Вы находитесь на странице: 1из 7

PHYSICAL REVIEW E 80, 031131 共2009兲

Nonequilibrium thermodynamics of driven amorphous materials. I.


Internal degrees of freedom and volume deformation
Eran Bouchbinder
Racah Institute of Physics, Hebrew University of Jerusalem, Jerusalem 91904, Israel

J. S. Langer
Department of Physics, University of California, Santa Barbara, California 93106-9530, USA
共Received 9 March 2009; published 22 September 2009兲
This is the first of three papers devoted to the nonequilibrium thermodynamics of amorphous materials. Our
focus here is on the role of internal degrees of freedom in determining the dynamics of such systems. For
illustrative purposes, we study a solid whose internal degrees of freedom are vacancies that govern irreversible
volume changes. Using this model, we compare a thermodynamic theory based on the Clausius-Duhem in-
equality to a statistical analysis based directly on the law of increase of entropy. The statistical theory is used
first to derive the Clausius-Duhem inequality. We then use the theory to go beyond those results and obtain
detailed equations of motion, including a rate factor that is enhanced by deformation-induced noisy fluctua-
tions. The statistical analysis points to the need for understanding how both energy and entropy are shared by
the vacancies and their environments.

DOI: 10.1103/PhysRevE.80.031131 PACS number共s兲: 05.70.Ln, 62.20.F⫺, 63.50.Lm

I. INTRODUCTION Almost by definition, the irreversible responses of mate-


rials to applied forces are determined by internal degrees
This is the first of three papers describing our efforts to of freedom. Deforming amorphous solids are generally
develop a thermodynamically well-founded theory of non- described in terms of internal entities such as flow defects
equilibrium phenomena in amorphous materials. Specific or, of special interest here, STZs. Qualitatively similar inter-
goals of this project are to develop a thermodynamic under- nal structures appear in theories of dislocation motion in
standing of the effective disorder temperature and the role crystalline solids and in nonequilibrium theories of granular
that it plays in shear-transformation-zone 共STZ兲 theories of materials and complex fluids. Theories of these dissipative
amorphous plasticity 关1–5兴. While working toward these
phenomena necessarily invoke the second law of thermody-
goals, we have encountered a number of fundamental ques-
namics at the very least as a constraint on the equations of
tions. Those questions include: what is the most basic state-
motion for the internal variables.
ment of the second law of thermodynamics? How can we
reconcile the different approaches to nonequilibrium thermo- There is a very large body of literature on this subject. For
dynamics taken by engineers, applied mathematicians, and example, see monographs by Lubliner 关8兴, Maugin 关9兴, and
physicists? Are the dynamic roles played by internal degrees Nemat-Nasser 关10兴, which we have found to be especially
of freedom properly described by any of those theoretical useful. Essentially all of this literature is based on the postu-
approaches? Many recent developments in the physics of late that the Clausius-Duhem entropy-production inequality
glassy materials, including the STZ theory, are based on the is the fundamental statement of the second law of thermody-
idea that the state of disorder in such systems is described by namics; therefore, we refer to that theoretical starting point
an effective temperature that is not necessarily the same as as “conventional.” We recognize, however, that the body of
the ordinary temperature. In what sense is the effective tem- literature to which we are referring contains many different
perature a well-defined thermodynamic concept? points of view and that these points of view have continued
In this first paper, we focus on questions regarding inter- to evolve in recent decades, especially in the engineering and
nal degrees of freedom. For illustrative purposes, we address applied mathematics communities. Our motivation for devel-
that issue in the limited context of a simple model of a uni- oping a statistical approach based directly on the law of in-
form, not necessarily glassy, solid in which vacancies govern crease of entropy simply reflects the fact that we have not
irreversible volume changes. We originally developed this been able to take any version of the conventional approach
vacancy model as a way of studying irreversible changes in far enough to answer the questions that we are asking.
the volume of a glassy material subject to varying tempera- In this paper, we look at the issues concerning the second
tures and pressures. We hope to return to such applications in law and internal degrees of freedom in a simple but physi-
the future, but, for the present, we use the model purely as an cally realistic situation—the vacancy model mentioned ear-
aid for exploring theoretical ideas. In the second paper 关6兴, lier. We start in Sec. II by introducing the model and then, in
we use the insights gained here to define the effective disor- Sec. III, by briefly summarizing a conventional analysis. The
der temperature and to write equations of motion for it. Fi- first-principles statistical theory and the resulting nonequilib-
nally, in the third paper 关7兴, we reformulate the STZ theory rium equations of motion are presented in Secs. IV–VI. We
in a way that is consistent with the thermodynamic analysis conclude in Sec. VII with some remarks about the broader
presented here and in 关6兴. implications of our results.

1539-3755/2009/80共3兲/031131共7兲 031131-1 ©2009 The American Physical Society


ERAN BOUCHBINDER AND J. S. LANGER PHYSICAL REVIEW E 80, 031131 共2009兲

II. VACANCY MODEL


V̇in = v0Ṅv 共2.2兲
We consider a uniform viscoelastic solid of total volume
V, containing a small but extensive number of vacancies Nv. and that Nv is a dynamical variable that satisfies its own
It may be easiest to visualize this solid as being noncrystal- equation of motion.
line, but that assumption is not essential for present purposes.
To avoid the complications of position-dependent deforma- III. CONVENTIONAL THEORY
tion, we assume that this system remains spatially uniform at
all times and we work with extensive quantities rather than A conventional analysis of this model, described along
local densities. We further assume that the system is never lines laid out by Coleman, Noll, and Gurtin in the 1960s
too far from thermodynamic equilibrium, i.e., that we are not 关12,13兴, starts by writing the first law of thermodynamics in
dealing with extremely rapid nonequilibrium phenomena for the form
which local thermodynamic concepts would be invalid. This
quasiequilibrium condition is essential for our arguments in U̇ = − pV̇ + Q, 共3.1兲
Sec. IV.
Our model, like any model of a material subject only to where U is the internal energy, p is the pressure, −pV̇ is the
volume and not shape deformations, is technically viscoelas- work done on the system, and Q is the rate at which thermal
tic rather than elastoplastic. If it is allowed to equilibrate at a energy is entering the system. The conventional theory then
fixed temperature and pressure, it eventually returns to the postulates that there exists an entropy S and a temperature ␪
same equilibrium volume with the same number of vacan- defined by a continuity equation. For this spatially uniform
cies, whereas a true elastoplastic material would return to a system, that equation is simply
permanently deformed shape if subjected to shear. A second
difference is that a plastic material can undergo steady-state Q
Ṡ − = ⌺. 共3.2兲
shear flow and usually exhibits a yield stress that marks the ␪
onset of that behavior. No such steady-state deformation can
occur in the present case. Here, the temperature ␪ is expressed in energy units
The fundamental differences between these kinds of in- 共kB = 1兲, Q / ␪ is the rate at which entropy is entering the
elastic deformation are important, but are not the central is- system, and ⌺ is the entropy source, i.e., the rate at which
sues to be discussed here. Rather, the model of purely volu- entropy is being produced. The conventional statement of the
metric deformation is especially useful to us because there is second law, the Clausius-Duhem inequality, says that the en-
no need for anything other than an additive decomposition of tropy production rate is non-negative
arbitrarily large elastic and inelastic volume changes. No
⌺ ⱖ 0. 共3.3兲
special mathematical efforts are needed to compute the re-
sults of complex sequences of deformations, but the separa- These relations are taken to be axiomatic; they do not pre-
tion between elastic and inelastic deformations remains a suppose any statistical interpretation of S or ␪.
nontrivial topic of interest as discussed, for example, in 关11兴. Eliminating Q between Eqs. 共3.1兲 and 共3.2兲, we find
The volume V in this model consists of three additive
components
␪Ṡ − U̇ − pV̇ = ␪⌺ ⱖ 0, 共3.4兲
V = V0 + Vel + Vin . 共2.1兲
which is conveniently rewritten by transforming to the
Here, V0 is a reference volume, determined by the entropy Helmholtz free energy F共␪ , Vel , Nv兲 = U共S , Vel , Nv兲 − ␪S and
共or temperature兲. For simplicity, we neglect thermoelastic ef- then performing the partial differentiations

冉 冊 冉 冊 冉 冊
fects and assume that V0 is just a constant. The elastic vol-
⳵F ⳵F ⳵F
ume, Vel, is associated with reversible changes in the elastic + S ␪˙ + + p V̇el + + v0 p Ṅv = − ␪⌺ ⱕ 0.
energy. An increment ␦Vel is a change in the total volume at ⳵␪ ⳵ Vel ⳵ Nv
fixed entropy and fixed Nv; that is, it takes place with no 共3.5兲
change in the internal state of the system. Our central as-
sumption is that the inelastic volume associated with the va- Here we have used Eqs. 共2.1兲 and 共2.2兲.
cancies is simply Vin = v0Nv, where v0 is the effective volume In the spirit of Coleman and Noll 关12兴, we recognize that
of a vacancy. Vel and Vin are independently “variable” but not this expression consists of three separate, independent in-
independently “controllable.” In equilibrium, Vel is con- equalities because, as discussed above, the time derivatives
trolled directly by the pressure. On the other hand, Nv is a are unconstrained by each other. We satisfy the first inequal-
“hidden” internal variable so long as the system is not ity by identifying S = −⳵F / ⳵␪, thus recovering the familiar
coupled to a chemical-potential reservoir that controls the thermodynamic relation. The second inequality usually is
number of vacancies. Nevertheless, we must treat Nv as hav- satisfied by identifying p = −⳵F / ⳵Vel, i.e., using the equilib-
ing its own dynamics and being able at any time to change in rium relation between p and Vel. However, if we were inter-
ways that are not directly constrained by concurrent changes ested in thermoviscoelastic effects, then we would satisfy
in Vel or the entropy. Specifically, we assume that the rate of this inequality by writing a dissipative equation of motion for
inelastic volume deformation is Vel of the form

031131-2
NONEQUILIBRIUM… . I. INTERNAL DEGREES… PHYSICAL REVIEW E 80, 031131 共2009兲

V̇el = − ␥el 冉 冊
⳵ F̃el
⳵ Vel ␪,Nv
; F̃el = F + pVel , 共3.6兲
IV. STATISTICAL THEORY
The basic statistical statement of the second law is that the
system as a whole, including any thermal reservoir to which
the subsystem of primary interest may be coupled, must
where ␥el is a non-negative dissipation coefficient. In this move toward states of higher probability, i.e., to states of
way, we would account for the energy dissipation that ac- higher entropy. Although the Coleman-Noll procedure as-
companies the relaxation of the “viscous pressure” p signs no a priori statistical significance to the entropy, this
+ ⳵F / ⳵Vel. Since thermoviscoelasticity is not the topic of pri- principle lies at its heart. In their formulation, however, the
mary interest here, we simply adopt the equilibrium relation principal focus is on spatial heterogeneities. The entropy of
from here on. the system as a whole increases as heat flows between spa-
Finally, using Eq. 共2.2兲, we write the third inequality in tially separated elements, each of which is always in a state
the form of local equilibrium with its own local energy, entropy, and
temperature. It is conceptually easy, albeit mathematically

− 冉 冊
⳵ F̃v
⳵ Nv ␪,Vel
Ṅv ⱖ 0, F̃v = F + pv0Nv . 共3.7兲
more complicated, to add spatial heterogeneity to the va-
cancy model. We do not do this explicitly here but, never-
theless, anticipate the need to reinterpret our uniform model
as just one element of a larger, spatially inhomogeneous,
This is a specific realization of the Clausius-Duhem in- coarse-grained system.
equality, Eq. 共3.3兲. The same result was obtained in Rice’s Our strategy is to start with a statistical definition of en-
classic 1971 paper 关14兴 where, however, the free energy F tropy and to introduce a thermal reservoir, but otherwise to
was assumed to be a function of the total deformation rather stay as close as possible to the conventional analysis. There-
than the elastic part alone, so that the pressure was missing fore, in analogy to Eq. 共3.1兲, we begin by writing the first
in the expression for F̃v. The inequality of 共3.7兲 is satisfied law in the form
by
− pV̇ = U̇ + U̇R , 共4.1兲

Ṅv = − ␥v 冉 冊
⳵ F̃v
⳵ Nv ␪,Vel
, 共3.8兲
where UR = UR共SR兲 is the energy of the reservoir as a function
of its entropy SR. Similarly, in analogy to Eqs. 共3.2兲 and
共3.3兲, the second law is

where ␥v is again a non-negative dissipation coefficient. Ṡneq + ṠR ⱖ 0, 共4.2兲


More generally, any monotonically increasing function of Nv where Sneq is the entropy of a system that is not necessarily
that vanishes where ⳵F̃v / ⳵Nv = 0 can be used on the right- in thermal equilibrium.
hand side of Eq. 共3.8兲. The main question is what to use for Sneq. We propose,
Ever since Coleman and Noll introduced their axiomatic with several conditions to be listed below, that the correct
version of thermomechanics, physicists have been impressed choice of this entropy has the form
by its mathematical elegance, but have worried that it might Sneq共U,V,兵⌳␣其兲 = ln ⍀共U,V,兵⌳␣其兲, 共4.3兲
be incomplete because it does not start with a statistical defi-
nition of entropy. It is not clear what statistical interpretation where ⍀共U , V , 兵⌳␣其兲 is a constrained measure of the number
of the entropy S is implied by the preceding equations or, of states of the system with energy U, volume V, and speci-
conversely, how the internal energy U might depend on S. fied values of a set of internal variables 兵⌳␣其. The ⌳␣’s are
For example, it is not obvious in a conventional formulation out of equilibrium if their values are not the ones that maxi-
how to evaluate the free energy F̃v in Eq. 共3.7兲. More impor- mize Sneq. When all of them do maximize Sneq, i.e., when
tantly, the Coleman-Noll postulates operationally define a ⌳␣ = ⌳␣eq, then we require that the equilibrium entropy
temperature as well as an entropy. In the analysis presented Seq共U , V兲 = ln ⍀共U , V兲 be accurately approximated by
here, we are looking ahead to an effective temperature theory 1 1
in which there will be two different temperatures—a situa- Seq共U,V兲 ⬇ Sneq共U,V,兵⌳␣eq其兲. 共4.4兲
V V
tion that seems to be beyond the scope of the conventional
axiomatic formulation. This approximation must become an equality in the thermo-
Lastly, we note that the axiomatic approach makes no dynamic limit, V → ⬁. In general, Sneq ⬍ Seq because the con-
mention of a thermal reservoir. It seems to us that any theory strained entropy Sneq counts fewer states than the uncon-
of this kind ought to include a specific mechanism by which strained entropy Seq. We require that the difference between
the temperature is controlled. If that mechanism involves these quantities per unit volume becomes negligibly small as
coupling to a thermal reservoir, then the theory ought to pre- 兵⌳␣其 → 兵⌳␣eq其 and as the size of the system becomes indefi-
dict the rate at which heat is flowing between the system and nitely large. Without this condition, we would not have a
the reservoir. Conversely, the theory should be able to predict single, well-defined entropy upon which to base a self-
what happens if that flow is constrained, as in an adiabatic consistent thermomechanical theory.
process. But any coupling to a reservoir disappears when Q Validity of Eq. 共4.4兲 therefore requires that three condi-
is eliminated in Eq. 共3.4兲. tions be satisfied:

031131-3
ERAN BOUCHBINDER AND J. S. LANGER PHYSICAL REVIEW E 80, 031131 共2009兲

共1兲 The set of variables 兵⌳␣其 must be subextensive. If where ␪R = ⳵UR / ⳵SR is the reservoir temperature. We also use
there are N␣ such variables and there are N total degrees of Eq. 共2.2兲 to eliminate V̇in in favor of Ṅv. The result is
freedom in the system, then N␣ / N must vanish in the ther-
modynamic limit. Otherwise, the variations of the ⌳␣’s
would produce an extensive entropic correction to the equi- 冉 冊
W共p,Nv,Ṅv兲 − 1 −

U̇ ⱖ 0,
␪R R
共4.10兲
librium free energy and Eq. 共4.4兲 would not be correct. More
explicitly, note that we can compute Seq by integrating over where
each of the variables ⌳␣ in ⍀共U , V , 兵⌳␣其兲, obtaining a cor-
rection to ln ⍀ proportional to N␣. That correction must be
negligibly small compared to Sneq共U , V , 兵⌳␣eq其兲, which is pro-
冋 冉 冊 册
W共p,Nv,Ṅv兲 = − pv0 +
⳵U
⳵ Nv Sneq,Vel
Ṅv 共4.11兲

portional to N.
共2兲 We must be working in the quasiequilibrium limit, is the rate at which inelastic work is done on the system
where all the unconstrained degrees of freedom have rapidly minus the rate at which energy is stored by the vacancies. As
come to equilibrium and where their fluctuations have been will be seen, W is a dissipation rate that appears in various
accounted for in computing Sneq. forms throughout this series of papers.
共3兲 Condition 共1兲 requires that the ⌳␣’s be coarse-grained The appearance of U̇R in this inequality is important be-
variables. If there is only a subextensive number of these cause we control the temperature of the system by control-
variables, then each of them must be a sum over a statisti- ling the reservoir temperature. Thus the inequality in Eq.
cally large number of degrees of freedom. That is, the ⌳␣ 共4.10兲 must be satisfied for arbitrary variations of UR, inde-
themselves must be extensive. 共Of course, nothing prevents pendent of whatever else is happening in the system. We also
us from interpreting them as spatial averages over a macro- must satisfy this inequality for arbitrary variations of Nv. For
scopically large system.兲 It then follows that the entropies example, the vacancy population could be relaxing toward an
associated with each of the ⌳␣ must be included explicitly in equilibrium value while UR remains constant. Therefore, in
Sneq. For example, our single internal variable Nv describes the spirit of Coleman and Noll, we argue that the only way to
an extensive population of vacancies. The associated en- satisfy this combined inequality for all possible variations of
tropy, i.e., the logarithm of the number of ways in which the the system is to enforce two separate, independent inequali-
Nv vacancies can be arranged in the volume V, must be con- ties
tained in Sneq.
Accordingly, the entropy appearing in Eq. 共4.2兲 is W共p,Nv,Ṅv兲 ⱖ 0 共4.12兲

Sneq共U,Vel,Nv兲 = ln ⍀共U,Vel,Nv兲. 共4.5兲 and

For reasons discussed in Sec. II, we replace V by Vel as an


independent argument of Sneq. We then invert Sneq共U , Vel , Nv兲
冉 冊
− 1−

U̇ ⱖ 0.
␪R R
共4.13兲
to obtain U共Sneq , Vel , Nv兲. We identify
The first of these relations is essentially identical to the

冉 冊⳵U
⳵ Sneq Vel,Nv
=␪ 共4.6兲
Clausius-Duhem inequality in Eq. 共3.7兲. The differences are
that we have derived Eq. 共4.12兲 from statistical first prin-
ciples rather than postulated it and that we know exactly
and, as stated following Eq. 共3.6兲, we use the equilibrium what energy and entropy are involved in it.
thermodynamic relation for the pressure The second inequality is satisfied by requiring that U̇R be

冉 冊
a function of ␪ that changes sign only when ␪ = ␪R; therefore
⳵U we write
= − p. 共4.7兲
⳵ Vel Sneq,Nv
− U̇R = A共␪, ␪R兲共␪R − ␪兲 ⬅ Q, 共4.14兲
Therefore,
where A共␪ , ␪R兲 is a non-negative function of its arguments.
U̇ = − pV̇el + 冉 冊
⳵U
⳵ Nv Sneq,Vel
Ṅv + ␪Ṡneq . 共4.8兲
Here, Q has the same meaning that it had in Eq. 共3.1兲—the
rate at which heat is flowing into the system, in this case,
from the reservoir—but now, Q is a well-defined function of
The first law, Eq. 共4.1兲, becomes ␪ and Eq. 共4.14兲 is an equation, not an inequality. With this
definition of Q, Eq. 共4.9兲 becomes
− pV̇in − 冉 冊
⳵U
⳵ Nv Sneq,Vel
Ṅv − U̇R = ␪Ṡneq , 共4.9兲 ␪Ṡneq = W共p,Nv,Ṅv兲 + Q. 共4.15兲

where we have used V̇ = V̇el + V̇in to eliminate V̇el. V. SPECIFICS OF THE VACANCY MODEL
At this point, we depart from the strategy that led to Eq.
Because we have an unambiguous definition of the total
共3.4兲. Instead of eliminating the coupling to the thermal res-
entropy and because we know that the entropy of the vacan-
ervoir as was done there, we use Eq. 共4.9兲 to evaluate Ṡneq cies must be included in it, we can write Sneq共U , Vel , Nv兲 in
in the second law, Eq. 共4.2兲, and we identify ṠR = U̇R / ␪R, the form

031131-4
NONEQUILIBRIUM… . I. INTERNAL DEGREES… PHYSICAL REVIEW E 80, 031131 共2009兲

Sneq共U,Vel,Nv兲 = S0共Nv兲 + S1共U1兲 VI. EQUATION OF MOTION FOR Nv

= S0共Nv兲 + S1关U − e0Nv − Uel共Vel兲兴. At this point in the analysis, the standard procedure is to
postulate a general form for an equation of motion for Nv and
共5.1兲
to use the Clausius-Duhem inequality in Eq. 共5.8兲 to con-
Equivalently, we can invert this relation and write it as an strain the parameters that appear in it. In the present case,
expression for the internal energy U, there is no reason why Nv should do anything more compli-
cated than relax toward a stable equilibrium value. There-
U共Sneq,Vel,Nv兲 = U0共Nv兲 + U1共S1兲 + Uel共Vel兲 fore, for small departures from equilibrium, we write
= e0Nv + U1关Sneq − S0共Nv兲兴 + Uel共Vel兲.
␶0Ṅv = ˜⌫共Nv兲关Neq
v 共␪,p兲 − Nv兴, 共6.1兲
共5.2兲
where ␶0 is a time scale, ˜⌫ is a positive, dimensionless rate
Here, U0共Nv兲 is the energy of the vacancies, e0 is the forma- factor that we anticipate will be a function of Nv 共as well as
tion energy of a vacancy, S0共Nv兲 is the entropy of the vacan- ␪ and p兲, and Neq
v 共␪ , p兲 is the equilibrium value of Nv at the
cies, Uel共Vel兲 is the elastic energy, and S1 and U1 are, respec- given temperature and pressure. A convenient alternative
tively, the entropy and energy of all the other configurational, form of this equation is

冋 册
kinetic, and vibrational degrees of freedom in the system.
The structure of these relations, i.e., the arguments of U1 and
S1 in their second versions, describes the way the energy and
entropy are shared between the vacancies and the other de-
␶0
Ṅv ˜
Nv
Neq
Nv
Nv
= ⌫共Nv兲 v − 1 ⬵ − ˜⌫共Nv兲ln eq .
Nv
冉 冊 共6.2兲

grees of freedom. Note that the total entropy and energy in Having no information about nonlinear corrections to Eq.
Eqs. 共5.1兲 and 共5.2兲 are assumed to have very simple forms. 共6.1兲, we can use the second expression on the right-hand
For example, we have omitted a standard thermoelastic term side of Eq. 共6.2兲 just as well as the first and will do so from
proportional to S1Vel in Eq. 共5.2兲. here on. Other nonlinear equations of motion for Nv can
For specificity, we assume that the vacancies are very easily be incorporated into this analysis when justified by
dilute so that some physical mechanism.

冉 冊
To satisfy the inequality in Eq. 共5.8兲, we require that both
Nv
S0共Nv兲 = − Nv ln + Nv , 共5.3兲 ⳵Gv / ⳵Nv and the expression for Ṅv on the right-hand side of
N0 either Eq. 共6.1兲 or Eq. 共6.2兲 vanish at the same point, i.e., at
v 共␪ , p兲. Thus, Nv 共␪ , p兲 is the solution of
Nv = Neq eq
where N0 is the number of sites at which vacancies might
occur. Then, using Eq. 共5.2兲, we find that
冉 冊 ⳵ Gv
冉 冊
= e0 + pv0 + ␪ ln
Neq
v

冉 冊 冉 冊
=0 共6.3兲
⳵U d Nv ⳵ Nv ␪,p,Nv=Neq N0
= 关U0共Nv兲 − ␪S0共Nv兲兴 = e0 + ␪ ln . v
⳵ Nv Sneq,Vel dNv N0 and the equilibrium number of vacancies is proportional to a
共5.4兲 Boltzmann factor
If we write
v 共␪,p兲 = N0 exp −
Neq 冉 e0 + pv0

. 共6.4兲

冉 冊

⳵U
= ␪Ṡneq ⬅ CV␪˙ 共5.5兲
⳵t Vel,Nv
The inequality in Eq. 共5.8兲 is always satisfied so long as Ṅv is
a monotonically decreasing function of Nv, which is required
and interpret the extensive quantity CV to be the heat capac- for dynamic stability, and is true for both Eqs. 共6.1兲 and
ity at constant volume, then Eq. 共4.15兲 becomes 共6.2兲.

冋 冉 冊册 冉 冊
Retaining Q explicitly in this analysis has the added ben-
Nv ⳵ Gv efit of allowing us to deduce an expression for the rate factor
CV␪˙ + e0 + pv0 + ␪ ln Ṅv = CV␪˙ + Ṅv = Q,
N0 ⳵ Nv ␪,p ˜⌫共N 兲. The equation of motion for N , as shown in Eq. 共6.1兲,
v v
共5.6兲 is a detailed-balance relation in which the vacancy creation
v 共␪ , p兲. Therefore, according to Eq.
rate is proportional to Neq
where the vacancy-related Gibbs free energy Gv is 共6.4兲, the creation rate automatically contains the appropriate
Gv共␪,p,Nv兲 = e0Nv − ␪S0共Nv兲 + pv0Nv . 共5.7兲 Arrhenius activation factor and ˜⌫ / ␶0 can be interpreted as a
dimensionless attempt frequency or, equivalently, a noise
The Clausius-Duhem inequality, Eq. 共4.12兲, is strength.

冉 冊
Our experience with the STZ theory of plasticity leads us
⳵ Gv
− Ṅv ⱖ 0. 共5.8兲 to write ˜⌫ as the sum of two terms
⳵ Nv ␪,p
˜⌫ = ␳共␪兲 + ⌫共N 兲, 共6.5兲
Clearly, this term in Eq. 共5.6兲 is a non-negative rate of heat v

production associated with the relaxation of the internal vari- where ␳共␪兲 is the strength of the noise generated solely by
able Nv toward an equilibrium value. thermal fluctuations in the absence of mechanical deforma-

031131-5
ERAN BOUCHBINDER AND J. S. LANGER PHYSICAL REVIEW E 80, 031131 共2009兲

tion and ⌫共Nv兲 is the noise strength associated with irrevers- as the external work is converted into internal heat. In the
ible deformations, i.e., with nonzero Ṅv. For present pur- following paper, we will encounter an intermediate case in
poses, we could simply set ␳共␪兲 = 1 and let ␶0 be temperature which A is small but nonzero.
dependent, but we will need the explicit factor ␳共␪兲 for dis-
cussing glassy systems in the following papers. VII. CONCLUDING REMARKS
A hypothesis 共originally due to Pechenik 关15兴兲 that has
Although the statistical analysis described here has
worked well for the STZ theory is that ⌫ is proportional to
been developed primarily for use in the effective-tempera-
the total rate per vacancy at which heat is generated by the
ture theory of amorphous materials presented in the follow-
work done on the system. In the present case, this means that
ing paper 关6兴, the present results already point toward some
␪0 general conclusions. We have shown that the Clausius-
N ⌫共Nv兲 = ␪Ṡneq − Q = W, 共6.6兲
␶0 v Duhem entropy-production inequality, when applied to the
dynamics of internal degrees of freedom, can be derived di-
where ␪0 is an energy and W is the same non-negative dis- rectly from a statistical interpretation of the second law of
sipation rate that was defined in Eq. 共4.11兲. With this as- thermodynamics—but only if the conditions listed following
sumption and with the second form of Ṅv given in Eq. 共6.2兲, Eq. 共4.4兲 are satisfied. Perhaps the most important of these
Eq. 共5.6兲 becomes conditions is that the internal variables must be a small set of

冉 冊
extensive quantities in order that the statistical entropy be
Nv
− ␪Nv ln2 共␳ + ⌫兲 + ␪0⌫Nv = 0. 共6.7兲 well defined for nonequilibrium situations. This condition, in
Neq
v turn, means that entropies as well as energies associated with
Solving for ␳ + ⌫ 共a necessarily non-negative noise strength兲, the internal variables must be included in any dynamical
we find description of the system.
The quantity appearing in the Clausius-Duhem inequality
˜⌫ = ␳ + ⌫ = ␳共␪兲 in Eq. 共5.8兲 is interpreted by Lubliner 关17兴 as the “dissipation
. 共6.8兲
1 − 共␪/␪0兲ln2共Nv/Neq
v 兲
associated with the internal variables and their conjugate
forces.” Our analysis suggests a sharper and more physically
Thus, the mechanically generated noise enhances the rate intuitive interpretation that the rate of energy dissipation
factor, possibly quite substantially. The feature that ˜⌫ di- W is the difference between the inelastic power −pV̇in
verges when Nv is sufficiently far from its equilibrium value
= −pv0Ṅv and the rate of change of the free energy U0共Nv兲
simply means that the system is dynamically driven away
− ␪S0共Nv兲 that is stored in the internal degrees of freedom.
from such values of Nv and that ˜⌫ remains positive at all One example of this difference occurs in Rice’s 1971 pa-
times. per 关14兴, where he suggests that his internal variables repre-
Putting these pieces of the theory together, we have sent an extensive set of slips on slip planes distributed
throughout a polycrystalline material. However, he does not
Ṅv ␳共␪兲 ln共Nv/Neq
v 兲
=− . 共6.9兲 calculate the number of ways in which the total slip can be
Nv ␶0 1 − 共␪/␪0兲ln 共Nv/Neq
2
v 兲 realized as the sum of many individual slips and therefore
does not include the entropy associated with his internal “av-
Equation 共5.6兲 becomes

冉 冊
eraging variables” in dynamical formulas analogous to Eqs.
Nv 共5.7兲 and 共5.8兲.
CV␪˙ + ␪ ln eq Ṅv = Q = A共␪, ␪R兲共␪R − ␪兲. 共6.10兲 These issues persist in the more recent literature and, in
Nv
our opinion, are quite serious. For example, Anand and Su
The combination of Eqs. 共6.9兲 and 共6.10兲 allows us to 关18兴 implement something like Rice’s picture of frictional
compute time-dependent functions ␪共t兲 and Nv共t兲 given any slips on multiple slip planes by using a phenomenological,
driving force p共t兲 and reservoir temperature ␪R共t兲. The sim- nonlinear, rate-dependent relation between local flow and re-
plest case is the limit in which the coupling to the reservoir is solved stresses. Some memory of past deformation is carried
so strong that ␪ = ␪R and the heat capacity of the reservoir is by a “plastic volumetric strain” and by a related cohesion
so large that ␪R remains a constant independent of how much parameter that appears in the flow equation, but these are
heat is flowing to or from the system. This assumption, that scalar quantities that cannot contain information about the
the temperature is fixed by coupling to the reservoir, is im- directional history of shear flow. There is no dynamical yield
plicit in most thermodynamic theories, but it is actually a bit stress as in STZ theory, nor—so far as we can tell—is there
subtle. The quantity Q on the right-hand side of Eq. 共6.10兲 is any way of using the theory to predict what happens when
undefined in the limit A → ⬁, ␪ → ␪R, which means that it can the loading stresses are removed or reversed, partly because
assume whatever value is needed in order to keep ␪ = ␪R the plastic volumetric strain and the cohesion parameter are
= constant. With ␪˙ = 0, Eq. 共6.10兲 is just the Clausius-Duhem scalars, but more importantly because neither are properly
inequality again, now telling us that Q ⬍ 0 and that—as in constituted internal state variables. The STZ theory has been
the Kelvin-Planck statement of the second law 关16兴—we are developed explicitly to overcome such difficulties. It is the
not allowed to convert heat directly into work. topic of the third paper in this series 关7兴.
Alternatively, suppose that the process is adiabatic, i.e., A related question, which is sometimes raised but not an-
A = 0. Then Eq. 共6.10兲 determines how the temperature rises swered in the conventional literature, is what happens when

031131-6
NONEQUILIBRIUM… . I. INTERNAL DEGREES… PHYSICAL REVIEW E 80, 031131 共2009兲

the internal degrees of freedom do not relax in a simple flowing steady states. This behavior is discussed in detail in
manner. In the present case, our irreversible process can be the third paper in this series 关7兴. More generally, we know
described by a variational principle; that is, our single inter- that no such energy-minimization principles exist for many
nal variable Nv moves downhill in the one-dimensional free- open situations, where the system is being persistently driven
energy landscape defined by Gv共Nv兲 in Eq. 共5.7兲. This picture away from equilibrium and where there are multiple,
is generalized in the conventional literature by assuming that coupled, internal state variables. It seems to us that it will be
a system with multiple internal variables moves downhill in hard to predict a form for a generalized Clausius-Duhem
a multidimensional inelastic potential. The resulting fluxes inequality without starting from a first-principles, fully sta-
obey what is called a “normality condition” or sometimes a tistical and dynamical description of such systems.
“generalized normality condition” 关8兴 because they are as-
sumed to be perpendicular to surfaces of constant 共general-
ized兲 potential. ACKNOWLEDGMENTS
We already know that the picture cannot be so simple for
the STZ theory, where increasing shear stress drives the sys- J.S.L. acknowledges support from U.S. Department of
tem through an exchange of stability between jammed and Energy Grant No. DE-FG03-99ER45762.

关1兴 M. L. Falk and J. S. Langer, Phys. Rev. E 57, 7192 共1998兲. Cambridge, England, 2004兲.
关2兴 E. Bouchbinder, J. S. Langer, and I. Procaccia, Phys. Rev. E 关11兴 H. Xiao, O. T. Bruhns, and A. Meyers, Acta Mech. 182, 31
75, 036107 共2007兲. 共2006兲.
关3兴 E. Bouchbinder, J. S. Langer, and I. Procaccia, Phys. Rev. E 关12兴 B. D. Coleman and W. Noll, Arch. Ration. Mech. Anal. 13,
75, 036108 共2007兲. 167 共1963兲.
关4兴 E. Bouchbinder, Phys. Rev. E 77, 051505 共2008兲. 关13兴 B. D. Coleman and M. E. Gurtin, J. Chem. Phys. 47, 597
关5兴 J. S. Langer, Phys. Rev. E 77, 021502 共2008兲. 共1967兲.
关6兴 E. Bouchbinder and J. S. Langer, following paper, Phys. Rev. 关14兴 J. R. Rice, J. Mech. Phys. Solids 19, 433 共1971兲.
E 80, 031132 共2009兲. 关15兴 J. S. Langer and L. Pechenik, Phys. Rev. E 68, 061507 共2003兲;
关7兴 E. Bouchbinder and J. S. Langer, this issue, Phys. Rev. E 80, L. Pechenik, ibid. 72, 021507 共2005兲.
031133 共2009兲. 关16兴 For example, see the historical discussion on M. Zemansky,
关8兴 J. Lubliner, Plasticity Theory 共Macmillan, New York, 1990兲. Heat and Thermodynamics, 3rd ed. 共McGraw-Hill, New York,
关9兴 G. A. Maugin, The Thermomechanics of Nonlinear Irrevers- 1951兲, p. 147.
ible Behaviors 共World Scientific, Singapore, 1999兲. 关17兴 J. Lubliner, Int. J. Non-Linear Mech. 7, 237 共1972兲.
关10兴 S. Nemat-Nasser, Plasticity 共Cambridge University Press, 关18兴 L. Anand and C. Su, J. Mech. Phys. Solids 53, 1362 共2005兲.

031131-7

Вам также может понравиться