ONSAGER S VARIATIONAL PRINCIPLE AND ITS APPLICATIONS. Abstract

Similar documents
ensemble and the canonical ensemble. The microcanonical ensemble is of fundamental

Fundamentals of Fluid Dynamics: Elementary Viscous Flow

Hydrodynamics. Stefan Flörchinger (Heidelberg) Heidelberg, 3 May 2010

Notes on Entropy Production in Multicomponent Fluids

Quick Recapitulation of Fluid Mechanics

n v molecules will pass per unit time through the area from left to

Introduction. Statistical physics: microscopic foundation of thermodynamics degrees of freedom 2 3 state variables!

7 The Navier-Stokes Equations

Entropy and the Second Law of Thermodynamics

Introduction Statistical Thermodynamics. Monday, January 6, 14

Grand Canonical Formalism

Hydrodynamics, Thermodynamics, and Mathematics

V (r,t) = i ˆ u( x, y,z,t) + ˆ j v( x, y,z,t) + k ˆ w( x, y, z,t)

Entropy generation and transport

ChE 210B: Advanced Topics in Equilibrium Statistical Mechanics

Navier-Stokes Equation: Principle of Conservation of Momentum

CHEMICAL ENGINEERING THERMODYNAMICS. Andrew S. Rosen

2 Equations of Motion

Engineering Thermodynamics. Chapter 6. Entropy: a measure of Disorder 6.1 Introduction

Onsager theory: overview

Final Review Prof. WAN, Xin

G : Statistical Mechanics

Thermodynamics & Statistical Mechanics

4. The Green Kubo Relations

6.2 Governing Equations for Natural Convection

Chapter 5. The Differential Forms of the Fundamental Laws

On Fluid Maxwell Equations

Getting started: CFD notation

CHAPTER 8 ENTROPY GENERATION AND TRANSPORT

Fundamentals of compressible and viscous flow analysis - Part II

[S R (U 0 ɛ 1 ) S R (U 0 ɛ 2 ]. (0.1) k B

Math 575-Lecture Viscous Newtonian fluid and the Navier-Stokes equations

Thermochemical Properties

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

1 Garrod #5.2: Entropy of Substance - Equation of State. 2

Contents. 1 Introduction and guide for this text 1. 2 Equilibrium and entropy 6. 3 Energy and how the microscopic world works 21

1. Thermodynamics 1.1. A macroscopic view of matter

CH.9. CONSTITUTIVE EQUATIONS IN FLUIDS. Multimedia Course on Continuum Mechanics

Today s Lecture: Atmosphere finish primitive equations, mostly thermodynamics

Euler equation and Navier-Stokes equation

PHYSICS 715 COURSE NOTES WEEK 1

MACROSCOPIC VARIABLES, THERMAL EQUILIBRIUM. Contents AND BOLTZMANN ENTROPY. 1 Macroscopic Variables 3. 2 Local quantities and Hydrodynamics fields 4

Chemistry. Lecture 10 Maxwell Relations. NC State University

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013 Notes on the Microcanonical Ensemble

ESCI 341 Atmospheric Thermodynamics Lesson 12 The Energy Minimum Principle

The Second Law of Thermodynamics (Chapter 4)

Review of fluid dynamics

ChE 503 A. Z. Panagiotopoulos 1

5. Systems in contact with a thermal bath

Lifshitz Hydrodynamics

AA210A Fundamentals of Compressible Flow. Chapter 5 -The conservation equations

- Marine Hydrodynamics. Lecture 4. Knowns Equations # Unknowns # (conservation of mass) (conservation of momentum)

Non equilibrium thermodynamics: foundations, scope, and extension to the meso scale. Miguel Rubi

THERMODYNAMICS THERMOSTATISTICS AND AN INTRODUCTION TO SECOND EDITION. University of Pennsylvania

Molecular Modeling of Matter

Potential Descending Principle, Dynamic Law of Physical Motion and Statistical Theory of Heat

Physics 408 Final Exam

A Brief Introduction to Statistical Mechanics

Symmetry of the Dielectric Tensor

Collective Effects. Equilibrium and Nonequilibrium Physics

Thermodynamics of phase transitions

MME 2010 METALLURGICAL THERMODYNAMICS II. Fundamentals of Thermodynamics for Systems of Constant Composition

CH 240 Chemical Engineering Thermodynamics Spring 2007

Lecture. Polymer Thermodynamics 0331 L First and Second Law of Thermodynamics

A SHORT INTRODUCTION TO TWO-PHASE FLOWS Two-phase flows balance equations

Lattice Boltzmann Method

Continuum Mechanics Fundamentals

Chapter 3. Property Relations The essence of macroscopic thermodynamics Dependence of U, H, S, G, and F on T, P, V, etc.

(# = %(& )(* +,(- Closed system, well-defined energy (or e.g. E± E/2): Microcanonical ensemble

Fluid Equations for Rarefied Gases

Chapter 2. General concepts. 2.1 The Navier-Stokes equations

2. Thermodynamics. Introduction. Understanding Molecular Simulation

CHAPTER 4. Basics of Fluid Dynamics

4.1 LAWS OF MECHANICS - Review

Part1B(Advanced Physics) Statistical Physics

1.3 Molecular Level Presentation

fiziks Institute for NET/JRF, GATE, IIT-JAM, JEST, TIFR and GRE in PHYSICAL SCIENCES

Fundamentals of Linear Elasticity

In this section, thermoelasticity is considered. By definition, the constitutive relations for Gradθ. This general case

Fluid Equations for Rarefied Gases

Classical thermodynamics

Shell Balances in Fluid Mechanics

The mathematical description of the motion of Atoms, Molecules & Other Particles. University of Rome La Sapienza - SAER - Mauro Valorani (2007)

REGULARIZATION AND BOUNDARY CONDITIONS FOR THE 13 MOMENT EQUATIONS

Chapter 2: Fluid Dynamics Review

UNIVERSITY OF SOUTHAMPTON

Section 2: Lecture 1 Integral Form of the Conservation Equations for Compressible Flow

Lecture 4: Entropy. Chapter I. Basic Principles of Stat Mechanics. A.G. Petukhov, PHYS 743. September 7, 2017

DSMC-Based Shear-Stress/Velocity-Slip Boundary Condition for Navier-Stokes Couette-Flow Simulations

Thermodynamics! for Environmentology!

Basic hydrodynamics. David Gurarie. 1 Newtonian fluids: Euler and Navier-Stokes equations

HYDRODYNAMIC BOUNDARY CONDITIONS FOR ONE-COMPONENT LIQUID-GAS FLOWS ON NON-ISOTHERMAL SOLID SUBSTRATES

CONTENTS 1. In this course we will cover more foundational topics such as: These topics may be taught as an independent study sometime next year.

Physics 607 Exam 2. ( ) = 1, Γ( z +1) = zγ( z) x n e x2 dx = 1. e x2

PHYS First Midterm SOLUTIONS

UNIVERSITY OF SOUTHAMPTON

Outline Review Example Problem 1 Example Problem 2. Thermodynamics. Review and Example Problems. X Bai. SDSMT, Physics. Fall 2013

arxiv:comp-gas/ v1 28 Apr 1993

CARNOT CYCLE = T = S ( U,V )

Review of Fluid Mechanics

Transcription:

ONSAGER S VARIAIONAL PRINCIPLE AND IS APPLICAIONS iezheng Qian Department of Mathematics, Hong Kong University of Science and echnology, Clear Water Bay, Kowloon, Hong Kong (Dated: April 30, 2016 Abstract his manuscript is prepared for a Short Course on the variational principle formulated by Onsager based on his reciprocal symmetry for the kinetic coefficients in linear irreversible thermodynamic processes. Based on the reciprocal relations for kinetic coefficients, Onsager s variational principle is of fundamental importance to non-equilibrium statistical physics and thermodynamics in the linear response regime. For his discovery of the reciprocal relations, Lars Onsager was awarded the 1968 Nobel Prize in Chemistry. he purpose of this short course is to present Onsager s variational principle and its applications to first-year graduate students in physics and applied mathematics. he presentation consists of four units: 1. Review of thermodynamics 2. Onsagers reciprocal symmetry for kinetic coefficients 3. Onsagers variational principle 4. Applications: 4.1 Heat transport 4.2 Lorentz reciprocal theorem 4.3 Cross coupling in rarefied gas flows 4.4 Cross coupling in a mixture of fluids Corresponding author. Email: maqian@ust.hk 1

I. A BRIEF REVIEW OF HERMODYNAMICS he microcanonical ensemble is of fundamental value while the canonical ensemble is more widely used. For more details, see C. Kittel, Elementary Statistical Physics. he energy is a constant of motion for a conservative system. If the energy of the system is prescribed to be in the range δe at E 0, we may form a satisfactory ensemble by taking the density as equal to zero except in the selected narrow range δe at E 0 : P (E = constant for energy in δe at E 0 and P (E = 0 outside this range. his particular ensemble is known as the microcanonical ensemble. It is appropriate to the discussion of an isolated system because the energy of an isolated system is a constant. Let us consider the implications of the microcanonical ensemble. We are given an isolated classical system with constant energy E 0. At time t 0 the system is characterized by definite values of position and velocity for each particle in the system. he macroscopic average physical properties of the system could be calculated by following the motion of the particles over a reasonable interval of time. We do not consider the time average, but consider instead an average over an ensemble of systems each at constant energy within δe at E 0. he microcanonical ensemble is arranged with constant density in the region of phase space accessible to the system. Here we have made as a fundamental postulate the assumption of equal a priori probabilities for different accessible regions of equal volume in phase space. A. Entropy he entropy of the microcanonical ensemble is S = k B log Γ, where Γ is the number of states with energy distributed between E 0 and E 0 + E. Remember that the general definition of the entropy is S = k B n w n log w n, where k B is the Boltzmann constant and w n is the distribution function for the system, satisfying n w n = 1, with n being the set of all quantum numbers which denote the various stationary states of the system. he entropy of the microcanonical ensemble is obtained from a maximization of S = k B n w n log w n with respect to w n, subject to the normalization condition n w n = 1 and the energy condition that w n is nonzero only if the corresponding energy is in the selected narrow range δe at E 0. 2

B. Conditions for equilibrium he entropy is a maximum when a closed system is in equilibrium. he value of S for a system in equilibrium depends on the energy U of the system; on the number N i of each molecular species i of the system; and on external variables x ν, such as volume V, strain, magnetization. In other words, S = S(U, x ν, N i. We consider the condition for equilibrium in a system made up of two interconnected subsystems, as illustrated in figure 1. Initially the subsystems are separated from each other by a insulating, rigid, and non-permeable barrier. Weakly interacting quasi-closed subsystems: In the following discussion, we consider two weakly interacting quasi-closed subsystems 1 and 2. On the one hand, the two subsystems are quasi-closed, and hence S j = S j (U j, V j, N ij for j = 1, 2 and S = S 1 + S 2 from the statistical independence. On the other hand, weak coupling is assumed between the two subsystems. his allows the whole (closed system to relax towards complete equilibrium (e.g., via energy transfer between the two subsystems if the total entropy is not maximized (with respect to U 1 or U 2. 1. hermal equilibrium Imagine that the barrier is allowed to transmit energy (beginning at one instant of time, with other inhibitions remaining in effect. If the conditions of the two subsystems 1 and 2 do not change, then they are in thermal equilibrium and the entropy of the total system must be a maximum with respect to small transfer of energy from one subsystem to the other. Using the additive property of the entropy, S = S 1 + S 2, we have in equilibrium or δs = δs 1 + δs 2 = 0, S1 δs = δu 1 + U 1 ( S2 U 2 δu 2 = 0. 3

Because the total system is thermally closed and the total energy is constant, i.e., δu = δu 1 + δu 2 = 0, we have [ ] S1 S2 δs = δu 1 = 0. U 1 U 2 As δu 1 is an arbitrary variation, we obtain in thermal equilibrium Defining the temperature by S 1 U 1 = S 2 U 2. 1 S =, U x ν,n i we obtain 1 = 2 as the condition for thermal equilibrium. Suppose that the two subsystems were not originally in thermal equilibrium, but that 2 > 1. When thermal contact is established to allow energy transmission, the total entropy S will increase. (he removal of any constraint can only increase the volume of phase space accessible to the system. hus, after thermal contact is established, δs > 0, or [ ] S1 S2 δu 1 > 0, U 1 U 2 and [ 1 1 ] δu 1 > 0. 1 2 Assuming 2 > 1, we have δu 1 > 0. his means that energy passes from the system of high to the system of low. So is indeed a quantity that behaves qualitatively like a temperature. 2. Mechanical equilibrium Now imagine that the wall is allowed to move and also transmit energy, but does not pass particles. he volumes V 1 and V 2 of the two subsystems can readjust to (further maximize the entropy. In mechanical equilibrium S1 S2 S1 δs = δv 1 + δv 2 + δu 1 + V 1 V 2 U 1 After thermal equilibrium has been established, we have S1 S2 δu 1 + δu 2 = 0. U 1 U 2 4 ( S2 U 2 δu 2 = 0.

As the total volume V = V 1 + V 2 is constant, we have δv = δv 1 + δv 2 = 0, and [ ] S1 S2 δs = δv 1 = 0. V 1 V 2 As δv 1 is an arbitrary variation, we obtain in mechanical equilibrium S 1 V 1 = S 2 V 2. Defining the pressure Π by Π S =, V U,N i we see that for a system in thermal equilibrium (with 1 = 2, Π 1 = Π 2 is the condition for mechanical equilibrium. In general we define a generalized force X ν related to the coordinate x ν by the equation X ν S =. x ν U,N i It is interesting to note that from the defining equations for and Π, we obtain Π = (S/V U,N i (S/U V,Ni = ( U V his expression for the pressure will be derived again by regarding U as a function of S and V. S,N i. Suppose that the two subsystems in thermal equilibrium were not originally in mechanical equilibrium, but that Π 1 > Π 2. When the wall is allowed to move, the total entropy S will increase. (he removal of any constraint can only increase the volume of phase space accessible to the system. From [ ] S1 S2 δs = δv 1 = 1 V 1 V 2 [Π 1 Π 2 ] δv 1 > 0, we see that Π 1 > Π 2 requires δv 1 > 0: the subsystem of the higher pressure expands in volume. So Π is indeed a quantity that behaves qualitatively like a pressure. 3. Particle equilibrium Now imagine that the wall allows diffusion through it of molecules of the ith chemical species. Suppose thermal equilibrium and mechanical equilibrium have already been established. From δn i1 + δn i2 = 0 and [ ( S1 S2 δs = N i1 5 N i2 ] δn i1 = 0,

we obtain S 1 N i1 = S 2 N i2, as the condition for particle equilibrium. Defining the chemical potential µ i by µ i S = N i we see that for a system in both thermal equilibrium ( 1 = 2 and mechanical equilibrium (Π 1 = Π 2, µ i1 = µ i2 is the condition for particle equilibrium. It is easy to show that particles tend to move from a region of higher chemical potential to that of a lower chemical U,x ν, potential as the system approaches equilibrium (δs > 0. C. Connection between statistical and thermodynamical quantities For a system in equilibrium, S = S(U, x ν, N i, where U is the energy, x ν denotes the set of external parameters describing the system, and N i is the number of molecules of the i-th species. If the conditions are changed slightly, but reversibly in such a way that the resulting system is also in equilibrium, we have S ds = du + S dx ν + U x ν ν i which may be rewritten as ( S N i dn i = du + 1 X ν dx ν 1 µ i dn i, ν i du = ds ν X ν dx ν + i µ i dn i. Consider that the number of particles is fixed and the volume is the only external parameter: dn i = 0; x ν V ; X ν Π. hen, du = ds ΠdV. We see that the change of internal energy consists of two parts. he term ds represents the change in U when the external parameters are kept constant (dv = 0. his is what we mean by heat. hus DQ = ds is the quantity of heat added to the system in a reversible process. he symbol D is used instead of d because DQ is not an exact differential that is, Q is not a state function. he 6

term ΠdV is the change in internal energy caused by the change in external parameters; this is what we mean by mechanical work, and DW = ΠdV is the work done on the system through the volume change dv. By elementary mechanics DW = pdv, hence Π p. herefore the change in internal energy may be expressed as which is he First Law of hermodynamics. du = DQ + DW, hat ds = DQ/ is an exact differential in a reversible process is a statement of he Second Law of hermodynamics. hat is, DQ/ is a differential of a state function, entirely defined by the state of the system. Now we know that state function is the entropy. On S = S(U, V : Here the discussion on various thermodynamic quantities starts from S = S(U, V in equilibrium and ds = du/ + pdv/ for reversible processes. he evaluation of S as a function of U and V can be carried out using the microcanonical ensemble which is characterized by U and V. We are ready to express U as a function of S and V : U = U(S, V. Other quantities of interest are then obtained from U(S, V : U U du = ds pdv = ds + dv, S V V S from which U = ; S V and U p =. V S (Note that the expression for p has been obtained before by starting from S = S(U, V. However, S and V are often inconvenient independent variables it is more convenient to work with (, V or (, p. For this purpose we introduce some thermodynamic potentials: F and G. 1. Helmholtz free energy When the temperature is set, it is convenient to introduce the Helmholtz free energy F which is defined as F U S, regardless of whether the system is in equilibrium or not 7

(as long as the entropy can be defined via partial equlibria. For a system in equilibrium, the macroscopic state is completely determined by and V, and F may be expressed as a function of these two independent variables describing the equilibrium state, i.e., U where = S ds pdv and V F (, V U S, has been used to replace S as an independent variable. From du = df = du ds Sd = Sd pdv = F F d + dv, V V we obtain F S =, V and F p =. V herefore, if the Helmholtz free energy F is expressed as a function of independent variables and V, then S and p are readily calculated from its partial derivatives. On the Helmholtz free energy F : While the definition of F U S is applicable to arbitrary macroscopic states (so long as there is partial equilibrium and hence the entropy can be defined, the total differential df = Sd pdv is obtained for reversible processes only. hat is, S = (F / V and p = (F /V are relations for equilibrium states only. For an equilibrium system of fixed and V, F (, V can be obtained from the partition function of canonical ensemble. Consider a thermodynamic process at constant temperature. he change in F is given by F = U S = ( Q S + W. For a reversible process, Q = S, and hence F = W. For an irreversible process, however, Q < S according to the second law of thermodynamics, and hence F < W. hat W F means the maximum work which can be done by the system is the decrease of F. If the volume is fixed, then W = 0 and F 0. hat is, at constant temperature and volume, the Helmholtz free energy varies towards a minimum, which is just F (, V for the final equilibrium state. On Q S: Consider a system in contact with a large heat bath such that a constant temperature is maintained. If Q is the heat received by the system from the heat bath, then the heat received by the heat bath from the system is Q. Assuming that the (very small 8

change in the (very large heat bath is reversible, we have Q/ as the change of its entropy. hat the total entropy tends to increase means Q/ + S 0, i.e., Q S. 2. Gibbs free energy When the temperature and the pressure p are both set, it is convenient to introduce the Gibbs free energy G which is defined as G F + pv U S + pv, regardless of whether the system is in equilibrium or not (as long as the entropy can be defined via partial equlibria. For a system in equilibrium, the macroscopic state is completely determined by and p, and G may be expressed as a function of these two independent variables describing the equilibrium state, i.e., G(, p U S + pv, U U where = and p = has been used to replace S and V as the two S V V S independent variables. From du = ds pdv and G G dg = du ds Sd + pdv + V dp = Sd + V dp = d + dp, p p we obtain S = G, p and G V =. p herefore, if the Gibbs free energy G is expressed as a function of independent variables and p, then S and V are readily calculated from its partial derivatives. Consider a thermodynamic process at constant temperature and pressure p. he change in G is given by G = U S + p V = ( Q S + W + p V = Q S. For a reversible process, Q = S, and hence G = 0. For an irreversible process, Q < S, and hence G < 0. herefore, at constant temperature and pressure, the Gibbs free energy tends to a minimum. his minimum is G(, p for the equilibrium state at the given temperature and pressure. 9

II. ONSAGER S VARIAIONAL PRINCIPLE A. Reciprocal relations for linear irreversible thermodynamic processes he heat flux J induced by temperature gradient is given by the constitutive equations 3 J i = λ ij j (i = 1, 2, 3, j=1 where λ ij are coefficients of heat conductivity. he heat conductivity tensor is symmetric even in crystals of low symmetry (Stokes 1851. B. Onsager s reciprocal symmetry derived from microscopic reversibility For a closed system, consider the fluctuations of a set of (macroscopic variables α i (i = 1,..., n with respect to their most probable (equilibrium values. he entropy of the system S has a maximum S e at equilibrium so that S = S S e can be written in the quadratic form, S (α 1,..., α n = 1 2 n β ij α i α j, where β is symmetric and positive definite. he probability density at α i (i = 1,..., n is given by f (α 1,..., α n = f (0,..., 0 e S/k B where k B is the Boltzmann constant. he forces conjugate to α i are defined by i,j=1 X i = S α i n = β ij α i j=1 which are linear combinations of α i not far from equilibrium. Following the above definition of the forces, the equilibrium average (over the distribution function f (α 1,..., α n of α i X j is given by α i X j = k B δ ij. Microscopic reversibility leads to the equality α i (t α j (t + τ = α j (t α i (t + τ 10

for time correlation functions. In a certain domain not far from equilibrium, the macroscopic variables α i (i = 1,..., n satisfy the linear equations d dt α i (t = n M ik α k (t = k=1 n L ij X j (t. Here the Onsager kinetic coefficient matrix L is related to the rate coefficient matrix M via the relation Lβ = M. Onsager s hypothesis is that fluctuations evolve in the mean according to the same macroscopic laws. herefore, in evaluating the correlation function α i (t α j (t + τ for a short time interval τ (a hydrodynamic time scale which is macroscopically short but microscopically long, α j (t + τ is given by j=1 α j (t + τ = α j (t + τ d dt α j (t = α j (t + τ n L jk X k (t. It is worth pointing out that τ is macroscopically short for the linear expansion but microscopically long for the applicability of the macroscopic laws. It follows that α i (t α j (t + τ is given by α i (t α j (t + τ = α i (t α j (t + τ Similarly, α j (t α i (t + τ is given by k=1 n L jk α i (t X k (t k=1 = α i (t α j (t τk B L ji α j (t α i (t + τ = α j (t α i (t τk B L ij. Comparing the above two time correlation functions, we obtain the reciprocal relations L ij = L ji from microscopic reversibility. Note that α i (t α j (t = α j (t α i (t by definition. C. Onsager s variational principle based on the reciprocal symmetry For small deviation from equilibrium, the system is in the linear response regime, where the state variables α i (i = 1,..., n evolve according to the kinetic equations α i = L ij X j, (1 11

or equivalently X i = R ij α j, (2 where the kinetic coefficients L ij form a symmetric and positive definite matrix, and so do the coefficients R ij, with L ij R jk = δ ik. Off-diagonal entries L ij and R ij are referred to as cross-coupling coefficients between different irreversible processes labeled by i and j. Under the condition that the variables α are even, i.e., their signs remain invariant under time reversal operation, Onsager derived the reciprocal relations L ij = L ji, (3 and consequently R ij = R ji, from the microscopic reversibility. It is worth emphasizing that his derivation does not require detailed knowledge of the irreversible processes. Based on Onsager s reciprocal relations (3, the kinetic equations (2 can be used to formulate a variational principle governing the irreversible processes. his variational principle states that for a closed system, the state evolution equations can be obtained by maximizing the action function (or functional Ṡ (α, α Φ S ( α, α (4 with respect to the rates { α i }. Here Ṡ = X i α i is the rate of change of the entropy, and the dissipation function Φ S ( α, α = R ij α i α j /2 is half the rate of entropy production. For an open system, however, there is an additional term Ṡ, which is the rate of entropy given by the system to the environment, added to the action (4, leading to O = Ṡ + Ṡ Φ S ( α, α, (5 which is called the Onsager-Machlup action. Note that Ṡ+Ṡ is still linear in { α i }. Onsager s variational principle states that for an open system, the state evolution equations can be obtained by maximizing the Onsager-Machlup action O with respect to the rates { α i }. his principle serves as a general framework for describing irreversible processes in the linear response regime. Onsager s variational principle is an extension from Rayleigh s principle of least energy dissipation, and naturally it reduces to the latter for isothermal systems. In an isothermal system, the rate of entropy given by the system to the environment can be expressed as Ṡ = Q/ = U/, where is the system temperature, Q is the rate of heat transfer 12

from the environment to the system, and U is the rate of change of the system energy, with Q = U according to the first law of thermodynamics. Note that is constant here. he maximization of the Onsager-Machlup action (5 is equivalent to the minimization of the so-called Rayleighian R = F (α, α + Φ F ( α, α, (6 with respect to the rates { α i }. Here F U (Ṡ Ṡ = + Ṡ is the rate of change of the Helmholtz free energy of the system, and the dissipation function Φ F ( α, α Φ S ( α, α is half the rate of free energy dissipation. As F is linear while Φ F is quadratic in the rates { α i }, the principle of least energy dissipation leads to F = 2Φ F. For isothermal systems, the Rayleighian can be written as R = F α i α i + 1 2 ζ ij α i α j, (7 where the first term in the right hand side is F and the second term is Φ F ( α, α, which is in a quadratic form with the friction coefficients ζ ij forming a symmetric and positive definite matrix. Minimization of R with respect to the rates gives the kinetic equations F α i = ζ ij α j, (8 which can be interpreted as a balance between the reversible force F/α i and the dissipative force linear in the rates. It is worth emphasizing that although the variational principle is equivalent to the kinetic equations combined with the reciprocal relations, the former possesses a notable advantage: the variational form allows flexibility in the choice of state variables. Once these variables are chosen, the conjugate forces are generated automatically via calculus of variations. III. APPLICAIONS A. Heat transport Consider the heat transport in a crystal. related by the constitutive equations he forces and rates ( velocities are 1 2 i = X i = 13 3 R ij J j, j=1

where X i (i = 1, 2, 3 are the forces and the components of the heat flux J j (j = 1, 2, 3 are the rates. Here the matrix R is the inverse of the Onsager coefficient matrix L, which is also symmetric. he dissipation function ϕ J, J is introduced in the form of ϕ J, J 1 2 3 R ij J i J j. It is worth emphasizing that ϕ can be defined in this quadratic form because of the symmetry in the matrices R and L. It is observed that substituting the constitutive equations into the quadratic expression for ϕ yields 2ϕ J, J 3 R ij J i J j = i,j=1 i,j=1 3 J i X i = i=1 3 ( 1 J i i which equals the rate of entropy production per unit volume due to heat transport. Let s denote the local entropy density in the system. hen, under the assumption of local equilibrium, the rate of change of s is given by ds dt = 1 ( J, where J is the rate of local accumulation of heat. he rate of change of the total entropy S is the volume integral Ṡ = i=1 ( ds dt dv = 1 J dv. he rate of the entropy given off to the surrounding environment is given by the surface integral Ṡ Jn = da, where J n is the outward normal component of the heat flux at the boundary. It follows that ( Ṡ + Ṡ = 1 J Jn dv + da ( = 1 J J 1 dv + dv = J dv. It can be shown that the constitutive equations for heat transport can be derived from the variational principle ] Φ J, J [Ṡ J + Ṡ (J n = minimum, 14,

where the temperature distribution is prescribed, and the rates, i.e., the heat flux J, are varied. Here Φ J, J is defined by Ṡ J is defined by and Ṡ (J n is defined by with Φ J, J ( ϕ J, J 1 dv 2 Ṡ J 3 R ij J i J j dv, i,j=1 ( 1 J dv, Ṡ Jn Jn da, Ṡ J + Ṡ Jn 1 J dv. ] he variation of Φ J, J [Ṡ J + Ṡ (J n is given by { ]} δ Φ J, J [Ṡ J + Ṡ (J n = from which we have according to the variational principle. equations ϕ J, J 1 = k X k J k k [ ] ϕ J, J 1 k δj k dv, J k We note that these are exactly the constitutive ( 1 R kj J j = k j As shown already, inserting the constitutive equations into the quadratic expression for ϕ yields the equality and hence the integral form 2ϕ J, J = k J k. ϕ J, J = J k k 2Φ J, J = Ṡ J + Ṡ (J n. J k k ( 1 Note that Ṡ J + Ṡ (J n is the rate of change of the entropy in the system and the surrounding environment. herefore, the rate of entropy production 2Φ J, J is equal to the rate of change of the total entropy Ṡ J + Ṡ (J n in an irreversible process governed by the constitutive equations. 15,

B. Stokes equation and Navier slip boundary condition he variational principle is now fully employed to investigate the Stokes flows with boundary slip. Below is a brief review, showing that the Stokes equation and the Navier slip boundary condition can be derived from the principle of least energy dissipation. Consider an incompressible Newtonian fluid in a region Ω with a solid boundary Ω, and neglect the inertial effect. he incompressibility condition reads v = 0, and the boundary is impermeable at which the normal velocity v n Ω = 0. Here the rate is the velocity field v ( r and the free energy is constant in time. he dissipation in the bulk region is due to the viscosity η, and the corresponding dissipation function is [ ] 1 Φ v = dv 4 η ( iv j + j v i 2. (9 Ω If boundary slip occurs at the fluid-solid interface, then the corresponding dissipation function is given by Φ s = Ω ds [ 1 2 β ( v slip ] 2, (10 where β is the slip coefficient, and v slip is the slip velocity, defined as the tangential velocity of the fluid relative to the solid at the fluid-solid interface. Here the solid boundary is still, and hence v slip = v. he Rayleighian of the system is given by R =Φ v + Φ s [ ] 1 = dv 4 η ( iv j + j v i 2 + Ω Ω ds [ ] 1 2 β v2. Combining the principle of least energy dissipation with the incompressibility condition, we have δ [ R Ω dv π iv i ] = 0 for any v ( r v ( r + δ v ( r, with π being the Lagrange multiplier. he Euler-Lagrange equations are the Stokes equation in the bulk region, and the Navier boundary condition (11 π + [η ( v + v ] = 0 (12 ˆn σ vis τ + β v slip = 0 (13 at the solid boundary, where σ vis η ( v + v is the viscous stress tensor, and τ I ˆnˆn with ˆn being the outward pointing (from fluid into solid unit vector normal to Ω. Note that the Lagrange multiplier π is the pressure. he total stress is σ πi + σ vis. 16

C. Lorentz reciprocal theorem Prior to Onsager s general work, there existed a few specific reciprocal relations studied by Lord Kelvin and Helmholtz. In fluid dynamics, the hydrodynamic reciprocal relations, known as the Lorentz reciprocal theorem, are also regarded as a special form of Onsager s reciprocal relations. Consider an incompressible Stokes flow in a region Ω with a solid boundary Ω. he velocity field v ( r is governed by the Stokes equation (12, with the no-slip boundary condition at Ω. Suppose that in the same system, two velocity fields v (1 and v (2 are both the solutions to Eq. (12, with their corresponding stress fields denoted by (1 σ and σ (2, respectively. he Lorentz reciprocal theorem states that dsˆn (1 σ v (2 = dsˆn (2 σ v (1, (14 Ω where ˆn is the outward pointing (from fluid into solid unit vector normal to Ω. he proof is as follows. he left hand side of Eq. (14 can be expressed as dsˆn (1 σ v (2 = dv (σ (1 v (2 Ω Ω = dv ( (1 σ v (2 + (1 σ : v (2 Ω [ = dv p (1 δ ij i v (2 j + η Ω 2 = dv η ( ( i v (1 j + j v (1 i i v (2 j + j v (2 i Ω 2 Ω ( ( ] i v (1 j + j v (1 i i v (2 j + j v (2 i where the Stokes equation, v = 0, and the symmetry of σ are used. It is readily seen that the right hand side of of Eq. (14 leads to the same expression. Furthermore, the Lorentz reciprocal theorem (14 can be expressed as, (15 F (1 k ẋ(2 k = F (2 k ẋ(1 k, (16 where ẋ k are the generalized velocities of the solid objects and F k are the generalized dissipative forces conjugate to ẋ k. Note that the no-slip boundary condition is applied to move from v of the fluid to ẋ k of the solid. Due to the linearity of the Stokes flows, we have the linear dependence of the forces on the rates: F k = ζ kl ẋ l, (17 17

where the friction coefficients ζ kl form a positive definite matrix. It follows from Eq. (16 that the Lorentz reciprocal theorem can be expressed as ζ kl = ζ lk, (18 meaning that the matrix formed by the friction coefficients ζ kl is symmetric. In the above discussion, the Lorentz reciprocal theorem expressed in Eqs. (16 and (18 is derived from the Stokes equation and the no-slip boundary condition. We have already shown that the Stokes equation (12 and the Navier boundary condition (13 can be simultaneously obtained from the principle of least energy dissipation. It is therefore expected that the hydrodynamic reciprocal relations can be generalized to describe the Stokes flows with the Navier boundary condition. Consider the same system with the Navier boundary condition. he velocity of the solid boundary Ω is denoted by W. From the Navier boundary condition (13, we readily obtain dsˆn (1 σ v slip(2 = dsˆn (2 σ v slip(1. (19 Ω Ω Meanwhile, Eq. (14 still holds. Note that W = v v slip on Ω. By combining Eqs. (14 and (19, we obtain the generalized form of the hydrodynamic reciprocal relations dsˆn (1 σ W (2 = dsˆn (2 σ W (1. (20 Ω Ω Note that the no-slip limit is obtained as β with W = v on Ω. With the Lorentz reciprocal theorem generalized from Eq. (14 to (20, it can be further expressed as Eq. (16, which results in the symmetry in Eq. (18. We emphasize that in the presence of boundary slip, we need Eq. (20 in order to arrive at Eqs. (16 and (18. It is remarkable that the reciprocal symmetry is preserved in the Stokes flows with the Navier slip condition. o use Eq. (18 for the present study, we consider the solid boundary Ω consisting of the surfaces of N rigid bodies Ω i (i = 1,..., N, each in a motion described by the translational velocity V i and the angular velocity ω i. he solid velocity at r on Ω i can be expressed as W ( r = V i + ω i δ r i (21 where δ r i is measured relative to the center of mass of the i-th rigid body. hen we have dsˆn (1 (2 σ W = Ω N ( dsˆn (1 σ V i(2 + Ω i i=1 18 N [ dsδ r i (ˆn (1 ] σ ω i(2, Ω i i=1

where Ω i dsˆn σ is the total force by the i-th rigid body on the fluid, and Ω i dsδ r i (ˆn σ is the total torque by the i-th rigid body on the fluid. his leads to a specific form of Eq. (16, in which V i and ω i (i = 1,..., N are the generalized velocities of the rigid bodies ẋ k, and Ω i dsˆn σ and Ω i dsδ r i (ˆn σ are their conjugate generalized forces F k. Finally we emphasize that the Lorentz reciprocal theorem is valid only when the slip length l s = η/β is a material constant, which makes the Navier boundary condition linear. D. Nematic liquid crystals he Leslie-Ericksen hydrodynamic theory for nematic liquid crystals gives the dissipative stress tensor σ and the torque density by the director on the fluid Γ as ( σ =α 1 n n : A n n + α 2 n N + α 3N n + α4a + α5 n ( n A + α 6 ( n A n, (22 ( Γ = n γ 1N + γ2a n, (23 where α i (i = 1,..., 6 are phenomenological parameters, n is the director, A ( 1 2 v + v is the rate-of-strain tensor of flow, and N n ν n = ( ω ν n is the velocity of the director relative to the fluid, with ω being the angular velocity of the director and ν 1 2 v being the angular velocity of the fluid. In addition, γ 1 and γ 2 are given by γ 1 = α 3 α 2 and γ 2 = α 6 α 5. Only five of the six α i (i = 1,..., 6 parameters are independent because of the Parodi relation which results from the Onsager reciprocal relations. α 2 + α 3 = α 6 α 5, (24 E. Cross coupling in gas flows in micro-channels In rarefied gases, mass and heat transport processes interfere with each other, leading to the mechano-caloric effect and thermo-osmotic effect, which are of interest to both theoretical study and practical applications. We employ the unified gas-kinetic scheme to investigate these cross coupling effects in gas flows in micro-channels. Our numerical simulations cover channels of planar surfaces and also channels of ratchet surfaces, with Onsager s reciprocal relation verified for both cases. For channels of planar surfaces, simulations are performed in a wide range of Knudsen number and our numerical results show good agreement with 19

the literature results. For channels of ratchet surfaces, simulations are performed for both the slip and transition regimes and our numerical results not only confirm the theoretical prediction [Phys. Rev. Lett. 107, 164502 (2011] for Knudsen number in the slip regime but also show that the off-diagonal kinetic coefficients for cross coupling effects are maximized at a Knudsen number in the transition regime. Finally, a preliminary optimization study is carried out for the geometry of Knudsen pump based on channels of ratchet surfaces. In a closed system out of equilibrium, the rate of entropy production can be expressed as N ds dt = J i X i, (25 i=1 where S is the entropy, J i are the thermodynamic fluxes, and X i are the conjugate thermodynamic forces. For small deviation away from equilibrium, we have the linear relations between J i and X i : J i = N L ij X j, (26 j=1 where L ij are the kinetic coefficients. Onsager s reciprocal relations state that L ij and L ji are equal as a result of microscopic reversibility. Starting from the Gibbs equation, the thermodynamic fluxes and forces can be identified for gas flows, and the corresponding constitutive equations can be derived. A schematic illustration of the cross coupling in a channel of planar surfaces can be found in figure 2, where a long channel is confined by two parallel solid plates separated by a distance H and connected with two reservoirs. he left reservoir is maintained at pressure p 0 p/2 and temperature 0 /2 while the right reservoir is maintained at p 0 + p/2 and 0 + /2. We use p < 0 and > 0 in our simulations, with p/p 0 1 and / 0 1 to ensure the linear response. Usually, a mass flux to the right is generated by the pressure gradient due to p < 0 and a heat flux to the left is generated by the temperature gradient due to > 0. For rarefied gas, however, p also contributes to the heat flux and also contributes to the mass flux. hese cross coupling effects are called the mechano-caloric effect and thermo-osmotic effect respectively. For a single-component gas, the rate of entropy production can be expressed as ds ( dt = J M ν 1 + J E, (27 where ν is the chemical potential per unit mass, J E and J M are the energy flux and mass flux from the left reservoir to the right reservoir, and means the quantity on the right 20

minus the quantity on the left. Here ν and J E can be written as ν = h s, (28 J E = J Q + hj M, (29 where s and h are the entropy and enthalpy per unit mass, and J Q is the heat flux. ogether with the Gibbs-Duhem equation where ρ is the mass density. Equation (27 becomes dν = sd + dp/ρ, (30 ds dt = 1 ρ J M p 1 J Q. (31 2 According to equation (31, the thermodynamic forces and fluxes are connected in the form of with J M = L MM J Q L QM L MQ L QQ ρ 1 0 0 1 p, (32 0 2 L MQ = L QM, (33 due to Onsager s reciprocal relations. he detailed mechanism may vary with geometric configuration and rarefaction. Here and throughout the paper, the subscript 0 denotes the reference state from which various deviations (in pressure, temperature, etc are measured. In the free molecular regime and with specular reflection on plates, the gas molecules travel ballistically from on side to the other and the distribution function at any point can be treated as a combination of two half-space Maxwellians from the two reservoirs. he kinetic coefficients in equation (32 can be analytically derived in this case, given by L MM L QM L MQ L QQ = Hρ 0 0 4 8kB 0 πm ρ 0/p 0 1/2 1/2 9p 0 /4ρ 0 where k B is the Boltzmann constant and m is the molecular mass. If the temperature gradient is imposed on the plates and the gas molecules are diffusely reflected, then the mass flux due to the temperature gradient is generated by thermal creep 21. (34

on the plates. he kinetic coefficients in this case have been calculated by several authors using different methods. Assuming the length to height ratio of the channel is fixed and noting ρλ = constant and µ ρ 0 1/2 for hard-sphere molecules, the average velocity U induced by thermal creep can be estimated from the Maxwell slip boundary condition, U µ 0 ρ 0 0 Kn 0, (35 where λ is the mean free path, µ is the dynamic viscosity independent of the density, and Kn = λ 0 /H is the Knudsen number. In later sections, we will show that L MQ and L QM are equal and increase with the increasing Kn. F. Cross coupling in a mixture of fluids 1. Ideal fluids he continuity equation is given by t ρ + (ρ v = 0 where ρ is the mass density. he momentum equation is given by t (ρ v + (ρ v v = ρ d v = p dt where d v v + v v is the rate of change of the velocity of a given fluid particle, and dt t p is the pressure. his is Euler s equation and is one of the fundamental equations of fluid dynamics. he motion of an ideal fluid is adiabatic and in adiabatic motion the entropy of any fluid particle remains constant as the particle moves about in space. herefore, the entropy equation is given by where s is the entropy per unit mass. (ρs + (ρs v = 0 t Now we are ready to derive the energy equation. he kinetic energy density is 1 2 ρ v2 and the internal energy density is ρε where ε is the internal energy per unit mass. Using the continuity equation and the momentum equation, we have 1 t 2 ρ v2 = v p 22 [( 1 2 ρ v2 ] v

for the kinetic energy. As to the internal energy, we have d (ρε = ρ ds + hdρ where h = ε + p/ρ is the enthalpy per unit mass. It follows that for an ideal fluid we have t (ρε = ρ t s + h ρ = ρ v s h (ρ v t by use of the entropy equation and the continuity equation. Putting the two energies together, we have t ( 1 2 ρ v2 + ρε = v p [ ] 1 2 ρ v2 v ρ v s h (ρ v. o proceed, we note that from thermodynamics we have dh = ds + (1/ρdp, ρdh = ρ ds + ( dp, and hence ρ h = ρ s + p. It follows that 1 t 2 ρ v2 + ρε can be expressed as [ ] 1 1 t 2 ρ v2 + ρε = 2 ρ v2 + ρh v, which shows that the energy flux is given by ( 1 2 ρ v2 + ρh v. Note that we have ρh v rather than ρε v in the energy flux because the work done by pressure forces is to be included. 2. Viscous fluids Now we turn to viscous fluids. momentum equation is given by he continuity equation remains unchanged and the t (ρ v + (ρ v v = ρ d σ v = p + dt where σ is the viscous stress tensor. he energy equation is given by [ ] 1 1 ( σ t 2 ρ v2 + ρε = 2 ρ v2 + ρh v + v q where q is the heat flux. Note that the energy equation for viscous fluids takes into account the work done by viscous forces and the thermal conduction, which are absent in ideal fluids. he entropy equation can be derived by use of the above equations and thermodynamic relations. he derivation is straightforward and the procedure can be outlined as follows. Using the continuity equation and the momentum equation, we can obtain an equation for ( 1 ρ v2 (. Combining the energy equation and the equation for 1 ρ v2, we can obtain an t 2 t 2 equation for t (ρε: t (ρε = (ρh v + σ : v q + v p. 23

Combining the above equation with (ρε = ρ s + h ρ, we obtain t t t ρ t s = σ : v q + v p ρ v h, which becomes ρ s + v s = σ : v q, t with the help of ρ h + p = ρ s. Finally we obtain the entropy equation q (ρs + (ρs v = + 1 1 σ : v + q t, where q is the entropy flux, 1 σ dissipation, and q 1 : v is the rate of entropy production due to viscous is the rate of entropy production due to thermal conduction. 3. A mixture of fluids Let us start from the continuity equations of two miscible components, with one labeled by the subscript 1 and the other labeled by the subscript 2. hey read and t ρ 1 + (ρ 1 v 1 = 0 t ρ 2 + (ρ 2 v 2 = 0, in which v i (i = 1, 2 is the velocity of a particular species. he mass density ρ and velocity v of the mixture are defined by ρ = ρ 1 + ρ 2 and ρ v = ρ 1 v 1 + ρ 2 v 2. Physically, v is the massaveraged velocity which is a field variable that enters into the hydrodynamic momentum equation. he local relative concentration c is defined by ρc = ρ 1 ρ 2. It follows that ρc v equals (ρ 1 ρ 2 v. Adding the continuity equations for the two components gives the continuity equation t ρ + (ρ v = 0 for ρ. he diffusive flux j is defined through the equation ρ 1 v 1 ρ 2 v 2 = ρc v + j. It follows that j is given by j = ρ 1 ( v 1 v ρ 2 ( v 2 v. Note that ρ 1 ( v 1 v + ρ 2 ( v 2 v = 0 the diffusion discussed here is defined relative to the motion of the center of mass of a fluid 24

element. Rewriting the difference between the continuity equations for the two components as we obtain t (ρc + ρc v + j = 0, ρ c + v c = ρ d t dt c = j. By introducing µ as an appropriately defined chemical potential of the mixture, we have thermodynamic equations d (ρε = ρ ds + hdρ + ρµdc and ρdh = ρ ds + dp + ρµdc for ε and h. hey will be used when we derive the entropy equation from the energy equation. o proceed, we employ the momentum equation and the energy equation t ( 1 2 ρ v2 + ρε t (ρ v + (ρ v v = ρ d σ v = p + dt = [ ] 1 ( σ 2 ρ v2 + ρh v + v q, which look the same as those used for viscous fluids of one component. We would like to point out that while σ is still the viscous stress tensor, the physical meaning of q is not clear at the moment it is expected to represent the total energy flux due to thermal conduction and diffusion. his is to be clarified by the explicit form of the entropy equation. In order to derive the entropy equation, we first combine the continuity equation and ( the momentum equation to obtain an equation for 1 ρ v2. We then combine the energy t 2 ( equation and the equation for 1 ρ v2 to obtain an equation for (ρε: t 2 t t (ρε = (ρh v + σ : v q + v p. Combining the above equation with (ρε = ρ s + h ρ + ρµ c from thermodynamics, t t t t we obtain ρ t s = σ : v q + v p ρ v h ρµ t c, which becomes ρ s + v s = σ : v q ρµ c + v c, t t 25

with the help of ρ h + p = ρ s ρµ c. Using ρ ( c + v c = j, we have t ρ s + v s = σ : v q + µ j = σ : v q µ j j µ, t which gives the entropy equation in which q µ j (ρs + (ρs v = t ( q µ j + π, is the entropy flux and π is the rate of entropy production per unit volume, given by π = 1 σ : v + q µ j 1 j µ. Here 1 σ : v is the rate of entropy production due to viscous dissipation, and q µ j 1 j µ is the rate of entropy production due to thermal conduction and diffusion. he latter can be written as ( q µ j ( + j µ 2 in which q µ j and j are the fluxes associated with thermal conduction and diffusion, and and µ 2 equations are the conjugate forces. Now we are ready to write down the constitutive µ j = α β 2 2 µ q µ j = δ γ 2 2 with the reciprocal relation β 2 = δ for the cross coupling between thermal conduction and diffusion. FIGURES 26

1 2 FIG. 1. A closed system made up of two interconnected subsystems. p 0 p/2 0 /2 H mass flux J M heat flux J Q p 0 + p/2 0 + /2 ( p < 0 ( > 0 L FIG. 2. A schematic illustration of the cross coupling in a channel of planar surfaces. 27