<<

PHYSICAL REVIEW E 89, 012108 (2014)

Thermodynamic induction effects exhibited in nonequilibrium systems with variable kinetic coefficients

S. N. Patitsas University of Lethbridge, 4401 University Drive, Lethbridge AB, Canada, T1K3M4 (Received 23 May 2013; revised manuscript received 22 November 2013; published 8 January 2014) A nonequilibrium thermodynamic theory demonstrating an induction effect of a statistical nature is presented. We have shown that this thermodynamic induction can arise in a class of systems that have variable kinetic coefficients (VKC). In particular if a kinetic coefficient associated with a given thermodynamic variable depends on another thermodynamic variable then we have derived an expression that can predict the extent of the induction. The amount of induction is shown to be proportional to the square of the driving force. The nature of the intervariable coupling for the induction effect has similarities with the Onsager symmetry relations, though there is an important sign difference as well as the magnitudes not being equal. Thermodynamic induction adds nonlinear terms that improve the stability of stationary states, at least within the VKC class of systems. Induction also produces a term in the expression for the rate of production that could be interpreted as self-organization. Many of these results are also obtained using a variational approach, based on maximizing entropy production, in a certain sense. Nonequilibrium quantities analogous to the free of equilibrium are introduced.

DOI: 10.1103/PhysRevE.89.012108 PACS number(s): 05.70.−a, 05.40.−a, 05.65.+b, 68.43.−h

I. INTRODUCTION into nonlinear realms include a nonlinear analysis developed to describe nonelectrolyte transport in kidney tubules and a In systems containing two or more irreversible transport phenomenological approach to analyze nonlinear aspects of mechanisms, reciprocal relations among transport coefficients electrokinetic effects [18,19]. have been observed for quite some time now. Early examples Further theoretical developments, partly aimed at extending include Kelvin’s analysis of thermoelectric phenomena and the Onsager relations to the nonlinear domain, included a study Helmholtz’s investigations into the conductivity of elec- of the time-reversal properties of Onsager’s relations [20], as trolytes. well as formulation of a generalized function [21]. In 1931, Onsager published his seminal studies concerning On the question of how to extend the Onsager relations, a clear the approach to thermodynamic equilibrium [1,2]. This is answer has remained elusive. In contrast to Prigogine’s prin- specifically a near-equilibrium linear theory where assump- ciple of minimum entropy production, Zeigler has proposed tions of local equilibrium apply. Such a linearized approach a principle of maximum entropy production. This was based naturally involves constant kinetic coefficients. Onsager on studies of dissipation in thermomechanics and is a linear showed that there exists very general symmetries among such response theory [22]. Since then, Zeigler’s principle has stood coefficients. This was further developed theoretically in seemingly direct contradiction to Prigogine’s principle, over the next couple of decades (see Refs. [3–7] for examples), without clear resolution (see Ref. [23] for a recent review). Bor- while from the experimental side, many systems were studied del has discussed Zeigler’s principle in the context of (linear) in detail, including systems exhibiting particle , information theory [24]. In this work we explore the possibility thermal conduction, electrical conduction, thermoelectricity, that this minimum vs maximum controversy might be settled thermomagnetic, thermomechanical and galvanomagnetic by looking at nonlinear systems approaching equilibrium, with effects, electrolytic transference, liquid helium fountain the nonlinearity embedded into the kinetic coefficients. effects, and chemical reactions. The experimental tests for To help illustrate the question of what Onsager’s reciprocal the validity of the Onsager relations have been reviewed relations might look like in the nonlinear realm, we consider extensively by Miller [8,9]. In his review Miller points out two thermodynamic variables, a and a . Then the custom- that linearization is required in certain systems. In particular, 1 2 ary equations for the linear nonequilibrium dynamics are systems involving chemical reactions are often highly [10,25–27]: nonlinear and linear modeling is expected to be inaccurate. The linear theory for approach to equilibrium was developed a˙1 = L11X1 + L12X2, (1) further when Prigogine studied stationary states and proved a˙ = L X + L X . (2) an important theorem on entropy production rates, i.e., the 2 21 1 22 2 minimum entropy production principle [10,11]. The famous result by Onsager stipulates that L12 = L21.Zero Examples of early attempts to deal with nonlinearities in external magnetic fields are assumed throughout this work. nonequilibrium thermodynamics included nonlinear convec- In order to emphasize our results reported here, we further tion and accounting for relativistic effects [12,13]. Nonlinear suppose the off-diagonal Onsager coefficients are zero to linear thermodynamics was developed as well order: [14,15]. This chemical work has continued on: for example a˙ = L X , (3) with studies of cooperative electron transfer [16,17]. Other 1 11 1 examples of work related to Onsager relations and extension a˙2 = L22X2. (4)

1539-3755/2014/89(1)/012108(14) 012108-1 ©2014 American Physical Society S. N. PATITSAS PHYSICAL REVIEW E 89, 012108 (2014)

Up to linear order these two variables are now uncoupled. If we equation is next allow for the L11 kinetic coefficient to have a functional dv dependence on variable a2 then extra nonlinear terms will m =+F (t). (7) dt be generated. If a2 happens to be the same as X2 within a ∗ constant, then by keeping only the next leading order term, The force F (t) varies on a time scale τ . To integrate this ∗ and making the direct substitution: L11 → L11 + cX2,wemay equation we must take a time step much larger than τ .The modify Eq. (3) to look like: average value of v will change much more slowly than F (t), so we take care to choose a time step t, which is large enough so = + ∗ a˙1 L11X1 (cX1)X2, (5) that t  τ , but small enough that the change inv ¯ is small. The result is where L11 is now strictly evaluated at X2 = 0. The cX1X2 t+t term is what we refer to as the direct contribution to nonlinear   m[v(t + t) − v(t)]= F (t )dt , (8) dynamics. One may then ask whether Eq. (4) also gets t modified. Put another way, can the dependence of L on 11 where the ensemble average has been taken on both sides. a indirectly induce an extra term into Eq. (4), which would 2 These ensemble averages are not over equilibrium states since change the way a changes with time? Naive application of 2 F  = 0. The states of the reservoir environment ( bath) the Onsager symmetry would suggest L = cX , adding the 0 12 1 respond quickly to the particle so when the mean velocityv ¯ term (cX )X to Eq. (4). However there is no known theoretical 1 1 is nonzero, the environment responds and thus changes F . justification for this, since this theory is inherently nonlinear Since the response of the environment is fast, local equilibrium and the proofs justifying the Onsager symmetry are not valid. conditions apply. We invoke the factor exp(S/k ) to describe We will show below that adding the term cX2 is on the right B 1 deviations from the equilibrium and convert the averaging track, but the magnitude will be smaller and more importantly procedure to evaluating equilibrium averages, i.e., there is an added minus sign. The actual equation for a˙2  S   becomes:  = kB ≈ +  F (t ) e F (t ) 0 (1 S/kB )F (t ) 0 a˙ =−(rcX )X + L X , (6) 1 2 1 1 22 2 = (S)F (t) , (9) k 0 where 0

012108-2 THERMODYNAMIC INDUCTION EFFECTS EXHIBITED IN . . . PHYSICAL REVIEW E 89, 012108 (2014) in a linear fashion, where L is constant and γ also a (small) III. GENERAL THEORY FOR THERMODYNAMIC constant. This dependence creates nonlinear effects and is what INDUCTION EFFECT couples our particle to the dynamical reservoir. The approach Consider an described by n thermodynamic to equilibrium creates entropy at a rate σ = MX2 so that in variables x with equilibrium values x . In this work we follow a time interval, dt, an amount of entropy σdt is created. The i i0 closely the approach presented in Ref. [10], in particular by linear component of this entropy change, LX2dt, will occur considering only discrete variables, and leaving the treatment regardless of the state of the particle and so will have no of continuous variables for future work. The discrete approach effect on the particle dynamics. As we will see, the nonlinear taken here suffices to clearly demonstrate the thermodynamic component will have a significant effect. We consider the induction effect. In fact even two variables will suffice, as we nonlinear part of the entropy change will show below. t t The total entropy ST may be written as a function of the n S (t) = dtγX(t)2v(t) ≈ γvX2 dtv(t), nonlin variables xi . The change ST , from equilibrium, of the total t t entropy, may be written as a function of the n-state variables (13) = − ai xi xi0 . Generalized forces (affinities) are defined by ∂ST Xi = . These conjugate parameters are interrelated as: where we have pulled the slowly varying function X(t) out of ∂ai the integral, and inserted into Eq. (9) to obtain: n X =− g a , (20)   | = 1     i ij j F (t ) nonlin [Snonlin(t )]F (t ) 0 j=1 kB 2 t and γX    = dt v(t )F (t )0. (14) n kB t =− −1 ai gij Xj , (21) We convert to the force-force correlation function as: j=1 t where g = g , i.e., the g matrix is symmetric, and with     =     ij ji v(t )F (t ) 0 dt v˙(t )F (t ) 0 the same symmetry holding for the inverse matrix. In order −∞ to ensure thermodynamic stability, both the g matrix and its t 1    inverse must be positive definite. These relations place the X = dt K(t − t ). (15) i m −∞ variables on equal footing with the ai variables. We note that these forces are defined as differential quantities, and that they Using Eq. (11) we find ∂S are not the same as T . See Ref. [29] for an example worked ∂ai  2 ∗ −(t−t)/τ ∗ out with two variables, i.e., n = 2. For each variable a we F (t )|nonlin = γX βTτ [1 − e ], (16) k are to think of a transfer of some quantity from one region to ∗ and assuming t  τ we find: another. The entropy of one region decreases while the entropy t+t of the other increases by a yet slightly larger amount, with net   2 ∗ dt F (t )|nonlin = γX τ Tβ(t). (17) increase Sk. Though many of these transfer processes may t share the same regions of space, there may in all be as many Putting together the linear and nonlinear terms: as 2n distinct regions. When considering relaxation towards equilibrium, the cus- 2 ∗ m[v(t + t) − v(t)]=−mβv¯(t)(t) + γX Tβτ (t), tomary approach is to write dynamical equations that relate the (18) time derivatives a˙i to linear combinations of the forces Xj .The kinetic coefficients are closely related to transport coefficients or such as thermal conductivity, electrical conductivity, etc. mv˙(t)=−βmv¯(t) + γX2Tβτ∗. (19) These transport processes are irreversible: the internal entropy increases with time as the system approaches equilibrium. We see as the result an extra force term, constantly pushing Before analyzing the induction effect, which is inherently the particle toward greater (lesser) velocity if γ is positive nonlinear, we briefly review the general approach. Since the (negative). This is an example of what we in general call variables treated here are statistical, a dynamical approach thermodynamic induction. Fluctuations play an important role, should use coarse-grained averages of the time derivatives. as there is no direct external force on the particle. Instead, For variable xi the time step ti chosen for the dynamics must by having the particle randomly accessing states, there is a be much larger than the relevant force (fluctuation) correlation ∗ statistical bias towards those particle states that allow the time τi , which is typically very short. We closely follow the dynamical reservoir to create entropy at a greater rate. The notation in Ref. 10, and use a bar to denote the coarse-grained ∗ induced force is small, containing factors γ , X, and τ ,all time derivative as a˙¯i . The definition is of which are assumed to be small in some sense. Though this t+ti 1  1 example is somewhat contrived, it does make the point of a˙¯i ≡ a˙i dt = ai (t + ti ) − ai (t). (22) making a case for the induction effect for a very well known ti t ti system. In order to discover more plausible physical examples We note that the symbols  denote ensemble averages over demonstrating this interesting induction effect we proceed to accessible microstates of the system. Addition of the subscript, a more general thermodynamic analysis. 0, will refer to equilibrium states in particular.

012108-3 S. N. PATITSAS PHYSICAL REVIEW E 89, 012108 (2014)

In order to evaluate this coarse-grained derivative we where closely follow the approach in Ref. [28]. In this approach, 1 0 a nonequilibrium average value is determined by evaluating Lij = dsKij (s), (31) an equilibrium average accompanied by the insertion of the kB −∞ ubiquitous exp(ST /kB ) probability weighting factor [10,27]. and Kij are the cross-correlation functions defined by Kij ≡ We emphasize that this must be the total system entropy a˙i (t)a˙j (t + s)0. The Onsager reciprocal symmetry Lij = Lji change. This weighting factor has played an important role is obtained by using time-reversal symmetry and assumption in thermodynamics for a long time and continues to receive of zero external magnetic field [28]. In the linear regime, the 1 attention. For example, the factor has been firmly established kinetic coefficients Lij are constants. recently as a consequence of the fluctuation theorem of Evans One actually solves for the system dynamics by combining and Searles [30]. The theorem is derived from exact, detailed Eqs. (21), (30) to obtain microscopic nonequilibrium dynamics, and the exp(ST /kB ) n factor is recovered in the short-time limit of the theorem. ˙ =− Explicitly then: Xi Aij Xj , (32) j=1 −   ST (t t) k n a˙i (t )= a˙i (t )e B , (23) = 0 where Aij k=1 gikLkj . The eigenvalues of the A matrix −   have units of s 1 and these define the n (relaxation) timescales where ST (t − t) ≡ ST (t ) − ST (t). Since the change in entropy is small, a linear expansion is warranted, leaving: τi for the system. In the special case where the matrices gkl and Akl are diagonal, then we have the following relation:   = 1    −  a˙i (t ) a˙i (t )ST (t t) 0, (24) gkkLkkτk = 1. (33) kB The rate of total entropy production is since a˙i 0 = 0. Integrating both sides of Eq. (24) over the time interval ti gives the coarse-grained time derivative: n n = t+ti σT Lij Xj Xi . (34) 1   = = a˙¯i = a˙i (t )dt i 1 j 1 ti t t+ti 1    B. Nonlinear case = dt a˙i (t )ST (t − t)0. (25) kB ti t Next we consider the case where the kinetic coefficients { } Making use of the generalized forces, the entropy change is are not constants but instead depend on the variables xi .We assume these dependencies to be weak so that the nonlinear n = terms generated are small. This restricts our analysis to a subset ST Xj aj . (26) of all possible systems which we term VKC. We understand j=1 that there may be systems that are nonlinear and yet do not fall = dST into this VKC class. In terms of the rate of entropy production, σT dt , we note that Our aim then is to adapt the basic approach used in the linear analysis by adding in these small nonlinear terms. The new n (variable) coefficients will be labeled Mij ({al}). In equilibrium, σT = Xj a˙j , (27) where al = 0, they take the (constant) values Lij , i.e., j=1 Mij ({al = 0}) = Lij . (35) so we may express ST as t+ti n t+ti The Onsager symmetry   ST = σdt = Xj a˙j dt . (28) Mij = Mji (36) t j=1 t still holds. One can follow the proof given in the previous A. Review of linear case section for Lij = Lji still holds even when these coefficients Substituting Eq. (28) into Eq. (25)leaves depend on another thermodynamic variable that is not ai or aj . Use of the coefficients Mkl will make the dynamical equations + + n 1 t ti t ti nonlinear. Our task then is to determine how Eq. (30)istobe a˙¯ = dt dta˙ (t) X a˙ (t) . i k t i j j 0 modified. B i t t j=1 (29) 1. Key assumptions Closely following Ref. [28] by taking the slowly varying We make four assumptions before proving Theorem 1, our main result. force functions Xj out of the integrals, and with a few more elementary steps one obtains n 1 a˙¯i = Lij Xj , (30) See Ref. [29] for an example with this symmetry explicitly verified j=1 inside of a specific model, i.e., classical effusion.

012108-4 THERMODYNAMIC INDUCTION EFFECTS EXHIBITED IN . . . PHYSICAL REVIEW E 89, 012108 (2014)

Assumption 1. We make linear expansions in the variables 2. Direct contributions { } al , and ignore higher-order terms. This defines a new set of In order to determine how Eqs. (30) are to be modified, constants γij,l such that we treat the slow and fast variables separately. In so doing, we M = L + γ a , (37) make a clear division between what we term direct and induced ij ij ij,l l contributions. Both types of contributions are essential and also l suffice to describe the system to leading order of nonlinearity. and The direct nonlinear contribution arises from simply sub- ∂M stituting in M for L in Eq. (30), while focusing on the γ = ij . (38) ij ij ij,l ∂a dynamics for the slow variables (i  m): l al =0 The Onsager symmetry condition Eq. (36) implies the follow- m n = ing similar symmetry: (a˙i )dir Xj γij,k ak. (40) j=1 k=m+1 γij,l = γji,l. (39) Making use of Eq. (21), The γij,l coefficients describe the variability of the kinetic coefficients, to leading order. This is the most reasonable m n =− −1 approach to account for the variability of kinetic coefficients. (a˙i )dir Xj γij,k gkk Xk, (41) = = + We assume that the coefficient γij,k for the linear order j 1 k m 1 corrections are not especially small somehow. In such a case or, we may have to treat correction terms up to quadratic order in a . n k = Assumption 2. Our second assumption is that of the n (a˙i )dir NikXk, (42) variables in question, m are of the slow variety while n − m k=m+1 are quickly varying in time. The slow variables are labeled where with indices i,j with index values ranging from 1 to m, and m m the fast variables are labeled with indices k,l with these index − N =−g 1 γ X =−L τ γ X , values ranging from m + 1ton. More exactly the assumption ik kk ij,l j kk k ij,l j j=1 j=1 is τk τi for all k>mand i  m. This is quite a restrictive condition. Fast and slow variables cannot be coupled to linear i  m, k > m, (43) order. This means that both the g and L matrices must both pq pq having made use of Eq. (33). The coefficients N are not have block-diagonal forms with two main diagonal blocks, ik constant because of the nonlinearity. We note that these slow and fast, i.e., g = 0, L = 0 when i>mand k  m. ik ik coefficients depend only on slowly varying forces X .Wealso The slow variables could represent the dynamical reservoir, j note that the direct contributions do not change the dynamics though we point out that the slow systems need not be large in for the fast variables. any sense. It is only their large characteristic timescales that matter here. Assumption 3. Our third assumption is that of no variability 3. Induced contributions in the kinetic coefficients Mkl associated with fast variables, We show that the fast variables are also affected, indirectly. i.e., γkl,q = 0 for all k,l > m. We note that this assumption Modification of the dynamical equations for fast variables means the fast diagonal block part of the L matrix is gives rise to effects we designate as thermodynamic induction. unchanged. Since gkl and Lkl are both symmetric and positive The induced contributions will play a key role in the fast definite, we may find a transformation to new variables for variable dynamics. We formulate the following theorem for which both gkl and Lkl are diagonal. This may involve both ro- this induction effect. tation and stretching transformations, similar to the procedure Theorem 1 (principle of thermodynamic induction). used to solve for normal modes in a system of many masses and springs [31]. Here, we will assume that these transformations m have already been accomplished for the fast variables. Though (a˙¯k)ind = NkiXi , (44) the physical interpretation of these transformed fast variables i=1 may be difficult, the transformation helps in the following analysis, as these variables are decoupled from each other. where Thus, with no loss of generality we take the fast block of both m = ∗  matrices gkl and Lkl as diagonal. Nki Lkkτk γij,k Xj ,k>m,i m. (45) Assumption 4. Our fourth assumption is that the Mij coef- j=1 ficients (for the slow variables) depend only on fast variables, The key step in our proof is to revisit Eqs. (25) and make a i.e., γij,q = 0forq  m. We note that this assumption means the slow diagonal block part of the L matrix is unchanged. distinction between linear and nonlinear contributions to ST : This last assumption is mostly for tidying up the algebra. In ST = Slin + Snonlin. (46) the case of only one slow variable it makes no difference. In this case the one slow variable a1 would solely play the role The treatment for Slin is just as described above and leads to of dynamical reservoir. Eq. (30). For the nonlinear treatment we again start from the

012108-5 S. N. PATITSAS PHYSICAL REVIEW E 89, 012108 (2014) expression for the rate of entropy production: allows us to express the thermodynamic induction theorem in t the following form:  =  Snonlin(t ) σnonlindt . (47) =− t Nki rkNik. (55) Into this equation we insert Eq. (34) with the substitution The induced terms are of opposite sign than the corresponding Lij → Mij .UsingEq.(37) and extracting the nonlinear terms direct terms. They are also smaller in magnitude since we have gives assumed rk 1. m m n σnonlin = γij,l Xi Xj al, (48) 4. Dynamical equations i=1 j=1 l=m+1 Equations (30), (42), (44) combine to give the following where we have made use of Assumptions 3 and 4. Substituting dynamical equations: Eq. (48) into Eqs. (25) and (47)gives: n  = + t+tk t a˙¯p (Lpq Npq)Xq , (56) 1    (a˙¯k)ind = dt dt a˙k(t ) q=1 kB tk t t m m n with the nonlinear kinetic coefficients Npq specified by  × γij,l al(t )Xi Xj 0, (49) Eqs. (43), (45), directly in terms of forces Xi . We note that i=1 j=1 l=m+1 all of the nonlinear terms reside in the off-diagonal blocks of the N matrix, direct terms in the upper-right block, induced  ∗ pq where the coarse-graining time scale tk τk . Pulling out 2 terms in the lower-left block. Insertion of Eq. (21)givesn the slowly varying Xi , Xj functions: coupled nonlinear first-order differential equations as: m m n 1 n n ¯ = (a˙k)ind Xi Xj γji,l − −1 = + k t gpq Xq (Lpq Npq)Xq . (57) B k i=1 j=1 l=m+1 q=1 q=1  t+tk t     × dt dt a˙k(t )al(t )0. (50) These equations could be used to solve for the transient t t response after the system suffers a large fluctuation. The Proceeding similarly as in Sec. II: procedure would involve solving these equations subject to a set of initial conditions X (0). Thinking of L + N as a m m n + i pq pq 1 t tk (a˙¯ ) = X X γ dt matrix we see that the diagonal blocks do not change at this k ind k t i j ji,l order of analysis. The off-diagonal blocks contain nonconstant B k i=1 j=1 l=m+1 t elements, which create nonlinear dynamics. We do note that t t ×      the variations in time for these kinetic coefficients will be slow. dt dt a˙k(t )a˙l(t ) 0. (51) These coefficients will appear to be almost constant as far as t −∞     the fast variables are concerned. Noting that a˙k(t )a˙l(t )0 = Kkl(t − t ) is the correlation function and making use of Eq. (31), we see that with the fast 5. Entropy production block part of the M matrix being diagonal, that Kkl = 0 when k = l,so For each variable we may define m m + ≡ 1 t tk σp a˙pXp. (58) (a˙¯ ) = X X γ dt k ind k t i j ji,k B k i=1 j=1 t Adding up these terms for slow variables only gives t t m     × dt dt Kkk(t − t ). (52) σslow ≡ a˙i Xi . (59) −∞ t i=1  ∗ With the assumptions tk τk we find For fast variables: m m n (a˙¯ ) = X X γ L τ ∗, (53) k ind i j ji,k kk k σfast ≡ a˙kXk (60) = = i 1 j 1 k=m+1 and we have our result, i.e., Eqs. (44) and (45). Defining the so that σT = σslow + σfast. Using Eqs. (34), (42), (44), dimensionless ratios rk as ∗ m m n τk σ = L X X + σ , (61) rk ≡ , (54) slow ij i j k,dir τk i=1 j=1 k=m+1 where we have defined 2This procedure easily allows for the possibility of the parameters m m { } γkl,q depending on the slow parameters ai , though here we will σk,dir ≡−τkLkkXk Xi Xj . (62) assume these are strictly constant. i=1 j=1

012108-6 THERMODYNAMIC INDUCTION EFFECTS EXHIBITED IN . . . PHYSICAL REVIEW E 89, 012108 (2014)

We see that σslow is affected by thermodynamic induction. power supply will hold the applied bias indefinitely over time Induction also affects the fast variables: and would create stationary conditions. n n Following Ref. [10] and defining the fluxes Jp ≡ a˙¯p, then = 2 + σfast LkkXk σk,ind, (63) Eq. (56) becomes: k=m+1 k=m+1 n = + where we have defined the induced rate of entropy production Jp (Lpq Npq)Xq . (65) q= for variable ak as 1 m m We consider the case where t = 0 corresponds to a system ≡ ∗ =− σk,ind τk LkkXk Xi Xj rkσk,dir. (64) state where all of the slow variables are significantly far away i=1 j=1 from equilibrium (perhaps caused by a large fluctuation), and all of the fast variables have equilibrium values ({Xk = 0}). C. Physical argument for thermodynamic induction For t>0 we wait until all fast variables have had enough time to respond and for all transient response on fast time scales τ Below, we will make an argument that the induction term k to disappear. Fluxes for the fast variables will also essentially contributes in a significant way to the stability of stationary disappear. Subsequently, these variables will evolve slowly, as states. As well, we provide here a physical argument, as quasistationary states, while tracking the slow variables. In the thermodynamic induction may at first seem a strange concept: limit where all slow timescales τ go to infinity, then the fast If two subsystems A and B are completely decoupled then i variables are truly stationary. Thus we define quasistationary subsystem B could never be shifted away from equilibrium states by the conditions: J = 0, for k>m. Explicitly: simply due to a flux caused by subsystem A being out of k equilibrium. In VKC systems the coupling is indirect, and yet m we find that a shift in subsystem B can indeed occur because the Jk = LkkXk + NkiXi = 0. (66) = coupling enters into the expression for σT . The induction effect i 1 is compatible with proper implementation of the all important Solving and using Eq. (45)gives: factor exp(ST /kB ). We cite the example of adsorption where m m m it is well known that the of an adsorbed 1 X | =− N X =−τ ∗ γ X X . species is not the same as the binding . The difference k qss ki i k ij,k i j (67) Lkk = = = lies in entropic factors, since the adsorbates would have more i 1 i 1 j 1 entropy in the gas . Thus we see entropic-based factors When a given fast variable xk is quasistationary, it is out of similar to exp(ST /kB ) in expressions for the desorption rate equilibrium, and the entropy of the subsystem associated with (flux) [32]. In chemical reactions the ST value for a reaction that variable is lowered from the equilibrium value by a net plays an important role in deciding the yield. Reactants can amount Sk given by be induced to participate in a reaction to a higher extent if 1 1 the total system entropy is increased, even if the mechanical S = X a =− g−1X2 k 2 k k 2 kk k variables directly associated with that given reactant are not ⎡ ⎤ affected by the entropy increase, but rather the increase in 2 1 m m total entropy is distributed among the product states and/or =− g−1 ⎣τ ∗ γ X X ⎦ < 0. (68) 2 kk k ij,k i j the environment. For example when two reactant molecules i=1 j=1 undergo an exothermic reaction at a surface, the excess heat of reaction could wind up in the substrate. The ensemble of This expression [not to be confused with ST from Eq. (26)] reactants is (statistically) influenced into participating even is the entropy change for the fast variable considered as an though it is another system (substrate) that benefits by having isolated system, so by the second law of thermodynamics its entropy increased. this must be a negative definite quadratic form. The fast subsystem can exist in a condition of lower-than-equilibrium entropy for sustained periods of time, as long as at least one IV. QUASISTATIONARY STATES of the relevant slow subsystems remains out of equilibrium. The concept of stationary states, as described in This sustained state would be impossible in thermodynamic Refs. [10,15] in the linear regime, is still applicable in this equilibrium. Indeed, it is possible that −Sk could be large nonlinear context. For this discussion, we feel it necessary enough, so that while in equilibrium, such states would be to make a subtle distinction between (completely) stationary sampled only for extremely short periods of time during very states and quasistationary states. For quasistationary condi- rare, extreme, fluctuations, i.e., events that are often considered tions, the variables ai evolve slowly in time. If we formally take to be thermodynamically inaccessible. This quartic form in the limit where all of the slow time scales τi go to infinity, then the slow variables {Xi } provides a good measure for how we have the conditions for stationary states to exist. Both types far the fast subsystem can be pushed away from equilibrium of states likely fall under what Prigogine intended his definition under sustained conditions. Furthermore, we point out that the of stationary states to refer to. Here we feel it is necessary to expression in Eq. (68) is a differential quantity. carefully distinguish between the two types. For a physical The actual subsystem in question could consist of two example, an electrical circuit involving a capacitor slowly distinct spatial regions with an imbalance in the relevant fast draining charge (with large RC time constant) corresponds variables. For example, the imbalance could be in thermal to quasistationary conditions. Replacing the capacitor with a energy between two regions of space. If we focus on just

012108-7 S. N. PATITSAS PHYSICAL REVIEW E 89, 012108 (2014) one fast variable xk, then the transfer will lead to one Using Eq. (21): region (for example the one losing ) having ˙ =−L11 − a decrease in entropy |δS |. In particular, using Eq. (67)for X1 X1 (αX1)X2, (72) k g11 the quasistationary state we obtain: L X˙ =+r(αX )X − 22 X . (73) ∂S ∂S 2 1 1 2 = | =− −1 | g22 δSk ak qss gkk Xk qss ∂ak ∂ak The quasistationary state for the fast variable corresponds ∂S m m to =− −1 ∗ gkk τk γij,k Xi Xj . (69) ∂ak 2 = i=1 j=1 rτ2αX1 X2ss . (74) This quantity can be positive or negative, depending on the In order to facilitate the analysis we assume the slow signs of the γij,k coefficients. We use the phrase free system is almost static and we ignore the time variation in for these δS terms. In the thermal energy transfer example, X1. We then ask what happens if we make a small change δX2 k = + TδSk would be the actual amount of heat induced to transfer in X2 away from X2ss , i.e., X2 X2ss δX2. Ignoring the between the two regions of the fast subsystem. Note that no linear (stabilizing terms) and focusing on the nonlinear terms, heat needs to be transferred between fast and slow subsystems. then:3 If the magnitude of this quantity becomes significant, relative X˙ ≈−(αX )δX . (75) to the equilibrium entropy of the region in question, then the 1 1 2 quasistationary states could be very interesting, and perhaps Over a short time t: very difficult to attain otherwise. For example, these states X (t) − X (0) ≈−(αX )δX t, (76) could involve some level of self-organization [33]. 1 1 1 2 We feel the δS is a better measure than the differential 2 ≈ 2 − 2 k X1(t) X1(0) 2αX1(0)δX2t, (77) entropy expression of Eq. (68), in assessing the potential L for producing such interesting effects. Multiplying δS by ˙ =+ 2 − 2 − 22 + k X2 rα X1(0) 2αX1(0)δX2t (X2ss δX2) the local temperature gives a new type of free energy that g22 L provides one with an energy budget available for overcoming =− 2 2 − 22 2rα X1(0)δX2t δX2. (78) barriers keeping this subsystem from making transitions to g22 these interesting quasistationary states. The reduction in entropy caused by shifting the fast For positive r we see that the nonlinear term reinforces the variables into quasistationary states suggests that the entropy stabilizing effect of the linear term. However, for negative r production associated with the slow variables may be in- the quasistationary state can become unstable. Technically, for creased. We will indeed find this to be the case, but beforehand, small enough values of t the stability is there, but realistically an investigation of the stability of quasistationary states is this time may have to be extremely small. For reasonable warranted. and relevant time scales the nonlinear term could win out and cause instability. It is not very difficult to set up an example problem with explicit numerical solution illustrating A. Stability of quasistationary states instability. For example, X1(0) = 1, L11 = 0.051 g11 = g22 = −5 The key point for the linear stability analysis is realizing, 1, L22 = 0.2, α = 0.1, r =−1, δX2 = 10 gives the solution ≡ − from Eq. (53), that no fast variables Xk are present in the for X2 X2(t) X2ss shown in Fig. 1. The quasistationary induction terms. Thus if a small change in a single fast variable state occurs at X2 =−0.5. We see that for very small times = Xl is made, i.e., δXl, then the change in the flux Jl is the same on the order of t 0.001 s the response appears to be =− as it is in the linear case. Since Jl = 0 in the quasistationary a stabilizing return back to zero, i.e., X2 0.5. But the state, we have Jl = LllδXl. The argument for stability follows system never returns to the quasistationary state. At around exactly as it does for the linear case (see Ref. [10]), i.e., the t = 0.003 s, X2(t) turns around and then responds in a way positivity of Lll, as well as the positive-definite character of consistent with instability. We conclude that if there is to be an gpq, guarantees that if al is pushed away from the stationary induction term then we have indeed obtained the correct sign, =− state, then the sign of Jl = a˙l will be such as to return i.e., r>0. The r 1 option clearly causes problems with the variable back towards the quasistationary state. Thus, it stability. Even smaller, negative values of r will also cause appears that the fast states are all stable when quasistationary. instabilities. In a certain sense, then r = 0 can be thought of as a situation neither stable or unstable, i.e., neutral. We feel that 1. Nonlinear stability analysis this further justifies the case for the induction effect: it may be true in general that in systems approaching equilibrium This argument can fail when the linear kinetic coefficients and containing kinetic coefficients that depend on other are very small. In such a case it is possible that the nonlinear thermodynamic variables, the induction term is necessary for terms dominate and decide the issue of stability. To illustrate we consider the case with one slow and one fast variable, i.e., n = 2, m = 1. The dynamical equations (56) become 3We also make the special condition that M is very small (but a˙1 = L11X1 + (cX1)X2, (70) 11 still positive) in the stationary state, i.e., X˙ 1 ≈ 0 in the stationary a˙2 =−(rcX1)X1 + L22X2. (71) state.

012108-8 THERMODYNAMIC INDUCTION EFFECTS EXHIBITED IN . . . PHYSICAL REVIEW E 89, 012108 (2014)

1.001x10-5 gradually relax towards equilibrium. Substituting Eq. (67)into Eq. (65), and with all the fast variables in quasistationary states, the fluxes for the slow variables are given by: m n Ji = Lij Xj + NikXk = = + 1.0005x10-5 j 1 k m 1 m n 1 m = Lij Xj − Nik Nkj Xj . (81) 2

X Lkk

Δ j=1 k=m+1 j=1

With all fast variables quasistationary, σfast = 0. Using 1x10-5 Eqs. (33), (55), (64), (67) we find m m | = = + σT {Jk =0} σslow Lij Xi Xj σextra, (82) i=1 j=1

i.e., the desired result. The extra entropy production rate σextra -6 9.995x10 is positive definite in the quasistationary state. 0 0.005 0.01 0.015 The physical meaning for this result is that the entire system t produces entropy faster when the fast variables are allowed to − relax by moving away from equilibrium values and achieving FIG. 1. (Color online) Plot of X2(t) X2ss with parameter r having the incorrect sign (r =−1). This solution displays an unstable quasistationary status. In the quasistationary state we may quasistationary state. think of the quantity σextra as the increase in σslow on top of the linear contribution. This result constitutes a nonequilibrium version of Le Chatelier’s principle. In the traditional Le providing stable approaches to equilibrium via quasistationary Chatelier’s principle, which is an equilibrium thermodynamics states. principle, when a given thermodynamic variable is pushed away from equilibrium, other thermodynamic variables (that B. Fluxes and entropy production are coupled by the g matrix) relax to new equilibrium values, so If we focus on one fast variable ak while in a quasistationary that the total entropy is again maximized, and the new relaxed state, we see that σk = 0 and yet the variable is not in its entropy is always greater than the unrelaxed entropy [27]. 2 equilibrium state. This means that LkkXk > 0. We refer to Here, when we push a slow variable away from equilibrium, such a term as a dissipative rate of entropy production,or fast variables (that are coupled by γij,k ) will temporarily relax dissipation rate, for short. The term is apt since its physical away from equilibrium to quasistationary states, and the new origin is strict relaxation. While in the stationary state, relaxed rate of entropy production is always greater than the the entropic induction continually pushes the variable in a unrelaxed rate. Note that our distinction between slow and fast direction opposite to dissipation, as far as entropy production states is essential in arriving at this new principle. We point goes. Referring to Eq. (64) we note that, unlike the dissipation out that the stationary state limit, where all variables become rate, the induced rate of entropy production for variable ak frozen in time, is not thermodynamic equilibrium, since the can be positive or negative, depending on the precise state slow fluxes Ji are not zero. This is worth pointing out since and also on the signs of the γji,k coefficients. However, in the one might mistakenly conclude thermodynamic equilibrium if quasistationary state there is a balance between induction and one focuses only on the fast variables. =− 2 dissipation: σk,ind,qss LkkXk . The quasistationary induced Since we have shown that allowing the fast variables to relax rate of entropy production is always negative. Given the to quasistationary states leads to increased overall entropy generality of this thermodynamic approach, there will likely production, we are led to formulate a variational principle, be many physical interpretations for such a negative value. which maximizes entropy production in a certain sense. We Self-organization may well be one of them. can use this variational principle to better understand why | fast variables would shift away from their equilibrium values Theorem 2. If σT {Xk =0} is the total entropy production with | a = 0. all fast variables at their equilibrium values, and σT {Jk =0} is the k total entropy production with all fast variables quasistationary then V. VARIATIONAL PRINCIPLES FOR | − | =  ENTROPY PRODUCTION σT {Jk =0} σT {Xk =0} σextra 0, (79) where We follow the approach taken by Prigogine for linear systems, where the choice is made to hold constant some, n m 2 = ∗ but not all, variables while leaving the rest free to vary [10]. σextra gkkτk NikXi . (80) In our case the slow variables play the role of Prigogine’s k=m+1 i=1 fixed variables, as viewed by the fast variables, which play the To prove this theorem we note that quasistationary states role of the free variables. For the linear system one minimizes do actually evolve slowly over time while the slow variables the total rate of entropy production [10]. For the nonlinear

012108-9 S. N. PATITSAS PHYSICAL REVIEW E 89, 012108 (2014) case considered here there are some differences. Using σslow state with fixed X1,X2,...,Xm (with mm. k important restrictions. The extent of the adjustment of the To prove this theorem we first substitute into Eq. (83)for fast variables must have limitations. Maximizing σ or the fluxes, using Eq. (65)togive slow σT without any constraints on the forces Xk wouldgivean m m m n unphysical runaway result. The fast variables relax until they = − −1 Lij Xi Xj 2 rk NkiXi Xk become quasistationary and an important dynamical balance i=1 j=1 i=1 k=m+1 is achieved. This balance is the reason for the minus signs in n front of the fast variable σk terms in Eq. (83). − −1 2 While the fast variables adjust themselves so that the whole rk LkkXk . (84) k=m+1 system gets to equilibrium faster, they may spend considerable time in states with lower entropy than their equilibrium states Maximizing with respect to each fast variable, and recalling (a = 0) would have, i.e., S < 0. The complex states and that the N coefficients do not depend on any fast variables k k ik interesting structures suggested above could be created as the X , results in n-m conditions: k fast variables sample their phase space and seek configurations ∂ m that maximize the rate of entropy production of the slow =−2r−1 N X − 2r−1L X =−2r−1J = 0, = ∂X k ik i k kk k k k system (with the constraints σk 0fork>m). This effect k i=1 should be enhanced if the slow system is large while the fast (85) system is small and has a gating, or bottlenecking, property of strongly controlling the pertinent kinetic coefficients of the where comparison to Eq. (53) has been made. Thus, we have large system. proven that stationary states maximize . When all of the n-m fast states are stationary (definition of completely stationary) A. Stationary states then = σslow = σT . As shown in Sec. IV, this total rate of = entropy production with all Jk 0 is larger than if all fast If we take the stationary limit, τi →∞, where the slow variables Xk were zero. variables are held constant, then the coefficients Npq become Corollary 1. One can easily show, using the method of constants. The slow variables are essentially projected out of Lagrange multipliers, that the quantity σslow is maximized the problem, acting as passive reservoirs. We note that at least when the fast variables Xk take their quasistationary values, if one slow variable is required to play the essential role of = we also add the n-m constraints: σk 0fork>m, i.e. for all dynamical reservoir. We note that Theorems 1–3 and corol- −1 fast states. The Lagrange multipliers are identified as rk . laries still apply in the stationary limit. The thermodynamic Corollary 2: (principle of maximum entropy production). induction effect persists as constant terms in the remaining Also, the total entropy production σT is maximized when n-m dynamical equations for the fast variables. These terms the fast variables Xk take their quasistationary values, with serve to drive the fast variables away from equilibrium, and = the same n-m constraints: σk 0fork>m. In this case, the towards the stationary states. The dynamical equations for the + −1 Lagrange multipliers are 1 rk . fast variables become linear. It is remarkable that a problem The quasistationary states are very important since they that begins unavoidably as nonlinear, becomes linear in this maximize the total entropy production, as long as we under- particular limit. stand the constraints and that only fast variables are involved in the maximization procedure. In this sense, we may now refer VI. CASE OF TWO VARIABLES to quasistationary states also as states of maximum entropy production. To help illustrate these concepts with an example, we con- Thus far we have formulated maximum entropy production sider the simplest system possible that exhibits thermodynamic principles in three ways. A fourth formulation is obtained as induction, in particular, the case where n = 2 and m = 1, i.e., follows. If we take the limiting procedure where slow state one slow variable acting as the dynamical reservoir, and one variables are actually fixed then we arrive at the following fast variable. This is an important case to consider since it principle: when a system described by n variables is held in a likely suffices to cover many applications of thermodynamic

012108-10 THERMODYNAMIC INDUCTION EFFECTS EXHIBITED IN . . . PHYSICAL REVIEW E 89, 012108 (2014) induction. This case illustrates the essential features of the of what is actually a slowly varying force X1: reservoir variable interacting with a variable that has its −1 r2γ11,2g22 2 dynamics coupled to the reservoir. In the simplest case, these X2 =− X , (94) L 1 two variables would be completely uncoupled except that the 22 = r2γ11,2 2 kinetic coefficient for the slow variable (i 1) happens to a2 = X . (95) L g2 1 depend on a2. This means the two variables are uncoupled up 22 22 to linear order, i.e., g12 = g21 = 0 and L12 = L21 = 0. This By Eq. (86) the free entropy rate is must be the case since if, for example, g12 was nonzero, 2 max = σT |{J =0} = L11X + σextra, (96) then we could not have one very slow time scale τ1 and k 1 one very fast time scale τ2. We can also see this as a where by Eq. (80) consequence of Assumption 2. The coupling at the nonlinear = ∗ 2 = ∗ −1 2 2 level will be described by the coefficient γ11,2 as prescribed σextra g22τ2 (N21X1) g22τ2 γ11,2g22 X1 . (97) in Assumption 1. We note that Assumption 2 guarantees that By Eq. (64) the induced rate of entropy production is γ12,1 = γ21,10 = γ12,2 = γ21,2 = 0, while Assumption 4 sets = = = =− ∗ 2 −1 4 =− γ11,1 0. Lastly, γ22,1 γ22,2 0byAssumption3.Thus, σind,2 τ2 r2γ11,2g22 X1 r2σex. (98) under our set of assumptions, only γ11,2 can be nonzero. For =− −1 In this state the differential change in entropy for the fast the slow variable, Eq. (43)givesN12 γ11,2g22 X1, which creates the coupling between the two variables, and Eqs. (56) variable is evaluated using Eq. (68): become 2 2 r2 γ11,2 4 ∗ S2 = X2a2 =− X =−τ σextra. (99) ¯ = − −1 L2 g3 1 2 a˙ 1 L11X1 γ11,2g22 X1X2 (87) 22 22 and and we verify that this is negative. ¯ = + −1 2 a˙ 2 L22X2 r2γ11,2g22 X1. (88) B. Solution of coupled differential equations Thus we verify our claims made in Eqs. (5), (6), For the two variable example discussed here, Eqs. (87), (88) =− −1 (with c γ11,2g22 ). The induction term manifests itself as are easily solved numerically. Here we present an example −1 2 +r2γ11,2g X and affects the dynamics of a2. Thermody- with the following two dynamical equations, with coupling 22 1 ± = namic equilibrium corresponds to a1 = a2 = 0. If the slow parameters equal to 0.03, (r 1): variable a1 is pushed away from zero (perhaps by a large X˙ 1 =−0.003X1 − 0.03X1X2 fluctuation), we see that a2 will be induced to also move (100) ˙ =+ 2 − away from zero. After a long time passes, both variables X2 0.03X1 X2. will relax back to equilibrium. We note that in the linear Solutions for given initial conditions X1(0) = 1, X2(0) = 0, limit we ignore γ11,2 and the two time scales are identified are shown in Fig. 2, which plots X1(t) (solid curve) and X2(t) as τ1 = τslow = 1/(L11g11) and τ2 = τfast = 1/(L22g22) τ1. For the entropy production rates: (dot-dashed). Also shown (short-dashed), is the solution for X with γ set to zero. The coupling increases the rate = = ¯ = 2 − −1 2 1 11,2 σslow σ1 a˙ 1X1 L11X1 γ11,2g22 X1X2, (89) of approach towards equilibrium for the slow state i.e. the σ = σ = a˙¯ X = L X2 + r γ g−1X2X , (90) dynamical reservoir. Variable 2 responds quickly and gets fast 2 2 2 22 2 2 11,2 22 1 2 pushed away from equilibrium. After t ≈ 5 the system is in a σT = σslow + σfast good approximation to a stationary state. In Fig. 3 we see how = 2 + 2 − − −1 2 the entropy production rates vary. Without the coupling terms, L11X1 L22X2 (1 r2)γ11,2g22 X1X2. (91) σ2 would be zero and σ1(t) would follow the long dashed curve. For the free entropy production: With the couplings, the entropy production for system 2 can be negative for some time. Of course, the total entropy production −1 2 −1 2 = σslow − r σfast = L11X − r L22X 2 1 2 2 never becomes negative. In fact, σT (short-dashed) increases −1 −1 2 to higher levels than for the dashed curve. This is consistent − 1 + r r2 γ11,2g X X2. (92) 2 22 1 with the total system approaching equilibrium faster with the We verify that coupling terms. ∂ =− −1 − −1 2 =− −1 2r2 L22X2 2γ11,2g22 X1 2r2 a˙2. (93) VII. NUMERICAL SIMULATION ∂X2 Equation (23) has the attractive feature of simplicity; though It is clear that maximizing , by varying X2, is equivalent analytically difficult to deal with without making simplifying to setting J2 = 0, i.e., the principle of maximum free entropy production is verified. approximations, it is readily amenable for numerical sim- ulation. It is quite straightforward to numerically simulate outcomes for one slow variable (#1) coupled to one fast A. Quasistationary state variable (#2). In the simulations presented here, very slow The one quasistationary state available for this system variable 1 is out of equilibrium and is producing entropy at comes from setting a˙¯2 = 0. Thus we maximize the function arateσ1 while relaxing towards equilibrium. This rate of (X1,X2) by holding X1 constant while varying X2.Interms entropy production depends on the value of variable 2. We

012108-11 S. N. PATITSAS PHYSICAL REVIEW E 89, 012108 (2014)

1 0.05

0.8 0.04

0.6 0.03 x x 1 2

0.4 0.02

0.2 0.01

0 0 0 5 10 15 20 FIG. 4. For the simulation described here, subsystem 2 has t 3000 particles, so that 3001 macrostates are available. With no external magnetic field the equilibrium value is 1500. Because of the FIG. 2. (Color online) Plots of X1(t) (solid curve) and X2(t) (dot- prescribed variation in subsystem 1 entropy production, the variable dashed). Also shown is the solution for X1 with γ11,2 set to zero j2 is pushed upwards away from equilibrium to a value near 1600. (short-dashed). during this time step, due to subsystem 1, is specified by: make system 2 very simple: N noninteracting particles, each 2 S = σ + c(j − N /2), (101) residing in one of two states (two level system such as the spin 1 10 2 2 1/2 paramagnet [25,34]). The state of subsystem 2 is described where c is a constant describing the strength of the entropic by one discrete variable j2 which is the number of particles induction effect. Positive values of c give a statistical prefer- with spin up. This variable may take integer values from 0 ence for j2 values above the equilibrium value. The key step to N2. In a zero magnetic field environment, the equilibrium is to numerically calculate weighting factors eS1/kB for each value for j2 would be N2/2, (N2 an even integer) if not coupled possible outcome. These important weighting factors are what to system 1. In our simulations we take time steps (one second gives the statistical preference. During each time step more each) during which we allow for system 2 to change i2 value microstates are created and sampled (hence more weighting) if by one, either upwards or downwards. The change in entropy S1 is larger. Implementing Eq. (101) into Eq. (23), along with a standard relaxation term for system 2 allows for calculation of the probability of subsystem 2 either making a transition -5 0.004 5x10 upwards or downwards. In Fig. 4 we present a simulation in which subsystem 2 contains N2 = 3000 particles, and begins 0.0035 0 in its equilibrium state. As we can see the value of j2 is pushed away from equilibrium, completely as a result of 0.003 simple statistics. Variable j2 attains a new mean value. Also -5x10-5 visible are fluctuations, both in magnitude and in time scale 0.0025 σ σ (τ2), in variable j2, which are consistent with the strength of 2 the dissipation constant trying to push subsystem 2 towards 0.002 -0.0001 equilibrium.

0.0015 -0.00015 VIII. SIMPLE PHYSICAL EXAMPLE: THERMAL CONDUCTION 0.001

-0.0002 We consider the simple two variable case where both 0.0005 variables represent energy imbalance, i.e., x1 = U1 and x2 = U2. So, 0 -0.00025 0 5 10 15 20 ∂S 1 1 t = = =− X1   2 T1. (102) ∂U1 T1 T1 FIG. 3. (Color online) Plots of σ (t) (dot-dashed curve), σ (t) 1 2 =−1 (solid), and σ (t) (short-dashed). Also shown is σ (t) (long-dashed) Similarly X2 2 T2, and we identify the conjugate forces T T T2 solved with γ11,2 couplings set to zero. with temperature differentials [10]. A schematic for this system

012108-12 THERMODYNAMIC INDUCTION EFFECTS EXHIBITED IN . . . PHYSICAL REVIEW E 89, 012108 (2014)

see rapid changes with temperature, possibly giving very large values for κ. Examples of such systems include Fe3O4 T+ΔT with a transition temperature near 122 K and BaVS3 with 1 T a transition near 70 K [36]. With these types of materials T2 may be increased from the expression in Eq. (106)for J copper by an order of magnitude or even more. Fabrication 2 of structures on length scales of 10 nm is currently not easy; it may become accessible with technology in the near future. FIG. 5. (Color online) Schematic diagram for two variable ther- Integrated circuit structures on the order of 30 nm are currently mal transfer system. Shaded areas represent subsystem 1 with large being produced on a wide scale. We note the interesting (slow dynamics), while the small hatched rectangle possibility of effectively shifting the transition temperature represents subsystem 2, a bottleneck composed of material having a significant temperature variance in the thermal conductivity. Arrows of a material through the application of a generalized force denote fluxes in both subsystems. associated with another subsystem variable. Fabrication of very small structures capable of producing significant values of T2 could have important applications in microelectronics. is provided in Fig. 5, which also illustrates how subsystem 2 Entropic induction could be used to cool small regions of an can act like a bottleneck for the heat transfer in system 1. The integrated circuit. Such a type of cooling could provide an = = heat capacities C1 and C2 can be used as: a1 C1T1, a2 alternative to cooling using the Peltier effect. = 2 = 2 C2T2. We note that g11 1/(C1T1 ), g22 1/(C2T2 ). To Note that the induction effect may become prominent in = ˙ =− 1 =− linear order: a˙1 L11X1, i.e., U1 L11 2 T1 k1T1, systems that are not microscopic, for example models for T1 ≡ 2 traffic flow [37], as an example where small changes in where k1 L11/T1 , and we see that L11 is proportional to the ˙ =− 1 certain parameters can create a bottleneck effect. Further thermal conductivity coefficient. Also U1 L11 2 U1 C1T1 2 examples of test systems may be found by considering systems C1T1 which allows one to identify the timescale τ1 = . L11 where , not energy, is out of equilibrium. The We assume that the thermal conductivity coefficient λ1 generalized force would be chemical potential difference, as depends to some degree on temperature, so therefore on T2. opposed to temperature difference. Also, we point out that the T2 ∂λ1 If we define a dimensionless constant κ ≡ |a =0, then λ1 ∂T2 2 generalized force X1 does not have to be created artificially. T2 ∂M11 ∂M11 κ ≡ |a =0 and γ11,2 = |T =0 = κL11T2g22. For example, very small systems exist naturally, which are L11 ∂T2 2 ∂U2 2 When a2 is in a stationary state, then from Eq. (94): composed of atoms and molecules on the verge of chemical ∗ reaction. In this case the generalized driving force is the affinity r2γ11,2 2 τ2 κL11T2 2 X =− X =− X . (103) [10,15]. In these very small systems the time scale τ2 can very 2 1 1 ∗ g22L22 τ2L22 easily be almost as small as τ2 . Thus, in these systems the In terms of temperature differentials, assuming T1 ≈ T2 ≡ T , induction effect can likely be very significant. and using L22g22τ2 = 1: We point out that since the induction effect does not even exist to linear order, then the results presented here represent T k (T )2 2 = τ ∗κ 1 1 . (104) the leading term in the response (for fast variables). Thus, even 2 2 T C2 T if κ and the driving force X1 are not small and the accuracy If the characteristic length scale of the bottleneck region is of the theory presented here [Eq. (103)] is not high, it still k1 = λ1 represents a good starting point and the best estimate currently l, then one can show that 2 where λ1 is the material C2 c2l available. thermal conductivity of subsystem 1, while c2 is the volumetric specific heat of subsystem 2. Thus, T λ (T )2 IX. CONCLUSIONS 2 = τ ∗κ 1 1 . (105) 2 2 2 T c2l T We have developed a nonequilibrium thermodynamic − theory that demonstrates an induction effect of a statistical For copper at room temperature, λ1 = 1.1 × 10 4 m2/s[35]. c2 nature. We have shown that this thermodynamic induction − = λ1 = × 12 1 If we take l 10 nm, then 2 1.1 10 s .Ifwe c2l can arise in systems that are naturally nonlinear through ∗ = × −13 take τ2 1.5 10 s, which is the characteristic time having nonconstant kinetic coefficients, i.e., the VKC class. scale for atomic vibrations and fluctuations [32,35] then In particular if a kinetic coefficient associated with a given ∗ λ1 ≈ τ 2 0.2. For copper, the temperature variation of the thermodynamic variable depends on another, faster, variable 2 c2l thermal conductivity is rather small [35]: at room temperature then we have derived an expression that can predict the κ =−300 × 0.0039 =−1.2. So extent of the induction. The induction is proportional to the square of the driving force. The nature of the intervariable T (T )2 2 ≈−0.24 1 . (106) coupling for the induction effect has similarities with the T T 2 Onsager symmetry relations, though there is an important Thus, even for an ordinary material such as copper, induction sign difference as well as the magnitudes not being equal. effects could be observed for very small systems. For example We have found that the nonlinear effects from the induction (T1) = T2 =− if T 0.1 then we predict that T 0.0024. can enhance the stability of stationary states, as the system If however, the material in the junction is near a metal approaches equilibrium. The induction effect gives an entropy insulator transition, then the thermal conductivity can also production rate term that opposes dissipation, which we

012108-13 S. N. PATITSAS PHYSICAL REVIEW E 89, 012108 (2014) refer to as the induced rate of entropy production. The depending on whether one wants to focus on the free entropy key step in identifying the thermodynamic induction effect production, the slow variable rate of entropy production, or the was in identifying certain variables to act as the dynamical total rate of entropy production. reservoir. At least one such, slow, variable is essential in the We anticipate that in some systems, thermodynamic in- analysis. The dynamical reservoir also plays a key role in duction effects are not merely small corrections to a linear arriving at a new nonequilibrium version of Le Chatelier’s response, but that the effects may be very significant. We have principle. shown that there exist nonequilibrium quantities analogous We have also developed a variational approach, based on to the free energies of equilibrium thermodynamics. These optimizing entropy production. On the question of resolving newly defined free entropies, which are nonzero only in whether entropy production is minimized or maximized, we nonequilibrium conditions, can quantify the significance of conclude that, at least for the nonlinear systems considered the induction effects. here, it is the free entropy production that is maximized. The Finally, we have discussed some schemes directed towards maximization occurs while the fast variables are quasistation- discovering experimental evidence for entropic induction, ary. Thus, the stationary states of Prigogine, introduced in the including a possible application to specialized cooling in context of the minimum entropy production principle, are still integrated circuits. Detailed calculations show that the entropic very useful, at least within the VKC class of systems. The proof induction effect is most likely to be realized if a key component we have provided is simple and provides predictive power such of the system is very small. Inside such a small region, or junc- as in establishing the values of the Legendre coefficients, as tion, fluctuations play an important role behind the induction. well as prescribing just how much faster the entire system approaches equilibrium in the quasistationary states. Such ACKNOWLEDGMENTS predictive power has been absent in previous discussions of the maximum entropy production principle [23]. The maximum We thank Cathy J. Meyer for her assistance in editing the entropy production principle can be expressed in various ways, manuscript.

[1] L. Onsager, Phys. Rev. 37, 405 (1931). [22] H. Zeigler, An Introduction to Thermomechanics (North [2] L. Onsager, Phys. Rev. 38, 2265 (1931). Holland, New York, 1977). [3] H. Casimir, Rev. Mod. Phys. 17, 343 (1945). [23] L. Martyushev and V. Seleznev, Phys. Rep. 426, 1 (2006). [4] H. Callen, Phys. Rev. 73, 1349 (1948). [24] S. Bordel, Phys. Stat. Mech. App. 389, 4564 (2010). [5] S. De Groot and P. Mazur, Phys. Rev. 94, 218 (1954). [25] F. Reif, Fundamentals of Statistical and Thermal Physics [6] S. De Groot and P. Mazur, Phys. Rev. 94, 224 (1954). (McGraw Hill, New York, 1965). [7] S. de Groot and N. van Kampen, Physica 21, 39 (1954). [26] R. Pathria, (Butterworth-Heinemann, [8] S. de Groot, Physica 13, 555 (1947). Boston, 1996). [9] D. Miller, Chem. Rev. 60, 15 (1960). [27] L. Landau and E. Lifshitz, Statistical Physics, Part 1 (Elsevier, [10] S. de Groot, Thermodynamics of Irreversible Processes (North New York, 1980). Holland, Amsterdam, 1966). [28] F. Reif, Fundamentals of Statistical and Thermal Physics [11] E. Jaynes, Ann. Rev. Phys. Chem. 31, 579 (1980). (McGraw Hill, New York, 1965), Ch. 15. [12] P. Glansdorff and I. Prigogine, Physica 30, 351 [29] S. Patitsas, Am. J. Phys. (to be published). (1964). [30] D. Evans and D. Searles, Adv. Phys. 51, 1529 (2002). [13] S. de Groot, C. van Weert, W. Hermens, and W. van Leeuwen, [31] K. R. Symon, Mechanics (Addison-Wesley, Don Mills, 1971). Phys. Lett. 26, 345 (1968). [32] A. Zangwill, Physics at Surfaces (Cambridge University Press, [14] I. Prigogine, Physica 31, 719 (1965). Cambridge, 1988). [15] I. Prigogine, Introduction to Thermodynamics of Irreversible [33] G. Nicolis and I. Prigogine, Self-Organization in Nonequilib- Processes (Wiley, New York, 1967). rium Systems: From Dissipative Structures to Order through [16] H. Tributsch and L. Pohlmann, Chem. Phys. Lett. 188, 338 Fluctuations (Wiley, New York, 1977). (1992). [34] D. V. Schroeder, An Introduction to Thermal Physics (Addison [17] H. Tributsch, Electrochim. Acta 52, 2302 (2007). Wesley Longman, Don Mills, Ontario, 2000). [18] F. Sauer, Am. J. Physiol. Ren. Physiol. 46, 399 (1973). [35] N. W. Ashcroft and D. N. Mermin, Solid State Physics,1sted. [19] C. F. Gonzalez-Fernandez, J. Colloid & Interface Sci. 166, 302 (Thomson Learning, Toronto, 1976). (1994). [36] M. Imada, A. Fujimori, and A. Tokura, Rev. Mod. Phys. 70, [20] W. Muschik, J. Non-Equil. Thermodyn. 2, 109 (1977). 1039 (1998). [21] J. Bataille, D. Edelin, and J. Kestin, J. Non-Equil. Thermodyn. [37] B. S. Kerner, S. L. Klenov, and M. Schreckenberg, Phys. Rev. 3, 153 (1978). E 84, 046110 (2011).

012108-14