A consistence-stability approach to hydrodynamic limit of interacting particle systems on lattices
Abstract.
This is a review based on the presentation done at the seminar Laurent Schwartz in December . It is announcing results in the forthcoming [MMM22]. This work presents a new simple quantitative method for proving the hydrodynamic limit of a class of interacting particle systems on lattices. We present here this method in a simplified setting, for the zero-range process and the Ginzburg-Landau process with Kawasaki dynamics, in the parabolic scaling and in dimension . The rate of convergence is quantitative and uniform in time. The proof relies on a consistence-stability approach in Wasserstein distance, and it avoids the use of the “block estimates”.
1. The general method
We consider the hydrodynamic limit of interacting particle systems on a lattice. The problem is to show that under an appropriate scaling of time and space, the local particle densities of a stochastic lattice gas converge to the solution of a macroscopic partial differential equation. We first present our method abstractly and then sketch applications to two concrete models: the zero-range process (ZRP) and the Ginzburg Landau process with Kawasaki dynamics (GLK). The hydrodynamic limit is known at a qualitative level for all these models under both hyperbolic and parabolic scalings for the ZRP and under parabolic scaling for the GLK, see [GPV88, Yau91, Rez91, KL99]. However finding quantitative error estimates had remained an important opened question, as well as understanding the long-time behaviour of the hydrodynamic limit. First results towards quantitative error, in the particular case of the Ginzburg-Landau process with Kawasaki dynamics in dimension , were obtained in the two-parts work [DMOWa, DMOWb], which builds upon partial progresses in [GOVW09].
1.1. Set up and notation
We denote by the state space at a given site (number of particles, spin, etc.), which will in this paper be (ZRP) or (GLK). Consider the discrete torus and the corresponding phase space of particle configurations . Variables in are called microscopic and denoted by , whereas variables in the limit continuous torus are called macroscopic and denoted by ; finally particle configurations in are denoted by . The canonical embeddding , means the macroscopic distance between sites of the lattice is . Given a particle configuration , we define the empirical measure
(1.1) |
where denotes the value of at , and is the space of positive Radon measures on the torus, and denotes the “average sum”, here .
At the microscopic level, the interacting particle system evolves through a stochastic process and the time-dependent probability measure describing the law of is denoted by . We consider a linear operator generating uniquely a Feller semigroup on (see [Lig85, Chapter 1]) so that given the solution satisfies
(1.2) |
where denotes continuous bounded real-valued functions and denotes the duality bracket between and .
At the macroscopic level, we consider a map (in general unbounded and nonlinear) and the evolution problem
(1.3) |
A measure is called invariant for (1.2) if
We also denote the Lipschitz functions with respect to the (normalised) norm: for every , , and we denote the smallest such constant by .
1.2. Abstract assumptions
(H0) Local equilibrium structure. There are depending on (Conv denotes the convex hull) and so that: (i) is invariant on for each , and (ii) for any , . We then define, given a macroscopic profile on , the local Gibbs measure
The two maps and allow comparisons between the microscopic and macroscopic scales, as summarized in Figure 1.

(H1) Microscopic stability. The semigroup satisfies
(1.4) |
(H2) Macroscopic stability. There is a Banach space so that (1.3) is locally well-posed in ; given the maximal time of existence we denote for , when (1.3) has a unique stationary solution with mass , otherwise we denote .
(H3) Consistency. There is a consistency error as so that for
for any , where is an equilibrium measure.
1.3. The abstract strategy
Theorem 1.1.
2. Concrete applications
We apply the abstract result to two archetypical models, the zero-range process (ZRP), and the Ginzburg-Landau process with Kawasaki dynamics (GLK).
2.1. The ZRP
In this case, the state space at each site is . Given the choice of a transition function with and a jump rate function , the base generator writes
(2.1) |
where is defined as before. The local equilibrium structure of (H0) is given by
(2.2) | ||||
(2.3) |
denoting . The pair thus constructed satisfies . When is constant, the local Gibbs measure is invariant with average number of particles . The mean transition rate is defined by . When , the first non-zero asymptotic dynamics as is given by the hyperbolic scaling , and the corresponding expected limit equation is . When , the first non-zero asymptotic dynamics as is the given by the parabolic scaling , and the corresponding limit equation is formally
(2.4) |
We make the following assumptions on the jump rate function .
(HZRP) The jump rate satisfies , for all , is non-decreasing, uniformly Lipschitz , and there are and such that for any .
The main result on the ZRP is:
2.2. The GLK
In this case, the state space at each site is . Given the choice of a single-site potential , the base generator writes
(2.6) |
where denotes neighbouring sites. The local equilibrium structure is given by
When is constant, the local Gibbs measure is invariant with average spin . The hyperbolic scaling formally leads to zero and the parabolic scaling formally leads to
(2.7) |
We assume that the single-site potential satisfies
(HGLK) The potential is and decomposes as with for all for some and .
3. The abstract strategy
In this section we sketch the proof of Theorem 1.1. Let be a solution to (1.3). Given , we denote by for the local -average .
Denote by and the densities with respect to , and write
so that Duhamel’s formula yields
Take with and integrate the above equation to get
(H1) implies and (H3) implies , which implies the conclusion of Theorem 1.1.
4. Proof for the ZRP
In this section we prove Theorem 2.1) (hydrodynamical limit for the ZRP). Note for this model is symmetric with respect to equilibrium measures.
Given with , , and , the density of the local Gibbs measure relatively to the invariant measure with mass is:
(4.1) |
where the function is defined by and the partition function is defined in (2.2), with denoting the radius of convergence of the series.
It is proved in [KL99, Chapter 2, Section 3] that assumption (HZRP) on implies that is well-defined and strictly increasing, with
Then the building block of the Gibbs measure satisfies . Moreover (HZRP) implies that the function is with uniform bound on all derivatives on , with Lipschitz constant less than , and with (in particular ), and that the invariant measure has exponential moment bounds, see [KL99, Corollary 3.6].
4.1. Microscopic Stability – (H1)
4.2. Macroscopic stability – (H2)
In the parabolic scaling the limit PDE is the nonlinear diffusion equation (2.4). We take with its standard infinity Banach norm. The proof that this norm remains uniformly bounded in time is classical in dimension (using the bounds on ), and for all times by maximal principle. Moreover exponentially fast as in .
4.3. Consistency estimate – (H3)
Let and the dimension .
Proposition 4.1.
Proof.
We start by computing
with (note that exponentially fast)
for some . Since (conservation of mass)
we can replace by
and use the Lipschitz bound on (microscopic stability) to get
with defined by (note that it has zero average against )
We then form sub-sum over non-overlapping cubes of size (this intermediate scale factor will be chosen later in terms of ). Let be a net of centers of non-overlapping cubes of the form . Then
with the defined by
where , for , denotes taking the average over the cube . Note that the average of against is . Then
where projects on the average over configurations with the same mass in the cube (and does not touch the other sites):
(4.3) |
for a function on . To estimate the first term we first approximate the measure on by the equilibrium measure with local mass , and denote it by (note that the approximation is made differently for each cube and depends on , even if it is written explicitly). This produces an error (using the Lipschitz regularity of and the exponential convergence to get uniform in time bounds). We then apply the Poincaré inequality [LSV96, Theorem 1.1] in the cube (whose constant is independent of the number of particles and proportional to the size of the cube) and the law of large number (using uniform bounds on the second moment of ):
where is the Dirichlet form on the cube with respect to the measure :
Then we change back the measure in each box, which produces (using the Lipschitz regularity of ) an error ), and we compute
(with denoting the Dirichlet form for ), where the last error accounts for the small default of self-adjointness. We deduce (in dimension ) that
To control the second term , we first use the equivalence of ensemble in [KL99, Appendix II, Corollary 1.7] on the local equilibrium measure (using uniform exponential moment bounds):
(4.4) |
Second we remark that the Lipschitz regularity of implies that , and since the average of with respect to is , we can write
Third, we remark that the Lipschitz regularity of (with constant ) implies a Lipschitz regularity of its averaged projection with constant , with respect to the local mass. Indeed, given , pick any pair of configuration with , and (such configuration trivially exists since ). Then we consider the initial coupling on which has cost . Then we evolve it along the flow of the coupling operator . The marginals respectively converge to and (convergence to equilibrium of the oiriginal evolution). Since the evolution by the coupling operator does not increase the Wasserstein distance, we deduce . An optimal coupling associated to this distance thus satisfies
where the first inequality follows from Jensen’s inequality. Thus the Jensen’s inequality is saturated which implies that the cost does not change sign on the support of , i.e. in the support. We then compute
and since on the support of , and
We deduce (using (5.3))
which yields by Taylor formula, the approximation of by , and the law of large numbers
Combining all estimates we get (optimizing )
(4.5) |
∎
5. Proof for the GLK
In this section we prove Theorem 2.2 (hydrodynamic limit for the GLK). Note again that for this model is symmetric with respect to equilibrium measures. Given and , the density of the local Gibbs measure relatively to the invariant measure with mass is:
(5.1) |
where the function is defined by and the partition function is defined on . The uniform convexity of at infinity easily implies bounds on some exponential moments of the invariant measure and (HGLK) implies that there exists so that (see [GOVW09, Lemma 41] and [DMOWa, Lemma 5.1]).
5.1. Microscopic stability – (H1)
We define a “coupling generator” by
(5.2) |
where is a constant to be chosen later and the adjoint is taken in so
Then for any there is (depending on ) so that
by using the assumptions on the potential: uniformly strictly convex and . This implies the weak contraction of the evolution in (-Wasserstein distance) for any , and thus by limit in . By duality this implies that the evolution is weakly contractive for the dual Lipschitz norm.
5.2. Macroscopic stability - (H2)
The limit equation is similar to that of the ZRP and (H2) is proved in the same way.
5.3. Consistency estimate - (H3)
Let the dimension .
Proposition 5.1.
Proof.
The proof follows the same structure as for the ZRP. We start by computing
with (note again that exponentially fast)
for some . By conservation of mass we replace again by
and use the Lipschitz bound (H1) on to get
with defined by (note that it has zero average against )
We again form sub-sum over non-overlapping cubes of size , with a net of centers of cubes . Then
with the defined by (and again denotes the average over the cube )
(Note again that the average of against is .) Then
where again averages over (and does not touch the other site) as in (4.3).
To estimate the first term we again approximate the measure on by the equilibrium measure with local mass , and denote it by (note that the approximation is made differently for each cube and depends on , even if it is written explicitly). This produces an error (using the Lipschitz regularity of and the exponential convergence to get uniform in time bounds). We then apply the Poincaré inequality [LY93, Theorem 2] in the cube (whose constant is independent of the number of particles and proportional to the size of the cube) and the law of large number (using uniform bounds on the second moment of )
where is the Dirichlet form on the cube with respect to the measure :
Then we use the entropy production
as before to deduce that
To control the second term , we first use the equivalence of ensemble in [LPY02, Corollary 5.3] on the local equilibrium measure (using bounds on some exponential moments):
(5.3) |
Second we remark that the Lipschitz regularity of implies that , and since the average of with respect to is , we can write
Third, we prove again that the Lipschitz regularity of (with constant ) implies a Lipschitz regularity of its averaged projection with constant , with respect to the local mass. Indeed, given , pick any pair of configuration with , and (such configuration trivially exists since ). Then consider the coupling on given by a product of smooth probability distributions localised around respectively and , so that the support of this coupling only contains strictly ordered . Then we evolve it along the flow of the coupling operator . The marginals respectively converge to and (convergence to equilibrium of the oiriginal evolution). Arguing as for the ZRP, we deduce that , and a corresponding optimal coupling associated to this distance is so that the cost does not change sign on its support, i.e. in the support. We deduce as for the ZRP that is -Lipschitz.
References
- [DMOWa] D. Dizdar, G. Menz, F. Otto, and T. Wu. The quantitative hydrodynamic limit of the Kawasaki dynamics. Preprint. arXiv:1807.09850.
- [DMOWb] D. Dizdar, G. Menz, F. Otto, and T. Wu. Toward a quantitative theory of the hydrodynamic limit. Preprint. arXiv:1807.09857.
- [Fat13] M. Fathi. A two-scale approach to the hydrodynamic limit part II: local Gibbs behavior. ALEA Lat. Am. J. Probab. Math. Stat., 10(2):625–651, 2013.
- [GOVW09] Ñ. Grunewald, F. Otto, C. Villani, and M. Westdickenberg. A two-scale approach to logarithmic Sobolev inequalities and the hydrodynamic limit. Ann. Inst. Henri Poincaré Probab. Stat., 45(2):302–351, 2009.
- [GPV88] M. Z. Guo, G. C. Papanicolaou, and S. R. S. Varadhan. Nonlinear diffusion limit for a system with nearest neighbor interactions. Comm. Math. Phys., 118(1):31–59, 1988.
- [KL99] C. Kipnis and C. Landim. Scaling limits of interacting particle systems, volume 320 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999.
- [Lig85] T. Liggett. Interacting Particle Systems. Springer Berlin Heidelberg, 1985.
- [LPY02] C. Landim, G. Panizo, and H. T. Yau. Spectral gap and logarithmic Sobolev inequality for unbounded conservative spin systems. Ann. Inst. H. Poincaré Probab. Statist., 38(5):739–777, 2002.
- [LSV96] C. Landim, S. Sethuraman, and S. Varadhan. Spectral gap for zero-range dynamics. Ann. Probab., 24(4):1871–1902, 1996.
- [LY93] Sheng Lin Lu and Horng-Tzer Yau. Spectral gap and logarithmic Sobolev inequality for Kawasaki and Glauber dynamics. Comm. Math. Phys., 156(2):399–433, 1993.
- [MMM22] D. Marahrens, A. Menegaki, and C. Mouhot. Quantitative hydrodynamic limit of interacting particle systems on lattices. Forthcoming, 2022.
- [Rez91] F. Rezakhanlou. Hydrodynamic limit for attractive particle systems on . Comm. Math. Phys., 140(3):417–448, 1991.
- [Yau91] H.-T. Yau. Relative entropy and hydrodynamics of Ginzburg-Landau models. Lett. Math. Phys., 22(1):63–80, 1991.