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

Diffusive limit of random walks on tessellations via generalized gradient flows

Anastasiia Hraivoronska, Oliver Tse
Abstract.

We study asymptotic limits of reversible random walks on tessellations via a variational approach, which relies on a specific generalized-gradient-flow formulation of the corresponding forward Kolmogorov equation. We establish sufficient conditions on sequences of tessellations and jump intensities under which a sequence of random walks converges to a diffusion process with a possibly spatially-dependent diffusion tensor.

Key words and phrases:
Random walks, tessellations, diffusive limits, generalized gradient flows, evolutionary convergence

1. Introduction

In this paper, we are interested in the limiting behavior of random walks on graphs corresponding to tessellations in the diffusive regime, known as the diffusive limit. A well-known example of such convergence is that of random walks on lattices to the Brownian motion (for instance, as a consequence of Donsker’s theorem [6, Theorem 14.1]). Many generalizations of Donsker’s theorem have appeared in the literature, including scaling limits of the random conductance model [4, 7], limit theorems for percolation clusters [25, 31], diffusion limits for continuous-time random walks [34, 41], Brownian motion as a limit of deterministic dynamics [29], and others [11, 13, 45]. The techniques used in these references are mainly probabilistic, and the underlying state space is usually the lattice or d\mathbb{R}^{d}. On the other hand, not much is known about diffusive limits of random walks on general geometric graphs and tessellations. This paper aims to contribute to filling this gap by exploiting modern variational techniques.

Recently, there has been renewed interest in studying such limits for families of tessellations from the viewpoint of numerical schemes, for instance, finite-volume methods [5, 16, 19] and flux discretization schemes [18, 26, 28] for parabolic equations such as the Fokker–Planck equation (see below (1.1)). These methods are known to converge for a restrictive class of tessellations. From the variational perspective, an approach similar to ours was used in [15] (one-dimension) and [21] (multi-dimension) to prove convergence of the finite-volume method for the Fokker–Planck equation.

The goal of this paper is twofold: (a) to provide sufficient conditions on the family of tessellations and transitions intensities of the random walk such that diffusive limits exist, and (b) to study the impact of these assumptions on the limit process. We believe that the outcome and the methodology used in this work can help with making advances in e.g. proving the convergence of more general numerical schemes, and in studying evolution equations in random environment.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 1. Examples of tessellations.

To clarify the goal, we now introduce our setup: Let {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0} be a family of finite tessellations of a bounded and convex set Ωd\Omega\subset\mathbb{R}^{d}, where 𝒯h\mathcal{T}^{h} is the family of cells and Σh𝒯h×𝒯h\Sigma^{h}\subset\mathcal{T}^{h}\times\mathcal{T}^{h} is identified with the set of faces (cf. Section 2.1 for the precise definition of cells and faces), and κh:Σh[0,)\kappa^{h}:\Sigma^{h}\to[0,\infty) be transition kernels for the random walk. Examples of tessellations are shown in Figure 1. The small parameter h>0h>0 stands for the characteristic size of the tessellation, i.e. the maximal diameter of the cells. The (time) marginal law of the random walk with initial law ρ¯h\bar{\rho}^{h} is known to satisfy the forward Kolmogorov equation (see for example [43, Section 6.3])

(fKh) tρth=Qhρth,ρ0h=ρ¯h,\partial_{t}\rho_{t}^{h}=Q^{*}_{h}\rho_{t}^{h},\qquad\rho_{0}^{h}=\bar{\rho}^{h},

with QhQ^{*}_{h} being the dual of the generator QhQ_{h} given, for any bounded function φB(𝒯h)\varphi\in B(\mathcal{T}^{h}), by

(Qhφ)(K)=L𝒯Kh[φ(L)φ(K)]κh(K,L),K𝒯h,(Q_{h}\varphi)(K)=\sum_{L\in\mathcal{T}_{K}^{h}}[\varphi(L)-\varphi(K)]\kappa^{h}(K,L),\qquad K\in\mathcal{T}^{h},

where the sum is taken over all elements in the set of adjacent cells 𝒯Kh={L𝒯h:(K,L)Σh}\mathcal{T}^{h}_{K}=\{L\in\mathcal{T}^{h}:(K,L)\in\Sigma^{h}\}. Here, we restrict ourselves to random walks satisfying detailed balance, i.e. random walks admitting a stationary measure πh𝒫(𝒯h)\pi^{h}\in\mathcal{P}(\mathcal{T}^{h}) such that

πh(K)κh(K,L)=πh(L)κh(L,K)(K,L)Σh.\pi^{h}(K)\kappa^{h}(K,L)=\pi^{h}(L)\kappa^{h}(L,K)\qquad\forall(K,L)\in\Sigma^{h}.

The questions we set out to answer are then:

  1. (a)

    Under which sufficient conditions on {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0} and {κh}h>0\{\kappa^{h}\}_{h>0} does the family of solutions {tρth}h>0\{t\mapsto\rho_{t}^{h}\}_{h>0} to (fKh) converge to a non-degenerate diffusion process?

  2. (b)

    What equation does the limiting object satisfy?

For this purpose, we employ a variational formulation based on a generalized gradient structure for the forward Kolmogorov equation (fKh)\eqref{eq_Kolmogorov}, which we describe briefly in the following (see Section 3.1 for more details).

A gradient structure is completely defined by the driving energy h:𝒫(𝒯h)[0,+]\mathcal{E}_{h}:\mathcal{P}(\mathcal{T}^{h})\to[0,+\infty] and the dual dissipation potential h:𝒫(𝒯h)×(Σh)[0,+]\mathcal{R}^{*}_{h}:\mathcal{P}(\mathcal{T}^{h})\times\mathcal{B}(\Sigma^{h})\to[0,+\infty]. Here, 𝒫(X)\mathcal{P}(X) and (X)\mathcal{B}(X) denote the spaces of probability measures and bounded measurable functions on XX respectively. In the case of the random walk, the evolution is driven by the relative entropy with respect to the stationary measure πh\pi^{h}:

h(ρh):=Ent(ρh|πh)={K𝒯hϕ(uh(K))πh(K)if ρhπh with uh:=dρhdπh,+otherwise,\mathcal{E}_{h}(\rho^{h}):=\operatorname{\text{Ent}}(\rho^{h}|\pi^{h})=\begin{cases}\displaystyle\sum_{K\in\mathcal{T}^{h}}\phi\bigl{(}u^{h}(K)\bigr{)}\pi^{h}(K)&\text{if }\rho^{h}\ll\pi^{h}\text{ with }u^{h}:=\dfrac{\mathop{}\!\mathrm{d}\rho^{h}}{\mathop{}\!\mathrm{d}\pi^{h}},\\ +\infty&\text{otherwise,}\end{cases}

with the energy density ϕ(s)=slogss+1\phi(s)=s\log s-s+1. As for the dual dissipation potential h\mathcal{R}_{h}^{*}, a range of choices can give rise to a gradient structure for (fKh). The general form of h\mathcal{R}^{*}_{h} studied in [39] is

h(ρh,ξh)=12(K,L)ΣhΨ(ξh(K,L))α(uh(K),uh(L))κh(K,L)πh(K).\mathcal{R}^{*}_{h}(\rho^{h},\xi^{h})=\frac{1}{2}\sum_{(K,L)\in\Sigma^{h}}\Psi^{*}\bigl{(}\xi^{h}(K,L)\bigr{)}\alpha(u^{h}(K),u^{h}(L))\kappa^{h}(K,L)\pi^{h}(K).

In this work, we make use of the so-called ‘cosh’ gradient structure, for which

Ψ(ξ)=4(cosh(ξ/2)1)andα(u,v)=uv.\Psi^{*}(\xi)=4(\cosh(\xi/2)-1)\quad\text{and}\quad\alpha(u,v)=\sqrt{uv}.

This choice first appeared in [24], was later derived from the large-deviation characterization in [38], and received significant attention in literature thereafter.

Another well-studied choice is the quadratic gradient structure. In particular, the quadratic structure was used to prove the convergence for the finite-volume discretization of the Fokker–Planck equation in [21]. In comparison, the adoption of the cosh-type gradient structure allows us to consider a more general class of tessellations, including a tilted d\mathbb{Z}^{d} tessellation (Example 2.9), and we can dispense with the orthogonality assumption used in [21] and the finite-volume methods mentioned above.

With the introduced h\mathcal{E}_{h} and h\mathcal{R}^{*}_{h}, one can express (fKh) in the form of a continuity equation (CEh) and the force-flux relation (FFh) for the density-flux pair (ρh,jh)(\rho^{h},j^{h}):

(CEh) tρth+div¯jth\displaystyle\partial_{t}\rho^{h}_{t}+\overline{\text{div}}\,j^{h}_{t} =0on (0,T)×𝒯h.\displaystyle=0\qquad\text{on }(0,T)\times\mathcal{T}^{h}.
(FFh) jth\displaystyle j^{h}_{t} =D2h(ρth,¯h(ρth)),\displaystyle=D_{2}\mathcal{R}^{*}_{h}\bigl{(}\rho^{h}_{t},-\overline{\nabla}\mathcal{E}_{h}^{\prime}(\rho^{h}_{t})\bigr{)},

where the flux jhj^{h} is written in terms of h\mathcal{R}^{*}_{h} and h\mathcal{E}_{h} acting on ρh\rho^{h}. Here, D2D_{2} denotes the derivative in the second variable, ¯\overline{\nabla} is the graph gradient and div¯\overline{\text{div}}\, is the graph divergence (defined in Section 3.1). Using Legendre–Fenchel duality, one obtains a variational characterization of the solutions to (FFh) given by the energy-dissipation balance, i.e. for any T>0T>0, the pair (ρh,jh)(\rho^{h},j^{h}) satisfies

(EDBh) h(ρh,jh):=0Th(ρth,jth)+h(ρth,¯h(ρth))dt+h(ρTh)h(ρ0h)=0.\mathcal{I}_{h}(\rho^{h},j^{h}):=\int_{0}^{T}\mathcal{R}_{h}(\rho^{h}_{t},j^{h}_{t})+\mathcal{R}_{h}^{*}\bigl{(}\rho^{h}_{t},-\overline{\nabla}\mathcal{E}_{h}^{\prime}(\rho^{h}_{t})\bigr{)}\,\mathop{}\!\mathrm{d}t+\mathcal{E}_{h}(\rho^{h}_{T})-\mathcal{E}_{h}(\rho^{h}_{0})=0.

with h\mathcal{R}_{h} and h\mathcal{R}_{h}^{*} being Legendre–Fenchel duality pairs w.r.t. the second variable.

When the following chain rule applies for all pairs (νh,ηh)(\nu^{h},\eta^{h}) satisfying (CEh)

(CRh) dmissingdth(νth)=ηth,¯h(νth)for almost every t(0,T),\displaystyle\frac{\mathop{}\!\mathrm{d}missing}{\mathop{}\!\mathrm{d}t}\mathcal{E}_{h}(\nu_{t}^{h})=\langle\eta_{t}^{h},\overline{\nabla}\mathcal{E}_{h}^{\prime}(\nu_{t}^{h})\rangle\qquad\text{for almost every $t\in(0,T)$},

then one also has that h(νh,ηh)0\mathcal{I}_{h}(\nu^{h},\eta^{h})\geq 0. In particular, a pair (ρh,jh)(\rho^{h},j^{h}) satisfying (CEh) and (EDBh) is a minimizer of h\mathcal{I}_{h}, which we use to define generalized gradient flow (GGF) solutions to (fKh):

(ρh,jh)satisfies(EDBh)}(ρh,jh)=argminh:(ρh,jh) is GGF-solution of (fKh).\displaystyle\left.\begin{gathered}(\rho^{h},j^{h})\;\text{satisfies}\\ \text{\eqref{eq_edb_discrete}}\end{gathered}\;\right\}\iff(\rho^{h},j^{h})=\arg\min\mathcal{I}_{h}\iff:(\rho^{h},j^{h})\text{ is GGF-solution of \eqref{eq_Kolmogorov}}.

Outline of strategy

The variational framework described above allows us to prove the discrete-to-continuuum convergence result in question by employing tools from the Calculus of Variations. This form of convergence is known as evolutionary Γ\varGamma-convergence. It was introduced by Sandier and Serfaty in [42] and led to numerous subsequent studies surveyed in [36, 44], see also [37] for recent developments and various forms of EDP convergence.

Our strategy comprises the following main steps:

  1. (1)

    Prove compactness for the family (ρh,jh)(\rho^{h},j^{h}) satisfying (CEh) and (EDBh). This allows us to extract a subsequence converging to a limiting pair (ρ,j)(\rho,j).

  2. (2)

    Prove lim​ inf inequalities for all the functionals in the energy-dissipation functional h\mathcal{I}_{h} to recover a limiting energy-dissipation functional \mathcal{I}:

    (ρ,j)lim infh0h(ρh,jh).\mathcal{I}(\rho,j)\leq\liminf_{h\to 0}\mathcal{I}_{h}(\rho^{h},j^{h}).
  3. (3)

    Recover a limiting diffusion equation of the type

    (fK) tρt=Qρt.\partial_{t}\rho_{t}=Q^{*}\rho_{t}.

    with some generator QQ from the limiting energy-dissipation functional \mathcal{I}.

Let us outline the main ideas of the steps above: Step 1 raises the question of how to approach compactness in our discrete-to-continuous settings when density-flux pairs (ρh,jh)(\rho^{h},j^{h}) belong to different spaces for each h>0h>0. Moreover, in the diffusive limit, we expect to obtain a curve (ρ,j)(\rho,j) satisfying the continuity equation with the usual divergence operator:

(CE) tρt+divjt=0on (0,T)×d.\partial_{t}\rho_{t}+\text{div}\,j_{t}=0\qquad\text{on }(0,T)\times\mathbb{R}^{d}.

For this purpose, we introduce a continuous reconstruction procedure in Section 4 such that any pair (ρh,jh)(\rho^{h},j^{h}) satisfying (CEh) induces a pair (ρ^h,ȷ^h)(\hat{\rho}^{h},\hat{\jmath}^{h}) that satisfies (CE) exactly. This allows us to ‘embed’ the discrete objects into a common space of continuous objects, and serves as a link between the discrete and continuous problems (see Figure 2). Another important aspect of the reconstruction procedure is that it allows us to prove a compactness result for GGF-solutions of (fKh), thereby allowing us to extract a subsequence that converges to a limiting curve (ρ,j)(\rho,j) satisfying (CE).

Discrete problem(ρh,jh) is GGF-solution of (fKh){\begin{array}[]{c}\textbf{Discrete problem}\\[1.99997pt] (\rho^{h},j^{h})\text{ is GGF-solution of }\eqref{eq_Kolmogorov}\end{array}\qquad}Continuous limit(ρ,j) is GF-solution of (fK){\qquad\begin{array}[]{c}\textbf{Continuous limit}\\[1.99997pt] (\rho,j)\text{ is GF-solution of }\eqref{eq_limit_Kolmogorov}\end{array}}(ρh,jh)(CEh):h(ρh,jh)=0{\begin{array}[]{c}(\rho^{h},j^{h})\in\eqref{eq_CE_discrete}:\mathcal{I}_{h}(\rho^{h},j^{h})=0\end{array}\qquad}Continuous reconstructionDefine (ρ^h,ȷ^h)(CE):^h(ρ^h,ȷ^h)=0{\begin{array}[]{c}\textbf{Continuous reconstruction}\\[1.99997pt] \text{Define }(\hat{\rho}^{h},\hat{\jmath}^{h})\in\eqref{eq_CE}:\hat{\mathcal{I}}_{h}(\hat{\rho}^{h},\hat{\jmath}^{h})=0\end{array}\qquad}Continuous limit(ρ,j)(CE):(ρ,j)=0{\qquad\begin{array}[]{c}\textbf{Continuous limit}\\[1.99997pt] (\rho,j)\in\eqref{eq_CE}:\mathcal{I}(\rho,j)=0\end{array}}  Step 3  Step 1Step 2
Figure 2. To pass to the discrete-to-continuum limits (dashed arrow in the first line), we employ the definition of the GGF-solution of (fKh) and introduce continuous reconstructions of the discrete objects with ^h(ρ^h,ȷ^h):=h(ρh,jh)\hat{\mathcal{I}}_{h}(\hat{\rho}^{h},\hat{\jmath}^{h}):=\mathcal{I}_{h}(\rho^{h},j^{h}). Passing h0h\to 0 from the continuous reconstruction to the continuum limit comprises two ingredients: the compactness result and the liminf inequality for h\mathcal{I}_{h}. The energy-dissipation functional \mathcal{I} obtained in the limit gives rise to the limit equation (fK).

To prove the liminf inequality in Step 2, we study all the components of the energy-dissipation functional h\mathcal{I}_{h} separately. The lower semicontinuity of the driving energy h\mathcal{E}_{h} follows from standard result, whereas the challenging part lies in proving the result for the dissipation potential h\mathcal{R}_{h} and the Fisher information 𝒟h:=h(,Dh())\mathcal{D}_{h}:=\mathcal{R}_{h}^{*}(\cdot,-D\mathcal{E}_{h}(\cdot)). For this purpose, we apply Γ\Gamma-convergence techniques to obtain and characterize the variational limits \mathcal{R} and 𝒟\mathcal{D} of {h}h>0\{\mathcal{R}_{h}\}_{h>0} and {𝒟h}h>0\{\mathcal{D}_{h}\}_{h>0} respectively. Here the assumptions placed on the family of tessellations {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0} and transition kernels {κh}h>0\{\kappa^{h}\}_{h>0} that we ask about in question (a) come into play, as the form of \mathcal{R} and 𝒟\mathcal{D} depends strongly on the relationship between {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0} and {κh}h>0\{\kappa^{h}\}_{h>0}.

To illustrate the idea, the discrete Fisher information 𝒟h\mathcal{D}_{h} for the ‘cosh’ gradient structure takes the form

𝒟h(ρh)=(K,L)Σh|(¯uh)(K,L)|2κh(K,L)πh(K) with uh=dρhdπh.\mathcal{D}_{h}(\rho^{h})=\sum_{(K,L)\in\Sigma^{h}}\left|(\overline{\nabla}\sqrt{u^{h}})(K,L)\right|^{2}\kappa^{h}(K,L)\pi^{h}(K)\quad\text{ with }u^{h}=\frac{\mathop{}\!\mathrm{d}\rho^{h}}{\mathop{}\!\mathrm{d}\pi^{h}}.

We prove that, under suitable assumptions on the families {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0}, {κh}h>0\{\kappa^{h}\}_{h>0} and {πh}h>0\{\pi^{h}\}_{h>0} , the family {𝒟h}h>0\{\mathcal{D}_{h}\}_{h>0} Γ\Gamma-converges to a limit functional of the form

𝒟(ρ)=Ωu,𝕋udπwith u=dρdπ,\mathcal{D}(\rho)=\int_{\Omega}\bigl{\langle}\nabla\sqrt{u},\mathbb{T}\nabla\sqrt{u}\bigr{\rangle}\mathop{}\!\mathrm{d}\pi\qquad\text{with }u=\frac{\mathop{}\!\mathrm{d}\rho}{\mathop{}\!\mathrm{d}\pi},

where 𝕋:dd×d\mathbb{T}:\mathbb{R}^{d}\to\mathbb{R}^{d\times d} is a symmetric and positive definite tensor, and π\pi is the limit of the sequence {πh}h>0\{\pi^{h}\}_{h>0}. All assumptions will be made precise in Section 2, where we also formulate and state our main result. The next step in our strategy provides the interpretation of 𝕋\mathbb{T} and π\pi as objects related to a diffusion process on d\mathbb{R}^{d}.

Morally, Step 3 is the ”reverse” procedure of formulating the forward Kolmogorov equation (fKh) as the generalized gradient flow characterised by the energy-dissipation balance (EDBh). Once we have identified the limit energy-dissipation functional \mathcal{I}, we can make use of classical gradient flow theory to deduce the form of the limit forward Kolmogorov equation (fK). In particular, we formally obtain the diffusion equation

(1.1) tρt=div(𝕋(ρt+ρtV)),\partial_{t}\rho_{t}=\text{div}\bigl{(}\mathbb{T}(\nabla\rho_{t}+\rho_{t}\nabla V)\bigr{)},

with V=log(dπ/dd)V=-\log(\mathop{}\!\mathrm{d}\pi/\mathop{}\!\mathrm{d}\mathscr{L}^{d}), thereby answering question (b). If 𝕋\mathbb{T} arises from a homogeneous random walk on a uniform lattice (see Example 2.8), then we arrive at (1.1) with 𝕋=Id\mathbb{T}=\text{Id}, the identity tensor.

The techniques we use to prove the lim inf\liminf inequalities in Step 2 are similar to those used in [21]; however, the philosophy and results have considerable differences. The authors of [21] prove the convergence of the finite-volume discretization of the equation (1.1) with 𝕋=Id\mathbb{T}=\text{I}_{d} to the original equation. We, on the other hand, start with a more general discrete evolution equation (fKh) and, consequently, recover the diffusion equation (1.1) with variable coefficients 𝕋\mathbb{T}.

Outline of the paper

The paper is organized as follows. In Section 2, we introduce assumptions on the sequence of tessellations and jump intensities that allow us to realize the described strategy. After that, we formulate the main result in Section 2.3. Moreover, in Section 2.4, we discuss several examples that illustrate the applicability of our main result to specific families of tessellations. Section (3) summarizes the definitions of the continuity equations and (generalized) GF-solutions of (fKh) and (fK). In Section 4, we specify the continuous reconstruction procedure and prove compactness result for the GGF-solutions of (fKh). Section 5 is devoted to the Γ\Gamma-convergence results for the Fisher information and the dual dissipation potential. Finally, we conclude with the proof of the main result in Section 6.

Acknowledgments

The authors acknowledge support from NWO Vidi grant 016.Vidi.189.102 ”Dynamical-Variational Transport Costs and Application to Variational Evolution”.

2. Assumptions and Main Results

In this section, we specify our assumptions on the families of tessellations, transition kernels, and stationary measures. After that we formulate our main result in Theorem A.

2.1. Tessellations

Let Ωd\Omega\subset\mathbb{R}^{d} be an open bounded convex set. A tessellation (𝒯h,Σh)(\mathcal{T}^{h},\Sigma^{h}) of Ω\Omega consists of a family 𝒯h\mathcal{T}^{h} of mutually disjoint cells (usually denoted by KK or LL) that are open and convex sets in Ω\Omega, and a family Σh\Sigma^{h} of pairs of neighboring cells {(K,L)𝒯h×𝒯h:d1(K¯L¯)>0}\{(K,L)\in\mathcal{T}^{h}\times\mathcal{T}^{h}:\mathscr{H}^{d-1}(\overline{K}\cap\overline{L})>0\}, where d1\mathscr{H}^{d-1} is the (d1)(d-1)-dimensional Hausdorff measure. Examples of suitable tessellations include Voronoi tessellations, and meshes commonly used in finite-volume methods. The common face of (K,L)Σh(K,L)\in\Sigma^{h} is denoted by (K|L)(K|L). The characterizing size of a tessellation is its maximum diameter:

h:=maxK𝒯h(diam(K)).h:=\max_{K\in\mathcal{T}^{h}}\left({\text{diam}(K)}\right).

The maximum diameter h>0h>0 gives an upper bound on the volumes of the cells |K|Cdhd|K|\leq C_{d}h^{d} and faces |(K|L)|Cd1hd1|(K|L)|\leq C_{d-1}h^{d-1}, where CdC_{d}, Cd1>0C_{d-1}>0 are universal constants depending only on the spatial dimension d1d\geq 1. In our work, it is also necessary to assume lower bounds on the volumes of the cells to prevent degeneration of cells. We make the following non-degeneracy assumption.

Non-degeneracy. There exist ζ(0,1)\zeta\in(0,1) such that (i) For each K𝒯hK\in\mathcal{T}^{h} there is an inner ball B(xK,ζh)KB(x_{K},\zeta h)\subset K with xK=\intbarKxdxx_{K}=\intbar_{K}x\mathop{}\!\mathrm{d}x; (ii) For every (K,L)Σh(K,L)\in\Sigma^{h} it holds that |(K|L)|ζhd1|(K|L)|\geq\zeta h^{d-1}.

Remark 2.1.

The non-degeneracy assumption implies particularly that |K|Cd(ζh)d|K|\geq C_{d}(\zeta h)^{d} for all K𝒯hK\in\mathcal{T}^{h}, and also provides a uniform bound on the cardinality of neighboring cells (cf. [22, Lemma 2.12]):

C𝒩:=suph>0supK𝒯hcard 𝒯Kh<,C_{\mathcal{N}}:=\sup_{h>0}\sup_{K\in\mathcal{T}^{h}}\text{card\,}\mathcal{T}^{h}_{K}<\infty,

which follows from the following calculations:

L𝒯KhCd(ζh)dL𝒯Kh|L||B(xK,2h)|Cd(2h)dcard 𝒯Kh2dζd.\sum_{L\in\mathcal{T}^{h}_{K}}C_{d}(\zeta h)^{d}\leq\sum_{L\in\mathcal{T}^{h}_{K}}|L|\leq\left|B(x_{K},2h)\right|\leq C_{d}(2h)^{d}\quad\Rightarrow\quad\text{card\,}\mathcal{T}^{h}_{K}\leq\frac{2^{d}}{\zeta^{d}}.

Here, card A\text{card\,}A is the cardinality of the set AA.

Remark 2.2.

While we closely follow the finite-volume setup when defining our tessellations, we remark that in contrast to [19, 22], we do not make the orthogonality assumption, i.e. requiring xKxLx_{K}-x_{L} to be orthogonal to (K|L)(K|L) for (K,L)Σh(K,L)\in\Sigma^{h}.

We now summarize the assumptions on the tessellations that is used within this paper.

Assumptions on 𝒯h\mathcal{T}^{h}. We assume the family of tessellations {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0} to be such that (Ass𝒯\mathcal{T}) {for any h>0 all cells K𝒯h are open, convex, and mutually disjoint;{(𝒯h,Σh)}h>0 is non-degenerate with some ζ(0,1) independent of h.\displaystyle\left\{\quad\begin{aligned} &\text{for any $h>0$ all cells $K\in\mathcal{T}^{h}$ are open, convex, and mutually disjoint;}\\ &\text{$\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0}$ is non-degenerate with some $\zeta\in(0,1)$ independent of $h$.}\end{aligned}\right.

2.2. Relations between jump intensities and tessellation

To obtain the diffusive limit, we need to properly relate the objects that define the dynamics, namely the jump kernels and the stationary measures {(κh,πh)}h>0\{(\kappa^{h},\pi^{h})\}_{h>0}, with the geometric properties of the tessellation. In this section, we emphasize all the assumptions we need to make on {(κh,πh)}h>0\{(\kappa^{h},\pi^{h})\}_{h>0} in relation to {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0}.

Assumptions on πh\pi^{h}. Let πh\pi^{h} be a stationary measure for (fKh) satisfying the detailed balance condition (DB) ϑh(K,L):=πh(K)κh(K,L)=πh(L)κh(L,K)(K,L)Σh.\vartheta^{h}(K,L):=\pi^{h}(K)\kappa^{h}(K,L)=\pi^{h}(L)\kappa^{h}(L,K)\qquad\forall(K,L)\in\Sigma^{h}. We assume πh\pi^{h} to have a density uniformly bounded from above and away from zero: (Bπ\pi) 0<πmininfh>0minK𝒯hπh(K)|K|suph>0maxK𝒯hπh(K)|K|πmax<.0<\pi_{\min}\leq\inf_{h>0}\min_{K\in\mathcal{T}^{h}}\frac{\pi^{h}(K)}{|K|}\leq\sup_{h>0}\max_{K\in\mathcal{T}^{h}}\frac{\pi^{h}(K)}{|K|}\leq\pi_{\max}<\infty. The continuous reconstruction π^h\hat{\pi}^{h} (cf. Section 4) converges in the following sense: dπ^h/dddπ/ddin L1(Ω).\mathop{}\!\mathrm{d}\hat{\pi}^{h}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\to\mathop{}\!\mathrm{d}\pi/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\quad\text{in }L^{1}(\Omega). We further assume that log(dπ/dd)Lipb(Ω)\log(\mathop{}\!\mathrm{d}\pi/\mathop{}\!\mathrm{d}\mathscr{L}^{d})\in\text{Lip}_{b}(\Omega).

Without loss of generality, and for simplicity, we assume πh\pi^{h} to have unit mass.

Example 2.3.

A stationary measure πh\pi^{h} satisfying the above mentioned assumptions can be obtained from a continuous measure π\pi. In practice, the stationary measure is usually given in terms of a potential V:ΩV:\Omega\to\mathbb{R}, i.e. π=eVd\pi=e^{-V}\mathscr{L}^{d}. In this case, we assume VLipb(Ω)V\in\text{Lip}_{b}(\Omega) and set

πh(K):=π(K)=KeV(x)dx.\pi^{h}(K):=\pi(K)=\int_{K}e^{-V(x)}\mathop{}\!\mathrm{d}x.

Then πh\pi^{h} converges to π\pi in the sense specified in Section 4.1, since

dπ^hdddπddL1(Ω)\displaystyle\left\|\frac{\mathop{}\!\mathrm{d}\hat{\pi}^{h}}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}-\frac{\mathop{}\!\mathrm{d}\pi}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}\right\|_{L^{1}(\Omega)} =K𝒯hK|\intbarKeV(y)dyeV(x)|dx\displaystyle=\sum_{K\in\mathcal{T}^{h}}\int_{K}\left|\intbar_{K}e^{-V(y)}\,\mathop{}\!\mathrm{d}y-e^{-V(x)}\right|\,\mathop{}\!\mathrm{d}x
K𝒯hK\intbarK|eV(y)eV(x)|dydx\displaystyle\leq\sum_{K\in\mathcal{T}^{h}}\int_{K}\intbar_{K}\left|e^{-V(y)}-e^{-V(x)}\right|\mathop{}\!\mathrm{d}y\,\mathop{}\!\mathrm{d}x
CK𝒯hK\intbarK|yx|dydxCh|Ω|,\displaystyle\leq C\sum_{K\in\mathcal{T}^{h}}\int_{K}\intbar_{K}|y-x|\,\mathop{}\!\mathrm{d}y\,\mathop{}\!\mathrm{d}x\leq Ch|\Omega|,

and, therefore, πh\pi^{h} satisfies the required assumptions.

Now we introduce scaling assumptions on the joint measure ϑh\vartheta^{h} defined in (DB).

Scaling of ϑh\vartheta^{h}. We assume the existence of constants 0<Cl<Cu<0<C_{l}<C_{u}<\infty independent of hh: (Bϑ\vartheta) Cl|(K|L)||xLxK|ϑh(K,L)Cu|(K|L)||xLxK|(K,L)Σh.C_{l}\frac{|(K|L)|}{|x_{L}-x_{K}|}\leq\vartheta^{h}(K,L)\leq C_{u}\frac{|(K|L)|}{|x_{L}-x_{K}|}\qquad\forall(K,L)\in\Sigma^{h}.

Remark 2.4.

Combining (Bϑ\vartheta) with (Bπ\pi) and the non-degenerate assumption gives rise to many possible formulations of uniform lower and upper bounds. We mention a reformulation of the upper bound that appears frequently in the proofs.

Dividing (Bϑ\vartheta) by |K||K| and using the non-degeneracy assumption on 𝒯h\mathcal{T}^{h} yields

πh(K)|K|κh(K,L)Cu|(K|L)||K||xLxK|Cuζhd1Cd(ζh)d+1=CuCdζd1h2.\frac{\pi^{h}(K)}{|K|}\kappa^{h}(K,L)\leq C_{u}\frac{|(K|L)|}{|K||x_{L}-x_{K}|}\leq C_{u}\frac{\zeta h^{d-1}}{C_{d}(\zeta h)^{d+1}}=\frac{C_{u}}{C_{d}\zeta^{d}}\frac{1}{h^{2}}.

Taking into account that πh(K)/|K|πmin\pi^{h}(K)/|K|\geq\pi_{\min} due to (Bπ\pi), we arrive at

h2L𝒯Khκh(K,L)CuCdζdπmin(card 𝒯Kh).h^{2}\sum_{L\in\mathcal{T}^{h}_{K}}\kappa^{h}(K,L)\leq\frac{C_{u}}{C_{d}\zeta^{d}\pi_{\min}}(\text{card\,}\mathcal{T}^{h}_{K}).

Recalling that the non-degeneracy assumption provides a uniform upper bound on the cardinality of cells in 𝒯Kh\mathcal{T}^{h}_{K} (cf. Remark 2.1), we obtain

(UB) suph>0supK𝒯hh2L𝒯Khκh(K,L)CuC𝒩Cdζdπmin=:Cκ<.\sup_{h>0}\sup_{K\in\mathcal{T}^{h}}h^{2}\sum_{L\in\mathcal{T}^{h}_{K}}\kappa^{h}(K,L)\leq\frac{C_{u}C_{\mathcal{N}}}{C_{d}\zeta^{d}\pi_{\min}}=:C_{\kappa}<\infty.

We need one final assumption on the compatibility of the joint measure ϑh\vartheta^{h} and the geometry of the tessellation, namely the so-called zero-local-average assumption. Intuitively, this assumption ensures that the limiting system remains a gradient flow.

Zero-local-average. For all cells K𝒯hK\in\mathcal{T}^{h} not touching the boundary, i.e. K¯Ω=\overline{K}\cap\partial\Omega=\emptyset, (Aloc{}_{\text{loc}}) L𝒯Khϑh(K,L)(xKxL)=0.\sum_{L\in\mathcal{T}^{h}_{K}}\vartheta^{h}(K,L)(x_{K}-x_{L})=0.

Similar assumptions to (Aloc{}_{\text{loc}}) have emerged in finite-volume schemes as explained in [5, Section 5.2.6], and in (stochastic) homogenization to ensure that the corrector problem has a solution [20, 23, 30]. Later, we will see that (Aloc{}_{\text{loc}}) is only a sufficient condition for the proofs, and can be replaced by a weaker asymptotic assumption (see (AMin) in Section 5.3). We stress that the assumptions on 𝒯h\mathcal{T}^{h}, πh\pi^{h} and scaling of ϑh\vartheta^{h} are required in throughout this paper, but (Aloc{}_{\text{loc}}) or its asymptotic variant (AMin) are only required for the identification of the limit given in Section 5.3.

Remark 2.5.

If the tessellations {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0} are such that (xLxK)(K|L)(x_{L}-x_{K})\perp(K|L) for all (K,L)Σh(K,L)\in\Sigma^{h} and ϑh(K,L)=C|(K|L)|/|xLxK|\vartheta^{h}(K,L)=C|(K|L)|/|x_{L}-x_{K}|, then Aloc{}_{\text{loc}} is satisfied.

2.3. Main result

Definition 2.6 (Admissible continuous reconstruction).

We call a pair (ρ^h,ȷ^h)(\hat{\rho}^{h},\hat{\jmath}^{h}) an admissible continuous reconstruction for a pair (ρh,jh)(\rho^{h},j^{h}) satisfying (CEh) if

  1. (i)

    ρ^h\hat{\rho}^{h} is defined by the piecewise constant reconstruction of the density:

    dρ^thdd:=K𝒯hρth(K)|K|𝟙K;\frac{\mathop{}\!\mathrm{d}\hat{\rho}^{h}_{t}}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}:=\sum_{K\in\mathcal{T}^{h}}\frac{\rho_{t}^{h}(K)}{|K|}\mathbbm{1}_{K};
  2. (ii)

    ȷ^h\hat{\jmath}^{h} is such that (ρ^h,ȷ^h)(\hat{\rho}^{h},\hat{\jmath}^{h}) satisfies (CE).

We state our main result in the following theorem.

Theorem A.

Let {(𝒯h,Σh)}h>0\{(\mathcal{T}_{h},\Sigma_{h})\}_{h>0} be a family of tessellations satisfying (Ass𝒯\mathcal{T}). Further, let {(ρh,jh)}h>0\{(\rho^{h},j^{h})\}_{h>0} be a family of GGF-solutions to (fKh) with {(κh,πh)}h>0\{(\kappa^{h},\pi^{h})\}_{h>0} satisfying (Bπ\pi), (Bϑ\vartheta), and (Aloc{}_{\text{loc}}), and initial data {ρ¯h}h>0\{\bar{\rho}^{h}\}_{h>0} satisfying

suph>0h(ρ¯h)<.\sup\nolimits_{h>0}\mathcal{E}_{h}(\bar{\rho}^{h})<\infty.

Then there exists a (not relabelled) subsequence of admissible continuous reconstructions {(ρ^h,ȷ^h)}h>0\{(\hat{\rho}^{h},\hat{\jmath}^{h})\}_{h>0} and a limit pair (ρ,j)(\rho,j) such that

  1. (1)

    (ρ,j)(\rho,j) satisfies (CE) with the density u:=dρ/dπL1((0,T);L1(Ω,π))u:=\mathop{}\!\mathrm{d}\rho/\mathop{}\!\mathrm{d}\pi\in L^{1}((0,T);L^{1}(\Omega,\pi)) and

    1. (i)

      dρ^th/dπ^hu\mathop{}\!\mathrm{d}\hat{\rho}^{h}_{t}/\mathop{}\!\mathrm{d}\hat{\pi}^{h}\to u strongly in L1(Ω,π)L^{1}(\Omega,\pi) for any t[0,T]t\in[0,T];

    2. (ii)

      ȷ^thdtjtdt\int_{\cdot}\hat{\jmath}^{h}_{t}\mathop{}\!\mathrm{d}t\rightharpoonup^{*}\int_{\cdot}j_{t}\mathop{}\!\mathrm{d}t in ([0,T]×Ω)\mathcal{M}([0,T]\times\Omega).

  2. (2)

    the following liminf estimate holds:

    (ρ,j)lim infh0h(ρh,jh).\mathcal{I}(\rho,j)\leq\liminf_{h\to 0}\mathcal{I}_{h}(\rho^{h},j^{h}).
  3. (3)

    (ρ,j)(\rho,j) is the gradient flow solution with the energy-dissipation functional given as

    (ρ,j)=0TΩ|djtdρt|𝕋12dρt+Ω|ut|𝕋2dπdt+(ρT)(ρ0),\mathcal{I}(\rho,j)=\int_{0}^{T}\int_{\Omega}\left|\frac{\mathop{}\!\mathrm{d}j_{t}}{\mathop{}\!\mathrm{d}\rho_{t}}\right|^{2}_{\mathbb{T}^{-1}}\mathop{}\!\mathrm{d}\rho_{t}+\int_{\Omega}\left|\nabla\sqrt{u_{t}}\right|^{2}_{\mathbb{T}}\mathop{}\!\mathrm{d}\pi\mathop{}\!\mathrm{d}t+\mathcal{E}(\rho_{T})-\mathcal{E}(\rho_{0}),

    where 𝕋:dd×d\mathbb{T}:\mathbb{R}^{d}\to\mathbb{R}^{d\times d} is a symmetric and positive definite diffusion tensor.

Remark 2.7.

The equation corresponding to the energy-dissipation balance given by (ρ,j)\mathcal{I}(\rho,j) is

(2.1) tρt=div(𝕋(ρt+ρtV))on (0,T)×Ω,\partial_{t}\rho_{t}=\text{div}\bigl{(}\mathbb{T}(\nabla\rho_{t}+\rho_{t}\nabla V)\bigr{)}\qquad\text{on }(0,T)\times\Omega,

with the no-flux boundary conditions 𝕋(ρt+ρtV)n=0\mathbb{T}(\nabla\rho_{t}+\rho_{t}\nabla V)\cdot n=0 on Ω\partial\Omega, where V=log(dπ/dd)V=-\log(\mathop{}\!\mathrm{d}\pi/\mathop{}\!\mathrm{d}\mathscr{L}^{d}) is the potential corresponding to the limit stationary measure π\pi.

To characterize the diffusion tensor 𝕋\mathbb{T}, we introduce a tensor 𝕋h\mathbb{T}^{h}:

𝕋h(x)=K𝒯h𝟙K(x)L𝒯Khκh(K,L)(xLxK)(xLxK),xd,\mathbb{T}^{h}(x)=\sum_{K\in\mathcal{T}^{h}}\mathbbm{1}_{K}(x)\sum_{L\in\mathcal{T}^{h}_{K}}\kappa^{h}(K,L)(x_{L}-x_{K})\otimes(x_{L}-x_{K}),\qquad x\in\mathbb{R}^{d},

where xK=\intbarKxdxx_{K}=\intbar_{K}x\mathop{}\!\mathrm{d}x. Then 𝕋\mathbb{T} is obtained as a limit of 𝕋h\mathbb{T}^{h}.

2.4. Examples

To help with getting an intuition for our assumptions and the main result, we present several examples in which the diffusion tensor 𝕋\mathbb{T} can be calculated explicitly.

Example 2.8 (Lattice hdh\mathbb{Z}^{d}).

Consider the simplest tessellation which corresponds to the lattice 𝒯h=hd\mathcal{T}^{h}=h\mathbb{Z}^{d} with d2d\geq 2. We choose the uniform stationary measure πh(K)=|K|=hd\pi^{h}(K)=|K|=h^{d} for all K𝒯hK\in\mathcal{T}^{h} and the uniform joint measure ϑh(K,L)=hd2/2\vartheta^{h}(K,L)=h^{d-2}/2 for all (K,L)Σh(K,L)\in\Sigma^{h}. It follows that the transition kernel is κh(K,L)=1/(2h2)\kappa^{h}(K,L)=1/(2h^{2}). Let {e1,,ed}\{e_{1},\cdots,e_{d}\} be the basis vectors in d\mathbb{R}^{d}. One can always choose the orientation of the basis to be such that for any K𝒯hK\in\mathcal{T}^{h} the vectors (xLxK)(x_{L}-x_{K}) pointing to the neighboring cells are {he1,he1,he2,he2,,hed,hed}\{he_{1},-he_{1},he_{2},-he_{2},\cdots,he_{d},-he_{d}\}. Then the diffusion tensor 𝕋h\mathbb{T}^{h} becomes the identity matrix for all h>0h>0 and, consequently, 𝕋=Id\mathbb{T}=\text{Id}.

We can also make a slightly different choice of the joint measure:

ϑh(K,L)=ci2hd2 if the face (K|L) is orthogonal to ei,\vartheta^{h}(K,L)=\frac{c_{i}}{2}h^{d-2}\quad\text{ if the face }(K|L)\text{ is orthogonal to }e_{i},

where ci>0c_{i}>0 are independent of hh. In this way, we make the jump intensities different in the different directions, i.e. κh(K,L)=ci/(2h2)\kappa^{h}(K,L)=c_{i}/(2h^{2}) if the face (K|L)(K|L) is orthogonal to eie_{i}. Note that (Aloc{}_{\text{loc}}) is still satisfied. In this case,

𝕋=(c1000c2000cd).\mathbb{T}=\left(\begin{array}[]{cccc}c_{1}&0&\cdots&0\\ 0&c_{2}&\cdots&0\\ \cdots&\cdots&\ddots&\cdots\\ 0&0&\cdots&c_{d}\end{array}\right).

Unsurprisingly, this example illustrates that different transitions intensities for the same family of tessellations may lead to different limit diffusion tensors.

Example 2.9 (Tilted h2h\mathbb{Z}^{2}).

Let 𝒯h\mathcal{T}^{h} be a tilted version of the lattice h2h\mathbb{Z}^{2} as shown in the Figure 3. The tilt is given by the parameter α=cosγ\alpha=\cos\gamma, γ[0,π/2)\gamma\in[0,\pi/2), where α=cos(π/2)\alpha=\cos(\pi/2) corresponds to h2h\mathbb{Z}^{2}. Each cell K𝒯hK\in\mathcal{T}^{h} has four neighbors {Kr,Ku,Kl,Kd}\left\{K_{r},K_{u},K_{l},K_{d}\right\}, where the subscript stands for right, up, left, and down neighbors of KK.

Refer to caption
Figure 3. Tilted d\mathbb{Z}^{d} tessellation.

We fix the basis {e1,e2}\{e_{1},e_{2}\} such that (xKrxK)=he1(x_{K_{r}}-x_{K})=he_{1}. In this basis, we have that

(xKlxK)=he1,(xKuxK)=h(α2e1+(1α2)e2),\displaystyle(x_{K_{l}}-x_{K})=-he_{1},\qquad(x_{K_{u}}-x_{K})=h(\alpha^{2}e_{1}+(1-\alpha^{2})e_{2}),
(xKdxK)=h(α2e1+(1α2)e2).\displaystyle(x_{K_{d}}-x_{K})=-h(\alpha^{2}e_{1}+(1-\alpha^{2})e_{2}).

The tensor 𝕋h\mathbb{T}^{h} then takes the form

𝕋h(x)=[\displaystyle\mathbb{T}^{h}(x)=\big{[} (κh(K,Kl)+κh(K,Kr))h2e1e1\displaystyle\left(\kappa^{h}(K,K_{l})+\kappa^{h}(K,K_{r})\right)h^{2}e_{1}\otimes e_{1}
+(κh(K,Ku)+κh(K,Kd))h2(α2e1+(1α2)e2)(α2e1+(1α2)e2)].\displaystyle+\left(\kappa^{h}(K,K_{u})+\kappa^{h}(K,K_{d})\right)h^{2}\left(\alpha^{2}e_{1}+(1-\alpha^{2})e_{2}\right)\otimes\left(\alpha^{2}e_{1}+(1-\alpha^{2})e_{2}\right)\big{]}.

Notice that for any nonnegative κh\kappa^{h}, we can never obtain 𝕋=Id\mathbb{T}=\text{I}_{d}. For uniform kernels κh=1/(2h2)\kappa^{h}=1/(2h^{2}), we get

𝕋(x)=(1+α4α2(1α2)α2(1α2)(1α2)2).\mathbb{T}(x)=\left(\begin{array}[]{cc}1+\alpha^{4}&\alpha^{2}(1-\alpha^{2})\\ \alpha^{2}(1-\alpha^{2})&(1-\alpha^{2})^{2}\end{array}\right).

Analogous to the previous example, this example illustrates that the same transition intensities for different families of tessellations may lead to different limit diffusion tensors.

Example 2.10.

Let Ω=[1,1]\Omega=[-1,1]\subset\mathbb{R}. Consider the tessellation 𝒯h=𝒯h𝒯+h\mathcal{T}^{h}=\mathcal{T}^{h}_{-}\cup\mathcal{T}^{h}_{+} consisting of cells with length h/2h/2 on (,0](-\infty,0], i.e. 𝒯h={(kh/2,(k1)h/2),k}\mathcal{T}^{h}_{-}=\{(-kh/2,-(k-1)h/2),k\in\mathbb{N}\} and the cells with the length hh on [0,)[0,\infty), i.e. 𝒯+h={((k1)h,kh),k}\mathcal{T}^{h}_{+}=\{((k-1)h,kh),k\in\mathbb{N}\} (see Figure 4).

Refer to caption
Figure 4. A tessellation on [1,1][-1,1] consisting of cells with length h/2h/2 in the left half and cells with length hh in the right half.

Consider ϑh(K,L)=1/|xLxK|\vartheta^{h}(K,L)=1/|x_{L}-x_{K}| for (K,L)Σh(K,L)\in\Sigma^{h}, which immediately implies that (Aloc{}_{\text{loc}}) is satisfied, with the uniform stationary measure πh(K)=h\pi^{h}(K)=h for all K𝒯hK\in\mathcal{T}^{h}. Then the tensor 𝕋h\mathbb{T}^{h} reads

𝕋h(x)=L𝒯Khϑh(K,L)πh(K)|xLxK|2=L𝒯Kh|xLxK|πh(K)for xK.\mathbb{T}^{h}(x)=\sum_{L\in\mathcal{T}^{h}_{K}}\frac{\vartheta^{h}(K,L)}{\pi^{h}(K)}|x_{L}-x_{K}|^{2}=\sum_{L\in\mathcal{T}^{h}_{K}}\frac{|x_{L}-x_{K}|}{\pi^{h}(K)}\qquad\text{for }x\in K.

In particular, for x(1,h/2)x\in(-1,-h/2), 𝕋h(x)=1\mathbb{T}^{h}(x)=1, and for x(h,1)x\in(h,1), 𝕋h(x)=2\mathbb{T}^{h}(x)=2. Therefore, in the limit h0h\to 0, we obtain 𝕋(x)=2𝟙(0,1)(x)+𝟙(1,0)(x)\mathbb{T}(x)=2\mathbbm{1}_{(0,1)}(x)+\mathbbm{1}_{(-1,0)}(x).

This last example illustrates how one obtains spatially inhomogeneous diffusion tensors in the limit from the inhomogeneity in the tessellations.

3. Gradient structures: discrete and continuous

In this section, we collect all the necessary definitions and statements regarding the gradient flow formulation of the discrete random walk and of the postulated continuous diffusion governed by (fKh) and (fK) respectively.

3.1. Generalized gradient structure for random walks

In the introduction we outlined the generalized gradient flow formulation for random walks. Gradient structures for Markov jump processes on graphs were first introduced in the independent works of Maas [32], Mielke [35], and Chow, Huang and Zho [12]. Motivated by large-deviation theory, a different form of gradient structure for discrete random walks was discovered by Mielke, Peletier and Renger in [38]. Unlike the earlier gradient structures, these large-deviation inspired gradient structures did not fit into the classical framework of gradient flow theory (see Section 3.2). Based on the energy-dissipation balance, a new framework for these structures, now known as generalized gradient structures, was recently established in [39]. This section collects rigorous definitions and concepts following the framework developed in [39].

We use the graph gradient and graph divergence defined respectively as

¯:(𝒯h)(Σh),\displaystyle\overline{\nabla}:\mathcal{B}(\mathcal{T}^{h})\to\mathcal{B}(\Sigma^{h}),\qquad (¯φh)(K,L)=φh(L)φh(K)for all (K,L)Σh;\displaystyle(\overline{\nabla}\varphi^{h})(K,L)=\varphi^{h}(L)-\varphi^{h}(K)\qquad\text{for all }(K,L)\in\Sigma^{h};
div¯:(Σh)(𝒯h),\displaystyle\overline{\text{div}}\,:\mathcal{M}(\Sigma^{h})\to\mathcal{M}(\mathcal{T}^{h}),\qquad (div¯j)(K)=L𝒯Kh[j(K,L)j(L,K)]for all K𝒯h;\displaystyle(\overline{\text{div}}\,j)(K)=\sum_{L\in\mathcal{T}^{h}_{K}}\bigl{[}j(K,L)-j(L,K)\bigr{]}\qquad\text{for all }K\in\mathcal{T}^{h};

where (X)\mathcal{B}(X) denote the space bounded measurable functions on XX, and (X)\mathcal{M}(X) the space of finite (signed) measures equipped with the topology of weak convergence, i.e. convergence against 𝒞0(X)\mathcal{C}_{0}(X), the space of continuous functions that vanish at infinity. Furthermore, we denote by |ν||\nu| the total variation measure of a measure ν(X)\nu\in\mathcal{M}(X), and by 𝒫(X)\mathcal{P}(X) the space of probability measures equipped with the topology of narrow convergence, i.e. convergence against 𝒞b(X)\mathcal{C}_{b}(X), the space of continuous and bounded functions.

We begin by defining the class of solutions for the continuity equation (CEh).

Definition 3.1.

We call a pair (ρh,jh)(\rho^{h},j^{h}), where

  • ρh𝒞([0,T];𝒫(𝒯h))\rho^{h}\in\mathcal{C}([0,T];\mathcal{P}(\mathcal{T}^{h})) is a curve of measures defined on the tessellation 𝒯h\mathcal{T}^{h}, and

  • jhj^{h} is a measurable family of fluxes jh=(jth)t[0,T](Σh)j^{h}=(j^{h}_{t})_{t\in[0,T]}\subset\mathcal{M}(\Sigma^{h}) with 0T|jth|(Σh)dt<\int_{0}^{T}|j^{h}_{t}|(\Sigma^{h})\mathop{}\!\mathrm{d}t<\infty,

a solution of the discrete continuity equation

(CEh) tρh+div¯jh=0in (0,T)×𝒯h,\partial_{t}\rho^{h}+\overline{\text{div}}\,j^{h}=0\qquad\text{in }(0,T)\times\mathcal{T}^{h},

if for all φh(𝒯h)\varphi^{h}\in\mathcal{B}(\mathcal{T}^{h}) and [s,t][0,T][s,t]\subset[0,T],

(3.1) K𝒯hφh(K)ρth(K)K𝒯hφh(K)ρsh(K)=st(K,L)Σh(¯φh)(K,L)jrh(K,L)dr.\sum_{K\in\mathcal{T}^{h}}\varphi^{h}(K)\rho^{h}_{t}(K)-\sum_{K\in\mathcal{T}^{h}}\varphi^{h}(K)\rho^{h}_{s}(K)=\int_{s}^{t}\sum_{(K,L)\in\Sigma^{h}}(\overline{\nabla}\varphi^{h})(K,L)j^{h}_{r}(K,L)\mathop{}\!\mathrm{d}r.

We denote by 𝒞h(0,T)\mathcal{CE}_{h}(0,T) the set of solutions to the discrete continuity equation (CEh).

Following [39], we define a generalized gradient flow solution of (fKh) as follows:

Definition 3.2 (GGF solutions).

A curve ρh𝒞([0,T];𝒫(𝒯h))\rho^{h}\in\mathcal{C}([0,T];\mathcal{P}(\mathcal{T}^{h})) is said to be an (h,h,h)(\mathcal{E}_{h},\mathcal{R}_{h},\mathcal{R}^{*}_{h}) -generalized gradient flow solution of (fKh) with initial data ρ¯h𝒫(𝒯h)dom(h)\bar{\rho}^{h}\in\mathcal{P}(\mathcal{T}^{h})\cap\text{dom}(\mathcal{E}_{h}) if

  1. (i)

    ρ0h=ρ¯h\rho_{0}^{h}=\bar{\rho}^{h} in 𝒫(𝒯h)\mathcal{P}(\mathcal{T}^{h});

  2. (ii)

    there exists a measurable family (jth)t[0,T](Σh)(j^{h}_{t})_{t\in[0,T]}\subset\mathcal{M}(\Sigma^{h}) such that (ρh,jh)𝒞h(0,T)(\rho^{h},j^{h})\in\mathcal{CE}_{h}(0,T) with

    sth(ρrh,jrh)+𝒟h(ρrh)dr+h(ρth)=h(ρsh)for all [s,t][0,T];\int_{s}^{t}\mathcal{R}_{h}(\rho^{h}_{r},j^{h}_{r})+\mathcal{D}_{h}(\rho^{h}_{r})\,\mathop{}\!\mathrm{d}r+\mathcal{E}_{h}(\rho^{h}_{t})=\mathcal{E}_{h}(\rho^{h}_{s})\quad\text{for all }[s,t]\subset[0,T];

    where

    𝒟h(ρ):=inf{lim infnh(ρn,¯h(ρn)):ρnρ,supn0h(ρn)<,ρn>0},\mathcal{D}_{h}(\rho):=\inf\Bigl{\{}\liminf_{n\to\infty}\mathcal{R}^{*}_{h}(\rho_{n},-\overline{\nabla}\mathcal{E}_{h}^{\prime}(\rho_{n})):\rho_{n}\rightharpoonup\rho,\quad\sup_{n\geq 0}\mathcal{E}_{h}(\rho_{n})<\infty,\quad\rho_{n}>0\Bigr{\}},

    i.e. 𝒟h\mathcal{D}_{h} is a lower-semicontinuous envelope of ρ(ρ,¯h(ρ))\rho\mapsto\mathcal{R}^{*}(\rho,-\overline{\nabla}\mathcal{E}_{h}^{\prime}(\rho)).

  3. (iii)

    the chain rule holds, i.e.

    (CRh) dmissingdth(ρth)=jth,¯h(ρth)for almost every t(0,T).\displaystyle\frac{\mathop{}\!\mathrm{d}missing}{\mathop{}\!\mathrm{d}t}\mathcal{E}_{h}(\rho_{t}^{h})=\langle j_{t}^{h},\overline{\nabla}\mathcal{E}_{h}^{\prime}(\rho_{t}^{h})\rangle\qquad\text{for almost every $t\in(0,T)$}.

We now make specific choices for all the components of the energy-dissipation functional (EDBh) introduced in Section 1.

The driving energy h:𝒫(𝒯h)[0,+]\mathcal{E}_{h}:\mathcal{P}(\mathcal{T}^{h})\to[0,+\infty] is taken to be the relative entropy with respect to the stationary measure πh\pi^{h}, i.e.

h(ρh):=Ent(ρh|πh)={K𝒯hϕ(uh(K))πh(K)if ρhπh with uh:=dρhdπh,+otherwise,\mathcal{E}_{h}(\rho^{h}):=\operatorname{\text{Ent}}(\rho^{h}|\pi^{h})=\begin{cases}\displaystyle\sum_{K\in\mathcal{T}^{h}}\phi\bigl{(}u^{h}(K)\bigr{)}\pi^{h}(K)&\text{if }\rho^{h}\ll\pi^{h}\text{ with }u^{h}:=\frac{\mathop{}\!\mathrm{d}\rho^{h}}{\mathop{}\!\mathrm{d}\pi^{h}},\\ +\infty&\text{otherwise,}\end{cases}

with the energy density ϕ(s)=slogss+1\phi(s)=s\log s-s+1.

The dual dissipation potential h:𝒫(𝒯h)×(Σh)[0,)\mathcal{R}_{h}^{*}:\mathcal{P}(\mathcal{T}^{h})\times\mathcal{B}(\Sigma^{h})\to[0,\infty), as defined in the introduction, takes the form

h(ρh,ξh)=12(K,L)ΣhΨ(ξh(K,L))uh(K)uh(L)θh(K,L),\mathcal{R}^{*}_{h}(\rho^{h},\xi^{h})=\frac{1}{2}\sum_{(K,L)\in\Sigma^{h}}\Psi^{*}\left(\xi^{h}(K,L)\right)\sqrt{u^{h}(K)u^{h}(L)}\,\theta^{h}(K,L),

where Ψ(ξ)=4(cosh(ξ/2)1).\Psi^{*}(\xi)=4\left(\cosh{(\xi/2)}-1\right).

The dissipation potential h:𝒫(𝒯h)×(Σh)[0,+]\mathcal{R}_{h}:\mathcal{P}(\mathcal{T}^{h})\times\mathcal{M}(\Sigma^{h})\to[0,+\infty] is the Legendre–Fenchel dual of \mathcal{R}^{*} w.r.t. to its second variable. In particular, it takes the explicit form

(3.2) h(ρh,jh)={12(K,L)Σ~hΨ(wh(K,L)uh(K)uh(L))uh(K)uh(L)ϑh(K,L)if |j|(Σh\Σ~h)=0,+if |j|(Σh\Σ~h)>0,\mathcal{R}_{h}(\rho^{h},j^{h})=\begin{cases}\displaystyle\frac{1}{2}\sum_{(K,L)\in\tilde{\Sigma}^{h}}\Psi\left(\frac{w^{h}(K,L)}{\sqrt{u^{h}(K)u^{h}(L)}}\right)\sqrt{u^{h}(K)u^{h}(L)}\,\vartheta^{h}(K,L)&\text{if }|j|(\Sigma^{h}\backslash\tilde{\Sigma}^{h})=0,\\ +\infty&\text{if }|j|(\Sigma^{h}\backslash\tilde{\Sigma}^{h})>0,\end{cases}

where wh:=djh/dϑhw^{h}:=\mathop{}\!\mathrm{d}j^{h}/\mathop{}\!\mathrm{d}\vartheta^{h}, Σ~h:={(K,L)Σh:uh(K)uh(L)>0}\tilde{\Sigma}^{h}:=\{(K,L)\in\Sigma^{h}:u^{h}(K)\,u^{h}(L)>0\}, and

Ψ(s)=2slog(s+s2+42)s2+4+4.\Psi(s)=2s\log\left(\frac{s+\sqrt{s^{2}+4}}{2}\right)-\sqrt{s^{2}+4}+4.

The Fisher information 𝒟h:𝒫(𝒯h)[0,+]\mathcal{D}_{h}:\mathcal{P}(\mathcal{T}^{h})\to[0,+\infty] is defined as

𝒟h(ρh)=(K,L)Σh|(¯uh)(K,L)|2ϑh(K,L) with uh=dρhdπh.\mathcal{D}_{h}(\rho^{h})=\sum_{(K,L)\in\Sigma^{h}}\left|\left(\overline{\nabla}\sqrt{u^{h}}\right)(K,L)\right|^{2}\vartheta^{h}(K,L)\quad\text{ with }u^{h}=\frac{\mathop{}\!\mathrm{d}\rho^{h}}{\mathop{}\!\mathrm{d}\pi^{h}}.

With the introduced choice of the energy-dissipation functional, the chain-rule estimate holds [39, Corollary 5.6] for any admissible curve with finite dissipation:

Proposition 3.3 (Chain-rule estimate).

For any curve (ρh,jh)𝒞h(0,T)(\rho^{h},j^{h})\in\mathcal{CE}_{h}(0,T) with finite dissipation, i.e.

0Th(ρth,jth)+𝒟h(ρth)dt<,\int_{0}^{T}\mathcal{R}_{h}(\rho^{h}_{t},j^{h}_{t})+\mathcal{D}_{h}(\rho^{h}_{t})\,\mathop{}\!\mathrm{d}t<\infty,

the chain rule (CRh) holds, thus leading to

h(ρh,jh)=0Th(ρth,jth)+𝒟(ρth)dt+(ρTh)(ρ0h)0.\mathcal{I}_{h}(\rho^{h},j^{h})=\int_{0}^{T}\mathcal{R}_{h}(\rho^{h}_{t},j^{h}_{t})+\mathcal{D}(\rho^{h}_{t})\mathop{}\!\mathrm{d}t+\mathcal{E}(\rho^{h}_{T})-\mathcal{E}(\rho^{h}_{0})\geq 0.

In the next lemma, we list the properties of (Ψ,Ψ)(\Psi,\Psi^{*}) that will be used in the proof of Lemma 4.4.

Lemma 3.4.

The Legendre-Fenchel pair (Ψ,Ψ)(\Psi,\Psi^{*}) are such that

  1. (i)

    Ψ\Psi is even and convex, Ψ(0)=0\Psi(0)=0, and Ψ\Psi is strictly increasing for s>0s>0.

  2. (ii)

    For s,p>0s,p>0 the mapping ssΨ(p/s)s\mapsto s\Psi(p/s) is decreasing.

  3. (iii)

    For s>0s>0, Ψ(s)\Psi(s) has a strictly increasing inverse Ψ1:[0,][0,]\Psi^{-1}:[0,\infty]\to[0,\infty] satisfying

    Ψ1(r)rξ+Ψ(ξ)ξfor all ξ>0.\Psi^{-1}(r)\leq\frac{r}{\xi}+\frac{\Psi^{*}(\xi)}{\xi}\quad\text{for all }\xi>0.
  4. (iv)

    Ψ(ξ)ξ2cosh(ξ/2)\Psi^{*}(\xi)\leq\xi^{2}\cosh(\xi/2).

Proof.

(ii) By convexity for 0<s<t0<s<t it holds that:

Ψ(stps)stΨ(ps)tΨ(pt)sΨ(ps).\Psi\left(\frac{s}{t}\cdot\frac{p}{s}\right)\leq\frac{s}{t}\Psi\left(\frac{p}{s}\right)\quad\Rightarrow\quad t\Psi\left(\frac{p}{t}\right)\leq s\Psi\left(\frac{p}{s}\right).

(iii) Since (Ψ,Ψ)(\Psi,\Psi^{*}) are convex conjugate, then

Ψ(s)sξΨ(ξ)for any s,ξ>0.\Psi(s)\geq s\xi-\Psi^{*}(\xi)\quad\text{for any }s,\xi>0.

Therefore,

sΨ1(sξΨ(ξ))Ψ1(r)r+Ψ(ξ)ξ,\displaystyle s\geq\Psi^{-1}\left(s\xi-\Psi^{*}(\xi)\right)\quad\Rightarrow\quad\Psi^{-1}(r)\leq\frac{r+\Psi^{*}(\xi)}{\xi},

thus concluding the proof. ∎

3.2. Gradient structure for continuous diffusion

We noted in Remark 2.7 that the limit forward Kolmogorov equation (also known as the Fokker–Planck equation) takes the form

(fK) tρt=div(𝕋(ρt+ρtV))on (0,T)×Ω,\partial_{t}\rho_{t}=\text{div}\left(\mathbb{T}(\nabla\rho_{t}+\rho_{t}\nabla V)\right)\qquad\text{on }(0,T)\times\Omega,

where Ωd\Omega\subset\mathbb{R}^{d} is a bounded convex domain and VLipb(Ω)V\in\text{Lip}_{b}(\Omega), the space of Lipschitz bounded functions. Such type of equations have been extensively studied over the last century, but the uncovering of their gradient structure in the space of measures only happened about two decades ago in the seminal work [27] by Jordan, Kinderlehrer and Otto, where the 2-Wasserstein metric played a central role. Shortly after, a general framework for gradient flows in metric spaces was developed by Ambrosio, Gigli and Savaré in [3], and the study of gradient flows for various evolution equations in spaces of measures has been an active area of research ever since. While there exist several ways to define gradient flow solutions to (1.1), we take the same approach as for GGF-solution for (fKh), namely, the approach based on the energy-dissipation balance.

The class of curves we consider is the solutions of the continuity equation (CE) in the sense of the following definition.

Definition 3.5.

The set of solutions 𝒞(0,T)\mathcal{CE}(0,T) is given by all pairs (ρ,j)(\rho,j), where

  • ρ𝒞([0,T];𝒫(Ω))\rho\in\mathcal{C}([0,T];\mathcal{P}(\Omega)) is a curve of positive measures defined on Ω\Omega, and

  • jj is a measurable family of fluxes j=(jt)t[0,T](Ω;d)j=(j_{t})_{t\in[0,T]}\subset\mathcal{M}(\Omega;\mathbb{R}^{d}) with

    0TΩ|djtdρt|2dρtdt<,\int_{0}^{T}\int_{\Omega}\left|\frac{\mathop{}\!\mathrm{d}j_{t}}{\mathop{}\!\mathrm{d}\rho_{t}}\right|^{2}\mathop{}\!\mathrm{d}\rho_{t}\mathop{}\!\mathrm{d}t<\infty,

satisfying the continuity equation

(CE) tρ+divj=0in (0,T)×Ω,\partial_{t}\rho+\text{div}j=0\qquad\text{in }(0,T)\times\Omega,

in the following sense:

(3.3) φ,ρtφ,ρs=stφ,jrdrfor all φ𝒞c(d) and [s,t][0,T].\langle\varphi,\rho_{t}\rangle-\langle\varphi,\rho_{s}\rangle=\int_{s}^{t}\langle\nabla\varphi,j_{r}\rangle\mathop{}\!\mathrm{d}r\qquad\text{for all $\varphi\in\mathcal{C}_{c}^{\infty}(\mathbb{R}^{d})$ and $[s,t]\subset[0,T]$.}
Remark 3.6.

It is known that if ρ\rho solves (CE) with

0TΩ|djtdρt|2dρtdt<,\int_{0}^{T}\int_{\Omega}\left|\frac{\mathop{}\!\mathrm{d}j_{t}}{\mathop{}\!\mathrm{d}\rho_{t}}\right|^{2}\mathop{}\!\mathrm{d}\rho_{t}\mathop{}\!\mathrm{d}t<\infty,

then ρ\rho is an absolutely continuous curve in 𝒫(Ω)\mathcal{P}(\Omega) w.r.t. the 2-Wasserstein distance [3, Chapter 8].

Definition 3.7.

A curve ρ𝒞([0,T];𝒫(Ω))\rho\in\mathcal{C}([0,T];\mathcal{P}(\Omega)) is said to be an (,,)(\mathcal{E},\mathcal{R},\mathcal{R}^{*})-gradient flow solution of (fK) with initial data ρ¯𝒫(Ω)dom()\bar{\rho}\in\mathcal{P}(\Omega)\cap\text{dom}(\mathcal{E}) if

  1. (i)

    ρ0=ρ¯\rho_{0}=\bar{\rho} in 𝒫(Ω)\mathcal{P}(\Omega);

  2. (ii)

    there exists a measurable family (jt)t[0,T](Ω;d)(j_{t})_{t\in[0,T]}\subset\mathcal{M}(\Omega;\mathbb{R}^{d}) such that (ρ,j)𝒞(0,T)(\rho,j)\in\mathcal{CE}(0,T) with

    stΩ(ρr,jr)+𝒟(ρr)dr+(ρt)=(ρs)for all [s,t][0,T],\int_{s}^{t}\int_{\Omega}\mathcal{R}(\rho_{r},j_{r})+\mathcal{D}(\rho_{r})\,\mathop{}\!\mathrm{d}r+\mathcal{E}(\rho_{t})=\mathcal{E}(\rho_{s})\quad\text{for all }[s,t]\subset[0,T],

    where

    𝒟(ρ):=inf{lim infn(ρn,¯(ρn)):ρnρ,supn0(ρn)<,ρn>0},\mathcal{D}(\rho):=\inf\left\{\liminf_{n\to\infty}\mathcal{R}^{*}(\rho_{n},-\overline{\nabla}\mathcal{E}^{\prime}(\rho_{n})):\rho_{n}\rightharpoonup\rho,\quad\sup\nolimits_{n\geq 0}\mathcal{E}(\rho_{n})<\infty,\quad\rho_{n}>0\right\},

    i.e. 𝒟\mathcal{D} is a lower-semicontinuous envelope of ρ(ρ,¯(ρ))\rho\mapsto\mathcal{R}^{*}(\rho,-\overline{\nabla}\mathcal{E}^{\prime}(\rho)).

  3. (iii)

    the following chain rule inequality holds:

    dmissingdt(ρt)(ρt,jt)+𝒟(ρt)for almost every t(0,T).-\frac{\mathop{}\!\mathrm{d}missing}{\mathop{}\!\mathrm{d}t}\mathcal{E}(\rho_{t})\leq\mathcal{R}(\rho_{t},j_{t})+\mathcal{D}(\rho_{t})\qquad\text{for almost every $t\in(0,T)$.}

According to the strategy explained in Section 1, we will obtain the energy-dissipation functional \mathcal{I} by proving the corresponding lim inf\liminf inequality for the discrete energy-dissipation functional h\mathcal{I}_{h} introduced in Section 3.1. For a family of GGF solutions {ρh}h>0\{\rho^{h}\}_{h>0} of (fKh), we will immediately have

(ρ,j)lim infh0h(ρh,jh)=0.\mathcal{I}(\rho,j)\leq\liminf_{h\to 0}\mathcal{I}_{h}(\rho^{h},j^{h})=0.

Then to prove that the limit curve ρ\rho indeed satisfies Defintion 3.7, it is left to show that (ρ,j)0\mathcal{I}(\rho,j)\geq 0, which is established by proving the chain rule inequality (iii) (cf. Theorem 6.4).

4. Continuous reconstruction and compactness

In this section, we first introduce our continuous reconstruction procedure for the density-flux pairs (ρh,jh)𝒞h(0,T)(\rho^{h},j^{h})\in\mathcal{CE}_{h}(0,T) (cf. Section 4.1). We then provide a compactness result for the sequence of continuous reconstructions {(ρ^h,ȷ^h)}h>0\{(\hat{\rho}^{h},\hat{\jmath}^{h})\}_{h>0} in Section 4.2.

4.1. Continuous reconstruction

Throughout this paper we will extensively use two operations: projecting functions supported on Ω\Omega on the tessellation 𝒯h\mathcal{T}^{h} and lifting discrete functions supported on 𝒯h\mathcal{T}^{h} to Ω\Omega. Specifically, we define the following operators

h:L1(Ω)(𝒯h),vh(K)=hv(K)=1|K|Kv(x)dx,K𝒯h,\displaystyle\mathbb{P}_{h}:L^{1}(\Omega)\to\mathcal{B}(\mathcal{T}^{h}),~{}\qquad v^{h}(K)=\mathbb{P}_{h}v\,(K)=\frac{1}{|K|}\int_{K}v(x)\mathop{}\!\mathrm{d}x,\quad K\in\mathcal{T}^{h},
𝕃h:(𝒯h)PC(𝒯h),v^h:=𝕃hvh=K𝒯hvh(K)𝟙K,\displaystyle\mathbb{L}_{h}:\mathcal{B}(\mathcal{T}^{h})\to\text{PC}(\mathcal{T}^{h}),\qquad\hat{v}^{h}:=\mathbb{L}_{h}v^{h}=\sum_{K\in\mathcal{T}^{h}}v^{h}(K)\mathbbm{1}_{K},

where PC(𝒯h)L1(Ω)\text{PC}(\mathcal{T}^{h})\subset L^{1}(\Omega) is the set of functions that are piecewise-constant on cells K𝒯hK\in\mathcal{T}^{h}.

The motivating idea for the reconstruction procedure is to embed the curve (ρh,jh)𝒞h(0,T)(\rho^{h},j^{h})\in\mathcal{CE}_{h}(0,T) into the continuous space in such a way that the lifted curve (ρ^h,ȷ^h)(\hat{\rho}^{h},\hat{\jmath}^{h}) belong to 𝒞(0,T)\mathcal{CE}(0,T). Assuming that φh=hφ\varphi^{h}=\mathbb{P}_{h}\varphi, we transform the left-hand side of (3.1) into

K𝒯hφh(K)ρth(K)=K𝒯hρth(K)|K|Kφ(x)dx\displaystyle\sum_{K\in\mathcal{T}^{h}}\varphi^{h}(K)\rho_{t}^{h}(K)=\sum_{K\in\mathcal{T}^{h}}\frac{\rho_{t}^{h}(K)}{|K|}\int_{K}\varphi(x)\mathop{}\!\mathrm{d}x =Ωφ(x)(K𝒯hρth(K)|K|𝟙K(x))dx.\displaystyle=\int_{\Omega}\varphi(x)\left(\sum_{K\in\mathcal{T}^{h}}\frac{\rho_{t}^{h}(K)}{|K|}\mathbbm{1}_{K}(x)\right)\mathop{}\!\mathrm{d}x.

Defining the reconstructed measure ρ^h\hat{\rho}^{h} via its density as

(4.1) dρ^thdd:=K𝒯hρth(K)|K|𝟙K,\frac{\mathop{}\!\mathrm{d}\hat{\rho}^{h}_{t}}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}:=\sum_{K\in\mathcal{T}^{h}}\frac{\rho_{t}^{h}(K)}{|K|}\mathbbm{1}_{K},

we then obtain equality in the first parts of (3.1) and (3.3):

(4.2) K𝒯hφh(K)ρth(K)K𝒯hφh(K)ρsh(K)=Ωφ(x)ρ^th(dx)Ωφ(x)ρ^sh(dx).\sum_{K\in\mathcal{T}^{h}}\varphi^{h}(K)\rho^{h}_{t}(K)-\sum_{K\in\mathcal{T}^{h}}\varphi^{h}(K)\rho^{h}_{s}(K)=\int_{\Omega}\varphi(x)\hat{\rho}^{h}_{t}(\mathop{}\!\mathrm{d}x)-\int_{\Omega}\varphi(x)\hat{\rho}^{h}_{s}(\mathop{}\!\mathrm{d}x).

In what follows we will also frequently use the formulation in terms of density with respect to the stationary measure uh:=dρh/dπhu^{h}:=\mathop{}\!\mathrm{d}\rho^{h}/\mathop{}\!\mathrm{d}\pi^{h}:

K𝒯hφh(K)ρth(K)\displaystyle\sum_{K\in\mathcal{T}^{h}}\varphi^{h}(K)\rho_{t}^{h}(K) =Ωφ(x)(K𝒯huth(K)πh(K)|K|𝟙K(x))dx=Ωφ(x)u^th(x)π^h(dx),\displaystyle=\int_{\Omega}\varphi(x)\left(\sum_{K\in\mathcal{T}^{h}}u_{t}^{h}(K)\frac{\pi^{h}(K)}{|K|}\mathbbm{1}_{K}(x)\right)\mathop{}\!\mathrm{d}x=\int_{\Omega}\varphi(x)\,\hat{u}^{h}_{t}(x)\hat{\pi}^{h}(\mathop{}\!\mathrm{d}x),

with

u^h=𝕃huhanddπ^hdd:=K𝒯hπh(K)|K|𝟙K.\hat{u}^{h}=\mathbb{L}_{h}u^{h}\qquad\text{and}\qquad\frac{\mathop{}\!\mathrm{d}\hat{\pi}^{h}}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}:=\sum_{K\in\mathcal{T}^{h}}\frac{\pi^{h}(K)}{|K|}\mathbbm{1}_{K}.

Assuming the same relation between the test functions φh=hφ\varphi^{h}=\mathbb{P}_{h}\varphi, we now look for a reconstruction formula for the flux that gives the equality in the right-hand sides of (3.1) and (3.3). For this purpose, we find a relation between the corresponding gradients of functions.

Lemma 4.1.

Let φh:=hφ\varphi^{h}:=\mathbb{P}_{h}\varphi be the projection of φ\varphi on 𝒯h\mathcal{T}^{h}. Then there exists a vector-valued measure σKL(Ω;d)\sigma_{KL}\in\mathcal{M}(\Omega;\mathbb{R}^{d}) such that

(4.3) (¯φh)(K,L)=Ω(φ)(x)σKL(dx),(K,L)Σh,φ𝒞b1(Ω).(\overline{\nabla}\varphi^{h})(K,L)=\int_{\Omega}(\nabla\varphi)(x)\cdot\sigma_{KL}(\mathop{}\!\mathrm{d}x),\qquad\forall(K,L)\in\Sigma^{h},\quad\forall\varphi\in\mathcal{C}_{b}^{1}(\Omega).

Furthermore, |σKL|(Ω)2dh.|\sigma_{KL}|(\Omega)\leq 2dh.

Before presenting the proof of this lemma, let us show its application to the definition of the reconstructed fluxes. Applying Lemma 4.1 and the definitions of 𝒞h(0,T)\mathcal{CE}_{h}(0,T) and 𝒞(0,T)\mathcal{CE}(0,T), we note that

(K,L)Σh(¯φh)(K,L)jth(K,L)\displaystyle\sum_{(K,L)\in\Sigma^{h}}(\overline{\nabla}\varphi^{h})(K,L)j^{h}_{t}(K,L) =(K,L)Σhjth(K,L)Ω(φ)(x)σKL(dx)\displaystyle=\sum_{(K,L)\in\Sigma^{h}}j^{h}_{t}(K,L)\int_{\Omega}(\nabla\varphi)(x)\sigma_{KL}(\mathop{}\!\mathrm{d}x)
=Ω(φ)(x)(K,L)Σhjth(K,L)σKL(dx).\displaystyle=\int_{\Omega}(\nabla\varphi)(x)\sum_{(K,L)\in\Sigma^{h}}j^{h}_{t}(K,L)\sigma_{KL}(\mathop{}\!\mathrm{d}x).

Therefore, we define

(4.4) ȷ^th:=(K,L)Σhjth(K,L)σKL.\hat{\jmath}^{h}_{t}:=\sum_{(K,L)\in\Sigma^{h}}j^{h}_{t}(K,L)\sigma_{KL}.

Then for a given (ρh,jh)𝒞h(0,T)(\rho^{h},j^{h})\in\mathcal{CE}_{h}(0,T), the pair (ρ^h,ȷ^h)(\hat{\rho}^{h},\hat{\jmath}^{h}) defined in (4.1) and (4.4) solves (CE)

tρ^th+ȷ^th=0in (0,T)×Ω,\partial_{t}\hat{\rho}^{h}_{t}+\nabla\cdot\hat{\jmath}^{h}_{t}=0\quad\text{in }(0,T)\times\Omega,

in the sense of Definition 3.5.

Proof of Lemma 4.1.

For any pair of neighboring cells (K,L)Σh(K,L)\in\Sigma^{h}:

(¯φh)(K,L)=φh(L)φh(K)\displaystyle(\overline{\nabla}\varphi^{h})(K,L)=\varphi^{h}(L)-\varphi^{h}(K) =Ωφ(y)𝟙L(y)|L|dyΩφ(x)𝟙K(x)|K|dx\displaystyle=\int_{\Omega}\varphi(y)\frac{\mathbbm{1}_{L}(y)}{|L|}\mathop{}\!\mathrm{d}y-\int_{\Omega}\varphi(x)\frac{\mathbbm{1}_{K}(x)}{|K|}\mathop{}\!\mathrm{d}x
=Ω×Ω(φ(y)φ(x))γKLh(dxdy),\displaystyle=\iint_{\Omega\times\Omega}(\varphi(y)-\varphi(x))\,\gamma^{h}_{KL}(\mathop{}\!\mathrm{d}x\,\mathop{}\!\mathrm{d}y),

where γKLh\gamma_{KL}^{h} is an arbitrary coupling between the measures 𝔪K=|K|1d|K\mathfrak{m}_{K}=|K|^{-1}\mathscr{L}^{d}|_{K} and 𝔪L=|L|1d|L\mathfrak{m}_{L}=|L|^{-1}\mathscr{L}^{d}|_{L}. We assume that the coupling is produced by a transport map TKLT_{KL}, meaning that for all xKx\in K there exist unique yLy\in L such that TKLx=yT_{KL}x=y. In this case, the coupling has the form

γKLh=(Id×TKL)#𝔪Kh.\gamma_{KL}^{h}=(\text{I}_{d}\times T_{KL})_{\#}\mathfrak{m}_{K}^{h}.

For a smooth φ\varphi the fundamental theorem of calculus gives:

φ(y)φ(x)=01(φ)(x+τ(yx))dτ(yx).\varphi(y)-\varphi(x)=\int_{0}^{1}(\nabla\varphi)(x+\tau(y-x))\mathop{}\!\mathrm{d}\tau\cdot(y-x).

Rewriting the coupling in terms of the transport map yields:

Ω×Ω\displaystyle\iint_{\Omega\times\Omega} (φ(y)φ(x))γKLh(dxdy)\displaystyle(\varphi(y)-\varphi(x))\gamma^{h}_{KL}(\mathop{}\!\mathrm{d}x\,\mathop{}\!\mathrm{d}y)
=Ω×Ω01(φ)(x+τ(yx))dτ(yx)((Id×TKL)#𝔪K)(dxdy)\displaystyle=\iint_{\Omega\times\Omega}\int_{0}^{1}(\nabla\varphi)(x+\tau(y-x))\mathop{}\!\mathrm{d}\tau\cdot(y-x)\left((\text{I}_{d}\times T_{KL})_{\#}\mathfrak{m}_{K}\right)(\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}y)
=01Ω(φ)(x+τ(TKLxx))(TKLxx)𝔪K(dx)dτ.\displaystyle=\int_{0}^{1}\int_{\Omega}(\nabla\varphi)(x+\tau(T_{KL}x-x))\cdot(T_{KL}x-x)\mathfrak{m}_{K}(\mathop{}\!\mathrm{d}x)\mathop{}\!\mathrm{d}\tau.

Introducing the notation rKL(x):=TKLxxr_{KL}(x):=T_{KL}x-x and ΦKLτ(x):=x+τrKL(x)\Phi^{\tau}_{KL}(x):=x+\tau r_{KL}(x) we proceed:

01Ω(φ)(x+τrKL(x))rKL(x)𝔪K(dx)dτ\displaystyle\int_{0}^{1}\int_{\Omega}(\nabla\varphi)(x+\tau r_{KL}(x))\cdot r_{KL}(x)\mathfrak{m}_{K}(\mathop{}\!\mathrm{d}x)\mathop{}\!\mathrm{d}\tau =01Ω(φ)(x)[(ΦKLτ)#(rKL𝔪K)](dx)dτ\displaystyle=\int_{0}^{1}\int_{\Omega}(\nabla\varphi)(x)\left[(\Phi^{\tau}_{KL})_{\#}\left(r_{KL}\mathfrak{m}_{K}\right)\right](\mathop{}\!\mathrm{d}x)\mathop{}\!\mathrm{d}\tau
=Ω(φ)(x)[01(ΦKLτ)#(rKL𝔪K)dτ](dx).\displaystyle=\int_{\Omega}(\nabla\varphi)(x)\left[\int_{0}^{1}(\Phi^{\tau}_{KL})_{\#}\left(r_{KL}\mathfrak{m}_{K}\right)\mathop{}\!\mathrm{d}\tau\right](\mathop{}\!\mathrm{d}x).

Denoting by σKL\sigma_{KL} the measure 01((ΦKLτ)#(rKL𝔪K))dτ\int_{0}^{1}\left((\Phi^{\tau}_{KL})_{\#}(r_{KL}\mathfrak{m}_{K})\right)\mathop{}\!\mathrm{d}\tau we obtain (4.3).

To estimate the total variation of σKL\sigma_{KL}, we notice that

|01Ωf(x+τrKL(x))rKLi(x)𝔪K(dx)dτ|fLsupxK|rKLi(x)|for f(Ω),\left|\int_{0}^{1}\int_{\Omega}f(x+\tau r_{KL}(x))r^{i}_{KL}(x)\mathfrak{m}_{K}(\mathop{}\!\mathrm{d}x)\mathop{}\!\mathrm{d}\tau\right|\leq\|f\|_{L^{\infty}}\sup_{x\in K}|r^{i}_{KL}(x)|\qquad\text{for }f\in\mathcal{B}(\Omega),

where supxK|rKLi(x)|supxK,yL|xy|2h\displaystyle\sup_{x\in K}|r^{i}_{KL}(x)|\leq\sup_{x\in K,y\in L}|x-y|\leq 2h. Therefore,

|σKL|(Ω)=i=1d|σKLi|(Ω)2dh.|\sigma_{KL}|(\Omega)=\sum_{i=1}^{d}|\sigma^{i}_{KL}|(\Omega)\leq 2dh.

Remark 4.2.

One can notice that the measures σKL\sigma_{KL} constructed in the proof are not uniquely defined due to the freedom in choosing transport maps TKLT_{KL}. However, we will see that the compactness result in Lemma 4.4 does not depend on the specific choice of TKLT_{KL}.

In the case of a lattice, the measure σKL\sigma_{KL} can be calculated explicitly.

Example 4.3.

Consider the tessellation hdh\mathbb{Z}^{d}. For any pair of neighboring cells KK and LL the optimal transport map is TKLx=x+hnKLT_{KL}x=x+hn_{KL}, with nKLn_{KL} being the (outward) normal on the cell face (K|L)(K|L), and, respectively, rKL(x)=hnKLr_{KL}(x)=hn_{KL}. The function ΦKLτ(x)=x+τhnKL\Phi_{KL}^{\tau}(x)=x+\tau hn_{KL} has an inverse (ΦKLτ)1(y)=yτhnKL(\Phi_{KL}^{\tau})^{-1}(y)=y-\tau hn_{KL}. Therefore, it is possible to calculate the measure σKL\sigma_{KL} explicitly:

df(x)σKL(dx)\displaystyle\int_{\mathbb{R}^{d}}f(x)\,\sigma_{KL}(\mathop{}\!\mathrm{d}x) =hnKLd01f(x+τhnKL)𝟙K(x)|K|dτdx\displaystyle=hn_{KL}\int_{\mathbb{R}^{d}}\int_{0}^{1}f(x+\tau hn_{KL})\frac{\mathbbm{1}_{K}(x)}{|K|}\mathop{}\!\mathrm{d}\tau\mathop{}\!\mathrm{d}x
=hnKLdf(x)01𝟙K+τhnKL(x)|K|dτdx.\displaystyle=hn_{KL}\int_{\mathbb{R}^{d}}f(x)\int_{0}^{1}\frac{\mathbbm{1}_{K+\tau hn_{KL}}(x)}{|K|}\mathop{}\!\mathrm{d}\tau\mathop{}\!\mathrm{d}x.

Notice that for any xKx\in K the indicator function 𝟙K+τhnKL(x)\mathbbm{1}_{K+\tau hn_{KL}}(x) is equal to 1 for 1h(hdist(x,(K|L))\frac{1}{h}(h-\text{dist}(x,(K|L)) and equal to 0 afterwards. Therefore, for xKx\in K:

01𝟙K+τhnKL(x)|K|dτ=1h|K|(hdist(x,(K|L)).\int_{0}^{1}\frac{\mathbbm{1}_{K+\tau hn_{KL}}(x)}{|K|}\mathop{}\!\mathrm{d}\tau=\frac{1}{h|K|}(h-\text{dist}(x,(K|L)).

A similar property holds for xLx\in L:

01𝟙K+τhnKL(x)|K|dτ=1|K|1hdist(x,(K|L))1dτ=1h|K|(hdist(x,(K|L)).\int_{0}^{1}\frac{\mathbbm{1}_{K+\tau hn_{KL}}(x)}{|K|}\mathop{}\!\mathrm{d}\tau=\frac{1}{|K|}\int_{\frac{1}{h}\text{dist}(x,(K|L))}^{1}\mathop{}\!\mathrm{d}\tau=\frac{1}{h|K|}(h-\text{dist}(x,(K|L)).

We conclude that

σKL(dx)=nKL|K|(hdist(x,(K|L))dx.\sigma_{KL}(\mathop{}\!\mathrm{d}x)=\frac{n_{KL}}{|K|}(h-\text{dist}(x,(K|L))\mathop{}\!\mathrm{d}x.

4.2. Compactness

Throughout this section, we consider a family {(ρh,jh)}h>0\{(\rho^{h},j^{h})\}_{h>0} of the GGF-solutions to (fKh) with initial data {ρ0h}h>0\{\rho^{h}_{0}\}_{h>0} satisfying suph>0h(ρ0h)<\sup_{h>0}\mathcal{E}_{h}(\rho^{h}_{0})<\infty. With the non-degeneracy assumption on {(𝒯h,Σh)}h>0\{(\mathcal{T}^{h},\Sigma^{h})\}_{h>0}, and the assumptions (Bπ\pi), (Bϑ\vartheta) on {πh}h>0\{\pi^{h}\}_{h>0}, and {ϑh}h>0\{\vartheta^{h}\}_{h>0}, we deduce compactness for the continuous reconstructions of the solutions.

Lemma 4.4.

Let (ȷ^th)t(0,T)(Ω;d)(\hat{\jmath}_{t}^{h})_{t\in(0,T)}\subset\mathcal{M}(\Omega;\mathbb{R}^{d}), h>0h>0, be defined as in (4.4). Then

  1. (1)

    the family

    {ȷ^thdt}h>0is (sequentially) weakly- compact in ((0,T)×Ω;d);\left\{\int_{\cdot}\,\hat{\jmath}_{t}^{h}\,\mathop{}\!\mathrm{d}t\right\}_{h>0}\quad\text{is (sequentially) weakly-$*$ compact in $\mathcal{M}((0,T)\times\Omega;\mathbb{R}^{d})$;}
  2. (2)

    the family {t|ȷ^th|(Ω)}h>0\{t\mapsto|\hat{\jmath}_{t}^{h}|(\Omega)\}_{h>0} is equi-integrable.

In particular, there exists a Borel family (jt)t(0,T)(Ω;d)(j_{t})_{t\in(0,T)}\subset\mathcal{M}(\Omega;\mathbb{R}^{d}) such that

ȷ^thdtjtdtweakly- in ((0,T)×Ω;d)for a (not relabelled) subsequence.\int_{\cdot}\,\hat{\jmath}_{t}^{h}\,\mathop{}\!\mathrm{d}t\rightharpoonup^{*}\int_{\cdot}\,j_{t}\,\mathop{}\!\mathrm{d}t\quad\text{weakly-$*$ in $\mathcal{M}((0,T)\times\Omega;\mathbb{R}^{d})$}\qquad\text{for a (not relabelled) subsequence.}
Proof.

Recall that for almost every t(0,T)t\in(0,T),

ȷ^th=(K,L)Σhjth(K,L)σKL,withσKL=01(ΦKLτ)#(rKL𝔪K)dτ.\hat{\jmath}_{t}^{h}=\sum_{(K,L)\in\Sigma^{h}}j_{t}^{h}(K,L)\sigma_{KL},\quad\text{with}\quad\sigma_{KL}=\int_{0}^{1}(\Phi^{\tau}_{KL})_{\#}\left(r_{KL}\mathfrak{m}_{K}\right)\mathop{}\!\mathrm{d}\tau.

For any measurable set A[0,T]A\subset[0,T] and any (K,L)Σh(K,L)\in\Sigma^{h} denote:

QKLi(A×Ω):=Aϑρth(K,L)|σKLi|(Ω)dtwith ϑρth(K,L):=ρth(K)κh(K,L).Q^{i}_{KL}(A\times\Omega):=\int_{A}\vartheta_{\rho_{t}^{h}}(K,L)|\sigma^{i}_{KL}|(\Omega)\mathop{}\!\mathrm{d}t\qquad\text{with }\vartheta_{\rho_{t}^{h}}(K,L):=\rho_{t}^{h}(K)\kappa^{h}(K,L).

Note that Qi(A×Ω):=(K,L)ΣhQKLi(A×Ω)Q^{i}(A\times\Omega):=\displaystyle\sum_{(K,L)\in\Sigma^{h}}Q^{i}_{KL}(A\times\Omega) multiplied by hh is uniformly bounded because of (UB):

hQi(A×Ω)=h(K,L)ΣhQKLi(A×Ω)2Crh2L𝒯Khκh(K,L)1(A)2CrCκ1(A).hQ^{i}(A\times\Omega)=h\sum_{(K,L)\in\Sigma^{h}}Q^{i}_{KL}(A\times\Omega)\leq 2C_{r}h^{2}\sum_{L\in\mathcal{T}^{h}_{K}}\kappa^{h}(K,L)\mathscr{L}^{1}(A)\leq 2C_{r}C_{\kappa}\mathscr{L}^{1}(A).

Setting Jh:=ȷ^thdtJ^{h}:=\int_{\cdot}\hat{\jmath}_{t}^{h}\,\mathop{}\!\mathrm{d}t, we will show that the sequence of measures {Jh(×Ω)}h>0(0,T)\{J^{h}(\cdot\times\Omega)\}_{h>0}\subset\mathcal{M}(0,T) is uniformly integrable. The properties collected in Lemma 3.4(i) together with (UB) provide the following estimate:

Ψ(h|Jh,i|(A×Ω)hQi(A×Ω))\displaystyle\Psi\left(\frac{h|J^{h,i}|(A\times\Omega)}{hQ^{i}(A\times\Omega)}\right) Ψ(1hQi(A×Ω)A(K,L)Σhh|jth(K,L)||σKLi|(Ω)dt)\displaystyle\leq\Psi\left(\frac{1}{hQ^{i}(A\times\Omega)}\int_{A}\sum_{(K,L)\in\Sigma^{h}}h\left|j_{t}^{h}(K,L)\right||\sigma^{i}_{KL}|(\Omega)\,\mathop{}\!\mathrm{d}t\right)
1hQi(A×Ω)A(K,L)ΣhΨ(hjth(K,L)ϑρth(K,L))ϑρth(K,L)|σKLi|(Ω)dt\displaystyle\leq\frac{1}{hQ^{i}(A\times\Omega)}\int_{A}\sum_{(K,L)\in\Sigma^{h}}\Psi\left(h\frac{j_{t}^{h}(K,L)}{\vartheta_{\rho_{t}^{h}}(K,L)}\right)\vartheta_{\rho_{t}^{h}}(K,L)|\sigma^{i}_{KL}|(\Omega)\,\mathop{}\!\mathrm{d}t
Crh2hQi(A×Ω)A(K,L)ΣhΨ(jth(K,L)ϑρth(K,L))ϑρth(K,L)dt\displaystyle\leq\frac{C_{r}h^{2}}{hQ^{i}(A\times\Omega)}\int_{A}\sum_{(K,L)\in\Sigma^{h}}\Psi\left(\frac{j_{t}^{h}(K,L)}{\vartheta_{\rho_{t}^{h}}(K,L)}\right)\vartheta_{\rho_{t}^{h}}(K,L)\,\mathop{}\!\mathrm{d}t
=Crh2hQi(A×Ω)Ah(ρth,jth)dt.\displaystyle=\frac{C_{r}h^{2}}{hQ^{i}(A\times\Omega)}\int_{A}\mathcal{R}_{h}(\rho_{t}^{h},j_{t}^{h})\mathop{}\!\mathrm{d}t.

Since hQi(A×Ω)Cκ1(A)hQ^{i}(A\times\Omega)\leq C_{\kappa}\mathscr{L}^{1}(A), Lemma 3.4(ii) gives that:

Cκ1(A)Ψ(h|Jh,i|(A×Ω)C1(A))hQi(A×Ω)Ψ(h|Jh,i|(A×Ω)hQi(A×Ω))Crh2Ah(ρth,jth)dt.C_{\kappa}\mathscr{L}^{1}(A)\Psi\left(\frac{h|J^{h,i}|(A\times\Omega)}{C\mathscr{L}^{1}(A)}\right)\leq hQ^{i}(A\times\Omega)\Psi\left(\frac{h|J^{h,i}|(A\times\Omega)}{hQ^{i}(A\times\Omega)}\right)\leq C_{r}h^{2}\int_{A}\mathcal{R}_{h}(\rho_{t}^{h},j_{t}^{h})\mathop{}\!\mathrm{d}t.

Taking the inverse yields:

|Jh,i|(A×Ω)Cκ1(A)hΨ1(Crh2C1(A)Ah(ρth,jth)dt).|J^{h,i}|(A\times\Omega)\leq\frac{C_{\kappa}\mathscr{L}^{1}(A)}{h}\Psi^{-1}\left(\frac{C_{r}h^{2}}{C\mathscr{L}^{1}(A)}\int_{A}\mathcal{R}_{h}(\rho_{t}^{h},j_{t}^{h})\mathop{}\!\mathrm{d}t\right).

Since (ρh,jh)(\rho^{h},j^{h}) are generalized gradient flow solutions in the sense of Definition 3.2, the integral of h\mathcal{R}_{h} is bounded uniformly in hh under the assumption on the initial conditions:

Ah(ρth,jth)dt0Th(ρth,jth)dtsuph>0h(ρ0h)=:M0<.\int_{A}\mathcal{R}_{h}(\rho_{t}^{h},j_{t}^{h})\mathop{}\!\mathrm{d}t\leq\int_{0}^{T}\mathcal{R}_{h}(\rho_{t}^{h},j_{t}^{h})\mathop{}\!\mathrm{d}t\leq\sup_{h>0}\mathcal{E}_{h}(\rho^{h}_{0})=:M_{0}<\infty.

Now we use the upper bound from Lemma 3.4(iii) to obtain

|Jh,i|(A×Ω)\displaystyle|J^{h,i}|(A\times\Omega) Cκ1(A)h(1ξCrh2Cκ1(A)M0+Ψ(ξ)ξ)\displaystyle\leq\frac{C_{\kappa}\mathscr{L}^{1}(A)}{h}\left(\frac{1}{\xi}\frac{C_{r}h^{2}}{C_{\kappa}\mathscr{L}^{1}(A)}M_{0}+\frac{\Psi^{*}(\xi)}{\xi}\right)
CrhξM0+Cκ1(A)hΨ(ξ)ξfor any ξ>0.\displaystyle\leq C_{r}\frac{h}{\xi}M_{0}+\frac{C_{\kappa}\mathscr{L}^{1}(A)}{h}\frac{\Psi^{*}(\xi)}{\xi}\qquad\text{for any $\xi>0$.}

Choosing ξ=βh\xi=\beta h, β>0\beta>0, and using the property Ψ(ξ)ξ2cosh(ξ/2)\Psi^{*}(\xi)\leq\xi^{2}\cosh(\xi/2), we find

suph(0,1)|Jh,i|(A×Ω)CrβM0+Cκ1(A)βcosh(β2)for any β>0.\sup_{h\in(0,1)}|J^{h,i}|(A\times\Omega)\leq\frac{C_{r}}{\beta}M_{0}+C_{\kappa}\mathscr{L}^{1}(A)\beta\cosh\left(\frac{\beta}{2}\right)\qquad\text{for any $\beta>0$.}

Now let ε>0\varepsilon>0 be arbitrary. By choosing β>0\beta>0 such that CrM0/β<ε/2C_{r}M_{0}/\beta<\varepsilon/2, and subsequently 1(A)\mathscr{L}^{1}(A) such that Cκ1(A)βcosh(β/2)<ε/2C_{\kappa}\mathscr{L}^{1}(A)\beta\cosh(\beta/2)<\varepsilon/2, we then conclude that

suph(0,1)|Jh,i|(A×Ω)<ε,i=1,,d.\sup_{h\in(0,1)}|J^{h,i}|(A\times\Omega)<\varepsilon,\qquad i=1,\ldots,d.

Moreover, by applying the estimate above to A=[0,T]A=[0,T], we simply obtain

suph>0|Jh|([0,T]×Ω)(CrM0+Cκ)T,\sup_{h>0}|\,J^{h}|([0,T]\times\Omega)\leq\left(C_{r}M_{0}+C_{\kappa}\right)\sqrt{T},

i.e. the total variation of JhJ^{h} is uniformly bounded, which allows us to extract a converging subsequence (not relabelled) and some JJ such that JhJJ^{h}\rightharpoonup^{*}J holds.

The equi-integrability of t|ȷ^th|(Ω)t\mapsto|\hat{\jmath}_{t}^{h}|(\Omega) readily follows from the estimate above. Since the limit JJ also satisfies the inequality above (weakly-* lower-semicontinuity of the total variation), we conclude that |J|(×Ω)|J|(\cdot\times\Omega) on [0,T][0,T] has Lebesgue density. By disintegration, JJ has the representation J=jtdtJ=\int_{\cdot}j_{t}\mathop{}\!\mathrm{d}t for a Borel family (jt)(Ω;d)(j_{t})\subset\mathcal{M}(\Omega;\mathbb{R}^{d}). ∎

As a consequence of the previous lemma, we obtain the following result for density-flux pairs.

Lemma 4.5.

There exist a (not relabelled) subsequence of pairs (ρ^h,ȷ^h)(\hat{\rho}^{h},\hat{\jmath}^{h}) defined as in (4.1) and (4.4) and a pair (ρ,j)𝒞(0,T)(\rho,j)\in\mathcal{CE}(0,T) such that

ρ^thρtweakly- in 𝒫(Ω) for all t[0,T].\begin{array}[]{ll}\hat{\rho}^{h}_{t}\rightharpoonup^{*}\rho_{t}&\text{weakly-$*$ in }\mathcal{P}(\Omega)\text{ for all }t\in[0,T].\end{array}
Proof.

Since (ρh,jh)(\rho^{h},j^{h}) satisfies (CEh)\eqref{eq_CE_discrete}, then for all h>0h>0 and all [s,t][0,T][s,t]\subset[0,T] we have that

|φ,ρ^thφ,ρ^sh|\displaystyle\left|\langle\varphi,\hat{\rho}^{h}_{t}\rangle-\langle\varphi,\hat{\rho}^{h}_{s}\rangle\right| =|stΩφ(x)ȷ^rh(dx)dr|φLipsuph>0|ȷ^h|([s,t]×Ω).\displaystyle=\left|\int_{s}^{t}\int_{\Omega}\nabla\varphi(x)\cdot\hat{\jmath}^{h}_{r}(\mathop{}\!\mathrm{d}x)\mathop{}\!\mathrm{d}r\right|\leq\|\varphi\|_{\text{Lip}}\sup_{h>0}|\,\hat{\jmath}^{h}|([s,t]\times\Omega).

Hence, the bounded Lipschitz distance is uniformly bounded:

suph>0dBL(ρ^sh,ρ^th)=suph>0supφ{|φ,ρ^thφ,ρ^sh|}suph>0|ȷ^h|([s,t]×Ω),\sup_{h>0}d_{BL}(\hat{\rho}^{h}_{s},\hat{\rho}^{h}_{t})=\sup_{h>0}\sup_{\varphi}\left\{\bigl{|}\langle\varphi,\hat{\rho}^{h}_{t}\rangle-\langle\varphi,\hat{\rho}^{h}_{s}\rangle\bigr{|}\right\}\leq\sup_{h>0}|\,\hat{\jmath}^{h}|([s,t]\times\Omega),

where the supremum is taken over all 11-Lipschitz functions φ\varphi.

From the equi-integrability of t|ȷ^th|(Ω)t\mapsto|\,\hat{\jmath}^{h}_{t}|(\Omega) it follows that {ρ^th}\{\hat{\rho}^{h}_{t}\} satisfies the refined version of Ascoli-Arzelá theorem ([3, Theorem 3.3.1]) and there exist a (not relabelled) subsequence {ρ^h}h>0\{\hat{\rho}^{h}\}_{h>0} and a limit curve ρ𝒞([0,T];𝒫(Ω))\rho\in\mathcal{C}([0,T];\mathcal{P}(\Omega)), such that the asserted convergence holds. ∎

In the next lemma, we provide the uniform bound on the BV-norm for the reconstructed densities u^h:=𝕃h(dρh/dπh)\hat{u}^{h}:=\mathbb{L}_{h}(\mathop{}\!\mathrm{d}\rho^{h}/\mathop{}\!\mathrm{d}\pi^{h}). As a preparation, we state the following property of non-degenerate tessellations 𝒯h\mathcal{T}^{h} [22, Lemma 2.12(ii)].

Proposition 4.6.

Let 𝒯h\mathcal{T}^{h} satisfy the non-degeneracy assumption, and xKx\in K, yLy\in L be arbitrary with K,L𝒯hK,L\in\mathcal{T}^{h}. The cells KK and LL can be connected by a path (Ki)i=0n1𝒯h(K_{i})_{i=0}^{n-1}\subset\mathcal{T}^{h} with K0=KK_{0}=K, Kn1=LK_{n-1}=L, (Ki,Ki+1)Σh(K_{i},K_{i+1})\in\Sigma^{h}, and [x,y](K|L)[x,y]\cap(K|L)\neq\emptyset, and nCζ|xy|/hn\leq C_{\zeta}|x-y|/h, where Cζ>0C_{\zeta}>0 depends only on ζ\zeta.

Lemma 4.7.

Let ρh𝒫(𝒯h)\rho^{h}\in\mathcal{P}(\mathcal{T}^{h}) with 𝒟h(ρh)<\mathcal{D}_{h}(\rho^{h})<\infty. Then u^h=𝕃h(dρh/dπh)\hat{u}^{h}=\mathbb{L}_{h}(\mathop{}\!\mathrm{d}\rho^{h}/\mathop{}\!\mathrm{d}\pi^{h}) satisfies

u^hL1(Ω)1πmin,|Du^h|(Ω)2CκCl𝒟h(ρh).\|\hat{u}^{h}\|_{L^{1}(\Omega)}\leq\frac{1}{\pi_{\min}},\qquad|D\hat{u}^{h}|(\Omega)\leq 2\frac{\sqrt{C_{\kappa}}}{C_{l}}\sqrt{\mathcal{D}_{h}(\rho^{h})}.
Proof.

For a fixed ψ𝒞c1(Ω)\psi\in\mathcal{C}_{c}^{1}(\Omega) we consider any ηd\eta\in\mathbb{R}^{d} such that 0<|η|<dist(supp(ψ),Ω)0<|\eta|<\text{dist}(\text{supp}(\psi),\partial\Omega), then

Ωu^h(x)ψ(x+η)ψ(x)|η|dx\displaystyle\int_{\Omega}\hat{u}^{h}(x)\frac{\psi(x+\eta)-\psi(x)}{|\eta|}\mathop{}\!\mathrm{d}x =1|η|Ωu^h(x)(ψ(x+η)ψ(x))dx\displaystyle=\frac{1}{|\eta|}\int_{\Omega}\hat{u}^{h}(x)(\psi(x+\eta)-\psi(x))\mathop{}\!\mathrm{d}x
=1|η|Ω(u^h(xη)u^h(x))ψ(x)dx\displaystyle=\frac{1}{|\eta|}\int_{\Omega}\left(\hat{u}^{h}(x-\eta)-\hat{u}^{h}(x)\right)\psi(x)\mathop{}\!\mathrm{d}x
1|η|ψLsupp(ψ)|u^h(xη)u^h(x)|dx.\displaystyle\leq\frac{1}{|\eta|}\|\psi\|_{L^{\infty}}\int_{\text{supp}(\psi)}\left|\hat{u}^{h}(x-\eta)-\hat{u}^{h}(x)\right|\mathop{}\!\mathrm{d}x.

Note that

supp(ψ)|u^h(xη)u^h(x)|dx=K𝒯hKsupp(ψ)|u^h(xη)u^h(x)|dx.\displaystyle\int_{\text{supp}(\psi)}\left|\hat{u}^{h}(x-\eta)-\hat{u}^{h}(x)\right|\mathop{}\!\mathrm{d}x=\sum_{K\in\mathcal{T}^{h}}\int_{K\cap\text{supp}(\psi)}\left|\hat{u}^{h}(x-\eta)-\hat{u}^{h}(x)\right|\mathop{}\!\mathrm{d}x.

Since |η|<dist(supp(ψ),Ω)|\eta|<\text{dist}(\text{supp}(\psi),\partial\Omega), we have that xηΩx-\eta\in\Omega for any xsupp(ψ)x\in\text{supp}(\psi). Therefore, we can find a unique cell L𝒯hL\in\mathcal{T}^{h} such that xηLx-\eta\in L. The line segment [x,xη][x,x-\eta] between the points xx and xηx-\eta defines a path between cell KK and cell LL, consisting of pairs (Ki,Ki+1)Σh(K_{i},K_{i+1})\in\Sigma^{h} such that [x,xη](Ki|Ki+1)[x,x-\eta]\cap(K_{i}|K_{i+1})\neq\emptyset. We denote this sequence of pairs by {(K0=K,K1),(K1,K2),,(Kn1,Kn=L)}\{(K_{0}=K,K_{1}),~{}(K_{1},K_{2}),~{}\dots,~{}(K_{n-1},K_{n}=L)\}. We further define the sets

CylΣh(x,η)\displaystyle\text{Cyl}_{\Sigma^{h}}(x,\eta) :={(M~,L~)Σh:[x,xη](M~|L~)},xΩ,\displaystyle:=\left\{(\tilde{M},\tilde{L})\in\Sigma^{h}\,:\,[x,x-\eta]\cap(\tilde{M}|\tilde{L})\neq\emptyset\right\},\qquad x\in\Omega\,,
CylΩ((K,L),η)\displaystyle\text{Cyl}_{\Omega}((K,L),\eta) :={xΩ:[x,xη](K|L)},(K,L)Σh.\displaystyle:=\Bigl{\{}x\in\Omega\,:\,[x,x-\eta]\cap(K|L)\neq\emptyset\Bigr{\}}\,,\qquad(K,L)\in\Sigma^{h}\,.

Applying the triangle inequality, we have

supp(ψ)|u^h(xη)u^h(x)|dx\displaystyle\int_{\text{supp}(\psi)}|\hat{u}^{h}(x-\eta)-\hat{u}^{h}(x)|\mathop{}\!\mathrm{d}x K𝒯hKsupp(ψ)i=0n1|uh(Ki+1)uh(Ki)|dx\displaystyle\leq\sum_{K\in\mathcal{T}^{h}}\int_{K\cap\text{supp}(\psi)}\sum_{i=0}^{n-1}\left|u^{h}(K_{i+1})-u^{h}(K_{i})\right|\mathop{}\!\mathrm{d}x
K𝒯hKsupp(ψ)(M,L)Σh|uh(L)uh(M)|𝟙CylΣh(x,η)(M,L)dx\displaystyle\leq\sum_{K\in\mathcal{T}^{h}}\int_{K\cap\text{supp}(\psi)}\sum_{(M,L)\in\Sigma^{h}}\left|u^{h}(L)-u^{h}(M)\right|\mathbbm{1}_{\text{Cyl}_{\Sigma^{h}}(x,\eta)}(M,L)\,\mathop{}\!\mathrm{d}x
=(M,L)Σh|uh(L)uh(M)|supp(ψ)𝟙CylΩ((M,L),η)(x)dx\displaystyle=\sum_{(M,L)\in\Sigma^{h}}\left|u^{h}(L)-u^{h}(M)\right|\int_{\text{supp}(\psi)}\mathbbm{1}_{\text{Cyl}_{\Omega}((M,L),\eta)}(x)\,\mathop{}\!\mathrm{d}x
(K,L)Σh|uh(L)uh(M)||(K|L)||η|,\displaystyle\leq\sum_{(K,L)\in\Sigma^{h}}\left|u^{h}(L)-u^{h}(M)\right||(K|L)||\eta|,

where the last inequality follows from the geometric argument that 𝟙CylΩ((K,L),η)(x)=1\mathbbm{1}_{\text{Cyl}_{\Omega}((K,L),\eta)}(x)=1 if and only if the point xΩx\in\Omega is in the cylinder CylΩ((K|L),η)\text{Cyl}_{\Omega}((K|L),\eta) with base (K|L)(K|L) and axis parallel to η\eta.

Applying the lower bound from (Bϑ\vartheta) and then the Hölder inequality, we then obtain

supp(ψ)|u^h(xη)u^h(x)|dx\displaystyle\int_{\text{supp}(\psi)}|\hat{u}^{h}(x-\eta)-\hat{u}^{h}(x)|\mathop{}\!\mathrm{d}x |η|Cl(K,L)Σh|uh(L)uh(K)|hϑh(K,L)\displaystyle\leq\frac{|\eta|}{C_{l}}\sum_{(K,L)\in\Sigma^{h}}\left|u^{h}(L)-u^{h}(K)\right|h\vartheta^{h}(K,L)
2|η|Cl(K𝒯hρh(K)L𝒯Khh2κh(K,L))1/2𝒟h(ρh)\displaystyle\leq 2\frac{|\eta|}{C_{l}}\left(\sum_{K\in\mathcal{T}^{h}}\rho^{h}(K)\sum_{L\in\mathcal{T}^{h}_{K}}h^{2}\kappa^{h}(K,L)\right)^{1/2}\sqrt{\mathcal{D}_{h}(\rho^{h})}
2|η|CκCl𝒟h(ρh).\displaystyle\leq 2\frac{|\eta|\sqrt{C_{\kappa}}}{C_{l}}\sqrt{\mathcal{D}_{h}(\rho^{h})}.

Therefore,

Ωu^h(x)ψ(x+η)ψ(x)|η|dx2CκClψL𝒟h(ρh).\int_{\Omega}\hat{u}^{h}(x)\frac{\psi(x+\eta)-\psi(x)}{|\eta|}\mathop{}\!\mathrm{d}x\leq 2\frac{\sqrt{C_{\kappa}}}{C_{l}}\|\psi\|_{L^{\infty}}\sqrt{\mathcal{D}_{h}(\rho^{h})}.

Taking the limit superior as |η|0|\eta|\to 0, and applying the dominated convergence theorem, we obtain

Ωu^h(x)(ηψ)(x)dx2CκClψL𝒟h(ρh).\int_{\Omega}\hat{u}^{h}(x)(\partial_{\eta}\psi)(x)\mathop{}\!\mathrm{d}x\leq 2\frac{\sqrt{C_{\kappa}}}{C_{l}}\|\psi\|_{L^{\infty}}\sqrt{\mathcal{D}_{h}(\rho^{h})}\,.

Finally, we take the supremum over ψ𝒞c1(Ω)\psi\in\mathcal{C}_{c}^{1}(\Omega) satisfying ψL1\|\psi\|_{L^{\infty}}\leq 1 and use the variational characterization of the BV-seminorm to obtain

|Duh|(Ω)2CκCl𝒟h(ρh)for all h>0.|Du^{h}|(\Omega)\leq 2\frac{\sqrt{C_{\kappa}}}{C_{l}}\sqrt{\mathcal{D}_{h}(\rho^{h})}\qquad\text{for all $h>0$}\,.

The bound on the L1L^{1}-norm follows directly from assumption (Bπ\pi). ∎

With the BV-bound proven in Lemma 4.7, we are now prepared to prove the compactness result for the GGF-solutions of (fKh).

Theorem 4.8 (Strong compactness).

Let the family of curves {ρh}h>0\{\rho^{h}\}_{h>0} be the GGF-solutions of (fKh) with suph>0h(ρ0h)<\sup_{h>0}\mathcal{E}_{h}(\rho^{h}_{0})<\infty. Then there exists uL1((0,T);L1(Ω))u\in L^{1}((0,T);L^{1}(\Omega)) and a (not relabelled) subsequence such that

u^thutstrongly in L1(Ω) for 1-a.e. t(0,T).\hat{u}^{h}_{t}\to u_{t}\quad\text{strongly in $L^{1}(\Omega)$\; for $\mathscr{L}^{1}$-a.e. $t\in(0,T)$.}
Proof.

We first notice that the BV bound from Lemma 4.7 holds for almost every t[0,T]t\in[0,T]. Therefore, {tu^th}\{t\mapsto\hat{u}^{h}_{t}\} is tight with respect to the BV-norm in the sense that

suph>00Tu^thBV(Ω)2dt2C2(T+suph>00T𝒟h(ρth)dt)2C2(T+suph>0h(ρ0h)).\sup_{h>0}\int_{0}^{T}\|\hat{u}^{h}_{t}\|_{BV(\Omega)}^{2}\mathop{}\!\mathrm{d}t\leq 2C^{2}\left(T+\sup_{h>0}\int_{0}^{T}\mathcal{D}_{h}(\rho^{h}_{t})\mathop{}\!\mathrm{d}t\right)\leq 2C^{2}\left(T+\sup_{h>0}\mathcal{E}_{h}(\rho^{h}_{0})\right).

Moreover, Lemma 4.5 provides weak integral equicontinuity, i.e.

limτ0suph>00TτdBL(ρ^t+τh,ρ^th)dt\displaystyle\lim_{\tau\to 0}\sup_{h>0}\int_{0}^{T-\tau}d_{BL}(\hat{\rho}^{h}_{t+\tau},\hat{\rho}^{h}_{t})\mathop{}\!\mathrm{d}t limτ0suph>00Tτ|ȷ^h|([t,t+τ]×Ω)dt\displaystyle\leq\lim_{\tau\to 0}\sup_{h>0}\int_{0}^{T-\tau}|\,\hat{\jmath}^{h}|\left([t,t+\tau]\times\Omega\right)\mathop{}\!\mathrm{d}t
limτ00TτCτdt=0.\displaystyle\leq\lim_{\tau\to 0}\int_{0}^{T-\tau}C\tau\mathop{}\!\mathrm{d}t=0.

Together, the tightness condition and the weak integral equicontinuity yield the relative compactness of {u^h}h>0\{\hat{u}^{h}\}_{h>0} in ((0,T);L1(Ω))\mathcal{M}((0,T);L^{1}(\Omega)) [40, Theorem 2]. The relative compactness in ((0,T);L1(Ω))\mathcal{M}((0,T);L^{1}(\Omega)) combined with the uniform integrability estimate

suph>0Au^thL1(Ω)dtπmin1|A|for any 1-measurable setA[0,T].\sup_{h>0}\int_{A}\|\hat{u}^{h}_{t}\|_{L^{1}(\Omega)}\mathop{}\!\mathrm{d}t\leq\pi_{\min}^{-1}|A|\qquad\text{for any $\mathscr{L}^{1}$-measurable set}A\subset[0,T].

provides that {u^h}h>0\{\hat{u}^{h}\}_{h>0} is relatively compact in L1((0,T);L1(Ω))L^{1}((0,T);L^{1}(\Omega)) [40, Proposition 1.10]. Therefore, there exists some uL1((0,T);L1(Ω))u\in L^{1}((0,T);L^{1}(\Omega)) and a subsequence of u^h\hat{u}^{h} (not relabelled) such that u^thut\hat{u}_{t}^{h}\to u_{t} in L1(Ω)L^{1}(\Omega) for almost every t(0,T)t\in(0,T). ∎

5. Gamma-convergence results

This section contains convergence results for the Fisher information 𝒟h\mathcal{D}_{h} and the dual dissipation potential h\mathcal{R}^{*}_{h}. Since h\mathcal{R}^{*}_{h} and 𝒟h\mathcal{D}_{h} are closely related, we will first introduce the results that hold for both functionals using a generic notation (cf. Section 5.1). Then we deal with the dual dissipation potential in Section 5.2, where we show the asymptotic upper bound. In Section 5.3 we apply the general results from Section 5.1 to prove the Γ\Gamma-convergence of the Fisher information.

5.1. General Gamma-convergence results

The notation throughout the first part of this section is as follows:

  1. (i)

    Let 𝒪\mathcal{O} be the family of all open subsets of Ω\Omega with Lipschitz boundary. We denote by 𝒯h|A\mathcal{T}^{h}|_{A} the restriction of 𝒯h\mathcal{T}^{h} to AA, i.e. 𝒯h|A:={K𝒯h:KA}\mathcal{T}^{h}|_{A}:=\left\{K\in\mathcal{T}^{h}:\,K\cap A\neq\emptyset\right\}. Furthermore, we introduce the set A𝒯h:=ΩintK𝒯h|AK¯A_{\mathcal{T}^{h}}:=\Omega\cap\text{int}\bigcup_{K\in\mathcal{T}^{h}|_{A}}\overline{K}, which can be larger then AA (see Figure 5). In what follows, we will use the convergence of the domain A𝒯hA_{\mathcal{T}^{h}} to AA in the following sense:

Proposition 5.1.

For any A𝒪A\in\mathcal{O} the indicator functions 𝟙A𝒯h\mathbbm{1}_{A_{\mathcal{T}^{h}}} converge pointwise d\mathscr{L}^{d}-a.e. to 𝟙A\mathbbm{1}_{A}.

Refer to caption
Figure 5. An example of a triangular tessellation on the square domain Ω\Omega: Consider a set AA that is colored in dark gray. Then the set A𝒯hA_{\mathcal{T}^{h}} is the union of the set AA and the light gray area.
  1. (ii)

    {μh}h>0\{\mu^{h}\}_{h>0} is a family of probability measures on 𝒯h\mathcal{T}^{h} such that dμ^h/ddL1(Ω)\mathop{}\!\mathrm{d}\hat{\mu}^{h}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\in L^{1}(\Omega) for all h>0h>0, where we use the reconstruction procedure defined in (4.1), i.e.

    dμ^hdd=K𝒯hμh(K)|K|𝟙K,\frac{\mathop{}\!\mathrm{d}\hat{\mu}^{h}}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}=\sum_{K\in\mathcal{T}^{h}}\frac{\mu^{h}(K)}{|K|}\mathbbm{1}_{K},

    and there exists μ𝒫(Ω)\mu\in\mathcal{P}(\Omega) with Lebesgue density dμ/ddL1(Ω)\mathop{}\!\mathrm{d}\mu/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\in L^{1}(\Omega) such that

    dμ^hdddμddin L1(Ω) and pointwise d-a.e. as h0.\frac{\mathop{}\!\mathrm{d}\hat{\mu}^{h}}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}\to\frac{\mathop{}\!\mathrm{d}\mu}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}\quad\text{in }L^{1}(\Omega)\text{ and pointwise }\mathscr{L}^{d}\text{-a.e.\ as }h\to 0.
  2. (iii)

    For each h>0h>0, the measure μh\mu^{h} plays the role of the reference measure for the functional

    (5.1) hμ(vh):=(K,L)Σh|(¯vh)(K,L)|2κh(K,L)μh(K),vh(𝒯h).\mathcal{F}_{h}^{\mu}(v^{h}):=\sum_{(K,L)\in\Sigma^{h}}\left|(\overline{\nabla}v^{h})(K,L)\right|^{2}\kappa^{h}(K,L)\,\mu^{h}(K),\qquad\,v^{h}\in\mathcal{B}(\mathcal{T}^{h}).
  3. (iv)

    We introduce a localized version of the functional hμ\mathcal{F}^{\mu}_{h}:

    (5.2) hμ(vh,A):=(K,L)Σh|A|(¯vh)(K,L)|2κh(K,L)μh(K),A𝒪,\mathcal{F}_{h}^{\mu}(v^{h},A):=\sum_{(K,L)\in\Sigma^{h}|_{A}}\left|(\overline{\nabla}v^{h})(K,L)\right|^{2}\kappa^{h}(K,L)\mu^{h}(K),\qquad A\in\mathcal{O},

    where the summation goes over the restriction of Σh\Sigma^{h} to AA, i.e.

    Σh|A={(K,L)Σh:K,L𝒯h|A}.\Sigma^{h}|_{A}=\left\{(K,L)\in\Sigma^{h}:\,K,L\in\mathcal{T}^{h}|_{A}\right\}.
  4. (v)

    Eventually, we will prove Γ\Gamma-convergence with respect to the L2L^{2}-topology. Therefore, we embed the discrete functional into the full L2(Ω)L^{2}(\Omega) space as:

    (5.3) ~hμ(v,A):={hμ(vh,A)if vPC(𝒯h),+otherwise.\tilde{\mathcal{F}}_{h}^{\mu}(v,A):=\begin{cases}\mathcal{F}_{h}^{\mu}(v^{h},A)&\text{if }v\in\text{PC}(\mathcal{T}^{h}),\\ +\infty&\text{otherwise.}\end{cases}
Remark 5.2.

This generic notation relates to the Fisher information 𝒟h\mathcal{D}_{h} in the following way:

𝒟h(ρh)=hπ(uh)with uh=dρhdπh.\mathcal{D}_{h}(\rho^{h})=\mathcal{F}_{h}^{\pi}\Bigl{(}\sqrt{u^{h}}\Bigr{)}\quad\text{with }u^{h}=\frac{\mathop{}\!\mathrm{d}\rho^{h}}{\mathop{}\!\mathrm{d}\pi^{h}}.

The relation of hμ\mathcal{F}_{h}^{\mu} with h\mathcal{R}^{*}_{h} is more subtle. We show in Lemma 5.14 below that for a smooth φ\varphi and a specific choice of approximating sequence φhφ\varphi^{h}\to\varphi it holds that

h(ρh,¯φh)=14hρ(φh)+o(1)|h0.\mathcal{R}^{*}_{h}(\rho^{h},\overline{\nabla}\varphi^{h})=\frac{1}{4}\mathcal{F}_{h}^{\rho}(\varphi^{h})+o(1)|_{h\to 0}.

With the notation at hand, we outline the steps to prove Γ\Gamma-convergence for hμ\mathcal{F}^{\mu}_{h} by means of the localization technique:

  1. (i)

    The family of functionals {~hμ(,A)}h>0\{\tilde{\mathcal{F}}_{h}^{\mu}(\cdot,A)\}_{h>0} has a subsequential Γ\Gamma-limit μ(,A)\mathcal{F}^{\mu}(\cdot,A) for all A𝒪A\in\mathcal{O} (cf. Lemma 5.4).

  2. (ii)

    The functionals μ(v,A)\mathcal{F}^{\mu}(v,A) and, in particular, μ(v,Ω)\mathcal{F}^{\mu}(v,\Omega) have an integral representation:

    μ(v,A)=Af(x,v)dμ.\mathcal{F}^{\mu}(v,A)=\int_{A}f(x,\nabla v)\mathop{}\!\mathrm{d}\mu.

    For this, we need to prove that μ(v,)\mathcal{F}^{\mu}(v,\cdot) satisfies several properties as a set function, namely that μ(v,)\mathcal{F}^{\mu}(v,\cdot) is a measure and is local (cf. Proposition 5.10).

  3. (iii)

    For vH1(Ω,μ)v\in H^{1}(\Omega,\mu), the integrand has an explicit upper bound f(x,v)v,𝕋vf(x,\nabla v)\leq\langle\nabla v,\mathbb{T}\nabla v\rangle (cf. Lemma 5.13) with some tensor 𝕋\mathbb{T} that comprises the properties of the tessellations and the kernel (cf. Lemma 5.12). For a measure μ\mu with the density dμ/dd\mathop{}\!\mathrm{d}\mu/\mathop{}\!\mathrm{d}\mathscr{L}^{d} bounded away from zero, we prove the exact integral representation, i.e. f(x,v)=v,𝕋vf(x,\nabla v)=\langle\nabla v,\mathbb{T}\nabla v\rangle (cf. Theorem 5.20).

Definitions and compactness

We define

(5.4) infμ(,A):=Γ-lim infh0~hμ(,A)supμ(,A):=Γ-lim suph0~hμ(,A)}for every A𝒪,\displaystyle\left.\begin{aligned} \mathcal{F}^{\mu}_{\inf}(\cdot,A)&:=\Gamma\text{-}\liminf_{h\to 0}\tilde{\mathcal{F}}^{\mu}_{h}(\cdot,A)\\ \mathcal{F}^{\mu}_{\sup}(\cdot,A)&:=\Gamma\text{-}\limsup_{h\to 0}\tilde{\mathcal{F}}^{\mu}_{h}(\cdot,A)\end{aligned}\qquad\right\}\quad\text{for every }A\in\mathcal{O},

where, by the usual definition,

infμ(,A)=inf{lim infh0~hμ(vh,A):vhv}supμ(,A)=inf{lim suph0~hμ(vh,A):vhv}}for every vL2(Ω,μ),A𝒪.\displaystyle\left.\begin{aligned} &\mathcal{F}^{\mu}_{\inf}(\cdot,A)=\inf\big{\{}\liminf_{h\to 0}\tilde{\mathcal{F}}^{\mu}_{h}(v_{h},A)~{}:~{}v_{h}\to v\big{\}}\\ &\mathcal{F}^{\mu}_{\sup}(\cdot,A)=\inf\big{\{}\limsup_{h\to 0}\tilde{\mathcal{F}}^{\mu}_{h}(v_{h},A)~{}:~{}v_{h}\to v\big{\}}\end{aligned}\qquad\right\}\quad\text{for every }v\in L^{2}(\Omega,\mu),~{}A\in\mathcal{O}.

Dealing with the functionals on the product space L2(Ω,μ)×𝒪L^{2}(\Omega,\mu)\times\mathcal{O} has a few subtleties due to the set dependence. We proceed in accordance with the theory presented in [14, Chapters 16-20]. Since ~hμ(,A)\tilde{\mathcal{F}}^{\mu}_{h}(\cdot,A) are increasing functionals, i.e. ~hμ(,A)~hμ(,A)\tilde{\mathcal{F}}^{\mu}_{h}(\cdot,A^{\prime})\leq\tilde{\mathcal{F}}^{\mu}_{h}(\cdot,A) for AAA^{\prime}\subset A, we can apply the next definition

Definition 5.3.

We say that hμ\mathcal{F}^{\mu}_{h} Γ¯\overline{\Gamma}-converges to μ\mathcal{F}^{\mu} (in L2(Ω,μ)L^{2}(\Omega,\mu)) if μ\mathcal{F}^{\mu} is the inner regular envelope of both functionals infμ\mathcal{F}^{\mu}_{\inf} and supμ\mathcal{F}^{\mu}_{\sup}.

The compactness result is standard.

Lemma 5.4.

The family of functionals {~hμ}h>0\{\tilde{\mathcal{F}}_{h}^{\mu}\}_{h>0} defined in (5.3) is sequentially Γ¯\overline{\Gamma}-compact, i.e. there exists a functional μ:L2(Ω,μ)×𝒪[0,+]\mathcal{F}^{\mu}:L^{2}(\Omega,\mu)\times\mathcal{O}\to[0,+\infty] such that Γ¯\overline{\Gamma}-limh0~hμ=μ\displaystyle\lim_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}=\mathcal{F}^{\mu} for some subsequence.

Proof.

The compactness theorem for localized functionals is similar to the standard compactness theorem for the Γ\Gamma-convergence (see [14, Proposition 16.9]). ∎

Remark 5.5.

If one knows a priori that supμ\mathcal{F}^{\mu}_{\sup} is inner regular, then the Γ¯\overline{\Gamma}-limit is equivalently characterized by the usual Γ\Gamma-limits for all A𝒪A\in\mathcal{O} ([14, Proposition 16.4, Remark 16.5]):

  1. (Γinf\Gamma_{\inf})

    for every vL2(Ω)v\in L^{2}(\Omega), for every A𝒪A\in\mathcal{O}, and for every sequence vhvv_{h}\to v in L2(Ω)L^{2}(\Omega) it holds that

    (v,A)lim infh0~hμ(vh,A);\mathcal{F}(v,A)\leq\liminf_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(v_{h},A);
  2. (Γsup\Gamma_{\sup})

    for every vL2(Ω)v\in L^{2}(\Omega) and for every A𝒪A\in\mathcal{O}, there exists a sequence vhvv_{h}\to v in L2(Ω)L^{2}(\Omega) such that

    (v,A)lim suph0~hμ(vh,A).\mathcal{F}(v,A)\geq\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(v_{h},A).

Integral representation

Since the subsequential Γ¯\overline{\Gamma}-limit μ\mathcal{F}^{\mu} exists, it is equal to the inner regular envelope of both functionals infμ\mathcal{F}^{\mu}_{\inf} and supμ\mathcal{F}^{\mu}_{\sup}. Therefore, it suffices to show that supμ\mathcal{F}^{\mu}_{\sup} is inner regular to conclude that μ=supμ\mathcal{F}^{\mu}=\mathcal{F}^{\mu}_{\sup}. We will establish inner regularity together with other properties of supμ\mathcal{F}^{\mu}_{\sup} as a set function in Propositon 5.9. All these properties, as well as possible integral representation, rely on growth conditions for supμ\mathcal{F}^{\mu}_{\sup}.

To prove the growth conditions with respect to integrating against a possibly unbounded measure μ\mu, we fix a suitable definition of H1(Ω,μ)H^{1}(\Omega,\mu):

Definition 5.6.

We define H1(Ω,μ)H^{1}(\Omega,\mu) to be the completion of 𝒞b2(Ω)\mathcal{C}_{b}^{2}(\Omega) w.r.t. the norm

fH1(Ω,μ)2:=fL2(Ω,μ)2+fL2(Ω,μ)2.\|f\|_{H^{1}(\Omega,\mu)}^{2}:=\|f\|_{L^{2}(\Omega,\mu)}^{2}+\|\nabla f\|_{L^{2}(\Omega,\mu)}^{2}\,.

A useful observation is the convergence of the discrete approximations 𝕃hhv\mathbb{L}_{h}\mathbb{P}_{h}v to vv.

Lemma 5.7.

Let vH1(Ω,μ)v\in H^{1}(\Omega,\mu), then v^h:=𝕃hhvv\hat{v}^{h}:=\mathbb{L}_{h}\mathbb{P}_{h}v\to v in L2(Ω,μ)L^{2}(\Omega,\mu).

Proof.

By density arguments, it suffices to consider v𝒞b2(Ω)v\in\mathcal{C}_{b}^{2}(\Omega).

The fact that v^hL2(Ω,μ)\hat{v}^{h}\in L^{2}(\Omega,\mu) follows directly from the boundedness of vv, and the convergence follows from the following inequality:

v^hvL2(Ω,μ)2\displaystyle\|\hat{v}^{h}-v\|^{2}_{L^{2}(\Omega,\mu)} =K𝒯hK|\intbarKv(y)dyv(x)|2μ(dx)\displaystyle=\sum_{K\in\mathcal{T}^{h}}\int_{K}\left|\intbar_{K}v(y)\mathop{}\!\mathrm{d}y-v(x)\right|^{2}\mu(\mathop{}\!\mathrm{d}x)
K𝒯hK\intbarK|v(y)v(x)|2dyμ(dx)h2vL2μ(Ω).\displaystyle\leq\sum_{K\in\mathcal{T}^{h}}\int_{K}\intbar_{K}\left|v(y)-v(x)\right|^{2}\mathop{}\!\mathrm{d}y\,\mu(\mathop{}\!\mathrm{d}x)\leq h^{2}\|\nabla v\|_{L^{\infty}}^{2}\,\mu(\Omega)\,.

Passing to the limit h0h\to 0 yields the statement. ∎

Now we establish the Sobolev upper bound for supμ\mathcal{F}^{\mu}_{\sup}.

Lemma 5.8.

For any vH1(Ω,μ)v\in H^{1}(\Omega,\mu) and A𝒪A\in\mathcal{O},

supμ(v,A)4CκA|v|2dμ,\mathcal{F}_{\sup}^{\mu}(v,A)\leq 4C_{\kappa}\int_{A}|\nabla v|^{2}\mathop{}\!\mathrm{d}\mu\,,

where CκC_{\kappa} is as defined in (UB).

Proof.

For any v𝒞b2(Ω)v\in\mathcal{C}_{b}^{2}(\Omega) and h>0h>0, we set v^h:=𝕃hhvPC(𝒯h)\hat{v}^{h}:=\mathbb{L}_{h}\mathbb{P}_{h}v\in\text{PC}(\mathcal{T}^{h}). Then

hμ(hv,A)\displaystyle\mathcal{F}^{\mu}_{h}(\mathbb{P}_{h}v,A) =(K,L)Σh|A|hv(L)hv(K)|2κ(K,L)μh(K)\displaystyle=\sum_{(K,L)\in\Sigma^{h}|_{A}}\left|\mathbb{P}_{h}v(L)-\mathbb{P}_{h}v(K)\right|^{2}\kappa(K,L)\mu^{h}(K)
=(K,L)Σh|Aκ(K,L)μh(K)|v(y)𝔪L(dy)v(x)𝔪K(dx)|2\displaystyle=\sum_{(K,L)\in\Sigma^{h}|_{A}}\kappa(K,L)\mu^{h}(K)\left|\int v(y)\mathfrak{m}_{L}(\mathop{}\!\mathrm{d}y)-\int v(x)\mathfrak{m}_{K}(\mathop{}\!\mathrm{d}x)\right|^{2}
(K,L)Σh|Aκ(K,L)μh(K)|v(y)v(x)|2γKL(dxdy),\displaystyle\leq\sum_{(K,L)\in\Sigma^{h}|_{A}}\kappa(K,L)\mu^{h}(K)\iint\left|v(y)-v(x)\right|^{2}\gamma_{KL}(\mathop{}\!\mathrm{d}x\,\mathop{}\!\mathrm{d}y)\,,

where γKL\gamma_{KL} is a coupling between 𝔪K=d|K\mathfrak{m}_{K}=\mathscr{L}^{d}|_{K} and 𝔪L=d|L\mathfrak{m}_{L}=\mathscr{L}^{d}|_{L}. Since vv is smooth and xx and yy are in neighboring cells, it holds that |v(y)v(x)|2|v(x)|h+O(h2)\left|v(y)-v(x)\right|\leq 2\left|\nabla v(x)\right|h+O(h^{2}), therefore,

hμ(hv,A)\displaystyle\mathcal{F}^{\mu}_{h}(\mathbb{P}_{h}v,A) (K,L)Σh|Aκ(K,L)μh(K)(4|v(x)|2h2γKL(dxdy)+O(h4))\displaystyle\leq\sum_{(K,L)\in\Sigma^{h}|_{A}}\kappa(K,L)\mu^{h}(K)\left(4\iint\left|\nabla v(x)\right|^{2}h^{2}\gamma_{KL}(\mathop{}\!\mathrm{d}x\,\mathop{}\!\mathrm{d}y)+O(h^{4})\right)

Applying (UB) yields

hμ(hv,A)\displaystyle\mathcal{F}^{\mu}_{h}(\mathbb{P}_{h}v,A) 4CκK𝒯h|Aμh(K)(K|v(x)|2𝔪K(dx)+O(h2))\displaystyle\leq 4C_{\kappa}\sum_{K\in\mathcal{T}^{h}|_{A}}\mu^{h}(K)\left(\int_{K}\left|\nabla v(x)\right|^{2}\mathfrak{m}_{K}(\mathop{}\!\mathrm{d}x)+O(h^{2})\right)
4Cκ(A𝒯h|v(x)|2μ^h(dx)+O(h2)μ^h(A𝒯h)).\displaystyle\leq 4C_{\kappa}\left(\int_{A_{\mathcal{T}^{h}}}\left|\nabla v(x)\right|^{2}\hat{\mu}^{h}(\mathop{}\!\mathrm{d}x)+O(h^{2})\hat{\mu}^{h}\bigl{(}A_{\mathcal{T}^{h}}\bigr{)}\right).

The second term vanishes in the limit h0h\to 0 since lim suph>0|μ^h|(Ω)<\limsup_{h>0}|\hat{\mu}^{h}|(\Omega)<\infty. For the first term, we also notice that 𝟙A𝒯h𝟙A\mathbbm{1}_{A_{\mathcal{T}^{h}}}\to\mathbbm{1}_{A} pointwise d\mathscr{L}^{d}-a.e. (see Proposition 5.1) and |v||\nabla v| is bounded on Ω\Omega. Thus, we can apply the generalized dominated convergence theorem [17, Theorem 1.20] to obtain

limh0A𝒯h|v|2dμ^h=A|v|2dμ.\displaystyle\lim_{h\to 0}\int_{A_{\mathcal{T}^{h}}}\left|\nabla v\right|^{2}\mathop{}\!\mathrm{d}\hat{\mu}^{h}=\int_{A}\left|\nabla v\right|^{2}\mathop{}\!\mathrm{d}\mu.

Altogether, we obtain the following bound for any v𝒞b2(Ω)v\in\mathcal{C}_{b}^{2}(\Omega):

supμ(v,A)lim suph0hμ(hv,A)4CκA|v(x)|2μ(dx).\mathcal{F}^{\mu}_{\sup}(v,A)\leq\limsup_{h\to 0}\mathcal{F}^{\mu}_{h}(\mathbb{P}_{h}v,A)\leq 4C_{\kappa}\int_{A}\left|\nabla v(x)\right|^{2}\mu(\mathop{}\!\mathrm{d}x).

For arbitrary vH1(Ω,μ)v\in H^{1}(\Omega,\mu), we consider a sequence {vn}n𝒞b2(Ω)\{v_{n}\}_{n\in\mathbb{N}}\subset\mathcal{C}_{b}^{2}(\Omega) such that vnvv_{n}\to v in H1(Ω,μ)H^{1}(\Omega,\mu), then the lower semicontinuity of sup\mathcal{F}_{\sup} yields

supμ(v,A)lim infnsupμ(vn,A)limn4CκA|vn(x)|2μ(dx)=4CκA|v(x)|2μ(dx),\mathcal{F}^{\mu}_{\sup}(v,A)\leq\liminf_{n\to\infty}\mathcal{F}^{\mu}_{\sup}(v_{n},A)\leq\lim_{n\to\infty}4C_{\kappa}\int_{A}\left|\nabla v_{n}(x)\right|^{2}\mu(\mathop{}\!\mathrm{d}x)=4C_{\kappa}\int_{A}\left|\nabla v(x)\right|^{2}\mu(\mathop{}\!\mathrm{d}x)\,,

thereby concluding the proof. ∎

The properties of supμ\mathcal{F}_{\sup}^{\mu} as a set function, namely, inner regularity, subadditivity, and locality, play a crucial role for the integral representation. The proofs of these properties follow the strategy of De Giorgi’s cut-off functions argument [14]. For the discrete functionals, the proofs were established in [1, 21] and can be applied with minor modification to our settings. For completeness, we include the proofs adapted to our projections and reconstruction procedures in Appendix A.

Proposition 5.9 (Properties of supμ\mathcal{F}_{\sup}^{\mu}).

The functional supμ\mathcal{F}_{\sup}^{\mu} defined in (5.2) has the following properties:

  1. (i)

    Inner regularity: For any vH1(Ω,μ)v\in H^{1}(\Omega,\mu) and for any A𝒪A\in\mathcal{O} it holds that

    supA⊂⊂Asupμ(v,A)=supμ(v,A);\sup_{A^{\prime}\subset\joinrel\subset A}\mathcal{F}^{\mu}_{\sup}(v,A^{\prime})=\mathcal{F}^{\mu}_{\sup}(v,A);
  2. (ii)

    Subadditivity: For any vH1(Ω,μ)v\in H^{1}(\Omega,\mu) and for any A,A,B,B𝒪A,A^{\prime},B,B^{\prime}\in\mathcal{O} such that A⊂⊂AA^{\prime}\subset\joinrel\subset A and B⊂⊂BB^{\prime}\subset\joinrel\subset B it holds that:

    supμ(v,AB)supμ(v,A)+supμ(v,B);\mathcal{F}^{\mu}_{\sup}(v,A^{\prime}\cup B^{\prime})\leq\mathcal{F}^{\mu}_{\sup}(v,A)+\mathcal{F}^{\mu}_{\sup}(v,B);
  3. (iii)

    Locality: For any A𝒪A\in\mathcal{O} and any v,ψH1(Ω,μ)v,\psi\in H^{1}(\Omega,\mu) such that v=ψv=\psi μ\mu-a.e. on AA there holds

    supμ(v,A)=supμ(v,A).\mathcal{F}^{\mu}_{\sup}(v,A)=\mathcal{F}^{\mu}_{\sup}(v,A).

A direct consequence of Lemma 5.8 and Proposition 5.9 is the integral representation of supμ\mathcal{F}_{\sup}^{\mu} [14].

Proposition 5.10 (Properties of the Γ\Gamma-limit).

Let μ:L2(Ω,μ)×𝒪[0,+]\mathcal{F}^{\mu}:L^{2}(\Omega,\mu)\times\mathcal{O}\to[0,+\infty] be Γ¯\overline{\Gamma}-limit of (~hμ)(\tilde{\mathcal{F}}^{\mu}_{h}). For every vH1(Ω,μ)v\in H^{1}(\Omega,\mu) and every A𝒪A\in\mathcal{O} the following properties hold:

  1. (i)

    supμ(v,A)=μ(v,A)\mathcal{F}^{\mu}_{\sup}(v,A)=\mathcal{F}^{\mu}(v,A);

  2. (ii)

    μ(v+c,A)=μ(v,A)\mathcal{F}^{\mu}(v+c,A)=\mathcal{F}^{\mu}(v,A) for every cc\in\mathbb{R};

  3. (iii)

    μ(v,)\mathcal{F}^{\mu}(v,\cdot) is the restriction to 𝒪\mathcal{O} of a Radon measure;

  4. (iv)

    μ(,A)\mathcal{F}^{\mu}(\cdot,A) is L2(Ω,μ)L^{2}(\Omega,\mu)-lower semicontinuous;

  5. (v)

    μ(,A)\mathcal{F}^{\mu}(\cdot,A) is local, which means μ(v,A)=μ(w,A)\mathcal{F}^{\mu}(v,A)=\mathcal{F}^{\mu}(w,A) if v=wv=w μ\mu-a.e. on AA;

  6. (vi)

    μ(v,A)\mathcal{F}^{\mu}(v,A) satisfies the growth condition:

    0μ(v,A)CA|v|2dμ,0\leq\mathcal{F}^{\mu}(v,A)\leq C\int_{A}\left|\nabla v\right|^{2}\mathop{}\!\mathrm{d}\mu,

    with some C>0C>0.

  7. (vii)

    μ(v,A)\mathcal{F}^{\mu}(v,A) has the integral representation

    μ(v,A)=Af(x,v)dμ\mathcal{F}^{\mu}(v,A)=\int_{A}f\left(x,\nabla v\right)\mathop{}\!\mathrm{d}\mu

    where v|AH1(A,μ)v|_{A}\in H^{1}(A,\mu).

Proof.

(i) Since supμ\mathcal{F}_{\sup}^{\mu} is inner regular, we conclude by definition that =supμ\mathcal{F}=\mathcal{F}_{\sup}^{\mu}.

(ii) It is easily seen that the equality holds for any h>0h>0 and arbitrary μh,vh,A\mu^{h},v^{h},A, and cc\in\mathbb{R}.

To conclude (iii) it is enough to show that supμ\mathcal{F}_{\sup}^{\mu} is subadditive, superadditive, and inner regular on 𝒪\mathcal{O} (see, for instance, [14, Theorem 14.23]). Proposition 5.9 provides subadditivity and inner regularity, and here we only comment on superadditivity. By definition of hμ\mathcal{F}_{h}^{\mu}, for any A,B𝒪A,B\in\mathcal{O} such that AB=A\cap B=\emptyset and dist(A,B)>0\text{dist}(A,B)>0 there exists small enough h0>0h_{0}>0 such that for any h<h0h<h_{0}:

hμ(v,AB)hμ(v,A)+hμ(v,B).\mathcal{F}^{\mu}_{h}(v,A\cup B)\geq\mathcal{F}^{\mu}_{h}(v,A)+\mathcal{F}^{\mu}_{h}(v,B).

If dist(A,B)=0\text{dist}(A,B)=0, then the required property follows from inner regularity.

Properties (iv)-(vi) directly follow from (i) and Proposition 5.9.

(vii) Properties (ii)-(vi) allow to conclude the integral representation [14, Theorem 20.1]. ∎

Remark 5.11.

The integrand f(x,ξ)f(x,\xi) can be obtained for all ξd\xi\in\mathbb{R}^{d} and a.e. xΩx\in\Omega as

(5.5) f(x,ξ)=μ(φξ,Qε(x))|Qε(x)|,f(x,\xi)=\frac{\mathcal{F}^{\mu}(\varphi_{\xi},Q_{\varepsilon}(x))}{|Q_{\varepsilon}(x)|},

where φξ(z)=ξ,z\varphi_{\xi}(z)=\langle\xi,z\rangle (for details see [10, Remark 4.5]).

Upper bound for the integral representation

We derive an upper bound for ff using the representation (5.5). For a fixed ξd\xi\in\mathbb{R}^{d} the projection of φξ\varphi_{\xi} on 𝒯h\mathcal{T}^{h} is

hφξ=K𝒯hξ,xK 1KwithxK=\intbarKxdx.\mathbb{P}_{h}\varphi_{\xi}=\sum_{K\in\mathcal{T}^{h}}\langle\xi,x_{K}\rangle\,\mathbbm{1}_{K}\quad\text{with}\quad x_{K}=\intbar_{K}x\mathop{}\!\mathrm{d}x.

Substituting hφξ\mathbb{P}_{h}\varphi_{\xi} into hμ\mathcal{F}^{\mu}_{h} yields

hμ(hφξ,A)\displaystyle\mathcal{F}^{\mu}_{h}(\mathbb{P}_{h}\varphi_{\xi},A) =(K,L)Σh|A|ξ,xLxK|2κh(K,L)μh(K)\displaystyle=\sum_{(K,L)\in\Sigma^{h}|_{A}}\left|\langle\xi,x_{L}-x_{K}\rangle\right|^{2}\kappa^{h}(K,L)\mu^{h}(K)
=K𝒯h|Aξ,L𝒯Kh|Aκh(K,L)(xLxK)(xLxK)ξμh(K)\displaystyle=\sum_{K\in\mathcal{T}^{h}|_{A}}\Big{\langle}\xi,\sum_{L\in\mathcal{T}^{h}_{K}|_{A}}\kappa^{h}(K,L)(x_{L}-x_{K})\otimes(x_{L}-x_{K})\xi\Big{\rangle}\,\mu^{h}(K)
=K𝒯h|AKξ,𝕋h(x)ξμ^h(dx)=ξ,Ω𝕋h(x)𝟙A𝒯hμ^h(dx)ξ,\displaystyle=\sum_{K\in\mathcal{T}^{h}|_{A}}\int_{K}\langle\xi,\mathbb{T}^{h}(x)\,\xi\rangle\,\hat{\mu}^{h}(\mathop{}\!\mathrm{d}x)=\Big{\langle}\xi,\int_{\Omega}\mathbb{T}^{h}(x)\mathbbm{1}_{A_{\mathcal{T}^{h}}}\hat{\mu}^{h}(\mathop{}\!\mathrm{d}x)\,\xi\Big{\rangle},

where we denoted by 𝕋h\mathbb{T}^{h} the tensor

(5.6) 𝕋h(x):=K𝒯h𝟙K(x)L𝒯Khκh(K,L)(xLxK)(xLxK).\mathbb{T}^{h}(x):=\sum_{K\in\mathcal{T}^{h}}\mathbbm{1}_{K}(x)\sum_{L\in\mathcal{T}^{h}_{K}}\kappa^{h}(K,L)(x_{L}-x_{K})\otimes(x_{L}-x_{K}).

When passing h0h\to 0, we expect 𝕋h\mathbb{T}^{h} converge to the diffusion tensor, therefore, we establish a number of useful properties of 𝕋h\mathbb{T}^{h}.

Lemma 5.12 (Properties of 𝕋h\mathbb{T}^{h}).

The diffusion tensor (5.6) has the following properties:

  1. (i)

    𝕋h(x)\mathbb{T}^{h}(x) is symmetric and positive-definite for any xΩx\in\Omega;

  2. (ii)

    {𝕋h}h>0\{\mathbb{T}^{h}\}_{h>0} is bounded in L(Ω;d×d)L^{\infty}(\Omega;\mathbb{R}^{d\times d}):

    for all the components 𝕋ijh\mathbb{T}^{h}_{ij} it holds that suph>0𝕋ijhL(Ω)<\displaystyle\sup_{h>0}\|\mathbb{T}^{h}_{ij}\|_{L^{\infty}(\Omega)}<\infty;
  3. (iii)

    {𝕋h}h>0\{\mathbb{T}^{h}\}_{h>0} has a weakly-* limit in the σ(L,L1)\sigma(L^{\infty},L^{1}) topology, i.e. there exist a subsequence and a tensor 𝕋L(Ω;d×d)\mathbb{T}\in L^{\infty}(\Omega;\mathbb{R}^{d\times d}) such that

    limh0Ω𝕋ijhfdx=Ω𝕋ijfdxfor all fL1(Ω).\lim_{h\to 0}\int_{\Omega}\mathbb{T}^{h}_{ij}f\mathop{}\!\mathrm{d}x=\int_{\Omega}\mathbb{T}_{ij}f\mathop{}\!\mathrm{d}x\qquad\text{for all }f\in L^{1}(\Omega).
Proof.

(i) Symmetry and positive-definiteness follow directly from the definition.

(ii) Fix any xΩx\in\Omega and consider the tensor 𝕋h\mathbb{T}^{h} component-wise:

𝕋ijh(x)=L𝒯Khκh(K,L)(xLixKi)(xLjxKj).\mathbb{T}^{h}_{ij}(x)=\sum_{L\in\mathcal{T}^{h}_{K}}\kappa^{h}(K,L)(x_{L}^{i}-x_{K}^{i})(x_{L}^{j}-x_{K}^{j}).

The bound |xLixKi|2h|x_{L}^{i}-x_{K}^{i}|\leq 2h and (UB) gives |𝕋ijh(x)|2h2L𝒯Khκh(K,L)2Cκ|\mathbb{T}^{h}_{ij}(x)|\leq 2h^{2}\sum_{L\in\mathcal{T}^{h}_{K}}\kappa^{h}(K,L)\leq 2C_{\kappa}. Consequently, suph>0𝕋ijhL(Ω)2Cκ\sup_{h>0}\|\mathbb{T}^{h}_{ij}\|_{L^{\infty}(\Omega)}\leq 2C_{\kappa}.

(iii) The weak-* convergence follows from (ii) and the duality of L1L^{1} and LL^{\infty} (see for instance [2, Theorem 8.5, Examples 8.6(1)]). ∎

The next lemma provides an upper bound for the integral representation of μ\mathcal{F}^{\mu}

Lemma 5.13.

Let μ:L2(Ω,μ)×𝒪[0,+]\mathcal{F}^{\mu}:L^{2}(\Omega,\mu)\times\mathcal{O}\to[0,+\infty] be the Γ¯\overline{\Gamma}-limit of {~hμ}h>0\{\tilde{\mathcal{F}}^{\mu}_{h}\}_{h>0}, then for every vH1(Ω,μ)v\in H^{1}(\Omega,\mu) and every A𝒪A\in\mathcal{O},

μ(v,A)Av,𝕋vdμ,\mathcal{F}^{\mu}(v,A)\leq\int_{A}\langle\nabla v,\mathbb{T}\nabla v\rangle\mathop{}\!\mathrm{d}\mu\,,

where 𝕋\mathbb{T} is defined in Lemma 5.12(iii).

Proof.

Lemma 5.12(iii) gives, in particular, that if 𝕋h𝕋\mathbb{T}^{h}\rightharpoonup^{*}\mathbb{T} in L(Ω)L^{\infty}(\Omega) for some (not relabelled) subsequence, then

limh0Ω𝕋h(x)𝟙A𝒯hμ^h(dx)=A𝕋(x)μ(dx),\lim_{h\to 0}\int_{\Omega}\mathbb{T}^{h}(x)\mathbbm{1}_{A_{\mathcal{T}^{h}}}\hat{\mu}^{h}(\mathop{}\!\mathrm{d}x)=\int_{A}\mathbb{T}(x)\mu(\mathop{}\!\mathrm{d}x),

which holds due to Proposition 5.1 and the generalized dominated convergence theorem. Therefore,

μ(φξ,Qε(x))lim infh0~hμ(hφξ,Qε(x))ξ,Qε(x)𝕋(z)μ(dz)ξ.\mathcal{F}^{\mu}(\varphi_{\xi},Q_{\varepsilon}(x))\leq\liminf_{h\to 0}\tilde{\mathcal{F}}^{\mu}_{h}(\mathbb{P}_{h}\varphi_{\xi},Q_{\varepsilon}(x))\leq\Big{\langle}\xi,\int_{Q_{\varepsilon}(x)}\mathbb{T}(z)\mu(\mathop{}\!\mathrm{d}z)\,\xi\Big{\rangle}.

The representation formula (5.5) yields

f(x,ξ)ξ,limε0\intbarQε(x)𝕋(z)μ(dz)ξ=ξ,𝕋(x)dμdd(x)ξ.f(x,\xi)\leq\Big{\langle}\xi,~{}\lim_{\varepsilon\to 0}\intbar_{Q_{\varepsilon}(x)}\mathbb{T}(z)\mu(\mathop{}\!\mathrm{d}z)\,\xi\Big{\rangle}=\Big{\langle}\xi,~{}\mathbb{T}(x)\frac{\mathop{}\!\mathrm{d}\mu}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}(x)\,\xi\Big{\rangle}.

Thus, for and vH1(Ω,μ)v\in H^{1}(\Omega,\mu) and A𝒪A\in\mathcal{O},

μ(v,A)Av,𝕋vdμ,\mathcal{F}^{\mu}(v,A)\leq\int_{A}\langle\nabla v,\mathbb{T}\nabla v\rangle\,\mathop{}\!\mathrm{d}\mu\,,

as required. ∎

5.2. Dual dissipation potential

Recall that the dual dissipation potential has the form:

h(ρh,¯φh)=12(K,L)ΣhΨ((¯φh)(K,L))uh(K)uh(L)ϑh(K,L),where uh=dρhdπh,\displaystyle\mathcal{R}^{*}_{h}(\rho^{h},\overline{\nabla}\varphi^{h})=\frac{1}{2}\sum_{(K,L)\in\Sigma^{h}}\Psi^{*}\left((\overline{\nabla}\varphi^{h})(K,L)\right)\sqrt{u^{h}(K)u^{h}(L)}\,\vartheta^{h}(K,L),\quad\text{where }u^{h}=\frac{\mathop{}\!\mathrm{d}\rho^{h}}{\mathop{}\!\mathrm{d}\pi^{h}},

with Ψ(ξ)=4(cosh(ξ/2)1)\Psi^{*}(\xi)=4\left(\cosh{(\xi/2)}-1\right).

In Lemma 5.13, we derived an upper bound for the integral representation of the Γ¯\overline{\Gamma}-limit of {~hρ}h>0\{\tilde{\mathcal{F}}^{\rho}_{h}\}_{h>0}. We now prove that the same bound applies asymptotically to h\mathcal{R}^{*}_{h} for some specific choice of approximating functions φh\varphi^{h}.

Lemma 5.14.

Let φ𝒞b2(Ω)\varphi\in\mathcal{C}_{b}^{2}(\Omega) and assume that {ρh}h>0\{\rho^{h}\}_{h>0} is a family of probability measures on 𝒯h\mathcal{T}^{h} such that dρ^h/dddρ/dd\mathop{}\!\mathrm{d}\hat{\rho}^{h}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\to\mathop{}\!\mathrm{d}\rho/\mathop{}\!\mathrm{d}\mathscr{L}^{d} in L1(Ω)L^{1}(\Omega) (cf. Section 5.1(ii)). Moreover, let {φh}h>0\{\varphi^{h}\}_{h>0} be the family of discrete functions on {𝒯h}h>0\{\mathcal{T}^{h}\}_{h>0} defined by φh(K):=φ(xK)\varphi^{h}(K):=\varphi(x_{K}) for K𝒯hK\in\mathcal{T}^{h}.

Then 𝕃hφhφ\mathbb{L}_{h}\varphi^{h}\to\varphi in L2(Ω,ρ)L^{2}(\Omega,\rho), and

lim suph0h(ρh,¯φh)14Ωφ,𝕋φdρ.\limsup_{h\to 0}\mathcal{R}^{*}_{h}(\rho^{h},\overline{\nabla}\varphi^{h})\leq\frac{1}{4}\int_{\Omega}\langle\nabla\varphi,~{}\mathbb{T}\nabla\varphi\rangle\mathop{}\!\mathrm{d}\rho\,.
Proof.

We first observe that 𝕃hφh=:φ^hφ\mathbb{L}_{h}\varphi^{h}=:\hat{\varphi}^{h}\to\varphi in L2(Ω,ρ)L^{2}(\Omega,\rho). This follows directly from estimate

Ω|φ^h(x)φ(x)|2ρ(dx)\displaystyle\int_{\Omega}|\hat{\varphi}^{h}(x)-\varphi(x)|^{2}\rho(\mathop{}\!\mathrm{d}x) =K𝒯hK|φ(xK)φ(x)|2ρ(dx)\displaystyle=\sum_{K\in\mathcal{T}^{h}}\int_{K}|\varphi(x_{K})-\varphi(x)|^{2}\rho(\mathop{}\!\mathrm{d}x)
K𝒯hK(φL2h2+o(h2))ρ(dx)\displaystyle\leq\sum_{K\in\mathcal{T}^{h}}\int_{K}\left(\|\nabla\varphi\|^{2}_{L^{\infty}}h^{2}+o(h^{2})\right)\rho(\mathop{}\!\mathrm{d}x)
φL2h2+o(h2).\displaystyle\leq\|\nabla\varphi\|^{2}_{L^{\infty}}h^{2}+o(h^{2}).

Now, we show that {φ^h}h>0\{\hat{\varphi}^{h}\}_{h>0} realises the upper bound for the Γ¯\overline{\Gamma}-limit ρ\mathcal{F}^{\rho} proven in Lemma 5.13, i.e.

lim suph0~hρ(φ^h,A)Aφ,𝕋φdρ.\limsup_{h\to 0}\tilde{\mathcal{F}}^{\rho}_{h}(\hat{\varphi}^{h},A)\leq\int_{A}\langle\nabla\varphi,~{}\mathbb{T}\nabla\varphi\rangle\mathop{}\!\mathrm{d}\rho.

Since φ\varphi is smooth, the discrete gradient for φh\varphi^{h} can be approximated by

(¯φh)(K,L)=φ(xL)φ(xK)\displaystyle(\overline{\nabla}\varphi^{h})(K,L)=\varphi(x_{L})-\varphi(x_{K}) =01(φ)(xK+τ(xLxK)),xLxKdτ\displaystyle=\int_{0}^{1}\langle(\nabla\varphi)(x_{K}+\tau(x_{L}-x_{K})),~{}x_{L}-x_{K}\rangle\mathop{}\!\mathrm{d}\tau
=(φ)(xK),xLxK+o(h).\displaystyle=\langle(\nabla\varphi)(x_{K}),~{}x_{L}-x_{K}\rangle+o(h).

Moreover, the difference between (φ)(xK)(\nabla\varphi)(x_{K}) and \intbarK(φ)(x)dx\intbar_{K}(\nabla\varphi)(x)\mathop{}\!\mathrm{d}x is of a small order:

\intbarK(iφ)(x)dx(iφ)(xK)\displaystyle\intbar_{K}(\partial_{i}\varphi)(x)\mathop{}\!\mathrm{d}x-(\partial_{i}\varphi)(x_{K}) =\intbarK01(iφ)(xK+τ(xxK)),xxKdτdx\displaystyle=\intbar_{K}\int_{0}^{1}\langle(\nabla\partial_{i}\varphi)(x_{K}+\tau(x-x_{K})),~{}x-x_{K}\rangle\mathop{}\!\mathrm{d}\tau\mathop{}\!\mathrm{d}x
=(iφ)(xK),\intbarK(xxK)dx+o(h)=o(h),\displaystyle=\left\langle(\nabla\partial_{i}\varphi)(x_{K}),\intbar_{K}(x-x_{K})\mathop{}\!\mathrm{d}x\right\rangle+o(h)=o(h),

which implies that

(¯φh)(K,L)=\intbarK(φ)(x),xLxKdx+o(h).(\overline{\nabla}\varphi^{h})(K,L)=\intbar_{K}\langle(\nabla\varphi)(x),~{}x_{L}-x_{K}\rangle\mathop{}\!\mathrm{d}x+o(h).

Substituting ¯φh\overline{\nabla}\varphi^{h} into hρ\mathcal{F}^{\rho}_{h} yields:

hρ(φh,A)\displaystyle\mathcal{F}^{\rho}_{h}(\varphi^{h},A) =(K,L)Σh|A|(¯φh)(K,L)|2κh(K,L)ρh(K)\displaystyle=\sum_{(K,L)\in\Sigma^{h}|_{A}}\left|(\overline{\nabla}\varphi^{h})(K,L)\right|^{2}\kappa^{h}(K,L)\rho^{h}(K)
=(K,L)Σh|A|(φ)(x),xLxKdx|2κh(K,L)ρh(K)+o(1)|h0,\displaystyle=\sum_{(K,L)\in\Sigma^{h}|_{A}}\left|\langle(\nabla\varphi)(x),~{}x_{L}-x_{K}\rangle\mathop{}\!\mathrm{d}x\right|^{2}\kappa^{h}(K,L)\rho^{h}(K)+o(1)|_{h\to 0},

where we used (UB). Applying Jensen’s inequality we get:

hρ(φh,A)\displaystyle\mathcal{F}^{\rho}_{h}(\varphi^{h},A) (K,L)Σh|A\intbarK|(φ)(x),xLxKdx|2dxκh(K,L)ρh(K)+o(1)|h0\displaystyle\leq\sum_{(K,L)\in\Sigma^{h}|_{A}}\intbar_{K}\left|\langle(\nabla\varphi)(x),~{}x_{L}-x_{K}\rangle\mathop{}\!\mathrm{d}x\right|^{2}\mathop{}\!\mathrm{d}x\,\kappa^{h}(K,L)\rho^{h}(K)+o(1)|_{h\to 0}
A𝒯h(φ)(x),𝕋h(x)(φ)(x)ρ^h(dx)+o(1)|h0.\displaystyle\leq\int_{A_{\mathcal{T}^{h}}}\big{\langle}(\nabla\varphi)(x),~{}\mathbb{T}^{h}(x)(\nabla\varphi)(x)\big{\rangle}\hat{\rho}^{h}(\mathop{}\!\mathrm{d}x)+o(1)|_{h\to 0}.

By Lemma 5.12(iii) and Proposition 5.1 one can pass h0h\to 0 on the right-hand side to obtain:

lim suph0hρ(φh,A)A(φ)(x),𝕋(x)(φ)(x)ρ(x)dx.\displaystyle\limsup_{h\to 0}\mathcal{F}^{\rho}_{h}(\varphi^{h},A)\leq\int_{A}\big{\langle}(\nabla\varphi)(x),~{}\mathbb{T}(x)(\nabla\varphi)(x)\big{\rangle}\rho(x)\mathop{}\!\mathrm{d}x.

The last step we need to take is to show that h(ρh,¯φh)=hρ(φh)+o(1)|h0\mathcal{R}^{*}_{h}(\rho^{h},\overline{\nabla}\varphi^{h})=\mathcal{F}_{h}^{\rho}(\varphi^{h})+o(1)|_{h\to 0}. We consider the expansion of h\mathcal{R}^{*}_{h}:

h(ρh,¯φh)\displaystyle\mathcal{R}^{*}_{h}(\rho^{h},\overline{\nabla}\varphi^{h}) =12(K,L)ΣhΨ((¯φh)(K,L))uh(K)uh(L)ϑh(K,L)\displaystyle=\frac{1}{2}\sum_{(K,L)\in\Sigma^{h}}\Psi^{*}\left((\overline{\nabla}\varphi^{h})(K,L)\right)\sqrt{u^{h}(K)u^{h}(L)}\,\vartheta^{h}(K,L)
14(K,L)Σh(|(¯φh)(K,L)|2+g((¯φh)(K,L)))κh(K,L)ρh(K)\displaystyle\leq\frac{1}{4}\sum_{(K,L)\in\Sigma^{h}}\left(\left|(\overline{\nabla}\varphi^{h})(K,L)\right|^{2}+g\left((\overline{\nabla}\varphi^{h})(K,L)\right)\right)\kappa^{h}(K,L)\rho^{h}(K)
14hρ(φh)+14(K,L)Σhg((¯φh)(K,L))κh(K,L)ρh(K),\displaystyle\leq\frac{1}{4}\mathcal{F}^{\rho}_{h}(\varphi^{h})+\frac{1}{4}\sum_{(K,L)\in\Sigma^{h}}g\Bigl{(}(\overline{\nabla}\varphi^{h})(K,L)\Bigr{)}\,\kappa^{h}(K,L)\rho^{h}(K),

where

g(r)=Ψ(r)r22=k=24(2k)!(r2)2k=O(r4).g(r)=\Psi^{*}(r)-\frac{r^{2}}{2}=\displaystyle\sum_{k=2}^{\infty}\frac{4}{(2k)!}\left(\frac{r}{2}\right)^{2k}=O(r^{4}).

Since |(¯φh)(K,L)|CφLh|(\overline{\nabla}\varphi^{h})(K,L)|\leq C\|\nabla\varphi\|_{L^{\infty}}h and recalling (UB) once again, we conclude the proof. ∎

5.3. Fisher information

In this section, we prove the Γ\Gamma-convergence for the family of discrete Fisher information {𝒟h}h>0\{\mathcal{D}_{h}\}_{h>0} defined as

(5.7) 𝒟h(ρh)=(K,L)Σh|¯uh(K,L)|2ϑh(K,L), with uh=dρhdπh.\mathcal{D}_{h}(\rho^{h})=\sum_{(K,L)\in\Sigma^{h}}\left|\overline{\nabla}\sqrt{u^{h}}(K,L)\right|^{2}\vartheta^{h}(K,L),\qquad\text{ with }u^{h}=\frac{\mathop{}\!\mathrm{d}\rho^{h}}{\mathop{}\!\mathrm{d}\pi^{h}}.

We state the main result of this section.

Theorem 5.15.

Up to passing to a subsequence, the family of functionals {𝒟h}h>0\{\mathcal{D}_{h}\}_{h>0} has a Γ\Gamma-limit 𝒟\mathcal{D} w.r.t. the L2L^{2}-topology taking the form

(5.8) 𝒟(ρ)={Ωu,𝕋udπif dρdπ=:uH1(Ω),+otherwise,\mathcal{D}(\rho)=\begin{cases}\displaystyle\int_{\Omega}\big{\langle}\nabla\sqrt{u},\mathbb{T}\nabla\sqrt{u}\big{\rangle}\mathop{}\!\mathrm{d}\pi&\text{if }\sqrt{\frac{\mathop{}\!\mathrm{d}\rho}{\mathop{}\!\mathrm{d}\pi}}=:\sqrt{u}\in H^{1}(\Omega),\\ +\infty&\text{otherwise,}\end{cases}

where 𝕋\mathbb{T} defined in Lemma 5.12.

Remark 5.16.

Since we assume that the densities {dπ^h/dd}h>0\{\mathop{}\!\mathrm{d}\hat{\pi}^{h}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\}_{h>0} are uniformly bounded from above and away from 0, and dπ^h/dddπ/dd\mathop{}\!\mathrm{d}\hat{\pi}^{h}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\to\mathop{}\!\mathrm{d}\pi/\mathop{}\!\mathrm{d}\mathscr{L}^{d} in L1(Ω)L^{1}(\Omega), then π\pi is bounded in the same way. Consequently, the norms in Lp(A,π)L^{p}(A,\pi) and Lp(A)L^{p}(A) are equivalent.

In Theorem 5.15 we implicitly consider 𝒟\mathcal{D} to depend on dρ/dπL2(Ω)\sqrt{\mathop{}\!\mathrm{d}\rho/\mathop{}\!\mathrm{d}\pi}\in L^{2}(\Omega) for all ρπ\rho\ll\pi and take the Γ\Gamma-limit in the corresponding topology. To simplify the notation we set vh:=uhv^{h}:=\sqrt{u^{h}} and consider again the localized functional:

hπ(vh,A)=(K,L)Σh|A|¯vh(K,L)|2ϑh(K,L).\mathcal{F}^{\pi}_{h}(v^{h},A)=\sum_{(K,L)\in\Sigma^{h}|_{A}}\left|\overline{\nabla}v^{h}(K,L)\right|^{2}\vartheta^{h}(K,L).

Notice that hπ(vh,Ω)=𝒟h(uhπh)\mathcal{F}^{\pi}_{h}(v^{h},\Omega)=\mathcal{D}_{h}(u^{h}\pi^{h}). In what follows we set π:=Γ\mathcal{F}^{\pi}:=\Gamma-limhπ\lim\mathcal{F}^{\pi}_{h}, infπ:=Γ\mathcal{F}^{\pi}_{\inf}:=\Gamma-lim infhπ\liminf\mathcal{F}^{\pi}_{h}, and supπ:=Γ\mathcal{F}^{\pi}_{\sup}:=\Gamma-lim suphπ\limsup\mathcal{F}^{\pi}_{h}.

Proposition 5.10 provides the existence of an integral representation:

π(v,A)=Af(x,v)dπ,vH1(Ω).\mathcal{F}^{\pi}(v,A)=\int_{A}f(x,\nabla v)\mathop{}\!\mathrm{d}\pi,\qquad v\in H^{1}(\Omega).

Unlike in Section 5.2, we are interested in the exact Γ\Gamma-limit for hπ\mathcal{F}^{\pi}_{h}. In fact, Proposition 5.10 provides almost all necessary properties to apply a general representation theorem [8, Theorem 2], except for a Sobolev lower bound. Therefore, we show the lower bound in Lemma 5.17, and then proceed to the representation.

Lemma 5.17.

For any vH1(Ω)v\in H^{1}(\Omega) and A𝒪A\in\mathcal{O} it holds that:

supπ(v,A)ClCζπmaxA|v|2dπClπminCζπmaxA|v|2dx.\mathcal{F}^{\pi}_{\sup}(v,A)\geq\frac{C_{l}}{C_{\zeta}\pi_{\max}}\int_{A}\left|\nabla v\right|^{2}\mathop{}\!\mathrm{d}\pi\geq\frac{C_{l}\pi_{\min}}{C_{\zeta}\pi_{\max}}\int_{A}\left|\nabla v\right|^{2}\mathop{}\!\mathrm{d}x.
Proof.

Let {vh}h>0L2(Ω)\{v_{h}\}_{h>0}\in L^{2}(\Omega) be a sequence with vhvv_{h}\to v in L2(Ω)L^{2}(\Omega) such that

supπ(v,A)=lim suph0~hπ(vh,A).\mathcal{F}^{\pi}_{\sup}(v,A)=\limsup_{h\to 0}\tilde{\mathcal{F}}^{\pi}_{h}(v_{h},A).

In particular, vhPC(𝒯h|A)v_{h}\in PC(\mathcal{T}^{h}|_{A}) for all h>0h>0.

To prove the lower bound, one can repeat some elements of the proof of Lemma 4.7. We fix an arbitrary ε>0\varepsilon>0 and denote Aε:={xA:dist(x,A)>ε}A_{\varepsilon}:=\left\{x\in A\,:\,\text{dist}(x,\partial A)>\varepsilon\right\}. Let ηd\eta\in\mathbb{R}^{d} be such that |η|<ε|\eta|<\varepsilon, then

Aε|vh(x+η)\displaystyle\int_{A_{\varepsilon}}|v_{h}(x+\eta) vh(x)|2dx=K𝒯h|AεKAε|vh(x+η)vh(x)|2dx\displaystyle-v_{h}(x)|^{2}\mathop{}\!\mathrm{d}x=\sum_{K\in\mathcal{T}^{h}|_{A_{\varepsilon}}}\int_{K\cap A_{\varepsilon}}\left|v_{h}(x+\eta)-v_{h}(x)\right|^{2}\,\mathop{}\!\mathrm{d}x
K𝒯h|AεKAεni=0n1|vh(Ki+1)vh(Ki)|2dx\displaystyle\leq\sum_{K\in\mathcal{T}^{h}|_{A_{\varepsilon}}}\int_{K\cap A_{\varepsilon}}n\sum_{i=0}^{n-1}\left|v_{h}(K_{i+1})-v_{h}(K_{i})\right|^{2}\,\mathop{}\!\mathrm{d}x
Cζ|η|hK𝒯h|AεKAε(M,L)Σh|vh(L)vh(M)|2𝟙CylΣh(x,η)(M,L)dx.\displaystyle\leq\frac{C_{\zeta}|\eta|}{h}\sum_{K\in\mathcal{T}^{h}|_{A_{\varepsilon}}}\int_{K\cap A_{\varepsilon}}\sum_{(M,L)\in\Sigma^{h}}\left|v_{h}(L)-v_{h}(M)\right|^{2}\mathbbm{1}_{\text{Cyl}_{\Sigma^{h}}(x,\eta)}(M,L)\,\mathop{}\!\mathrm{d}x.

where we used Proposition 4.6 to assert that nCζ|η|/hn\leq C_{\zeta}|\eta|/h for some constant Cζ>0C_{\zeta}>0, independent of xx. Now one can repeat the transformations from the proof Lemma 4.7 to obtain:

Aε|vh(x+η)\displaystyle\int_{A_{\varepsilon}}|v_{h}(x+\eta) vh(x)|2dxCζCl|η|2𝒟h(vh,Aε).\displaystyle-v_{h}(x)|^{2}\mathop{}\!\mathrm{d}x\leq\frac{C_{\zeta}}{C_{l}}|\eta|^{2}\mathcal{D}_{h}(v_{h},A_{\varepsilon}).

Passing to the limit superior as h0h\to 0 then yields

supπ(v,Aε)ClCζv(+η)vL2(Aε)2|η|2.\mathcal{F}^{\pi}_{\sup}(v,A_{\varepsilon})\geq\frac{C_{l}}{C_{\zeta}}\frac{\left\|v(\cdot+\eta)-v\right\|^{2}_{L^{2}(A_{\varepsilon})}}{|\eta|^{2}}.

For vH1(Ω)v\in H^{1}(\Omega), passing |η|0|\eta|\to 0 yields

supπ(v,Aε)ClCζAε|v|2dxClCζπmaxAε|v|2π(dx).\mathcal{F}^{\pi}_{\sup}(v,A_{\varepsilon})\geq\frac{C_{l}}{C_{\zeta}}\int_{A_{\varepsilon}}\left|\nabla v\right|^{2}\mathop{}\!\mathrm{d}x\geq\frac{C_{l}}{C_{\zeta}\pi_{\max}}\int_{A_{\varepsilon}}\left|\nabla v\right|^{2}\pi(\mathop{}\!\mathrm{d}x).

Since supπ\mathcal{F}^{\pi}_{\sup} is inner regular (Proposition 5.9), then

supπ(v,A)=supε>0supπ(v,Aε)ClCζπmaxA|v|2π(dx),\mathcal{F}^{\pi}_{\sup}(v,A)=\sup_{\varepsilon>0}\mathcal{F}^{\pi}_{\sup}(v,A_{\varepsilon})\geq\frac{C_{l}}{C_{\zeta}\pi_{\max}}\int_{A}\left|\nabla v\right|^{2}\pi(\mathop{}\!\mathrm{d}x)\,,

where the inequality follows from the monotone convergence theorem. ∎

Now we are in the position to use the following proposition from [8, Theorem 2].

Proposition 5.18.

Let :H1(Ω)×𝒪[0,+]\mathcal{F}:H^{1}(\Omega)\times\mathcal{O}\to[0,+\infty] be a functional satisfying:

  1. (i)

    (v,)\mathcal{F}(v,\cdot) is the restriction to 𝒪\mathcal{O} of a Radon measure;

  2. (ii)

    (,A)\mathcal{F}(\cdot,A) is L2(Ω)L^{2}(\Omega) lower semicontinuous;

  3. (iii)

    (,A)\mathcal{F}(\cdot,A) is local, which means (v,A)=(w,A)\mathcal{F}(v,A)=\mathcal{F}(w,A) if v=wv=w d\mathscr{L}^{d}-a.e. on AA;

  4. (iv)

    (v+c,A)=(v,A)\mathcal{F}(v+c,A)=\mathcal{F}(v,A) for every cc\in\mathbb{R};

  5. (v)

    (v,A)\mathcal{F}(v,A) satisfies the growth condition:

    1CA|v|2dx(v,A)CA(1+|v|2)dx\frac{1}{C}\int_{A}\left|\nabla v\right|^{2}\mathop{}\!\mathrm{d}x\leq\mathcal{F}(v,A)\leq C\int_{A}\left(1+\left|\nabla v\right|^{2}\right)\mathop{}\!\mathrm{d}x

    for some C>0C>0.

Then for every vH1(Ω)v\in H^{1}(\Omega) and A𝒪A\in\mathcal{O}

(v,A)=Af(x,v,v)dx,\mathcal{F}(v,A)=\int_{A}f(x,v,\nabla v)\mathop{}\!\mathrm{d}x,

where

f(x0,v0,ξ):=lim supε0+1|Qε(x)|M(v0+ξ(x),Qε(x))f(x_{0},v_{0},\xi):=\limsup_{\varepsilon\to 0+}\frac{1}{\left|Q_{\varepsilon}(x)\right|}\text{M}(\langle v_{0}+\xi\cdot(\cdot-x),Q_{\varepsilon}(x))

for all x0Ωx_{0}\in\Omega, v0dv_{0}\in\mathbb{R}^{d}, ξd\xi\in\mathbb{R}^{d}, and where, for (v,A)H1(Ω)×𝒪(v,A)\in H^{1}(\Omega)\times\mathcal{O},

M(v,A):=inf{(w,A):wH1(A) with v=w in a neighborhood of A}.\text{M}(v,A):=\inf\left\{\mathcal{F}(w,A)~{}:~{}w\in H^{1}(A)\text{ with }v=w\text{ in a neighborhood of }\partial A\right\}.

Applying Proposition 5.18 to our setting, we notice that ff does not depend explicitly on vv (since μ(v+c,A)=μ(v,A)\mathcal{F}^{\mu}(v+c,A)=\mathcal{F}^{\mu}(v,A) for every cc\in\mathbb{R} Proposition 5.10). This gives

(5.9) π(v,A)=Afπ(x,v)dxwith fπ(x,ξ)=lim supε01|Qε(x)|M(ξ(x),Qε(x)).\mathcal{F}^{\pi}(v,A)=\int_{A}f_{\pi}(x,\nabla v)\mathop{}\!\mathrm{d}x\quad\text{with }f_{\pi}(x,\xi)=\limsup_{\varepsilon\to 0}\frac{1}{\left|Q_{\varepsilon}(x)\right|}\text{M}(\xi\cdot(\cdot-x),Q_{\varepsilon}(x)).

The identification formula (5.9) suggests looking for the minimizer of ψπ(ψ,A)\psi\mapsto\mathcal{F}^{\pi}(\psi,A) w.r.t. the Dirichlet boundary condition. To relate this minimization problem to our discrete formulation, we follow similar approach as in [21], specifically, we define

Mh(φh,A):=infψh{hπ(ψh,A):ψh on 𝒯h|Awithψh=φh on 𝒯h|Ac}.\text{M}_{h}(\varphi^{h},A):=\inf_{\psi^{h}}\Bigl{\{}\mathcal{F}^{\pi}_{h}(\psi^{h},A)~{}:~{}\psi^{h}\text{ on }\mathcal{T}^{h}|_{A}\quad\text{with}\quad\psi^{h}=\varphi^{h}\text{ on }\mathcal{T}^{h}|_{A^{c}}\Bigr{\}}.

We also make use of the following definition of M

M(φ,A)=infψ{π(ψ,A):ψH1(A)withψφH01(A)},\text{M}(\varphi,A)=\inf_{\psi}\Bigl{\{}\mathcal{F}^{\pi}(\psi,A)~{}:~{}\psi\in H^{1}(A)\quad\text{with}\quad\psi-\varphi\in H^{1}_{0}(A)\Bigr{\}},

that was proven to be equivalent to the one from Proposition 5.18 in [21, Remark 7.4].

Note that the Γ\Gamma-convergence of hπ\mathcal{F}^{\pi}_{h} to π\mathcal{F}^{\pi} does not suffice to conclude the convergence of MhM_{h} to MM. Hence, we define

π,φ(v,A):={π(v,A)if vφH01(A),+otherwise,\mathcal{F}^{\pi,\varphi}(v,A):=\begin{cases}\mathcal{F}^{\pi}(v,A)&\text{if }v-\varphi\in H^{1}_{0}(A),\\ +\infty&\text{otherwise,}\end{cases}

and the corresponding discrete counterpart of π,φ\mathcal{F}^{\pi,\varphi} is

hπ,φ(vh,A):={hπ(vh,A)if vh=hφ=:φh on A𝒯hc,+otherwise.\mathcal{F}^{\pi,\varphi}_{h}(v^{h},A):=\begin{cases}\mathcal{F}^{\pi}_{h}(v^{h},A)&\text{if }v^{h}=\mathbb{P}_{h}\varphi=:\varphi^{h}\text{ on }A^{c}_{\mathcal{T}^{h}},\\ +\infty&\text{otherwise.}\end{cases}

Similarly to [21, Lemma 7.9], the next proposition claims that hπ,φ(,A)Γπ,φ(,A)\mathcal{F}^{\pi,\varphi}_{h}(\cdot,A)\xrightarrow{\Gamma}\mathcal{F}^{\pi,\varphi}(\cdot,A) for any A𝒪A\in\mathcal{O} with Lipschitz boundary. For completeness, we include the proof in Appendix A.

Proposition 5.19.

Let A𝒪A\in\mathcal{O} and φLip(Ω)\varphi\in\text{Lip}(\Omega). For any sequence {hπ,φ(,A)}h>0\{\mathcal{F}^{\pi,\varphi}_{h}(\cdot,A)\}_{h>0}, there exists a subsequence that Γ\Gamma-converges in the L2(Ω)L^{2}(\Omega)-topology to π,φ(,A)\mathcal{F}^{\pi,\varphi}(\cdot,A).

Now we comment on the assumptions on tessellations and kernels needed for proving the representation result. It suffices to use (Bϑ\vartheta) and (Aloc{}_{\text{loc}}). The last assumption (Aloc{}_{\text{loc}}) was not necessary for any preceding statements and its role here is to ensure that the discrete functions

φhx,ξ(K):=ξ,xKx for all K𝒯h\varphi_{h}^{x,\xi}(K):=\langle\xi,x_{K}-x\rangle\quad\text{ for all }K\in\mathcal{T}^{h}

are minimizers for hπ,φ(,Qε(x))\mathcal{F}^{\pi,\varphi}_{h}(\cdot,Q_{\varepsilon}(x)), where xK=\intbarKxdxx_{K}=\intbar_{K}x\mathop{}\!\mathrm{d}x. To relax (Aloc{}_{\text{loc}}) we can introduce an asymptotic assumption involving almost minimizers. Namely, we assume that

(AMin) limh0(hπ(φhx,ξ,Qε(x))Mh(φhx,ξ,Qε(x)))=0.\lim_{h\to 0}\left(\mathcal{F}^{\pi}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x))-\text{M}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x))\right)=0.

Finally, we state the representation result.

Theorem 5.20.

Let hπ:L2(Ω,π)×𝒪[0,+]\mathcal{F}^{\pi}_{h}:L^{2}(\Omega,\pi)\times\mathcal{O}\to[0,+\infty] be the Γ¯\overline{\Gamma}-limit of {~h}h>0\{\tilde{\mathcal{F}}_{h}\}_{h>0} defined as in Lemma 5.4, then the functional π(v,A)\mathcal{F}^{\pi}(v,A) has the integral representation

π(v,A)={Av,𝕋vdπ,if vH1(A,π),+,otherwise.\mathcal{F}^{\pi}(v,A)=\begin{cases}\displaystyle\int_{A}\langle\nabla v,\mathbb{T}\nabla v\rangle\mathop{}\!\mathrm{d}\pi,&\text{if }v\in H^{1}(A,\pi),\\ +\infty,&\text{otherwise.}\end{cases}

with the tensor 𝕋\mathbb{T} defined in Lemma 5.12.

Proof.

Proposition 5.19 and the theorem fundamental on convergence of minimizers (see, for instance, [9, Theorem 1.21]) together with (AMin) provides:

M(φx,ξ,Qε(x))=limh0Mh(φhx,ξ,Qε(x))=limh0hπ(φhx,ξ,Qε(x)).\text{M}(\varphi^{x,\xi},Q_{\varepsilon}(x))=\lim_{h\to 0}\text{M}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x))=\lim_{h\to 0}\mathcal{F}^{\pi}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x)).

Substituting φhx,ξ(K)=ξ,xxK\varphi_{h}^{x,\xi}(K)=\langle\xi,x-x_{K}\rangle into hπ\mathcal{F}^{\pi}_{h} yields

hπ(φhx,ξ,Qε(x))\displaystyle\mathcal{F}^{\pi}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x)) =(K,L)Σh|Qε(x)ϑh(K,L)|ξ,xLxK|2=ξ,𝕋h,ε(x)ξ,\displaystyle=\sum_{(K,L)\in\Sigma^{h}|_{Q_{\varepsilon}(x)}}\vartheta^{h}(K,L)\left|\langle\xi,x_{L}-x_{K}\rangle\right|^{2}=\bigl{\langle}\xi,\mathbb{T}^{h,\varepsilon}(x)\xi\bigr{\rangle},

with

𝕋h,ε(x):=(K,L)Σh|Qε(x)ϑh(K,L)(xLxK)(xLxK).\mathbb{T}^{h,\varepsilon}(x):=\sum_{(K,L)\in\Sigma^{h}|_{Q_{\varepsilon}(x)}}\vartheta^{h}(K,L)(x_{L}-x_{K})\otimes(x_{L}-x_{K}).

Since 𝕋h\mathbb{T}^{h} defined in Lemma  5.12 is piecewise constant on the tessellation, we can rewrite 𝕋h,ε\mathbb{T}^{h,\varepsilon} as

𝕋h,ε(x)=K𝒯h|Qε(x)πh(K)\intbarK𝕋h(z)dz=[Qε(x)]𝒯h𝕋h(z)π^h(dz).\displaystyle\mathbb{T}^{h,\varepsilon}(x)=\sum_{K\in\mathcal{T}^{h}|_{Q_{\varepsilon}(x)}}\pi^{h}(K)\intbar_{K}\mathbb{T}^{h}(z)\mathop{}\!\mathrm{d}z=\int_{[Q_{\varepsilon}(x)]_{\mathcal{T}^{h}}}\mathbb{T}^{h}(z)\,\hat{\pi}^{h}(\mathop{}\!\mathrm{d}z).

Using that 𝟙Qε(x)𝒯hdπ^h/dd𝟙Qε(x)dπ/dd\mathbbm{1}_{Q_{\varepsilon}(x)_{\mathcal{T}^{h}}}\mathop{}\!\mathrm{d}\hat{\pi}^{h}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\to\mathbbm{1}_{Q_{\varepsilon}(x)}\mathop{}\!\mathrm{d}\pi/\mathop{}\!\mathrm{d}\mathscr{L}^{d} in L1(Ω)L^{1}(\Omega) and Lemma 5.12(iii), we then obtain

limh0𝕋h,ε(x)=Qε(x)𝕋(z)π(dz).\lim_{h\to 0}\mathbb{T}^{h,\varepsilon}(x)=\int_{Q_{\varepsilon}(x)}\mathbb{T}(z)\,\pi(\mathop{}\!\mathrm{d}z).

Therefore,

M(φx,ξ,Qε(x))=limh0hπ(φhx,ξ,Qε(x))=limh0ξ,𝕋h,ε(x)ξ=ξ,Qε(x)𝕋(z)π(dz)ξ.\displaystyle\text{M}(\varphi^{x,\xi},Q_{\varepsilon}(x))=\lim_{h\to 0}\mathcal{F}^{\pi}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x))=\lim_{h\to 0}\bigl{\langle}\xi,\mathbb{T}^{h,\varepsilon}(x)\,\xi\bigr{\rangle}=\left\langle\xi,\int_{Q_{\varepsilon}(x)}\mathbb{T}(z)\,\pi(\mathop{}\!\mathrm{d}z)\,\xi\right\rangle.

Finally, we substitute M into the expression (5.9) for ff, and obtain for almost every xΩx\in\Omega:

f(x,ξ)\displaystyle f(x,\xi) =lim supε0+1|Qε(x)|M(φx,ξ,Qε(x))\displaystyle=\limsup_{\varepsilon\to 0+}\frac{1}{\left|Q_{\varepsilon}(x)\right|}\text{M}(\varphi^{x,\xi},Q_{\varepsilon}(x))
=ξ,limε0+\intbarQε(x)𝕋(z)π(dz)ξ=ξ,𝕋(x)dπdd(x)ξ,\displaystyle=\left\langle\xi,\lim_{\varepsilon\to 0+}\intbar_{Q_{\varepsilon}(x)}\mathbb{T}(z)\pi(\mathop{}\!\mathrm{d}z)\,\xi\right\rangle=\left\langle\xi,\mathbb{T}(x)\frac{\mathop{}\!\mathrm{d}\pi}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}(x)\,\xi\right\rangle,

thereby concluding the proof. ∎

Corollary 5.21.

If (Aloc{}_{\text{loc}}) holds, then the functions φhx,ξ\varphi^{x,\xi}_{h} are minimizers for hπ(,Qε(x))\mathcal{F}^{\pi}_{h}(\cdot,Q_{\varepsilon}(x)), i.e.

hπ(φhx,ξ,Qε(x))=Mh(φhx,ξ,Qε(x))for any (x,ξ)Ω×d,h>0,ε>0.\mathcal{F}^{\pi}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x))=\text{M}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x))\quad\text{for any }(x,\xi)\in\Omega\times\mathbb{R}^{d},\,h>0,\,\varepsilon>0.

In particular, the conclusion of Theorem 5.20 holds true.

Proof.

Computing the first variation for hπ(φh,Qε(x))\mathcal{F}^{\pi}_{h}(\varphi^{h},Q_{\varepsilon}(x)) gives

δhπ(φh,Qε(x))[wh]=2(K,L)Σh|Qε(x)ϑh(K,L)(φh(L)φh(K))(wh(L)wh(K)),\displaystyle\delta\mathcal{F}^{\pi}_{h}(\varphi^{h},Q_{\varepsilon}(x))[w_{h}]=2\sum_{(K,L)\in\Sigma^{h}|_{Q_{\varepsilon}(x)}}\vartheta^{h}(K,L)\left(\varphi^{h}(L)-\varphi^{h}(K)\right)\left(w^{h}(L)-w^{h}(K)\right),

where whw^{h} satisfies the boundary condition wh=0w^{h}=0 on [Qεc(x)]𝒯h[Q^{c}_{\varepsilon}(x)]_{\mathcal{T}^{h}}.

Substituting φhx,ξ\varphi_{h}^{x,\xi} and then using the symmetry, we have

δhπ(φhx,ξ,Qε(x))[wh]\displaystyle\delta\mathcal{F}^{\pi}_{h}(\varphi^{x,\xi}_{h},Q_{\varepsilon}(x))[w_{h}] =2(K,L)Σh|Qε(x)ϑh(K,L)ξ,xLxK(wh(L)wh(K))\displaystyle=2\sum_{(K,L)\in\Sigma^{h}|_{Q_{\varepsilon}(x)}}\vartheta^{h}(K,L)\langle\xi,x_{L}-x_{K}\rangle\left(w_{h}(L)-w_{h}(K)\right)
=4K𝒯h|Qε(x)wh(K)ξ,L𝒯Kh|Qε(x)ϑh(K,L)(xLxK).\displaystyle=4\sum_{K\in\mathcal{T}^{h}|_{Q_{\varepsilon}(x)}}w_{h}(K)\Big{\langle}\xi,\sum_{L\in\mathcal{T}^{h}_{K}|_{Q_{\varepsilon}(x)}}\vartheta^{h}(K,L)(x_{L}-x_{K})\Big{\rangle}.

Notice that the boundary condition implies that the summation goes over cells KQε(x)K\subset Q_{\varepsilon}(x) strictly contained within Qε(x)Q_{\varepsilon}(x). This means that the inside sum goes over all the neighbors of the cell KK. This allows us to apply assumption (Aloc{}_{\text{loc}}) to obtain

L𝒯Khϑh(K,L)(xKxL)=0,\sum_{L\in\mathcal{T}^{h}_{K}}\vartheta^{h}(K,L)\left(x_{K}-x_{L}\right)=0,

and conclude that δhπ(φhx,ξ,Qε(x))[wh]=0\delta\mathcal{F}^{\pi}_{h}(\varphi_{h}^{x,\xi},Q_{\varepsilon}(x))[w_{h}]=0 for all whw_{h} with wh0w_{h}\equiv 0 on [Qε(x)]𝒯hc[Q_{\varepsilon}(x)]^{c}_{\mathcal{T}^{h}}. Since hπ(,Qε(x))\mathcal{F}^{\pi}_{h}(\cdot,Q_{\varepsilon}(x)) is convex, this implies that φhx,ξ\varphi^{x,\xi}_{h} is the minimizer. ∎

Proof of Theorem 5.15.

The result readily follows from Theorem 5.20. ∎

Lemma 5.22.

The diffusion tensor 𝕋\mathbb{T} is uniformly elliptic and uniformly bounded:

λ|ξ|2ξ,𝕋(x)ξΛ|ξ|2for any xΩ and ξd,\lambda|\xi|^{2}\leq\langle\xi,~{}\mathbb{T}(x)\,\xi\rangle\leq\Lambda|\xi|^{2}\quad\text{for any }x\in\Omega\text{ and }\xi\in\mathbb{R}^{d},

with some λ,Λ>0\lambda,\Lambda>0.

Proof.

The upper bound follows from Lemma 5.12(2) with Λ=2Cκ\Lambda=2C_{\kappa}. The lower bound can be deduced from Theorem 5.20. Indeed, since

ξ,𝕋(x)ξ\displaystyle\langle\xi,\mathbb{T}(x)\,\xi\rangle =ξ,limε0\intbarQε(x)𝕋(z)dzξ=limε0\intbarQε(x)ξ,𝕋(z)ξdz\displaystyle=\left\langle\xi,\lim_{\varepsilon\to 0}\intbar_{Q_{\varepsilon}(x)}\mathbb{T}(z)\mathop{}\!\mathrm{d}z~{}\xi\right\rangle=\lim_{\varepsilon\to 0}\intbar_{Q_{\varepsilon}(x)}\langle\xi,\mathbb{T}(z)\,\xi\rangle\mathop{}\!\mathrm{d}z
=limε01|Qε(x)|π(,ξ,Qε(x)),\displaystyle=\lim_{\varepsilon\to 0}\frac{1}{|Q_{\varepsilon}(x)|}\mathcal{F}^{\pi}\left(\langle\cdot,\xi\rangle,Q_{\varepsilon}(x)\right),

applying the lower bound for π\mathcal{F}^{\pi} from Lemma 5.17 gives

ξ,𝕋(x)ξlimε0ClCζπmax2|Qε(x)|Qε(x)|ξ|2dπ2ClπminCζπmax|ξ|2,\displaystyle\langle\xi,\mathbb{T}(x)\,\xi\rangle\geq\lim_{\varepsilon\to 0}\frac{C_{l}}{C_{\zeta}\pi_{\max}}\frac{2}{|Q_{\varepsilon}(x)|}\int_{Q_{\varepsilon}(x)}|\xi|^{2}\mathop{}\!\mathrm{d}\pi\geq 2\frac{C_{l}\pi_{\min}}{C_{\zeta}\pi_{\max}}|\xi|^{2},

from which the lower bound follows. ∎

6. Convergence result

With the results above, we are now in the position of proving our main result, Theorem A. We begin by recalling our reconstruction procedure for discrete density-flux pairs from the beginning of Section 4. We then proceed to show lim inf\liminf inequalities for the functionals h\mathcal{E}_{h}, 𝒟h\mathcal{D}_{h} and h\mathcal{R}_{h}, therewith establishing the lim inf\liminf inequality for the energy-dissipation functional h\mathcal{I}_{h} (cf. Theorem 6.2). The chain rule is established in Section 6.2, which is essential in guaranteeing the nonnegativity of the limit energy-dissipation functional \mathcal{I}. Finally, we conclude with the proof of Theorem A.

6.1. Lim inf inequalities

Given a density-flux pair (ρh,jh)𝒞h(0,T)(\rho^{h},j^{h})\in\mathcal{CE}_{h}(0,T) we define

(6.1) dρ^hdd:=K𝒯hρh(K)|K|𝟙K,ȷ^h:=(K,L)Σhjh(K,L)σKL,\frac{\mathop{}\!\mathrm{d}\hat{\rho}^{h}}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}:=\sum_{K\in\mathcal{T}^{h}}\frac{\rho^{h}(K)}{|K|}\mathbbm{1}_{K},\qquad\hat{\jmath}^{h}:=\sum_{(K,L)\in\Sigma^{h}}j^{h}(K,L)\sigma_{KL},

where σKL(Ω;d)\sigma_{KL}\in\mathcal{M}(\Omega;\mathbb{R}^{d}) is defined in the way that (ρ^h,ȷ^h)𝒞(0,T)(\hat{\rho}^{h},\hat{\jmath}^{h})\in\mathcal{CE}(0,T) (we constructed σKL\sigma_{KL} explicitly in Lemma 4.1).

Definition 6.1 (Density-flux convergence).

A discrete density-flux pair (ρh,jh)𝒞h(0,T)(\rho^{h},j^{h})\in\mathcal{CE}_{h}(0,T) is said to converge to a density-flux pair (ρ,j)𝒞(0,T)(\rho,j)\in\mathcal{CE}(0,T) if the pair of reconstructions (ρ^h,ȷ^h)𝒞(0,T)(\hat{\rho}^{h},\hat{\jmath}^{h})\in\mathcal{CE}(0,T) defined as in (6.1) converges in the following sense

  1. (1)

    dρ^th/dddρt/dd\mathop{}\!\mathrm{d}\hat{\rho}^{h}_{t}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\to\mathop{}\!\mathrm{d}\rho_{t}/\mathop{}\!\mathrm{d}\mathscr{L}^{d} in L1(Ω)L^{1}(\Omega) for almost every t[0,T]t\in[0,T],

  2. (2)

    ȷ^thdtjtdt\int_{\cdot}\hat{\jmath}_{t}^{h}\,\mathop{}\!\mathrm{d}t\rightharpoonup^{*}\int_{\cdot}j_{t}\,\mathop{}\!\mathrm{d}t in ([0,T]×Ω)\mathcal{M}([0,T]\times\Omega).

We now summarize the lower bounds for all components of the energy-dissipation functional h\mathcal{I}_{h} defined in Section 3.1. The form of the lower bounds is already suggested by Lemma 5.14 for the dissipation potential \mathcal{R} and by Theorem 5.15 for the Fisher information 𝒟\mathcal{D}. Let us first give the definitions of \mathcal{R}, \mathcal{R}^{*}, 𝒟\mathcal{D}, and \mathcal{E} and then summarize the corresponding lim inf\liminf inequalities in Theorem 6.2.

The dual dissipation potential :𝒫(Ω)×𝒞c2(Ω)[0,)\mathcal{R}^{*}:\mathcal{P}(\Omega)\times\mathcal{C}_{c}^{2}(\Omega)\to[0,\infty) takes the form

(ρ,φ)=14Ωφ,𝕋φdρ.\mathcal{R}^{*}(\rho,\varphi)=\frac{1}{4}\int_{\Omega}\langle\nabla\varphi,\mathbb{T}\nabla\varphi\rangle\mathop{}\!\mathrm{d}\rho.

The dissipation potential :𝒫(Ω)×(Ω;d)[0,+]\mathcal{R}:\mathcal{P}(\Omega)\times\mathcal{M}(\Omega;\mathbb{R}^{d})\to[0,+\infty] is

(ρ,j)={14Ωdjdρ,𝕋1djdρdρif jρ,+otherwise\mathcal{R}(\rho,j)=\begin{cases}\displaystyle\frac{1}{4}\int_{\Omega}\Big{\langle}\frac{\mathop{}\!\mathrm{d}j}{\mathop{}\!\mathrm{d}\rho},\mathbb{T}^{-1}\frac{\mathop{}\!\mathrm{d}j}{\mathop{}\!\mathrm{d}\rho}\Big{\rangle}\mathop{}\!\mathrm{d}\rho&\text{if $j\ll\rho$},\\ +\infty&\text{otherwise}\end{cases}

The Fisher information 𝒟:𝒫(Ω)[0,+]\mathcal{D}:\mathcal{P}(\Omega)\to[0,+\infty] is defined as

𝒟(ρ)={Ωu,𝕋udπif dρdπ=:uH1(Ω),+otherwise.\mathcal{D}(\rho)=\begin{cases}\displaystyle\int_{\Omega}\big{\langle}\nabla\sqrt{u},\mathbb{T}\nabla\sqrt{u}\big{\rangle}\mathop{}\!\mathrm{d}\pi&\text{if }\sqrt{\frac{\mathop{}\!\mathrm{d}\rho}{\mathop{}\!\mathrm{d}\pi}}=:\sqrt{u}\in H^{1}(\Omega),\\ +\infty&\text{otherwise.}\end{cases}

The energy functional :𝒫(Ω)[0,+]\mathcal{E}:\mathcal{P}(\Omega)\to[0,+\infty] is given by (ρ)=Ent(ρ|π)\mathcal{E}(\rho)=\operatorname{\text{Ent}}(\rho|\pi).

Theorem 6.2.

Let (ρh,jh)𝒞h(0,T)(\rho^{h},j^{h})\in\mathcal{CE}_{h}(0,T) converge to (ρ,j)𝒞(0,T)(\rho,j)\in\mathcal{CE}(0,T) in the sense of Definition 6.1. Then the following lower bounds hold for

  1. (i)

    the dissipation potential:

    lim infh00Th(ρth,jth)dt0T(ρt,jt)dt;\liminf_{h\to 0}\int_{0}^{T}\mathcal{R}_{h}(\rho^{h}_{t},j^{h}_{t})\mathop{}\!\mathrm{d}t\geq\int_{0}^{T}\mathcal{R}(\rho_{t},j_{t})\mathop{}\!\mathrm{d}t;
  2. (ii)

    the Fisher information:

    lim infh00T𝒟h(ρth)dt0T𝒟(ρt)dt;\liminf_{h\to 0}\int_{0}^{T}\mathcal{D}_{h}(\rho^{h}_{t})\mathop{}\!\mathrm{d}t\geq\int_{0}^{T}\mathcal{D}(\rho_{t})\mathop{}\!\mathrm{d}t;
  3. (iii)

    the energy functional:

    lim infh0h(ρth)(ρt)for all t[0,T].\liminf_{h\to 0}\mathcal{E}_{h}(\rho^{h}_{t})\geq\mathcal{E}(\rho_{t})\qquad\text{for all $t\in[0,T]$.}
Proof.

(i) The dissipation potential. We employ the dual formulation of \mathcal{R}. Let χ𝒞c((0,T))\chi\in\mathcal{C}^{\infty}_{c}((0,T)) and φ𝒞c(Ω)\varphi\in\mathcal{C}_{c}^{\infty}(\Omega) be arbitrary. Then from the weak-convergence of ȷ^thdt\int_{\cdot}\hat{\jmath}_{t}^{h}\,\mathop{}\!\mathrm{d}t and Lemma 5.14 we obtain

0Tχ(t)φ,jt(ρt,χ(t)φ)dtlimh00Tχ(t)φ,ȷ^thdtlim suph00Th(ρth,χ(t)¯φh)dt,\displaystyle\int_{0}^{T}\langle\chi(t)\nabla\varphi,j_{t}\rangle-\mathcal{R}^{*}(\rho_{t},\chi(t)\nabla\varphi)\mathop{}\!\mathrm{d}t\leq\lim_{h\to 0}\int_{0}^{T}\langle\chi(t)\nabla\varphi,\,\hat{\jmath}_{t}^{h}\rangle\mathop{}\!\mathrm{d}t-\limsup_{h\to 0}\int_{0}^{T}\mathcal{R}_{h}^{*}(\rho_{t}^{h},\chi(t)\overline{\nabla}\varphi^{h})\mathop{}\!\mathrm{d}t,

where φh(K)=φ(xK)\varphi^{h}(K)=\varphi(x_{K}) for all K𝒯hK\in\mathcal{T}^{h}. For the first term on the right-hand side, we have

0Tχ(t)φ,ȷ^thdt=0Tχ(t)¯hφ,jthdt=0Tχ(t)¯φh,jthdt+o(h),\int_{0}^{T}\langle\chi(t)\nabla\varphi,\,\hat{\jmath}_{t}^{h}\rangle\mathop{}\!\mathrm{d}t=\int_{0}^{T}\langle\chi(t)\overline{\nabla}\mathbb{P}_{h}\varphi,j_{t}^{h}\rangle\mathop{}\!\mathrm{d}t=\int_{0}^{T}\langle\chi(t)\overline{\nabla}\varphi^{h},j_{t}^{h}\rangle\mathop{}\!\mathrm{d}t+o(h),

owing to the regularity of φ\varphi and Lemma 4.4, and therefore

limh00Tχ(t)φ,ȷ^thdt=limh00Tχ(t)¯φh,jthdt.\lim_{h\to 0}\int_{0}^{T}\langle\chi(t)\nabla\varphi,\hat{\jmath}_{t}^{h}\rangle\mathop{}\!\mathrm{d}t=\lim_{h\to 0}\int_{0}^{T}\langle\chi(t)\overline{\nabla}\varphi^{h},j_{t}^{h}\rangle\mathop{}\!\mathrm{d}t.

Consequently, we obtain

0Tχ(t)φ,jt(ρt,χ(t)φ)dt\displaystyle\int_{0}^{T}\langle\chi(t)\nabla\varphi,j_{t}\rangle-\mathcal{R}^{*}(\rho_{t},\chi(t)\nabla\varphi)\mathop{}\!\mathrm{d}t limh00Tχ(t)¯φh,jthdtlim suph00Th(ρth,χ(t)¯φh)dt\displaystyle\leq\lim_{h\to 0}\int_{0}^{T}\langle\chi(t)\overline{\nabla}\varphi^{h},j_{t}^{h}\rangle\mathop{}\!\mathrm{d}t-\limsup_{h\to 0}\int_{0}^{T}\mathcal{R}_{h}^{*}(\rho_{t}^{h},\chi(t)\overline{\nabla}\varphi^{h})\mathop{}\!\mathrm{d}t
lim infh00Tχ(t)¯φh,jthh(ρth,χ(t)¯φh)dt\displaystyle\leq\liminf_{h\to 0}\int_{0}^{T}\langle\chi(t)\overline{\nabla}\varphi^{h},j_{t}^{h}\rangle-\mathcal{R}_{h}^{*}(\rho_{t}^{h},\chi(t)\overline{\nabla}\varphi^{h})\mathop{}\!\mathrm{d}t
lim infh00Th(ρth,jth)dt.\displaystyle\leq\liminf_{h\to 0}\int_{0}^{T}\mathcal{R}_{h}(\rho_{t}^{h},j_{t}^{h})\mathop{}\!\mathrm{d}t.

To conclude, we will make use of Legendre–Fenchel’s duality. In what follows, we set P:=ρtdtP:=\int_{\cdot}\rho_{t}\,\mathop{}\!\mathrm{d}t, J:=jtdtJ:=\int_{\cdot}j_{t}\,\mathop{}\!\mathrm{d}t, and 𝒱\mathcal{V} as the closure in L2(Ω,P;d)L^{2}(\Omega,P;\mathbb{R}^{d}) of the subspace V:={𝕋1/2(χφ):χ𝒞c((0,T)),φ𝒞c(Ω)}V:=\{\mathbb{T}^{1/2}\nabla(\chi\varphi)\,:\,\chi\in\mathcal{C}_{c}^{\infty}((0,T)),\,\varphi\in\mathcal{C}_{c}^{\infty}(\Omega)\}, where 𝕋1/2\mathbb{T}^{1/2} denotes the square root of the positive definite matrix 𝕋\mathbb{T}. Writing the term on the left in the previous inequality as

(0,T)×Ω𝕋1/2(χφ)𝕋1/2dJdPdP12𝕋1/2(χφ)L2(Ω,P;d)2,\iint_{(0,T)\times\Omega}\mathbb{T}^{1/2}\nabla(\chi\varphi)\cdot\mathbb{T}^{-1/2}\frac{\mathop{}\!\mathrm{d}J}{\mathop{}\!\mathrm{d}P}\,\mathop{}\!\mathrm{d}P-\frac{1}{2}\|\mathbb{T}^{1/2}\nabla(\chi\varphi)\|_{L^{2}(\Omega,P;\mathbb{R}^{d})}^{2},

the Fenchel–Moreau duality theorem then gives

supψV{(0,T)×Ω𝕋1/2(χφ)𝕋1/2dJdPdP12𝕋1/2(χφ)L2(Ω,P;d)2}\displaystyle\sup_{\psi\in V}\left\{\iint_{(0,T)\times\Omega}\mathbb{T}^{1/2}\nabla(\chi\varphi)\cdot\mathbb{T}^{-1/2}\frac{\mathop{}\!\mathrm{d}J}{\mathop{}\!\mathrm{d}P}\,\mathop{}\!\mathrm{d}P-\frac{1}{2}\|\mathbb{T}^{1/2}\nabla(\chi\varphi)\|_{L^{2}(\Omega,P;\mathbb{R}^{d})}^{2}\right\}
=supψ𝒱{(0,T)×Ωψ𝕋1/2dJdPdP12ψL2(Ω,P;d)2}\displaystyle\qquad=\sup_{\psi\in\mathcal{V}}\left\{\iint_{(0,T)\times\Omega}\psi\cdot\mathbb{T}^{-1/2}\frac{\mathop{}\!\mathrm{d}J}{\mathop{}\!\mathrm{d}P}\,\mathop{}\!\mathrm{d}P-\frac{1}{2}\|\psi\|_{L^{2}(\Omega,P;\mathbb{R}^{d})}^{2}\right\}
=12𝕋1/2dJdPL2(Ω,P;d)2=0T(ρt,jt)dt,\displaystyle\qquad=\frac{1}{2}\left\|\mathbb{T}^{-1/2}\frac{\mathop{}\!\mathrm{d}J}{\mathop{}\!\mathrm{d}P}\right\|_{L^{2}(\Omega,P;\mathbb{R}^{d})}^{2}=\int_{0}^{T}\mathcal{R}(\rho_{t},j_{t})\mathop{}\!\mathrm{d}t\,,

where the last equality follows from the fact that (dJ/dP)(t,x)=(djt/dρt)(t,x)(\mathop{}\!\mathrm{d}J/\mathop{}\!\mathrm{d}P)(t,x)=(\mathop{}\!\mathrm{d}j_{t}/\mathop{}\!\mathrm{d}\rho_{t})(t,x) for PP-almost every (t,x)(0,T)×Ω(t,x)\in(0,T)\times\Omega.

(ii) The Fisher information. Since dρ^th/dddρt/dd\mathop{}\!\mathrm{d}\hat{\rho}_{t}^{h}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\to\mathop{}\!\mathrm{d}\rho_{t}/\mathop{}\!\mathrm{d}\mathscr{L}^{d} strongly in L1(Ω)L^{1}(\Omega) for almost every t(0,T)t\in(0,T), we have by Theorem 5.15 that

lim infh0𝒟h(ρ^th)𝒟(ρt)for almost every t(0,T).\liminf_{h\to 0}\mathcal{D}_{h}(\hat{\rho}^{h}_{t})\geq\mathcal{D}(\rho_{t})\qquad\text{for almost every $t\in(0,T)$.}

Applying Fatou’s lemma then yields

lim infh00T𝒟h(ρth)dt0Tlim infh0𝒟h(ρth)dt.\liminf_{h\to 0}\int_{0}^{T}\mathcal{D}_{h}(\rho^{h}_{t})\mathop{}\!\mathrm{d}t\geq\int_{0}^{T}\liminf_{h\to 0}\mathcal{D}_{h}(\rho^{h}_{t})\mathop{}\!\mathrm{d}t.

(iii) The energy functional. As the following calculations hold for any t[0,T]t\in[0,T], we drop the subscript tt. We recall that

h(ρh)={K𝒯hϕ(uh(K))πh(K) if ρhπh, with uh(K)=ρh(K)πh(K),+otherwise,\mathcal{E}_{h}(\rho^{h})=\begin{cases}\displaystyle\sum_{K\in\mathcal{T}^{h}}\phi\left(u^{h}(K)\right)\pi^{h}(K)&\text{ if }\rho^{h}\ll\pi^{h},\text{ with }u^{h}(K)=\frac{\rho^{h}(K)}{\pi^{h}(K)},\\ +\infty&\text{otherwise,}\end{cases}

where ϕ(z)=zlogzz+1\phi(z)=z\log z-z+1. Since ρh\rho^{h} and πh\pi^{h} are probability measures,

h(ρh)=K𝒯huh(K)log(uh(K))πh(K)if ρhπh.\mathcal{E}_{h}(\rho^{h})=\sum_{K\in\mathcal{T}^{h}}u^{h}(K)\log\left(u^{h}(K)\right)\pi^{h}(K)\qquad\text{if $\rho^{h}\ll\pi^{h}$.}

Our piecewise constant reconstruction provides that

h(ρh)=Ωu^h(x)log(u^h(x))π^h(dx)=Ent(ρ^h|π^h).\mathcal{E}_{h}(\rho^{h})=\int_{\Omega}\hat{u}^{h}(x)\log\bigl{(}\hat{u}^{h}(x)\bigr{)}\hat{\pi}^{h}(\mathop{}\!\mathrm{d}x)=\operatorname{\text{Ent}}(\hat{\rho}^{h}|\hat{\pi}^{h}).

The narrow convergence of ρh\rho^{h} and πh\pi^{h} in 𝒫(Ω)\mathcal{P}(\Omega), along with the joint lower semicontinuity of the relative entropy [3, Lemma 9.4.3] then gives

lim infh0h(ρh)=lim infh0Ent(ρ^h|π^h)Ent(ρ|π)=(ρ),\liminf_{h\to 0}\mathcal{E}_{h}(\rho^{h})=\liminf_{h\to 0}\operatorname{\text{Ent}}(\hat{\rho}^{h}|\hat{\pi}^{h})\geq\operatorname{\text{Ent}}(\rho|\pi)=\mathcal{E}(\rho),

as required. ∎

6.2. Chain rule

In this section we aim to establish the chain rule inequality:

dmissingdt(ρt)(ρt,jt)+𝒟(ρt)for almost every t(0,T),-\frac{\mathop{}\!\mathrm{d}missing}{\mathop{}\!\mathrm{d}t}\mathcal{E}(\rho_{t})\leq\mathcal{R}(\rho_{t},j_{t})+\mathcal{D}(\rho_{t})\qquad\text{for almost every $t\in(0,T)$,}

from which we establish the nonnegativity of the limit energy-dissipation functional, i.e.

(ρ,j)=0T{(ρt,jt)+𝒟(ρt)}dt+(ρT)(ρ0)0.\mathcal{I}(\rho,j)=\int_{0}^{T}\left\{\mathcal{R}(\rho_{t},j_{t})+\mathcal{D}(\rho_{t})\right\}\mathop{}\!\mathrm{d}t+\mathcal{E}(\rho_{T})-\mathcal{E}(\rho_{0})\geq 0.

We will show that this inequality can be obtained from the chain rule for the relative entropy Ent(ρ|π)\operatorname{\text{Ent}}(\rho|\pi) along W2W_{2}-absolutely continuous curves.

We begin by rewriting Ent(ρ|π)\operatorname{\text{Ent}}(\rho|\pi) in a more convenient form for the purpose of this section. We denote V:=log(dπ/dd)V:=-\log\left(\mathop{}\!\mathrm{d}\pi/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\right). If measures ρ\rho and π\pi have Lebesgue densities, then it holds that

Ent(ρ|π)=Ωdρdπlogdρdπdπ\displaystyle\operatorname{\text{Ent}}(\rho|\pi)=\int_{\Omega}\frac{\mathop{}\!\mathrm{d}\rho}{\mathop{}\!\mathrm{d}\pi}\log\frac{\mathop{}\!\mathrm{d}\rho}{\mathop{}\!\mathrm{d}\pi}\mathop{}\!\mathrm{d}\pi =Ωdρddlogdρdddd+ΩVdρ=Ent(ρ|d)+ΩV𝑑ρ,\displaystyle=\int_{\Omega}\frac{\mathop{}\!\mathrm{d}\rho}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}\log\frac{\mathop{}\!\mathrm{d}\rho}{\mathop{}\!\mathrm{d}\mathscr{L}^{d}}\mathop{}\!\mathrm{d}\mathscr{L}^{d}+\int_{\Omega}V\mathop{}\!\mathrm{d}\rho=\operatorname{\text{Ent}}(\rho|\mathscr{L}^{d})+\int_{\Omega}V\,d\rho,

which can be justified by monotone convergence. Recall that by the assumptions on {πh}h0\{\pi^{h}\}_{h\geq 0} (Section 2.2) VLipb(Ω)V\in\text{Lip}_{b}(\Omega), and, therefore, Ent(ρ|d)\operatorname{\text{Ent}}(\rho|\mathscr{L}^{d}) is finite whenever Ent(ρ|π)\operatorname{\text{Ent}}(\rho|\pi) is finite.

Now we extend the definition of the energy for all measures ρ\rho with Lebesgue densities. First, we define an extended potential VE:d(,+]V_{E}:\mathbb{R}^{d}\to(-\infty,+\infty] by

VE(x):={V(x)if xΩ¯,+otherwise.V_{E}(x):=\begin{cases}V(x)&\text{if }x\in\overline{\Omega},\\ +\infty&\text{otherwise.}\end{cases}

Since V𝒞b(Ω¯)V\in\mathcal{C}_{b}(\overline{\Omega}), VEV_{E} is a lower semicontinuous on d\mathbb{R}^{d}. Then for any ρ𝒫(d)\rho\in\mathcal{P}(\mathbb{R}^{d}) we consider the extended energy functional E:𝒫2(d)(,+]\mathcal{E}_{E}:\mathcal{P}_{2}(\mathbb{R}^{d})\to(-\infty,+\infty] defined by

E(ρ):={Ent(ρ|d)+dVEdρfor ρd,+otherwise.\mathcal{E}_{E}(\rho):=\begin{cases}\displaystyle\;\operatorname{\text{Ent}}(\rho|\mathscr{L}^{d})+\int_{\mathbb{R}^{d}}V_{E}\mathop{}\!\mathrm{d}\rho&\text{for $\rho\ll\mathscr{L}^{d}$,}\\ \;+\infty&\text{otherwise.}\end{cases}

We remark that the functionals E\mathcal{E}_{E} and Ent(ρ|π)\operatorname{\text{Ent}}(\rho|\pi) coincide on their sublevel sets. We also mention that Ent(ρ|d)>\operatorname{\text{Ent}}(\rho|\mathscr{L}^{d})>-\infty if ρ𝒫2(d)\rho\in\mathcal{P}_{2}(\mathbb{R}^{d}) [3].

The following lemma results from a minor modification of [3, Theorem 10.4.13]. In our case, VV is not λ\lambda-convex, but the result remains true due to the regularity assumed on VV, i.e. VLipb(Ω)V\in\text{Lip}_{b}(\Omega).

Lemma 6.3.

A measure ρ=ϱddom(E)\rho=\varrho\mathscr{L}^{d}\in\text{dom}(\mathcal{E}_{E}) belongs to dom(E)\text{dom}(\partial\mathcal{E}_{E}) if and only if ϱWloc1,1(Ω)\varrho\in W^{1,1}_{loc}(\Omega) and

(6.2) ϱw=ϱ+ϱVEfor some wL2(d,ρ;d).\varrho w=\nabla\varrho+\varrho\nabla V_{E}\qquad\text{for some $w\in L^{2}(\mathbb{R}^{d},\rho;\mathbb{R}^{d})$}.

In this case, ww is the minimal selection in E\partial\mathcal{E}_{E}.

Theorem 6.4 (Chain rule).

Let (ρ,j)𝒞(0,T)(\rho,j)\in\mathcal{CE}(0,T) be such that

0T{(ρt,jt)+𝒟(ρt)}dt<andsupt[0,T](ρt)<.\int_{0}^{T}\left\{\mathcal{R}(\rho_{t},j_{t})+\mathcal{D}(\rho_{t})\right\}\mathop{}\!\mathrm{d}t<\infty\quad\text{and}\quad\sup\nolimits_{t\in[0,T]}\mathcal{E}(\rho_{t})<\infty.

Then the map t(ρt)t\mapsto\mathcal{E}(\rho_{t}) is absolutely continuous, and

dmissingdt(ρt)(ρt,jt)+𝒟(ρt)for almost every t(0,T).-\frac{\mathop{}\!\mathrm{d}missing}{\mathop{}\!\mathrm{d}t}\mathcal{E}(\rho_{t})\leq\mathcal{R}(\rho_{t},j_{t})+\mathcal{D}(\rho_{t})\qquad\text{for almost every $t\in(0,T)$.}

In particular, this implies

(ρ,j)=0T{(ρt,jt)+𝒟(ρt)}dt+(ρT)(ρ0)0.\mathcal{I}(\rho,j)=\int_{0}^{T}\left\{\mathcal{R}(\rho_{t},j_{t})+\mathcal{D}(\rho_{t})\right\}\mathop{}\!\mathrm{d}t+\mathcal{E}(\rho_{T})-\mathcal{E}(\rho_{0})\geq 0.
Proof.

From the continuity equation and finiteness of 0T(ρt,jt)dt\int_{0}^{T}\mathcal{R}(\rho_{t},j_{t})\,\mathop{}\!\mathrm{d}t, we deduce from [3, Theorem 8.3.1] that [0,T]tρt[0,T]\ni t\mapsto\rho_{t} is W2W_{2}-absolutely continuous (cf. Remark 3.6).

Furthermore, it is not difficult to show that the extended functional E\mathcal{E}_{E} defined above is a regular functional (according to [3, Definition 10.1.4]) satisfying the properties in [3, Equations (10.1.1a,b)]. In particular, [3, E. Chain rule in Section 10.1.2 ] applies, i.e. we have that

d~dtE(ρt)=dwt,djtdρtdρtfor all wtE(ρt) and tA,\frac{\tilde{\text{d}}}{\mathop{}\!\mathrm{d}t}\mathcal{E}_{E}(\rho_{t})=\int_{\mathbb{R}^{d}}\left\langle w_{t},\frac{\mathop{}\!\mathrm{d}j_{t}}{\mathop{}\!\mathrm{d}\rho_{t}}\right\rangle\mathop{}\!\mathrm{d}\rho_{t}\qquad\text{for all $w_{t}\in\partial\mathcal{E}_{E}(\rho_{t})$ and $t\in A$},

where A(0,T)A\subset(0,T) is the set of points satisfying the properties in [3, (a,b,c) of E. Chain rule in Section 10.1.2]. In the following, we show that the set (0,T)A(0,T)\setminus A is 1\mathscr{L}^{1}-negligible.

Due to the λ\lambda-convexity of ρEnt(ρ|d)\rho\mapsto\operatorname{\text{Ent}}(\rho|\mathscr{L}^{d}) w.r.t. the W2W_{2}-metric [3] (see also [33]), we have that

tEnt(ρt|d)is absolutely continuous.t\mapsto\operatorname{\text{Ent}}(\rho_{t}|\mathscr{L}^{d})\qquad\text{is absolutely continuous.}

On the other hand, the Lipschitz continuity of VV gives

|ΩVdρtΩVdρs|Ω×Ω|V(x)V(y)|πst(dxdy)VL(Ω)W2(ρt,ρs).\left|\int_{\Omega}V\,\mathop{}\!\mathrm{d}\rho_{t}-\int_{\Omega}V\,\mathop{}\!\mathrm{d}\rho_{s}\right|\leq\iint_{\Omega\times\Omega}|V(x)-V(y)|\,\pi_{s}^{t}(\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}y)\leq\|\nabla V\|_{L^{\infty}(\Omega)}W_{2}(\rho_{t},\rho_{s}).

Altogether, we find that

tE(ρt)is absolutely continuous.t\mapsto\mathcal{E}_{E}(\rho_{t})\qquad\text{is absolutely continuous.}

In particular, the map tE(ρt)t\mapsto\mathcal{E}_{E}(\rho_{t}) is differentiable for almost every t(0,T)t\in(0,T).

We now show that ϱt=dρt/ddWloc1,1(Ω)\varrho_{t}=\mathop{}\!\mathrm{d}\rho_{t}/\mathop{}\!\mathrm{d}\mathscr{L}^{d}\in W_{loc}^{1,1}(\Omega) and that (6.2) holds. Notice that if 𝒟(ρt)<\mathcal{D}(\rho_{t})<\infty, then

B|ϱt|dx\displaystyle\int_{B}|\nabla\varrho_{t}|\,\mathop{}\!\mathrm{d}x B|ut|dπ+But|V|dπ\displaystyle\leq\int_{B}|\nabla u_{t}|\,\mathop{}\!\mathrm{d}\pi+\int_{B}u_{t}|\nabla V|\,\mathop{}\!\mathrm{d}\pi
2λ1/2ρt(B)𝒟(ρt)+VL(Ω)ρt(B)<for any Borel set BΩ,\displaystyle\leq 2\lambda^{-1/2}\sqrt{\rho_{t}(B)}\sqrt{\mathcal{D}(\rho_{t})}+\|\nabla V\|_{L^{\infty}(\Omega)}\,\rho_{t}(B)<\infty\qquad\text{for any Borel set $B\subset\Omega$,}

thus implying that |ϱt|dρt=ϱtd|\nabla\varrho_{t}|\,\mathscr{L}^{d}\ll\rho_{t}=\varrho_{t}\mathscr{L}^{d} and ϱtWloc1,1(Ω)\varrho_{t}\in W^{1,1}_{loc}(\Omega). Here we used the fact that

𝒟(ρ)=Bu,𝕋udπλB|u|2dπ.\mathcal{D}(\rho)=\int_{B}\langle\nabla\sqrt{u},\mathbb{T}\nabla\sqrt{u}\rangle\,\mathop{}\!\mathrm{d}\pi\geq\lambda\int_{B}|\nabla\sqrt{u}|^{2}\,\mathop{}\!\mathrm{d}\pi.

We now define

wt:=ϱtϱt+VEρt-almost everywhere.w_{t}:=\frac{\nabla\varrho_{t}}{\varrho_{t}}+\nabla V_{E}\qquad\text{$\rho_{t}$-almost everywhere.}

Then, with a similar computation as above, we obtain

wtL2(ρt)2\displaystyle\|w_{t}\|_{L^{2}(\rho_{t})}^{2} =Ω|ϱtϱt+VE|2dρt={ϱt>0}|utut|2dρt=4Ω|ut|2dπ4λ1𝒟(ρt).\displaystyle=\int_{\Omega}\left|\frac{\nabla\varrho_{t}}{\varrho_{t}}+\nabla V_{E}\right|^{2}\mathop{}\!\mathrm{d}\rho_{t}=\int_{\{\varrho_{t}>0\}}\left|\frac{\nabla u_{t}}{u_{t}}\right|^{2}\,\mathop{}\!\mathrm{d}\rho_{t}=4\int_{\Omega}\left|\nabla\sqrt{u_{t}}\right|^{2}\mathop{}\!\mathrm{d}\pi\leq 4\lambda^{-1}\mathcal{D}(\rho_{t}).

Hence, Lemma 6.3 implies that ρtdom(E)\rho_{t}\in\text{dom}(\partial\mathcal{E}_{E}) and wtE(ρt)w_{t}\in\partial\mathcal{E}_{E}(\rho_{t}) is a minimal selection.

We then conclude that (0,T)A(0,T)\setminus A is 1\mathscr{L}^{1}-negligible and for all almost every t(0,T)t\in(0,T):

dmissingdtE(ρt)\displaystyle\frac{\mathop{}\!\mathrm{d}missing}{\mathop{}\!\mathrm{d}t}\mathcal{E}_{E}(\rho_{t}) =Ωϱtϱt+V,djtdρtdρt=Ω𝕋1/2(utut),𝕋1/2djtdρtdρt\displaystyle=\int_{\Omega}\left\langle\frac{\nabla\varrho_{t}}{\varrho_{t}}+\nabla V,\frac{\mathop{}\!\mathrm{d}j_{t}}{\mathop{}\!\mathrm{d}\rho_{t}}\right\rangle\mathop{}\!\mathrm{d}\rho_{t}=\int_{\Omega}-\left\langle-\mathbb{T}^{1/2}\left(\frac{\nabla u_{t}}{u_{t}}\right),\mathbb{T}^{-1/2}\frac{\mathop{}\!\mathrm{d}j_{t}}{\mathop{}\!\mathrm{d}\rho_{t}}\right\rangle\mathop{}\!\mathrm{d}\rho_{t}
12Ωdjtdρt,𝕋1djtdρtdρt12Ωutut,𝕋(utut)dρt=(ρt,jt)𝒟(ρt).\displaystyle\geq-\frac{1}{2}\int_{\Omega}\left\langle\frac{\mathop{}\!\mathrm{d}j_{t}}{\mathop{}\!\mathrm{d}\rho_{t}},\mathbb{T}^{-1}\frac{\mathop{}\!\mathrm{d}j_{t}}{\mathop{}\!\mathrm{d}\rho_{t}}\right\rangle\mathop{}\!\mathrm{d}\rho_{t}-\frac{1}{2}\int_{\Omega}\left\langle\frac{\nabla u_{t}}{u_{t}},\mathbb{T}\left(\frac{\nabla u_{t}}{u_{t}}\right)\right\rangle\mathop{}\!\mathrm{d}\rho_{t}=-\mathcal{R}(\rho_{t},j_{t})-\mathcal{D}(\rho_{t}).

We finally obtain the asserted inequality after integrating over time and rearranging the terms. ∎

6.3. Proof of Theorem A

We now have all the ingredients to summarize the proof of Theorem A.

Proof of Theorem A.

Consider a family {(ρh,jh)}h>0\{(\rho^{h},j^{h})\}_{h>0} of GGF-solutions to (fKh) according to Definition 3.2. Let {(ρ^h,ȷ^h)}h>0\{(\hat{\rho}^{h},\hat{\jmath}^{h})\}_{h>0} be defined as in (6.1). Then, the existence of a subsequential limit pair (ρ,j)CE(\rho,j)\in CE and the convergence specified in Theorem A(1) follow from Lemma 4.4 and Theorem 4.8.

The lim inf\liminf inequality from assertion (2) is proven in Theorem 6.2, and it immediately follows that (ρ,j)lim infh0h(ρh,jh)=0\mathcal{I}(\rho,j)\leq\liminf_{h\to 0}\mathcal{I}_{h}(\rho^{h},j^{h})=0. On the other hand, (ρ,j)0\mathcal{I}(\rho,j)\geq 0 by the chain rule estimate proven in Theorem 6.4. Therefore, the limit pair (ρ,j)(\rho,j) is the (,,)(\mathcal{E},\mathcal{R},\mathcal{R}^{*})-gradient flow solution of (2.1) in the sense of Definition 3.7. ∎

Appendix A Properties of Gamma-limits as set functions

Proposition 5.9.

The functional supμ\mathcal{F}_{\sup}^{\mu} defined in (5.1) has the following properties:

  1. (i)

    Inner regularity:For any vH1(Ω,μ)v\in H^{1}(\Omega,\mu) and for any A𝒪A\in\mathcal{O} it holds that

    supA⊂⊂Asupμ(v,A)=supμ(v,A);\sup_{A^{\prime}\subset\joinrel\subset A}\mathcal{F}^{\mu}_{\sup}(v,A^{\prime})=\mathcal{F}^{\mu}_{\sup}(v,A);
  2. (ii)

    Subadditivity: For any vH1(Ω,μ)v\in H^{1}(\Omega,\mu) and for any A,A,B,B𝒪A,A^{\prime},B,B^{\prime}\in\mathcal{O} such that A⊂⊂AA^{\prime}\subset\joinrel\subset A and B⊂⊂BB^{\prime}\subset\joinrel\subset B it holds that:

    supμ(v,AB)supμ(v,A)+supμ(v,B);\mathcal{F}^{\mu}_{\sup}(v,A^{\prime}\cup B^{\prime})\leq\mathcal{F}^{\mu}_{\sup}(v,A)+\mathcal{F}^{\mu}_{\sup}(v,B);
  3. (iii)

    Locality: For any A𝒪A\in\mathcal{O} and any v,ψH1(Ω,μ)v,\psi\in H^{1}(\Omega,\mu) such that v=wv=w μ\mu-a.e. on AA there holds

    supμ(v,A)=supμ(w,A).\mathcal{F}^{\mu}_{\sup}(v,A)=\mathcal{F}^{\mu}_{\sup}(w,A).
Proof.

(i) Inner regularity. It is enough to prove that

supA⊂⊂Asupμ(v,A)supμ(v,A),\sup_{A^{\prime}\subset\joinrel\subset A}\mathcal{F}^{\mu}_{\sup}(v,A^{\prime})\geq\mathcal{F}^{\mu}_{\sup}(v,A),

because the opposite inequality holds since supμ(v,)\mathcal{F}^{\mu}_{\sup}(v,\cdot) is an increasing set function. Note that supμ(v,A)\mathcal{F}^{\mu}_{\sup}(v,A) is finite for any vH1(Ω,μ)v\in H^{1}(\Omega,\mu) (Lemma 5.8).

First we need to choose two sequences (v^h)(\hat{v}^{h}) and (w^h)(\hat{w}^{h}) which converge to vv in L2(Ω,μ)L^{2}(\Omega,\mu). To define the first sequence we fix some δ>0\delta>0 and choose A′′⊂⊂AA^{\prime\prime}\subset\joinrel\subset A such that

A\A′′¯|v|2dμδ.\int_{A\backslash\overline{A^{\prime\prime}}}|\nabla v|^{2}\mathop{}\!\mathrm{d}\mu\leq\delta.

Then by definition of supμ\mathcal{F}^{\mu}_{\sup} there exists a sequence (v^h)(\hat{v}^{h}) such that

lim suph0~hμ(v^h,A\A′′¯)=supμ(v,A\A′′¯)CA\A′′¯|v|2dμCδ,\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A\backslash\overline{A^{\prime\prime}})=\mathcal{F}^{\mu}_{\sup}(v,A\backslash\overline{A^{\prime\prime}})\leq C\int_{A\backslash\overline{A^{\prime\prime}}}|\nabla v|^{2}\mathop{}\!\mathrm{d}\mu\leq C\delta,

where the upper bound was shown in Lemma 5.8. To define the second sequence (w^h)(\hat{w}^{h}) let A𝒪A^{\prime}\in\mathcal{O} be such that A′′⊂⊂A⊂⊂AA^{\prime\prime}\subset\joinrel\subset A^{\prime}\subset\joinrel\subset A. Again by definition, we find a sequence (w^h)(\hat{w}^{h}) such that w^hv\hat{w}^{h}\to v in L2(Ω,μ)L^{2}(\Omega,\mu) and

lim suph0~hμ(w^h,A)=supμ(v,A).\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A^{\prime})=\mathcal{F}^{\mu}_{\sup}(v,A^{\prime}).

Notice that both ~hμ(v^h,A\A′′¯)\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A\backslash\overline{A^{\prime\prime}}) and ~hμ(w^h,A)\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A^{\prime}) are finite for h1h\ll 1 sufficiently small, which necessarily implies that v^h\hat{v}^{h} and w^h\hat{w}^{h} are piecewise constant functions on 𝒯h\mathcal{T}^{h}. Everywhere in this proof, we use notation with ”hats” and superscript hh (for example, v^h\hat{v}^{h}) for functions in PC(𝒯h)\text{PC}(\mathcal{T}^{h}) and superscript hh (for example, vhv^{h}) for the corresponding discrete functions on 𝒯h\mathcal{T}^{h}.

Next, we construct a sequence that ”interpolates” between (v^h)(\hat{v}^{h}) and (w^h)(\hat{w}^{h}). Set ε:=dist(A′′,Ac)\varepsilon:=\text{dist}\left(A^{\prime\prime},A^{\prime c}\right) and define sets Ai:={xA:dist(x,A′′)<iε/N}A_{i}:=\{x\in A^{\prime}:~{}\text{dist}(x,A^{\prime\prime})<i\varepsilon/N\} for i{1,,N}i\in\{1,\dots,N\}. Note that the following inclusions hold A′′⊂⊂A1⊂⊂⊂⊂AN⊂⊂AA^{\prime\prime}\subset\joinrel\subset A_{1}\subset\joinrel\subset\dots\subset\joinrel\subset A_{N}\subset\joinrel\subset A^{\prime}. Denote by φiN\varphi_{i}^{N} a cut-off function between AiA_{i} and Ai+1A_{i+1}, i.e. φiN𝒞c(Ai+1)\varphi_{i}^{N}\in\mathcal{C}_{c}^{\infty}(A_{i+1}), 0φiN10\leq\varphi_{i}^{N}\leq 1 on Ω\Omega, and φiN=1\varphi_{i}^{N}=1 in a neighborhood of Ai¯\overline{A_{i}}. with φiNsup2N/ε\|\nabla\varphi_{i}^{N}\|_{\sup}\leq 2N/\varepsilon. It has a piecewise constant approximation φ^iN,h:=𝕃hhφiN\hat{\varphi}^{N,h}_{i}:=\mathbb{L}_{h}\mathbb{P}_{h}\varphi_{i}^{N}, which satisfies φ^iN,h=1\hat{\varphi}^{N,h}_{i}=1 on (Ai1)𝒯h(A_{i-1})_{\mathcal{T}^{h}} and φ^iN,h=0\hat{\varphi}^{N,h}_{i}=0 on (AAi+2¯)𝒯h(A\setminus\overline{A_{i+2}})_{\mathcal{T}^{h}} for h<h0N:=ε/(3N)h<h_{0}^{N}:=\varepsilon/(3N). Now define

w^iN,h:=φ^iN,hw^h+(1φ^iN,h)v^h,i=1,,N.\hat{w}^{N,h}_{i}:=\hat{\varphi}^{N,h}_{i}\hat{w}^{h}+(1-\hat{\varphi}^{N,h}_{i})\hat{v}^{h},\qquad i=1,\ldots,N.

Observe that since φ^iN,h\hat{\varphi}^{N,h}_{i} converges pointwisely uniformly to φi\varphi_{i} as h0h\to 0, the sequence (w^iN,h)(\hat{w}^{N,h}_{i}) still converges to vv in L2(Ω,μ)L^{2}(\Omega,\mu) as h0h\to 0 for any ii\in\mathbb{N}.

For h<h0Nh<h_{0}^{N} sufficiently small, the following holds:

~hμ(w^iN,h,A)\displaystyle\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i},A) =~hμ(w^h,Ai1)+~hμ(v^h,AAi+2¯)+~hμ(w^iN,h,Ai+2¯Ai1)\displaystyle=\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A_{i-1})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A\setminus\overline{A_{i+2}})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i},\overline{A_{i+2}}\setminus A_{i-1})
~hμ(w^h,A)+~hμ(v^h,A\A′′)+~hμ(w^iN,h,GiN,h),\displaystyle\leq\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A^{\prime})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A\backslash A^{\prime\prime})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}_{i}^{N,h},G_{i}^{N,h}),

where

GiN,h:=int(Ai+2¯Ai1)+Bh(0)=(Ai+2Ai1¯)+Bh(0)Ai+3\Ai2¯.G_{i}^{N,h}:=\text{int}\,(\overline{A_{i+2}}\setminus A_{i-1})+B_{h}(0)=(A_{i+2}\setminus\overline{A_{i-1}})+B_{h}(0)\subset A_{i+3}\backslash\overline{A_{i-2}}.

We are now left to estimate the last term in the previous inequality.

We begin by bounding the discrete gradient of wihw_{i}^{h} by

|¯wiN,h|(K,L)\displaystyle|\overline{\nabla}w_{i}^{N,h}|(K,L) =|¯(φiN,hwh+(1φiN,h)vh)|(K,L)\displaystyle=\left|\overline{\nabla}(\varphi^{N,h}_{i}w^{h}+(1-\varphi^{N,h}_{i})v^{h})\right|(K,L)
=|(wh(K)vh(K))¯φiN,h(K,L)+φiN,h(L)¯wh(K,L)+(1φiN,h(L))¯vh(K,L)|\displaystyle=\left|\bigr{(}w^{h}(K)-v^{h}(K)\bigl{)}\overline{\nabla}\varphi^{N,h}_{i}(K,L)+\varphi^{N,h}_{i}(L)\overline{\nabla}w^{h}(K,L)+(1-\varphi^{N,h}_{i}(L))\overline{\nabla}v^{h}(K,L)\right|
|¯φiN,h|(K,L)|wh(K)vh(K)|+|¯wh|(K,L)+|¯vh|(K,L).\displaystyle\leq|\overline{\nabla}\varphi^{N,h}_{i}|(K,L)|w^{h}(K)-v^{h}(K)|+|\overline{\nabla}w^{h}|(K,L)+|\overline{\nabla}v^{h}|(K,L).

By Lemma 4.1 and since φsup2N/ε\|\nabla\varphi\|_{\sup}\leq 2N/\varepsilon, then |¯φiN,h|2CrNh/ε|\overline{\nabla}\varphi^{N,h}_{i}|\leq 2C_{r}Nh/\varepsilon, therefore

(K,L)Σh|Gi|¯φiN,h(K,L)|2|vh(K)wh(K)|2μh(K)κh(K,L)\displaystyle\sum_{(K,L)\in\Sigma^{h}|_{G_{i}}}|\overline{\nabla}\varphi^{N,h}_{i}(K,L)|^{2}|v^{h}(K)-w^{h}(K)|^{2}\mu^{h}(K)\kappa^{h}(K,L)
4Cr2N2ε2h2(K,L)Σh|Giκh(K,L)K|v^h(x)w^h(x)|2μ(dx)\displaystyle\hskip 60.00009pt\leq\frac{4C_{r}^{2}N^{2}}{\varepsilon^{2}}h^{2}\sum_{(K,L)\in\Sigma^{h}|_{G_{i}}}\kappa^{h}(K,L)\int_{K}|\hat{v}^{h}(x)-\hat{w}^{h}(x)|^{2}\mu(\mathop{}\!\mathrm{d}x)
4Cr2N2ε2Cκv^hw^hL2(Ω,μ)2,\displaystyle\hskip 60.00009pt\leq\frac{4C_{r}^{2}N^{2}}{\varepsilon^{2}}C_{\kappa}\|\hat{v}^{h}-\hat{w}^{h}\|^{2}_{L^{2}(\Omega,\mu)},

where we used the upper bound assumption (UB). On the other hand, for any η>0\eta>0, we can choose h=hN,η<h0Nh=h^{N,\eta}<h_{0}^{N} such that

~hμ(w^h,GiN,h)+~hμ(v^h,GiN,h)~hμ(w^h,Ai+2Ai1¯)+~hμ(v^h,Ai+2Ai1¯)+η.\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},G_{i}^{N,h})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},G_{i}^{N,h})\leq\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A_{i+2}\setminus\overline{A_{i-1}})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A_{i+2}\setminus\overline{A_{i-1}})+\eta.

In particular, we can choose η=ηN\eta=\eta_{N} depending on NN such that ηN0\eta_{N}\to 0 as NN\to\infty.

Making use of these estimates gives

~hμ(w^iN,h,GiN,h)\displaystyle\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i},G_{i}^{N,h}) 3[~hμ(w^h,Ai+2Ai1¯)+~hμ(v^h,Ai+2Ai1¯)+CNε2v^hw^hL2(Ω,μ)2+η],\displaystyle\leq 3\left[\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A_{i+2}\setminus\overline{A_{i-1}})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A_{i+2}\setminus\overline{A_{i-1}})+\frac{C_{N}}{\varepsilon^{2}}\|\hat{v}^{h}-\hat{w}^{h}\|^{2}_{L^{2}(\Omega,\mu)}+\eta\right],

with CN=4CκCr2N2C_{N}=4C_{\kappa}C_{r}^{2}N^{2}. Choosing i(h){1,,N3}i(h)\in\{1,\dots,N-3\} such that

~hμ(w^i(h)N,h,A)1N3j=1N3~hμ(w^jN,h,A),\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i(h)},A)\leq\frac{1}{N-3}\sum_{j=1}^{N-3}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{j},A),

we then obtain

~hμ(w^i(h)N,h,A)\displaystyle\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i(h)},A) ~hμ(w^h,A)+~hμ(v^h,A\A′′)+1N3j=1N3~hμ(w^jN,h,GjN,h).\displaystyle\leq\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A^{\prime})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A\backslash A^{\prime\prime})+\frac{1}{N-3}\sum_{j=1}^{N-3}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{j},G^{N,h}_{j}).

Combining the estimates together, we have

1N3j=1N3~hμ(w^jN,h,GjN,h)\displaystyle\frac{1}{N-3}\sum_{j=1}^{N-3}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{j},G^{N,h}_{j}) 3[~hμ(w^h,A\A′′)+~hμ(v^h,A\A′′)N3+CNε2v^hw^hL2(Ω,μ)2+ηN].\displaystyle\leq 3\left[\frac{\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A^{\prime}\backslash A^{\prime\prime})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A\backslash A^{\prime\prime})}{N-3}+\frac{C_{N}}{\varepsilon^{2}}\|\hat{v}^{h}-\hat{w}^{h}\|^{2}_{L^{2}(\Omega,\mu)}+\eta_{N}\right].

Taking the limit superior gives

supμ(v,A)\displaystyle\mathcal{F}^{\mu}_{\sup}(v,A) lim suph0~hμ(w^i(h)N,h,A)\displaystyle\leq\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i(h)},A)
supμ(v,A)+supμ(v,A\A′′)+3N3[lim suph0~hμ(w^h,A\A′′)+supμ(v,A\A′′)]+3ηN\displaystyle\leq\mathcal{F}^{\mu}_{\sup}(v,A^{\prime})+\mathcal{F}^{\mu}_{\sup}(v,A\backslash A^{\prime\prime})+\frac{3}{N-3}\left[\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A^{\prime}\backslash A^{\prime\prime})+\mathcal{F}^{\mu}_{\sup}(v,A\backslash A^{\prime\prime})\right]+3\eta_{N}
supA⊂⊂Asupμ(v,A)+Cδ+3N3[lim suph0~hμ(w^h,A\A′′)+Cδ]+3ηN.\displaystyle\leq\sup_{A^{\prime}\subset\joinrel\subset A}\mathcal{F}^{\mu}_{\sup}(v,A^{\prime})+C\delta+\frac{3}{N-3}\left[\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A^{\prime}\backslash A^{\prime\prime})+C\delta\right]+3\eta_{N}.

By sending δ0\delta\to 0 and NN\to\infty, we eventually conclude

supμ(v,A)supA⊂⊂Asupμ(v,A),\mathcal{F}^{\mu}_{\sup}(v,A)\leq\sup_{A^{\prime}\subset\joinrel\subset A}\mathcal{F}^{\mu}_{\sup}(v,A^{\prime}),

thereby concluding the proof of inner regularity.

(ii) Subadditivity. The proof follows in a similar fashion as in (1). We begin by choosing two sequences (v^h)(\hat{v}^{h}) and (w^h)(\hat{w}^{h}) converging to vv in L2(Ω,μ)L^{2}(\Omega,\mu) such that

lim suph0~hμ(v^h,A)=supμ(v,A)andlim suph0~hμ(w^h,B)=supμ(v,B).\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A)=\mathcal{F}^{\mu}_{\sup}(v,A)\qquad\text{and}\qquad\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},B)=\mathcal{F}^{\mu}_{\sup}(v,B).

Set ε:=dist(A,Ac)\varepsilon:=\text{dist}\left(A^{\prime},A^{c}\right) and define sets Ai:={xA:dist(x,A)<iε/N}A_{i}:=\{x\in A:~{}\text{dist}(x,A^{\prime})<i\varepsilon/N\} for i{1,,N}i\in\{1,\dots,N\}. Note that the following inclusions hold A⊂⊂A1⊂⊂⊂⊂AN⊂⊂AA^{\prime}\subset\joinrel\subset A_{1}\subset\joinrel\subset\dots\subset\joinrel\subset A_{N}\subset\joinrel\subset A. Let φiN\varphi_{i}^{N} be a cut-off function between AiA_{i} and Ai+1A_{i+1} with φiN2N/ε\|\nabla\varphi_{i}^{N}\|\leq 2N/\varepsilon. We use the piecewise constant approximation φ^iN,h\hat{\varphi}^{N,h}_{i} to define the sequence:

w^iN,h:=φ^iN,hv^h+(1φ^iN,h)w^h.\hat{w}^{N,h}_{i}:=\hat{\varphi}^{N,h}_{i}\hat{v}^{h}+(1-\hat{\varphi}^{N,h}_{i})\hat{w}^{h}.

For h1h\ll 1 sufficiently small, it holds that

~hμ(w^iN,h,AB)\displaystyle\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i},A^{\prime}\cup B^{\prime}) ~hμ(v^h,A)+~hμ(w^h,B)+~hμ(w^iN,h,GiN,h),\displaystyle\leq\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A)+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},B)+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i},G_{i}^{N,h}),

with GiN,hG_{i}^{N,h} as in (1). The last term on the right-hand side may be estimated as in (1) to obtain

~hμ(w^iN,h,GiN,h)\displaystyle\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i},G_{i}^{N,h}) 3[~hμ(w^h,Ai+2Ai1¯)+~hμ(v^h,Ai+2Ai1¯)+CNε2v^hw^hL2(Ω,μ)2+ηN].\displaystyle\leq 3\left[\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A_{i+2}\setminus\overline{A_{i-1}})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A_{i+2}\setminus\overline{A_{i-1}})+\frac{C_{N}}{\varepsilon^{2}}\|\hat{v}^{h}-\hat{w}^{h}\|^{2}_{L^{2}(\Omega,\mu)}+\eta_{N}\right].

Choosing i(h)i(h) such that

~hμ(w^i(h)N,h,AB)1Nj=1N~hμ(w^jN,h,AB),\displaystyle\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i(h)},A^{\prime}\cup B^{\prime})\leq\frac{1}{N}\sum_{j=1}^{N}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{j},A^{\prime}\cup B^{\prime}),

we then obtain

~hμ(w^i(h)N,h,AB)\displaystyle\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i(h)},A^{\prime}\cup B^{\prime}) ~hμ(v^h,A)+~hμ(w^h,B)+1Nj=1N~hμ(w^jN,h,GjN,h),\displaystyle\leq\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A)+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},B)+\frac{1}{N}\sum_{j=1}^{N}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{j},G_{j}^{N,h}),

where the last term may be estimated by

1Nj=1N~hμ(w^jN,h,GjN,h)\displaystyle\frac{1}{N}\sum_{j=1}^{N}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{j},G_{j}^{N,h}) 3[~hμ(w^h,A\A)+~hμ(v^h,A\A)N+CNε2v^hw^hL2(Ω,μ)2+ηN].\displaystyle\leq 3\left[\frac{\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{h},A\backslash A^{\prime})+\tilde{\mathcal{F}}_{h}^{\mu}(\hat{v}^{h},A\backslash A^{\prime})}{N}+\frac{C_{N}}{\varepsilon^{2}}\|\hat{v}^{h}-\hat{w}^{h}\|^{2}_{L^{2}(\Omega,\mu)}+\eta_{N}\right].

Taking the limit superior as h0h\to 0 gives

supμ(v,AB)\displaystyle\mathcal{F}^{\mu}_{\sup}(v,A^{\prime}\cup B^{\prime}) lim suph0~hμ(w^i(h)N,h,AB)\displaystyle\leq\limsup_{h\to 0}\tilde{\mathcal{F}}_{h}^{\mu}(\hat{w}^{N,h}_{i(h)},A^{\prime}\cup B^{\prime})
supμ(v,A)+supμ(v,B)+o(1)|N.\displaystyle\leq\mathcal{F}^{\mu}_{\sup}(v,A)+\mathcal{F}^{\mu}_{\sup}(v,B)+o(1)|_{N\to\infty}.

By sending NN\to\infty and applying the inner regularity property, we conclude

supμ(v,AB)supμ(v,A)+supμ(v,B).\mathcal{F}^{\mu}_{\sup}(v,A\cup B)\leq\mathcal{F}^{\mu}_{\sup}(v,A)+\mathcal{F}^{\mu}_{\sup}(v,B).

(iii) Locality. We first prove supμ(v,A)supμ(w,A)\mathcal{F}^{\mu}_{\sup}(v,A)\leq\mathcal{F}^{\mu}_{\sup}(w,A). The argument is similar to the previous points. For a fixed δ>0\delta>0 there exists Aδ⊂⊂AA_{\delta}\subset\joinrel\subset A such that A\Aδ¯|v|2dμ<δ\int_{A\backslash\overline{A_{\delta}}}|\nabla v|^{2}\mathop{}\!\mathrm{d}\mu<\delta. We choose two sequences (v^h)(\hat{v}^{h}) and (w^h)(\hat{w}^{h}) such that v^hv\hat{v}^{h}\to v, w^hw\hat{w}^{h}\to w in L2(Ω,μ)L^{2}(\Omega,\mu) satisfying

lim suph0~hμ(v^h,A\Aδ¯)\displaystyle\limsup_{h\to 0}\tilde{\mathcal{F}}^{\mu}_{h}(\hat{v}^{h},A\backslash\overline{A_{\delta}}) =supμ(v,A)CA\Aδ¯|v|2dμ<Cδ,\displaystyle=\mathcal{F}^{\mu}_{\sup}(v,A)\leq C\int_{A\backslash\overline{A_{\delta}}}|\nabla v|^{2}\mathop{}\!\mathrm{d}\mu<C\delta,
lim suph0~hμ(w^h,A)\displaystyle\limsup_{h\to 0}\tilde{\mathcal{F}}^{\mu}_{h}(\hat{w}^{h},A) =supμ(w,A).\displaystyle=\mathcal{F}^{\mu}_{\sup}(w,A).

Set ε:=dist(Aδ,Ac)\varepsilon:=\text{dist}\left(A_{\delta},A^{c}\right) and define sets Ai:={xA:dist(x,Aδ)<iε/N}A_{i}:=\{x\in A:~{}\text{dist}(x,A_{\delta})<i\varepsilon/N\} for i{1,,N}i\in\{1,\dots,N\}. Denote by φiN\varphi_{i}^{N} a cut-off function between AiA_{i} and Ai+1A_{i+1} with φiNsup2N/ε\|\nabla\varphi_{i}^{N}\|_{\sup}\leq 2N/\varepsilon. It has a piecewise constant approximation φ^iN,h:=𝕃hhφiN\hat{\varphi}^{N,h}_{i}:=\mathbb{L}_{h}\mathbb{P}_{h}\varphi_{i}^{N}. Then, we define

w^iN,h:=φ^iN,hw^h+(1φ^iN,h)v^h,\hat{w}^{N,h}_{i}:=\hat{\varphi}^{N,h}_{i}\hat{w}^{h}+(1-\hat{\varphi}^{N,h}_{i})\hat{v}^{h},

with (w^iN,h)(\hat{w}^{N,h}_{i}) still converging to ww in L2(Ω,μ)L^{2}(\Omega,\mu) as h0h\to 0 for any i=1,,Ni=1,\ldots,N. Similar to the proof of inner regularity, we obtain the existence of some i(h){1,,N}i(h)\in\{1,\ldots,N\} such that

lim suph0~hμ(w^i(h)N,h,A)\displaystyle\limsup_{h\to 0}\tilde{\mathcal{F}}^{\mu}_{h}(\hat{w}^{N,h}_{i(h)},A) supμ(w,A)+O(δ)|δ0+o(1)|N.\displaystyle\leq\mathcal{F}^{\mu}_{\sup}(w,A)+O(\delta)|_{\delta\to 0}+o(1)|_{N\to\infty}.

Passing to the limits δ0\delta\to 0 and NN\to\infty then yields

supμ(v,A)supμ(w,A).\mathcal{F}^{\mu}_{\sup}(v,A)\leq\mathcal{F}^{\mu}_{\sup}(w,A).

The assertion follows as we can swap the roles of vv and ww. ∎

We recall that

π,φ(v,A):={π(v,A)if vφH01(A),+otherwise,\mathcal{F}^{\pi,\varphi}(v,A):=\begin{cases}\mathcal{F}^{\pi}(v,A)&\text{if }v-\varphi\in H^{1}_{0}(A),\\ +\infty&\text{otherwise,}\end{cases}

and the corresponding discrete counterpart of π,φ\mathcal{F}^{\pi,\varphi} is

hπ,φ(vh,A):={hπ(vh,A)if vh=hφ=:φh on 𝒯h|Ac,+otherwise.\mathcal{F}^{\pi,\varphi}_{h}(v^{h},A):=\begin{cases}\mathcal{F}^{\pi}_{h}(v^{h},A)&\text{if }v^{h}=\mathbb{P}_{h}\varphi=:\varphi^{h}\text{ on }\mathcal{T}^{h}|_{A^{c}},\\ +\infty&\text{otherwise.}\end{cases}
Proposition 5.19.

Let A𝒪A\in\mathcal{O} be arbitrary with Lipschitz boundary and φH1(Ω)\varphi\in H^{1}(\Omega). For any sequence (hπ,φ(,A))(\mathcal{F}^{\pi,\varphi}_{h}(\cdot,A)) there exists a subsequence that Γ\Gamma-converges in the L2(Ω)L^{2}(\Omega)-topology to π,φ(,A)\mathcal{F}^{\pi,\varphi}(\cdot,A).

Proof.

Let us first prove the Γ\Gamma-lim inf\liminf inequality. We consider a sequence 𝕃hvhv\mathbb{L}_{h}v^{h}\to v in L2(Ω)L^{2}(\Omega) such that suph>0hπ,φ(vh,A)<\sup_{h>0}\mathcal{F}^{\pi,\varphi}_{h}(v^{h},A)<\infty. This implies that vh=φhv^{h}=\varphi^{h} on 𝒯h|Ac\mathcal{T}^{h}|_{A^{c}} and hπ,φ(vh,A)=hπ(vh,A)\mathcal{F}^{\pi,\varphi}_{h}(v^{h},A)=\mathcal{F}^{\pi}_{h}(v^{h},A). Consequently, we also have that suph>0hπ(vh,A)<\sup_{h>0}\mathcal{F}^{\pi}_{h}(v^{h},A)<\infty. By the same argument as in Lemma 5.17, we deduce that vH1(A)v\in H^{1}(A). Since Γ\Gamma-limhπ(,A)=π(,A)\lim\mathcal{F}^{\pi}_{h}(\cdot,A)=\mathcal{F}^{\pi}(\cdot,A) it remains to prove that vφH01(A)v-\varphi\in H_{0}^{1}(A).

Notice that AcA𝒯hcA^{c}\subset A^{c}_{\mathcal{T}^{h}} for all h>0h>0. Since vh=φhv^{h}=\varphi^{h} on 𝒯h|Ac\mathcal{T}^{h}|_{A^{c}}, their piecewise reconstructions satisfy 𝕃hvh=𝕃hφh\mathbb{L}_{h}v^{h}=\mathbb{L}_{h}\varphi^{h} on A𝒯hcA^{c}_{\mathcal{T}^{h}}, and hence also on AcA^{c} for all h>0h>0. Using the fact that 𝕃hvhv\mathbb{L}_{h}v^{h}\to v and 𝕃hφhφ\mathbb{L}_{h}\varphi^{h}\to\varphi in L2(Ω)L^{2}(\Omega), we easily deduce that v=φv=\varphi in L2(ΩA)L^{2}(\Omega\setminus A). The deduced regularity vH1(A)v\in H^{1}(A) and assumed regularity φH1(Ω)\varphi\in H^{1}(\Omega) then allows to conclude that vφH01(A)v-\varphi\in H_{0}^{1}(A).

Thus, the lim inf\liminf inequality follows:

lim infh0hπ,φ(vh,A)π,φ(v,A).\liminf_{h\to 0}\mathcal{F}^{\pi,\varphi}_{h}(v^{h},A)\geq\mathcal{F}^{\pi,\varphi}(v,A)\,.

Now we show the approximate lim sup\limsup inequality. Let vH1(Ω)v\in H^{1}(\Omega) such that supp(vφ)⊂⊂A\text{supp}(v-\varphi)\subset\joinrel\subset A. There exists a recovery sequence vhvv^{h}\to v in L2(Ω)L^{2}(\Omega) such that limh0hπ(vh,A)=π(v,A)=π,φ(v,A)\displaystyle\lim_{h\to 0}\mathcal{F}^{\pi}_{h}(v^{h},A)=\mathcal{F}^{\pi}(v,A)=\mathcal{F}^{\pi,\varphi}(v,A).

Set ε:=dist(supp(vφ),Ac)\varepsilon:=\text{dist}\left(\text{supp}(v-\varphi),A^{c}\right) and define sets Ai:={xA:dist(x,supp(vφ))<iε/N}A_{i}:=\{x\in A:~{}\text{dist}(x,\text{supp}(v-\varphi))<i\varepsilon/N\} for i{1,,N}i\in\{1,\dots,N\}. Denote by φiN\varphi_{i}^{N} a cut-off function between AiA_{i} and Ai+1A_{i+1} with φiNsup2N/ε\|\nabla\varphi_{i}^{N}\|_{\sup}\leq 2N/\varepsilon. It has a piecewise constant approximation φ^iN,h:=𝕃hhφiN\hat{\varphi}^{N,h}_{i}:=\mathbb{L}_{h}\mathbb{P}_{h}\varphi_{i}^{N}. Then, we define

w^iN,h:=φ^iN,hv^h+(1φ^iN,h)φ,\hat{w}^{N,h}_{i}:=\hat{\varphi}^{N,h}_{i}\hat{v}^{h}+(1-\hat{\varphi}^{N,h}_{i})\varphi,

with (w^iN,h)(\hat{w}^{N,h}_{i}) still converging to vv in L2(Ω)L^{2}(\Omega) as h0h\to 0 for any i=1,,Ni=1,\ldots,N. Similar to the proof of inner regularity, we obtain the existence of some i(h){1,,N}i(h)\in\{1,\ldots,N\} such that

lim suph0~hπ(w^i(h)N,h,A)lim suph0~hπ(vh,A)+o(1)|N.\limsup_{h\to 0}\tilde{\mathcal{F}}^{\pi}_{h}(\hat{w}^{N,h}_{i(h)},A)\leq\limsup_{h\to 0}\tilde{\mathcal{F}}^{\pi}_{h}(v^{h},A)+o(1)|_{N\to\infty}.

Passing NN\to\infty yields

lim suph0~hπ,φ(w^i(h)N,h,A)π,φ(v,A).\limsup_{h\to 0}\tilde{\mathcal{F}}^{\pi,\varphi}_{h}(\hat{w}^{N,h}_{i(h)},A)\leq\mathcal{F}^{\pi,\varphi}(v,A).

References

  • [1] R. Alicandro and M. Cicalese. A general integral representation result for continuum limits of discrete energies with superlinear growth. SIAM Journal on Mathematical Analysis, 36(1):1–37, 2004.
  • [2] H. W. Alt. Linear Functional Analysis: An Application-Oriented Introduction. Springer, 2016.
  • [3] L. Ambrosio, N. Gigli, and G. Savaré. Gradient Flows: in Metric Spaces and in the Space of Probability Measures. Springer Science & Business Media, 2008.
  • [4] S. Andres, A. Chiarini, and M. Slowik. Quenched local limit theorem for random walks among time-dependent ergodic degenerate weights. Probability Theory and Related Fields, 179(3):1145–1181, 2021.
  • [5] T. Barth, R. Herbin, and M. Ohlberger. Finite volume methods: foundation and analysis. Encyclopedia of Computational Mechanics Second Edition, pages 1–60, 2018.
  • [6] P. Billingsley. Convergence of Probability Measures. John Wiley & Sons, 1999.
  • [7] M. Biskup. Recent progress on the random conductance model. Probability Surveys, 8:294–373, 2011.
  • [8] G. Bouchitté, I. Fonseca, G. Leoni, and L. Mascarenhas. A global method for relaxation in W1,pW^{1,p} and in SBVpSBV_{p}. Archive for Rational Mechanics and Analysis, 165(3):187–242, 2002.
  • [9] A. Braides. Γ\Gamma-Convergence for Beginners, volume 22. Clarendon Press, 2002.
  • [10] A. Braides. A handbook of Γ\Gamma-convergence. In Handbook of Differential Equations: Stationary Partial Differential Equations, volume 3, pages 101–213. Elsevier, 2006.
  • [11] P. Caputo, A. Faggionato, and T. Prescott. Invariance principle for Mott variable range hopping and other walks on point processes. Ann. Inst. Henri Poincaré Probab. Stat., 49(3):654–697, 2013.
  • [12] S.-N. Chow, W. Huang, Y. Li, and H. Zhou. Fokker–Planck equations for a free energy functional or Markov process on a graph. Archive for Rational Mechanics and Analysis, 203(3):969–1008, 2012.
  • [13] D. A. Croydon and B. M. Hambly. Local limit theorems for sequences of simple random walks on graphs. Potential Analysis, 29(4):351–389, 2008.
  • [14] G. Dal Maso. An Introduction to Γ\Gamma-Convergence. Springer, 1993.
  • [15] K. Disser and M. Liero. On gradient structures for Markov chains and the passage to Wasserstein gradient flows. Networks & Heterogeneous Media, 10(2):233, 2015.
  • [16] J. Droniou, R. Eymard, T. Gallouët, C. Guichard, and R. Herbin. The Gradient Discretisation Method, volume 82 of Mathématiques & Applications (Berlin). Springer, Cham, 2018.
  • [17] L. C. Evans and R. F. Gariepy. Measure Theory and Fine Properties of Functions. CRC press, 2015.
  • [18] R. Eymard, J. Fuhrmann, and K. Gärtner. A finite volume scheme for nonlinear parabolic equations derived from one-dimensional local Dirichlet problems. Numerische Mathematik, 102(3):463–495, 2006.
  • [19] R. Eymard, T. Gallouët, and R. Herbin. Finite volume methods. Handbook of Numerical Analysis, 7:713–1018, 2000.
  • [20] A. Faggionato. Random walks and exclusion processes among random conductances on random infinite clusters: homogenization and hydrodynamic limit. Electronic Journal of Probability, 13:2217–2247, 2008.
  • [21] D. Forkert, J. Maas, and L. Portinale. Evolutionary Γ\Gamma-convergence of entropic gradient flow structures for Fokker-Planck equations in multiple dimensions. arXiv preprint arXiv:2008.10962, 2020.
  • [22] P. Gladbach, E. Kopfer, and J. Maas. Scaling limits of discrete optimal transport. SIAM Journal on Mathematical Analysis, 52(3):2759–2802, 2020.
  • [23] A. Gloria and F. Otto. An optimal variance estimate in stochastic homogenization of discrete elliptic equations. Ann. Probab., 39(3):779–856, 2011.
  • [24] M. Grmela. Weakly nonlocal hydrodynamics. Physical Review E, 47(1):351, 1993.
  • [25] B. Hambly and M. Barlow. Parabolic Harnack inequality and local limit theorem for percolation clusters. Electronic Journal of Probability, 14:1–26, 2009.
  • [26] M. Heida. Convergences of the square-root approximation scheme to the Fokker–Planck operator. Mathematical Models and Methods in Applied Sciences, 28(13):2599–2635, 2018.
  • [27] R. Jordan, D. Kinderlehrer, and F. Otto. The variational formulation of the Fokker–Planck equation. SIAM Journal on Mathematical Analysis, 29(1):1–17, 1998.
  • [28] M. Kantner. Generalized Scharfetter–Gummel schemes for electro-thermal transport in degenerate semiconductors using the Kelvin formula for the Seebeck coefficient. Journal of Computational Physics, 402:109091, 2020.
  • [29] P. M. Kotelenez. From discrete deterministic dynamics to Brownian motions. Stochastics and Dynamics, 5(03):343–384, 2005.
  • [30] S. M. Kozlov. The averaging method and walks in inhomogeneous environments. Uspekhi Mat. Nauk, 40(2(242)):61–120, 238, 1985.
  • [31] T. Kumagai. Random Walks on Disordered Media and Their Scaling Limits. Springer, 2014.
  • [32] J. Maas. Gradient flows of the entropy for finite Markov chains. Journal of Functional Analysis, 261(8):2250–2292, 2011.
  • [33] R. J. McCann. A convexity principle for interacting gases. Advances in Mathematics, 128(1):153–179, 1997.
  • [34] M. M. Meerschaert and H.-P. Scheffler. Limit theorems for continuous-time random walks with infinite mean waiting times. Journal of Applied Probability, 41(3):623–638, 2004.
  • [35] A. Mielke. Geodesic convexity of the relative entropy in reversible markov chains. Calculus of Variations and Partial Differential Equations, 48(1):1–31, 2013.
  • [36] A. Mielke. On evolutionary Γ\Gamma-convergence for gradient systems. In Macroscopic and Large Scale Phenomena: Coarse Graining, Mean Field Limits and Ergodicity, pages 187–249. Springer, 2016.
  • [37] A. Mielke, A. Montefusco, and M. A. Peletier. Exploring families of energy-dissipation landscapes via tilting: three types of EDP convergence. Continuum Mechanics and Thermodynamics, 33(3):611–637, 2021.
  • [38] A. Mielke, M. A. Peletier, and D. R. M. Renger. On the relation between gradient flows and the large-deviation principle, with applications to Markov chains and diffusion. Potential Analysis, 41(4):1293–1327, 2014.
  • [39] M. A. Peletier, R. Rossi, G. Savaré, and O. Tse. Jump processes as generalized gradient flows. Calculus of Variations and Partial Differential Equations, 61(1):1–85, 2022.
  • [40] R. Rossi and G. Savaré. Tightness, integral equicontinuity and compactness for evolution problems in Banach spaces. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 2(2):395–431, 2003.
  • [41] T. Sandev, R. Metzler, and A. Chechkin. From continuous-time random walks to the generalized diffusion equation. Fractional Calculus and Applied Analysis, 21(1):10–28, 2018.
  • [42] E. Sandier and S. Serfaty. Γ\Gamma-convergence of gradient flows with applications to Ginzburg–Landau. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 57(12):1627–1672, 2004.
  • [43] C. Schuette and P. Metzner. Markov chains and jump processes. Freie Universitaet Berlin, 2009.
  • [44] S. Serfaty. Γ\Gamma-convergence of gradient flows on Hilbert and metric spaces and applications. Discrete & Continuous Dynamical Systems, 31(4):1427, 2011.
  • [45] S. R. S. Varadhan and H.-T. Yau. Diffusive limit of lattice gas with mixing conditions. Asian Journal of Mathematics, 1(4):623–678, 1997.