THE JOURNAL OF CHEMICAL PHYSICS 124, 224103 共2006兲 Statistical mechanical theory for steady state systems. V. Nonequilibrium probability density Phil Attarda兲 School of Chemistry F11, University of Sydney, New South Wales 2006, Australia 共Received 15 March 2006; accepted 17 April 2006; published online 12 June 2006兲 The phase space probability density for steady heat flow is given. This generalizes the Boltzmann distribution to a nonequilibrium system. The expression includes the nonequilibrium partition function, which is a generating function for statistical averages and which can be related to a nonequilibrium free energy. The probability density is shown to give the Green-Kubo formula in the linear regime. A Monte Carlo algorithm is developed based upon a Metropolis sampling of the probability distribution using an umbrella weight. The nonequilibrium simulation scheme is shown to be much more efficient for the thermal conductivity of a Lennard-Jones fluid than the Green-Kubo equilibrium fluctuation method. The theory for heat flow is generalized to give the generic nonequilibrium probability densities for hydrodynamic transport, for time-dependent mechanical work, and for nonequilibrium quantum statistical mechanics. © 2006 American Institute of Physics. 关DOI: 10.1063/1.2203069兴 I. INTRODUCTION II. HEAT FLOW This paper is the culmination of a series on the theory of steady state systems.1–4 The goal of the research has been to develop a nonequilibrium phase space probability distribution. The earlier papers laid much of the ground work in terms of general theory and specific application to the problem of heat flow. The present paper gives an explicit formula for the probability density in the case of steady heat flow and extends this formula to more general cases such as nonequilibrium work and nonequilibrium quantum statistical mechanics. The focus on the nonequilibrium phase space probability distribution is motivated by the status of the equilibrium theory. The Boltzmann distribution is central to equilibrium statistical mechanics, and the theory of equilibrium statistical mechanics is entirely dependent on that result. Accordingly, one can say that one has a theory for nonequilibrium statistical mechanics if, and only if, one can write down explicitly the probability distribution. The paper is set out as follows: Section II gives the steady state probability distribution for heat flow and explores its mathematical properties. Section III gives the details of the nonequilibrium Metropolis Monte Carlo algorithm and presents results for the thermal conductivity of a Lennard-Jones fluid. Section IV gives the probability distribution for hydrodynamic transport in general, for nonequilibrium mechanical work on a subsystem in contact with a thermal reservoir, and for nonequilibrium quantum statistical mechanics. a兲 Electronic mail: [email protected] 0021-9606/2006/124共22兲/224103/13/$23.00 A. Steady state probability In Paper I it was shown that the appropriate thermodynamic variables for the case of heat flow were the zeroth E0 and first E1 energy moments, and their respective thermodynamic conjugates, the zeroth T0 and first T1 temperature.1 The first temperature is related to the temperature gradient. The idea that moments and gradients are conjugate is due to Onsager.5 Consider a subsystem of length L in thermal contact with two heat reservoirs of temperature T± located at z = ± L / 2. The zeroth and first inverse temperatures of the reservoirs are1 0 = 冋 册 1 1 1 + , 2kB T+ T− 1 = 冋 册 1 1 1 − . k BL T + T − 共1兲 One sees that the zeroth temperature T0 = kB / 0 is essentially the average temperature of the reservoirs and that the first temperature T1 = kB / 1 is essentially the imposed thermal 2 gradient T−1 1 = −共ⵜT兲 / T0. The zeroth energy moment is the ordinary Hamiltonian N E0共⌫兲 = H共⌫兲 = 兺 ⑀i , 共2兲 i=1 and the first energy moment in the z direction is just N E1共⌫兲 = 兺 ⑀izi , 共3兲 i=1 where ⑀i is the total energy of particle i. The rate of change of the energy moment, Ė1共⌫兲 = ⌫˙ · ⵜE1共⌫兲, is related to the heat flux by J = Ė1 / V, where V = AL is the subsystem volume. Note that previously1,3,4 this was denoted as Ė01; then, as now, it means the adiabatic evolution. 共Adiabatic means natural or 124, 224103-1 © 2006 American Institute of Physics Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-2 J. Chem. Phys. 124, 224103 共2006兲 Phil Attard Hamiltonian motion, which is to say that no heat flows to the isolated system.兲 In the above ⌫ = 兵qN , pN其 is a point in the phase space of the subsystem, with q and p denoting position and momentum, respectively. The conjugate point in phase space is the one with all the velocities reversed ⌫† = 共qN,共− p兲N兲. 共4兲 The conjugation operation is synonymous with time reversal. Both moments have an even parity, E0共⌫兲 = E0共⌫†兲 and E1共⌫兲 = E1共⌫†兲, since it is assumed that there are no velocity dependent forces in the Hamiltonian. The rate of change of the energy moment necessarily has an odd parity, Ė1共⌫†兲 = −Ė1共⌫兲. Any function f共x兲 can be split into even and odd functions, f ±共x兲 = 关f共x兲 ± f共−x兲兴 / 2, and any non-negative function can be expressed as the exponential of f共x兲 and consequently can be factorized as g共x兲 = exp关f +共x兲兴exp关f −共x兲兴 ⬅ ge共x兲go共x兲. Since the steady state probability distribution is necessarily non-negative, it can be formally written as 㜷ss共⌫兩0, 1兲 = 㜷e共⌫兩0, 1兲㜷o共⌫兩1兲. 共5兲 The reservoirs can only enter via the energy moment derivatives of their entropy, 0 and 1. The probabilities must be exponentials of linear functions of 0 and 1 times their conjugate energy moment or a linear functional of that moment. These requirements follow from the recognition that the exponents represent that part of the reservoir entropy that depends on the subsystem and that the subsystem is very much smaller than the reservoir and so a linear Taylor expansion suffices.6 In the limit 1 → 0, 㜷ss must reduce to the equilibrium Boltzmann distribution, which is even, and hence the odd exponent must be independent of 0. The even term 㜷e is identical with the structural probability given previously,1 㜷e共⌫兩0, 1兲 = 1 Z̃ss共0, 1兲 e−0E0共⌫兲e−1E1共⌫兲 . 共6兲 Here Z̃ss is the normalizing partition function; top and bottom may be multiplied by the traditional but immaterial factor of N!h3N to make it dimensionless, and its logarithm may be taken to form a free energy or total entropy.6 This result was originally derived by making an explicit physical identification in terms of the change in entropy of the reservoir.1 It is not possible to add any additional even term without violating that physical interpretation and the requirements of linearity in the inverse temperatures and of extensivity in the energies. Monte Carlo simulations based on this probability distribution and a comparison with certain predictions based on a bulk equation of state provide convincing evidence that this is indeed correct for the structure.1 In view of the above discussion the odd probability may be written as 㜷o共⌫兩0, 1兲 = e Wmir 1 1Wmir 1 共⌫兲 共7兲 , mir † Wmir 1 共⌫ 兲 = −W1 共⌫兲. with having an odd parity, The label “mir” stands for mirror, which reflects the asymmetric nature of this term. The fact that Wmir 1 is multiplied by 1 means that it must be a linear function 共or functional兲 of the conjugate thermodynamic variable E1. However, it also has to have an odd parity, which suggests that it is a linear function or functional of Ė1. In Ref. 3 the ansatz Wmir 1 共⌫兲 = ␣Ė1共⌫兲 was explored for different values of the time constant ␣. The results suggested that this was a reasonable approximation in certain regimes, but that this form itself was not exact. In order to identify Wmir 1 one has to give the change in the reservoirs’ entropy associated with the point in phase space ⌫. The difficulty is that E1 is not a conserved variable and the total change in energy moment arises from its adiabatic evolution, which costs no reservoir entropy, and from perturbations by the reservoir, which do. The total change in moment is already accounted for by the even exponent, and the quantity Wmir 1 is essentially the internal change that must be subtracted. This particular physical interpretation of Wmir 1 is discussed in more detail at the end of Sec. II E below. It turns out that 1⌬Wmir 1 may also be interpreted as a type of thermodynamic work, or change in reservoir entropy, over an interval. Since the system is in a steady state, over long enough time scales the total change in energy moment must be zero, which is to say that the internal change in moment must be canceled by the change in the subsystem moment induced by the reservoir, ⌬0E1 = −⌬s共r兲E1. But by energy conservation, the change in the subsystem moment induced by the reservoir must be equal and opposite to the change in the reservoir moment, ⌬s共r兲E1 = −⌬E1r. Hence mir 0 ⌬Wmir 1 = ⌬ E1 = ⌬E1r, and one has 1⌬W1 = ⌬Sr / kB. Accordmir ingly, one may indeed interpret ⌬W1 as a type of thermodynamic work performed by the reservoir, which accounts for the notation used for this term. These physical interpretations can explain and rationalize Wmir 1 , but they do not replace a precise definition. This is now given, and one should judge the validity of this forin the light of its mathematical consemulation of Wmir 1 quences. Let ⌫0共t 兩 ⌫兲 be the Hamiltonian 共adiabatic兲 trajectory of the isolated subsystem that passes through ⌫ at t = 0. Denote the isolated subsystem past and future moments by E±1 共⌫兲 ⬅ E1共⌫0共± 兩 ⌫兲兲, for some time interval ⬎ 0. From the time reversible nature of the equations of motion, ⌫0共t 兩 ⌫兲 = 关⌫0共−t 兩 ⌫†兲兴†, it follows that E+1 共⌫†兲 = E−1 共⌫兲. With these definitions Wmir 1 may be written in several equivalent forms, 1 + − Wmir 1 共⌫兲 = 关E1 共⌫兲 − E1 共⌫兲兴 2 = = 1 2 1 2 冕 冕 − dt⬘Ė1共⌫0共t⬘兩⌫兲兲 0 − dt⬘关Ė1共⌫0共t⬘兩⌫兲兲 − Ė1共⌫0共t⬘兩⌫†兲兲兴. 共8兲 The factor of a half outside the integral accounts for double counting, as the future cost now will also be included as a past cost in the future; equivalently, the second equality shows that only the even part of the integrand contributes, and the past contribution to date is exactly half the total contribution. The third equality demonstrates that the first Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-3 J. Chem. Phys. 124, 224103 共2006兲 Statistical mechanical theory for steady state systems. V two equalities do not really violate causality. Clearly Wmir 1 has the necessary odd parity, and because of this it may be called the mirror thermodynamic work. With this result, the steady state probability density is explicitly e−0E0共⌫兲e−1E1共⌫兲  Wmir共⌫兲 e 1 1 , 㜷ss共⌫兩0, 1兲 = 3N h N!Zss共0, 1兲 共9兲 where h is Planck’s constant and N is the number of atoms in the subsystem. The first group of factors corresponds to 㜷e, the final factor corresponds to 㜷o, and the denominator is the normalizing partition function, Zss共0, 1兲 = 1 h3NN! 冕 mir d⌫e−0E0共⌫兲e−1E1共⌫兲e1W1 共⌫兲 . 共10兲 The nonequilibrium unconstrained total entropy is Stotal,ss共0 , 1兲 = kB ln Zss共0 , 1兲, and its derivatives generate nonequilibrium statistical mechanical averages, just as in the equilibrium case.6 The nonequilibrium constrained free energy is −T0 times the constrained total entropy, which is the subsystem-dependent part of the reservoir entropy 共the exponent of the probability density兲 plus the entropy of the isolated subsystem constrained to be in the macrostate. That is, F共E0,E1,E±1 兩0, 1, 兲 = E0 + 1关E1 − 共E+1 − E−1 兲/2兴/0 − T0S共E0,E1,E+1 ,E−1 兩兲, 共11兲 where the final term is the entropy of the isolated subsystem 共see also Sec. II F 2 below兲. The constrained free energy is strictly greater than −kBT0 ln Zss. In the thermodynamic limit the minimum value of the constrained free energy 共with respect to the constraints, which are written to the left of the vertical bar兲 is equal to −kBT0 ln Zss with a negligible error, and this defines the most likely thermodynamic state. Its derivatives in the most likely macrostate generate various relationships between thermodynamic quantities, just as in the equilibrium case.6 B. Nature of W1mir The nonzero contributions to this integral come from about the origin 共from the so-called inertial region兲. This may be seen from the asymptote of the integrand in the intermediate regime2 Ė1共⌫0共t兩⌫兲兲 ⬃ sign共t兲Ė1共E1共⌫兲兲, short ⱗ 兩t兩 ⱗ long . discussed below is related to this by ss = −1 / 2VT20. Since the asymptote is odd in time,2,3 it may be added to the integrand without changing the result. Hence Wmir can be 1 equivalently written as Wmir 1 共⌫兲 = Ė1共E1兲 = − VkBT20 具E21典0 E1 ⬅ − cE1 共future兲, 共13兲 where is the thermal conductivity. The transport coefficient 冕 − dt⬘关Ė1共⌫0共t⬘兩⌫兲兲 − sign共t⬘兲Ė1共E1共⌫兲兲兴, 共14兲 which shows explicitly that Wmir 1 is dominated by the brief inertial part of the trajectory. In other words, Wmir 1 is independent of for ⲏ short. In so far as the time integral that is Wmir 1 is dominated by short times, then the temporally even part of Ė1共t兲 must vanish by 兩t兩 ⱗ short, the inertial time. The simplest approxima2 2 tion is to take Ė1共⌫0共t 兩 ⌫兲兲even ⬇ Ė1共⌫兲e−nt /short, for some n of order unity. In this case Wmir 1 共⌫兲 ⬇ 冑 shortĖ1共⌫兲. 4n 共15兲 This result is consistent with the ansatz discussed previously1 and which was shown to be a reasonable approximation in certain regimes.3 An earlier work gives the inertial time approximately as2 short = = − 2具Ė1共t兲E1共0兲典0 具Ė21典0 2VkBT20 具Ė21典0 , t Ⰷ short . 共16兲 This result will be used shortly. The long time limit is given by2 long = 具E21典0 2VkBT20 . 共17兲 C. Relationship with Green-Kubo It is now shown that the above result for the steady state probability density gives the Green-Kubo expression for the thermal conductivity. Fourier’s law gives the thermal conductivity as the ratio of the heat flux to the applied temperature gradient, 共12兲 共This is valid if 兩tĖ1兩 Ⰶ 兩E1兩, which in practice is always the case and will be assumed throughout.兲 It is not explicitly required here, but the most likely flux going forward in time is2,3 1 2 = 具Ė1共⌫兲典ss . VkBT201 共18兲 Hence using the steady state probability distribution in the linear regime 1 Ⰶ 1, the average of Ė1 will yield an expression for the thermal conductivity that can be compared with the Green-Kubo expression.7 The steady state average of the rate of change of the first energy moment is Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-4 J. Chem. Phys. 124, 224103 共2006兲 Phil Attard mir 具Ė1共⌫兲典ss = = 兰 d⌫e−0E0共⌫兲e−1E1共⌫兲e1W1 共⌫兲 mir 兰 d⌫e−0E0共⌫兲e−1E1共⌫兲e1W1 兰 E0共⌫兲 = E0共⌫†兲 = E0共⌫*兲, Ė1共⌫兲 共⌫兲 E1共⌫兲 = E1共⌫†兲 = − E1共⌫*兲, d⌫e−0E0共⌫兲关1 − 1E1共⌫兲 + 1Wmir 1 共⌫兲兴Ė1共⌫兲 兰 d⌫e−0E0共⌫兲关1 − 1E1共⌫兲 + 1Wmir 1 共⌫兲兴 = 1兰 d⌫e−0E0共⌫兲Wmir 1 共⌫兲Ė1共⌫兲 兰 d⌫e−0E0共⌫兲 = 1具Wmir 1 共⌫兲Ė1共⌫兲典0 . 共21兲 Ė1共⌫兲 = − Ė1共⌫ 兲 = − Ė1共⌫ 兲, † * mir † mir * Wmir 1 共⌫兲 = − W1 共⌫ 兲 = − W1 共⌫ 兲, 共19兲 Since Ė1 is of an odd parity, the integral of it times the two even terms from the linearization of the exponents vanishes. The integral over −1E1共⌫兲 in the denominator vanishes because it is odd in z. Inserting the definition of Wmir 1 and using the Fourier result for the thermal conductivity, one has the result and, of course, d⌫ = d⌫† = d⌫*. It is assumed that the equilibrium system is disordered or that it is mirror symmetric about the midplane. Expanding the exponents to a cubic order in 1 and observing that any product of terms that has a total odd parity under either operation must vanish, one obtains 具Ė1共⌫兲典ss/1 = 具Wmir 1 Ė1典0 + 21 2 mir 关具E1W1 Ė1典0 2 − 具E21典0具Wmir 1 Ė1典0兴 + 共兲 = = 1 冕 冕 2VkBT20 1 VkBT20 − mir 2 − 具共Wmir 1 兲 典0具W1 Ė1典0兴. dt⬘具Ė1共⌫0共t⬘兩⌫兲兲Ė1共⌫兲典0 dt⬘具Ė1共⌫0共t⬘兩⌫兲兲Ė1共⌫兲典0 . 21 3 关具共Wmir 1 兲 Ė1典0/3 2 共20兲 0 The second equality follows because the time correlation function of two quantities of the same parity is an even function of time. In the intermediate regime, this may be recognized as the Green-Kubo expression for the thermal conductivity,7 which in turn is equivalent to the Onsager expression for the transport coefficients.2 This result is a very stringent test of the present expression for the steady state probability distribution. There is one, and only one, exponent that is odd, linear in 1, and that satisfies the Green-Kubo relation. Conversely, the GreenKubo relation alone does not determine the steady state probability distribution, since it is possible to add any even function to the exponent without changing the integral. In other words, the Green-Kubo relation represents a necessary condition that the steady state probability distribution must satisfy, but it does not provide a sufficient condition to determine the steady state probability distribution. This observation has implications for the constraints invoked in the nonequilibrium molecular dynamics technique: demanding that the Green-Kubo relation be satisfied is not sufficient to determine them uniquely. The first neglected term is of the order 41. Amongst other things the nonlinear terms account for the coupling of the induced moment to the heat flux. E. Adiabatic evolution of W1mir Let ⌫⬘ = ⌫ + ⌬t⌫˙ , be the adiabatic evolution of ⌫ after an infinitesimal time step. The adiabatic evolution of Wmir 1 can be obtained from Wmir 1 共⌫兲 = 1 2 = 1 2 = 1 2 冕 冕 冕 − dt⬘Ė1共⌫0共t⬘兩⌫兲兲 −⌬t −−⌬t −⌬t −−⌬t dt⬙Ė1共⌫0共t⬙ + ⌬t兩⌫兲兲 dt⬙Ė1共⌫0共t⬙兩⌫⬘兲兲 = Wmir 1 共⌫⬘兲 − ⌬t 关Ė1共⌫0共兩⌫⬘兲兲 − Ė1共⌫0共− 兩⌫⬘兲兲兴 2 = Wmir 1 共⌫⬘兲 − ⌬tĖ1共E1共⌫⬘兲兲. The present steady state probability density is not restricted to the linear regime. The first correction to GreenKubo may be obtained as follows. Define the z-mirror operation as ⴱ: qz → −qz, pz → −pz, and under it and the conjugation one has the useful parities 共23兲 Hence the adiabatic rate of change of Wmir 1 is Ẇmir 1 共⌫兲 = Ė1共E1共⌫兲兲, D. Nonlinear conductivity 共22兲 共24兲 which has an even parity, as it ought to have. This result is exact for in the intermediate regime. Despite first appearances, this exact result for Ẇmir 1 is, in fact, consistent with the approximate result 关Eq. 共15兲兴, 冑 Wmir 1 共⌫兲 ⬇ shortĖ1共⌫兲 / 4n, even though its derivative, mir Ẇ1 共⌫兲 ⬇ shortË1共⌫兲冑 / 4n, does not look at all like the exact result. Consider, however, the quantity Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-5 冕 ⬁ dt⬘具Ẇmir 1 共⌫0共t⬘兩⌫兲兲Ė1共⌫兲典0 . 共25兲 0 Using the approximate expression, it is given by − 冑 short具Ė21典0 , 4n 共26兲 since 具Ė1共⌫0共t⬘ 兩 ⌫兲兲Ė1共⌫兲典0 → 0, t⬘ → ⬁. Using the exact expression and the fact that Ė1 = −cE1 关Eq. 共13兲兴, it is given by 冕 ⬁ dt⬘具− cE1共⌫0共t⬘兩⌫兲兲Ė1共⌫兲典0 0 =c 冕 ⬁ dt⬘具Ė1共⌫0共t⬘兩⌫兲兲E1共⌫兲典0 = − c具E21典0 , 共27兲 0 since the time correlation of functions of different parities is odd in time and since 具E1共⌫0共t⬘ 兩 ⌫兲兲E1共⌫兲典0 → 0, t⬘ → ⬁. Equating these two, one must have 冑 J. Chem. Phys. 124, 224103 共2006兲 Statistical mechanical theory for steady state systems. V c具E21典0 VkBT20 = . short = 4n 具Ė21典0 具Ė21典0 共28兲 If one takes n = , this is equivalent to the result obtained previously2 关Eq. 共16兲兴. Such a value is not unreasonable. The consistency of these two different approaches gives one some confidence in both the theory and the approximations. This result for the adiabatic evolution of Wmir 1 provides a physical interpretation of the steady state probability 关Eq. 共9兲兴. The difficulty with heat flow is the lack of conserved variables; the change in energy moment ⌬E1 occurs both internally by the adiabatic evolution of the isolated subsystem ⌬0E1 and externally by the exchange with the reservoir ⌬s共r兲E1 = −⌬E1r. This is in contrast to the canonical Boltzmann distribution where the total energy is conserved, so that ⌬E0 = ⌬0E0 + ⌬s共r兲E0 = −⌬E0r, which is why the Boltzmann exponent is ⌬Sr / kB = −0E0. In the present case, the change in entropy of the reservoirs is ⌬Sr/kB = 0⌬E0r + 1⌬E1r ⬇ − 0⌬E0 − 1关⌬E1 − ⌬tĖ1兴 ⬇ − 0⌬E0 − 1关⌬E1 − An important aspect of any statistical system is the transitions between states, and nowhere is this more evident than for nonequilibrium systems. Indeed, Papers II and IV in the present series focused on this aspect of the problem and established a number of results and theorems that the transition probability must satisfy.2,4 Relevant to the present analysis is that the transition probability should 共1兲 yield a stationary steady state probability, 共2兲 obey the so-called microscopic transition theorem,4 and 共3兲 reduce to the second entropy expression for macrostate transitions.2 In the present case of steady heat flow, and indeed in any realistic statistical problem, it is essential to account for the stochastic perturbations from the reservoir in the transitions between states. Consider then the transition between the microstates ⌫ → ⌫⬙ in an infinitesimal time step ⌬t. This can be considered as comprising a deterministic transition ⌫ → ⌫⬘ due to the internal forces of the subsystem, followed by a stochastic transition ⌫⬘ → ⌫⬙ due to the perturbations by the reservoir. The deterministic transition is just the adiabatic evolution of the isolated subsystem, ⌫⬘ = ⌫ + ⌬t⌫˙ . In terms of the conditional transition probability this is ⌳d共⌫⬘兩⌫兲 = ␦共⌫⬘ − ⌫ − ⌬t⌫˙ 兲. 共30兲 The stochastic transition probability can be written as the product of an even and an odd function, ⌳s共⌫⬙兩⌫⬘兲 = ⌳e共⌫⬙兩⌫⬘兲⌳o共⌫⬙兩⌫⬘兲. 共31兲 In the case of transitions, however, parity now refers to the reversibility or irreversibility of the transition. That is, even transition probabilities satisfy ⌳e共⌫⬙兩⌫⬘兲 = ⌳e共⌫⬘†兩⌫⬙†兲, 共32兲 and odd transition probabilities satisfy ⌳o共⌫⬙兩⌫⬘兲 = 1/⌳o共⌫⬘†兩⌫⬙†兲. 共33兲 In Ref. 4, the odd transition probability for steady heat flow was given as = − 0⌬E0 − 1关⌬E1 − ⌬0E1兴 ⌬Wmir 1 兴. F. Transition probability 共29兲 From the final approximate expression, one sees that subtracting the change in Wmir 1 from the total change in moment is equivalent to identifying that part of the change in energy moment that is due to the reservoir. It is only this external influence that affects the reservoir entropy and that contributes to the steady state probability. In Sec. II F 1, which follows shortly, it is shown that the steady state probability distribution is approximately stationary during the adiabatic evolution of the isolated subsystem. This confirms the above interpretation that the exponent reflects the entropy of the reservoirs only and that the contribution from internal changes of the subsystem has been correctly removed. ⌳o共⌫⬙兩⌫⬘兲 = e−0共E0⬙−E0⬘兲/2e−1共E1⬙−E1⬘兲/2 . 共34兲 Notice that the exponent is the difference between two phase space functions of an even parity, and that this is necessary for the transition itself to be odd. Hopefully this will not cause confusion. The even stochastic transition probability may be taken as ⌳e共⌫⬙兩⌫⬘兲 = ⍜⌬共兩⌫⬙ − ⌫⬘兩兲. 共35兲 Here ⍜⌬ is a short ranged, even function 共such as a Heaviside step function or a Gaussean兲, with a range ⌬ that represents the strength of the stochastic perturbations from the reservoir. It is normalized and does not depend on the temperatures to a leading order. Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-6 J. Chem. Phys. 124, 224103 共2006兲 Phil Attard evolution, E0共⌫⬘兲 = E0共⌫兲, one can concentrate on the moments. For an infinitesimal time step ⌬t the first moment changes to One can give a straightforward physical interpretation to this result for the stochastic transition probability. The change in entropy of the reservoirs during the transition ⌫ → ⌫⬙ is ⌬S/kB = − 0⌬E0 − 1关⌬E1 − ⌬0E1兴 E1共⌫⬘兲 = E1共⌫兲 + ⌬tĖ1共⌫兲, = − 0共E0⬙ − E0兲 − 1共E1⬙ − E1 − ⌬tĖ1兲 = − 0共E0⬙ − E0⬘兲 − 1共E1⬙ − E1⬘兲. 共36兲 共39兲 and the mirror work becomes Hence the stochastic conditional transition probability is just the exponential of half the change in the reservoir entropy, mir Wmir 1 共⌫⬘兲 = W1 共⌫兲 + ⌬tĖ1共E1共⌫兲兲. ⌳s共⌫⬙兩⌫⬘兲 = ⍜⌬共兩⌫⬙ − ⌫⬘兩兲e−0共E0⬙−E0⬘兲/2e−1共E1⬙−E1⬘兲/2 = ⍜⌬共兩⌫⬙ − ⌫⬘兩兲e ⌬S/2kB The overbar signifies the most likely state, and it is in such states that the steady state probability is significantly greater than zero. Conversely, states with Ė1 ⫽ Ė1 occur rarely, with the probability getting smaller as the difference increases. 共Fluctuations away from the equilibrium or the steady state are relatively negligible.6兲 The formal result is 共37兲 . It is possible to multiply the even stochastic transition mir⬙ 共40兲 mir⬘ probability by a factor of e1共W1 −W1 兲/2, since the difference between the two exponents of the odd parity gives a transition of the even parity. All of the results given below would be unchanged if this factor were to be added. This includes the microscopic transition theorem established in Ref. 4, which is independent of the even transition probability. mir E1共⌫兲 − Wmir 1 共⌫兲 = E1共⌫⬘兲 − W1 共⌫⬘兲 − ⌬t关Ė1共⌫⬘兲 − Ẇmir 1 共⌫⬘兲兴. 共41兲 1. Stationary steady state probability In the most likely state the final term vanishes, and for nearby states it is small. Hence the present steady state probability density is almost stationary during its adiabatic evolution, For steady heat flow the probability density does not depend explicitly on time, and so it must be stationary under the transition probability given above. Consider then the evolution of the probability in time ⌬t, which is given by 㜷̃ss共⌫⬙兩0, 1兲 = = 冕 冕 㜷ss共⌫兩0, 1兲 = 㜷ss共⌫⬘兩0, 1兲e1⌬t关Ė1共⌫⬘兲−Ẇ1 mir d⌫⬘d⌫⌳s共⌫⬙兩⌫⬘兲⌳d共⌫⬘兩⌫兲㜷ss共⌫兩01兲 and it is necessary to show that the left hand side is indeed 㜷ss. First the deterministic evolution of the probability is explored. Since the energy is conserved during the Hamiltonian d⌫⬘⌳s共⌫⬙兩⌫⬘兲㜷ss共⌫⬘兩0, 1兲e1⌬t关Ė1共⌫⬘兲−Ẇ1 mir = 冕 d⌫⬘⍜⌬共兩⌫⬙ − ⌫⬘兩兲e mir⬙ e−0E0⬙e−1E1⬙e1W1 = Zss = 㜷ss共⌫⬙兩0, 1兲 冕 共42兲 共⌫⬘兲兴 −0共E0⬙−E0⬘兲/2 −1共E1⬙−E1⬘兲/2 e 冕 . 共To the leading order one can use ⌫ or ⌫⬘ in the final term.兲 The second part of the proof consists in showing that the present steady state probability density is stationary during the stochastic step. The proof follows that given previously in steady state4 and in equilibrium8 contexts, although some attention has to be paid to the change in probability during the adiabatic evolution. The stochastic transition integral becomes d⌫⬘⌳s共⌫⬙兩⌫⬘兲㜷ss共⌫0共− ⌬t兩⌫⬘兲兩0, 1兲, 共38兲 冕 共⌫⬘兲兴 ⬘ −0E0⬘ −1E1⬘ 1Wmir 1 e e e Zss mir⬘ d⌫⬘⍜⌬共兩⌫⬙ − ⌫⬘兩兲e1共W1 mir⬘ d⌫⬘⌳s共⌫⬘兩⌫⬙兲 ⫻ e1共W1 e1⌬t关Ė1共⌫⬘兲−Ẇ1 mir 共⌫⬘兲兴 mir ⬙ −Wmir 1 兲 −0共E0⬘−E0⬙兲/2 −1共E1⬘−E1⬙兲/2 1⌬t关Ė1共⌫⬘兲−Ẇ1 共⌫⬘兲兴 e e mir ⬙ −Wmir 1 兲e 1⌬t关Ė1共⌫⬘兲−Ẇ1 共⌫⬘兲兴 e = 㜷ss共⌫⬙兩0, 1兲. 共43兲 Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-7 This is the required result which shows the stationarity of the steady state probability under the present transition probability. The passage to the final equality requires comment. The normalization of the conditional transition probability, 兰d⌫⬘⌳s共⌫⬘ 兩 ⌫⬙兲 = 1, is unchanged to the order ⌬2 by the factor of e J. Chem. Phys. 124, 224103 共2006兲 Statistical mechanical theory for steady state systems. V mir mir 1共W1 ⬘−W1 ⬙兲 . Furthermore, it is assumed that 2 1⌬t关Ė1共⌫⬘兲 − Ẇmir 1 共⌫⬘兲兴 ⬃ O⌬ , 共44兲 2. Unconditional forward and reverse transitions In Paper IV, the so-called transition theorem and its reverse were given explicitly for the case of the mechanical work.4 In the present case of steady heat flow, the unconditional microscopic transition probability is ss共0, 1兲 e , 共45兲 for an infinitesimal ⌬t. The Ė1 in the final term can be replaced by 共Ė1⬙ + Ė1兲 / 2 to this order. Consider the forward transition ⌫ → ⌫⬘ → ⌫⬙, and its reverse, ⌫⬙† → ⌫† → ⌫†. Note that ⌫† ⫽ ⌫⬘†, but that 兩⌫⬙† − ⌫⬘兩 = 兩⌫† − ⌫†兩. The ratio of the forward to the reverse transition probabilities is 㜷共⌫⬙ ← ⌫兩⌬t兲 mir mir = e1共W1 共⌫⬙兲+W1 共⌫兲兲e⌬t1共Ė1共⌫⬙兲+Ė1共⌫兲兲/2 , † † 㜷共⌫ ← ⌫⬙ 兩⌬t兲 共46兲 the even terms canceling. The exponent of the final term is part of the change in the reservoir entropy 共the thermodynamic work done兲 on the trajectory. Consider a trajectory 关⌫兴 = 兵⌫0 , ⌫1 , . . . , ⌫ f 其, over a time interval t f = f⌬t, and its reverse, 关⌫‡兴 = 兵⌫†f , ⌫†f−1 , . . . , ⌫†0其. The adiabatic change in E1 over the trajectory is f−1 ⌬0E1关⌫兴 = ⌬t 关Ė1共⌫0兲 + Ė1共⌫ f 兲兴 + ⌬t 兺 Ė1共⌫i兲. 2 i=1 Clearly, ⌬0E1关⌫‡兴 = −⌬0E1关⌫兴. Then 0E /2 − 共Wmir−Wmir兲/2 1 1 1,f 1,0 e ⫻ 关㜷ss共⌫ f 兩0, 1兲㜷ss共⌫0兩0, 1兲兴1/2 , 共48兲 since E1,i−1 ⬘ = E1,i−1 + ⌬tĖ1,i−1. The exponent ⌬0E1 should really be replaced by ⌬0E1 + ⌬t关Ė1共⌫0兲 − Ė1共⌫ f 兲兴 / 2, but this correction is negligible for large f, and in any case it is fixed up by the ratio that is taken shortly. Briefly digress to consider a closely related problem, namely, that instead of the steady state the system is initially in the so-called static state, which has an even probability density 㜷st共⌫兩0, 1兲 = e−0E0共⌫兲e−1E1共⌫兲 . Zst共0, 1兲 共49兲 In this case the trajectory probability during the evolution to the steady state is f 0E /2 1 共50兲 The ratio of the probability of the forward and reverse trajectories in the steady state is mir 㜷关⌫兴 0 0 1Wmir 1 共⌫ f 兲e 1W1 共⌫0兲e 1⌬ E1关⌫兴 ⬇ e 1⌬ E1关⌫兴 ‡ =e 㜷关⌫ 兴 ⫻关㜷ss共⌫⬙兩0, 1兲㜷ss共⌫兩0, 1兲兴1/2 ˙ −Wmir 1 兲/2 ⌬t1E1/2 f = 兿 关⍜⌬共兩⌫i+1 − ⌫i⬘兩兲兴e1⌬ ⫻ 关㜷st共⌫ f 兩0, 1兲㜷st共⌫0兩0, 1兲兴1/2 . = ⍜⌬共兩⌫⬙ − ⌫⬘兩兲 mir⬙ ⬘ 兲/2兴㜷 共⌫ 兩 ,  兲 ⫻ e−1共E1,i−E1,i−1 ss 0 0 1 i=1 −0共E0⬙+E0兲/2 mir ˙ e−1共E1⬙+E1−⌬tE1兲/2e1W1 /Z ⫻ e−1共W1 i=1 㜷st关⌫兴 = 兿 关⍜⌬共兩⌫i+1 − ⌫i⬘兩兲兴e1⌬ 㜷共⌫⬙ ← ⌫兩⌬t兲 = ⌳s共⌫⬙兩⌫⬘兲㜷ss共⌫兩0, 1兲 ⫻ ⬘ 兲/2 㜷关⌫兴 = 兿 关⍜⌬共兩⌫i+1 − ⌫i⬘兩兲e−0共E0,i−E0,i−1 i=1 which is again second order and neglectable. The rationale for this result is that the fluctuations away from the most likely state are relatively negligible in the thermodynamic limit. Alternatively, one can arrange the second order terms in the normalization to cancel exactly this term. = ⍜⌬共兩⌫⬙ − ⌫⬘兩兲e f 共47兲 ⬇ e⌬S关⌫兴/kB , 共51兲 the even terms canceling. The final two approximations are valid for large t f , since the retained terms scale with f⌬t. The final exponent is the change in the entropy of the reservoirs 关Eq. 共36兲兴. This result holds exactly for trajectories evolving to the steady state from the static state, 㜷st关⌫兴 0 = e1⌬ E1关⌫兴 . 㜷st关⌫‡兴 共52兲 The result for the ratio of the probabilities of the forward and reverse trajectories may be called the reverse transition theorem.4 The probability of observing the entropy of the reservoir change by ⌬S over a period t is related to the probability of observing the opposite change by 㜷共⌬S兩0, 1,t f 兲 = 冕 d关⌫兴␦共⌬S − ⌬S关⌫兴兲共㜷关⌫兴兲 ⬇ kB−1 冕 d关⌫‡兴␦共⌬S/kB − 1⌬0E1关⌫兴兲 ⫻ 㜷关⌫‡兴e1⌬ 0E 关⌫兴 1 = e⌬S/kB㜷共− ⌬S兩0, 1,t f 兲. 共53兲 This result says in essence that the probability of a positive Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-8 J. Chem. Phys. 124, 224103 共2006兲 Phil Attard increase in entropy is exponentially greater than the probability of a decrease in entropy during heat flow, with the exponent being the entropy change under consideration. The change in entropy of the reservoirs may be regarded as a type of a thermodynamic work performed by the reservoirs during heat flow. Hence, in essence, this is the temperature gradient version of the fluctuation theorem that was originally derived by Evans et al.9 and Evan10 A derivation has also been given by Crooks,11,12 and the theorem has been verified experimentally.13 The present derivation based on the microscopic transition probability closely follows that given by Attard.4 Closely related to the fluctuation theorem is the work theorem due to Jarzynski,14 which has also been rederived in different fashions15,16,10 and verified experimentally.17 In this context one is motivated to consider the average of the exponential of the heat flux, 具e−1⌬ 0E 1 典ss,t f = = 冕 冕 d关⌫兴e−1⌬ mir 冕 冕 = 冉 = ln 冕 d⌫ exp关− 0E0共⌫兲 − 1E1共⌫兲 + 1兵E+1 共⌫兲 − E−1 共⌫兲其/2兴␦共E1 − E1共⌫兲兲 ⫻␦共E+1 − E+1 共⌫兲兲␦共E−1 − E−1 共⌫兲兲 + − + − = S共2兲 0 共E1 ,E1,E1 兩兲/kB − 1E1 + 1兵E1 − E1 其/2. + − S共2兲 0 共E1 ,E1,E1 兩兲 = 共⌫ f 兲+Wmir 1 共⌫0兲兲 d⌫†f 㜷共⌫†f 兩0, 1兲e ⫻ 共2兲 共E+1 ,E1,E−1 兩兲/kB Stotal Zst共0, 1兲 Zss共0, 1兲 † −1Wmir 1 共⌫ f 兲 mir 冊 共⌫†0兲 2 . ss + ss 共E − E1兲2 + 共E1 − E−1 兲2 2 1 2 + d⌫†0㜷共⌫†0兩0, 1兲e−1W1 共55兲 The second entropy of the isolated system during the adiabatic sequential transition is 㜷共关⌫兴兩0, 1,t f 兲 d关⌫‡兴㜷共关⌫‡兴兩0, 1,t f 兲 ⫻e1共W1 ⬇ 0E 关⌫兴 1 regarded simply as the difference in two energies over a time interval of 2.兲 The exponent represents the reservoir contribution to the total entropy, since a subsystem microstate, a point in its phase space, has zero entropy. Hence the total second entropy for this sequential transition is just the logarithm of the sum over the macrostate of the weights of the microstates, the latter being the unnormalized steady state probability, 共54兲 Here it has been assumed that the trajectory is long enough that the ends are uncorrelated. This result shows that this particular average is not extensive in time, 共i.e., it does not scale with t f 兲. In essence the right hand side is the exponential of twice the difference in free energies of the static and the steady state systems. 3. Second entropy In Paper II the so-called second entropy was introduced, and this was said to be the appropriate entropy for transitions between macrostates, and from it the Onsager reciprocal relations were deduced as well as the Green-Kubo theory for the linear transport coefficients.2 In Paper IV the second entropy for macrostates was used to formulate a transition probability for microstates that became the basis for the microscopic transition theorem and the microscopic reverse transition theorem.4 Now the relationship between the steady state probability density and the second entropy is derived for the present case of heat flow. Three relevant quantities appear in the steady state probability density 关Eq. 共9兲兴: E1, E+1 , and E−1 , since it will be recalled that the mirror work term can be written as Wmir 1 = 共E+1 − E−1 兲 / 2. Accordingly, the steady state probability density really describes the sequential transition, E−1 → E1 → E−1 , with each interval of duration . 共Note that Wmir 1 should always be interpreted in this fashion, and it should never be S共1兲 2 E, 2 1 ⬎ short . 共56兲 This generalizes slightly the result given previously2 from a single to a sequential transition. The final term can be replaced by the first or ordinary entropy S共1兲共E1兲, which in turn to the leading order can be replaced by 关S共1兲共E−1 兲 + 2S共1兲共E1兲 + S共1兲共E+1 兲兴 / 4 without changing the following results for Ė1 over the entire interval. The derivatives of the total second entropy are 共2兲 Stotal 共E+1 ,E1,E−1 兩兲 E+1 共2兲 Stotal 共E+1 ,E1,E−1 兩兲 E−1 = ss E+1 − E1 1 + , 2 = − ss E1 − E−1 1 − , 2 共57兲 共2兲 共E+1 ,E1,E−1 兩兲 Stotal E+ − E1 E1 − E−1 = − ss 1 + ss − 1 E1 + S共1兲E1 . Maximizing the second entropy by setting these derivatives to zero, ones sees that the most likely flux is ⴰ ⴰ E+1 = E−1 = − −1 ss 1 , 2 共58兲 and that the most likely moment is ⴰ E1 = 共S共1兲兲−11 . 共59兲 共The circle denotes the coarse-grained velocity,2 ˱1 ⬅ ± 共E±1 − E1兲 / .兲 These agree with earlier results.1,2 As mentioned above, the second entropy form of the transport coefficient is related to the thermal conductivity by ss = −1 / 2VT20. Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-9 J. Chem. Phys. 124, 224103 共2006兲 Statistical mechanical theory for steady state systems. V One may conclude two things from this analysis. First, the present steady state probability distribution is entirely consistent with the second entropy analysis of the heat transport. This demonstrates again the necessity of the mirror work term Wmir 1 . Second, the steady state probability distribution really describes transitions rather than phase space microstates per se, since the exponent is so closely related to the second entropy. This surprising result contrasts with equilibrium systems where it is the first entropy or state weight that is relevant to the probability density. III. NONEQUILIBRIUM MONTE CARLO SIMULATION A. Algorithm Similar to earlier papers in the series,1,3,4 a LennardJones fluid was simulated. All quantities were made dimensionless using the welldepth ⑀LJ, the diameter LJ, and the 2 time constant LJ = 冑 共mLJLJ / ⑀LJ兲, where mLJ is the mass. In addition, Boltzmann’s constant was set equal to unity. The Lennard-Jones potential between atoms was cut and shifted at Rcut = 2.5, and no tail correction was invoked. A spatial neighbor table was used with cubic cells of sidelength ⬇0.6, which reduces the number of neighbors required for a force calculation by almost a factor of 3 compared to the conventional cells of length Rcut.1 Periodic boundary conditions and the minimum image convention were used. Both a uniform bulk fluid and an inhomogeneous fluid were simulated. The latter was in the form of a slit pore, terminated in the z direction by uniform Lennard-Jones walls. The distance between the walls for a given number of atoms was chosen so that the uniform density in the center of the cell was equal to the nominal bulk density. Umbrella sampling Monte carlo simulations were performed in 6N-dimensional phase space, where N = 120– 500 atoms. The Metropolis algorithm was used with an umbrella weight density ˙ 共⌫兲 = e−0E0共⌫兲e−1E1共⌫兲e␣1E1共⌫兲 . 共60兲 It is emphasized that this is the umbrella weight used in the Metropolis sampling scheme; the exact steady state probability density 关Eq. 共9兲兴 was used to calculate the averages 共see below兲. The final term obviously approximates 1Wmir 1 , but is about a factor of 400 faster to evaluate. In the simulations reported here ␣ was fixed at 0.08. It would be possible to optimize this choice or to determine ␣ on the fly. 共See, for example, Ref. 3.兲 A trial move of an atom consisted of a small displacement in its position and momentum simultaneously. The step lengths were chosen to give an acceptance rate of about 50%. A cycle consisted of one trial move of each atom. Averages were collected after every 50 cycles. Labeling the current configuration used for an average by i, the Hamiltonian trajectory ⌫0共t 兩 ⌫i兲 was generated forward and backward in time using a second order rule and a time step of ⌬t = 10−2, which gave a satisfactory energy conservation. The running integral for Wmir 1 共⌫i ; t兲 was calculated along the trajectory using both the trapezoidal rule and Simpson’s rule, with indistinguishable results. The average flux was calculated as a function of the time interval, FIG. 1. The dependence of the thermal conductivity on the time interval for 2 the mirror work Wmir 1 共⌫ ; 兲. The curves are 共兲 = 具Ė1共0兲典 / VkBT01 for densities of, from bottom to top, 0.3, 0.5, 0.6, and 0.8, and T0 = 2. ˙ 具Ė1典 = mir 兺 iĖ1共⌫i兲e−␣1E1共⌫i兲e1W1 兺 ie 共⌫i;兲 −␣1E˙1共⌫i兲 1Wmir 1 共⌫i;兲 e . 共61兲 Notice how the umbrella weight used in the Metropolis scheme is canceled here. The thermal conductivity is reported below as 共兲 = 具Ė1典 / 1VkBT20. Not only is the umbrella method orders of magnitude faster in generating configurations, but it also allows results as a function of to be collected, and it reduces the correlation between consecutive costly trajectories by inserting many cheap umbrella steps. Prior to the generation of each trajectory the velocities of the particles were scaled and shifted at a constant kinetic energy to give a zero total z momentum. In the inhomogeneous system, a constraint force was added to keep the total z momentum zero on a trajectory. Of the order of 50 000 trajectories were generated for each case studied. B. Results for heat flow Figure 1 tests the dependence of the thermal conductivity on the time interval used to calculate Wmir 1 共⌫ ; 兲. It can be seen that the thermal conductivity is independent of the integration limit for Wmir 1 for ⲏ 1. This asymptotic or plateau value is the thermal conductivity. The value of required to reach the respective plateaus here appears comparable to straight Green-Kubo equilibrium calculations,3 but the present steady state simulations used about one third of the number of trajectories for a comparable statistical error. In the text it was assumed that the change in moment over the relevant time scales was negligible, 兩Ė1兩 Ⰶ 兩E1兩. In the case of = 0.8 at the largest value of in Fig. 1, 具E1典ss = −432 and 具Ė1典ss = 161, and so this assumption is valid in this case. Indeed, the reason for making this assumption was that on long time scales the moment must return to zero and the rate of change of moment must begin to decrease. There is no evidence of this occurring in any of the cases over the full interval shown in Fig. 1. Table I shows the values of the relaxation time calculated using Eqs. 共16兲 and 共17兲. Both the inertial time and the long time decrease with the increasing density. This is in agreement with the trend of the curves in Fig. 1. Indeed, the actual estimates of the relaxation times in Table I are in semiquantitative agreement with the respective boundaries of Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-10 J. Chem. Phys. 124, 224103 共2006兲 Phil Attard TABLE I. Thermal conductivity and relaxation times for various densities at T0 = 2. The standard errors of the last few digits are in parentheses. short Eq. 共16兲 long Eq. 共17兲 0.3 0.5 0.6 0.8 1.63共8兲 2.78共13兲 3.76共16兲 7.34共18兲 0.404共19兲 0.233共11兲 0.197共9兲 0.167共4兲 3.22共16兲 5.31共34兲 3.41共18兲 1.36共3兲 the plateaux in Fig. 1. The estimate of long, the upper limit on that may be used in the present theory, is perhaps a little conservative. Figure 2 compares the thermal conductivity obtained from the present nonequilibrium Monte Carlo simulations with previous NEMD results.18,19 The good agreement between the two approaches validates the present phase space probability distribution. The number of time steps required for an error of about 0.1 was about 3 ⫻ 107 共typically 2 ⫻ 105 independent trajectories, each of about 75 time steps forward and backward to get into the intermediate regime兲. This obviously depends on the size of the applied thermal gradient, 共the statistical error decreases with the increasing gradient兲 but appears comparable to that required by NEMD simulations.19 No attempt was made to optimize the present algorithm in terms of the number of Monte Carlo cycles between trajectory evaluations or the value of the umbrella parameter. Figure 2 also shows results for the thermal conductivity obtained for the slit pore, where the simulation cell was terminated by uniform Lennard-Jones walls. The results are consistent with those obtained for a bulk system using periodic boundary conditions. This indicates that the density inhomogeneity induced by the walls has little effect on the thermal conductivity. For the bulk system, the minimum image convention was used for all separations that appeared in the expression for Ė1.3 Figure 3 explores the nonlinear dependence of the heat flux on the applied temperature gradient. The increase in with the increasing 1 is due primarily to the first nonlinear term 关Eq. 共22兲兴. This represents a coupling of the induced density gradient to the heat flux. Nonlinear effects appear FIG. 3. The nonlinear thermal conductivity 共1兲 = 具Ė1共0兲典1 / VkBT201, for T0 = 2 共mainly for a cubic cell with N = 120兲. In 共a兲, the density is = 0.8 and the fitted quadratic is 共1兲 = 7.21+ 41321. In 共b兲, the density is = 0.6 and the fitted quadratic is 共1兲 = 3.61+ 55421. greater for = 0.6 than for = 0.8. The largest gradient shown corresponds to 1011 K / m in argon, which may be difficult to achieve in the laboratory. In this case the temperature discontinuity across the periodic z boundaries is 1kB / ⑀LJ, and it is not clear how this discontinuity and other finite size effects affect the nonlinear conductivity. IV. GENERALIZATIONS A. Hydrodynamic transport Heat flow is a particular example of a hydrodynamic transport. The theory for nonequilibrium statistical mechanics and the expression for the steady state probability distribution given above for heat flow may be readily generalized to other types of flow. It may also be generalized beyond the single gradient situation to include reservoirs that impose variations on short spatial wavelengths. 1. Flow in general Let A0 and A1 be the zeroth and first moments, respectively, of a set of linear additive variables that the subsystem may exchange with the reservoirs, 共energy, number, volume, charge, etc.兲. These could represent vectors whose components correspond to two or three spatial dimensions, or the components could represent different exchangeable variables. Let FIG. 2. Thermal conductivity at T0 = 2. The circles and squares are the present steady state results for bulk and inhomogeneous systems, respectively 共horizontally offset by ±0.015 for clarity兲, and the triangles are NEMD 共Refs. 18 and 19兲 results. b␣ ⬅ S , A␣ ␣ = 0,1 共62兲 be the conjugate thermodynamic variables 共one divided by temperature, minus chemical potential divided by tempera- Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-11 ture, etc.兲. Then the steady state probability distribution is e 㜷ss共⌫兩b0,b1兲 = −b0·A0共⌫兲/kB −b1·A1共⌫兲/kB b1·Wmir 1 共⌫兲/kB e e h N!Zss共b0,b1兲 3N , 共63兲 where Zss is the normalizing partition function, and where the mirror work is Wmir 1 共⌫兲 = 1 2 冕 − dtȦ1共⌫0共t兩⌫兲兲. 共64兲 By the conservation rules for the linear additive variables, Ȧ0 = 0 for an isolated system, and so one could formally include Wmir 0 = 0 without changing this result for the probability distribution 共see next兲. 2. Short spatial wavelengths The preceding subsection and the body of the text are applicable when the reservoirs impose only long wavelength variations in the field parameters. As shown in Sec. II of Ref. 1, the first moment is just the leading term in a polynomial expansion that can accommodate variations on all spatial wavelengths. In particular, if b共r兲 is the value of the field variables imposed by the reservoirs at position r, and if A共r兲 is the macrostate of the isolated subsystem consisting of specified values of the conjugate thermodynamic variables at that position, then − 冕 J. Chem. Phys. 124, 224103 共2006兲 Statistical mechanical theory for steady state systems. V drb共r兲 · A共r兲 = − 兺 bn · An 共65兲 n is the reservoir entropy associated with the subsystem macrostate. Here the summand represents the product of coefficients in an appropriate polynomial expansion of which the first two correspond in essence to the zeroth and first moments.1 Accordingly, the steady state probability distribution is 㜷ss共⌫兩关b兴兲 = 兰drb共r兲·Wmir共⌫,r兲/k e−兰drb共r兲·A共⌫,r兲/kBe h3NN!Zss关b兴 B , 1 2 冕 − dtȦ共⌫0共t兩⌫兲,r兲. ˜ 㜷̃共⌫兩,t兲 = Z̃−1e−H共⌫−,t−兲 = Z̃−1e−H共⌫,t兲eW共⌫,t兲 . 共68兲 where ⌫− = ⌫共t − 兩 ⌫ , t兲 is the starting point of the Hamiltonian 共adiabatic兲 trajectory, and the work done is W̃共⌫ , t兲 = H共⌫ , t兲 − H共⌫− , t − 兲. This is called the YamadaKawasaki distribution.20,21 The problem with it is that it does not take into account the influence of the heat reservoir while the work is being performed. A modified thermostatted form of the Yamada-Kawasaki distribution has been given, but it is said to be computationally intractable.22–24 Based on calculations performed with the Yamada-Kawasaki distribution, some have cast doubt on the very existence of a nonequilibrium probability distribution, since, they argue, that such a distribution is fractal in nature and that it shrinks onto a low-dimensional attractor.25–27 In view of these difficulties with the Yamada-Kawasaki distribution and its modifications, one seeks an alternative nonequilibrium probability distribution that is not restricted to an isolated system or that does not invoke artificial, deterministic thermostats. The distribution should properly account for the influence of the heat reservoir while the nonequilibrium work is being performed. This influence is stochastic, and its probability distribution is analogous to that described above for the case of heat flow down an imposed thermal gradient. Hence one requires a “mirror work” that has an odd parity. To obtain this one must continue the work path into the future by making it even about t, mir t 共t⬘兲 ⬅ 再 共t⬘兲, t⬘ 艋 t 共2t − t⬘兲, t⬘ ⬎ t. ⌫⬘ mir ⬅ ⌫mir共t⬘兩⌫,t兲 = t 共67兲 Note that, in effect, the zeroth mirror work has been included in the probability distribution by the integration over all space, but, as mentioned above, due to the conservation laws this is zero, and no error is introduced by doing this. t 共69兲 再 ⌫共t⬘兩⌫,t兲, t⬘ 艋 t ⌫共2t − t⬘兩⌫†,t兲† , t⬘ ⬎ t. 冎 共70兲 With these the mirror work is 1 Wmir共⌫,t兲 = 关Hmir共⌫+,t + 兲 − Hmir共⌫−,t − 兲兴 t t 2 = B. Nonequilibrium mechanical work The nonequilibrium probability distribution is now formulated for the case that a time-dependent work is performed on a subsystem while it is in contact with a thermal reservoir. Let  be the inverse temperature of the thermal reservoir, and consider a time-dependent Hamiltonian H共⌫ , t兲, where 共t兲 is the work parameter. If the system were isolated from the thermal reservoir during its evolution, and if the system were Boltzmann dis- 冎 Denote the corresponding Hamiltonian 共adiabatic兲 trajectory that is at ⌫ at t by 共66兲 where the mirror work is Wmir共⌫,r兲 = tributed at t − , then the probability distribution at time t would be = 冕 冕 1 2 1 2 t+ t− dt⬘Ḣmir共⌫⬘ mir,t⬘兲, t t t t− dt⬘关Ḣ共⌫共t⬘兩⌫,t兲,t⬘兲 − Ḣ共⌫共t⬘兩⌫†,t兲,t⬘兲兴, 共71兲 which clearly has an odd parity Wmir共⌫ , t兲 = −Wmir共⌫† , t兲. With it the nonequilibrium probability distribution for the subsystem of the thermal reservoir upon which work is being performed is Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-12 J. Chem. Phys. 124, 224103 共2006兲 Phil Attard mir e−H共⌫,t兲eW 㜷共⌫兩,t兲 = Z共,t兲 共⌫,t兲 共72兲 . The probability distribution is normalized by Z共 , t兲, which is a time-dependent partition function whose logarithm gives the nonequilibrium total entropy, which may be used as a generating function. Asymptotically one expects ¯ Ḣmir共⌫mir共t⬘兩⌫,t兲,t⬘兲 ⬃ sign共t⬘兲Ḣ共t兲, t t 兩t⬘兩 ⲏ short , 共73兲 which is the most likely rate of doing work at time t. This assumes that the change in energy is negligible on the rel¯ evant time scales 兩Ḣ兩 Ⰶ 兩H兩. Since this asymptote is odd in time, one concludes that the mirror work is independent of for in the intermediate regime, and that Wmir is dominated by the region t⬘ ⬇ t. In view of this and the fact that d⌫共t⬘ 兩 ⌫ , t兲 / dt = 0, the rate of change of the mirror work along a Hamiltonian trajectory is ¯ Ẇmir共⌫,t兲 = Ḣ共t兲. 共74兲 As for the heat flow, the probability distribution is stationary during the adiabatic evolution on the most likely points of the phase space. This result is significant in the context of the time evolution of the probability distribution that includes the stochastic perturbations from the thermal reservoir. As above, use a single prime to denote the adiabatic development in time ⌬t, ⌫ → ⌫⬘, and a double prime to denote the final stochastic position due to the influence of the reservoir, ⌫⬘ → ⌫⬙. The conditional transition probability may be taken to be ⌳共⌫⬙兩⌫,t兲 = ⍜⌬共兩⌫⬙ − ⌫⬘兩兲e−共H⬙ −H⬘ 兲/2 . 共75兲 The final, odd term is identical to that given previously,4 and hence this transition probability obeys the microscopic transition theorem 关see Eqs. 共9兲 and 共11兲 of Ref. 4兴. Hence this transition probability for the case of nonequilibrium work yields the fluctuation theorem9 and the work theorem,14 as was shown in Sec. I C of Ref. 4. This transition probability preserves the nonequilibrium phase space probability density 关Eq. 共72兲兴 during its time evolution, 㜷共⌫⬙兩,t + ⌬t兲 = 冕 d⌫⌳共⌫⬙兩⌫,t兲㜷共⌫兩,t兲. 共76兲 This result may be readily confirmed using the fact that H⬘ − Wmir⬘ = H − Wmir 共at least for those phase points most likely to occur兲, together with the usual normalization requirements on the transition probability. 关See the discussion surrounding Eq. 共44兲 above.兴 C. Nonequilibrium quantum statistical mechanics Consider a quantum system with a time-dependent Hamiltonian operator Ĥ共t兲. Nonequilibrium quantum statis- tical mechanics follows from a development analogous to the classical case. Define the mirror work operator Ŵmir共t兲 = 关Ê+共t兲 − Ê−共t兲兴/2, 共77兲 where the past and future energy operators are ʱ共t兲 = ⍜̂共⫿ ;t兲Ĥmir t 共t ± 兲⍜̂共± ;t兲 and where the time-shift operator is 冋 ⍜̂共 ;t兲 = exp −i ប 冕 t+ 册 dt⬘Ĥmir t 共t⬘兲 . t 共78兲 共79兲 The mirror Hamiltonian operator has been continued into the future, Ĥmir t 共t⬘兲 ⬅ 再 Ĥ共t⬘兲, t⬘ 艋 t Ĥ共2t − t⬘兲, t⬘ ⬎ t, 冎 共80兲 and the manipulation of the operators derived from it is famir cilitated by the symmetry about t, Ĥmir t 共t⬘兲 = Ĥt 共2t − t⬘兲. With these definitions, the nonequilibrium density operator for a subsystem of a thermal reservoir of an inverse temperature  is ˆ 共t兲 = 1 exp − 关Ĥ共t兲 − Ŵmir共t兲兴, Z共t兲 共81兲 where Z共t兲 is the normalization factor. Accordingly, the average of an observable at time t is 具Ô典t = tr兵ˆ 共t兲Ô共t兲其, 共82兲 and the present density operator can be said to provide a basis for nonequilibrium quantum statistical mechanics. V. CONCLUSION This paper has been concerned with developing a nonequilibrium probability distribution. A specific steady state system was studied in detail, namely, the heat flow down an imposed temperature gradient. The key difference from the Boltzmann distribution for an equilibrium system was a term of the odd phase space parity, the mirror work, which reflects the arrow of time inherent in nonequilibrium systems. The steady state probability distribution that was derived here was shown to be consistent with the Green-Kubo formula for the thermal conductivity and with the second entropy formula for macrostate transitions. A Monte Carlo algorithm was developed based upon the steady state probability distribution. This was shown to be computationally efficient and to yield known values for the thermal conductivity. Estimates of various relaxation times based on fluctuation formulas were shown to be semiquantitative. The nonequilibrium theory was extended generically beyond the heat flow. The probability distribution for the general hydrodynamic flow due to an imposed thermodynamic gradient was given, and this was further generalized to imposed fields with arbitrary spatial variations. The theory was also applied to the case of a time-dependent mechanical work, and a mirror work term with an odd phase space parity was identified for the probability density. The associated Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 224103-13 J. Chem. Phys. 124, 224103 共2006兲 Statistical mechanical theory for steady state systems. V transition probability satisfied the microscopic transition theorem,4 and hence the fluctuation9 and work theorems,14 and conserved the probability density. The analogous density operator for nonequilibrium quantum statistical mechanics based on a mirror Hamiltonian operator was also given. The expressions obtained in this paper appear to be formally exact for the steady state, although they may not be restricted only to the steady state. At a minimum they are likely to be good approximations for the quasi-steady-state 共thermodynamic gradients or rates of doing work that change relatively slowly with time兲, but they may prove even more general than this. For example, in the case of a nonequilibrium mechanical work, the present probability density can be explicitly shown to be correct for t ⬍ 0 and for t ⬎ , for a sudden change in the Hamiltonian at t = 0; it may even prove correct for all t. That it works in these regimes for this extreme example of a transient nonequilibrium behavior hints at the broad applicability of the present approach. The present probability densities possibly represent the optimum generic formulation for the universal nonequilibrium state with an arbitrary time dependence when there is an ideal thermal contact with the heat reservoir, and all other details of the reservoir can be ignored. ACKNOWLEDGMENT This work was financially supported by the Australian Research Council. P. Attard, J. Chem. Phys. 121, 7076 共2004兲. P. Attard, J. Chem. Phys. 122, 154101 共2005兲. 3 P. Attard, J. Chem. Phys. 122, 244105 共2005兲. 4 P. Attard, J. Chem. Phys. 124, 024109 共2006兲. 5 L. Onsager, Phys. Rev. 37, 405 共1931兲; 38, 2265 共1931兲. 6 P. Attard, Thermodynamics and Statistical Mechanics: Equilibrium by Entropy Maximisation 共Academic, London, 2002兲. 7 M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids 共Oxford University Press, Oxford, 1987兲. 8 P. Attard, J. Chem. Phys. 116, 9616 共2002兲. 9 D. J. Evans, E. G. D. Cohen, and G. P. Morriss, Phys. Rev. Lett. 71, 2401 共1993兲. 10 D. J. Evans, Mol. Phys. 101, 1551 共2003兲. 11 G. E. Crooks, Phys. Rev. E 60, 2721 共1999兲. 12 G. E. Crooks, Phys. Rev. E 61, 2361 共2000兲. 13 G. M. Wang, E. M. Sevick, E. Mittag, D. J. Searles, and D. J. Evans, Phys. Rev. Lett. 89, 050601 共2002兲. 14 C. Jarzynski, Phys. Rev. Lett. 78, 2690 共1997兲. 15 C. Jarzynski, Phys. Rev. E 56, 5018 共1997兲. 16 G. E. Crooks, J. Stat. Phys. 90, 1481 共1998兲. 17 J. Liphardt, S. Dumont, S. B. Smith, I. Tinoco, Jr., and C. Bustamante, Science 296, 1832 共2002兲. 18 D. J. Evans, Phys. Rev. A 34, 1449 共1986兲. 19 P. J. Daivis and D. J. Evans, Phys. Rev. E 48, 1058 共1993兲. 20 T. Yamada and K. Kawasaki, Prog. Theor. Phys. 38, 1031 共1967兲. 21 T. Yamada and K. Kawasaki, Prog. Theor. Phys. 53, 111 共1975兲. 22 G. P. Morriss and D. J. Evans, Mol. Phys. 54, 629 共1985兲. 23 G. P. Morriss and D. J. Evans, Phys. Rev. A 37, 3605 共1988兲. 24 D. J. Evans and G. P. Morriss, Statistical Mechanics of Nonequilibrium Liquids 共Academic, London, 1990兲. 25 B. L. Holian, G. Ciccotti, W. G. Hoover, B. Moran, and H. A. Posch, Phys. Rev. A 39, 5414 共1989兲. 26 G. P. Morriss, Phys. Rev. A 39, 4811 共1989兲. 27 D. J. Evans and D. J. Searles, Phys. Rev. E 52, 5839 共1995兲. 1 2 Downloaded 14 Jun 2006 to 129.78.64.100. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp
© Copyright 2026 Paperzz