This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\stackMath

A consistence-stability approach to hydrodynamic limit of interacting particle systems on lattices

Angeliki Menegaki Institut des hautes études Scientifiques, 35 Rte de Chartres, 91440, Bures-sur-Yvette, France [email protected]  and  Clément Mouhot DPMMS, University of Cambridge, Wilberforce Road, Cambridge CB3 0WA, UK [email protected]
Abstract.

This is a review based on the presentation done at the seminar Laurent Schwartz in December 20212021. 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 11. 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 11, 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 𝖷\mathsf{X} the state space at a given site (number of particles, spin, etc.), which will in this paper be \mathbb{N} (ZRP) or \mathbb{R} (GLK). Consider the discrete torus 𝕋Nd\mathbb{T}_{N}^{d} and the corresponding phase space of particle configurations 𝖷N:=𝖷𝕋Nd\mathsf{X}_{N}:=\mathsf{X}^{\mathbb{T}^{d}_{N}}. Variables in 𝕋Nd\mathbb{T}^{d}_{N} are called microscopic and denoted by x,y,zx,y,z, whereas variables in the limit continuous torus 𝕋d\mathbb{T}^{d} are called macroscopic and denoted by uu; finally particle configurations in 𝖷N\mathsf{X}_{N} are denoted by η\eta. The canonical embeddding 𝕋Nd𝕋d\mathbb{T}_{N}^{d}\to\mathbb{T}^{d}, xxNx\mapsto\frac{x}{N} means the macroscopic distance between sites of the lattice is 1N\frac{1}{N}. Given a particle configuration η𝖷N\eta\in\mathsf{X}_{N}, we define the empirical measure

(1.1) αηN:=\stackon[3.8ex] x𝕋NdηxδxN+(𝕋d).\alpha_{\eta}^{N}:=\mathop{\mathchoice{\stackon[-3.8ex]{\displaystyle\sum}{\smash{\rule{0.4pt}{17.22217pt}}}}{\stackon[-2.6ex]{\textstyle\sum}{\smash{\rule{0.4pt}{12.48604pt}}}}{\stackon[-1.9ex]{\scriptstyle\sum}{\smash{\rule{0.4pt}{9.47217pt}}}}{\stackon[-1.4ex]{\scriptscriptstyle\sum}{\smash{\rule{0.4pt}{7.3194pt}}}}}_{x\in\mathbb{T}^{d}_{N}}\eta_{x}\delta_{\frac{x}{N}}\in\mathcal{M}_{+}(\mathbb{T}^{d}).

where ηx\eta_{x} denotes the value of η\eta at x𝕋Ndx\in\mathbb{T}^{d}_{N}, and +(𝕋d)\mathcal{M}_{+}(\mathbb{T}^{d}) is the space of positive Radon measures on the torus, and \stackon[2.6ex] \mathop{\mathchoice{\stackon[-3.8ex]{\displaystyle\sum}{\smash{\rule{0.4pt}{17.22217pt}}}}{\stackon[-2.6ex]{\textstyle\sum}{\smash{\rule{0.4pt}{12.48604pt}}}}{\stackon[-1.9ex]{\scriptstyle\sum}{\smash{\rule{0.4pt}{9.47217pt}}}}{\stackon[-1.4ex]{\scriptscriptstyle\sum}{\smash{\rule{0.4pt}{7.3194pt}}}}} denotes the “average sum”, here Ndx𝕋NdN^{-d}\sum_{x\in\mathbb{T}^{d}_{N}}.

At the microscopic level, the interacting particle system evolves through a stochastic process and the time-dependent probability measure describing the law of η\eta is denoted by μtNP(𝖷N)\mu_{t}^{N}\in P(\mathsf{X}_{N}). We consider a linear operator N:Cb(𝖷N)Cb(𝖷N)\mathcal{L}_{N}:C_{b}(\mathsf{X}_{N})\rightarrow C_{b}(\mathsf{X}_{N}) generating uniquely a Feller semigroup etNe^{t\mathcal{L}_{N}} on P(𝖷N)P(\mathsf{X}_{N}) (see [Lig85, Chapter 1]) so that given μ0NP(𝖷N)\mu_{0}^{N}\in P(\mathsf{X}_{N}) the solution μtN=etNμ0NP(𝖷N)\mu_{t}^{N}=e^{t\mathcal{L}_{N}}\mu^{N}_{0}\in P(\mathsf{X}_{N}) satisfies

(1.2) ΦCb(𝖷N),ddtΦ,μtN=NΦ,μtN,\forall\,\Phi\in C_{b}(\mathsf{X}_{N}),\quad\frac{{\rm d}}{{\rm d}t}\langle\Phi,\mu^{N}_{t}\rangle=\langle\mathcal{L}_{N}\Phi,\mu^{N}_{t}\rangle,

where Cb(𝖷N)C_{b}(\mathsf{X}_{N}) denotes continuous bounded real-valued functions and ,\langle\cdot,\cdot\rangle denotes the duality bracket between Cb(𝖷N)C_{b}(\mathsf{X}_{N}) and P(𝖷N)P(\mathsf{X}_{N}).

At the macroscopic level, we consider a map :+(𝔾)+(𝔾)\mathcal{L}_{\infty}:\mathcal{M}_{+}(\mathbb{G}_{\infty})\rightarrow\mathcal{M}_{+}(\mathbb{G}_{\infty}) (in general unbounded and nonlinear) and the evolution problem

(1.3) tft=ft,ft=0=f0.\partial_{t}f_{t}=\mathcal{L}_{\infty}f_{t},\quad f_{t=0}=f_{0}.

A measure μNP(𝖷N)\mu^{N}\in P(\mathsf{X}_{N}) is called invariant for (1.2) if

ΦCb(𝖷N),μN,NΦ=0.\forall\,\Phi\in C_{b}(\mathsf{X}_{N}),\quad\big{\langle}\mu^{N},\mathcal{L}_{N}\Phi\big{\rangle}=0.

We also denote Lip(𝖷N)\operatorname{Lip}(\mathsf{X}_{N}) the Lipschitz functions Φ:𝖷N\Phi:\mathsf{X}_{N}\to\mathbb{R} with respect to the (normalised) 1\ell^{1} norm: for every η,ζ𝖷N\eta,\zeta\in\mathsf{X}_{N}, |Φ(η)Φ(ζ)|CΦ\stackon[2.6ex] x𝕋Nd|ηxζx||\Phi(\eta)-\Phi(\zeta)|\leq C_{\Phi}\mathop{\mathchoice{\stackon[-3.8ex]{\displaystyle\sum}{\smash{\rule{0.4pt}{17.22217pt}}}}{\stackon[-2.6ex]{\textstyle\sum}{\smash{\rule{0.4pt}{12.48604pt}}}}{\stackon[-1.9ex]{\scriptstyle\sum}{\smash{\rule{0.4pt}{9.47217pt}}}}{\stackon[-1.4ex]{\scriptscriptstyle\sum}{\smash{\rule{0.4pt}{7.3194pt}}}}}_{x\in\mathbb{T}_{N}^{d}}|\eta_{x}-\zeta_{x}|, and we denote the smallest such constant CΦC_{\Phi} by [Φ]Lip(𝖷N)+[\Phi]_{\operatorname{Lip}(\mathsf{X}_{N})}\in\mathbb{R}_{+}.

1.2. Abstract assumptions

We make the following assumptions on (1.2)-(1.3):

(H0) Local equilibrium structure. There are nλ:Conv(𝖷)+n_{\lambda}:\text{Conv}(\mathsf{X})\to\mathbb{R}_{+} depending on λ\lambda\in\mathbb{R} (Conv denotes the convex hull) and σ:Conv(𝖷)\sigma:\text{Conv}(\mathsf{X})\to\mathbb{R} so that: (i) nλ𝕋Ndn_{\lambda}^{\otimes\mathbb{T}^{d}_{N}} is invariant on 𝖷N\mathsf{X}_{N} for each λ\lambda, and (ii) for any ρConv(𝖷)\rho\in\text{Conv}(\mathsf{X}), 𝔼nσ(ρ)[ηx]=ρ\mathbb{E}_{n_{\sigma(\rho)}}[\eta_{x}]=\rho. We then define, given a macroscopic profile ff on 𝕋d\mathbb{T}^{d}, the local Gibbs measure

ϑfN(η):=νσ(f(N))N(η)whereνFN(η):=x𝔾NnF(x)(η(x)).\vartheta^{N}_{f}(\eta):=\nu_{\sigma\left(f\left(\frac{\cdot}{N}\right)\right)}^{N}(\eta)\quad\text{where}\quad\nu_{F}^{N}(\eta):=\prod_{x\in\mathbb{G}_{N}}n_{F(x)}(\eta(x)).

The two maps ηαηN\eta\mapsto\alpha_{\eta}^{N} and fϑfNf\mapsto\vartheta^{N}_{f} allow comparisons between the microscopic and macroscopic scales, as summarized in Figure 1.

Refer to caption
Figure 1. The functional setting.

(H1) Microscopic stability. The semigroup etNe^{t\mathcal{L}_{N}} satisfies

(1.4) ΦLip(𝖷N),[etNΦ]Lip(𝖷N)[Φ]Lip(𝖷N).\displaystyle\forall\,\Phi\in\operatorname{Lip}(\mathsf{X}_{N}),\quad\left[e^{t\mathcal{L}_{N}}\Phi\right]_{\operatorname{Lip}(\mathsf{X}_{N})}\leq\left[\Phi\right]_{\operatorname{Lip}(\mathsf{X}_{N})}.

(H2) Macroscopic stability. There is a Banach space 𝔅+(𝔾)\mathfrak{B}\subset\mathcal{M}_{+}(\mathbb{G}_{\infty}) so that (1.3) is locally well-posed in 𝔅\mathfrak{B}; given the maximal time of existence Tm(0,+]T_{m}\in(0,+\infty] we denote for t[0,Tm)t\in[0,T_{m}), R(t):=ftf𝔅R(t):=\left\|f_{t}-f_{\infty}\right\|_{\mathfrak{B}} when (1.3) has a unique stationary solution f𝔅f_{\infty}\in\mathfrak{B} with mass 𝕋df=𝕋df0\int_{\mathbb{T}^{d}}f_{\infty}=\int_{\mathbb{T}^{d}}f_{0}, otherwise we denote R(t):=ft𝔅R(t):=\left\|f_{t}\right\|_{\mathfrak{B}}.

(H3) Consistency. There is a consistency error ϵ(N)0\epsilon(N)\to 0 as NN\to\infty so that for T[0,Tm)T\in[0,T_{m})

1T0T0t(e(ts)NΦ),[N(dϑfsNdνN)dds(dϑfsNdνN)]dνNdsdtϵ(N)[Φ]Lip(𝖷N)0tR(s)ds\frac{1}{T}\int_{0}^{T}\int_{0}^{t}\left\langle\left(e^{(t-s)\mathcal{L}_{N}}\Phi\right),\left[\mathcal{L}_{N}^{*}\left(\frac{{\rm d}\vartheta_{f_{s}}^{N}}{{\rm d}\nu^{N}_{\infty}}\right)-\frac{{\rm d}}{{\rm d}s}\left(\frac{{\rm d}\vartheta_{f_{s}}^{N}}{{\rm d}\nu^{N}_{\infty}}\right)\right]{\rm d}\nu^{N}_{\infty}\right\rangle\,{\rm d}s\,{\rm d}t\leq\epsilon(N)[\Phi]_{\operatorname{Lip}(\mathsf{X}_{N})}\int_{0}^{t}R(s)\,{\rm d}s

for any ΦLip(𝖷N)\Phi\in\operatorname{Lip}(\mathsf{X}_{N}), where νN\nu^{N}_{\infty} is an equilibrium measure.

1.3. The abstract strategy

Theorem 1.1.

Consider (1.2)-(1.3) with the assumptions (H0)–(H1)–(H2)–(H3). Let ϕC(𝕋d)\phi\in C^{\infty}(\mathbb{T}^{d}), μ0NP1(XN)\mu_{0}^{N}\in P_{1}(X_{N}) for all N1N\geq 1, f0f_{0}\in\mathcal{B}. Then

(1.5) T[0,Tm),1T0TμtNϑftNLipdtϵ(N)0TR(s)ds+μ0Nϑf0NLip.\forall\,T\in[0,T_{m}),\quad\frac{1}{T}\int_{0}^{T}\left\|\mu_{t}^{N}-\vartheta_{f_{t}}^{N}\right\|_{\operatorname{Lip}^{*}}\,{\rm d}t\lesssim\epsilon(N)\int_{0}^{T}R(s)\,{\rm d}s+\left\|\mu_{0}^{N}-\vartheta_{f_{0}}^{N}\right\|_{\operatorname{Lip}^{*}}.
Remark 1.

Note that μtNϑftNLip0\|\mu_{t}^{N}-\vartheta_{f_{t}}^{N}\|_{\operatorname{Lip}^{*}}\to 0 as NN\to\infty implies that the empirical measure (1.1) sampled from the law μtN\mu^{N}_{t} satisfies

(1.6) ϕCb(𝔾),ϵ>0,t0,limNμtN({|αηN,φft,φ|>ϵ})=0\forall\,\phi\in C_{b}(\mathbb{G}),\ \forall\,\epsilon>0,\ \forall\,t\geq 0,\quad\lim_{N\rightarrow\infty}\mu^{N}_{t}\left(\left\{|\langle\alpha^{N}_{\eta},\varphi\rangle-\langle f_{t},\varphi\rangle|>\epsilon\right\}\right)=0

with a rate of convergence (thus recovering quantitatively results from [GPV88]):

μtN({|αηN,φft,φ|>ϵ})\displaystyle\mu^{N}_{t}\left(\left\{|\langle\alpha^{N}_{\eta},\varphi\rangle-\langle f_{t},\varphi\rangle|>\epsilon\right\}\right)
μtN({αηN,φft,φ+ϵ})+μtN({αηN,φft,φϵ})\displaystyle\leq\mu^{N}_{t}\left(\left\{\langle\alpha^{N}_{\eta},\varphi\rangle\geq\langle f_{t},\varphi\rangle+\epsilon\right\}\right)+\mu^{N}_{t}\left(\left\{\langle\alpha^{N}_{\eta},\varphi\rangle\leq\langle f_{t},\varphi\rangle-\epsilon\right\}\right)
𝖷N[Fϵ+(ϕ,αηN)Fϵ+(ϕ,ft)]dμtN+𝖷N[Fϵ(ϕ,αηN)Fϵ(ϕ,ft)]dμtN\displaystyle\leq\int_{\mathsf{X}_{N}}\left[F_{\epsilon}^{+}\left(\left\langle\phi,\alpha_{\eta}^{N}\right\rangle\right)-F_{\epsilon}^{+}\left(\left\langle\phi,f_{t}\right\rangle\right)\right]\,{\rm d}\mu_{t}^{N}+\int_{\mathsf{X}_{N}}\left[F_{\epsilon}^{-}\left(\left\langle\phi,\alpha_{\eta}^{N}\right\rangle\right)-F^{-}_{\epsilon}\left(\left\langle\phi,f_{t}\right\rangle\right)\right]\,{\rm d}\mu_{t}^{N}

where Fϵ±F_{\epsilon}^{\pm} are mollified version of the characteristic functions of respectively {zϕ,ft+ϵ}\{z\geq\left\langle\phi,f_{t}\right\rangle+\epsilon\} and {zϕ,ftϵ}\{z\leq\left\langle\phi,f_{t}\right\rangle-\epsilon\}, which yields

supt[0,T]μtN({|αηN,φft,φ|>ϵ})ϵ1μtNϑftNLip+ϵ2Nd.\sup_{t\in[0,T]}\mu^{N}_{t}\left(\left\{|\langle\alpha^{N}_{\eta},\varphi\rangle-\langle f_{t},\varphi\rangle|>\epsilon\right\}\right)\lesssim\epsilon^{-1}\|\mu_{t}^{N}-\vartheta_{f_{t}}^{N}\|_{\operatorname{Lip}^{*}}+\epsilon^{-2}N^{-d}.

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 𝖷=\mathsf{X}=\mathbb{N}. Given the choice of a transition function pP(𝕋Nd)p\in P(\mathbb{T}^{d}_{N}) with p(0)=0p(0)=0 and a jump rate function g:+g:\mathbb{N}\rightarrow\mathbb{R}_{+}, the base generator ^N\hat{\mathcal{L}}_{N} writes

(2.1) ΦCb(𝖷N),η𝖷N,^NΦ(η):=x,y𝕋Ndp(yx)g(ηx)[Φ(ηxy)Φ(η)]\forall\,\Phi\in C_{b}(\mathsf{X}_{N}),\ \forall\,\eta\in\mathsf{X}_{N},\quad\hat{\mathcal{L}}_{N}\Phi(\eta):=\sum_{x,y\in\mathbb{T}^{d}_{N}}p(y-x)g(\eta_{x})\left[\Phi(\eta^{xy})-\Phi(\eta)\right]

where ηxy\eta^{xy} is defined as before. The local equilibrium structure of (H0) is given by

(2.2) nλ(k):=λkg(k)!Z(λ) with Z(λ):=k=0+λkg(k)!\displaystyle n_{\lambda}(k):=\frac{\lambda^{k}}{g(k)!Z(\lambda)}\quad\text{ with }\quad Z(\lambda):=\sum_{k=0}^{+\infty}\frac{\lambda^{k}}{g(k)!}
(2.3) σ is defined implicitely by σ(ρ)Z(σ(ρ))Z(σ(ρ))ρ\displaystyle\sigma\ \text{ is defined implicitely by }\ \sigma(\rho)\frac{Z^{\prime}(\sigma(\rho))}{Z(\sigma(\rho))}\equiv\rho

denoting g(k)!:=g(k)g(k1)g(1)g(k)!:=g(k)g(k-1)\cdots g(1). The pair (g,σ)(g,\sigma) thus constructed satisfies 𝔼nσ(α)[g]=σ(α)\mathbb{E}_{n_{\sigma(\alpha)}}[g]=\sigma(\alpha). When fρ[0,+)f\equiv\rho\in[0,+\infty) is constant, the local Gibbs measure ϑρN=νσ(ρ)N\vartheta_{\rho}^{N}=\nu_{\sigma(\rho)}^{N} is invariant with average number of particles ρ\rho. The mean transition rate is defined by γ:=xdxp(x)d\gamma:=\sum_{x\in\mathbb{Z}^{d}}xp(x)\in\mathbb{R}^{d}. When γ0\gamma\not=0, the first non-zero asymptotic dynamics as NN\to\infty is given by the hyperbolic scaling N:=N^N\mathcal{L}_{N}:=N\hat{\mathcal{L}}_{N}, and the corresponding expected limit equation is tf=γ[σ(f)]\partial_{t}f=\gamma\cdot\nabla[\sigma(f)]. When γ=0\gamma=0, the first non-zero asymptotic dynamics as NN\to\infty is the given by the parabolic scaling N:=N2^N\mathcal{L}_{N}:=N^{2}\hat{\mathcal{L}}_{N}, and the corresponding limit equation is formally

(2.4) tf=Δa[σ(f)] with Δa:=i,j=1daijij2andaij:=xdp(x)xixj.\partial_{t}f=\Delta_{a}[\sigma(f)]\quad\text{ with }\quad\Delta_{a}:=\sum_{i,j=1}^{d}a_{ij}\partial^{2}_{ij}\quad\text{and}\quad a_{ij}:=\sum_{x\in\mathbb{Z}^{d}}p(x)x_{i}x_{j}.

We make the following assumptions on the jump rate function g:[0,)g:\mathbb{N}\rightarrow[0,\infty).

(HZRP) The jump rate gg satisfies g(0)=0g(0)=0, g(n)>0g(n)>0 for all n>0n>0, is non-decreasing, uniformly Lipschitz supn0|g(n+1)g(n)|<+\sup_{n\geq 0}|g(n+1)-g(n)|<+\infty, and there are n0>0n_{0}>0 and β>0\beta>0 such that g(n)g(n)βg(n^{\prime})-g(n)\geq\beta for any nn+n0n^{\prime}\geq n+n_{0}.

The main result on the ZRP is:

Theorem 2.1 (Hydrodynamic limit for the ZRP).

Consider ^N\hat{\mathcal{L}}_{N} defined in (2.1) with gg satisfying (HZRP). Let d=1d=1, f0C3(𝕋)f_{0}\in C^{3}(\mathbb{T}) with f0δ>0f_{0}\geq\delta>0, and μ0NP1(𝖷N)\mu_{0}^{N}\in P_{1}(\mathsf{X}_{N}) for all N1N\geq 1. Assume γ=0\gamma=0, define μtN=etN2^N\mu^{N}_{t}=e^{tN^{2}\hat{\mathcal{L}}_{N}} and ftC([0,T),C3(𝕋d))f_{t}\in C([0,T),C^{3}(\mathbb{T}^{d})) solution to (2.4), then the following convergence holds (with quantitative constants)

(2.5) supT01T0TμtNϑftNLipdtN18+μ0Nϑf0NLip.\sup_{T\geq 0}\frac{1}{T}\int_{0}^{T}\left\|\mu_{t}^{N}-\vartheta_{f_{t}}^{N}\right\|_{\operatorname{Lip}^{*}}\,{\rm d}t\lesssim N^{-\frac{1}{8}}+\left\|\mu_{0}^{N}-\vartheta_{f_{0}}^{N}\right\|_{\operatorname{Lip}^{*}}.

2.2. The GLK

In this case, the state space at each site is 𝖷=\mathsf{X}=\mathbb{R}. Given the choice of a single-site potential VC2()V\in C^{2}(\mathbb{R}), the base generator ^N\hat{\mathcal{L}}_{N} writes

(2.6) ^NΦ(η):=12xy𝕋Nd(ηxηy)212xy𝕋Nd[V(ηx)V(ηy)](ηxηy)\hat{\mathcal{L}}_{N}\Phi(\eta):=\frac{1}{2}\sum_{x\sim y\in\mathbb{T}^{d}_{N}}\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right)^{2}-\frac{1}{2}\sum_{x\sim y\in\mathbb{T}^{d}_{N}}\left[V^{\prime}(\eta_{x})-V^{\prime}(\eta_{y})\right]\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right)

where xyx\sim y denotes neighbouring sites. The local equilibrium structure is given by

nλ(r):=eλrV(r)Z(λ) with Z(λ):=eλrV(r)dr\displaystyle n_{\lambda}(r):=\frac{e^{\lambda r-V(r)}}{Z(\lambda)}\quad\text{ with }\quad Z(\lambda):=\int_{\mathbb{R}}e^{\lambda r-V(r)}\,{\rm d}r
σ is defined implicitely by Z(σ(ρ))Z(σ(ρ))ρ.\displaystyle\sigma\ \text{ is defined implicitely by }\ \frac{Z^{\prime}(\sigma(\rho))}{Z(\sigma(\rho))}\equiv\rho.

When fρf\equiv\rho\in\mathbb{R} is constant, the local Gibbs measure ϑρN=νσ(ρ)N\vartheta_{\rho}^{N}=\nu_{\sigma(\rho)}^{N} is invariant with average spin ρ\rho. The hyperbolic scaling formally leads to zero and the parabolic scaling N:=N2^N\mathcal{L}_{N}:=N^{2}\hat{\mathcal{L}}_{N} formally leads to

(2.7) tf=2Δ[σ(f)].\partial_{t}f=2\Delta[\sigma(f)].

We assume that the single-site potential satisfies

(HGLK) The potential VV is C2C^{2} and decomposes as V(u)=V0(u)+V1(u)V(u)=V_{0}(u)+V_{1}(u) with V0′′(u)κV_{0}^{\prime\prime}(u)\geq\kappa for all uu\in\mathbb{R} for some κ>0\kappa>0 and V1W1,()1\|V_{1}\|_{W^{1,\infty}(\mathbb{R})}\lesssim 1.

This assumption is similar with those in [GOVW09, DMOWa, Fat13]. One can take for example a double-well potential, provided it is uniformly convex at infinity.

Theorem 2.2 (Hydrodynamic limit for the GLK).

Consider N\mathcal{L}_{N} defined in (2.6) with VV satisfying (HGLK). Let d=1d=1, f0C3(𝕋d)f_{0}\in C^{3}(\mathbb{T}^{d}) and μ0NP1(𝖷N)\mu_{0}^{N}\in P_{1}(\mathsf{X}_{N}) for all N1N\geq 1. Define μtN=etN2^N\mu^{N}_{t}=e^{tN^{2}\hat{\mathcal{L}}_{N}} and ftC([0,+),C3(𝕋d))f_{t}\in C([0,+\infty),C^{3}(\mathbb{T}^{d})) the global solution to (2.7), then the following convergence holds (with quantitative constants)

(2.8) supT01T0TμtNϑftNLipdtN18+μ0Nϑf0NLip.\sup_{T\geq 0}\frac{1}{T}\int_{0}^{T}\left\|\mu_{t}^{N}-\vartheta_{f_{t}}^{N}\right\|_{\operatorname{Lip}^{*}}\,{\rm d}t\lesssim N^{-\frac{1}{8}}+\left\|\mu_{0}^{N}-\vartheta_{f_{0}}^{N}\right\|_{\operatorname{Lip}^{*}}.

3. The abstract strategy

In this section we sketch the proof of Theorem 1.1. Let ftf_{t} be a solution to (1.3). Given 0<<N0<\ell<N, we denote by η\eta^{\ell} for the local \ell-average ηx:=\stackon[2.6ex] |yx|ηy\eta^{\ell}_{x}:=\mathop{\mathchoice{\stackon[-3.8ex]{\displaystyle\sum}{\smash{\rule{0.4pt}{17.22217pt}}}}{\stackon[-2.6ex]{\textstyle\sum}{\smash{\rule{0.4pt}{12.48604pt}}}}{\stackon[-1.9ex]{\scriptstyle\sum}{\smash{\rule{0.4pt}{9.47217pt}}}}{\stackon[-1.4ex]{\scriptscriptstyle\sum}{\smash{\rule{0.4pt}{7.3194pt}}}}}_{|y-x|\leq\ell}\eta_{y}.

Denote by FtN:=dμtNdνNF^{N}_{t}:=\frac{{\rm d}\mu^{N}_{t}}{{\rm d}\nu^{N}_{\infty}} and GtN:=dϑftNdνNG^{N}_{t}:=\frac{{\rm d}\vartheta^{N}_{f_{t}}}{{\rm d}\nu^{N}_{\infty}} the densities with respect to νN\nu^{N}_{\infty}, and write

ddt(FtNGtN)=N(FtNGtN)+(NGtNtGtN)\displaystyle\frac{\,{\rm d}}{\,{\rm d}t}\Big{(}F^{N}_{t}-G_{t}^{N}\Big{)}=\mathcal{L}_{N}^{*}\Big{(}F^{N}_{t}-G_{t}^{N}\Big{)}+\left(\mathcal{L}_{N}^{*}G_{t}^{N}-\partial_{t}G_{t}^{N}\right)

so that Duhamel’s formula yields

FtNGtN=etN(F0NG0N)+0te(ts)N(NGsNsGsN)ds.\displaystyle F^{N}_{t}-G_{t}^{N}=e^{t\mathcal{L}_{N}^{*}}\left(F^{N}_{0}-G_{0}^{N}\right)+\int_{0}^{t}e^{(t-s)\mathcal{L}_{N}^{*}}\left(\mathcal{L}_{N}^{*}G_{s}^{N}-\partial_{s}G_{s}^{N}\right)\,{\rm d}s.

Take ΦLip(𝖷N)\Phi\in\operatorname{Lip}(\mathsf{X}_{N}) with ΦLip(𝖷N)1\|\Phi\|_{\operatorname{Lip}(\mathsf{X}_{N})}\leq 1 and integrate the above equation to get

𝖷NΦ(FtNGtN)dνN\displaystyle\int_{\mathsf{X}_{N}}\Phi\Big{(}F^{N}_{t}-G_{t}^{N}\Big{)}\,{\rm d}\nu^{N}_{\infty}
=𝖷N(etNΦ)(F0NG0N)dνNI1(t)+𝖷N0t(e(ts)NΦ)(NGsNsGsN)dνNdsI2(t).\displaystyle=\underbrace{\int_{\mathsf{X}_{N}}\left(e^{t\mathcal{L}_{N}}\Phi\right)\Big{(}F^{N}_{0}-G_{0}^{N}\Big{)}\,{\rm d}\nu^{N}_{\infty}}_{I_{1}(t)}+\underbrace{\int_{\mathsf{X}_{N}}\int_{0}^{t}\left(e^{(t-s)\mathcal{L}_{N}}\Phi\right)\left(\mathcal{L}_{N}G_{s}^{N}-\partial_{s}G_{s}^{N}\right)\,{\rm d}\nu^{N}_{\infty}\,{\rm d}s}_{I_{2}(t)}.

(H1) implies I1(t)μ0Nϑf0NLipI_{1}(t)\lesssim\|\mu^{N}_{0}-\vartheta_{f_{0}}^{N}\|_{\operatorname{Lip}^{*}} and (H3) implies 1T0TI2(t)dtϵ(N)0TR(s)ds\frac{1}{T}\int_{0}^{T}I_{2}(t)\,{\rm d}t\leq\epsilon(N)\int_{0}^{T}R(s)\,{\rm d}s, 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 N=N\mathcal{L}_{N}=\mathcal{L}_{N}^{*} is symmetric with respect to equilibrium measures.

Given ftC3(𝕋d)f_{t}\in C^{3}(\mathbb{T}^{d}) with f>δf>\delta, δ>0\delta>0, and ρ:=𝕋df\rho:=\int_{\mathbb{T}^{d}}f, the density of the local Gibbs measure relatively to the invariant measure with mass ρ\rho is:

(4.1) GtN(η):=dϑfN(η)dϑρN(η)=x𝕋Nd(σ(ft(xN))σ(ρ))η(x)(Z(σ(ft(xN)))Z(σ(ρ)))1.G_{t}^{N}(\eta):=\frac{{\rm d}\vartheta_{f}^{N}(\eta)}{{\rm d}\vartheta_{\rho}^{N}(\eta)}=\prod_{x\in\mathbb{T}^{d}_{N}}\left(\frac{\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)}{\sigma(\rho)}\right)^{\eta(x)}\left(\frac{Z(\sigma\left(f_{t}\left(\frac{x}{N}\right)\right))}{Z(\sigma(\rho))}\right)^{-1}.

where the function σ(r)\sigma(r) is defined by nσ(r),η(x)=r\langle n_{\sigma(r)},\eta(x)\rangle=r and the partition function Z:[0,λ)Z:[0,\lambda^{\ast})\rightarrow\mathbb{R} is defined in (2.2), with λ[0,+]\lambda^{*}\in[0,+\infty] denoting the radius of convergence of the series.

It is proved in [KL99, Chapter 2, Section 3] that assumption (HZRP) on gg implies that σ=R1:[0,)[0,)\sigma=R^{-1}:[0,\infty)\rightarrow[0,\infty) is well-defined and strictly increasing, with

R(λ)=λλlog(Z(λ))=1Z(λ)n0nλng(n)!.R(\lambda)=\lambda\partial_{\lambda}\log(Z(\lambda))=\frac{1}{Z(\lambda)}\sum_{n\geq 0}\frac{n\lambda^{n}}{g(n)!}.

Then the building block nρn_{\rho} of the Gibbs measure satisfies nσ(ρ),g(η(x))=σ(ρ)\langle n_{\sigma(\rho)},g(\eta(x))\rangle=\sigma(\rho). Moreover (HZRP) implies that the function σ\sigma is CC^{\infty} with uniform bound on all derivatives on +\mathbb{R}_{+}, with Lipschitz constant less than gg^{\ast}, and with infλ>0λ1σ(λ)>0\inf_{\lambda>0}\lambda^{-1}\sigma(\lambda)>0 (in particular σ(0)>0\sigma^{\prime}(0)>0), and that the invariant measure has exponential moment bounds, see [KL99, Corollary 3.6].

4.1. Microscopic Stability – (H1)

We use the coupling of [Lig85, Rez91]. Let

(4.2) ~NΨ(η,ζ):=x,y𝕋Ndp(yx)(g(ηx)g(ζx))[Ψ(ηxy,ζxy)Ψ(η,ζ)]+x,y𝕋Ndp(yx)(g(ηx)g(ηx)g(ζx))[Ψ(ηxy,ζ)Ψ(η,ζ)]+x,y𝕋Ndp(yx)(g(ζx)g(ηx)g(ζx))[Ψ(η,ζxy)Ψ(η,ζ)].\begin{split}\widetilde{\mathcal{L}}_{N}\Psi(\eta,\zeta):=&\sum_{x,y\in\mathbb{T}^{d}_{N}}p(y-x)\Big{(}g(\eta_{x})\wedge g(\zeta_{x})\Big{)}\Big{[}\Psi(\eta^{xy},\zeta^{xy})-\Psi(\eta,\zeta)\Big{]}\\ &+\sum_{x,y\in\mathbb{T}^{d}_{N}}p(y-x)\Big{(}g(\eta_{x})-g(\eta_{x})\wedge g(\zeta_{x})\Big{)}\Big{[}\Psi(\eta^{xy},\zeta)-\Psi(\eta,\zeta)\Big{]}\\ &+\sum_{x,y\in\mathbb{T}^{d}_{N}}p(y-x)\Big{(}g(\zeta_{x})-g(\eta_{x})\wedge g(\zeta_{x})\Big{)}\Big{[}\Psi(\eta,\zeta^{xy})-\Psi(\eta,\zeta)\Big{]}.\end{split}

for a two-variable test function Ψ(η,ζ)\Psi(\eta,\zeta). Then ~NΦ(η)=^NΦ(η)\widetilde{\mathcal{L}}_{N}\Phi(\eta)=\hat{\mathcal{L}}_{N}\Phi(\eta) and ~NΦ(ζ)=^NΦ(ζ)\widetilde{\mathcal{L}}_{N}\Phi(\zeta)=\hat{\mathcal{L}}_{N}\Phi(\zeta), and (H1) follows from the fact that et~Ne^{t\widetilde{\mathcal{L}}_{N}} preserves sign and the inequality

~N(z𝕋Nd|ηzζz|)0.\widetilde{\mathcal{L}}_{N}\left(\sum_{z\in\mathbb{T}^{d}_{N}}|\eta_{z}-\zeta_{z}|\right)\leq 0.

To prove the latter inequality, we compute

~N(z𝕋Nd|ηzζz|)\displaystyle\widetilde{\mathcal{L}}_{N}\left(\sum_{z\in\mathbb{T}^{d}_{N}}|\eta_{z}-\zeta_{z}|\right) =x,y𝕋Ndp(yx)(g(ηx)g(ηx)g(ζx))\displaystyle=\sum_{x,y\in\mathbb{T}^{d}_{N}}p(y-x)\Big{(}g(\eta_{x})-g(\eta_{x})\wedge g(\zeta_{x})\Big{)}
×[|ηxxyζx|+|ηyxyζy||ηxζx||ηyζy|]\displaystyle\hskip 56.9055pt\times\Big{[}|\eta^{xy}_{x}-\zeta_{x}|+|\eta^{xy}_{y}-\zeta_{y}|-|\eta_{x}-\zeta_{x}|-|\eta_{y}-\zeta_{y}|\Big{]}
+x,y𝕋Ndp(yx)(g(ζx)g(ηx)g(ζx))\displaystyle\quad+\sum_{x,y\in\mathbb{T}^{d}_{N}}p(y-x)\Big{(}g(\zeta_{x})-g(\eta_{x})\wedge g(\zeta_{x})\Big{)}
×[|ηxζxxy|+|ηyζyxy||ηxζx||ηyζy|].\displaystyle\hskip 56.9055pt\times\Big{[}|\eta_{x}-\zeta^{xy}_{x}|+|\eta_{y}-\zeta^{xy}_{y}|-|\eta_{x}-\zeta_{x}|-|\eta_{y}-\zeta_{y}|\Big{]}.

When g(ηx)g(ηx)g(ζx)>0g(\eta_{x})-g(\eta_{x})\wedge g(\zeta_{x})>0 necessarily ηxζy1\eta_{x}-\zeta_{y}\geq 1 and

[|ηxxyζx|+|ηyxyζy||ηxζx||ηyζy|]0.\Big{[}|\eta^{xy}_{x}-\zeta_{x}|+|\eta^{xy}_{y}-\zeta_{y}|-|\eta_{x}-\zeta_{x}|-|\eta_{y}-\zeta_{y}|\Big{]}\leq 0.

When g(ζx)g(ηx)g(ζx)>0g(\zeta_{x})-g(\eta_{x})\wedge g(\zeta_{x})>0 necessarily ζxηx1\zeta_{x}-\eta_{x}\geq 1 and

[|ηxζxxy|+|ηyζyxy||ηxζx||ηyζy|]0.\Big{[}|\eta_{x}-\zeta^{xy}_{x}|+|\eta_{y}-\zeta^{xy}_{y}|-|\eta_{x}-\zeta_{x}|-|\eta_{y}-\zeta_{y}|\Big{]}\leq 0.

4.2. Macroscopic stability – (H2)

In the parabolic scaling the limit PDE is the nonlinear diffusion equation (2.4). We take =C3\mathcal{B}=C^{3} with its standard infinity Banach norm. The proof that this norm remains uniformly bounded in time is classical in dimension d=1d=1 (using the bounds on σ\sigma), and ft[δ,1δ]f_{t}\in[\delta,1-\delta] for all times by maximal principle. Moreover ftρf_{t}\to\rho exponentially fast as tt\to\infty in \mathcal{B}.

4.3. Consistency estimate – (H3)

Let γ=0\gamma=0 and the dimension d=1d=1.

Proposition 4.1.

Given the solution ftC3(𝕋d)f_{t}\in C^{3}(\mathbb{T}^{d}) to (2.4) with fδf\geq\delta, δ>0\delta>0, and ρ:=𝕋df\rho:=\int_{\mathbb{T}^{d}}f, and GtNG_{t}^{N} defined in (4.1), we have for every ΦLip(𝖷N)\Phi\in\operatorname{Lip}(\mathsf{X}_{N}) with [Φ]Lip(𝖷N)1[\Phi]_{\operatorname{Lip}(\mathsf{X}_{N})}\leq 1

1T0TItNdt:=1T0T0t(e(ts)NΦ),[NGsNddsGsN]dνNdsdt=𝒪(N18)\displaystyle\frac{1}{T}\int_{0}^{T}I_{t}^{N}\,{\rm d}t:=\frac{1}{T}\int_{0}^{T}\int_{0}^{t}\left\langle\left(e^{(t-s)\mathcal{L}_{N}}\Phi\right),\left[\mathcal{L}_{N}G^{N}_{s}-\frac{{\rm d}}{{\rm d}s}G^{N}_{s}\right]{\rm d}\nu^{N}_{\infty}\right\rangle\,{\rm d}s\,{\rm d}t=\mathcal{O}\left(N^{-\frac{1}{8}}\right)

where the constant depends on the estimates in (H2).

Proof.

We start by computing

NGsNddsGsN=x𝕋NdAxNGsN\mathcal{L}_{N}G^{N}_{s}-\frac{{\rm d}}{{\rm d}s}G^{N}_{s}=\sum_{x\in\mathbb{T}^{d}_{N}}A_{x}^{N}G^{N}_{s}

with (note that ftρf_{t}\to\rho exponentially fast)

AxN\displaystyle A_{x}^{N} :=N2y𝕋Ndp(yx)g(ηx)(σ(ft(yN))σ(ft(xN))1)ηxσ(ft(xN))σ(ft(xN))Δa[σ(f)](xN)\displaystyle:=N^{2}\sum_{y\in\mathbb{T}^{d}_{N}}p(y-x)g(\eta_{x})\left(\frac{\sigma\left(f_{t}\left(\frac{y}{N}\right)\right)}{\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)}-1\right)-\eta_{x}\frac{\sigma^{\prime}\left(f_{t}\left(\frac{x}{N}\right)\right)}{\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)}\Delta_{a}[\sigma(f)]\left(\frac{x}{N}\right)
=g(ηx)σ(ft(xN))Δa[σ(f)](xN)ηxσ(ft(xN))σ(ft(xN))Δa[σ(f)](xN)+𝒪(eCsN)\displaystyle=\frac{g(\eta_{x})}{\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)}\Delta_{a}[\sigma(f)]\left(\frac{x}{N}\right)-\eta_{x}\frac{\sigma^{\prime}\left(f_{t}\left(\frac{x}{N}\right)\right)}{\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)}\Delta_{a}[\sigma(f)]\left(\frac{x}{N}\right)+\mathcal{O}\left(\frac{e^{-Cs}}{N}\right)

for some C>0C>0. Since (conservation of mass)

𝖷N(x𝕋NdAxNGsN)dνN=𝖷N(x𝕋NdAxN)dϑfsN=0,\int_{\mathsf{X}_{N}}\left(\sum_{x\in\mathbb{T}^{d}_{N}}A_{x}^{N}G^{N}_{s}\right)\,{\rm d}\nu_{\infty}^{N}=\int_{\mathsf{X}_{N}}\left(\sum_{x\in\mathbb{T}^{d}_{N}}A_{x}^{N}\right)\,{\rm d}\vartheta_{f_{s}}^{N}=0,

we can replace Φts:=e(ts)NΦ\Phi_{t-s}:=e^{(t-s)\mathcal{L}_{N}}\Phi by

Φ~t,s:=e(ts)NΦ𝐄ϑfsN[e(ts)NΦ]\tilde{\Phi}_{t,s}:=e^{(t-s)\mathcal{L}_{N}}\Phi-{\bf E}_{\vartheta_{f_{s}}^{N}}[e^{(t-s)\mathcal{L}_{N}}\Phi]

and use the Lipschitz bound on e(ts)NΦe^{(t-s)\mathcal{L}_{N}}\Phi (microscopic stability) to get

ItN=0t𝖷NΦ~t,s(η)(x𝕋NdA~xN)dϑfsN+𝒪(1N)I_{t}^{N}=\int_{0}^{t}\int_{\mathsf{X}_{N}}\tilde{\Phi}_{t,s}(\eta)\left(\sum_{x\in\mathbb{T}^{d}_{N}}\tilde{A}^{N}_{x}\right)\,{\rm d}\vartheta_{f_{s}}^{N}+\mathcal{O}\left(\frac{1}{N}\right)

with A~xN\tilde{A}^{N}_{x} defined by (note that it has zero average against dϑfsN{\rm d}\vartheta_{f_{s}}^{N})

A~xN:={g(ηx)σ(ft(xN))σ(ft(xN))[ηxft(xN)]}Δa[σ(f)](xN)σ(ft(xN)).\tilde{A}^{N}_{x}:=\left\{g(\eta_{x})-\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)-\sigma^{\prime}\left(f_{t}\left(\frac{x}{N}\right)\right)\left[\eta_{x}-f_{t}\left(\frac{x}{N}\right)\right]\right\}\frac{\Delta_{a}[\sigma(f)]\left(\frac{x}{N}\right)}{\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)}.

We then form sub-sum over non-overlapping cubes of size {1,,N}\ell\in\{1,\dots,N\} (this intermediate scale factor \ell will be chosen later in terms of NN). Let Nd𝕋Nd\mathcal{R}^{d}_{N}\subset\mathbb{T}^{d}_{N} be a net of centers of non-overlapping cubes of the form 𝒞x:={y𝕋Nd:xy}\mathcal{C}_{x}:=\{y\in\mathbb{T}^{d}_{N}\ :\ \|x-y\|_{\infty}\leq\ell\}. Then

ItN\displaystyle I_{t}^{N} =xNd0t𝖷NΦ~t,s(η)(y𝒞xA~yN)dϑfsN+𝒪(1N)\displaystyle=\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\tilde{\Phi}_{t,s}(\eta)\left(\sum_{y\in\mathcal{C}_{x}}\tilde{A}^{N}_{y}\right)\,{\rm d}\vartheta_{f_{s}}^{N}+\mathcal{O}\left(\frac{1}{N}\right)
=(2+1)dxNd0t𝖷NΦ~t,s(η)A^xNdϑfsN+𝒪(1N)\displaystyle=(2\ell+1)^{d}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\tilde{\Phi}_{t,s}(\eta)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}+\mathcal{O}\left(\frac{1}{N}\right)

with the A^xN\hat{A}^{N}_{x} defined by

A^xN:={g(η)𝒞xσ(ft(xN))σ(ft(xN))[η𝒞xft(xN)]}Δa[σ(f)](xN)σ(ft(xN))\hat{A}^{N}_{x}:=\left\{\langle g(\eta)\rangle_{\mathcal{C}_{x}}-\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)-\sigma^{\prime}\left(f_{t}\left(\frac{x}{N}\right)\right)\left[\langle\eta\rangle_{\mathcal{C}_{x}}-f_{t}\left(\frac{x}{N}\right)\right]\right\}\frac{\Delta_{a}[\sigma(f)]\left(\frac{x}{N}\right)}{\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)}

where F(η)𝒞x\langle F(\eta)\rangle_{\mathcal{C}_{x}}, for F=F(ηx)F=F(\eta_{x}), denotes taking the average over the cube 𝒞x\mathcal{C}_{x}. Note that the average of A^xN\hat{A}^{N}_{x} against dϑfsN{\rm d}\vartheta_{f_{s}}^{N} is 𝒪(eCs/N)\mathcal{O}(e^{-Cs}\ell/N). Then

xNd0t𝖷NΦ~t,sA^xNdϑfsN\displaystyle\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\tilde{\Phi}_{t,s}\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}
=xNd0t𝖷N(Φ~t,sΠxNΦ~t,s)A^xNdϑfsN+xNd0t𝖷NΠxNΦ~t,sA^xNdϑfsN\displaystyle=\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left(\tilde{\Phi}_{t,s}-\Pi_{x}^{N}\tilde{\Phi}_{t,s}\right)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}+\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\Pi_{x}^{N}\tilde{\Phi}_{t,s}\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}
=xNd0t𝖷N(ΦtsΠxNΦts)A^xNdϑfsN\displaystyle=\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left(\Phi_{t-s}-\Pi_{x}^{N}\Phi_{t-s}\right)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}
+xNd0t𝖷N(ΠxNΦts𝐄ϑfsN[ΠxNΦts])A^xNdϑfsN=:JtN+J~tN\displaystyle\hskip 71.13188pt+\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left(\Pi_{x}^{N}\Phi_{t-s}-{\bf E}_{\vartheta_{f_{s}}^{N}}[\Pi_{x}^{N}\Phi_{t-s}]\right)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}=:J^{N}_{t}+\tilde{J}^{N}_{t}

where ΠxN\Pi_{x}^{N} projects on the average over configurations Ωm:={η~:η~𝒞x=m}\Omega_{m}:=\left\{\tilde{\eta}\,:\,\langle\tilde{\eta}\rangle_{\mathcal{C}_{x}}=m\right\} with the same mass in the cube 𝒞x\mathcal{C}_{x} (and does not touch the other sites):

(4.3) ΠxNφ(η)=[ΠxNφ](η𝒞x)=Ωη𝒞xφ(η~)dν,η𝒞x(η~)\Pi_{x}^{N}\varphi(\eta)=[\Pi_{x}^{N}\varphi](\langle\eta\rangle_{\mathcal{C}_{x}})=\int_{\Omega_{\langle\eta\rangle_{\mathcal{C}_{x}}}}\varphi(\tilde{\eta})\,{\rm d}\nu^{\ell,\langle\eta\rangle_{\mathcal{C}_{x}}}(\tilde{\eta})

for a function φ\varphi on 𝖷𝒞x\mathsf{X}^{\mathcal{C}_{x}}. To estimate the first term JtNJ^{N}_{t} we first approximate the measure ϑfsN\vartheta_{f_{s}}^{N} on 𝒞x\mathcal{C}_{x} by the equilibrium measure with local mass ft(x/N)f_{t}(x/N), and denote it by ϑ¯fs\bar{\vartheta}_{f_{s}} (note that the approximation is made differently for each cube and depends on xx, even if it is written explicitly). This produces an error 𝒪(d+1/N)\mathcal{O}(\ell^{d+1}/N) (using the Lipschitz regularity of Φts\Phi_{t-s} and the exponential convergence ftρf_{t}\to\rho to get uniform in time bounds). We then apply the Poincaré inequality [LSV96, Theorem 1.1] in the cube 𝒞x\mathcal{C}_{x} (whose constant is independent of the number of particles and proportional to the size of the cube) and the law of large number A^xNL2(ϑ¯fsN)=𝒪(eCsd/2)\|\hat{A}^{N}_{x}\|_{L^{2}(\bar{\vartheta}^{N}_{f_{s}})}=\mathcal{O}(e^{-Cs}\ell^{-d/2}) (using uniform bounds on the second moment of ϑ¯fs\bar{\vartheta}_{f_{s}}):

JtN\displaystyle J^{N}_{t} xNd0tΦtsΠxNΦtsL2(ϑ¯fsN)A^xNL2(ϑ¯fsN)ds+𝒪(d+1N)\displaystyle\leq\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\|\Phi_{t-s}-\Pi_{x}^{N}\Phi_{t-s}\|_{L^{2}(\bar{\vartheta}_{f_{s}}^{N})}\|\hat{A}^{N}_{x}\|_{L^{2}(\bar{\vartheta}_{f_{s}}^{N})}\,{\rm d}s+\mathcal{O}\left(\frac{\ell^{d+1}}{N}\right)
1d2xNd0tD¯x(Φts)eCsds+𝒪(d+1N)\displaystyle\lesssim\ell^{1-\frac{d}{2}}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\sqrt{\bar{D}^{\ell}_{x}\left(\Phi_{t-s}\right)}e^{-Cs}\,{\rm d}s+\mathcal{O}\left(\frac{\ell^{d+1}}{N}\right)
1d2Nd20t(xNdD¯x(Φts))12eCsds+𝒪(d+1N)\displaystyle\lesssim\ell^{1-\frac{d}{2}}N^{\frac{d}{2}}\int_{0}^{t}\left(\sum_{x\in\mathcal{R}_{N}^{d}}\bar{D}^{\ell}_{x}\left(\Phi_{t-s}\right)\right)^{\frac{1}{2}}e^{-Cs}\,{\rm d}s+\mathcal{O}\left(\frac{\ell^{d+1}}{N}\right)

where D¯x(Φ)\bar{D}^{\ell}_{x}(\Phi) is the Dirichlet form on the cube 𝒞x\mathcal{C}_{x} with respect to the measure ϑ¯fsN\bar{\vartheta}_{f_{s}}^{N}:

D¯x(Φ):=y,z𝒞x𝖷Np(zy)g(ηy)[Φ(ηyz)Φ(η)]2dϑ¯fsN.\bar{D}^{\ell}_{x}(\Phi):=\sum_{y,z\in\mathcal{C}_{x}}\int_{\mathsf{X}_{N}}p(z-y)g(\eta_{y})\left[\Phi(\eta^{yz})-\Phi(\eta)\right]^{2}\,{\rm d}\bar{\vartheta}_{f_{s}}^{N}.

Then we change back the measure ϑ¯fsN\bar{\vartheta}_{f_{s}}^{N} in each box, which produces (using the Lipschitz regularity of Φ\Phi) an error 3/2N1/2\ell^{3/2}N^{-1/2}), and we compute

12N2ddt𝖷NΦts(η)2dϑfsNxNdDx(Φts)+𝒪(1N2)\frac{1}{2N^{2}}\frac{{\rm d}}{{\rm d}t}\int_{\mathsf{X}_{N}}\Phi_{t-s}(\eta)^{2}\,{\rm d}\vartheta_{f_{s}}^{N}\leq-\sum_{x\in\mathcal{R}_{N}^{d}}D^{\ell}_{x}\left(\Phi_{t-s}\right)+\mathcal{O}\left(\frac{1}{N^{2}}\right)

(with DD^{\ell} denoting the Dirichlet form for ϑfsN\vartheta_{f_{s}}^{N}), where the last error accounts for the small default of self-adjointness. We deduce (in dimension d=1d=1) that

0TJtNdtT12(N)1d2+𝒪(Td+1N).\int_{0}^{T}J_{t}^{N}\,{\rm d}t\lesssim T^{\frac{1}{2}}\left(\frac{\ell}{N}\right)^{1-\frac{d}{2}}+\mathcal{O}\left(\frac{T\ell^{d+1}}{N}\right).

To control the second term J~tN\tilde{J}_{t}^{N}, we first use the equivalence of ensemble in [KL99, Appendix II, Corollary 1.7] on the local equilibrium measure ϑ¯fsN\bar{\vartheta}_{f_{s}}^{N} (using uniform exponential moment bounds):

(4.4) g(η)𝒞x=σ(η𝒞x)+𝒪(1d).\langle g(\eta)\rangle_{\mathcal{C}_{x}}=\sigma\left(\langle\eta\rangle_{\mathcal{C}_{x}}\right)+\mathcal{O}\left(\frac{1}{\ell^{d}}\right).

Second we remark that the Lipschitz regularity of Φts\Phi_{t-s} implies that ΠxNΦts𝐄ϑfsN[ΠxNΦts]=𝒪(dNd)\Pi_{x}^{N}\Phi_{t-s}-{\bf E}_{\vartheta_{f_{s}}^{N}}[\Pi_{x}^{N}\Phi_{t-s}]=\mathcal{O}(\ell^{d}N^{-d}), and since the average of A^xN\hat{A}_{x}^{N} with respect to ϑfsN\vartheta_{f_{s}}^{N} is 𝒪(/N)\mathcal{O}(\ell/N), we can write

J~tN=xNd0t𝖷N(ΠxNΦts[η𝒞x]ΠxNΦts[fs(xN)])A^xNdϑfsN+𝒪(N).\tilde{J}_{t}^{N}=\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left(\Pi_{x}^{N}\Phi_{t-s}[\langle\eta\rangle_{\mathcal{C}_{x}}]-\Pi_{x}^{N}\Phi_{t-s}\left[f_{s}\left(\frac{x}{N}\right)\right]\right)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}+\mathcal{O}\left(\frac{\ell}{N}\right).

Third, we remark that the Lipschitz regularity of Φts\Phi_{t-s} (with constant NdN^{-d}) implies a Lipschitz regularity of its averaged projection ΠxNΦts\Pi_{x}^{N}\Phi_{t-s} with constant dNd\ell^{d}N^{-d}, with respect to the local mass. Indeed, given 0=mm<+0=m\leq m^{\prime}<+\infty, pick any pair of configuration (η0,ζ0)(\eta_{0},\zeta_{0}) with η0𝒞x=m\langle\eta_{0}\rangle_{\mathcal{C}_{x}}=m, ζ0𝒞x=m\langle\zeta_{0}\rangle_{\mathcal{C}_{x}}=m^{\prime} and η0ζ0\eta_{0}\leq\zeta_{0} (such configuration trivially exists since mmm\leq m^{\prime}). Then we consider the initial coupling δ(η0,ζ0)\delta_{(\eta_{0},\zeta_{0})} on Ωm×Ωm\Omega_{m}\times\Omega_{m^{\prime}} which has 1\ell^{1} cost mmm^{\prime}-m. Then we evolve it along the flow of the coupling operator et~Ne^{t\tilde{\mathcal{L}}_{N}}. The marginals respectively converge to ν,m\nu^{\ell,m} and ν,m\nu^{\ell,m^{\prime}} (convergence to equilibrium of the oiriginal evolution). Since the evolution by the coupling operator does not increase the Wasserstein distance, we deduce W1(ν,m,ν,m)mmW_{1}(\nu^{\ell,m},\nu^{\ell,m^{\prime}})\leq m^{\prime}-m. An optimal coupling Π\Pi associated to this distance thus satisfies

mmΩm×Ωm(\stackon[3.8ex] x𝕋Nd|ηxζx|)Π(η,ζ)mmm^{\prime}-m\leq\int_{\Omega_{m}\times\Omega_{m^{\prime}}}\left(\mathop{\mathchoice{\stackon[-3.8ex]{\displaystyle\sum}{\smash{\rule{0.4pt}{17.22217pt}}}}{\stackon[-2.6ex]{\textstyle\sum}{\smash{\rule{0.4pt}{12.48604pt}}}}{\stackon[-1.9ex]{\scriptstyle\sum}{\smash{\rule{0.4pt}{9.47217pt}}}}{\stackon[-1.4ex]{\scriptscriptstyle\sum}{\smash{\rule{0.4pt}{7.3194pt}}}}}_{x\in\mathbb{T}^{d}_{N}}|\eta_{x}-\zeta_{x}|\right)\Pi(\eta,\zeta)\leq m^{\prime}-m

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 Π\Pi, i.e. ηζ\eta\leq\zeta in the support. We then compute

ΠxNΦts(m)ΠxNΦts(m)\displaystyle\Pi_{x}^{N}\Phi_{t-s}(m^{\prime})-\Pi_{x}^{N}\Phi_{t-s}(m) =ΩmΦts(ζ)dν,m(ζ)ΩmΦts(η)dν,m(η)\displaystyle=\int_{\Omega_{m^{\prime}}}\Phi_{t-s}(\zeta)\,{\rm d}\nu^{\ell,m^{\prime}}(\zeta)-\int_{\Omega_{m}}\Phi_{t-s}(\eta)\,{\rm d}\nu^{\ell,m}(\eta)
=Ωm×Ωm[Φts(ζ)Φ(η)]dΠ(η,ζ)\displaystyle=\int_{\Omega_{m}\times\Omega_{m^{\prime}}}\left[\Phi_{t-s}(\zeta)-\Phi(\eta)\right]\,{\rm d}\Pi(\eta,\zeta)

and since ηζ\eta\leq\zeta on the support of Π\Pi, ζη1(𝒞x)=(mm)d\|\zeta-\eta\|_{\ell^{1}(\mathcal{C}_{x})}=(m^{\prime}-m)\ell^{d} and

|ΠxNΦts(m)ΠxNΦts(m)|dNd|mm|.\left|\Pi_{x}^{N}\Phi_{t-s}(m^{\prime})-\Pi_{x}^{N}\Phi_{t-s}(m)\right|\leq\frac{\ell^{d}}{N^{d}}|m^{\prime}-m|.

We deduce (using (5.3))

J~tNdNdxNd0t𝖷N|η𝒞xfs(xN)|×|σ(η𝒞x)σ(fs(xN))σ(fs(xN))[η𝒞xfs(xN)]|dϑfsNeCsds+1NdxNd0t𝖷N|η𝒞xfs(xN)|dϑfsNeCsds+𝒪(N)\tilde{J}_{t}^{N}\lesssim\frac{\ell^{d}}{N^{d}}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left|\langle\eta\rangle_{\mathcal{C}_{x}}-f_{s}\left(\frac{x}{N}\right)\right|\times\\ \left|\sigma(\langle\eta\rangle_{\mathcal{C}_{x}})-\sigma\left(f_{s}\left(\frac{x}{N}\right)\right)-\sigma^{\prime}\left(f_{s}\left(\frac{x}{N}\right)\right)\left[\langle\eta\rangle_{\mathcal{C}_{x}}-f_{s}\left(\frac{x}{N}\right)\right]\right|\,{\rm d}\vartheta_{f_{s}}^{N}e^{-Cs}\,{\rm d}s\\ +\frac{1}{N^{d}}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left|\langle\eta\rangle_{\mathcal{C}_{x}}-f_{s}\left(\frac{x}{N}\right)\right|\,{\rm d}\vartheta^{N}_{f_{s}}e^{-Cs}\,{\rm d}s+\mathcal{O}\left(\frac{\ell}{N}\right)

which yields by Taylor formula, the approximation of ϑfsN\vartheta^{N}_{f_{s}} by ϑ¯fsN\bar{\vartheta}^{N}_{f_{s}}, and the law of large numbers

J~tN\displaystyle\tilde{J}_{t}^{N} dNdxNd0t𝖷N|η𝒞xfs(xN)|3dϑfsNeCsds\displaystyle\lesssim\frac{\ell^{d}}{N^{d}}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left|\langle\eta\rangle_{\mathcal{C}_{x}}-f_{s}\left(\frac{x}{N}\right)\right|^{3}\,{\rm d}\vartheta_{f_{s}}^{N}e^{-Cs}\,{\rm d}s
+1NdxNd0t𝖷N|η𝒞xfs(xN)|dϑfsNeCsds+𝒪(N)eCsds\displaystyle\hskip 56.9055pt+\frac{1}{N^{d}}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left|\langle\eta\rangle_{\mathcal{C}_{x}}-f_{s}\left(\frac{x}{N}\right)\right|\,{\rm d}\vartheta^{N}_{f_{s}}e^{-Cs}\,{\rm d}s+\mathcal{O}\left(\frac{\ell}{N}\right)e^{-Cs}\,{\rm d}s
dNdxNd0t𝖷N|η𝒞xfs(xN)|3dϑ¯fsNeCsds\displaystyle\lesssim\frac{\ell^{d}}{N^{d}}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left|\langle\eta\rangle_{\mathcal{C}_{x}}-f_{s}\left(\frac{x}{N}\right)\right|^{3}\,{\rm d}\bar{\vartheta}_{f_{s}}^{N}e^{-Cs}\,{\rm d}s
+1NdxNd0t𝖷N|η𝒞xfs(xN)|dϑ¯fsNeCsds+𝒪(N)\displaystyle\hskip 56.9055pt+\frac{1}{N^{d}}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left|\langle\eta\rangle_{\mathcal{C}_{x}}-f_{s}\left(\frac{x}{N}\right)\right|\,{\rm d}\bar{\vartheta}^{N}_{f_{s}}e^{-Cs}\,{\rm d}s+\mathcal{O}\left(\frac{\ell}{N}\right)
𝒪(3d2)+𝒪(N).\displaystyle\lesssim\mathcal{O}\left(\ell^{-\frac{3d}{2}}\right)+\mathcal{O}\left(\frac{\ell}{N}\right).

Combining all estimates we get (optimizing :=N1/4\ell:=N^{1/4})

(4.5) 1T0TItNdt(1N+1+d2N1d2+1+2dN+1d2+N)1N18.\frac{1}{T}\int_{0}^{T}I_{t}^{N}\,{\rm d}t\lesssim\left(\frac{1}{N}+\frac{\ell^{1+\frac{d}{2}}}{N^{1-\frac{d}{2}}}+\frac{\ell^{1+2d}}{N}+\frac{1}{\ell^{\frac{d}{2}}}+\frac{\ell}{N}\right)\lesssim\frac{1}{N^{\frac{1}{8}}}.

5. Proof for the GLK

In this section we prove Theorem 2.2 (hydrodynamic limit for the GLK). Note again that for this model N=N\mathcal{L}_{N}=\mathcal{L}_{N}^{*} is symmetric with respect to equilibrium measures. Given ftC3(𝕋d)f_{t}\in C^{3}(\mathbb{T}^{d}) and ρ:=𝕋df\rho:=\int_{\mathbb{T}^{d}}f\in\mathbb{R}, the density of the local Gibbs measure relatively to the invariant measure with mass ρ\rho is:

(5.1) GtN(η):=dϑfN(η)dϑρN(η)=x𝕋Nde[σ(ft(xN))σ(ρ)]ηxZ(σ(ρ))Z(σ(ft(xN))).G_{t}^{N}(\eta):=\frac{{\rm d}\vartheta_{f}^{N}(\eta)}{{\rm d}\vartheta_{\rho}^{N}(\eta)}=\prod_{x\in\mathbb{T}^{d}_{N}}e^{\left[\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)-\sigma(\rho)\right]\eta_{x}}\frac{Z(\sigma(\rho))}{Z\left(\sigma\left(f_{t}\left(\frac{x}{N}\right)\right)\right)}.

where the function σ(r)\sigma(r) is defined by nσ(r),ηx=r\langle n_{\sigma(r)},\eta_{x}\rangle=r and the partition function Z(λ)=eλrV(r)drZ(\lambda)=\int_{\mathbb{R}}e^{\lambda r-V(r)}\,{\rm d}r is defined on \mathbb{R}. The uniform convexity of VV at infinity easily implies bounds on some exponential moments of the invariant measure and (HGLK) implies that there exists C>0C>0 so that 0<1CσC<0<\frac{1}{C}\leq\sigma^{\prime}\leq C<\infty (see [GOVW09, Lemma 41] and [DMOWa, Lemma 5.1]).

5.1. Microscopic stability – (H1)

We define a “coupling generator” ~N:Cb(𝖷N2)Cb(𝖷N2)\widetilde{\mathcal{L}}_{N}:C_{b}(\mathsf{X}_{N}^{2})\to C_{b}(\mathsf{X}_{N}^{2}) by

(5.2) ~NΨ(η,ζ):=xy([(ηxηy)(ηxηy)1]Ψ(η,ζ)+[1(ζxζy)(ζxζy)]Ψ(η,ζ)+K(ηxηy)(ζxζy)Ψ(η,ζ))\begin{split}\widetilde{\mathcal{L}}_{N}\Psi(\eta,\zeta):=\sum_{x\sim y}\bigg{(}&\left[\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right)^{*}\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right)\otimes 1\right]\Psi(\eta,\zeta)\\ &+\left[1\otimes\left(\frac{\partial}{\partial\zeta_{x}}-\frac{\partial}{\partial\zeta_{y}}\right)^{*}\left(\frac{\partial}{\partial\zeta_{x}}-\frac{\partial}{\partial\zeta_{y}}\right)\right]\Psi(\eta,\zeta)\\ &+K\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right)\otimes\left(\frac{\partial}{\partial\zeta_{x}}-\frac{\partial}{\partial\zeta_{y}}\right)\Psi(\eta,\zeta)\bigg{)}\end{split}

where K>0K>0 is a constant to be chosen later and the adjoint is taken in L2(dϑρN)L^{2}({\rm d}\vartheta^{N}_{\rho}) so

^N\displaystyle\hat{\mathcal{L}}_{N} =xy(ηxηy)2(V(ηx)V(ηy))(ηxηy)\displaystyle=\sum_{x\sim y}\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right)^{2}-\left(V^{\prime}(\eta_{x})-V^{\prime}(\eta_{y})\right)\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right)
=xy(ηxηy)(ηxηy).\displaystyle=\sum_{x\sim y}\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right)^{*}\left(\frac{\partial}{\partial\eta_{x}}-\frac{\partial}{\partial\eta_{y}}\right).

Then for any p(1,2]p\in(1,2] there is K=K(p)>0K=K(p)>0 (depending on pp) so that

~N(x𝕋Nd|ηxζx|p)\displaystyle\widetilde{\mathcal{L}}_{N}\left(\sum_{x\in\mathbb{T}^{d}_{N}}|\eta_{x}-\zeta_{x}|^{p}\right) =2p(p1)(2+4d)x𝕋Nd|ηxζx|p2\displaystyle=2p(p-1)(2+4d)\sum_{x\in\mathbb{T}^{d}_{N}}|\eta_{x}-\zeta_{x}|^{p-2}
2(p1)xy[V0(ηx)V0(ζx)](ηxζx)|ηxζx|p1\displaystyle\qquad-2(p-1)\sum_{x\sim y}\left[V_{0}^{\prime}(\eta_{x})-V_{0}^{\prime}(\zeta_{x})\right](\eta_{x}-\zeta_{x})|\eta_{x}-\zeta_{x}|^{p-1}
2(p1)xy[V1(ηx)V1(ζx)](ηxζx)|ηxζx|p1\displaystyle\qquad-2(p-1)\sum_{x\sim y}\left[V_{1}^{\prime}(\eta_{x})-V_{1}^{\prime}(\zeta_{x})\right](\eta_{x}-\zeta_{x})|\eta_{x}-\zeta_{x}|^{p-1}
+Kp(p1)(2+4d)x𝕋Nd|ηxζx|p20\displaystyle\qquad+Kp(p-1)(2+4d)\sum_{x\in\mathbb{T}^{d}_{N}}|\eta_{x}-\zeta_{x}|^{p-2}\leq 0

by using the assumptions on the potential: V0V_{0} uniformly strictly convex and V1W1,V_{1}\in W^{1,\infty}. This implies the weak contraction of the evolution in WpW_{p} (pp-Wasserstein distance) for any p(1,2]p\in(1,2], and thus by limit in W1W_{1}. 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 d=1d=1.

Proposition 5.1.

Given the solution ftC3(𝕋d)f_{t}\in C^{3}(\mathbb{T}^{d}) to (2.4), and ρ:=𝕋df\rho:=\int_{\mathbb{T}^{d}}f, and GtNG_{t}^{N} defined in (5.1), we have for every ΦLip(𝖷N)\Phi\in\operatorname{Lip}(\mathsf{X}_{N}) with [Φ]Lip(𝖷N)1[\Phi]_{\operatorname{Lip}(\mathsf{X}_{N})}\leq 1

1T0TItNdt:=1T0T0t(e(ts)NΦ),[NGsNddsGsN]dνNdsdt=𝒪(N18)\displaystyle\frac{1}{T}\int_{0}^{T}I_{t}^{N}\,{\rm d}t:=\frac{1}{T}\int_{0}^{T}\int_{0}^{t}\left\langle\left(e^{(t-s)\mathcal{L}_{N}}\Phi\right),\left[\mathcal{L}_{N}G^{N}_{s}-\frac{{\rm d}}{{\rm d}s}G^{N}_{s}\right]{\rm d}\nu^{N}_{\infty}\right\rangle\,{\rm d}s\,{\rm d}t=\mathcal{O}\left(N^{-\frac{1}{8}}\right)

where the constant depends on the estimates in (H2).

Proof.

The proof follows the same structure as for the ZRP. We start by computing

NGsNddsGsN=x𝕋NdAxNGsN\mathcal{L}_{N}G^{N}_{s}-\frac{{\rm d}}{{\rm d}s}G^{N}_{s}=\sum_{x\in\mathbb{T}^{d}_{N}}A_{x}^{N}G^{N}_{s}

with (note again that ftρf_{t}\to\rho exponentially fast)

AxN\displaystyle A_{x}^{N} =N22yx[2σ(fs(xN))(σ(fs(xN))σ(fs(yN)))\displaystyle=\frac{N^{2}}{2}\sum_{y\sim x}\Bigg{[}2\sigma\left(f_{s}\left(\frac{x}{N}\right)\right)\left(\sigma\left(f_{s}\left(\frac{x}{N}\right)\right)-\sigma\left(f_{s}\left(\frac{y}{N}\right)\right)\right)
2V(ηx)σ(fs(xN))σ(fs(yN))]\displaystyle\hskip 85.35826pt-2V^{\prime}(\eta_{x})\sigma\left(f_{s}\left(\frac{x}{N}\right)\right)-\sigma\left(f_{s}\left(\frac{y}{N}\right)\right)\Bigg{]}
x(ηxfs(xN))σ(fs(xN))Δ[σ(f)](xN)\displaystyle\hskip 142.26378pt-\sum_{x}\left(\eta_{x}-f_{s}\left(\frac{x}{N}\right)\right)\sigma^{\prime}\left(f_{s}\left(\frac{x}{N}\right)\right)\Delta[\sigma(f)]\left(\frac{x}{N}\right)
=Δ[σ(f)](xN)[V(ηx)σ(fs(xN))σ(fs(xN))(ηxfs(xN))]+𝒪(eCsN)\displaystyle=\Delta[\sigma(f)]\left(\frac{x}{N}\right)\Bigg{[}V^{\prime}(\eta_{x})-\sigma\left(f_{s}\left(\frac{x}{N}\right)\right)-\sigma^{\prime}\left(f_{s}\left(\frac{x}{N}\right)\right)\left(\eta_{x}-f_{s}\left(\frac{x}{N}\right)\right)\Bigg{]}+\mathcal{O}\left(\frac{e^{-Cs}}{N}\right)

for some C>0C>0. By conservation of mass we replace again Φts:=e(ts)NΦ\Phi_{t-s}:=e^{(t-s)\mathcal{L}_{N}}\Phi by

Φ~t,s:=e(ts)NΦ𝐄ϑfsN[e(ts)NΦ]\tilde{\Phi}_{t,s}:=e^{(t-s)\mathcal{L}_{N}}\Phi-{\bf E}_{\vartheta_{f_{s}}^{N}}[e^{(t-s)\mathcal{L}_{N}}\Phi]

and use the Lipschitz bound (H1) on e(ts)NΦe^{(t-s)\mathcal{L}_{N}}\Phi to get

ItN=0t𝖷NΦ~t,s(η)(x𝕋NdA~xN)dϑfsN+𝒪(1N)I_{t}^{N}=\int_{0}^{t}\int_{\mathsf{X}_{N}}\tilde{\Phi}_{t,s}(\eta)\left(\sum_{x\in\mathbb{T}^{d}_{N}}\tilde{A}^{N}_{x}\right)\,{\rm d}\vartheta_{f_{s}}^{N}+\mathcal{O}\left(\frac{1}{N}\right)

with A~xN\tilde{A}^{N}_{x} defined by (note that it has zero average against dϑfsN{\rm d}\vartheta_{f_{s}}^{N})

A~xN:=Δ[σ(f)](xN)[V(ηx)σ(f(xN))σ(fs(xN))(ηxf(xN))].\tilde{A}^{N}_{x}:=\Delta[\sigma(f)]\left(\frac{x}{N}\right)\left[V^{\prime}(\eta_{x})-\sigma\left(f\left(\frac{x}{N}\right)\right)-\sigma^{\prime}\left(f_{s}\left(\frac{x}{N}\right)\right)\left(\eta_{x}-f\left(\frac{x}{N}\right)\right)\right].

We again form sub-sum over non-overlapping cubes of size {1,,N}\ell\in\{1,\dots,N\}, with Nd𝕋Nd\mathcal{R}^{d}_{N}\subset\mathbb{T}^{d}_{N} a net of centers of cubes 𝒞x:={y𝕋Nd:xy}\mathcal{C}_{x}:=\{y\in\mathbb{T}^{d}_{N}\ :\ \|x-y\|_{\infty}\leq\ell\}. Then

ItN\displaystyle I_{t}^{N} =xNd0t𝖷NΦ~t,s(η)(y𝒞xA~yN)dϑfsN+𝒪(1N)\displaystyle=\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\tilde{\Phi}_{t,s}(\eta)\left(\sum_{y\in\mathcal{C}_{x}}\tilde{A}^{N}_{y}\right)\,{\rm d}\vartheta_{f_{s}}^{N}+\mathcal{O}\left(\frac{1}{N}\right)
=(2+1)dxNd0t𝖷NΦ~t,s(η)A^xNdϑfsN+𝒪(1N)\displaystyle=(2\ell+1)^{d}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\tilde{\Phi}_{t,s}(\eta)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}+\mathcal{O}\left(\frac{1}{N}\right)

with the A^xN\hat{A}^{N}_{x} defined by (and F(η)𝒞x\langle F(\eta)\rangle_{\mathcal{C}_{x}} again denotes the average over the cube 𝒞x\mathcal{C}_{x})

A^xN:=Δ[σ(f)](xN)[V(η)𝒞xσ(f(xN))σ(fs(xN))(η𝒞xf(xN))].\hat{A}^{N}_{x}:=\Delta[\sigma(f)]\left(\frac{x}{N}\right)\left[\langle V^{\prime}(\eta)\rangle_{\mathcal{C}_{x}}-\sigma\left(f\left(\frac{x}{N}\right)\right)-\sigma^{\prime}\left(f_{s}\left(\frac{x}{N}\right)\right)\left(\langle\eta\rangle_{\mathcal{C}_{x}}-f\left(\frac{x}{N}\right)\right)\right].

(Note again that the average of A^xN\hat{A}^{N}_{x} against dϑfsN{\rm d}\vartheta_{f_{s}}^{N} is 𝒪(eCs/N)\mathcal{O}(e^{-Cs}\ell/N).) Then

xNd0t𝖷NΦ~t,sA^xNdϑfsN\displaystyle\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\tilde{\Phi}_{t,s}\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}
=xNd0t𝖷N(Φ~t,sΠxNΦ~t,s)A^xNdϑfsN+xNd0t𝖷NΠxNΦ~t,sA^xNdϑfsN\displaystyle\qquad=\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left(\tilde{\Phi}_{t,s}-\Pi_{x}^{N}\tilde{\Phi}_{t,s}\right)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}+\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\Pi_{x}^{N}\tilde{\Phi}_{t,s}\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}
=xNd0t𝖷N(ΦtsΠxNΦts)A^xNdϑfsN\displaystyle\qquad=\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left(\Phi_{t-s}-\Pi_{x}^{N}\Phi_{t-s}\right)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}
+xNd0t𝖷N(ΠxNΦts𝐄ϑfsN[ΠxNΦts])A^xNdϑfsN=:JtN+J~tN\displaystyle\hskip 99.58464pt+\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left(\Pi_{x}^{N}\Phi_{t-s}-{\bf E}_{\vartheta_{f_{s}}^{N}}[\Pi_{x}^{N}\Phi_{t-s}]\right)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}=:J^{N}_{t}+\tilde{J}^{N}_{t}

where ΠxN\Pi_{x}^{N} again averages over Ωm\Omega_{m} (and does not touch the other site) as in (4.3).

To estimate the first term JtNJ^{N}_{t} we again approximate the measure ϑfsN\vartheta_{f_{s}}^{N} on 𝒞x\mathcal{C}_{x} by the equilibrium measure with local mass ft(x/N)f_{t}(x/N), and denote it by ϑ¯fs\bar{\vartheta}_{f_{s}} (note that the approximation is made differently for each cube and depends on xx, even if it is written explicitly). This produces an error 𝒪(d+1/N)\mathcal{O}(\ell^{d+1}/N) (using the Lipschitz regularity of Φts\Phi_{t-s} and the exponential convergence ftρf_{t}\to\rho to get uniform in time bounds). We then apply the Poincaré inequality [LY93, Theorem 2] in the cube 𝒞x\mathcal{C}_{x} (whose constant is independent of the number of particles and proportional to the size of the cube) and the law of large number A^xNL2(ϑ¯fsN)=𝒪(eCsd/2)\|\hat{A}^{N}_{x}\|_{L^{2}(\bar{\vartheta}^{N}_{f_{s}})}=\mathcal{O}(e^{-Cs}\ell^{-d/2}) (using uniform bounds on the second moment of ϑ¯fsN\bar{\vartheta}^{N}_{f_{s}})

JtN\displaystyle J^{N}_{t} xNd0tΦtsΠxNΦtsL2(ϑ¯fsN)A^xNL2(ϑ¯fsN)ds+𝒪(d+1N)\displaystyle\leq\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\|\Phi_{t-s}-\Pi_{x}^{N}\Phi_{t-s}\|_{L^{2}(\bar{\vartheta}_{f_{s}}^{N})}\|\hat{A}^{N}_{x}\|_{L^{2}(\bar{\vartheta}_{f_{s}}^{N})}\,{\rm d}s+\mathcal{O}\left(\frac{\ell^{d+1}}{N}\right)
1d2xNd0tD¯x(Φts)eCsds+𝒪(d+1N)\displaystyle\lesssim\ell^{1-\frac{d}{2}}\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\sqrt{\bar{D}^{\ell}_{x}\left(\Phi_{t-s}\right)}e^{-Cs}\,{\rm d}s+\mathcal{O}\left(\frac{\ell^{d+1}}{N}\right)
1d2Nd20t(xNdD¯x(Φts))12eCsds+𝒪(d+1N)\displaystyle\lesssim\ell^{1-\frac{d}{2}}N^{\frac{d}{2}}\int_{0}^{t}\left(\sum_{x\in\mathcal{R}_{N}^{d}}\bar{D}^{\ell}_{x}\left(\Phi_{t-s}\right)\right)^{\frac{1}{2}}e^{-Cs}\,{\rm d}s+\mathcal{O}\left(\frac{\ell^{d+1}}{N}\right)

where D¯x(Φ)\bar{D}^{\ell}_{x}(\Phi) is the Dirichlet form on the cube 𝒞x\mathcal{C}_{x} with respect to the measure ϑ¯fsN\bar{\vartheta}_{f_{s}}^{N}:

D¯x(Φ):=yz𝒞x𝖷N[ηxΦ(η)ηyΦ(η)]2dϑ¯fsN.\bar{D}^{\ell}_{x}(\Phi):=\sum_{y\sim z\in\mathcal{C}_{x}}\int_{\mathsf{X}_{N}}\left[\partial_{\eta_{x}}\Phi(\eta)-\partial_{\eta_{y}}\Phi(\eta)\right]^{2}\,{\rm d}\bar{\vartheta}_{f_{s}}^{N}.

Then we use the entropy production

12N2ddt𝖷NΦts(η)2dϑfsNxNdDx(Φts)+𝒪(1N2)\frac{1}{2N^{2}}\frac{{\rm d}}{{\rm d}t}\int_{\mathsf{X}_{N}}\Phi_{t-s}(\eta)^{2}\,{\rm d}\vartheta_{f_{s}}^{N}\leq-\sum_{x\in\mathcal{R}_{N}^{d}}D^{\ell}_{x}\left(\Phi_{t-s}\right)+\mathcal{O}\left(\frac{1}{N^{2}}\right)

as before to deduce that

0TJtNdtT12(N)1d2+𝒪(Td+1N).\int_{0}^{T}J_{t}^{N}\,{\rm d}t\lesssim T^{\frac{1}{2}}\left(\frac{\ell}{N}\right)^{1-\frac{d}{2}}+\mathcal{O}\left(\frac{T\ell^{d+1}}{N}\right).

To control the second term J~tN\tilde{J}_{t}^{N}, we first use the equivalence of ensemble in [LPY02, Corollary 5.3] on the local equilibrium measure ϑ¯fsN\bar{\vartheta}_{f_{s}}^{N} (using bounds on some exponential moments):

(5.3) V(η)𝒞x=σ(η𝒞x)+𝒪(1d).\langle V^{\prime}(\eta)\rangle_{\mathcal{C}_{x}}=\sigma\left(\langle\eta\rangle_{\mathcal{C}_{x}}\right)+\mathcal{O}\left(\frac{1}{\ell^{d}}\right).

Second we remark that the Lipschitz regularity of Φts\Phi_{t-s} implies that ΠxNΦts𝐄ϑfsN[ΠxNΦts]=𝒪(dNd)\Pi_{x}^{N}\Phi_{t-s}-{\bf E}_{\vartheta_{f_{s}}^{N}}[\Pi_{x}^{N}\Phi_{t-s}]=\mathcal{O}(\ell^{d}N^{-d}), and since the average of A^xN\hat{A}_{x}^{N} with respect to ϑfsN\vartheta_{f_{s}}^{N} is 𝒪(/N)\mathcal{O}(\ell/N), we can write

J~tN=xNd0t𝖷N(ΠxNΦts[η𝒞x]ΠxNΦts[fs(xN)])A^xNdϑfsN+𝒪(N).\tilde{J}_{t}^{N}=\sum_{x\in\mathcal{R}_{N}^{d}}\int_{0}^{t}\int_{\mathsf{X}_{N}}\left(\Pi_{x}^{N}\Phi_{t-s}[\langle\eta\rangle_{\mathcal{C}_{x}}]-\Pi_{x}^{N}\Phi_{t-s}\left[f_{s}\left(\frac{x}{N}\right)\right]\right)\hat{A}^{N}_{x}\,{\rm d}\vartheta_{f_{s}}^{N}+\mathcal{O}\left(\frac{\ell}{N}\right).

Third, we prove again that the Lipschitz regularity of Φts\Phi_{t-s} (with constant NdN^{-d}) implies a Lipschitz regularity of its averaged projection ΠxNΦts\Pi_{x}^{N}\Phi_{t-s} with constant dNd\ell^{d}N^{-d}, with respect to the local mass. Indeed, given 0=m<m<+0=m<m^{\prime}<+\infty, pick any pair of configuration (η0,ζ0)(\eta_{0},\zeta_{0}) with η0𝒞x=m\langle\eta_{0}\rangle_{\mathcal{C}_{x}}=m, ζ0𝒞x=m\langle\zeta_{0}\rangle_{\mathcal{C}_{x}}=m^{\prime} and η0<ζ0\eta_{0}<\zeta_{0} (such configuration trivially exists since m<mm<m^{\prime}). Then consider the coupling on Ωm×Ωm\Omega_{m}\times\Omega_{m^{\prime}} given by a product of smooth probability distributions localised around respectively δη0\delta_{\eta_{0}} and δζ0\delta_{\zeta_{0}}, so that the support of this coupling only contains strictly ordered η<ζ\eta<\zeta. Then we evolve it along the flow of the coupling operator et~Ne^{t\tilde{\mathcal{L}}_{N}}. The marginals respectively converge to ν,m\nu^{\ell,m} and ν,m\nu^{\ell,m^{\prime}} (convergence to equilibrium of the oiriginal evolution). Arguing as for the ZRP, we deduce that W1(ν,m,ν,m)=mmW_{1}(\nu^{\ell,m},\nu^{\ell,m^{\prime}})=m^{\prime}-m, and a corresponding optimal coupling Π\Pi associated to this distance is so that the cost does not change sign on its support, i.e. ηζ\eta\leq\zeta in the support. We deduce as for the ZRP that ΠxNΦts\Pi_{x}^{N}\Phi_{t-s} is dNd\ell^{d}N^{-d}-Lipschitz.

We finally deduce from (5.3), the Taylor formula, the approximation of ϑfsN\vartheta^{N}_{f_{s}} by ϑ¯fsN\bar{\vartheta}^{N}_{f_{s}}, and the law of large numbers, the same estimate on J~tN\tilde{J}^{N}_{t} as for the ZRP, and finally the same estimate (4.5) on ItNI^{N}_{t}, which concludes the proof. ∎

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 𝐙d{\bf Z}^{d}. 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.