Philosophy of Science, 75 (December 2008) pp. 682–696. 0031-8248/2008/7505-0016$10.00 Copyright 2008 by the Philosophy of Science Association. All rights reserved. 682 Boltzmann, Gibbs, and the Concept of Equilibrium David A. Lavis†‡ The Boltzmann and Gibbs approaches to statistical mechanics have very different definitions of equilibrium and entropy. The problems associated with this are discussed, and it is suggested that they can be resolved, to produce a version of statistical me- chanics incorporating both approaches, by redefining equilibrium not as a binary prop- erty (being/not being in equilibrium) but as a continuous property (degrees of equilibrium) measured by the Boltzmann entropy and by introducing the idea of thermodynamic-like behavior for the Boltzmann entropy. The Kac ring model is used as an example to test the proposals. 1. Introduction. The object of study in the Gibbs formulation of statistical mechanics is an ensemble of systems, and the Gibbs entropy is a functional of the ensemble probability density function. Equilibrium is defined as the state where the probability density function is a time-independent solution of Liouville’s equation. The development of this approach has been very successful, but its extension to nonequilibrium presents con- tentious problems.1 To implement the Boltzmann approach (Lebowitz 1993; Bricmont 1995; Goldstein 2001) the phase space is divided into a set of macrostates. The Boltzmann entropy at a particular point in phase space is a measure of the volume of the macrostate in which the phase point is situated. The system is understood to be in equilibrium when the phase point is in a †To contact the author, please write to: Department of Mathematics, King’s College, London WC2R 2LS, U.K.; e-mail: david.lavis@kcl.ac.uk. ‡This is a modified version of the paper presented at the twentieth biennial meeting of the Philosophy of Science Association, Vancouver, Canada, November 2006. I am grateful to the organizers for giving me the opportunity to contribute to the conference and to Roman Frigg for many useful discussions. 1. The most developed program for doing this is that of the Brussels-Austin School of the late Ilya Prigogine. (For a comprehensive review, see Bishop 2004.) THE CONCEPT OF EQUILIBRIUM 683 particular region of phase space. The entropy and equilibrium are thus properties of a single system. The purpose of this article is to attempt to produce a synthesis of the Gibbs and Boltzmann approaches, which validates the Gibbs approach, as currently used in ‘equilibrium’ statistical mechanics and solid state physics, while at the same time endorsing the Boltzmann picture of the time evolution of entropy, including ‘the approach to equilibrium’. In order to do this we need to resolve in some way three questions, to which the current versions of the Gibbs and Boltzmann approaches offer ap- parently irreconcilable answers: (a) What is meant by equilibrium? (b) What is statistical mechanical entropy? and (c) What is the object of study? The attempt to produce conciliatory answers to a and b will occupy most of this article. However, we shall at the outset deal with c. As indicated above, ensembles are an intrinsic feature of the Gibbs approach (see, e.g., Prigogine 1994, 8). However we follow the Neo-Boltzmannian view of Lebowitz (1993, 38) that we “neither have nor do we need ensembles.” The object of study in statistical mechanics is a single system and all talk of ensembles can be understood as just a way of giving a relative frequency flavor to the probabilities of events occurring in that system. We now describe briefly the dynamics and thermodynamics of the sys- tem together with the statistical approach of Gibbs. The Boltzmann ap- proach is described in greater detail in Section 2. At the microscopic (dynamic) level, the system (taken to consist of N microsystems) is supposed to have a one-to-one autonomous dynamics , on its phase space . The system is reversible, meaning thatx r f x Gt N there exists a self-inverse operator on , such that ′J x � G f x p x rN t . Then . On the subsets of there is a′ �1f Jx p Jx f p (f ) p Jf J Gt �t t t N sigma-additive measure , such that (a) is finite, (b) is absolutelym m(G ) mN continuous with respect to the Lebesque measure on , and (c) isG mN preserved by ; that is , and measurable . This{f } m(fg) p m(g) G t g O Gt t N means that there will be no convergence to an attractor (which could in the dynamic sense be taken as an equilibrium state). At the phenomenological (thermodynamic) level, equilibrium is a state in which there is no perceptible change in macroscopic properties. It is such that a system (a) either is or is not in equilibrium (a binary property), (b) never evolves out of equilibrium, and (c) when not in equilibrium evolves toward it. At the statistical level (in the Gibbs approach), the phase point x � is distributed according to a probability density function in-G r(x; t)N 684 DAVID A. LAVIS variant under , meaning that it is a solution of Liouville’s equation.{f }t At equilibrium the Gibbs entropy is the functional S [r] :p �k r(x) ln [r(x)]dm (1)G B � GN of a time-independent probability density function. Problems arise when an attempt is made to extend the use of (1) to nonequilibrium situations, which are now perceived as being those where is time dependent.r 2. The Macroscopic Level–Boltzmann Approach. Here we must introduce a set of macroscopic variables at the observational level that give moreY detail than the thermodynamic variables, and a set of macrostates {m} defined so that: (i) every is in exactly one macrostate denoted byx � GN , (ii) each macrostate corresponds to a unique set of values for , (iii)m Yx is invariant under all permutations of the microsystems, and (iv) themx phase points and are in macrostates of the same size.2 The Boltzmannx Jx entropy, which is a function on the macrostates, and consequently also a function on the phase points in , isGN S (x) p S (m ) :p k ln [m(m )]. (2)B B x B x This is, of course, an extensive variable, and the quantity of interest is the dimensionless entropy per microsystem , which for thes :p S /(Nk )B B B sake of brevity we shall refer to as the Boltzmann entropy. Along a tra- jectory will not be a monotonically increasing function of time. RathersB we should like it to exhibit thermodynamic-like behavior, defined in an informal (preliminary) way as follows: Definition (TL1): The evolution of the system will be thermodynamic- like if spends most of the time close to its maximum value, fromsB which it exhibits frequent small fluctuations and rarer large fluctu- ations. This leads to two problems which we discuss in Sections 2.1 and 2.2. 2.1. How Do We Define Equilibrium? Is it possible to designate a part of as the equilibrium state? On the grounds of the system’s reversibilityGN and recurrence we can, of course, discount the possibility that such a 2. When is the direct product of the configuration space and the momentum space,GN and the macrostates are generated from a partition of the one-particle configuration space, the points and are in same macrostate. However, for discrete-time systemsx Jx phase space is just configuration space and the points and are usually in differentx Jx macrostates. THE CONCEPT OF EQUILIBRIUM 685 region is one which, once entered by the evolving phase point, will not be exited. As was well understood by both Maxwell and Boltzmann, equilibrium must be a state which admits the possibility of fluctuations out of equilibrium. Lebowitz (1993, 34) and Goldstein (2001, 8) refer to a particular macrostate as the equilibrium macrostate, and the remark by Bricmont (1995, 179), that “by far the largest volumes [of phase space] correspond to the equilibrium values of the macroscopic variables (and this is how ‘equilibrium’ should be defined)” is in a similar vein. So is there a single equilibrium macrostate? If so it must be that in which the phase point spends more time than in any other macrostate and, if the system were ergodic, it would be the largest macrostate (see Sec. 2.2),m Max with largest Boltzmann entropy. There is one immediate problem asso- ciated with this. Suppose we consider the set of entropy levels ,{s (m)}B . Then, as has been shown by Lavis (2005) for the baker’s gas,G m O GN associated with these levels there may be degeneracies , such that, forq(m) some with , . The effect of this ism m(m) ! m(m ) m(m)q(m) 1 m(m )Max Max that the entropy will be likely, in the course of evolution, to spend more time in a level less than the maximum (see Lavis 2005, Figs. 4 and 5). Another example, which has been used to discuss the evolution of Boltzmann’s entropy (see Bricmont 1995, App. 1) and which we shall use as an illustrative example in this article, is the Kac ring model (Kac 1959, 99).3 Macrostates in this model can be indexed ,k p 0, 1, . . . , (1/2)N where is the macrostate with up spins and downm (1/2)N � k (1/2)N � kk spins, giving N! Nm(m ) p with m(G ) p 2 . (3)k N([N/2] � k)!([N/2] � k)! Then, of course, with monotonically decreasing withm p m m(m )Max 0 k increasing . But although the maximum macrostate is unique,FkF , and the entropy level corresponding to the largestq(m ) p 2 G k ( 0k volume of is given, for , by the macrostate pair . It may beG N 1 2 mN �1 supposed that this question of degeneracy is an artifact of relevance only for small . It is certainly the case, both for the baker’s gas and KacN ring, that, if is a macrostate which maximizes , although′ ′ ′m m(m )q(m ) 3. This model consists of up or down spins distributed equidistantly around a circle.N Randomly distributed at some of the midpoints between the spins are spin flippers.m The dynamics consists of rotating the spins (but not the spin-flippers) one spin-site in the clockwise direction, with spins changing their direction when they pass through a spin flipper. consists of the points, corresponding to all combinations of the twoNG 2N spin states, and is decomposable into dynamically invariant cycles. If is even, them parity of is preserved along a cycle which has maximum size . If is odd, thek N m parity of alternates with steps along a cycle and the maximum cycle size is .k 2N 686 DAVID A. LAVIS , , as . So, maybe the union′ ′ ′m(m )q(m ) r� m(m ) s (m ) r s (m ) N r �Max B B Max of and all the equally sized macrostates with measure can be′m m(m )Max used as the equilibrium state. To test this possibility take the Kac ring and consider the partial sum k k m(B(N, k)) :p m(m ) p m(m ) � 2 m(m ), (4)� �j 0 j jp�k jp1 where is the union of all with . Then it is not difficultB(N, k) m j � [�k, k]j to show that, for fixed ,k m(B(N, k)) r 0, as N r �. (5) m(G )N The proportion of contained within the band of macrostatesG B(N, k)N decreases with .4 If we want the equilibrium region to satisfy the propertyN described in the quotation given above from Bricmont (1995), then this cannot be done by designating a band of a fixed number of macrostates in this way. To ensure that as becomes large “by far the largest volumesN [of phase space] correspond to the equilibrium values of the macroscopic variables” we need to choose a value of increasing with . Thus fork N example to create an ‘equilibrium band of macrostates’ containing 99.999% of we must choose , for ; , forG k p 22 N p 100 k p 70 N pN ; and for . An entropy profile for this last case1,000 k p 221 N p 10,000 is shown in Figure 1. The putative equilibrium state, representing 99.999% of is given by the region bounded by the horizontal lines in theG10,000 figure. If the system were ergodic (like the baker’s gas) we would expect the system to spend almost all of the 10,000 time steps in this region, whereas in this particular simulation only 4,658 steps satisfied this con- dition. This is to be expected as the Kac ring is not ergodic but has an ergodic decomposition into cycles of which this figure represents half a cycle.5 The proportion of ‘equilibrium states’ will differ between cycles. So the problems with defining an equilibrium region are: (i) Just choosing the largest macrostate as the equilibrium region does not guarantee that this region becomes an increasing proportion of phase space as increases. In fact the reverse is the case for theN Kac ring. 4. For later reference we note that, in particular, this result applies to withk p 0 for large . The proportion of phase space in the largest ma-�m(m )/m(G ) � 2/Np N0 N crostate decreases with .N 5. This is because is odd. The second half of the cycle in which the spins are reversem has an identical entropy profile. THE CONCEPT OF EQUILIBRIUM 687 Figure 1. Half a cycle of the plot of against time, for a Kac ring of 10,000 spins.sB There are spin-flippers, distributed randomly around the ring, and the evo-m p 509 lution is initiated in the low-entropy state . The upper and lower horizontalk p 4,993 lines correspond, respectively, to and .s (m ) p 0.69266408 s (m ) p 0.69168704B Max B 221 (ii) Any choice of a collection of macrostates to represent equilibrium is: (a) Arbitrary: leading to an arbitrary division between fluctuations within and out of equilibrium, as is shown in the profile in Figure 1. (b) Difficult: as we have shown in the Kac ring. Except for ergodic systems there is no clear way to make a choice which guarantees that the system will spend most of its time in equilibrium. The choice of a region consisting of 99.999% of phase space still yields an evolution where only about 47% of the points on the trajectory lie within it. But why define equilibrium in this binary way? We suggest that the quality which we are trying to capture is a matter of degree, rather than the two- valued property of either being in equilibrium or not in equilibrium. We, therefore, make the following proposal: Definition (C): All references to a system being, or not being, in equilibrium should be replaced by references to the commonness of the state of the system, with this property being measured by (some possibly-scaled form of ) the Boltzmann entropy. 2.2. We Need Thermodynamic-Like Behavior to Be Typical. By this we mean that most initial states of the system should lead to thermodynamic- 688 DAVID A. LAVIS like behavior. Before discussing the dynamic properties needed for this, we shall refer briefly to the more limited notion of typicality employed by the Neo-Boltzmannians and contained in the assertion (Lebowitz 1999, S348) “that will typically be increasing in a way which explains andSB describes qualitatively the evolution towards equilibrium of macroscopic systems.” The conditions for this to be the case were first discussed by Ehrenfest and Ehrenfest-Afanassjewa (1992, 32–34). Consider a macrostate divided into four parts , , ,(��) (��) (��)m m m m , where consists of those points in which have evolved from(��) (��)m m m a smaller macrostate and which evolve into a larger macrostate, with the other parts defined in a similar way.6 On grounds of symmetry, if andx are both in , then ; otherwise the macrostates will(��) (��)Jx m m(m ) p m(m ) be in reversal pairs with plus and minus signs interchanged. In order for forward evolution to a larger macrostate to be typical it must be the case that the overwhelmingly largest part of is . This was asserted without(��)m m proof by Ehrenfest and Ehrenfest-Afanassjewa (1992). More recent ar- guments have been advanced from the point of view that a macrostate is more likely to be surrounded by larger macrostates or that it is easier to ‘aim at’ a larger, rather than a smaller, neighboring macrostate. Even accepting this argument, it can at the most explain how, if the state of a system is assigned randomly to a point in a macrostate, then the subse- quent first transition is to a larger macrostate. As was pointed out by Lavis (2005), it gives no explanation for the entropy direction at the next transition, since the part of the macrostate occupied after the first tran- sition will be determined by the dynamics. In any case, I wish to argue that this is a too narrowly defined version of typicality, which should be applied to thermodynamic-like behavior over the whole evolution. I have already argued in Section 2.1 that commonness or ‘equilibrium- ness’ is a matter of degree, and it is clear that thermodynamic-like behavior is also a matter of degree, for which we need to propose a measure. A difference between these properties is that commonness is something which can be assessed at an instant of time, whereas thermodynamic-like be- havior is a temporally global property assessed over the whole trajectory. Let be a trajectory in identified (uniquely) by the property thatL Gx N it passes through the point and let be the proportion of time whichx T (g)x the phase point evolving along spends in the some .7 DefinitionL g O Gx N TL1 is an informal qualitative definition of thermodynamic-like behavior, 6. For simplicity we have excluded evolutions between macrostates of the same size. 7. It was shown by Birkhoff (1931) that exists and is independent of the locationT (g)x of on for almost all (f.a.a.) ; that is, except possibly for a set of -measurex L x � G mx N zero. From this it follows (see, e.g., Lavis 1977) that is a constant of motionT (g)x f.a.a. .x � GN THE CONCEPT OF EQUILIBRIUM 689 for which we need the entropy profile of , not only to be quite close tosB for most of its evolution but also for fluctuations(s ) :p s (m )B Max B Max around this value to be fairly small. We, therefore, propose the following definition: Definition (TL2): The degree to which the evolution of the system is thermodynamic-like along is measured by the extent to whichLx D [s ] :p FAs S � (s ) F, (6)x B B x B Max and 2� [ ]W [s ] :p A s � As S S , (7)x B B B x x are small, where As S :p T (m)s (m), (8)�B x x B {m} is the time average of along and is the standard deviations L W [s ]B x x B with respect to the time distribution. Of course, it could be regarded as unsatisfactory that two parameters are used as a measure of the degree of a property and it is a matter of judgment which is more important. For the Kac ring of 10,000 spins with the entropy profile shown in Figure 1, �2 �1D [s ] p 0.58122724 # 10 , W [s ] p 0.31802804 # 10 , (9)x B x B and, as a comparison, for the same ring with the flippers placed at every tenth site D [s ] p 0.20078055, W [s ] p 0.20632198. (10)x B x B It is clear (and unsurprising) that the random distribution of spin flippers leads to more thermodynamic-like behavior. To explore the consequences of TL2 we distinguish between four aspects of a system: (i) The number of microsystems and their degrees of freedom, to-N gether giving the phase space , with points representing micro-GN states. (ii) The measure on .m GN (iii) The mode of division of into the set of macrostates.G {m}N (iv) The -measure preserving dynamics of the system.m Having chosen i and ii, the choices for iii and then iv are not unique. In the case of the baker’s gas (Lavis 2005), is a -dimensional unitG 2NN hypercube with volume measure. Macrostates are specified by partitioning 690 DAVID A. LAVIS each square face of the hypercube. With such a setup it would now be possible to choose all manner of discrete-time dynamics.8 Whether a system is ergodic will be determined by i, ii, and iv and, if it is, m(g) T (g) p t(g), where t(g) :p , G g O G (11)x Nm(G )N is both the proportion of in and of the time spent in , f.a.a.G g g L �N x . This will be the case for the baker’s gas but not the Kac ring. ForGN any specification of i–iii we denote the results of computing ,As SB x , and using (11) by , , and ; that is to say, weD [s ] W [s ] As S D[s ] W[s ]x B x B B B B omit the unnecessary trajectory-identifying subscript . If we were ablex to devise a model with i–iii the same as the Kac ring, butN p 10,000 with an ergodic dynamics, we would have As S p 0.69261408,B (12)�4 �4D[s ] p 0.49997497 # 10 , W[s ] p 0.70707142 # 10 ,B B and it is not difficult to show that and are monotonicallyD[s ] W[s ]B B decreasing functions of . Ergodicity leads to more thermodynamic-likeN behavior, which becomes increasingly thermodynamic-like with increasing .9 This behavior is also typical, since it occurs f.a.a. .N L � Gx N Of course, the results (12) are not simply dependent on the putative ergodic dynamics of the system, but also on the way that the macrostates have been defined. If the time average of is to be close tos (x) (s )B B Max and if the fluctuations in are to be small, then most of must lies (x) GB N in macrostates with close to . In the case of the Kac ring,s (m) (s )B B Max with , 99.98% of lies in macrostates withN p 10,000 G s (m) 110000 B . However, of course, the Kac ring, although not ergodic, gives0.999(s )B Max every appearance, at least in the instances investigated (see Fig. 1), of behaving in a thermodynamic-like manner. Although ergodicity, with a suitable macrostate structure, is sufficient for thermodynamic-like behav- ior, it is clearly not necessary. Consider the case where can be ergodically decomposed; meaningGN that G p w G , (13)N a {a} where is invariant and indecomposable under . Then the time spentG {f }a t 8. Discrete time because of the absence of a momentum component in the phase space. (See note 2 above.) 9. This latter result can be proved for any division of into macrostates with theGN measure given by a combinatorial quantity like (3). THE CONCEPT OF EQUILIBRIUM 691 Figure 2. Plot of for the cycle of the Kac ring shown in Figure 1 togetherT (m ) aa k with the corresponding curve of for a system with ergodic dynamics.t(m )k in isg O GN m(g ∩ G )aT (g) p , f.a.a. L � G , (14)a x am(G )a and we, henceforth, identify a time average along a trajectory in usingGa with also replacing the subscript in (6) and (7). In the case of. . .A S a xa the Kac ring the index labels the cycles which form the ergodic decom-a position of . For a particular cycle, such as that shown in Figure 1,GN is obtained simply by counting the number of times the phase pointT (m)a visits each macrostate in a complete cycle. The data are then used to compute the results given in (9). A plot of against the microstateT (m )a k index is shown in Figure 2 with a comparison made with , thek t(m )k corresponding curve for a system with ergodic dynamics. This gives a graphic illustration of the suggestion that ergodic systems, with a suitable choice of macrostates, are likely to be more thermodynamic-like in their behavior than nonergodic systems. It might also be speculated that –� ergodic systems (Vranas 1998) show increasingly thermodynamic-like be- havior with decreasing .� Another advantage of an ergodic system is that, f.a.a. , the levelL � Gx N 692 DAVID A. LAVIS of thermodynamic-like behavior will be the same. This contrasts with an ergodic decomposition characterized by (13), where it is possible for dif- fering levels of thermodynamic-like behavior to be exhibited within dif- ferent members of the decomposition. To be precise, take small positive and and regard behavior along a trajectory as thermodynamic-like� �D W if and only if both and . Let be the union of(T)D [s ] ! � W [s ] ! � Ga B D a B W N all in which the behavior is thermodynamic-like with .(A) (T)G G p G \Ga N N N In the discussion of the Boltzmann approach, reference to probabilities has so far been avoided. To complete the discussion in this section and to relate the arguments to the Gibbs approach we shall need (see Lavis 2005) to introduce two sets of probabilities for a system with the ergodic decomposition (13). The first of these is . Thenp :p Pr (x � L ), G Ga a a thermodynamic-like behavior will be typical for the system if p K 1. (15)� a (A)G OGa Thus we have two levels of degree, the first represented by the choices of and and the second concerning the extent to which (15) is satisfied.� �D W 3. Reconciling Gibbs and Boltzmann. As we saw in Section 1, the Gibbs approach depends on defining a probability density function on ,r(x) GN for a system ‘at equilibrium’. Thus we must address more directly the question of probabilities. Assuming the ergodic decomposition (13), we take the time-average definition of von Plato (1989), for which Prob(x � gFx � G ) :p T (g), G g O G , (16)a a N where is given by (14). Thus the probability density function forT (g)a isGa 1/m(G ) x � Ga ar (x) p , (17)a {0 otherwise from which we have r(x) p ED r (x) :p r (x)p . (18)�a a a {a} If we assume that all points of are equally likely, then on Bayesian/GN Laplacean grounds, and consonant with the approach of Bricmont (2001), we should choose p p m(G )/m(G ), (19)a a N giving, from (18), , which is the microcanonical distributionr(x) p 1/m(G )N THE CONCEPT OF EQUILIBRIUM 693 and for which, from (1), s :p S /(Nk ) p ln [m(G )] p s (G ). (20)G G B N B N It also follows, from (2), (8), (11), and (19) that ED As S p As S. (21)B a B Lavis (2005) has proposed a general scheme for relating a phase function , defined on to a macro-function defined on the macrostatesf x � G FN and then to a thermodynamic function . The first step is to course{m} F grain over the macrostates to produce .10 The second step is tof (x) F(m) define the thermodynamic variable along the trajectory as . InF L AFSx x the case of the Boltzmann entropy, which is both a phase function and a macro-function, the first step in this procedure is unnecessary since it already, by definition, course grained over the macrostates. Then we pro- ceed to identify the dimensionless thermodynamic entropy per microsys- tem with . In the case of a system with an ergodic decomposition thisAs SB x definition would yield a different thermodynamic entropy for eachsa member of the decomposition, with, from (6), s :p As S p (s ) � D [s ]. (22)a B a B Max a B In the case where the behavior is thermodynamic-like in , differs fromG sa a by at most some small and, if (15) holds, this will be the case(s ) �B Max D for measurements along most trajectories. In the case of the Kac ring with and the trajectory investigated for Figures 1 and 2 theN p 10,000 actual difference is given in (9), a value which is likely to decrease with increasing .N It is often said that in “equilibrium [the Gibbs entropy] agrees with Boltzmann and Clausius entropies (up to terms that are negligible when the number of particles is large) and everything is fine” (Bricmont 1995, 188). Interpreted within the present context this means that the good approximation , for the entropy per microsystem of a system fors (m )B Max which thermodynamic-like behavior is typical, can be replace by s pG . The advantage of this substitution is obvious, since the first ex-s (G )B N pression is dependent on the division into macrostates and second is not. However, a little care is needed in justifying this substitution. It is not valid because, as asserted in the quotation from Bricmont (2001, 3), occupies an increasing proportion of as increases. Indeed, wem G NMax N have shown for the Kac ring that the reverse is the case. That proportion becomes vanishingly small as increases. However, the required substi-N 10. It is argued that is a good approximation to for the phase functions relevantF f to thermodynamics since their variation is small over the points in a macrostate. 694 DAVID A. LAVIS tution can still be made, since for that model s (m ) ln (N )B Max � 1 � , as N r �. (23) s (G ) 2N ln (2)B N Although it may seem that the incorrect intuition on the part of Bricmont (2001) concerning the growth in the relative size of the largest macrostate, leading as it does to the correct conclusion with respect to entropy, is easily modified and of no importance, we have shown in Section 2.1 that it has profound consequences for the attempt to define equilibrium in the Boltzmann approach. It should be emphasized that the Gibbs entropy (20) is no longer taken as that of some (we would argue) nonexistent equilibrium state, but as an approximation to the true thermodynamic entropy which is the time average over macrostates of the Boltzmann entropy. The use of a time- independent probability density function for the Gibbs entropy is not because the system is at equilibrium but because the underlying dynamics is autonomous.11 The thermodynamic entropy approximated by the Gibbs entropy (20) remains constant if the phase space remains unchanged but changes discontinuously if a change in external constraints leads to a change in . An example of this, for a perfect gas in a box when a partitionGN is removed, is considered by Lavis (2005), who shows that the Boltzmann entropy follows closely the step change in the Gibbs entropy. 4. Conclusion. In our program for reconciling the Boltzmann and Gibbs approaches to statistical mechanics we have made use both of ergodicity and ergodic decomposition, and there is deep (and justified) suspicion of the use of ergodic arguments, particularly among philosophers of physics. Earman and Rédei (1996, 75) argue that “ergodic theory in its traditional form is unlikely to play more than a cameo role in whatever the final explanation of the success of equilibrium statistical mechanics turns out to be.” In its ‘traditional form’ the ergodic argument goes something like this: (a) Measurement processes on thermodynamic systems take a long time compared to the time for microscopic processes in the system and thus can be effectively regarded as infinite time averages. (b) In an ergodic system the infinite time average can be shown, for all but a set of measure zero, to be equal to the macrostate average with respect to an invariant normalized measure which is unique.12 The traditional objections to this argument are also well known: (i) Measurements may be regarded as time 11. A nonautonomous dynamic system will not yield a time-independent solution to Liouville’s equation. 12. In the sense that it is the only invariant normalized measure absolutely continuous with respect to the Lebesque measure. THE CONCEPT OF EQUILIBRIUM 695 averages, but they are not infinite time averages. If they were one could not, by measurement, investigate a system not in equilibrium. In fact, traditional ergodic theory does not distinguish between systems in equi- librium and not in equilibrium. (ii) Ergodic results are all to within sets of measure zero and one cannot equate such sets with events with zero probability of occurrence. (iii) Rather few systems have been shown to be ergodic. So one must look for a reason for the success of equilibrium statistical mechanics for nonergodic systems and when it is found it will make the ergodicity of ergodic systems irrelevant as well. This use of ergodicity differs substantially from that described above, and it thus escapes wholly or partly the strictures applied to it. In respect of the question of equilibrium/nonequilibrium I argue that the reason this does not arise in ergodic arguments is that equilibrium does not exist. The phase point of the system, in its passage along a trajectory, passes through common (high entropy) and uncommon (low entropy) macro- states and that is all. So we cannot be charged with ‘blurring out’ the period when the system was not in equilibrium. The charge against ergodic arguments related to sets of measure zero is applicable only if one wants to argue that the procedure always works; that is that non-thermody- namic-like behavior never occurs. But we have, in this respect, taken a Boltzmann view. We need thermodynamic-like behavior to be typical, and I have proposed conditions for this to be the case. But I admit the pos- sibility of atypical behavior occurring with small but not-vanishing prob- ability. While the class of systems admitting a finite or denumerable er- godic decomposition is likely to be much larger than that of the purely ergodic systems, there remains the difficult question of determining general conditions under which the temporal behavior along a trajectory, mea- sured in terms of visiting times in macrostates, approximates, in most members of the ergodic decomposition, to thermodynamic-like behavior. REFERENCES Birkhoff, Garrett D. (1931), “Proof of the Ergodic Theorem”, Proceedings of the National Academy of Science USA 17: 656–660. Bishop, Robert C. (2004), “Nonequilibrium Statistical Mechanics Brussels-Austin Style”, Studies in History and Philosophy of Modern Physics 35: 1–30. Bricmont, Jean (1995), “Science of Chaos or Chaos in Science?”, Physicalia 17: 159–208. ——— (2001), “Bayes, Boltzmann and Bohm: Probabilities in Physics”, in J. Bricmont, D. Dürr, M. C. Galvotti, G. Ghirardi, F. Petruccione, and N. Zanghi (eds.), Chance in Physics: Foundations and Perspectives. Berlin: Springer, 3–21. Earman, John, and Miklós Rédei (1996), “Why Ergodic Theory Does Not Explain the Success of Equilibrium Statistical Mechanics”, British Journal for the Philosophy of Science 47: 63–78. Ehrenfest, Paul, and Tatiana Ehrenfest-Afanassjewa (1912), The Conceptual Foundations of the Statistical Approach in Mechanics. Ithaca, NY: Cornell University Press. Goldstein, Sheldon (2001), “Boltzmann’s Approach to Statistical Mechanics”, in J. Bric- 696 DAVID A. LAVIS mont, D. Dürr, M. C. Galvotti, G. Ghirardi, F. Petruccione, and N. Zanghi (eds.), Chance in Physics: Foundations and Perspectives. Berlin: Springer, 39–54. Kac, Mark (1959), Probability and Related Topics in the Physical Sciences. London: Interscience. Lavis, David A. (1977), “The Role of Statistical Mechanics in Classical Physics”, British Journal for the Philosophy of Science 28: 255–279. ——— (2005), “Boltzmann and Gibbs: An Attempted Reconciliation”, Studies in History and Philosophy of Modern Physics 36: 245–273. Lebowitz, Joel L. (1993), “Boltzmann’s Entropy and Time’s Arrow”, Physics Today 46: 32– 38. ——— (1999), “Statistical Mechanics: A Selective Review of Two Central Issues”, Reviews of Modern Physics 71: S346–S357. Prigogine, Ilya (1994), Les Lois du Chaos. Paris: Flammarion. von Plato, Jan (1989), “Probability in Dynamical Systems”, in J. E. Fenstad, I. T. Frolov, and R. Hilpinen (eds.), Logic, Methodology and Philosophy of Science VIII. Amsterdam: Elsevier, 427–443. Vranas, Peter B. M. (1998), “Epsilon-Ergodicity and the Success of Equilibrium Statistical Mechanics”, Philosophy of Science 65: 688–708.