Statistical mechanics of the fluctuating lattice Boltzmann equation

Statistical mechanics of the fluctuating lattice Boltzmann equation Statistical mechanics of the fluctuating lattice Boltzmann equation

12.07.2015 Views

PHYSICAL REVIEW E 76, 036704 2007Statistical mechanics of the fluctuating lattice Boltzmann equationBurkhard Dünweg and Ulf D. SchillerMax Planck Institute for Polymer Research, Ackermannweg 10, D-55128 Mainz, GermanyAnthony J. C. LaddMax Planck Institute for Polymer Research, Ackermannweg 10, D-55128 Mainz, Germany and Chemical Engineering Department,University of Florida, Gainesville, Florida 32611-6005, USAReceived 11 July 2007; published 12 September 2007We propose a derivation of the fluctuating lattice Boltzmann equation that is consistent with both equilibriumstatistical mechanics and fluctuating hydrodynamics. The formalism is based on a generalized lattice-gasmodel, with each velocity direction occupied by many particles. We show that the most probable state of thismodel corresponds to the usual equilibrium distribution of the lattice Boltzmann equation. Thermal fluctuationsabout this equilibrium are controlled by the mean number of particles at a lattice site. Stochastic collision rulesare described by a Monte Carlo process satisfying detailed balance. This allows for a straightforward derivationof discrete Langevin equations for the fluctuating modes. It is shown that all nonconserved modes should bethermalized, as first pointed out by Adhikari et al. Europhys. Lett. 71, 473 2005; any other choice violatesthe condition of detailed balance. A Chapman-Enskog analysis is used to derive the equations of fluctuatinghydrodynamics on large length and time scales; the level of fluctuations is shown to be thermodynamicallyconsistent with the equation of state of an isothermal, ideal gas. We believe this formalism will be useful indeveloping new algorithms for thermal and multiphase flows.DOI: 10.1103/PhysRevE.76.036704PACS numbers: 47.11.Qr, 47.57.sI. INTRODUCTIONThe collisions conserve mass and momentum, henceur,t = j r,t/r,t. 4 stress tensorLattice Boltzmann LB methods 1,2 have become apopular tool for simulating hydrodynamics, particularly in i = i c i =0.i icomplex geometries. The underlying model is a regular lattice5of sites r, combined with a small set of velocity vectorsc i , which, within one time step h, connect a given site withThe algorithm thus satisfies important requirements for simulatinghydrodynamic flows—mass and momentum conserva-some of its neighbors. The set of velocities is chosen to becompatible with the symmetry of the lattice. The basic dynamicalvariables are real-valued populations n i ; in thetion, and locality—but lacks Galilean invariance due to thefinite number of velocities. Full rotational symmetry is alsopresent paper, we will consider n i as the mass density associatedwith the velocity c i . The LB algorithm is then de-lost, but by a suitable choice of velocity set, isotropic momentumtransport can be recovered on sufficiently large hydrodynamiclength scales. Nevertheless, the finite number ofscribed by the update rulevelocities always confines the method to flows with smalln i r + c i h,t + h = n i r,t = n i r,t + i n i r,t, 1Mach number u/c s ≪1. The speed of sound c s is of orderb/h, where b is the lattice spacing, or of order c i .Most of the LB literature deals with deterministic collisionrules, with i describing a linear relaxation of the dis-where n i denotes the complete set of populations. Then i r,t at each site are first rearranged in a “collision” step, tribution n i toward the local equilibrium 3,4:described by i , and then propagated along their respectivelinks. The hydrodynamic fields, mass densityn eq i ,u = a c i1+ u · c i2+ u · c i 24− u2cr,t = n i r,t,2s 2c s 2c2, 6siwhere a c i0 is the weight associated with the speed c i . Theand momentum densityviscosity of the LB fluid is determined by the choice of relaxationrates.However, to simulate Brownian motion of suspended particles,thermal fluctuations must be included. At the hydro-j r,t = n i r,tc i3idynamic level, this means adding uncorrelated noise to thefluid stress tensor 5. In Refs. 6–8 an analogous fluctuatingLB model was introduced by making i a stochasticare moments of the discrete velocity distribution n i r,t,while the fluid velocity is given byvariable, but in such a way that the noise was only applied tothe modes linear combinations of n i related to the viscous1539-3755/2007/763/03670410036704-1©2007 The American Physical Society

PHYSICAL REVIEW E 76, 036704 2007<strong>Statistical</strong> <strong>mechanics</strong> <strong>of</strong> <strong>the</strong> <strong>fluctuating</strong> <strong>lattice</strong> <strong>Boltzmann</strong> <strong>equation</strong>Burkhard Dünweg and Ulf D. SchillerMax Planck Institute for Polymer Research, Ackermannweg 10, D-55128 Mainz, GermanyAnthony J. C. LaddMax Planck Institute for Polymer Research, Ackermannweg 10, D-55128 Mainz, Germany and Chemical Engineering Department,University <strong>of</strong> Florida, Gainesville, Florida 32611-6005, USAReceived 11 July 2007; published 12 September 2007We propose a derivation <strong>of</strong> <strong>the</strong> <strong>fluctuating</strong> <strong>lattice</strong> <strong>Boltzmann</strong> <strong>equation</strong> that is consistent with both equilibriumstatistical <strong>mechanics</strong> and <strong>fluctuating</strong> hydrodynamics. The formalism is based on a generalized <strong>lattice</strong>-gasmodel, with each velocity direction occupied by many particles. We show that <strong>the</strong> most probable state <strong>of</strong> thismodel corresponds to <strong>the</strong> usual equilibrium distribution <strong>of</strong> <strong>the</strong> <strong>lattice</strong> <strong>Boltzmann</strong> <strong>equation</strong>. Thermal fluctuationsabout this equilibrium are controlled by <strong>the</strong> mean number <strong>of</strong> particles at a <strong>lattice</strong> site. Stochastic collision rulesare described by a Monte Carlo process satisfying detailed balance. This allows for a straightforward derivation<strong>of</strong> discrete Langevin <strong>equation</strong>s for <strong>the</strong> <strong>fluctuating</strong> modes. It is shown that all nonconserved modes should be<strong>the</strong>rmalized, as first pointed out by Adhikari et al. Europhys. Lett. 71, 473 2005; any o<strong>the</strong>r choice violates<strong>the</strong> condition <strong>of</strong> detailed balance. A Chapman-Enskog analysis is used to derive <strong>the</strong> <strong>equation</strong>s <strong>of</strong> <strong>fluctuating</strong>hydrodynamics on large length and time scales; <strong>the</strong> level <strong>of</strong> fluctuations is shown to be <strong>the</strong>rmodynamicallyconsistent with <strong>the</strong> <strong>equation</strong> <strong>of</strong> state <strong>of</strong> an iso<strong>the</strong>rmal, ideal gas. We believe this formalism will be useful indeveloping new algorithms for <strong>the</strong>rmal and multiphase flows.DOI: 10.1103/PhysRevE.76.036704PACS numbers: 47.11.Qr, 47.57.sI. INTRODUCTIONThe collisions conserve mass and momentum, henceur,t = j r,t/r,t. 4 stress tensorLattice <strong>Boltzmann</strong> LB methods 1,2 have become apopular tool for simulating hydrodynamics, particularly in i = i c i =0.i icomplex geometries. The underlying model is a regular <strong>lattice</strong>5<strong>of</strong> sites r, combined with a small set <strong>of</strong> velocity vectorsc i , which, within one time step h, connect a given site withThe algorithm thus satisfies important requirements for simulatinghydrodynamic flows—mass and momentum conserva-some <strong>of</strong> its neighbors. The set <strong>of</strong> velocities is chosen to becompatible with <strong>the</strong> symmetry <strong>of</strong> <strong>the</strong> <strong>lattice</strong>. The basic dynamicalvariables are real-valued populations n i ; in <strong>the</strong>tion, and locality—but lacks Galilean invariance due to <strong>the</strong>finite number <strong>of</strong> velocities. Full rotational symmetry is alsopresent paper, we will consider n i as <strong>the</strong> mass density associatedwith <strong>the</strong> velocity c i . The LB algorithm is <strong>the</strong>n de-lost, but by a suitable choice <strong>of</strong> velocity set, isotropic momentumtransport can be recovered on sufficiently large hydrodynamiclength scales. Never<strong>the</strong>less, <strong>the</strong> finite number <strong>of</strong>scribed by <strong>the</strong> update rulevelocities always confines <strong>the</strong> method to flows with smalln i r + c i h,t + h = n i r,t = n i r,t + i n i r,t, 1Mach number u/c s ≪1. The speed <strong>of</strong> sound c s is <strong>of</strong> orderb/h, where b is <strong>the</strong> <strong>lattice</strong> spacing, or <strong>of</strong> order c i .Most <strong>of</strong> <strong>the</strong> LB literature deals with deterministic collisionrules, with i describing a linear relaxation <strong>of</strong> <strong>the</strong> dis-where n i denotes <strong>the</strong> complete set <strong>of</strong> populations. Then i r,t at each site are first rearranged in a “collision” step, tribution n i toward <strong>the</strong> local equilibrium 3,4:described by i , and <strong>the</strong>n propagated along <strong>the</strong>ir respectivelinks. The hydrodynamic fields, mass densityn eq i ,u = a c i1+ u · c i2+ u · c i 24− u2cr,t = n i r,t,2s 2c s 2c2, 6siwhere a c i0 is <strong>the</strong> weight associated with <strong>the</strong> speed c i . Theand momentum densityviscosity <strong>of</strong> <strong>the</strong> LB fluid is determined by <strong>the</strong> choice <strong>of</strong> relaxationrates.However, to simulate Brownian motion <strong>of</strong> suspended particles,<strong>the</strong>rmal fluctuations must be included. At <strong>the</strong> hydro-j r,t = n i r,tc i3idynamic level, this means adding uncorrelated noise to <strong>the</strong>fluid stress tensor 5. In Refs. 6–8 an analogous <strong>fluctuating</strong>LB model was introduced by making i a stochasticare moments <strong>of</strong> <strong>the</strong> discrete velocity distribution n i r,t,while <strong>the</strong> fluid velocity is given byvariable, but in such a way that <strong>the</strong> noise was only applied to<strong>the</strong> modes linear combinations <strong>of</strong> n i related to <strong>the</strong> viscous1539-3755/2007/763/03670410036704-1©2007 The American Physical Society


DÜNWEG, SCHILLER, AND LADDneq = n neq i c i c i ; ihere and denote Cartesian components and n i neq =n i−n i eq is <strong>the</strong> nonequilibrium distribution. Although this procedureis correct in <strong>the</strong> hydrodynamic limit 7,9, it providespoor <strong>the</strong>rmalization on smaller length scales, as was firstobserved by Adhikari et al. 10. They introduced a <strong>the</strong>rmalizationprocedure which applies to all nonconserved modes,with significantly improved numerical behavior at shortscales 10. The procedure was derived by considering a<strong>fluctuating</strong> LB model, making explicit use <strong>of</strong> <strong>the</strong> transformationbetween <strong>the</strong> populations n i and <strong>the</strong> modes 11.The purpose <strong>of</strong> <strong>the</strong> present paper is to rederive <strong>the</strong> stochasticupdating rule <strong>of</strong> Ref. 10 from a generalized <strong>lattice</strong>gasmodel. The difference in our formulation lies in <strong>the</strong> introduction<strong>of</strong> an ensemble <strong>of</strong> population densities at eachgrid point, so that a <strong>fluctuating</strong> LB simulation is a singlerealization <strong>of</strong> this ensemble. There follows naturally a probabilitydistribution, Pn i , for <strong>the</strong> set <strong>of</strong> populations n i ata position r and time t. The equilibrium distribution at asingle site can be derived by maximizing P subject to <strong>the</strong>constraints <strong>of</strong> fixed mass and momentum densities, and j .This distribution agrees with <strong>the</strong> standard equilibrium distributionfor LB models Eq. 6 up to terms <strong>of</strong> order u 2 .Asimilar procedure has been followed in deriving H-<strong>the</strong>oremsfor LB models 12–14, but <strong>the</strong>se papers were not concernedwith fluctuations.A coarse-graining <strong>of</strong> <strong>the</strong> microscopic collision operatorleads to a Langevin description for <strong>the</strong> nonconserved degrees<strong>of</strong> freedom. However, <strong>the</strong>se stochastic collisions may also beviewed as a Monte Carlo process 15, satisfying <strong>the</strong> principle<strong>of</strong> detailed balance governed by Pn i . The procedure<strong>of</strong> Refs. 7,9 can be shown to violate detailed balance, while<strong>the</strong> improved version <strong>of</strong> Ref. 10 satisfies it.In summary, our goal is to reconnect <strong>the</strong> <strong>lattice</strong> <strong>Boltzmann</strong><strong>equation</strong> with its <strong>lattice</strong> gas origins, and thus to establisha firm statistical mechanical foundation for stochasticLB simulations, in addition to <strong>the</strong> usual connection to <strong>fluctuating</strong>hydrodynamics 7,10. We believe this provides acomparable <strong>the</strong>oretical framework to that already availablefor o<strong>the</strong>r stochastic simulation methods, such as dissipativeparticle dynamics 16 and stochastic rotation dynamics 17.This formulation also <strong>of</strong>fers <strong>the</strong> possibility for future modificationsand generalizations, for example, to <strong>the</strong>rmal flows18, models with nonideal <strong>equation</strong>s <strong>of</strong> state 19,20, ormulticomponent mixtures 21.The paper is organized as follows. In Sec. II we describe<strong>the</strong> underlying <strong>lattice</strong>-gas model, derive <strong>the</strong> probability distributionPn i , and show that <strong>the</strong> most probable value forn i is equivalent to Eq. 6. In Sec. III we consider smallfluctuations around <strong>the</strong> equilibrium distribution. We show<strong>the</strong>y are approximately Gaussian distributed, with <strong>the</strong> level<strong>of</strong> <strong>the</strong>rmal fluctuations governed by <strong>the</strong> degree <strong>of</strong> coarsegraining: a given amount <strong>of</strong> mass on a <strong>lattice</strong> site can bedistributed between many particles, in which case <strong>the</strong> fluctuationsare small, or between few, in which case <strong>the</strong>y arelarge. In this way we can adjust <strong>the</strong> level <strong>of</strong> fluctuationswhile keeping <strong>the</strong> temperature fixed. In Sec. IV we construct7PHYSICAL REVIEW E 76, 036704 2007a stochastic collision operator such that detailed balance issatisfied. From this, we derive <strong>the</strong> random stresses at an individualsite. In Sec. V we apply <strong>the</strong> Chapman-Enskog procedure9 to <strong>the</strong> algorithm in order to find <strong>the</strong> behavior on<strong>the</strong> hydrodynamic scale; <strong>the</strong> deterministic and stochasticterms are here treated on an equal basis 22. We <strong>the</strong>n findthat, on <strong>the</strong> macroscopic scale, <strong>the</strong> procedure yields exactly<strong>the</strong> stress correlations given by Landau and Lifshitz 5. SectionVI discusses how to choose parameters for a coupledparticle-fluid system. Section VII summarizes our conclusions.II. SINGLE-SITE PROBABILITY DISTRIBUTIONHistorically, <strong>the</strong> <strong>lattice</strong>-<strong>Boltzmann</strong> model 3,23 was developedfrom earlier work on <strong>lattice</strong>-gas LG models24,25, in which each velocity direction was occupied by atmost one particle. Here, we imagine a generalized <strong>lattice</strong>-gasmodel GLG where each velocity direction can be occupiedby many particles. Each particle has <strong>the</strong> same mass, but differentvelocity directions may have different mean populations,even in a fluid at rest. The microscopic state <strong>of</strong> <strong>the</strong>system at any given site is specified by a set <strong>of</strong> integers i giving <strong>the</strong> occupancies <strong>of</strong> each direction. Then <strong>the</strong> update <strong>of</strong><strong>the</strong> GLG is analogous to <strong>the</strong> standard LG or LB models, butwith i an integer as opposed to a Boolean or real variable: i r + c i h,t + h = i r,t = i r,t + ˜ i i r,t,where ˜ i operates on i to compute <strong>the</strong> change in population i − i . While collisions may be both deterministic andmicroscopically reversible, we shall assume only that <strong>the</strong>collision operator satisfies detailed balance.Without considering <strong>the</strong> collision rules in detail, we constructan equilibrium distribution from <strong>the</strong> following thoughtexperiment. Consider a velocity direction, i, at a particularsite, r. Particles are drawn randomly from a large reservoirand assigned to <strong>the</strong> site r with velocity c i ; <strong>the</strong> number <strong>of</strong>particles in <strong>the</strong> reservoir is assumed to be much larger than<strong>the</strong> number <strong>of</strong> particles selected, i r. Under <strong>the</strong>se circumstances i r follows a Poisson distribution,P i = ¯i i i ! e−¯i,9with a mean number <strong>of</strong> particles ¯i, and a variance 2 i − i 2 = ¯i.10Let m p be <strong>the</strong> mass <strong>of</strong> a particle and =m p /b d , where b is<strong>the</strong> <strong>lattice</strong> spacing and d is <strong>the</strong> spatial dimension. Then n i= i , andn 2 i − n i 2 = n i .11The fluctuations in mass density at a site are controlled by<strong>the</strong> mass <strong>of</strong> an LB particle: small m p means that <strong>the</strong> mass isdistributed onto many particles, and <strong>the</strong>refore fluctuationsare small. For fixed m p , and <strong>the</strong>refore <strong>the</strong> level <strong>of</strong> fluctuationsbecomes large as b decreases. This is natural, sincea fine spatial resolution means fewer particles per cell, andlarger fluctuations relative to <strong>the</strong> mean.8036704-2


STATISTICAL MECHANICS OF THE FLUCTUATING …If we now imagine sampling each velocity with an independentreservoir, but taking only those sets <strong>of</strong> populationswhich produce specific values for <strong>the</strong> total mass and momentum,<strong>the</strong> probability density for <strong>the</strong> occupation numbers isexcept for normalizationP i i i ¯i i ! e−¯ii i − i c i − j .i12Using Stirling’s approximation for i ≫1, we can write <strong>the</strong>distribution in terms <strong>of</strong> <strong>the</strong> entropy associated with <strong>the</strong> occupationnumbers,S i =− i ln i − i − i ln ¯i + ¯i, 13iand <strong>the</strong> constraints:P i expS i i i − i c i − j .i14The equilibrium distribution, eq i , can be found by maximizingS, treating i as a continuous variable, and taking intoaccount <strong>the</strong> mass and momentum constraints via Lagrangemultipliers, and , j respectively:S i+ + j · c i =0, i − =0,i1516 i c i − j =0.17iIt should be noted that this procedure is closely related to <strong>the</strong>determination <strong>of</strong> an entropy function for <strong>the</strong> LB <strong>equation</strong>13. Equation 15 can be solved to give <strong>the</strong> equilibriumpopulations in terms <strong>of</strong> <strong>the</strong> Lagrange multipliers, i eq = ¯i exp + j · c i,18which are <strong>the</strong>n determined from <strong>the</strong> constraints, Eqs. 16and 17, substituting i eq for i .The mean populations in <strong>the</strong> absence <strong>of</strong> constraints, ¯i,can be expressed in terms <strong>of</strong> <strong>the</strong> mean number <strong>of</strong> particles ata site,¯i = ¯a c i,19where ¯ = i ¯i. The symmetry <strong>of</strong> <strong>the</strong> <strong>lattice</strong> constrains <strong>the</strong>weights, a c i, to be dependent on <strong>the</strong> speed <strong>of</strong> <strong>the</strong> particle, butnot its direction. Thus for a <strong>lattice</strong> with cubic symmetry, a c i =1,20i a c ic i =0,i21PHYSICAL REVIEW E 76, 036704 2007 a c ic i c i = 2 ,i22 a c ic i c i c i =0,23iwhere is <strong>the</strong> Kronecker delta, and 2 is a constant withunits b/h 2 .A solution <strong>of</strong> <strong>the</strong> nonlinear <strong>equation</strong>s 16–18 requiresan iterative numerical procedure, but it is more practical toseek an approximate expression for <strong>the</strong> equilibrium distributionin <strong>the</strong> limit that j ·c i is small 14. To second order in , j <strong>the</strong> mass and momentum constraints yield: 2 2¯e j 1+ = , 242¯e 2 j = u .25Inserting <strong>the</strong>se results into Eq. 18, we find <strong>the</strong> equilibriumdistribution can be written in <strong>the</strong> form <strong>of</strong> Eq. 6,n eq i = a c i1+ u · c i+ u · c i 22− u226 2 2 22 2.For <strong>the</strong> sake <strong>of</strong> completeness, we now briefly mention <strong>the</strong>procedure 4,9 to determine <strong>the</strong> weights a c i such that <strong>the</strong> LBmodel is consistent with hydrodynamics. This requires that<strong>the</strong> second moment <strong>of</strong> <strong>the</strong> equilibrium distribution, eq = n eq i c i c i ,27ishould equal <strong>the</strong> Euler stress p +u u , with <strong>the</strong> pressuregiven by <strong>the</strong> ideal gas <strong>equation</strong> <strong>of</strong> state, p=k B T/m p , wherek B is <strong>Boltzmann</strong>’s constant and T is <strong>the</strong> absolute temperature.For an iso<strong>the</strong>rmal gas <strong>of</strong> particles <strong>of</strong> mass m p , k B T=m p c 2 s ,and <strong>the</strong>refore <strong>the</strong> <strong>equation</strong> <strong>of</strong> state is also given by p=c 2 s ,with c s <strong>the</strong> speed <strong>of</strong> sound.To evaluate eq we require <strong>the</strong> fourth moment <strong>of</strong> a c i,which from cubic symmetry must be <strong>of</strong> <strong>the</strong> form a c ic i c i c i c i = 4 + 4 + + ,i28where is unity if all four indexes are <strong>the</strong> same and zeroo<strong>the</strong>rwise; 4 and 4 have units <strong>of</strong> b/h 4 . Consistency betweenEq. 27 and <strong>the</strong> Euler stress requires that 2 = c s 2 = k B T/m p , 4 = 2 2 ,2930 4 =0.31These conditions, toge<strong>the</strong>r with <strong>the</strong> normalization condition, i a c i=1, determine <strong>the</strong> weights uniquely for a model withthree different speeds. For example, for <strong>the</strong> D3Q19 model4 19 velocities on a three-dimensional simple cubic <strong>lattice</strong>,a 0 =1/3 for <strong>the</strong> stationary particles, a 1 =1/18 for <strong>the</strong>036704-3


DÜNWEG, SCHILLER, AND LADDsix nearest-neighbor directions, and a 2 =1/36 for <strong>the</strong> 12next-nearest-neighbor directions: <strong>the</strong> sound speed is <strong>the</strong>n c s2=1/3b/h 2 . In <strong>the</strong> D2Q9 model 4 nine velocities on atwo-dimensional square <strong>lattice</strong> <strong>the</strong> weights are a 0 =4/9, a 1=1/9, and a 2 =1/36; <strong>the</strong> sound speed is again c s 2 =1/3b/h 2 .III. SINGLE-SITE FLUCTUATIONSWe now consider <strong>the</strong> distribution <strong>of</strong> small fluctuations in<strong>the</strong> mass densities associated with each velocity direction,n i neq =n i −n i eq . Using <strong>the</strong> results <strong>of</strong> <strong>the</strong> Appendix to incorporate<strong>the</strong> constraints, and converting from fluctuations in i t<strong>of</strong>luctuations in n i ,Pn i neq exp− in i neq 22n ieqin neq i c i n neq i .i32The variance in <strong>the</strong> fluctuations depends on direction, but,since n neq i is already a small quantity in comparison with n eq i ,we will approximate <strong>the</strong> variance by <strong>the</strong> low-velocity limit,lim n eq i = a c i.33u→0The velocity dependence <strong>of</strong> <strong>the</strong> fluctuations in <strong>the</strong> GLGmodel is a consequence <strong>of</strong> <strong>the</strong> broken Galilean invariance,which is only entirely restored in <strong>the</strong> limit u→0. However,<strong>the</strong> approximation in Eq. 33 makes no difference to <strong>the</strong>macroscopic dynamics <strong>of</strong> <strong>the</strong> <strong>fluctuating</strong> LB model Eqs.82–85, since <strong>the</strong> stress fluctuations are already secondorderin <strong>the</strong> Chapman-Enskog expansion.We now introduce normalized fluctuations x i , via <strong>the</strong> definitionn neq i = ac i x i ,and transform Eq. 32 to <strong>the</strong> simplified expressionx i2i ac i x i ac i c i x i .i34Px i exp− 1 2 i35Equations 6, 34, and 35 define <strong>the</strong> statistics <strong>of</strong> our <strong>fluctuating</strong>LB model. Equation 35 is entirely consistent with<strong>the</strong> proposed GLG model, within <strong>the</strong> approximation expressedin Eq. 33.The LB collision operator can be conveniently representedin terms <strong>of</strong> modes, which are linear combinations <strong>of</strong><strong>the</strong> mass densities, n i 11, with basis vectors constructedfrom orthogonal polynomials in <strong>the</strong> velocity set c i . There ismore than one possible choice for <strong>the</strong>se basis vectors 26,and we use <strong>the</strong> “weighted” set 10,26, for which only <strong>the</strong>hydrodynamic modes mass density, momentum density, andstress have a projection on <strong>the</strong> equilibrium distribution. We<strong>the</strong>n write <strong>the</strong> nonequilibrium distribution as an orthonormaltransformation <strong>of</strong> <strong>the</strong> scaled variables, x i :m k = ê ki x i ,i36x i = ê ki m k ,37kwhere m k is <strong>the</strong> amplitude <strong>of</strong> <strong>the</strong> kth mode, and <strong>the</strong> basisvectors satisfy <strong>the</strong> orthonormality conditions ê ki ê li = kl .38iIt should be noted that <strong>the</strong> basis vectors ê ki are different from<strong>the</strong> e ki defined in Ref. 26, since <strong>the</strong>re <strong>the</strong> transformationwas for unscaled variables, n i , ra<strong>the</strong>r than <strong>the</strong> scaled variables,x i , used here. The essential physics <strong>of</strong> <strong>the</strong> transformationis, however, unchanged; <strong>the</strong> present expressions are justa reparametrization. The basis vectors ê ki are related to <strong>the</strong>weighted basis vectors used in Ref. 26:ê ki = ac iw ke ki ,39where w k is <strong>the</strong> length <strong>of</strong> <strong>the</strong> kth basis vector,w k = a c ie 2 ki . 40iThe hydrodynamic modes, mass density, momentum density,and stress, can be written in a model-independent form.Explicitly,ê 0i = ac i41for <strong>the</strong> mass mode, andê i = ac i2c c i,s =1, ...,d 42for <strong>the</strong> momentum modes. Note that in our formalism m 0 andm =1,...,d are zero.In addition to <strong>the</strong> conserved modes, <strong>the</strong>re are dd+1/2viscous modes: one bulk mode, d−1 shear modes involvingdiagonal elements <strong>of</strong> <strong>the</strong> c i c i tensor, and dd−1/2 <strong>of</strong>fdiagonalelements. The bulk stress mode is given by2d c 2 i − dc 2 s , 43where orthogonality to <strong>the</strong> mass mode is assured by Schmidtorthogonalization. There is a shear mode <strong>of</strong> <strong>the</strong> formê d+1,i = 1 c s2 ac iê d+2,i = 1 c s2 ac i2dd −1 dc 2 ix − c 2 i , 44and d−2 shear modes <strong>of</strong> <strong>the</strong> form d2 a c iê d+3,i =22c c 2 iy − c 2 iz , 45stoge<strong>the</strong>r with additional modes formed by cyclic permutations<strong>of</strong> <strong>the</strong> Cartesian indexes. The dd−1/2 <strong>of</strong>f-diagonalshear stresses are <strong>of</strong> <strong>the</strong> form a c iê 2d+1,i =2c c ixc iystoge<strong>the</strong>r with cyclic permutations.PHYSICAL REVIEW E 76, 036704 200746036704-4


STATISTICAL MECHANICS OF THE FLUCTUATING …TABLE I. Basis vectors <strong>of</strong> <strong>the</strong> D2Q9 model. Each row correspondsto a different basis vector, with <strong>the</strong> actual polynomial in ĉ ishown in <strong>the</strong> second column; <strong>the</strong> components <strong>of</strong> ĉ i =c i h/b arenormalized to unity. The orthonormal basis vectors ê ki can be obtainedfrom <strong>the</strong> table using Eq. 39, ê ki = ac i /w k e ki : <strong>the</strong> normalizingfactor for each basis vector is in <strong>the</strong> third column.k e ki w k0 1 11 ĉ ix 1/32 ĉ iy 1/33 3ĉ 2 i −2 44 2ĉ 2ix −ĉ i 4/95 ĉ ix ĉ iy 1/96 3ĉ 2 i −4ĉ ix 2/37 3ĉ 2 i −4ĉ iy 2/38 9ĉ 4 i −15ĉ 2 i +2 16All <strong>the</strong>se vectors are mutually orthogonal. Fur<strong>the</strong>r orthogonalvectors, whose span are <strong>the</strong> kinetic or “ghost”10 modes, may be constructed in terms <strong>of</strong> higher-orderpolynomials <strong>of</strong> c i 11; <strong>the</strong>se are model dependent. Completesets <strong>of</strong> basis vectors 26 for <strong>the</strong> D2Q9 and D3Q19 LB models4 are given in Tables I and II, respectively.TABLE II. Basis vectors <strong>of</strong> <strong>the</strong> D3Q19 model. Each row correspondsto a different basis vector, with <strong>the</strong> actual polynomial in ĉ ishown in <strong>the</strong> second column; <strong>the</strong> components <strong>of</strong> ĉ i =c i h/b arenormalized to unity. The orthonormal basis vectors ê ki can be obtainedfrom <strong>the</strong> table using Eq. 39, ê ki = ac i /w k e ki : <strong>the</strong> normalizingfactor for each basis vector is in <strong>the</strong> third column.k e ki w k0 1 11 ĉ ix 1/32 ĉ iy 1/33 ĉ iz 1/34 ĉ 2 i −1 2/35 3ĉ 2ix −ĉ i 4/36 ĉ 2iy −ĉ iz4/97 ĉ ix ĉ iy 1/98 ĉ iy ĉ iz 1/99 ĉ iz ĉ ix 1/910 3ĉ 2 i −5ĉ ix 2/311 3ĉ 2 i −5ĉ iy 2/312 3ĉ 2 i −5ĉ iz 2/313 ĉ iy −ĉ iz ĉ ix 2/914 ĉ iz −ĉ ix ĉ iy 2/915 ĉ ix −ĉ iy ĉ iz 2/916 3ĉ 4 i −6ĉ 2 i +1 217 2ĉ 2 i −33ĉ ix −ĉ 2 i 4/318 2ĉ 2 i −3ĉ iy −ĉ iz 4/9PHYSICAL REVIEW E 76, 036704 2007Equation 35 can be rewritten using Eqs. 37 and 38 togive <strong>the</strong> nonequilibrium probability distribution <strong>of</strong> <strong>the</strong> modesm k ,Pm k exp− 1 2 k47There is no contribution to P from <strong>the</strong> conserved modes.m k2 m i exp− 1 mid 2 k2.kdIV. STOCHASTIC COLLISIONS AS A MONTE CARLOPROCESSIn this section we construct a stochastic collision operator,viewed as a Monte Carlo process, and consider <strong>the</strong> localdynamics at <strong>the</strong> level <strong>of</strong> a single <strong>lattice</strong> site. In <strong>the</strong> nextsection Sec. V we will consider <strong>the</strong> global dynamics,through a Chapman-Enskog expansion. A deterministic collisionoperator at <strong>the</strong> microscopic level is quite complicatedto construct, even for <strong>the</strong> simplest three-dimensional LGmodels 27, and cannot be easily extended to <strong>the</strong> largernumber <strong>of</strong> particles in Eq. 8. Collision rules are mucheasier to construct at <strong>the</strong> <strong>Boltzmann</strong> level 3; <strong>the</strong> stochasticupdate from precollision to postcollision populations, n i→n i , is facilitated by making <strong>the</strong> transition between modes,m k →m k , since each degree <strong>of</strong> freedom is <strong>the</strong>n independent.Denoting a transition probability between <strong>the</strong> pre- and postcollisionstates <strong>of</strong> a particular mode m by m→m , <strong>the</strong>condition <strong>of</strong> detailed balance, governed by <strong>the</strong> distribution inEq. 47, readsm → m m → m = exp− m2 /2exp− m 2 /2 . 48A simulation at <strong>the</strong> hydrodynamic level does not need tosatisfy this condition, and typically does not, but it is essentialfor a proper <strong>the</strong>rmal equilibrium <strong>of</strong> <strong>the</strong> LB fluid.There are many possible realizations <strong>of</strong> Eq. 48: onewell-known example is <strong>the</strong> Metropolis method, involving atrial move followed by a stochastic acceptance or rejectionstep to enforce detailed balance. Here we consider <strong>the</strong> linearrelaxation model typically used in LB simulations, balancedby Gaussian noise:m = m + r,49where is related to an eigenvalue <strong>of</strong> <strong>the</strong> linearized collisionoperator, =1+ see Eq. 8 <strong>of</strong> Ref. 26, and r is a Gaussianrandom variable with zero mean and unit variance. Thedissipation parameter is restricted by <strong>the</strong> linear stabilitylimit, 1, with <strong>the</strong> case 0 corresponding to “overrelaxation”.Equation 49 has <strong>the</strong> technical advantage <strong>of</strong> beingrejection-free, and <strong>the</strong> conceptual advantage <strong>of</strong> enablingan analytic calculation to be made at <strong>the</strong> Chapman-Enskoglevel see Sec. V.The parameter must be adjusted to satisfy detailed balance,Eq. 48, using <strong>the</strong> relation Eq. 49 r= −1 m −m. Since <strong>the</strong> transition probability for m→m is identicalto <strong>the</strong> probability for generating <strong>the</strong> value <strong>of</strong> r that gives m from m,036704-5


DÜNWEG, SCHILLER, AND LADDm → m = 2 2 −1/2 exp− m − m 2 /2 2 . 50There is a similar expression for <strong>the</strong> reverse transition, m →m, with m and m interchanged. From Eq. 48, we <strong>the</strong>nfind that detailed balance is satisfied for = 1− 2 1/2 .51Thus <strong>the</strong> case =1 corresponds to a conserved mode, while=0 corresponds to m being entirely random, with nomemory <strong>of</strong> its previous value.Each mode, m k , in <strong>the</strong> LB model is assigned its own relaxationrate k , subject to <strong>the</strong> constraints <strong>of</strong> symmetry andconservation laws; <strong>the</strong> conserved modes kd require that k =1. For <strong>the</strong> bulk stress we choose a value b , and for <strong>the</strong>d+2d−1/2 shear stresses a single value s . In Refs.7,9,10 <strong>the</strong> kinetic modes were updated with k =0, but it ispossible to achieve more accurate boundary conditions witha proper tuning <strong>of</strong> <strong>the</strong> kinetic eigenvalues 26,28. Equation51 ensures that detailed balance is satisfied for all choices<strong>of</strong> k . A purely deterministic LB model is obtained by setting k =0 for all modes; physically, this corresponds to <strong>the</strong> limit<strong>of</strong> m p →0, or ¯i→.The original derivation <strong>of</strong> <strong>the</strong> <strong>fluctuating</strong> LB model 7,9is obtained by setting k = k =0 for all <strong>the</strong> kinetic modes, butchoosing <strong>the</strong> variance <strong>of</strong> <strong>the</strong> stresses according to Eq. 51.The kinetic modes are projected out at every time step bythis collision rule, and m k →m k =0=1. However, <strong>the</strong>re isno route back to <strong>the</strong> precollisional state, m k =0→m k =0,and detailed balance Eq. 48 is clearly violated. Never<strong>the</strong>less,this model still yields <strong>the</strong> correct <strong>fluctuating</strong> hydrodynamicsin <strong>the</strong> limit <strong>of</strong> large length scales 9, as is shown by<strong>the</strong> analysis in Sec. V. Treating all <strong>the</strong> nonconserved modeson an equal basis 10 satisfies detailed balance on all scales,and is entirely equivalent to Eqs. 48–51.As a general rule, proper <strong>the</strong>rmalization requires as manyrandom variables as <strong>the</strong>re are degrees <strong>of</strong> freedom in <strong>the</strong> systemnot counting <strong>the</strong> conserved variables. However, in <strong>the</strong>special case where k =0 for all kinetic modes, <strong>the</strong> deterministicLB model can be propagated forward in time from <strong>the</strong>mass, momentum, and stress at each <strong>lattice</strong> site. Thus it istempting to conclude that in this instance only <strong>the</strong> stressmodes should be <strong>the</strong>rmalized 6–8, since <strong>the</strong> n i play <strong>the</strong>role <strong>of</strong> auxiliary variables. However, this is incorrect; if both k =0 and k =0 for any mode, it is impossible to reconstruct<strong>the</strong> reverse trajectory and <strong>the</strong> system will not reach <strong>the</strong>rmalequilibrium. The number <strong>of</strong> degrees <strong>of</strong> freedom is only welldefined,<strong>the</strong>refore, when all k 0. Situations where some k =0 should be treated as a limit <strong>of</strong> finite k , and continuitytells us that <strong>the</strong> number <strong>of</strong> random variables should be <strong>the</strong>same.The update rule in Eq. 49, with 0, is an exact solution<strong>of</strong> a continuous Langevin <strong>equation</strong> 29,30,ddt m =−m + 52with t=0 and tt=2t−t. Integrating Eq.52 from t=0 to t=h i.e., one LB time step gives Eq. 49,with =exp−h. The standard first-order Euler approximationto Eq. 52 corresponds to =1−h, and is only validfor small h. By contrast Eq. 49 does not impose any restrictionon <strong>the</strong> time step.For <strong>the</strong> Chapman-Enskog analysis in Sec. V, we will need<strong>the</strong> collisional update <strong>of</strong> <strong>the</strong> nonequilibrium stress tensor, neq = i n neq i c i c i ; <strong>the</strong> equilibrium part <strong>of</strong> <strong>the</strong> stress is unchangedby <strong>the</strong> collision process. We first decompose neqinto a multiple <strong>of</strong> <strong>the</strong> unit tensor bulk stress, and a tracelesspart shear stresses, denoted by an overbar: neq = ¯ neq + 1 d neq ,53where we have used <strong>the</strong> Einstein summation convention for<strong>the</strong> Cartesian components. The change in <strong>the</strong> nonequilibriumstress tensor at a <strong>lattice</strong> site, due to collisions, can be determinedfrom Eqs. 49 and 51,¯ neq = s ¯ neq + R¯ ,54 neq = b neq + R .55The variables R are Gaussian random variables with zeromean; in addition R¯ is traceless. The covariance matrixR R is determined by <strong>the</strong> variances <strong>of</strong> <strong>the</strong> stochasticstress modes. The calculation can be simplified by observingthat <strong>the</strong> matrix is a fourth rank tensor and is <strong>the</strong>refore isotropicby <strong>the</strong> symmetries <strong>of</strong> <strong>the</strong> LB model,R R = R 1 + + R 2 . 56The unknown constants, R 1 and R 2 , can be determined fromspecial cases. For example, in <strong>the</strong> D3Q19 model defined inTable II, neq xy = 2 cs m 7 ,57and <strong>the</strong>refore, from Eqs. 49 and 54,R 2 xy = c 4 s 1− 2 s = R 1 , 58where <strong>the</strong> final equality follows from Eq. 56. Similarly, neq yy − neq zz =2 2 cs m 6 ,59and <strong>the</strong>reforeR yy − R zz 2 =4c 4 s 1− 2 s =4R 1 . 60This result is consistent with Eq. 56, which demonstratesthat <strong>the</strong> <strong>fluctuating</strong> stresses are indeed isotropic. Finally, <strong>the</strong>fluctuations in <strong>the</strong> trace, R , are related to b :R R =6c 4 s 1− 2 b =6R 1 +9R 2 . 61The general expression for <strong>the</strong> covariance in <strong>the</strong> randomstresses isR R c s4PHYSICAL REVIEW E 76, 036704 2007= 1− 2 s + + 2 d 2 s − 2 b .62This covariance matrix is different from <strong>the</strong> global fluctuationsin stress, which are superposed onto <strong>the</strong> hydrodynamicmodes Sec. V.036704-6


STATISTICAL MECHANICS OF THE FLUCTUATING …V. CHAPMAN-ENSKOG EXPANSIONIn order to determine <strong>the</strong> behavior on hydrodynamiclength and time scales, we apply <strong>the</strong> Chapman-Enskogmethod to <strong>the</strong> stochastic dynamics <strong>of</strong> <strong>the</strong> <strong>fluctuating</strong> LBmodel. We modify <strong>the</strong> derivation <strong>of</strong> Ref. 9 to include <strong>the</strong>rmalfluctuations: for an alternative procedure, see Ref. 31.Here, <strong>the</strong> expansion parameter is used to separate <strong>the</strong> <strong>lattice</strong>scale, r, from <strong>the</strong> hydrodynamic scale, r 1 =r. Thus = 1 , with <strong>the</strong> notation =/r .Since <strong>the</strong> collision operator is local in space and time, <strong>the</strong>nonequilibrium distribution is also taken to be <strong>of</strong> order :n i neq =n i 1 , with n i 1 <strong>of</strong> order unity in <strong>the</strong> expansion,n i = n i eq + n i 1 .63We use <strong>the</strong> usual multiple time scale expansion 32, t= t1 + 2 t2 , to separate <strong>the</strong> convective t 1 and diffusive t 2 relaxation processes. The left-hand side <strong>of</strong> Eq. 1 is expandedabout r,t in a Taylor series with respect to h: to firstorder in , t1 + c i 1 n i eq = h −1 i .64Multiplying this <strong>equation</strong> by one <strong>of</strong> <strong>the</strong> basis vectors andsumming over all <strong>the</strong> directions, we obtain <strong>the</strong> <strong>equation</strong>s for<strong>the</strong> dynamics <strong>of</strong> <strong>the</strong> <strong>fluctuating</strong> LB model on <strong>the</strong> t 1 timescale, t1 in i eq e ki + 1 in eq i c i e ki = h −1 i e ki .i65Note that we use <strong>the</strong> e ki basis vectors here, in conjunctionwith n i eq and n i neq , not <strong>the</strong> normalized basis vectors ê ki ,which are for <strong>the</strong> x i .When applied to <strong>the</strong> conserved degrees <strong>of</strong> freedom, kd, Eq. 65 leads to <strong>the</strong> inviscid fluid <strong>equation</strong>s: t1 + 1 j =0, t1 j + 1 c s 2 + u u =0.6667The equilibrium distribution contains polynomials in c i up tosecond order, and is thus automatically orthogonal to <strong>the</strong>kinetic modes, which are made up <strong>of</strong> third-order and fourthorderpolynomials in c i . Since <strong>the</strong> equilibrium distributionhas no projection on <strong>the</strong> kinetic modes, <strong>the</strong> time-derivative inEq. 65 vanishes identically for kd 2 +3d/2. However,<strong>the</strong> gradient term in Eq. 65 includes an additional factor <strong>of</strong>c i : thus third-order polynomials survive, making small equilibriumcontributions <strong>of</strong> order u 2 to <strong>the</strong> dynamics.At order 2 , <strong>the</strong> <strong>Boltzmann</strong> <strong>equation</strong> is t2 n i eq + t1 n i neq + c i 1 n i neq + h 2 t 1 t1 + c i 1 n ieq+ h 2 1 t1 + c i 1 n i eq c i =0, 71where <strong>the</strong> terms have been grouped to suggest <strong>the</strong> most expedientmeans <strong>of</strong> calculation. Since only <strong>the</strong> hydrodynamicmodes survive to <strong>the</strong> t 2 time scale, we consider just <strong>the</strong>modes up to k=d. It follows immediately from Eq. 71 and<strong>the</strong> conservations laws Eqs. 66 and 67 that <strong>the</strong> fluid isincompressible on <strong>the</strong> t 2 time scale, t2 =0.72The momentum <strong>equation</strong> can be written as t2 j + 1 neq + 1 2 − =0, 73where we can use Eq. 69 to substitute <strong>the</strong> velocity gradientsfor − . This is <strong>the</strong> usual <strong>lattice</strong> correction to <strong>the</strong> viscousmomentum flux 9. The kinetic modes make no contributionto <strong>the</strong> hydrodynamic variables, and j , at longtimes.The nonequilibrium stress can be calculated by combining<strong>the</strong> stress update rule, Eqs. 54 and 55, with Eq. 69. Forexample, from Eq. 54,and from Eq. 69PHYSICAL REVIEW E 76, 036704 2007 xy − eq xy = s xy − eq xy + R xy ,74Similarly, for <strong>the</strong> stress modes, dkd 2 +3d/2, we find: t1 c s 2 + u u + c s 2 1 j + 1 j + 1 j Eliminating xy xy − xy = hc 2 s 1 x u y + 1 y u x .from <strong>the</strong>se two <strong>equation</strong>s,75= h −1 − . 68Evaluating <strong>the</strong> time derivatives in Eq. 68 gives a simplifiedexpression for <strong>the</strong> nonequilibrium stress, apart from smallterms <strong>of</strong> order u 3 9, − = hc 2 s 1 u + 1 u . 69Thus, on <strong>the</strong> t 1 time scale, <strong>the</strong> viscous stresses fluctuatearound a mean value that is slaved to <strong>the</strong> velocity gradient, 1 u + 1 u .The kinetic modes fluctuate around zero on <strong>the</strong> t 1 timescale, with at most a small correction <strong>of</strong> order u 2 :m k − m k = Ou 2 .701− s neq xy + hc 2 s 1 x u y + 1 y u x = R xy . 76In <strong>the</strong> general case, we again decompose <strong>the</strong> stress into itstrace and traceless parts, neq =− hc 2s1− s 1 u + 1 u − 2 d 1 u − hc 2s1− b 2 d 1 u − Q ,77where <strong>the</strong> random stress tensor on <strong>the</strong> macroscopic level is1 1 1Q =− R¯ −1− s 1− b d R . 78Equation 73 can now be rewritten in terms <strong>of</strong> <strong>the</strong> viscousand <strong>fluctuating</strong> stresses036704-7


DÜNWEG, SCHILLER, AND LADD t2 j = 1 Q + 1 1 u + 1 u + − 2 d 1 u .79The deterministic part <strong>of</strong> <strong>the</strong> stress tensor has <strong>the</strong> desiredNewtonian form 5, with <strong>the</strong> usual expressions 9 for <strong>the</strong>shear viscosity and bulk viscosity :2 = hc s 1+ s,2 1− s280 = hc s 1+ b.81d 1− bCombining <strong>the</strong> momentum transport on <strong>the</strong> t 1 and t 2 timescales we obtain <strong>the</strong> <strong>equation</strong>s <strong>of</strong> <strong>fluctuating</strong> hydrodynamics5, t u + u u + c s 2 t + u =0,= Q + u + u + − 2 d u ,8283with random stresses Q . These are Gaussian variables withzero mean and a covariance matrix that can be calculatedfrom <strong>the</strong> analogous result on <strong>the</strong> microscopic level, Eq. 62:Q Q = 2m 2pc sb d + + − 2 .h d84This is <strong>the</strong> discrete analog <strong>of</strong> <strong>the</strong> covariance matrix <strong>of</strong> <strong>the</strong><strong>fluctuating</strong> stresses given by Landau and Lifshitz 5. Thedelta functions in space and time that appear in <strong>the</strong> continuum<strong>the</strong>ory are here converted into factors b −d and h −1 .Thus <strong>the</strong> stress fluctuations depend on <strong>the</strong> discretization <strong>of</strong>space and time. Equation 84 can be made consistent with<strong>the</strong> amplitude <strong>of</strong> <strong>fluctuating</strong> stresses in Ref. 5 by choosingk B T = m p c 2 s .85This is exactly <strong>the</strong> relation expected from <strong>the</strong> <strong>equation</strong> <strong>of</strong>state <strong>of</strong> an iso<strong>the</strong>rmal, ideal gas. In o<strong>the</strong>r words, our resultsare simultaneously consistent with macroscopic <strong>the</strong>rmodynamicsand <strong>fluctuating</strong> hydrodynamics.VI. CHOICE OF PARAMETERSThe <strong>fluctuating</strong> LB model has been used to simulate arange <strong>of</strong> s<strong>of</strong>t-matter physics, such as colloidal suspensions6 and polymer solutions 33,34. In such cases it is necessaryto match <strong>the</strong> LB parameters to <strong>the</strong> mass density, temperature,and viscosity <strong>of</strong> <strong>the</strong> molecular system. In addition<strong>the</strong>re are two parameters that control <strong>the</strong> accuracy <strong>of</strong> <strong>the</strong> LBsimulation without affecting <strong>the</strong> physics being simulated;namely <strong>the</strong> grid spacing, b, and <strong>the</strong> time step, h. The gridspacing must be related to <strong>the</strong> characteristic length scale <strong>of</strong>PHYSICAL REVIEW E 76, 036704 2007<strong>the</strong> physical system. For example, in coupling <strong>the</strong> LB fluid tos<strong>of</strong>t matter, like polymer chains, colloidal particles, or membranes,<strong>the</strong> length would be <strong>the</strong> size <strong>of</strong> <strong>the</strong> object. For flow incomplex geometries, it would be <strong>the</strong> channel width, while forsimulations <strong>of</strong> turbulent flow, it would be <strong>the</strong> Kolmogorovlength. This length scale, plus <strong>the</strong> desired spatial resolution,fixes <strong>the</strong> <strong>lattice</strong> spacing b in absolute units. Choosing a suitabletime step <strong>the</strong>n automatically sets <strong>the</strong> speed <strong>of</strong> soundc s =ĉ s b/h, where ĉ s is a dimensionless property <strong>of</strong> <strong>the</strong> LBmodel; for example, in <strong>the</strong> D2Q9 and D3Q19 models ĉ s= 1/3. Typically, <strong>the</strong> sound speed will be unrealisticallysmall for a dense liquid; however, this is not crucial since <strong>the</strong>LB method only runs in flow regimes where density fluctuationsare negligible.Once <strong>the</strong> length and time scales have been set, we canmatch <strong>the</strong> shear and bulk viscosities to <strong>the</strong> molecular system.Equations 80 and 81 suggest using b and h to computenondimensional viscosities from <strong>the</strong> reference values,ˆ =ˆ =2hc = hs b 2 2ĉ ,s2hc = hs b 2 2ĉ .sThe parameters s and b are <strong>the</strong>n set by ˆ and ˆ: s =2ˆ −12ˆ +1 ,868788dˆ −1 b = . 89dˆ +1Small time steps <strong>the</strong>refore imply that <strong>the</strong> LB simulation isrun in <strong>the</strong> over-relaxation regime. The relaxation rates <strong>of</strong> <strong>the</strong>kinetic modes can be chosen for convenience k =0 or toimprove <strong>the</strong> accuracy <strong>of</strong> <strong>the</strong> boundary conditions 26,28.The remaining LB parameter is <strong>the</strong> particle mass, m p ,which must be fixed, for a given b and h, so that <strong>the</strong> fluctuationsin <strong>the</strong> LB fluid are consistent with <strong>the</strong> temperature, Eq.85. The parameter =m p /b d determines <strong>the</strong> variance in <strong>the</strong>fluctuations Eq. 32, = k BTh 2ĉ 2 s b d+2 ,90from which we see that too fine a grid or too large a time stepwill cause an unacceptably high noise level. A stable simulationwill require that <strong>the</strong> time step scales as hb d/2+1 orb 5/2 in three dimensions, which is slightly more stringentthan <strong>the</strong> usual diffusive scaling, hb 2 .VII. CONCLUSIONSFor models <strong>of</strong> <strong>the</strong> D3Q19 type, our analysis has shownthat a <strong>fluctuating</strong> LB <strong>equation</strong> can be developed from statisticalmechanical considerations. We have shown that <strong>the</strong>fluctuations are governed by <strong>the</strong> degree <strong>of</strong> coarse graining,and that <strong>the</strong> relevant parameter is <strong>the</strong> mass <strong>of</strong> <strong>the</strong> LB particle,036704-8


STATISTICAL MECHANICS OF THE FLUCTUATING …m p , which, for a given temperature, is determined by <strong>the</strong>discretization <strong>of</strong> space, b, and time, h. The temperature appearingin <strong>the</strong> <strong>equation</strong> <strong>of</strong> state is identical to that whichcontrols <strong>the</strong> fluctuations, as it should be.The beauty <strong>of</strong> <strong>the</strong> present approach is that one only needsto take care that <strong>the</strong> statistical properties are correct at <strong>the</strong> LBlevel. The correct fluctuation-dissipation <strong>the</strong>orem at <strong>the</strong>Navier-Stokes level is <strong>the</strong>n an automatic consequence <strong>of</strong> <strong>the</strong>microscopic physics. We have introduced <strong>the</strong> principle <strong>of</strong>detailed balance into <strong>the</strong> LB model, which is <strong>the</strong> microscopiccounterpart <strong>of</strong> <strong>the</strong> fluctuation-dissipation <strong>the</strong>orem used inprevious work 7,9,10. We have demonstrated that all nonconservedmodes must be <strong>the</strong>rmalized 10 in order to satisfydetailed balance; early implementations <strong>of</strong> <strong>the</strong> <strong>fluctuating</strong> LBmodel 7,9 did not preserve detailed balance. On <strong>the</strong> o<strong>the</strong>rhand, all <strong>the</strong>se methods have been shown to be correct in <strong>the</strong>hydrodynamic limit. Only <strong>the</strong> stress fluctuations survive tolong times, and fluctuations in <strong>the</strong> kinetic modes becomeasymptotically irrelevant. Never<strong>the</strong>less, practical simulationsrarely probe <strong>the</strong> asymptotic limit, and <strong>the</strong>n a procedurewhich is statistically correct on all length scales is clearlypreferable.ACKNOWLEDGMENTSWe thank R. Adhikari, M. E. Cates, and A. J. Wagner forvery stimulating discussions on <strong>the</strong> subject. U.D.S. thanks<strong>the</strong> Volkswagen Foundation for support within <strong>the</strong> framework<strong>of</strong> <strong>the</strong> program “New conceptual approaches to modelingand simulation <strong>of</strong> complex systems.” A.J.C.L. thanks <strong>the</strong>Alexander von Humboldt Foundation for supporting his stayat <strong>the</strong> Max Planck Institute for Polymer Research.APPENDIX: CONSTRAINED DISTRIBUTIONSLet us consider a constrained probability distributionP i <strong>of</strong> <strong>the</strong> following general form:P i expS i j i ij − q j ,iA1where S is a function <strong>of</strong> i , and ij and q j are constants.The constraints can be eliminated by making use <strong>of</strong> <strong>the</strong> Fourierrepresentation <strong>of</strong> <strong>the</strong> delta function, x=2 −1 expikxdk:whereP i dk j expŜ i ,k j , A2jŜ i ,k j = S i + ijk j i ij − q j .iA3Now, let i 0 , k j 0 denote <strong>the</strong> saddle point <strong>of</strong> Ŝ, which can befound by solving <strong>the</strong> system <strong>of</strong> <strong>equation</strong>s:Ŝ=0⇔ S i i+ ijk j ij =0,A4Ŝ=0⇔k j i ij − q j =0. A5iThe solution 0 isatisfies <strong>the</strong> constraints in Eq. A1 and isidentical to <strong>the</strong> one obtained by maximizing S, taking intoaccount <strong>the</strong> constraints via Lagrange multipliers, ik j .The second-order Taylor expansion <strong>of</strong> Ŝ around <strong>the</strong> saddlepoint isŜ i ,k j = Ŝ i 0 ,k j 0 + ilwhere we have introduced <strong>the</strong> abbreviations il = 1 2 2 il i l + i ij i k j ,ijS, i l0 i i = i − i 0 ,k j = k j − k j 0 .A6A7A8A9The probability distribution for i is <strong>the</strong>n approximatelyGaussian.The expansion <strong>of</strong> Ŝ is now inserted into Eq. A2. Ignoring<strong>the</strong> constant term, which can be absorbed in <strong>the</strong> normalization<strong>of</strong> P, and transforming to <strong>the</strong> new variables k j ,wefind ij i P i jPHYSICAL REVIEW E 76, 036704 2007 dk j exp ik jiexp il i l .ilReintroducing delta functions, we obtain <strong>the</strong> final resultP i exp il i l iljA10 ij i . A11iAssuming <strong>the</strong> coefficients ij form a negative-definite matrixo<strong>the</strong>rwise <strong>the</strong> Gaussian approximation would not makesense, <strong>the</strong> saddle point is a maximum in P.1 S. Succi, The Lattice <strong>Boltzmann</strong> Equation for Fluid Dynamicsand Beyond Oxford University Press, New York, 2001.2 R. Benzi, S. Succi, and M. Vergassola, Phys. Rep. 222, 1451992.3 F. Higuera, S. Succi, and R. Benzi, Europhys. Lett. 9, 3451989.4 Y. H. Qian, D. d’Humières, and P. Lallemand, Europhys. Lett.17, 479 1992.036704-9


DÜNWEG, SCHILLER, AND LADD5 L. D. Landau and E. M. Lifshitz, Fluid Mechanics Addison-Wesley, London, 1959.6 A. J. C. Ladd, Phys. Rev. Lett. 70, 1339 1993.7 A. J. C. Ladd, J. Fluid Mech. 271, 285 1994.8 A. J. C. Ladd, J. Fluid Mech. 271, 3111994.9 A. J. C. Ladd and R. Verberg, J. Stat. Phys. 104, 1191 2001.10 R. Adhikari, K. Stratford, M. E. Cates, and A. J. Wagner, Europhys.Lett. 71, 473 2005.11 D. D’Humières, I. Ginzburg, M. Kraftzyck, P. Lallemand, andL. S. Luo, Philos. Trans. R. Soc. London, Ser. A 360, 4372002.12 A. J. Wagner, Europhys. Lett. 44, 144 1998.13 I. V. Karlin, A. Ferrante, and H. C. Öttinger, Europhys. Lett.47, 182 1999.14 B. M. Boghosian, P. J. Love, P. V. Coveney, I. V. Karlin, S.Succi, and J. Yepez, Phys. Rev. E 68, 025103R 2003.15 D. P. Landau and K. Binder, A Guide to Monte Carlo Simulationsin <strong>Statistical</strong> Physics Cambridge University Press, Cambridge,England, 2000.16 P. Español and P. Warren, Europhys. Lett. 30, 191 1995.17 A. Malevanets and R. Kapral, J. Chem. Phys. 110, 86051999.18 G. R. McNamara, A. L. Garcia, and B. J. Alder, J. Stat. Phys.87, 1111 1997.19 L.-S. Luo, Phys. Rev. E 62, 4982 2000.PHYSICAL REVIEW E 76, 036704 200720 L.-S. Luo and S. S. Girimaji, Phys. Rev. E 67, 036302 2003.21 M. R. Swift, E. Orlandini, W. R. Osborn, and J. M. Yeomans,Phys. Rev. E 54, 5041 1996.22 P. Ahlrichs, Ph.D. <strong>the</strong>sis, Johannes Gutenberg–Universität,Mainz, 2000 unpublished.23 G. R. McNamara and G. Zanetti, Phys. Rev. Lett. 61, 23321988.24 U. Frisch, B. Hasslacher, and Y. Pomeau, Phys. Rev. Lett. 56,1505 1986.25 U. Frisch, D. d’Humières, B. Hasslacher, P. Lallemand, Y.Pomeau, and J.-P. Rivet, Complex Syst. 1, 649 1987.26 B. Chun and A. J. C. Ladd, Phys. Rev. E 75, 066705 2007.27 M. Hénon, Complex Syst. 1, 763 1987.28 I. Ginzburg and D. d’Humières, Phys. Rev. E 68, 0666142003.29 S. Chandrasekar, Rev. Mod. Phys. 15, 11943.30 H. Risken, The Fokker-Planck Equation. Methods <strong>of</strong> Solutionand Applications Springer-Verlag, Berlin, 1984.31 M. Junk, A. Klar, and L.-S. Luo, J. Comput. Phys. 210, 6762005.32 E. J. Hinch, Perturbation Methods Cambridge UniversityPress, Cambridge, England, 1991.33 P. Ahlrichs and B. Dünweg, J. Chem. Phys. 111, 8225 1999.34 O. Berk Usta, A. J. C. Ladd, and J. E. Butler, J. Chem. Phys.122, 094902 2005.036704-10

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!