog
633: expansion of \cite{Garzo02}.
634: The lhs of the Boltzmann equations for the HCS can be rewritten in this case as: %\begin{equation}
635: $\frac{\partial f^{\mbox{\tiny{HCS}}}_{\alpha }}{\partial t}=\frac{\partial f^{\mbox{\tiny{HCS}}}_{\alpha }}{\partial
636: T}\frac{\partial T}{\partial t}=-\frac{\Gamma }{n}\frac{\partial f^{\mbox{\tiny{HCS}}}_{\alpha }
637: }{\partial T}$
638: %\end{equation}
639: and the two Boltzmann equations become:
640: \begin{equation}
641: -\frac{\Gamma }{n}\frac{\partial f^{\mbox{\tiny{HCS}}}_{\alpha}}{\partial T}=
642: \mathcal{B}\left(
643: f^{\mbox{\tiny{HCS}}}_{\alpha},f^{\mbox{\tiny{HCS}}}_{\alpha,}e_{\alpha\alpha}\right) +\mathcal{B}\left( f^{\mbox{\tiny{HCS}}}_{\alpha},f^{\mbox{\tiny{HCS}}}_{\beta,}
644: e_{\alpha\beta}\right) \ \ \ \ \ \ \ \ \alpha \ne \beta \label{boltHCSA}
645: \end{equation}
646: These equations need to be solved subject to the following `constraints'
647: which fix the values of the hydrodynamic fields:
648: $\int f^{\mbox{\tiny{HCS}}}_{A}d\mathbf{u} =n_{A}$, $\int f^{\mbox{\tiny{HCS}}}_{B}d\mathbf{u} =n_{B}$ and
649: $\int f^{\mbox{\tiny{HCS}}}_{A}m_{A}u^{2}d\mathbf{u+}\int f^{\mbox{\tiny{HCS}}}_{B}m_{B}u^{2}d\mathbf{u} =nT$.
650: The common way of solving Eqs. (\ref{boltHCSA}) is to represent
651: the functions, $\ f^{\mbox{\tiny{HCS}}}_{\alpha }$,
652: by Sonine polynomial series times Maxwellians in the respective non-dimensionalized
653: (peculiar) velocities. However, it can be shown \cite{Noskowicz06} that due to the exponential
654: tails of the HCS distributions these expansions are not convergent
655: yet they are asymptotic and Borel resummable. To overcome this difficulty, consider the modified expansion:
656: \begin{equation}
657: f^{\mbox{\tiny{HCS}}}_{\alpha }=f_{\alpha}^{M,\eta}\phi
658: _{\alpha } \equiv n_{\alpha }\left( \frac{\gamma
659: _{\alpha }}{\pi }\right) ^{\frac{3}{2}}e^{-\eta \gamma _{\alpha }u^{2}}\phi
660: _{\alpha } \label{expansion}
661: \end{equation}
662: where $\gamma _{\alpha }=\frac{3m_{\alpha }}{2T}$, $\phi _{\alpha
663: }\equiv\sum_{p=0}^{\infty }h_{\alpha }^{p}S_{\frac{1}{2}}^{p}\left(
664: \gamma _{\alpha }u^{2}\right) $, and $\eta>0$ is a constant, i.e. $f_{\alpha }$ is represented by
665: a truncated Sonine polynomial series (whose coefficients are denoted
666: by $ \{ h^p_\alpha \} $) times a Maxwellian in which the
667: temperature is replaced by a ``wrong'' temperature $\frac{T}{\eta }$. It
668: can be shown \cite{Noskowicz06}
669: that this method produces convergent series when $\eta
670: <0.5$.
671: Upon substituting the form (\ref{expansion}) in Eq. (\ref{boltHCSA}) and projecting on the Nth order Sonine polynomial $S_{\frac{1}{2}}^{N}\left( \gamma _{\alpha }u_{1}^{2}\right) $, one obtains the following non linear algebraic system for the coefficients $h_{\alpha}^p$:
672: \begin{eqnarray}
673: \frac{n_{\alpha}\Gamma }{nT}\sum\limits_{p}^{ }h_{\alpha}^{p}R_{pN}
674: &=&\frac{M_{\alpha }^{3}n_{\alpha }^{2}\sigma _{\alpha }^{2}}{
675: \pi ^{3}}\sqrt{\frac{2T}{3m_{0}}}\sum\limits_{p,q}B_{\alpha
676: \alpha }^{pqN}h_{\alpha }^{p}h_{\alpha }^{q} \nonumber\\
677: &+&\chi _{\alpha \beta }\frac{\left( M_{\beta }M_{\alpha }\right) ^{\frac{3}{
678: 2}}n_{\alpha }n_{\beta }\sigma _{\alpha \beta }^{2}}{\pi ^{3}}\sqrt{\frac{2T
679: }{3m_{0}}}\sum\limits_{p,q} B_{\alpha \beta }^{pqN}h_{\alpha
680: }^{p}h_{\beta }^{q}\label{Bolteqexp}
681: \end{eqnarray}
682: where $m_0 \equiv m_A+m_B$, $M_{\alpha} \equiv \frac{m_{\alpha}}{m_0}$, and $R_{pN}$ is defined as follows:
683: \begin{displaymath}
684: R_{pN}=\frac{1}{\pi ^{\frac{3}{2}}}\int e^{-\eta u_{\alpha}^{2}}S_{\frac{1}{2}
685: }^{N}\left( u_{\alpha}^{2}\right)
686: \end{displaymath}
687: \begin{equation}
688: \times
689: \left[ \frac{3}{2}S_{\frac{1}{2}}^{p}\left(
690: u_{\alpha}^{2}\right) -u^{2}\left( \eta S_{\frac{1}{2}}^{p}\left( u_{\alpha}^{2}\right) -\frac{
691: \partial }{\partial u_{\alpha}^{2}}S_{\frac{1}{2}}^{p}\left( u_{\alpha}^{2}\right)
692: \right) \right] d\mathbf{u_{\alpha}} \label{int_R}
693: \end{equation}
694: where $\mathbf{u}_{\alpha}\equiv \sqrt{\gamma_{\alpha}}\mathbf{u}$, and the coefficients (or coupling constants), $B_{\alpha \beta }^{pqN}$, are defined, in terms of the rescaled velocities $\tilde{\mathbf{u}} \equiv \sqrt{\frac{3m_0}{2T}}\mathbf{u}$, by:
695: \begin{eqnarray}
696: B_{\alpha \beta }^{pqN} &= &\int S_{\frac{1}{2}}^{N}\left( M _{\alpha }\tilde{u}_{1}^{2}\right)\nonumber \int \int_{\mathbf{\tilde{u}}_{12}\mathbf{\cdot k}>0}\bigg[\frac{1}{e_{\alpha \beta }^{2}}
697: e^{-\eta M _{\alpha }\tilde{u}_{1}^{\prime 2}-\eta M _{\beta
698: }\tilde{u}_{2}^{\prime 2}} S_{\frac{1}{2}}^{p}\left( M _{\alpha }\tilde{u}_{1}^{\prime 2}\right)
699: S_{\frac{1}{2}}^{q}\left( M _{\beta }\tilde{u}_{2}^{\prime 2}\right)\nonumber\\
700: & & -e^{-\eta
701: M _{\alpha }\tilde{u}_{1}^{2}-\eta M _{\beta }\tilde{u}_{2}^{2}}S_{\frac{1}{2}}^{p}\left( M _{\alpha }\tilde{u}_{1}^{2}\right) S_{
702: \frac{1}{2}}^{q}\left( M _{\beta }\tilde{u}_{2}^{2}\right)\bigg] \left( \mathbf{\tilde{u}}
703: _{12}\mathbf{\cdot k}\right) d\mathbf{\tilde{u}}_{1}d\mathbf{\tilde{u}}_{2}d\mathbf{k}
704: \label{int_B}
705: \end{eqnarray}
706: Though straightforward in principle, the computation of these coefficients
707: turns out to be forbiddingly tedious as the order of the Sonine expansions is
708: increased. In order to be able to carry out high order expansions, we made use of a computer-aided method exploiting the fact that
709: the Sonine polynomials can be derived from their respective
710: generating functions,
711: $G_m \left(s,x\right)$:
712: \begin {equation}
713: G_m\left( s,x\right)=\left( 1-s\right)
714: ^{-m-1}e^{-\frac{s}{1-s}x} =
715: \sum\limits_{p=0}^\infty s^{p}S_{m}^{p}\left( x\right)
716: \label{Sonine_gf}
717: \end{equation}
718: This fact enables one to define generating functions for the
719: integrals (\ref{int_R}) and (\ref{int_B}): $R\left( s,w\right)
720: \equiv \sum\limits_{p,N}s^{p}w^{N}R_{pN}$ and
721: $B_{\alpha
722: \beta }\left( w,s,t\right)
723: \equiv\sum\limits_{p,q,N}s^{p}t^{q}w^{N}B_{\alpha \beta }^{pqN}$
724: of
725: \ the coefficients, $B_{\alpha \beta }^{pqN}$, and $R_{pN}$,
726: respectively (where all sums here and below range from zero to infinity). The latter can then be computed (using a symbolic processor
727: such as MAPLE$^{\mbox{\tiny{TM}}}$) to any desired order by
728: taking successive derivatives of $R\left( s,w\right) $ and $B_{\alpha
729: \beta }\left( w,s,t\right)$.
730: %\begin{equation}
731: %R_{pN} = \frac{1}{p!N!}\frac{\partial^{p+N}R\left( w,s\right)}{\partial s^{p}\partial w^{N}}\rfloor s%=0,w=0}
732: %\end{equation}
733: %and
734: %\begin{equation}
735: %B_{\alpha \beta }^{pqN} = \frac{1}{p!q!N!}\frac{\partial^{p+q+N}B_{\alpha
736: %\beta }\left( w,s,t\right)}{\partial s^{p}\partial t^{q} \partial w^{N}}\rfloor _{s=0,t=0,w=0}
737: %\end{equation}
738: %where:
739: These generating functions are given by:
740: \begin{equation}
741: R\left( s,w\right) =\frac{3}{2}\frac{\left( 1-s\right) w}{\left( 1-ws+\left(
742: \eta -1\right) \left( 1-w\right) \left( 1-s\right) \right) ^{\frac{5}{2}}}
743: \label{geneR}
744: \end{equation}
745: and using Eq. (\ref{int_B})
746: \begin{eqnarray}
747: B_{\alpha \beta } &=&\left( 1-s\right) ^{-\frac{3}{2}
748: }\left( 1-t\right) ^{-\frac{3}{2}}\left( 1-w\right) ^{-\frac{3}{2}}\nonumber \\
749: &&\times \bigg[ \frac{1}{e_{\alpha \beta }^{2}} I_{\alpha \beta }\left(
750: \frac{M_{\alpha }w}{1-w},0,\left( \frac{s}{1-s}+\eta \right) M_{\alpha
751: },\left( \frac{t}{1-t}+\eta \right) M_{\beta }\right) \nonumber\\
752: &&\qquad-I_{\alpha \beta }\left( M_{\alpha }\left( \eta +\frac{w}{1-w}+\frac{s}{1-s
753: }\right) ,\left( \frac{t}{1-t}+\eta \right) M_{\beta }\right) \bigg]\nonumber
754: \end{eqnarray}
755: where $I_{\alpha \beta }\left( a,b,c,d,x,y,z\right) $ is the integral defined by:
756: \begin{equation}
757: I_{\alpha \beta }\left( a,b,c,d,x,y,z\right) \equiv \int_{\mathbf{u}_{12}
758: \mathbf{\cdot k}>0}d\mathbf{u}_{1}d\mathbf{u}_{2}d\mathbf{k}\left(
759: \mathbf{u}_{12} \cdot \mathbf{
760: k} \right) e^{-F}
761: \label{IAB}
762: \end{equation}
763: where
764: $
765: F=au_{1}^{2}+bu_{2}^{2}+cu_{1}^{\prime 2}+du_{2}^{\prime 2}+\frac{x}{2}%
766: \left( \mathbf{u}_{1}+\mathbf{u}_{1}^{\prime }\right) ^{2}+\frac{y}{2}\left(
767: \mathbf{u}_{1}\mathbf{+u}_{2}^{\prime }\right) ^{2}+\frac{z}{2}\left(
768: \mathbf{u}_{1}\mathbf{+u}_{2}\right) ^{2}
769: $ and $I_{\alpha\beta}\left( a, b \right) \equiv I_{\alpha\beta}\left(a,b,0,0
770: \right)$.
771: It is easy to calculate $I_{\alpha \beta } $, the result being:
772: \begin{equation}
773: I_{\alpha \beta }=\frac{2\pi ^{\frac{7}{2}}}{\lambda ^{\frac{3}{2}}}\frac{1}{
774: \mu _{\alpha \beta }\left( \nu _{\alpha \beta }+\mu _{\alpha \beta }\right) }
775: \end{equation}
776: where $\lambda=a+b+c+d+2x+2y+2z$, $\mu _{\alpha \beta }=R_{\alpha \beta }-\frac{K_{\alpha \beta }^{2}}{\lambda },\ \
777: \nu _{\alpha \beta }=S_{\alpha \beta }$\\
778: $$-\frac{t^{\alpha\beta}}{\lambda }\left(
779: \left( d\mathbf{+}y\right) M^{\alpha \beta }-\left( c+x\right) M^{\beta
780: \alpha }\right) \left( 2K_{\alpha\beta}+t^{\alpha\beta}\left( \left( d\mathbf{+}y\right) M^{\alpha \beta
781: }-\left( c+x\right) M^{\beta \alpha }\right) \right), $$ \\
782: $R_{\alpha \beta } =\left( a+c+2x\right) \left( M^{\beta \alpha }\right)
783: ^{2}+\left( b+d\right) \left( M^{\alpha \beta }\right) ^{2}+\frac{1}{2}%
784: \left( M^{\alpha \beta }-M^{\beta \alpha }\right) ^{2}\left( y+z\right)$, \\
785: $S_{\alpha \beta } =\frac{1}{2}t^{\alpha\beta}\left[ \left( t^{\alpha\beta}-2\right) \left( \left(
786: 2c+x\right) \left( M^{\beta \alpha }\right) ^{2}+\left( M^{\alpha \beta
787: }\right) ^{2}\left( 2d+y\right) \right) +2M^{\alpha \beta } M^{\beta\alpha} y
788: \right. $$\\
789: $$\left.
790: -2\left(
791: M^{\beta \alpha }\right) ^{2}x\right] $,
792: $K_{\alpha \beta }=\left( M^{\beta \alpha }\left( a+c+2x+y+z\right)
793: -M^{\alpha \beta }\left( b+d+y+z\right) \right) $, and
794: $t^{\alpha\beta}\equiv \frac{1+e_{\alpha\beta}}{e_{\alpha\beta}}$.
795: Next, using Eq.~(\ref{expansion}) for $f_{\alpha }$ in Eq. (\ref{GammaA}) and (\ref{GammaAB}), multiplying the result by $s^pt^q$ and summing over $p$ and $q$, one obtains:
796: $\Gamma =\frac{\sqrt{6\pi }}{9}\frac{n^{2}\sigma _{AB}^{2}T^{\frac{3}{2}}}{
797: \sqrt{m_{0}}}\widetilde{\Gamma }
798: $, where
799: \begin{eqnarray}
800: \widetilde{\Gamma } &\equiv&\varepsilon _{AA}\frac{n_{A}^{2}}{n^{2}}
801: \frac{\sigma _{A}^{2}}{\sigma _{AB}^{2}}M_{A}^{4}\sum_{p}\sum_{p}\Gamma _{AA}^{pq}h_{A}^{p}h_{A}^{q}
802: %\nonumber \\&&
803: +4\varepsilon _{AB}\frac{n_{A}n_{B}}{n^{2}}M_{A}^{\frac{5}{2}
804: }M_{B}^{\frac{5}{2}}\sum_{p}\sum_{q}\Gamma
805: _{AB}^{pq}h_{A}^{p}h_{B}^{q} \nonumber \\
806: &&+\chi _{BB}\varepsilon _{BB}\frac{n_{B}^{2}}{n^{2}}\frac{\sigma _{B}^{2}}{
807: \sigma _{AB}^{2}}M_{B}^{4}\sum_{p}\sum_{q}\Gamma
808: _{BB}^{pq}h_{B}^{p}h_{B}^{q},
809: \label{gammaexp}
810: \end{eqnarray}
811: $\Gamma _{\alpha \beta }^{pq} \equiv \frac{1}{p!q!}\frac{\partial ^{p+q}G_{\alpha
812: \beta }^{\Gamma }}{\partial s^{p}\partial t^{q}}\rfloor _{s=0,t=0}
813: $,
814: the generating function $G_{\alpha \beta }^{\Gamma }\left( s,t\right) $ being given by:
815: \begin{equation}
816: G_{\alpha \beta }^{\Gamma }\left( s,t\right) =\frac{\left( \frac{\left( 1-t\right) M_{\alpha }}{\left( t+\eta \left(
817: 1-t\right) \right) }+\frac{\left( 1-s\right) M_{\beta }}{\left( s+\eta
818: \left( 1-s\right) \right) }\right) ^{\frac{3}{2}}}{\left( s+\eta \left(
819: 1-s\right) \right) ^{\frac{3}{2}}\left( t+\left( 1-t\right) \eta \right) ^{
820: \frac{3}{2}}M_{\alpha }^{3}M_{\beta }^{3}}
821: \label{genegamma}
822: \end{equation}
823: %Finally, using the form (\ref{expansion}) the aforementioned
824: %kinematic constraints read:
825: %\begin{equation}
826: %\sum\limits_{p=0}^{\infty }O_{p}^{I}h_{A}^{p}=1;\ \ \sum\limits_{p=0}^{\infty }O_{p}^{I}h_{B}^{p}=1; %\ \
827: %\sum\limits_{p=0}^{\infty }O_{p}^{II}\left( \frac{n_{A}}{n}h_{A}^{p}+\frac{
828: %n_{B}}{n}h_{B}^{p}\right) =1
829: %\end{equation}
830: %where the coefficients $O_{p}^{I}$ and $O_{p}^{II}$ are obtained using their
831: %respective generating functions:
832: %\begin{equation}
833: %O^{I}\left( s\right) \equiv \sum\limits_{p=0}^{\infty }s^{p}O_{p}^{I}=\frac{1}{\left( s+\left( 1-s\r%ight) \eta
834: %\right) ^{\frac{3}{2}}}; \
835: %O^{II}\left( s\right) \equiv \sum\limits_{p=0}^{\infty }s^{p}O_{p}^{II}=\frac{\left( 1-s\right) }{\l%eft( s+\left(
836: %1-s\right) \eta \right) ^{\frac{5}{2}}}
837: %\end{equation}
838:
839: %As mentioned, a truncation of the above equations can be solved for any values
840: %of the parameters of the problem, by using standard solvers. Convergence is
841: %tested by checking that an increase in the order of truncation (we have
842: %checked up to the 25th order in the Sonine polynomial expansion in some cases)
843: %does not change the results (i.e., the desired coefficients).
844: With the coupling constants calculated, the algebraic equations, Eq.~(\ref{Bolteqexp}), are now solved to produce the distribution
845: functions, from which one can easily calculate the temperature ratio,
846: $T_A/T_B$.
847: Fig.~\ref{TR_fig}(a) shows a plot of the temperature ratio as a function of
848: $\frac{m_A}{m_A +m_B}$ for a system of particles of equal mass density, $c=0.5$ and $\varepsilon_{AA}=\varepsilon_{BB}=\varepsilon_{AB}=0.8$. Note that the temperature ratio can ``violate'' equipartition quite strongly.
849: In order to compute the heat conductivity and the thermal diffusion coefficient, we now turn to solve the Boltzmann equation, carrying out the Chapman-Enskog expansion to first order in gradients of the hydrodynamic fields, with the homogeneous cooling state distribution function serving as as a zeroth order solution. The distribution functions are expressed as $f_{\alpha }=f_{\alpha
850: }^{\mbox{\tiny{HCS}}}+f_{\alpha }^{K}$, where $f_{\alpha }^{K}$ is the first order
851: perturbation in a gradient expansion, written for convenience as $
852: f_{A}^{K}=f_{\alpha }^{M,\nu }\Phi _{\alpha }^{K},$ cf. Eq.~(\ref{expansion}), where the value of $\nu$ may be taken to be different from that of $\eta$. To first order in the gradients of
853: the hydrodynamic fields, the two coupled Boltzmann equations for the
854: perturbations $\Phi _{\alpha }^{K}$ read:
855: \begin{eqnarray*}
856: D^{K}f_{A} &=&\left( L_{AA}^{(1)}+L_{AA}^{(2)}+L_{AB}^{(1)}\right) \Phi
857: _{A}^{K}+L_{AB}^{(2)}\Phi _{B}^{K} \\
858: D^{K}f_{B} &=&L_{BA}^{(2)}\Phi _{A}^{K}+\left(
859: L_{BB}^{(1)}+L_{BB}^{(2)}+L_{BA}^{(1)}\right) \Phi _{B}^{K}
860: \end{eqnarray*}
861: where
862: \begin{eqnarray*}
863: L_{\alpha \beta }^{(1)}\Phi _{\alpha }^{K} &\equiv &\mathcal{B}\left(
864: f_{\alpha }^{M,\nu }\Phi _{\alpha }^{K},f_{\beta }^{\left( HCS\right)
865: },e_{\alpha \beta }\right) \ \ \ \alpha\ne\beta\\
866: L_{\alpha \beta }^{(2)}\Phi _{\beta }^{K} &\equiv &\mathcal{B}\left(
867: f_{\alpha }^{\left( HCS\right) },f_{\beta }^{M,\nu }\Phi _{\beta
868: }^{K},e_{\alpha \beta }\right) \ \ \ \alpha \ne \beta
869: \end{eqnarray*}
870: are the linearized Boltzmann operators,
871: and $D^{K}f_{\alpha }$ denotes the first order term in the expansion of $Df_{\alpha }$. Following the definition of the operator, $D$:
872: \begin{eqnarray*}
873: Df_{\alpha} &=&f_{\alpha }^{M,\eta }\bigg[ \phi_{\alpha} D\ln n_{\alpha}+2\gamma
874: _{\alpha}u_{i}\left( \eta \phi_{\alpha} -\phi_{\alpha} ^{\prime }\right) DV_{i}\\
875: &+&\left( \gamma
876: _{\alpha}u^{2}\left( \eta \phi_{\alpha} -\phi_{\alpha} ^{\prime }\right) -\frac{3}{2}\phi_{\alpha} \right)
877: D\ln T+c\frac{\partial \phi_{\alpha} }{\partial c}D\ln c\bigg] \\
878: &+&f_{\alpha }^{M,\nu }D\Phi _{\alpha}^{K}+\Phi _{\alpha}^{K}f_{\alpha }^{M,\nu }\left[
879: D\ln n_{\alpha}+2\gamma _{\alpha}u_{i}\nu DV_{i}+\left( \gamma _{\alpha}u^{2}\nu -\frac{3}{2%
880: }\right) D\ln T\right]
881: \end{eqnarray*}
882: where a prime superscript of a function denotes a derivative with respect
883: to its argument.
884: Carrying out the expansion to first order in gradients leads, on the basis
885: of tensorial considerations, to the following form for the function $\Phi
886: _{\alpha}^{K}$:
887: \begin{eqnarray}
888: \Phi _{\alpha}^{K}&=&\Phi _{\alpha}^{K,T}\left( \gamma _{\alpha}u^{2}\right) \sqrt{\gamma _{\alpha}%
889: }u_{j}\frac{\partial \ln T}{\partial x_{j}}+\Phi _{\alpha}^{K,n}\left( \gamma
890: _{\alpha}u\right) \sqrt{\gamma _{\alpha}}u_{j}\frac{\partial \ln n}{\partial x_{j}} \nonumber\\
891: &+&\Phi _{\alpha}^{K,c}\left( \gamma _{\alpha}u\right) \sqrt{\gamma _{\alpha}}u_{j}\frac{%
892: \partial \ln c}{\partial x_{j}}+\Phi _{\alpha}^{K,V}\left( \gamma _{\alpha}u\right)
893: \gamma _{\alpha}^{\frac{3}{2}}\overline{u_{i}u_{j}}\frac{\partial V_{i}}{\partial
894: x_{j}} \label{fiKform}
895: \end{eqnarray}
896: where for any tensor $A_{ij}$, $\overline{A_{ij}}\equiv \frac{1}{2}\left(A_{ij}+A_{ji}-\frac{2}{3}\delta_{ij}A_{kk}\right)$. $D^{K}f_{\alpha }$ is given by:
897: \begin{eqnarray}
898: D^{K}f_{\alpha } &=&\frac{\Gamma }{nT}f_{\alpha }^{M,\nu }\bigg[\left( \gamma
899: _{\alpha }u^{2}\left( \Phi _{\alpha }^{K,c}\right) ^{\prime }-\left( \gamma
900: _{\alpha }u^{2}\nu -2\right) \Phi _{\alpha }^{K,c}-c\frac{\partial \ln
901: \widehat{\Gamma }}{\partial c}\Phi _{\alpha }^{K,T}\right) \nonumber\\
902: & & \times \sqrt{\gamma
903: _{\alpha }}u_{j}\frac{\partial \ln c}{\partial x_{j}} \nonumber\\
904: &&+\left( \gamma _{\alpha }u^{2}\left( \Phi _{\alpha }^{K,n}\right) ^{\prime
905: }-\left( \gamma _{\alpha }u^{2}\nu -2\right) \Phi _{\alpha }^{K,n}-\Phi
906: _{\alpha }^{K,T}\right) \sqrt{\gamma _{\alpha }}u_{j}\frac{\partial \ln n}{%
907: \partial x_{j}}\nonumber \\
908: &&+\left( \gamma _{\alpha }u^{2}\left( \Phi _{\alpha }^{K,T}\right) ^{\prime
909: }-\left( \gamma _{\alpha }u^{2}\nu -2\right) \Phi _{\alpha }^{K,T}-\frac{1}{2%
910: }\Phi _{\alpha }^{K,T}\right) \sqrt{\gamma _{\alpha }}u_{j}\frac{\partial
911: \ln T}{\partial x_{j}}\nonumber \\
912: &&+\left( \gamma _{\alpha }u^{2}\left( \Phi _{\alpha }^{K,V}\right) ^{\prime
913: }-\left( \gamma _{\alpha }u^{2}\nu -3\right) \Phi _{\alpha }^{K,V}\right)
914: \gamma _{\alpha }^{\frac{3}{2}}\overline{u_{i}u_{j}}\frac{\partial V_{i}}{%
915: \partial x_{j}}\bigg] \nonumber\\
916: &&
917: +f_{\alpha }^{M,\eta }\bigg[2\gamma _{\alpha }\left( \eta \phi _{\alpha
918: }-\phi _{\alpha }^{\prime }\right) \overline{u_{i}u_{j}}\frac{\partial V_{i}%
919: }{\partial x_{j}}+\left( \phi _{\alpha }-\frac{nm_{\alpha}}{\rho }\left( \eta \phi _{\alpha
920: }-\phi _{\alpha }^{\prime }\right) \right) u_{i}\frac{\partial \ln n}{%
921: \partial x_{i}} \nonumber\\
922: &&+\left( \left( \gamma _{\alpha }u^{2}-\frac{nm_{\alpha }}{\rho }\right)
923: \left( \eta \phi _{\alpha }-\phi _{\alpha }^{\prime }\right) -\frac{3}{2}%
924: \phi _{\alpha }\right) u_{i}\frac{\partial \ln T}{\partial x_{i}} \nonumber\\
925: &&+f_{\alpha }^{M,\eta }\left( \phi _{\alpha }+c\frac{\partial \phi _{\alpha
926: }}{\partial c}\right) u_{j}\frac{\partial \ln c}{\partial x_{j}}\bigg]
927: \label{DKf}
928: \end{eqnarray}
929: %Employing the form (\ref{fiKform}) into the definitions of the heat and particle fluxes
930: %yields the following expressions for the heat conductivity and thermal diffusion coefficient%s:
931: Employing (\ref{fiKform}) in the definition of $\mathbf{Q}$ and $\mathbf{J}_A$, one obtains:
932: \begin{eqnarray*}
933: \mathbf{\kappa }_{A}^{T} &=&-\frac{1}{3}Tr\left[ \int f_{A}^{M,\nu}\left[
934: \Phi _{A}^{K,T}\left( \gamma _{A}u^{2}\right) \sqrt{\gamma _{A}}u_{j}\right]
935: u_{i}d\mathbf{u}\right] \\
936: \mathbf{\lambda }^{T} &=&-\frac{1}{3}Tr\left[ \int\left( m_{A}f_{A}^{M,\nu}\Phi
937: _{A}^{K,T}\left( \gamma _{A}u^{2}\right) \sqrt{\gamma _{A}} \right. \right.\\
938: && \left. \left. \qquad \qquad \ \ +m_{B} f_{B}^{M,\nu}\Phi _{B}^{K,T}\left( \gamma _{B}u^{2}\right) \sqrt{%
939: \gamma _{B}}\right)\right] u^{2}u_iu_jd\mathbf{u}
940: \end{eqnarray*}
941: where $\Phi _{\alpha }^{K,T}$ obey (with $\beta\ne\alpha$):
942: \begin{eqnarray}
943: &&\left( L_{\alpha \alpha }^{(1)}+L_{\alpha \alpha }^{(2)}+L_{\alpha \beta
944: }^{(1)}\right) \left( \Phi _{\alpha }^{K,T}\left( \gamma _{\alpha }u^{2}\right) \sqrt{\gamma _{\alpha }}\mathbf{u}%
945: \right) +L_{\alpha \beta }^{(2)}\left( \Phi _{\beta }^{K,T}\left( \gamma _{\alpha }u^{2}\right) \sqrt{\gamma
946: _{\beta }}\mathbf{u}\right) \nonumber\\
947: &-&\frac{\sqrt{6\pi }}{9}\frac{n\sigma _{\alpha
948: \beta }^{2}}{\sqrt{m_{0}}}\widetilde{\Gamma }\sqrt{T}\mathbf{H}_{\alpha
949: }^{T}\left( \Phi _{\alpha }^{K,T}\left( \gamma _{\alpha }u^{2}\right) \sqrt{\gamma _{\alpha }}\mathbf{u}\right) =%
950: \mathbf{R}_{\alpha }^{T}
951: \label{EqfiKT}
952: \end{eqnarray}
953:
954: where\ $\mathbf{H}_{\alpha}^{T}\left( \Phi _{\alpha }^{K,T}\sqrt{\gamma _{\alpha }}
955: \,\mathbf{u}\right) =f_{\alpha }^{M,\nu }\left( \gamma _{\alpha}u^{2}\left( \Phi
956: _{\alpha}^{K,T}\right) ^{\prime }-\left( \gamma _{\alpha}u^{2}\nu -\frac{3}{2}\right)
957: \Phi _{\alpha}^{K,T}\right) \sqrt{\gamma _{\alpha}}\,\mathbf{u}\,$ and $\mathbf{R}
958: _{\alpha}^{T}=f_{\alpha }^{M,\nu }\left( \left( \gamma _{\alpha}u^{2}-\frac{nm_{\alpha}}{
959: \rho }\right) \left( \eta \phi _{\alpha}-\phi _{\alpha}^{\prime }\right) -\frac{3}{2}
960: \phi _{\alpha}\right) \mathbf{u}$. The equations (\ref{EqfiKT}) are solved subject to the constraint $\int f_{A}^Km_{A}\mathbf{u}d\mathbf{u+}\int f_{B}^Km_{B}\mathbf{u}d\mathbf{u}=0$ using the same computer aided method applied to the solution of the HCS. The functions $%
961: \Phi _{\alpha }^{K,T}$ are represented by series of Sonine
962: polynomials $\Phi _{\alpha }^{K,T}=\sum_{p=0}\frac{\widehat{k}_{\alpha
963: }^{K,T,p}}{n\sigma _{AB}^{2}}S_{\frac{3}{2}}^{p}\left( \gamma _{\alpha
964: }u^{2}\right) $ (the factor $\nu <1$ in the function $f_{\alpha }^{M,\nu }$
965: ensuring the convergence of the series), which is substituted in (\ref{EqfiKT}).
966: \begin{figure}
967: \begin{center}
968: \begin{tabular}{ccc}
969: \subfigure[]{\includegraphics[height=0.25\hsize]{ptt_fig1.eps}}&
970: \subfigure[]{\includegraphics[height=0.25\hsize]{ptt_fig2.eps}}&
971: \subfigure[]{\includegraphics[height=0.25\hsize]{ptt_fig3.eps}}
972: \end{tabular}
973: \end{center}
974: \caption{Plot of the temperature ratio $T_A/T_B$ (a), thermal diffusion coefficient $\kappa_A^{T}$ (b), and thermal
975: conductivity $\lambda^{T}$ (c), as a
976: function of $\frac{m_A}{m_A+m_B}$ for grains made of the same material, with
977: $\frac{n_A}{n}=\frac{n_B}{n}=0.5$ and
978: $\varepsilon_{AA}=\varepsilon_{AB}=\varepsilon_{BB}=0.8$}
979: \label{TR_fig}
980: \end{figure}
981: % to yield:
982: %\begin{eqnarray}
983: %&\sum_{q}&\bigg[\left( L_{\alpha \alpha }^{(1)}+L_{\alpha \alpha }^{(2)}+L_{\alpha \beta
984: %}^{(1)}\right) \left( S_{\frac{3}{2}}^{q}\left( \gamma _{\alpha }u^{2}\right) \sqrt{\gamma _{\alpha }%}\mathbf{u}%
985: %\right) +L_{\alpha \beta }^{(2)}\left( S_{\frac{3}{2}}^{q}\left( \gamma _{\alpha }u^{2}\right)\sqrt{\%gamma
986: %_{\beta }}\mathbf{u}\right) \nonumber\\
987: %&-&\frac{\sqrt{6\pi }}{9}\frac{n\sigma _{\alpha
988: %\beta }^{2}}{\sqrt{m_{0}}}\widetilde{\Gamma }\sqrt{T}\mathbf{H}_{\alpha
989: %}^{T}\left(S_{\frac{3}{2}}^{q}\left( \gamma _{\alpha }u^{2}\right)\sqrt{\gamma _{\alpha }}\mathbf{u}\%right)\bigg] =%
990: %\mathbf{R}_{\alpha }^{T}
991: %\end{eqnarray}
992: The resulting equation is then projected on $S_{\frac{3}{2}}^{N}\left( \gamma _{\alpha }u^{2}\right)
993: \sqrt{\gamma _{\alpha }}\mathbf{u}$ to yield a linear system of equations for the
994: coefficients $\widehat{k}_{\alpha }^{K,T,p}$. The system of equations
995: is then solved (after evaluating the coupling constants using the
996: generating function method explained above). The resulting perturbations
997: are then used to calculate the transport coefficients. Results
998: for the thermal diffusion constant, $\kappa_A^{T}$, and the
999: heat conductivity, $\lambda^{T}$, are presented in Fig.~\ref{TR_fig}(b,c).
1000: Note the nontrivial dependence of these transport coefficients on
1001: $\frac{m_A}{m_A+m_B}$. Although clearly equipartition is strongly broken, these
1002: transport coefficients are well behaved (as are those not shown here).
1003: \section{Elastic energy and ``heat flux''}
1004:
1005: \label{sec:elasticity_derivation}
1006:
1007: Consider a set of particles whose binary interaction potential is harmonic (or
1008: approximated by): $
1009: U\left[\vec{r}_{ij}(t)\right]=\frac{1}{2}K_{ij}\left(|\vec{r}_{ij}|-l_{ij}\right)^{2}$,
1010: where $l_{ij}$ is the equilibrium separation of particles $i$ and $j$ and ${\bf
1011: r}_{ij} \equiv {\bf r}_i-{\bf r}_j$. The force on particle $i$ exerted by
1012: particle $j$ is given by $ \vec{f}_{i/j} =
1013: -\mbox{\boldmath$\nabla$}_{i}U\left(\vec{r}_{ij}\right) =
1014: -K_{ij}\left(|\vec{r}_{ij}|-l_{ij}\right)\hat{\vec{r}}_{ij}$, where
1015: $\hat{\vec{r}}_{ij} \equiv \vec{r}_{ij} / \left| \vec{r}_{ij} \right|$ is a
1016: unit vector. Consider, for simplicity, an unstressed reference configuration,
1017: in which all particle pairs are at their equilibrium separation
1018: ($\left|\vec{r}_{ij}^{0}\right|=l_{ij}$), where the superscript~$0$ denotes the
1019: reference configuration. The relative displacement of particles $i$ and $j$ is
1020: defined by: ${\bf u}_{ij} \equiv {\bf r}_{ij}-{\bf r}_{ij}^0$. To linear order
1021: in the displacements the corresponding force can be approximated by:
1022: \begin{eqnarray}
1023: \vec{f}_{i/j} & = & -K_{ij}\left(|\vec{r}_{ij}|-l_{ij}\right)\hat{\vec{r}}_{ij}=-K_{ij}\left(|\vec{r}_{ij}^{0}+\vec{u}_{ij}|-\left|\vec{r}_{ij}^{0}\right|\right)\hat{\vec{r}}_{ij}\nonumber \\
1024: & = &
1025: -K_{ij}\left(\hat{\vec{r}}_{ij}^{0}\cdot\vec{u}_{ij}\right)\hat{\vec{r}}_{ij}^{0}.\label{eq:harmonic_force_linear}\end{eqnarray}
1026:
1027: For atomic systems, the harmonic approximation is obtained from a Taylor
1028: expansion of the (effective) interatomic potential around its minimum. For
1029: granular materials, the contact interactions among the particles are typically
1030: described using contact mechanics~\cite{Gladwell80,Johnson85} (force models
1031: used in simulations of granular materials are reviewed
1032: in~\cite{Sadd93,Walton95,Schafer96,Wolf96,Herrmann98}). For the present
1033: discussion, it is assumed that the interparticle forces can be linearized for
1034: small deformations with respect to a reference
1035: configuration.
1036:
1037:
1038:
1039: Consider the contact stress, see Eq.~(\ref{contactdef}). Upon substituting
1040: the formula for the force to linear order in the displacements, Eq.~(\ref{eq:harmonic_force_linear}),
1041: in Eq.~(\ref{contactdef}) one obtains an expression for the linear elastic
1042: stress tensor:
1043: \begin{eqnarray} \sigma_{\alpha\beta}^{\mathrm{lin}}(\vec{r},t) & = &
1044: \frac{1}{2}\sum_{ij}K_{ij}\hat{r}_{ij\alpha}^{0}r_{ij\beta}^{0}\hat{r}_{ij\gamma}^{0}u_{ij\gamma}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}].\label{eq:lin_stress}
1045: \end{eqnarray}
1046: It is interesting to examine the potential energy density as read off
1047: Eq.~(\ref{e1}) and compare it to the relation used in continuum linear
1048: elasticity. In the quasistatic limit, the energy density reduces to the
1049: potential energy density {[}in this limit, the time of deformation goes to
1050: infinity ($t_f\to\infty$) and the velocities go to zero ($\vec{v}_{i}\to0$), so
1051: that the kinetic energy is zero while the displacement
1052: $\vec{u}_{i}\sim\vec{v}_{i} t_f$, is kept finite{]}. To lowest nonvanishing
1053: order in the strain, the potential energy is given by
1054: $\frac{1}{2}K_{ij}\left(\hat{\vec{r}}_{ij}^{0}\cdot\vec{u}_{ij}\right)^{2}$.
1055: Hence, at this order:
1056: \begin{equation}
1057: e^{\mathrm{el}}(\vec{r},t)=\frac{1}{4}\sum_{i,j}K_{ij}\left(\hat{\vec{r}}_{ij}^{0}\cdot\vec{u}_{ij}\right)^{2}\phi[\vec{r}-\vec{r}_{i}(t)].\label{eq:lin_elasticenergydensity}\end{equation}
1058: Classical linear elasticity identifies the elastic energy
1059: as $e^{\mathcal{\mathrm{el}}}=\frac{1}{2}\mbox{\boldmath$\sigma$}^{\mathrm{lin}}:\mbox{\boldmath$\epsilon$}^{\mathrm{lin}}$, where $\epsilon^{\mathrm{lin}}$ is defined in
1060: Eq.~(\ref{epslin}).
1061: Since
1062: the stress is symmetric for central forces (it is easily verified that
1063: Eq.~(\ref{eq:lin_stress}) is invariant to the exchange of $\alpha$
1064: and $\beta$), we can calculate
1065: $\frac{1}{2}\sigma_{\alpha\beta}^{\mathrm{lin}}\frac{\partial
1066: u_{\alpha}^{\mathrm{lin}}}{\partial r_{\beta}}=\frac{1}{2}\frac{\partial}{\partial
1067: r_{\beta}}
1068: \left(\sigma_{\alpha\beta}^{\mathrm{lin}}u_{\alpha}^{\mathrm{lin}}\right) $
1069: instead,
1070: the second equality following from the static equilibrium relation
1071: $\frac{\partial \sigma_{\alpha\beta}^{\mathrm{lin}}}{\partial r_\beta}=0$,
1072: see also Eq.~(\ref{second}). It follows, using
1073: Eq.~(\ref{contactdef}) and definition (\ref{linearu}) that:
1074: \begin{displaymath}
1075: \frac{1}{2} \sigma_{\alpha\beta}^{\mathrm{lin}} \epsilon^{\mathrm{lin}}_{\alpha\beta}
1076: = -\frac{1}{4}\frac{\partial}{\partial r_{\beta}}\sum_{ij}u_{\alpha}^{\mathrm{lin}} f_{i/j\alpha} r_{ij\beta}^{0}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1077: \end{displaymath}
1078: Next, using Eq.~(\ref{flu}) one obtains:
1079: %\end{document}
1080: \begin{displaymath}
1081: \frac{1}{2} \sigma_{\alpha\beta}^{\mathrm{lin}} \epsilon^{\mathrm{lin}}_{\alpha\beta} = -\frac{1}{4}\frac{\partial}{\partial r_{\beta}}\sum_{ij}\left[u_{i\alpha}- u'_{i\alpha}\right]
1082: f_{i/j\alpha} r_{ij\beta}^{0}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1083: \end{displaymath}
1084: \begin{displaymath}
1085: = -\frac{1}{4}\sum_{ij}u_{i\alpha} f_{i/j\alpha} r_{ij\beta}^{0}
1086: \frac{\partial}{\partial r_{\beta}}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1087: \end{displaymath}
1088: \begin{displaymath}
1089: +\frac{1}{4}\frac{\partial}{\partial r_{\beta}}\sum_{ij}u'_{i\alpha}f_{i/j\alpha}r_{ij\beta}^{0}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1090: \end{displaymath}
1091: \begin{displaymath}
1092: = \frac{1}{4}
1093: \sum_{ij}u_{i\alpha}f_{ij\alpha}
1094: \left(\phi\left[ {\bf r}-{\bf r}_i(t) \right]
1095: -\phi \left[ {\bf r}-{\bf r}_j(t) \right]
1096: \right)
1097: \end{displaymath}
1098: \begin{displaymath}
1099: +\frac{1}{4}\frac{\partial}{\partial r_{\beta}}\sum_{ij}u'_{i\alpha}f_{i/j\alpha}r_{ij\beta}^{0}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1100: \end{displaymath}
1101: In equilibrium $\sum_j f_{i/j\alpha}=0$, hence
1102: $\sum_j f_{i/j\alpha}\phi\left( {\bf r}-{\bf r}_i(t) \right)=0$ and it can be
1103: replaced by $-\sum_j f_{i/j\alpha}\phi\left( {\bf r}-{\bf r}_i(t) \right)$.
1104: Therefore
1105: \begin{displaymath}
1106: \frac{1}{2} \sigma_{\alpha\beta}^{\mathrm{lin}} \epsilon^{\mathrm{lin}}_{\alpha\beta}=
1107: -\frac{1}{4}
1108: \sum_{ij}u_{i\alpha}f_{ij\alpha}
1109: \left(\phi\left[ {\bf r}-{\bf r}_i(t) \right]
1110: +\phi \left[ {\bf r}-{\bf r}_j(t) \right]
1111: \right)
1112: \end{displaymath}
1113: \begin{displaymath}
1114: +\frac{1}{4}\frac{\partial}{\partial r_{\beta}}\sum_{ij}u'_{i\alpha}f_{i/j\alpha}r_{ij\beta}^{0}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1115: \end{displaymath}
1116: \begin{displaymath}
1117: = -\frac{1}{4}\sum_{ij}u_{ij\alpha}f_{i/j\alpha}\phi \left[ {\bf r}-{\bf r}_i(t) \right]
1118: +\frac{1}{4}\frac{\partial}{\partial r_{\beta}}\sum_{ij}u'_{i\alpha}f_{i/j\alpha}r_{ij\beta}^{0}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1119: \end{displaymath}
1120: Next, substituting the expression for the force, Eq.~(\ref{eq:harmonic_force_linear}), one obtains:
1121: \begin{displaymath}
1122: \frac{1}{2} \sigma_{\alpha\beta}^{\mathrm{lin}} \epsilon^{\mathrm{lin}}_{\alpha\beta} = \frac{1}{4}\sum_{ij}K_{ij}\left(\hat{\vec{r}}_{ij}^{0}\cdot\vec{u}_{ij}\right)^{2}
1123: \phi \left[ {\bf r}-{\bf r}_i(t) \right]
1124: \end{displaymath}
1125: \begin{displaymath}
1126: +\frac{1}{4}\frac{\partial}{\partial r_{\beta}}\sum_{ij}f_{i/j\alpha}u'_{i\alpha}r_{ij\beta}^{0}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1127: \end{displaymath}
1128: \begin{displaymath}
1129: = e^{\mathrm{{el}}}(\vec{r},t)+\frac{1}{4}\frac{\partial}{\partial r_{\beta}}\sum_{ij}f_{i/j\alpha}u'_{i\alpha}r_{ij\beta}^{0}\int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}]
1130: \end{displaymath}
1131: \begin{equation}
1132: = e^{\mathrm{{el}}}(\vec{r},t)+\frac{1}{8}\frac{\partial}{\partial r_{\beta}}\sum_{ij}f_{i/j\alpha}\left( u'_{i\alpha}
1133: + u'_{j\alpha}\right) r_{ij\beta}^{0}
1134: \int_{0}^{1}ds\,\phi[\vec{r}-\vec{r}_{i}^{0}+s\vec{r}_{ij}^{0}] \label{eee}
1135: \end{equation}
1136: This result is identical to that obtained on the basis of dynamical considerations in Appendix B, cf. Eqs.~(\ref{tiny},\ref{last}).
1137:
1138:
1139: It follows that the coarse grained elastic energy density is not precisely
1140: given by the classical expression
1141: $\frac{1}{2}\mbox{\boldmath$\sigma$}^{\mathrm{lin}}:\mbox{\boldmath$\epsilon$}^{\mathrm{lin}}$; the above
1142: additional term provides a correction to the classical expression. The
1143: correction represents the adiabatic limit of the divergence of the (time
1144: integral of) the heat flux or, in physical terms, the divergence of the
1145: fluctuating part of the work of the interparticle forces, i.e. the work done on
1146: the fluctuating parts of the displacements or the work which is unresolved by
1147: macroscopic elasticity. As this term is a divergence of a flux, its average
1148: over a volume, $\Omega$, whose linear dimension is $W$, is proportional to
1149: $W^{D-1}$, while the average over the first term is proportional to $W^{D}$.
1150: Therefore their ratio tends to zero as $1/W$, i.e., the \emph{average} of
1151: $\frac{1}{2}\mbox{\boldmath$\sigma$}^{\mathrm{lin}}:\mbox{\boldmath$\epsilon$}^{\mathrm{lin}}$ over a sufficiently
1152: large volume (not its `local value') begets the standard elastic energy
1153: density. It should be emphasized that both contributions to the elastic energy
1154: density are of the same (second) order in the displacements (or strain), i.e.
1155: the additional term is not a ``higher order'' correction but rather a
1156: correction that appears when the deformation is not affine (i.e., when ${\bf
1157: u}^\prime_{i} \ne 0$).
1158: \subsection{Numerical results: 2D disordered harmonic networks}
1159:
1160: In order to examine the correction to the classical elastic energy numerically,
1161: we use disordered harmonic networks in 2D, obtained as follows: ``particles''
1162: (of equal mass) are placed at the nodes of a triangular lattice, of lattice
1163: constant $d$. The system is square shaped. Nearest neighbors are connected by
1164: springs (with rest lengths equal to the lattice constant, ensuring a
1165: stress-free reference state). We introduce disorder in two ways: first, the
1166: spring network is diluted by randomly removing a fraction $c$ of the springs.
1167: Second, we employ spring constants which are uniformly distributed in the range
1168: $[K-\delta K,K+\delta K]$. The particle positions are the same as in the
1169: triangular lattice, hence the density remains uniform. We use periodic boundary
1170: conditions in both axes, and apply a specified global strain to the system by
1171: imposing a change in the period as well as (for shear deformation) Lees-Edwards
1172: boundary conditions.
1173:
1174: The (linear) static equilibrium equations are solved (by inversion of the
1175: dynamical matrix) for a given applied global strain,
1176: $\epsilon^{\mathrm{app}}_{\alpha\beta}$, yielding a set of corresponding
1177: displacements $\left\{ \vec{u}_{i}\right\} $. The latter are used for
1178: calculating the CG linear strain field, $\epsilon^{\mathrm{lin}}$ [using
1179: Eq.~(\ref{epslin})], the CG linear stress, $\sigma^{\mathrm{lin}}$
1180: [Eq.~(\ref{eq:lin_stress}], the CG potential energy density, $e^{\mathrm{el}}$
1181: [Eq.~(\ref{eq:lin_elasticenergydensity})], and the fluctuating energy flux,
1182: $\frac{1}{2}\vec{Q}^{\mathrm{{force,rev,I}}}$ [Eq.~(\ref{last})]. The coarse
1183: graining function used~\cite{Goldhirsch02} is a 2D Gaussian,
1184: $\phi(\vec{r})=\frac{1}{\pi w^{2}}e^{-(|\vec{r}|/w)^{2}}$.
1185:
1186: We verified that for a lattice with uniform spring constants (i.e., $c=\delta
1187: K=0$), the strain components are uniform and equal to the applied global strain
1188: for $w\gtrsim d$, as expected, since in for this system, under homogeneous
1189: deformation, the particle displacements are affine. The resulting
1190: stress components are uniform as well, and are consistent with the continuum
1191: isotropic elastic moduli for a triangular lattice (Lam\'{e} constants
1192: $\lambda=\mu=\frac{\sqrt{3}K}{4}$). As expected, in this case the CG energy
1193: density conforms to the classical expression,
1194: $e^{\mathrm{el}}=\frac{1}{2}\mbox{\boldmath$\sigma$}^{\mathrm{lin}}:\mbox{\boldmath$\epsilon$}^{\mathrm{lin}}$.
1195:
1196: For disordered systems, even under a ``homogeneous applied strain'' as
1197: described above, there is a correction to the classical expression for the
1198: energy density, as detailed above. Fig.~\ref{fig:dErel} presents a contour plot
1199: of the relative energy difference $\Delta \tilde{e}\equiv 2
1200: \frac{e^{\mathrm{el}}}{\mbox{\boldmath$\sigma$}^{\mathrm{lin}}:\mbox{\boldmath$\epsilon$}^{\mathrm{lin}}}-1=-
1201: \frac{\mathrm{div}\, {\bf
1202: Q}^{\mathrm{{force,rev,I}}}}{\mbox{\boldmath$\sigma$}^{\mathrm{lin}}:\mbox{\boldmath$\epsilon$}^{\mathrm{lin}}}$,
1203: for a system of $30\times 30$ particles, with $c=2\%$ and $\delta K/K=25\%$
1204: subject to an applied strain $\epsilon^{\mathrm{app}}_{xx}=-\epsilon^{\mathrm
1205: app}_{yy}=5\cdot10^{-3}, \epsilon^{\mathrm{app}}_{xy}=0$, calculated with a
1206: CG width, $w=4d$. The deviation from the classical expression is non-negligible
1207: ($\Delta \tilde{e}\simeq 1\%$), even thought the present system has a relatively
1208: small disorder [we verified that in the same system, a linear stress-strain
1209: relation holds to within $4\%$~\cite{Goldhirsch02}]. It is interesting to try
1210: to relate the fluctuating energy flux to a local measure of the disorder in the
1211: particle displacements. A possible field for characterizing the fluctuating
1212: displacements is the ``noise'' field~\cite{Goldenberg06c}, $\eta({\bf r})\equiv
1213: \sum_i m_i |\vec{u}'_i(\vec{r})|^2 \phi[\vec{r}-\vec{r}_i(t)]$, defined in an
1214: analogous way to the kinetic temperature. In Fig.~\ref{fig:Q_gradN} we compare
1215: the fluctuating energy flux with the gradient of the noise field. It appears
1216: that their magnitude (but not their directions) exhibits quite a strong
1217: correlation, suggesting that the fluctuating energy flux may indeed be related
1218: to gradients of the noise, but the corresponding ``conductivity'' may be
1219: anisotropic. This is to be expected, since the system is locally anisotropic
1220: [as the local elastic tensor has been checked to be
1221: anisotropic~\cite{Goldhirsch02}]. It is possible that a tensorial, rather than
1222: scalar, characterization of the displacement fluctuations is required,
1223: analogous to an ``anisotropic temperature''.
1224: %
1225: \begin{figure}
1226: \begin{center}\begin{tabular}{cc}
1227: \subfigure[]{\includegraphics[%
1228: height=0.42\hsize,clip]{ptt_fig4.eps}\label{fig:dErel}}&
1229: \subfigure[]{\includegraphics[%
1230: height=0.39\hsize,clip]{ptt_fig5.eps}\label{fig:Q_gradN}}\tabularnewline
1231: \end{tabular}\end{center}
1232: \caption{(a) A contour plot of the relative difference between the elastic
1233: energy density, $e^{\mathrm{el}}$, and its ``classical'' value,
1234: $\frac{1}{2}\mbox{\boldmath$\sigma$}^{\mathrm{lin}}:\mbox{\boldmath$\epsilon$}^{\mathrm{lin}}$, in a disordered
1235: harmonic network (see text). (b) The fluctuating energy flux,
1236: $\frac{1}{2}\vec{Q}^{\mathrm{{force,rev,I}}}$ (black) and the (negative)
1237: gradient of the ``noise'', $-\mbox{\boldmath$\nabla$} \eta$ (gray), in the same system (arbitrary
1238: scaling).}
1239: \label{fig:dErel_Q_gradN}
1240: \end{figure}
1241: \section{Conclusion}
1242: We have shown that the classical definition of granular temperature as a
1243: measure of the fluctuating kinetic energy is useful for the description of
1244: granular gases even when this temperature is not the (inverse of the)
1245: thermodynamic conjugate potential of the energy, as in (local) equilibrium. An
1246: analysis of the case of a near-elastic granular gas shows that indeed one can
1247: venture into the nonequilibrium domain starting from local equilibrium (for the
1248: near elastic case). The example of a binary granular gas mixture in which the
1249: temperatures of the components can be very different from each other and still
1250: the Chapman-Enskog method produces sensible results, is a compelling case in
1251: favor of the use of the granular temperature field for far-from-equilibrium
1252: systems, at least in the fluidized phase. In the `opposite' limit of a
1253: granular solid we have shown that the heat flux (which is always present, even
1254: in athermal system, as shown in Appendix A) is not `lost' when one performs a
1255: deformation of an elastic solid: its effect appears in the form of a correction
1256: to the classical elastic energy density. This correction is strongly correlated
1257: with the gradient of an `elastic temperature' which is a measure of the degree
1258: of non-affinity induced by the randomness of the system. Note that the system
1259: considered is not pre-stressed whereas many granular systems are. This fact
1260: does not change the fundamental conclusion that in elastic solids there is an
1261: accumulation of energy as a result of work done on the non-affine (or
1262: fluctuating) degrees of freedom. As the physics of this finding is very general
1263: we believe it applies (with a proper modification) to granular solids beyond
1264: the elastic regime. Indeed a strong correlation of the displacement `noise'
1265: with the appearance of localized plastic events has been suggested in
1266: \cite{Goldenberg06c}. Another connection of this kind has been proposed in
1267: \cite{Lemaitre02}. At this stage we do not know whether there is a direct link
1268: between this `new' contribution that stems from the heat flux and (gradients
1269: of) `configurational temperatures'. However, since there is an obvious link
1270: between diffusion and non-affinity (see e.g., \cite{Utter07}) and a seemingly
1271: established link between diffusion and a `configurational temperature' (through
1272: an FD relation) it comes to reason that the `extra' term in the elastic energy
1273: is related to a configurational measure of disorder. Following the spirit of a
1274: suggestion by Savage \cite{Savage98} one may perhaps conjecture that an
1275: effective temperature which combines the configurational and kinetic
1276: contributions (perhaps just as a sum) can serve as a generalized granular
1277: temperature valid for all granular phases (with appropriate extensions to other
1278: systems). At this stage it seems that much more research is necessary to
1279: establish whether this or similar conjectures are of any physical value. As a
1280: final comment we would like to mention the rather well known fact that (even
1281: relatively simple) non-equilibrium systems exhibit rich behavior and cannot
1282: usually be characterized by the same number of macroscopic variables or fields
1283: as (near) equilibrium systems. Therefore, even if it turns out that many
1284: properties of granular matter can be correlated with a generalized temperature
1285: or temperatures there is a good chance that more characterizations will be
1286: needed for this class of systems to fully describe or explain their behavior.
1287:
1288: \appendix
1289: \section{Continuum Mechanics: a microscopic derivation}
1290:
1291: The equations of continuum mechanics are derived below from
1292: microscopic considerations.
1293: For sake of simplicity it is assumed that the constituents experience
1294: binary interactions whose reversible part is determined by a potential. The
1295: derivation below is a slight extension of what can be found
1296: in \cite{Glasser01,Babic97} but unlike in \cite{Glasser01} we do not invoke
1297: temporal coarse graining, for simplicity.
1298:
1299: Consider a set of particles of mass $m_i$, whose (center of mass)
1300: positions and velocities are
1301: ${\bf r}_i(t)$ and ${\bf v}_i(t) \equiv \dot{\bf r}_i(t)$,
1302: respectively. Let the binary interaction potential be denoted by
1303: $U({\bf r}_{ij})$, where ${\bf r}_{ij} \equiv {\bf r}_i - {\bf r}_j$. It
1304: is assumed that $U=0$ for $i=j$. In
1305: addition, let $\phi({\bf r})$ denote a spatial coarse graining function, i.e.
1306: $\phi$ is normalized (its integral over space equals unity), it is
1307: positive semidefinite and has a single maximum at ${\bf r}=0$. It is often
1308: convenient to employ a Gaussian distribution for $\phi$. The width
1309: of $\phi$ is the ``coarse graining scale''.
1310:
1311: The coarse-grained (CG) mass density, $\rho({\bf r},t)$, at position ${\bf r}$
1312: and time $t$ is given by: $ \rho({\bf r},t) \equiv \sum_i m_i \phi\left[{\bf
1313: r}-{\bf r}_i(t)\right] $. Note that while ${\bf r}_i(t)$ is the position
1314: of particle $i$ at time $t$, ${\bf r}$ is merely a considered position in
1315: space. Similarly, define the CG momentum density, ${\bf p}({\bf r},t)$, by: $
1316: {\bf p}({\bf r},t) \equiv \sum_i m_i {\bf v}_i (t) \phi\left[{\bf r}-{\bf
1317: r}_i(t)\right] $. The CG energy density, $e({\bf r},t)$ is given by:
1318: \begin{equation}
1319: e({\bf r},t) \equiv \sum_i \frac{1}{2} m_i v^2_i \phi\left[ {\bf r}-{\bf r}_i(t)\right]+ \frac{1}{2} \sum_{i,j} U({\bf r}_{ij}(t))\phi\left[ {\bf r}-{\bf r}_i(t)\right] \label{e1}
1320: \end{equation}
1321: By taking the time derivative of the density one obtains:
1322: \begin{equation}
1323: \dot{\rho}({\bf r},t)= -\frac{\partial}{\partial r_\alpha}
1324: \sum_i m_i v_{i\alpha} (t) \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1325: = -\mathrm{div}\, {\bf p}({\bf r},t)\label{rhodot1}
1326: \end{equation}
1327: where the summation convention has been invoked and Greek letters denote
1328: Cartesian coordinates. Defining the velocity field, ${\bf V}({\bf r},t)$,
1329: as: $
1330: %\begin{equation}
1331: {\bf V}({\bf r},t) \equiv \frac{{\bf p}({\bf r},t)}{\rho({\bf r},t)} \label{veldef}$,
1332: %\end{equation}
1333: one obtains from Eq.~(\ref{rhodot1}) the equation of continuity:
1334: $\dot{\rho} =-\mathrm{div}\, (\rho {\bf V})$.
1335: Next, upon taking the derivative of the momentum field one obtains:
1336: \begin{equation}
1337: \dot{ p}_\alpha ({\bf r},t) = -\frac{\partial}{\partial r_\beta}
1338: \sum_i m_i v_{i\alpha}(t) v_{i\beta} (t) \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1339: + \sum_i m_i \dot{v}_{i\alpha} \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1340: \label{p1}
1341: \end{equation}
1342: Denote by ${\bf f}_i(t)$ the force experienced by particle $i$ and by
1343: ${\bf f}_{i/j}(t)$ the force exerted by particle $j$ on $i$.
1344: Using Newton's second and third laws one obtains:
1345: %\begin{displaymath}
1346: $
1347: \sum_i m_i \dot{ v}_{i\alpha}(t) \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1348: =\sum_{i,j} f_{i/j\alpha}(t) \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1349: %\end{displaymath}
1350: %\begin{displaymath}
1351: =\sum_{i,j} f_{j/i\alpha}(t) \phi\left[ {\bf r}-{\bf r}_j(t)\right]
1352: = -\sum_{i,j} f_{i/j\alpha}(t) \phi\left[ {\bf r}-{\bf r}_j(t)\right]
1353: %\end{displaymath}
1354: $.
1355: It follows that:
1356: \begin{displaymath}
1357: \sum_i m_i \dot{ v}_{i\alpha}(t) \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1358: =
1359: \end{displaymath}
1360: \begin{displaymath}
1361: =\frac{1}{2}\sum_{ij} f_{i/j\alpha}(t) \left( \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1362: -\phi\left[ {\bf r}-{\bf r}_j(t)\right]
1363: \right)
1364: \end{displaymath}
1365: \begin{displaymath}
1366: = -\frac{1}{2} \sum_{ij} f_{i/j\alpha} (t) \int_0^1 ds \frac{\partial}{\partial s}
1367: \phi\left[ {\bf r}- {\bf r}_i (t) + s{\bf r}_{ij}(t) \right]
1368: \end{displaymath}
1369: \begin{equation}
1370: =-\frac{\partial}{\partial r_\beta}
1371: \frac{1}{2} \sum_i m_i f_{i/j\alpha} r_{ij\beta}(t) \int_0^1
1372: ds \phi\left[ {\bf r}-{\bf r}_i(t)+s {\bf r}_{ij}(t)\right]
1373: \end{equation}
1374: Substituting the latter result in Eq.~(\ref{p1}) one obtains:
1375: %\begin{equation}
1376: $
1377: \dot{p}_\alpha({\bf r},t) = -\frac{\partial}{\partial r_\beta} P_{\alpha\beta}
1378: ({\bf r},t) \label{p2}
1379: %\end{equation}
1380: $,
1381: where the pressure tensor, $\mathbf P$, is given by:
1382: \begin{displaymath}
1383: P_{\alpha\beta}({\bf r},t) =
1384: \sum_i m_i v_{i\alpha}(t) v_{i\beta}(t)
1385: \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1386: \end{displaymath}
1387: \begin{equation}
1388: + \frac{1}{2}\sum_{ij} f_{i/j\alpha}({\bf r},t) r_{ij\beta}({\bf r},t)
1389: \int_0^1 ds \phi\left[ {\bf r}-{\bf r}_i(t)+ s{\bf r}_{ij}(t) \right] \label{prev}
1390: \end{equation}
1391: Next, define the fluctuating part of the velocity of a particle (when
1392: coarse graining is around the position ${\bf r}$ at time $t$) as:
1393: ${\bf v}^\prime_{i}({\bf r},t) \equiv {\bf v}_i(t) - {\bf V}({\bf r},t)$.
1394: Substituting this definition in Eq.~(\ref{prev}) and performing straightforward algebra (using $\sum_i v^\prime_i({\bf r},t) \phi\left(
1395: {\bf r}-{\bf r}_i(t)\right) =0 $, following the definitions of
1396: ${\bf v}^\prime$ and the velocity field ${\bf V}$), one obtains:
1397: \begin{displaymath}
1398: P_{\alpha\beta}({\bf r},t) =
1399: \rho({\bf r},t) V_\alpha({\bf r},t) V_\beta ({\bf r},t)
1400: +\sum_{i} m_i v^\prime_{i\alpha} ({\bf r},t) v^\prime_{i\beta} ({\bf r},t)
1401: \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1402: \end{displaymath}
1403: \begin{equation}
1404: +\frac{1}{2} \sum_{ij} f_{i/j\alpha}({\bf r},t) r_{ij\beta}({\bf r},t)
1405: \int_0^1 ds \phi\left[ {\bf r}-{\bf r}_i(t)+ s {\bf r}_{ij} (t) \right]
1406: \end{equation}
1407: Identifying
1408: \begin{equation}
1409: \sigma^{\mathrm{{kin}}}_{\alpha\beta}({\bf r},t) = -\sum_{i}m_i v^\prime_{i\alpha} ({\bf r},t) v^\prime_{i\beta} ({\bf r},t)
1410: \phi\left[ {\bf r}-{\bf r}_i(t)\right] \label{kineticstress1}
1411: \end{equation}
1412: as the
1413: ``kinetic stress'' and
1414: \begin{equation}
1415: \sigma^{\mathrm{{cont}}}_{\alpha\beta}({\bf r},t) =
1416: -\frac{1}{2} \sum_{ij} f_{i/j\alpha}({\bf r},t) r_{ij\beta}({\bf r},t)
1417: \int_0^1 ds \phi\left[ {\bf r}-{\bf r}_i(t) +s {\bf r}_{ij} (t) \right]
1418: \label{contactdef}
1419: \end{equation}
1420: as the ``contact stress'', the stress
1421: tensor, $\mbox{\boldmath $\sigma$}$, is given by:
1422: \begin{displaymath}
1423: \sigma_{\alpha\beta}({\bf r},t) =
1424: -\sum_{i}m_i v^\prime_{i\alpha} ({\bf r},t) v^\prime_{i\beta} ({\bf r},t)
1425: \phi\left[ {\bf r}-{\bf r}_i(t) \right]
1426: \end{displaymath}
1427: \begin{equation}
1428: -\frac{1}{2} \sum_{ij} f_{i/j\alpha}({\bf r},t) r_{ij\beta}({\bf r},t)
1429: \int_0^1 ds \phi\left[ {\bf r}-{\bf r}_i(t)+s {\bf r}_{ij}(t)\right]
1430: \label{stress}
1431: \end{equation}
1432: Note that an opposite convention for the sign of the stress tensor is
1433: often employed in the field of granular matter.
1434: It follows from Eq.~(\ref{stress}) (employing
1435: ${\bf p} = \rho {\bf V}$ and invoking the equation of continuity)
1436: that:
1437: \begin{equation}
1438: \rho \frac{DV_\alpha ({\bf r},t)}{Dt} = \frac{\partial \sigma_{\alpha\beta}
1439: ({\bf r},t) }{\partial r_\beta} \label{velvel}
1440: \end{equation}
1441: where $\frac{D{\bf V}}{Dt}\equiv \frac{\partial {\bf V}} {\partial
1442: t} + {\bf V} \cdot \mbox{\boldmath$\nabla$} {\bf V}$ denotes the material derivative.
1443:
1444: Next, consider the energy density. Taking the derivative of
1445: Eq.~(\ref{e1}) one obtains after some simple algebra and the use of Newton's
1446: second law:
1447: \begin{displaymath}
1448: \dot{e}({\bf r},t) = -\frac{\partial}{\partial r_\alpha}
1449: \sum_i\bigg[ \frac{1}{2} m_i v_i^2 (t)
1450: + \frac{1}{2} \sum_{j} U({\bf r}_{ij}(t)) \bigg] v_{i\alpha}(t)
1451: \phi\left[{\bf r}-{\bf r}_i(t)\right]
1452: \end{displaymath}
1453: \begin{equation}
1454: + \sum_i {\bf f}_i(t) \cdot {\bf v}_i (t) \phi\left[{\bf r}-{\bf r}_i(t)\right]
1455: -\frac{1}{2}\sum_i {\bf f}^{\mathrm{{rev}}}_{i/j}(t) \cdot {\bf v}_{ij}(t)
1456: \phi\left[{\bf r}-{\bf r}_i(t)\right]\label{e3}
1457: \end{equation}
1458: where the reversible part of the force ${\bf f}_{i/j}$
1459: is given by ${\bf f}^{\mathrm{{rev}}}_{i/j}(t)=-\frac{\partial U({\bf r}_{ij}(t))}{\partial {\bf r}_{ij}(t)}$.
1460: Using the above decomposition of the velocity,
1461: ${\bf v}_i(t) = {\bf v}^\prime_i({\bf r},t) + {\bf V}({\bf r},t)$, one
1462: obtains that the term whose divergence appears in Eq.~(\ref{e3}) can be
1463: rewritten as:
1464: \begin{displaymath}
1465: \sum_i \frac{1}{2} m_i v_i^2 (t) v_{i\alpha}(t) \phi\left[{\bf r}-{\bf r}_i(t)
1466: \right] + \frac{1}{2} \sum_{ij} U({\bf r}_{ij}(t)) v_{i\alpha}(t)
1467: \phi\left[{\bf r}-{\bf r}_i(t)
1468: \right]
1469: \end{displaymath}
1470: \begin{displaymath}
1471: = e({\bf r},t) {\bf V}_\alpha({\bf r},t)
1472: + \frac{1}{2} \sum_i m_i v^{\prime 2}_{i}({\bf r},t)
1473: v^{\prime}_{i\alpha}({\bf r},t) \phi\left[{\bf r}-{\bf r}_i(t)
1474: \right]
1475: \end{displaymath}
1476: \begin{equation}
1477: -V_\beta({\bf r},t) \sigma^{\mathrm{{kin}}}_{\alpha\beta}({\bf r},t)
1478: +\frac{1}{2} \sum_{ij} U({\bf r}_{ij}(t)) v^{\prime}_{i\alpha}({\bf r},t) \phi\left[{\bf r}-{\bf r}_i(t)
1479: \right] \label{term1}
1480: \end{equation}
1481: Next, note that
1482: \begin{displaymath}
1483: \sum_i {\bf f}_i(t) \cdot {\bf v}_i(t)\phi\left[{\bf r}-{\bf r}_i(t)\right]
1484: =\sum_{ij} {\bf f}_{i/j} (t) \cdot {\bf v}_i(t)\phi\left[{\bf r}-{\bf r}_i(t)\right]
1485: \end{displaymath}
1486: \begin{displaymath}
1487: = -\sum_{ij} {\bf f}_{i/j} (t) \cdot {\bf v}_j(t)\phi\left[{\bf r}-{\bf r}_j(t)\right]
1488: \end{displaymath}
1489: \begin{displaymath}
1490: =\frac{1}{2} \sum_{ij} {\bf f}_{i/j} (t) \cdot {\bf v}_{ij}(t)
1491: \phi\left[{\bf r}-{\bf r}_i(t)\right]
1492: \end{displaymath}
1493: \begin{displaymath}
1494: +\frac{1}{4} \sum_{ij} {\bf f}_{i/j} (t) \cdot \left({\bf v}_i(t)+ {\bf v}_j(t)
1495: \right)
1496: \left( \phi\left[{\bf r}-{\bf r}_i(t)\right]
1497: -\phi\left[{\bf r}-{\bf r}_j(t)\right] \right)
1498: \end{displaymath}
1499: \begin{displaymath}
1500: =\frac{1}{2} \sum_{ij} {\bf f}_{i/j} (t) \cdot {\bf v}_{ij}(t)
1501: \phi\left[{\bf r}-{\bf r}_i(t)\right]
1502: \end{displaymath}
1503: \begin{equation}
1504: -\frac{\partial}{\partial r_\alpha}
1505: \frac{1}{4} \sum_{ij} {\bf f}_{i/j} (t) \cdot \left[{\bf v}_i(t)+ {\bf v}_j(t)
1506: \right] {\bf r}_{ij\alpha} (t)
1507: \int_0^1 ds \phi\left[{\bf r}-{\bf r}_i(t)+ s{\bf r}_{ij} (t) \right]
1508: \label{term2}
1509: \end{equation}
1510: Using velocity decomposition and Eq.~(\ref{contactdef}) one obtains from
1511: Eq.~({\ref{term2}}):
1512: \begin{displaymath}
1513: \sum_i {\bf f}_i(t) \cdot {\bf v}_i(t)\phi\left[{\bf r}-{\bf r}_i(t)\right] =
1514: \frac{1}{2} \sum_{ij}{\bf f}_{i/j} (t) \cdot {\bf v}_{ij} (t)
1515: \phi\left[{\bf r}-{\bf r}_j(t)\right]
1516: \end{displaymath}
1517: \begin{displaymath}
1518: -\frac{1}{4} \frac{\partial}{\partial r_\alpha}
1519: \left(
1520: \sum_{ij} {\bf f}_{i/j}(t)\cdot \left[ {\bf v}^\prime_i({\bf r},t)
1521: + {\bf v}^\prime_j({\bf r},t) \right] r_{ij\alpha}(t)
1522: \int_0^1 ds \phi\left[{\bf r}-{\bf r}_i(t) + s {\bf r}_{ij}(t)\right]
1523: \right)
1524: \end{displaymath}
1525: \begin{equation}
1526: + \frac{\partial}{\partial r_\alpha}\bigg( {\bf V}_\beta({\bf r},t)
1527: \sigma^{\mathrm{{cont}}}_{\beta\alpha}({\bf r},t)\bigg) \label{term4}
1528: \end{equation}
1529: Next, combining
1530: Eqs. (\ref{stress},\ref{e3},\ref{term4}) one obtains:
1531: \begin{equation}
1532: \dot{e}({\bf r},t) = - \frac{\partial} {\partial r_\alpha}
1533: \bigg(
1534: e({\bf r},t) {\bf V}_\alpha({\bf r},t)
1535: -{\bf V}_\beta({\bf r},t) \sigma_{\beta\alpha}({\bf r},t) +
1536: Q_\alpha({\bf r},t)
1537: \bigg)- \rho({\bf r},t) \Gamma({\bf r,t}) \label{energy}
1538: \end{equation}
1539: where ${\bf Q}$ is identified as the heat flux and is given by:
1540: \begin{displaymath}
1541: Q_\alpha =
1542: \frac{1}{2} \sum_{ij} \left(
1543: U({\bf r}_{ij}(t)) + m_i v_i^{\prime 2} v^\prime_{i\alpha} \right)
1544: v^\prime_{i\alpha}({\bf r},t) \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1545: \end{displaymath}
1546: \begin{equation}
1547: +\frac{1}{4} \sum_{ij} {\bf f}_{i/j}(t)\cdot \bigg( {\bf v}^\prime_i({\bf r},t)
1548: + {\bf v}^\prime_j({\bf r},t) \bigg) r_{ij\alpha}(t)
1549: \int_0^1 ds \phi\left[{\bf r}-{\bf r}_i(t) + s {\bf r}_{ij}(t)\right]
1550: \label{heatflux1}
1551: \end{equation}
1552: and the sink term, $-\rho\Gamma$ is given by:
1553: \begin{displaymath}
1554: -\rho({\bf r},t) \Gamma({\bf r},t) = \frac{1}{2} \sum_{ij} \bigg( {\bf f}_{i/j}({\bf r},t)
1555: - {\bf f}^{\mathrm{{rev}}}_{ij}({\bf r},t) \bigg)\cdot {\bf v}_{ij} ({\bf r},t) \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1556: \end{displaymath}
1557: \begin{equation}
1558: = \frac{1}{2} \sum_{ij} {\bf f}^{\mathrm{{irrev}}}_{ij}({\bf r},t)
1559: \cdot {\bf v}_{ij} ({\bf r},t) \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1560: \end{equation}
1561: where the dissipative part of the force,
1562: ${\bf f}^{\mathrm{{irrev}}}_{ij}$,
1563: is given by ${\bf f}_{i/j}-{\bf f}^{\mathrm{{rev}}}_{ij}$.
1564: Note that the heat flux represents the transport of energy by the fluctuating
1565: velocity and the power related to the fluctuating velocities. The
1566: energy sink term is entirely due to the power associated with the
1567: dissipative part of the forces.
1568:
1569: The energy density can be rewritten as follows:
1570: \begin{displaymath}
1571: e({\bf r},t) \equiv \sum_i \frac{1}{2} m_i v^{\prime 2}_i \phi\left[ {\bf r}-{\bf r}_i(t)\right]
1572: \end{displaymath}
1573: \begin{equation}
1574: +\frac{1}{2} \rho({\bf r}) {\bf V}^2 ({\bf r},t)
1575: +\frac{1}{2} \sum_{i,j} U({\bf r}_{ij}(t))\phi\left[ {\bf r}-{\bf r}_i(t)\right] \label{e10}
1576: \end{equation}
1577: Denote the internal specific energy by $\overline{u}$. The latter
1578: is given by
1579: \begin{equation}
1580: \rho({\bf r},t) \overline{u}({\bf r},t) \equiv \sum_i \frac{1}{2} m_i
1581: v^{\prime 2}_i \phi\left[ {\bf r}-{\bf r}_i(t)\right] +
1582: \frac{1}{2} \sum_{i,j} U({\bf r}_{ij}(t))\phi\left[ {\bf r}-{\bf r}_i(t)\right]
1583: \label{defT}
1584: \end{equation}
1585: and the energy density therefore equals $\frac{1}{2} \rho {\bf V}^2 + \rho \overline{u}$. Substituting this decomposition into Eq.~(\ref{energy}) and using Eqs.~(\ref{rhodot1},\ref{velvel}) one obtains:
1586: \begin{equation}
1587: \rho({\bf r},t) \frac{D\overline{u}({\bf r},t)}{Dt}
1588: = \frac{\partial V_\beta({\bf r},t)}{\partial r_\alpha} \sigma_{\beta\alpha}
1589: ({\bf r},t) - \mathrm{div}\, {\bf Q}({\bf r},t) -\rho({\bf r},t) \Gamma({\bf r},t) \label{internal}
1590: \end{equation}
1591: \section{The quasistatic limit of Continuum Mechanics}
1592: Consider the case of slow deformations of a near-static
1593: assembly of particles. To this end, let the particle velocities be denoted
1594: by ${\bf v}_i(t) = \delta {\bf \tilde v}_i(t)$, where $\delta$ is a small
1595: parameter and $\tilde{A}$ is considered to be $O(1)$
1596: for any dynamical variable, $A$. The velocity field, ${\bf V}$, as
1597: well as all other fields are
1598: rescaled in a similar way.
1599: Furthermore, let time be expressed as $\tilde{t}= \frac{t}{\delta}$, hence
1600: $\frac{\partial} {\partial t} = \delta \frac{\partial}{{\partial \tilde{ t}}}$. With this
1601: rescaling the equations of continuum mechanics, Eqs.~(\ref{rhodot1},\ref{velvel},\ref{energy})
1602: become:
1603: \begin{displaymath}
1604: \delta \frac{\partial \rho}{\partial \tilde{t}} =
1605: - \delta \mathrm{div}\,(\rho {\bf \tilde{V}})
1606: \end{displaymath}
1607: \begin{displaymath}
1608: \delta^2 \rho \frac{D\tilde{V}_\alpha}{D\tilde{t}} =
1609: \delta^2 \frac{\partial\tilde{\sigma}^{\mathrm{{kin}}}_{\alpha\beta}}{\partial r_\beta}
1610: + \frac{\partial \sigma^{\mathrm{{cont}}}_{\alpha\beta}}{\partial r_\beta}
1611: \end{displaymath}
1612: \begin{displaymath}
1613: \delta^3 \frac{\partial \tilde{e}^{\mathrm{{kin}}}}{\partial \tilde{t}}
1614: +\delta \frac{\partial \tilde{e}^{\mathrm{{pot}}}}{\partial \tilde{t}}
1615: =- \frac{\partial}{\partial r_\alpha}
1616: \bigg(
1617: \delta^3 \tilde{e}^{\mathrm{{kin}}}{\bf \tilde{V}}_\alpha
1618: + \delta e^{\mathrm{{pot}}} {\bf \tilde{V}}_\alpha
1619: -\delta^3 {\bf \tilde{V}}_\beta \tilde{\sigma}^{\mathrm{{kin}}}_{\beta\alpha}
1620: \end{displaymath}
1621: \begin{equation}
1622: -\delta {\bf \tilde{V}}_\beta {\sigma}^{\mathrm{{cont}}}_{\beta\alpha}
1623: + \delta \tilde{Q}^{\mathrm{{pot}}}_\alpha
1624: +\delta^4 \tilde{Q}^{\mathrm{{kin}}}_\alpha
1625: +\delta \tilde{Q}^{\mathrm{{force}}}_\alpha
1626: \bigg)
1627: - \delta \rho \tilde{\Gamma}
1628: \end{equation}
1629: where the superscript `kin' refers to the kinetic part, the
1630: superscript `pot' refers to the potential part (involving the potential
1631: $U$) and the superscript force refers to the fluctuating power contribution
1632: to the heat flux. To linear order in $\delta$ the above equations reduce to:
1633: \begin{equation}
1634: \frac{\partial \rho}{\partial \tilde{t}} =
1635: - \mathrm{div}\,(\rho {\bf \tilde{V}})\label{first}
1636: \end{equation}
1637: \begin{equation}
1638: \frac{\partial \sigma^{\mathrm{{cont}}}_{\alpha\beta}}{\partial r_\beta} =0
1639: \label{second}
1640: \end{equation}
1641: \begin{equation}
1642: \frac{\partial \tilde{e}^{\mathrm{{pot}}}}{\partial \tilde{t}}
1643: =- \frac{\partial}{\partial r_\alpha}
1644: \bigg(
1645: e^{\mathrm{{pot}}} {\bf \tilde{V}}_\alpha
1646: - {\bf \tilde{V}}_\beta {\sigma}^{\mathrm{{cont}}}_{\beta\alpha}
1647: + {Q}^{\mathrm{{pot}}}_\alpha
1648: + \tilde{Q}^{\mathrm{{force}}}_\alpha
1649: \bigg)
1650: -\rho \tilde {\Gamma}\label{slow11}
1651: \end{equation}
1652: Using Eq.~(\ref{second}), the equation for the energy
1653: density becomes to linear order in the `slowness', $\delta$ (and
1654: after reverting to the original time variable):
1655: \begin{equation}
1656: \frac{\partial {e}^{\mathrm{{pot}}}} {\partial {t}}
1657: =- \frac{\partial}{\partial r_\alpha}
1658: \bigg(
1659: e^{\mathrm{{pot}}} {\bf {V}}_\alpha
1660: + {Q}^{\mathrm{{pot}}}_\alpha
1661: + {Q}^{\mathrm{{force}}}_\alpha
1662: \bigg)
1663: + \frac{\partial {\bf {V}}_\beta}
1664: {\partial r_\alpha}
1665: {\sigma}^{\mathrm{ {cont}}}_{\beta\alpha}
1666: -\rho {\Gamma}\label{slow1}
1667: \end{equation}
1668: We further specialize to the case when the particle relative
1669: displacements are small and the potential is quadratic in the
1670: relative displacements (to leading order).
1671: Let ${\bf u}_i(t)$ be the displacement of particle $i$:
1672: $\dot{\bf u}_i(t) = {\bf v}_i$. Also, define the displacement field
1673: ${\bf u}({\bf r},t)$ via the Lagrangian relation:
1674: $\frac{\partial {\bf u}^{\mathrm{{La}}}({\bf R},t)}{\partial t}
1675: ={\bf V}^{\mathrm{{La}}}({\bf R},t)$, where ${\bf R}$ is the initial position
1676: of a pathline and the superscript `La' denotes the fact that one is employing
1677: Lagrangian coordinates. As shown in \cite{Goldhirsch02}
1678: \begin{equation}
1679: {\bf u}^{\mathrm{{lin}}}({\bf r},t) =
1680: \frac{\sum_i m_i {\bf u}_i(t) \phi\left[{\bf r}-{\bf r}_i(t)\right]}
1681: {\sum_j m_j \phi\left[{\bf r}-{\bf r}_j(t)\right]} \label{linearu}
1682: \end{equation}
1683: where the superscript `lin' denotes the linear order in the displacements. To
1684: this order the (linear) strain field, $\mbox{\boldmath $\epsilon$}^{\mathrm{{lin}}}$ is given by:
1685: \begin{equation}
1686: \epsilon^{\mathrm{{lin}}}_{\alpha\beta}({\bf r},t)=\frac{1}{2}\bigg(\frac{\partial
1687: u_\alpha ^{\mathrm{{lin}}}({\bf r},t)}{\partial r_\beta}
1688: +\frac{\partial
1689: u_\beta ^{\mathrm{{lin}}}({\bf r},t)}{\partial r_\alpha}\bigg)
1690: \label{epslin}
1691: \end{equation}
1692: Keeping in Eq.~(\ref{slow1}) only terms that are quadratic in the
1693: displacements one obtains (note that the potential is of second
1694: order in the displacements):
1695: \begin{equation}
1696: \frac{\partial {e}^{\mathrm{{pot}}}} {\partial {t}}
1697: = -\mathrm{div}\,{{\bf Q}}^{\mathrm{{force}}}
1698: + \frac{\partial {\bf {V}}_\beta}
1699: {\partial r_\alpha}
1700: {\sigma}^{\mathrm{ {cont}}}_{\beta\alpha}
1701: -\rho {\Gamma}\label{slow2.5}
1702: \end{equation}
1703: Since for quadratic potentials $\mbox{\boldmath $\sigma$}^{\mathrm{cont}}$ is
1704: symmetric (as is easy to check, see also Eq.~(\ref{eq:lin_stress})) one can rewrite
1705: Eq.~(\ref{slow2.5}) as follows:
1706: \begin{equation}
1707: \frac{\partial {e}^{\mathrm{{pot}}}} {\partial {t}}
1708: = -\mathrm{div}\,{{\bf Q}}^{\mathrm{{force}}}
1709: + \frac{1}{2} \bigg( \frac{\partial {\bf {V}}_\beta}
1710: {\partial r_\alpha}
1711: + \frac{\partial {\bf {V}}_\alpha}
1712: {\partial r_\beta}\bigg)
1713: {\sigma}^{\mathrm{ {cont}}}_{\beta\alpha}
1714: - \rho {\Gamma}\label{slow2}
1715: \end{equation}
1716: hence (to quadratic order in the displacements)
1717: \begin{equation}
1718: \frac{\partial {e}^{\mathrm{{pot}}}} {\partial {t}}
1719: = -\mathrm{div}\, {{\bf Q}}^{\mathrm{{force}}}
1720: +
1721: \frac{\partial \epsilon^{\mathrm{{lin}}}_{\alpha\beta}} {\partial {t}}
1722: {\sigma}^{\mathrm{ {cont}}}_{\beta\alpha}
1723: - \rho {\Gamma}\label{slow3}
1724: \end{equation}
1725: Consider next the following virtual dynamical evolution from a force free
1726: initial state to a final state at time $t_{\mathrm{{f}}}$:
1727: ${\bf u}_i(t) = {\bf u}_i(t_{\mathrm{{f}}})
1728: \frac{t}{t_{\mathrm{{f}}}}$.
1729: Assume that the reversible part of the force,
1730: ${\bf f}^{\mathrm{{rev}}}_{i/j}$ is linear in the displacement difference ${\bf u}_{ij}$ and
1731: the irreversible part of the force is linear in the velocity difference
1732: ${\bf v}_{ij}$. In this
1733: dynamics the velocities are constant, the strain field satisfies
1734: $\mbox{\boldmath$\epsilon$}^{\mathrm{{lin}}}({\bf r},t)=
1735: \mbox{\boldmath$\epsilon$}^{\mathrm{{lin}}}({\bf r},t_{\mathrm{{f}}})\frac{t}{t_{\mathrm{{f}}}}$, the reversible part of the contact stress,
1736: $\mbox{\boldmath$\sigma$}^{\mathrm{{cont,rev}}}$ (in which the force
1737: is replaced by the reversible part of the force) satisfies
1738: $\mbox{\boldmath$\sigma$}^{\mathrm{{cont,rev}}}({\bf r},t) =
1739: \mbox{\boldmath$\sigma$}^{\mathrm{{cont,rev}}}({\bf r},t_{\mathrm{{f}}})
1740: \frac{t}{t_{\mathrm{{f}}}}$, the irreversible part of the contact stress,
1741: $\mbox{\boldmath$\sigma$}^{\mathrm{{cont,irrev}}}({\bf r},t)$, is constant in time, and similarly the part of the heat flux which corresponds to the reversible parts of the
1742: forces satisfies
1743: ${\bf Q}^{\mathrm{force,rev}}({\bf r},t)=
1744: {\bf Q}^{\mathrm{force,rev}}({\bf r},t_{\mathrm{{f}}})
1745: \frac{t}{t_{\mathrm{{f}}}}$ and
1746: ${\bf Q}^{\mathrm{force,irrev}}({\bf r},t)$ is constant in time. In addition,
1747: $\rho \Gamma({\bf r},t)$ is constant in time. Note that within the
1748: approximation in which only the lowest nonvanishing order in the
1749: displacement is retained,
1750: ${\bf r}_{ij}(t)$ is replaced by ${\bf r}_{ij}(0)$. Upon substituting
1751: these expressions in Eq.~({\ref{slow3}) and integrating Eq.~(\ref{slow3})
1752: over time from zero to $t_{\mathrm{{f}}}$ one obtains:
1753: \begin{displaymath}
1754: e({\bf r}, t_{\mathrm{{f}}}) = \frac{1}{2} \epsilon^{\mathrm{{lin}}}_{\alpha\beta}
1755: ({\bf r}, t_{\mathrm{{f}}}) \sigma^{\mathrm{{cont,rev}}}_{\beta\alpha}
1756: ({\bf r}, t_{\mathrm{{f}}})
1757: + \epsilon^{\mathrm{{lin}}}_{\alpha\beta}
1758: ({\bf r}, t_{\mathrm{{f}}}) \sigma^{\mathrm{{cont,irrev}}}_{\beta\alpha}
1759: ({\bf r}, t_{\mathrm{{f}}})
1760: \end{displaymath}
1761: \begin{equation}
1762: -\frac{1}{2}\mathrm{div}\, {\bf Q}^{\mathrm{{force,rev}}}({\bf r}, t_{\mathrm{{f}}}) t_{\mathrm{{f}}}
1763: -\mathrm{div}\,{\bf Q}^{\mathrm{{force,irrev}}}({\bf r}, t_{\mathrm{{f}}})
1764: { t_{\mathrm{{f}}}}
1765: -\rho({\bf r}, t_{\mathrm{{f}}}) \Gamma({\bf r}, t_{\mathrm{{f}}})
1766: { t_{\mathrm{{f}}}} \label{almost}
1767: \end{equation}
1768: In order to keep the displacements fixed we replace the velocities
1769: that appear in the expressions for the various fields by $1/t_{\mathrm{{f}}}$ times the corresponding displacements. A field in which the velocities have
1770: been replaced by the corresponding displacements is marked by a superscript I. It follows that at $t=t_{\mathrm{{f}}}$:
1771: $\mbox{\boldmath$\sigma$}^{\mathrm{{cont,rev,I}}}=\mbox{\boldmath$\sigma$}^{\mathrm{{cont,rev}}}$,
1772: $\mbox{\boldmath$\sigma$}^{\mathrm{{cont,irrev,I}}} = t_{\mathrm{{f}}}
1773: \mbox{\boldmath$\sigma$}^{\mathrm{{cont,rev}}}$,
1774: ${\bf Q}^{\mathrm{{force,rev,I}}}= t_{\mathrm{{f}}}
1775: {\bf Q}^{\mathrm{{force,rev}}}$,
1776: ${\bf Q}^{\mathrm{{force,irrev,I}}}= t^2_{\mathrm{{f}}}{\bf Q}^{\mathrm{{force,irrev}}}$ (recall the assumption that ${\bf f}^{\mathrm{{irrev}}}_{i/j}$
1777: is linear
1778: in the particle velocities),
1779: and $(\rho\Gamma)^{\mathrm{{I}}} = t^2_{\mathrm{{f}}} \rho\Gamma$.
1780: Substituting these expressions in Eq.~(\ref{almost}), keeping the displacements
1781: fixed and letting $t_{\mathrm{{f}}}$ go to infinity, one obtains:
1782: \begin{equation}
1783: e({\bf r}, t_{\mathrm{{f}}}) = \frac{1}{2} \epsilon^{\mathrm{{lin}}}_{\alpha\beta}
1784: ({\bf r}, t_{\mathrm{{f}}}) \sigma^{\mathrm{{cont,rev}}}_{\beta\alpha}
1785: ({\bf r}, t_{\mathrm{{f}}})
1786: -\frac{1}{2}\mathrm{div}\, {\bf Q}^{\mathrm{{force,rev,I}}}({\bf r}, t_{\mathrm{{f}}})
1787: \label{tiny}
1788: \end{equation}
1789: where (changing $t_{\mathrm{{f}}}$ to $t$) if follows from the expression
1790: for the heat flux, Eq~(\ref{heatflux1}) and the above definitions that
1791: \begin{eqnarray}
1792: Q^{\mathrm{{force,rev,I}}}_\alpha &=&
1793: \frac{1}{4} \sum_{ij} {\bf f}^{\mathrm{{rev}}}_{ij}(t)\cdot \left( {\bf u}^\prime_i({\bf r},t)
1794: + {\bf u}^\prime_j({\bf r},t) \right) r_{ij\alpha}(0)\\
1795: &&\times \int_0^1 ds \phi\left[{\bf r}-{\bf r}_i(t) + s {\bf r}_{ij}(t)\right]
1796: \label{last}
1797: \end{eqnarray}
1798: where
1799: \begin{equation}
1800: \label{flu}
1801: {\bf u}^\prime_i({\bf r},t) \equiv {\bf u}_i(t) - {\bf u}^{\mathrm{{lin}}}
1802: ({\bf r},t)
1803: \end{equation}
1804: is the displacement fluctuation of particle $i$.
1805: We have thus obtained a correction to the classical formula for the
1806: elastic energy which represents the fluctuating work and stems from the integral
1807: of the heat flux over time. This result can be directly obtained from
1808: an analysis of the elastic energy, cf. Eq.~(\ref{eee}).
1809:
1810: \begin{thebibliography}{10}
1811: \expandafter\ifx\csname url\endcsname\relax
1812: \def\url#1{\texttt{#1}}\fi
1813: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1814:
1815: \bibitem{IO}
1816: I.~Oppenheim, Nonlinear response theory, in: J.~W. Halley (Ed.), Correlation
1817: functions and quasiparticle interactions in condensed matter, Plenum, NY,
1818: 1998, pp. 235--258.
1819:
1820: \bibitem{Mori1}
1821: H.~Mori, Time-correlation functions in the statistical mechanics of transport
1822: processes, Phys. \ Rev. 111 (1958) 694--706.
1823:
1824: \bibitem{Mori2}
1825: H.~Mori, Statistical mechanical theory for transport in fluids, Phys. \ Rev.
1826: 112 (1958) 1829--1842.
1827:
1828: \bibitem{vanKampen86}
1829: N.~G. van Kampen, I.~Oppenheim, Brownian motion as a problem of eliminating
1830: fast variables, Physica A 138 (1986) 231--248.
1831:
1832: \bibitem{HarrisX}
1833: S.~Harris, An Introduction to the theory of the Boltzmann equation, Holt,
1834: Rinehart and Winston, New-York, 1971.
1835:
1836: \bibitem{Goldhirsch07a}
1837: I.~Goldhirsch, A.~S. Peletminskii, S.~V. Peletminskii, A.~I. Sokolovsky,
1838: Application of {B}ogolyubov's approach to the derivation of kinetic equations
1839: for dissipative fluids, preprint (2007).
1840:
1841: \bibitem{Chapman70}
1842: S.~Chapman, T.~G. Cowling, The mathematical theory of nonuniform gases,
1843: Cambridge University Press, Cambridge, 1970.
1844:
1845: \bibitem{Goldhirsch03}
1846: I.~Goldhirsch, Rapid granular flows, Annu.\ Rev.\ Fluid Mech. 35 (2003)
1847: 267--293.
1848:
1849: \bibitem{Tan98}
1850: M.-L. Tan, I.~Goldhirsch, Rapid granular flows as mesoscopic systems, Phys.\
1851: Rev.\ Lett. 81 (1998) 3022--3025.
1852:
1853: \bibitem{Goldenberg02}
1854: C.~Goldenberg, I.~Goldhirsch, Force chains, microelasticity, and
1855: macroelasticity, Phys.\ Rev.\ Lett. 89 (2002) 084302.
1856:
1857: \bibitem{Goldenberg06b}
1858: C.~Goldenberg, A.~P.~F. Atman, P.~Claudin, G.~Combe, I.~Goldhirsch, Scale
1859: separation in granular packings: stress plateaus and fluctuations, Phys.\
1860: Rev.\ Lett. 96 (2006) 168001.
1861:
1862: \bibitem{Mehta89}
1863: A.~Mehta, S.~F. Edwards, Statistical mechanics of powder mixtures, Physica A
1864: 157 (1989) 1091--1097.
1865:
1866: \bibitem{Edwards90}
1867: S.~F. Edwards, The rheology of powders, Rheologica Acta 29 (1990) 493--499.
1868:
1869: \bibitem{Makse02}
1870: H.~A. Makse, J.~Kurchan, Testing the thermodynamic approach to granular matter
1871: with a numerical model of a decisive experiment, Nature 415 (2002) 614--617.
1872:
1873: \bibitem{Mayor05}
1874: P.~Mayor, G.~D'Anna, A.~Barrat, V.~Loreto, Observing {B}rownian motion and
1875: measuring temperatures in vibration-fluidized granular matter, New J. Phys. 7
1876: (2005) 1--16.
1877:
1878: \bibitem{Goldhirsch00}
1879: I.~Goldhirsch, T.~P.~C. van Noije, Green-{K}ubo relations for granular fluids,
1880: Phys.\ Rev.~E 61 (2000) 3241--3244.
1881:
1882: \bibitem{Brilliantov05}
1883: N.~V. Brilliantov, T.~P{\"o}schel, Self-difussion in granular gases:
1884: {G}reen-{K}ubo versus {C}hapman-{E}nskog, Chaos 15 (2005) 026108.
1885:
1886: \bibitem{Dufty06}
1887: A.~B. J.Dufty, J.~J. Brey, Linear response for granular fluids, J. Stat. Mech.
1888: Theor. and Expt. (2006) L08002.
1889:
1890: \bibitem{Puglisi02}
1891: A.~Puglisi, A.~Baldassarri, V.~Loreto, Fluctuation-dissipation relations in
1892: driven granular gases, Phys.\ Rev.~E (2002) 061305.
1893:
1894: \bibitem{Tsallis88}
1895: C.~Tsallis, Possible generalization of {B}oltzmann-{G}ibbs statistics, J. Stat.
1896: Phys. 52 (1988) 479--487.
1897:
1898: \bibitem{Beck06}
1899: C.~Beck, Stretched exponentials from superstatistics, Physica A 365 (2006)
1900: 96--101.
1901:
1902: \bibitem{Hutter94}
1903: K.~Hutter, K.~R. Rajagopal, On flows of granular materials, Continuum Mechanics
1904: and Thermodynamics 6 (1994) 81--139.
1905:
1906: \bibitem{Kumaran06}
1907: V.~Kumaran, The constitutive relation for the granular flow of rough particles,
1908: J.\ Fluid Mech. 561 (2006) 1--42.
1909:
1910: \bibitem{Sela98}
1911: N.~Sela, I.~Goldhirsch, Hydrodynamic equations for rapid flows of smooth
1912: inelastic spheres, to {B}urnett order, J.\ Fluid Mech. 361 (1998) 41--74.
1913:
1914: \bibitem{Brey98}
1915: J.~J. Brey, J.~W. Dufty, C.~S. Kim, A.~Santos, Hydrodynamics of granular flow
1916: at low density, Phys.\ Rev.~E 58 (1998) 4638--4653.
1917:
1918: \bibitem{Glasser01}
1919: B.~J. Glasser, I.~Goldhirsch, Scale dependence, correlations, and fluctuations
1920: of stresses in rapid granular flows, Phys.\ Fluids 13 (2001) 407--420.
1921:
1922: \bibitem{Goldshtein95}
1923: A.~Goldshtein, M.~Shapiro, Mechanics of collisional motion of granular
1924: materials, part {I}: General hydrodynamic equations, J.\ Fluid Mech. 282
1925: (1995) 75--114.
1926:
1927: \bibitem{Goldhirsch93a}
1928: I.~Goldhirsch, M.-L. Tan, G.~Zanetti, A molecular dynamical study of granular
1929: fluids: the unforced granular gas, J. \ Sci. \ Comp. 8 (1993) 1--40.
1930:
1931: \bibitem{Soto01}
1932: R.~Soto, M.~Mareschal, Statistical mechanics of fluidized granular media:
1933: short-range velocity correlations, Phys.\ Rev.~E 63 (2001) 041303.
1934:
1935: \bibitem{Luding00}
1936: S.~Luding, On the relevance of ``molecular chaos'' for granular flows, ZAMM. 80
1937: (2000) S9--S12.
1938:
1939: \bibitem{Noije00}
1940: T.~P.~C. van Noije, M.~H. Ernst, Cahn-{H}illiard theory for unstable granular
1941: fluids, Phys.\ Rev.~E 61 (1997) 1765--1982.
1942:
1943: \bibitem{Wakou02}
1944: J.~Wakou, R.~Brito, M.~H. Ernst, Towards a {L}andau-{G}inzburg type theory for
1945: granular fluids, J. Stat. Phys 107 (2002) 3--22.
1946:
1947: \bibitem{Noskowicz06}
1948: S.~H. Noskowicz, O.~Bar-Lev, D.~Serero, I.~Goldhirsch, Strongly inelastic
1949: granular gases, cond-mat/0612694.
1950:
1951: \bibitem{Garzo02}
1952: V.~Garzo, J.~W. Dufty, Hydrodynamics of a granular binary mixture at low
1953: density, Phys. Fluids 14 (2002) 1476--1490.
1954:
1955: \bibitem{Serero06}
1956: D.~Serero, I.~Goldhirsch, S.~H. Noskowicz, M.-L. Tan, Hydrodynamics of granular
1957: gases and granular gas mixtures, J.\ Fluid Mech. 554 (2006) 237--258.
1958:
1959: \bibitem{Gladwell80}
1960: G.~M.~L. Gladwell, Contact Problems in the Classical Theory of Elasticity,
1961: Sijthoff \& Noordhoff, The Netherlands, 1980.
1962:
1963: \bibitem{Johnson85}
1964: K.~L. Johnson, Contact Mechanics, Cambridge University Press, Cambridge, 1985.
1965:
1966: \bibitem{Sadd93}
1967: M.~H. Sadd, Q.~Tai, A.~Shukla, Contact law effects on wave propagation in
1968: particulate materials using distinct element modeling, Int.\ J.\ Non-Linear
1969: Mech. 28 (1993) 251--265.
1970:
1971: \bibitem{Walton95}
1972: O.~R. Walton, Force models for particle-dynamics simulations of granular
1973: materials, in: E.~Guazzelli, L.~Oger (Eds.), Mobile Particulate Systems,
1974: Kluwer, Dordrecht, 1995, pp. 367--380.
1975:
1976: \bibitem{Schafer96}
1977: J.~Sch{\"a}fer, S.~Dippel, D.~E. Wolf, Force schemes in simulations of granular
1978: materials, J. de Physique I 6 (1996) 5--20.
1979:
1980: \bibitem{Wolf96}
1981: D.~E. Wolf, Modeling and computer simulation of granular media, in: K.~H.
1982: Hoffmann, M.~Schreiber (Eds.), Computational Physics, Springer, Heidelberg,
1983: 1996, pp. 64--94.
1984:
1985: \bibitem{Herrmann98}
1986: H.~J. Herrmann, S.~Luding, Review article: Modeling granular media on the
1987: computer, Cont. Mech. and Thermodynamics 10 (1998) 189--231.
1988:
1989: \bibitem{Goldhirsch02}
1990: I.~Goldhirsch, C.~Goldenberg, On the microscopic foundations of elasticity,
1991: Eur.\ Phys.\ J.~E 9 (2002) 245--251.
1992:
1993: \bibitem{Goldenberg06c}
1994: C.~Goldenberg, A.~Tanguy, J.-L. Barrat, Particle displacements in the
1995: deformation of amorphous materials: local fluctuations vs. the non-affine
1996: field, cond-mat/0610518.
1997:
1998: \bibitem{Lemaitre02}
1999: A.~Lemaitre, Rearrangements and dilatancy for sheared dense materials, Phys.\
2000: Rev.\ Lett. 89 (2002) 195503.
2001:
2002: \bibitem{Utter07}
2003: B.~Utter, R.~P. Behringer, Experimental measures of affine and non-affine
2004: deformation in granular shear, cond-mat/0702334.
2005:
2006: \bibitem{Savage98}
2007: S.~B. Savage, Analyses of slow high-concentration flows of granular materials,
2008: J.\ Fluid Mech. 377 (1998) 1--26.
2009:
2010: \bibitem{Babic97}
2011: M.~Babic, Average balance equations for granular materials, Int.\ J.\ Eng.\
2012: Sci. 35 (1997) 523--548.
2013:
2014: \end{thebibliography}
2015:
2016: \end{document}
2017: