Diffusive limit of random walks on tessellations via generalized gradient flows
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 convergence1. 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 . 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.




To clarify the goal, we now introduce our setup: Let be a family of finite tessellations of a bounded and convex set , where is the family of cells and is identified with the set of faces (cf. Section 2.1 for the precise definition of cells and faces), and be transition kernels for the random walk. Examples of tessellations are shown in Figure 1. The small parameter 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 is known to satisfy the forward Kolmogorov equation (see for example [43, Section 6.3])
(fKh) |
with being the dual of the generator given, for any bounded function , by
where the sum is taken over all elements in the set of adjacent cells . Here, we restrict ourselves to random walks satisfying detailed balance, i.e. random walks admitting a stationary measure such that
The questions we set out to answer are then:
-
(a)
Under which sufficient conditions on and does the family of solutions to (fKh) converge to a non-degenerate diffusion process?
-
(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 , which we describe briefly in the following (see Section 3.1 for more details).
A gradient structure is completely defined by the driving energy and the dual dissipation potential . Here, and denote the spaces of probability measures and bounded measurable functions on respectively. In the case of the random walk, the evolution is driven by the relative entropy with respect to the stationary measure :
with the energy density . As for the dual dissipation potential , a range of choices can give rise to a gradient structure for (fKh). The general form of studied in [39] is
In this work, we make use of the so-called ‘cosh’ gradient structure, for which
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 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 and , one can express (fKh) in the form of a continuity equation (CEh) and the force-flux relation (FFh) for the density-flux pair :
(CEh) | ||||
(FFh) |
where the flux is written in terms of and acting on . Here, denotes the derivative in the second variable, is the graph gradient and 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 , the pair satisfies
(EDBh) |
with and being Legendre–Fenchel duality pairs w.r.t. the second variable.
When the following chain rule applies for all pairs satisfying (CEh)
(CRh) |
then one also has that . In particular, a pair satisfying (CEh) and (EDBh) is a minimizer of , which we use to define generalized gradient flow (GGF) solutions to (fKh):
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 -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)
-
(2)
Prove lim inf inequalities for all the functionals in the energy-dissipation functional to recover a limiting energy-dissipation functional :
-
(3)
Recover a limiting diffusion equation of the type
(fK) with some generator from the limiting energy-dissipation functional .
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 belong to different spaces for each . Moreover, in the diffusive limit, we expect to obtain a curve satisfying the continuity equation with the usual divergence operator:
(CE) |
For this purpose, we introduce a continuous reconstruction procedure in Section 4 such that any pair satisfying (CEh) induces a pair 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 satisfying (CE).
To prove the liminf inequality in Step 2, we study all the components of the energy-dissipation functional separately. The lower semicontinuity of the driving energy follows from standard result, whereas the challenging part lies in proving the result for the dissipation potential and the Fisher information . For this purpose, we apply -convergence techniques to obtain and characterize the variational limits and of and respectively. Here the assumptions placed on the family of tessellations and transition kernels that we ask about in question (a) come into play, as the form of and depends strongly on the relationship between and .
To illustrate the idea, the discrete Fisher information for the ‘cosh’ gradient structure takes the form
We prove that, under suitable assumptions on the families , and , the family -converges to a limit functional of the form
where is a symmetric and positive definite tensor, and is the limit of the sequence . 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 and as objects related to a diffusion process on .
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 , 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) |
with , thereby answering question (b). If arises from a homogeneous random walk on a uniform lattice (see Example 2.8), then we arrive at (1.1) with , the identity tensor.
The techniques we use to prove the 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 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 .
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 -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 be an open bounded convex set. A tessellation of consists of a family of mutually disjoint cells (usually denoted by or ) that are open and convex sets in , and a family of pairs of neighboring cells , where is the -dimensional Hausdorff measure. Examples of suitable tessellations include Voronoi tessellations, and meshes commonly used in finite-volume methods. The common face of is denoted by . The characterizing size of a tessellation is its maximum diameter:
The maximum diameter gives an upper bound on the volumes of the cells and faces , where , are universal constants depending only on the spatial dimension . 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 such that (i) For each there is an inner ball with ; (ii) For every it holds that .
Remark 2.1.
The non-degeneracy assumption implies particularly that for all , and also provides a uniform bound on the cardinality of neighboring cells (cf. [22, Lemma 2.12]):
which follows from the following calculations:
Here, is the cardinality of the set .
Remark 2.2.
We now summarize the assumptions on the tessellations that is used within this paper.
Assumptions on . We assume the family of tessellations to be such that (Ass)
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 , with the geometric properties of the tessellation. In this section, we emphasize all the assumptions we need to make on in relation to .
Assumptions on . Let be a stationary measure for (fKh) satisfying the detailed balance condition (DB) We assume to have a density uniformly bounded from above and away from zero: (B) The continuous reconstruction (cf. Section 4) converges in the following sense: We further assume that .
Without loss of generality, and for simplicity, we assume to have unit mass.
Example 2.3.
A stationary measure satisfying the above mentioned assumptions can be obtained from a continuous measure . In practice, the stationary measure is usually given in terms of a potential , i.e. . In this case, we assume and set
Then converges to in the sense specified in Section 4.1, since
and, therefore, satisfies the required assumptions.
Now we introduce scaling assumptions on the joint measure defined in (DB).
Scaling of . We assume the existence of constants independent of : (B)
Remark 2.4.
We need one final assumption on the compatibility of the joint measure 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 not touching the boundary, i.e. , (A)
Similar assumptions to (A) 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 (A) 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 , and scaling of are required in throughout this paper, but (A) 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 are such that for all and , then A is satisfied.
2.3. Main result
Definition 2.6 (Admissible continuous reconstruction).
We state our main result in the following theorem.
Theorem A.
Let be a family of tessellations satisfying (Ass). Further, let be a family of GGF-solutions to (fKh) with satisfying (B), (B), and (A), and initial data satisfying
Then there exists a (not relabelled) subsequence of admissible continuous reconstructions and a limit pair such that
-
(1)
satisfies (CE) with the density and
-
(i)
strongly in for any ;
-
(ii)
in .
-
(i)
-
(2)
the following liminf estimate holds:
-
(3)
is the gradient flow solution with the energy-dissipation functional given as
where is a symmetric and positive definite diffusion tensor.
Remark 2.7.
The equation corresponding to the energy-dissipation balance given by is
(2.1) |
with the no-flux boundary conditions on , where is the potential corresponding to the limit stationary measure .
To characterize the diffusion tensor , we introduce a tensor :
where . Then is obtained as a limit of .
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 can be calculated explicitly.
Example 2.8 (Lattice ).
Consider the simplest tessellation which corresponds to the lattice with . We choose the uniform stationary measure for all and the uniform joint measure for all . It follows that the transition kernel is . Let be the basis vectors in . One can always choose the orientation of the basis to be such that for any the vectors pointing to the neighboring cells are . Then the diffusion tensor becomes the identity matrix for all and, consequently, .
We can also make a slightly different choice of the joint measure:
where are independent of . In this way, we make the jump intensities different in the different directions, i.e. if the face is orthogonal to . Note that (A) is still satisfied. In this case,
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 ).
Let be a tilted version of the lattice as shown in the Figure 3. The tilt is given by the parameter , , where corresponds to . Each cell has four neighbors , where the subscript stands for right, up, left, and down neighbors of .

We fix the basis such that . In this basis, we have that
The tensor then takes the form
Notice that for any nonnegative , we can never obtain . For uniform kernels , we get
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 . Consider the tessellation consisting of cells with length on , i.e. and the cells with the length on , i.e. (see Figure 4).

Consider for , which immediately implies that (A) is satisfied, with the uniform stationary measure for all . Then the tensor reads
In particular, for , , and for , . Therefore, in the limit , we obtain .
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
where denote the space bounded measurable functions on , and the space of finite (signed) measures equipped with the topology of weak∗ convergence, i.e. convergence against , the space of continuous functions that vanish at infinity. Furthermore, we denote by the total variation measure of a measure , and by the space of probability measures equipped with the topology of narrow convergence, i.e. convergence against , 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 , where
-
•
is a curve of measures defined on the tessellation , and
-
•
is a measurable family of fluxes with ,
a solution of the discrete continuity equation
(CEh) |
if for all and ,
(3.1) |
We denote by the set of solutions to the discrete continuity equation (CEh).
Definition 3.2 (GGF solutions).
A curve is said to be an -generalized gradient flow solution of (fKh) with initial data if
-
(i)
in ;
-
(ii)
there exists a measurable family such that with
where
i.e. is a lower-semicontinuous envelope of .
-
(iii)
the chain rule holds, i.e.
(CRh)
We now make specific choices for all the components of the energy-dissipation functional (EDBh) introduced in Section 1.
The driving energy is taken to be the relative entropy with respect to the stationary measure , i.e.
with the energy density .
The dual dissipation potential , as defined in the introduction, takes the form
where
The dissipation potential is the Legendre–Fenchel dual of w.r.t. to its second variable. In particular, it takes the explicit form
(3.2) |
where , , and
The Fisher information is defined as
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).
In the next lemma, we list the properties of that will be used in the proof of Lemma 4.4.
Lemma 3.4.
The Legendre-Fenchel pair are such that
-
(i)
is even and convex, , and is strictly increasing for .
-
(ii)
For the mapping is decreasing.
-
(iii)
For , has a strictly increasing inverse satisfying
-
(iv)
.
Proof.
(ii) By convexity for it holds that:
(iii) Since are convex conjugate, then
Therefore,
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) |
where is a bounded convex domain and , 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 is given by all pairs , where
-
•
is a curve of positive measures defined on , and
-
•
is a measurable family of fluxes with
satisfying the continuity equation
(CE) |
in the following sense:
(3.3) |
Remark 3.6.
Definition 3.7.
A curve is said to be an -gradient flow solution of (fK) with initial data if
-
(i)
in ;
-
(ii)
there exists a measurable family such that with
where
i.e. is a lower-semicontinuous envelope of .
-
(iii)
the following chain rule inequality holds:
According to the strategy explained in Section 1, we will obtain the energy-dissipation functional by proving the corresponding inequality for the discrete energy-dissipation functional introduced in Section 3.1. For a family of GGF solutions of (fKh), we will immediately have
Then to prove that the limit curve indeed satisfies Defintion 3.7, it is left to show that , 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 (cf. Section 4.1). We then provide a compactness result for the sequence of continuous reconstructions in Section 4.2.
4.1. Continuous reconstruction
Throughout this paper we will extensively use two operations: projecting functions supported on on the tessellation and lifting discrete functions supported on to . Specifically, we define the following operators
where is the set of functions that are piecewise-constant on cells .
The motivating idea for the reconstruction procedure is to embed the curve into the continuous space in such a way that the lifted curve belong to . Assuming that , we transform the left-hand side of (3.1) into
Defining the reconstructed measure via its density as
(4.1) |
we then obtain equality in the first parts of (3.1) and (3.3):
(4.2) |
In what follows we will also frequently use the formulation in terms of density with respect to the stationary measure :
with
Assuming the same relation between the test functions , 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 be the projection of on . Then there exists a vector-valued measure such that
(4.3) |
Furthermore,
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 and , we note that
Therefore, we define
(4.4) |
Proof of Lemma 4.1.
For any pair of neighboring cells :
where is an arbitrary coupling between the measures and . We assume that the coupling is produced by a transport map , meaning that for all there exist unique such that . In this case, the coupling has the form
For a smooth the fundamental theorem of calculus gives:
Rewriting the coupling in terms of the transport map yields:
Introducing the notation and we proceed:
Denoting by the measure we obtain (4.3).
To estimate the total variation of , we notice that
where . Therefore,
∎
Remark 4.2.
One can notice that the measures constructed in the proof are not uniquely defined due to the freedom in choosing transport maps . However, we will see that the compactness result in Lemma 4.4 does not depend on the specific choice of .
In the case of a lattice, the measure can be calculated explicitly.
Example 4.3.
Consider the tessellation . For any pair of neighboring cells and the optimal transport map is , with being the (outward) normal on the cell face , and, respectively, . The function has an inverse . Therefore, it is possible to calculate the measure explicitly:
Notice that for any the indicator function is equal to 1 for and equal to 0 afterwards. Therefore, for :
A similar property holds for :
We conclude that
4.2. Compactness
Throughout this section, we consider a family of the GGF-solutions to (fKh) with initial data satisfying . With the non-degeneracy assumption on , and the assumptions (B), (B) on , and , we deduce compactness for the continuous reconstructions of the solutions.
Lemma 4.4.
Let , , be defined as in (4.4). Then
-
(1)
the family
-
(2)
the family is equi-integrable.
In particular, there exists a Borel family such that
Proof.
Recall that for almost every ,
For any measurable set and any denote:
Note that multiplied by is uniformly bounded because of (UB):
Setting , we will show that the sequence of measures is uniformly integrable. The properties collected in Lemma 3.4(i) together with (UB) provide the following estimate:
Since , Lemma 3.4(ii) gives that:
Taking the inverse yields:
Since are generalized gradient flow solutions in the sense of Definition 3.2, the integral of is bounded uniformly in under the assumption on the initial conditions:
Now we use the upper bound from Lemma 3.4(iii) to obtain
Choosing , , and using the property , we find
Now let be arbitrary. By choosing such that , and subsequently such that , we then conclude that
Moreover, by applying the estimate above to , we simply obtain
i.e. the total variation of is uniformly bounded, which allows us to extract a converging subsequence (not relabelled) and some such that holds.
The equi-integrability of readily follows from the estimate above. Since the limit also satisfies the inequality above (weakly- lower-semicontinuity of the total variation), we conclude that on has Lebesgue density. By disintegration, has the representation for a Borel family . ∎
As a consequence of the previous lemma, we obtain the following result for density-flux pairs.
Lemma 4.5.
Proof.
Since satisfies , then for all and all we have that
Hence, the bounded Lipschitz distance is uniformly bounded:
where the supremum is taken over all -Lipschitz functions .
From the equi-integrability of it follows that satisfies the refined version of Ascoli-Arzelá theorem ([3, Theorem 3.3.1]) and there exist a (not relabelled) subsequence and a limit curve , such that the asserted convergence holds. ∎
In the next lemma, we provide the uniform bound on the BV-norm for the reconstructed densities . As a preparation, we state the following property of non-degenerate tessellations [22, Lemma 2.12(ii)].
Proposition 4.6.
Let satisfy the non-degeneracy assumption, and , be arbitrary with . The cells and can be connected by a path with , , , and , and , where depends only on .
Lemma 4.7.
Let with . Then satisfies
Proof.
For a fixed we consider any such that , then
Note that
Since , we have that for any . Therefore, we can find a unique cell such that . The line segment between the points and defines a path between cell and cell , consisting of pairs such that . We denote this sequence of pairs by . We further define the sets
Applying the triangle inequality, we have
where the last inequality follows from the geometric argument that if and only if the point is in the cylinder with base and axis parallel to .
Applying the lower bound from (B) and then the Hölder inequality, we then obtain
Therefore,
Taking the limit superior as , and applying the dominated convergence theorem, we obtain
Finally, we take the supremum over satisfying and use the variational characterization of the BV-seminorm to obtain
The bound on the -norm follows directly from assumption (B). ∎
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 be the GGF-solutions of (fKh) with . Then there exists and a (not relabelled) subsequence such that
Proof.
We first notice that the BV bound from Lemma 4.7 holds for almost every . Therefore, is tight with respect to the BV-norm in the sense that
Moreover, Lemma 4.5 provides weak integral equicontinuity, i.e.
Together, the tightness condition and the weak integral equicontinuity yield the relative compactness of in [40, Theorem 2]. The relative compactness in combined with the uniform integrability estimate
provides that is relatively compact in [40, Proposition 1.10]. Therefore, there exists some and a subsequence of (not relabelled) such that in for almost every . ∎
5. Gamma-convergence results
This section contains convergence results for the Fisher information and the dual dissipation potential . Since and 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 -convergence of the Fisher information.
5.1. General Gamma-convergence results
The notation throughout the first part of this section is as follows:
-
(i)
Let be the family of all open subsets of with Lipschitz boundary. We denote by the restriction of to , i.e. . Furthermore, we introduce the set , which can be larger then (see Figure 5). In what follows, we will use the convergence of the domain to in the following sense:
Proposition 5.1.
For any the indicator functions converge pointwise -a.e. to .

-
(ii)
is a family of probability measures on such that for all , where we use the reconstruction procedure defined in (4.1), i.e.
and there exists with Lebesgue density such that
-
(iii)
For each , the measure plays the role of the reference measure for the functional
(5.1) -
(iv)
We introduce a localized version of the functional :
(5.2) where the summation goes over the restriction of to , i.e.
-
(v)
Eventually, we will prove -convergence with respect to the -topology. Therefore, we embed the discrete functional into the full space as:
(5.3)
Remark 5.2.
This generic notation relates to the Fisher information in the following way:
The relation of with is more subtle. We show in Lemma 5.14 below that for a smooth and a specific choice of approximating sequence it holds that
With the notation at hand, we outline the steps to prove -convergence for by means of the localization technique:
-
(i)
The family of functionals has a subsequential -limit for all (cf. Lemma 5.4).
-
(ii)
The functionals and, in particular, have an integral representation:
For this, we need to prove that satisfies several properties as a set function, namely that is a measure and is local (cf. Proposition 5.10).
- (iii)
Definitions and compactness
We define
(5.4) |
where, by the usual definition,
Dealing with the functionals on the product space has a few subtleties due to the set dependence. We proceed in accordance with the theory presented in [14, Chapters 16-20]. Since are increasing functionals, i.e. for , we can apply the next definition
Definition 5.3.
We say that converges to (in ) if is the inner regular envelope of both functionals and .
The compactness result is standard.
Lemma 5.4.
The family of functionals defined in (5.3) is sequentially -compact, i.e. there exists a functional such that - for some subsequence.
Proof.
The compactness theorem for localized functionals is similar to the standard compactness theorem for the -convergence (see [14, Proposition 16.9]). ∎
Remark 5.5.
If one knows a priori that is inner regular, then the -limit is equivalently characterized by the usual -limits for all ([14, Proposition 16.4, Remark 16.5]):
-
()
for every , for every , and for every sequence in it holds that
-
()
for every and for every , there exists a sequence in such that
Integral representation
Since the subsequential -limit exists, it is equal to the inner regular envelope of both functionals and . Therefore, it suffices to show that is inner regular to conclude that . We will establish inner regularity together with other properties of as a set function in Propositon 5.9. All these properties, as well as possible integral representation, rely on growth conditions for .
To prove the growth conditions with respect to integrating against a possibly unbounded measure , we fix a suitable definition of :
Definition 5.6.
We define to be the completion of w.r.t. the norm
A useful observation is the convergence of the discrete approximations to .
Lemma 5.7.
Let , then in .
Proof.
By density arguments, it suffices to consider .
The fact that follows directly from the boundedness of , and the convergence follows from the following inequality:
Passing to the limit yields the statement. ∎
Now we establish the Sobolev upper bound for .
Lemma 5.8.
Proof.
For any and , we set . Then
where is a coupling between and . Since is smooth and and are in neighboring cells, it holds that , therefore,
Applying (UB) yields
The second term vanishes in the limit since . For the first term, we also notice that pointwise -a.e. (see Proposition 5.1) and is bounded on . Thus, we can apply the generalized dominated convergence theorem [17, Theorem 1.20] to obtain
Altogether, we obtain the following bound for any :
For arbitrary , we consider a sequence such that in , then the lower semicontinuity of yields
thereby concluding the proof. ∎
The properties of 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 ).
The functional defined in (5.2) has the following properties:
-
(i)
Inner regularity: For any and for any it holds that
-
(ii)
Subadditivity: For any and for any such that and it holds that:
-
(iii)
Locality: For any and any such that -a.e. on there holds
Proposition 5.10 (Properties of the -limit).
Let be -limit of . For every and every the following properties hold:
-
(i)
;
-
(ii)
for every ;
-
(iii)
is the restriction to of a Radon measure;
-
(iv)
is -lower semicontinuous;
-
(v)
is local, which means if -a.e. on ;
-
(vi)
satisfies the growth condition:
with some .
-
(vii)
has the integral representation
where .
Proof.
(i) Since is inner regular, we conclude by definition that .
(ii) It is easily seen that the equality holds for any and arbitrary , and .
To conclude (iii) it is enough to show that is subadditive, superadditive, and inner regular on (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 , for any such that and there exists small enough such that for any :
If , 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.
Upper bound for the integral representation
We derive an upper bound for using the representation (5.5). For a fixed the projection of on is
Substituting into yields
where we denoted by the tensor
(5.6) |
When passing , we expect converge to the diffusion tensor, therefore, we establish a number of useful properties of .
Lemma 5.12 (Properties of ).
The diffusion tensor (5.6) has the following properties:
-
(i)
is symmetric and positive-definite for any ;
-
(ii)
is bounded in :
for all the components it holds that ; -
(iii)
has a weakly- limit in the topology, i.e. there exist a subsequence and a tensor such that
Proof.
(i) Symmetry and positive-definiteness follow directly from the definition.
(iii) The weak- convergence follows from (ii) and the duality of and (see for instance [2, Theorem 8.5, Examples 8.6(1)]). ∎
The next lemma provides an upper bound for the integral representation of
Lemma 5.13.
5.2. Dual dissipation potential
Recall that the dual dissipation potential has the form:
with .
In Lemma 5.13, we derived an upper bound for the integral representation of the -limit of . We now prove that the same bound applies asymptotically to for some specific choice of approximating functions .
Lemma 5.14.
Let and assume that is a family of probability measures on such that in (cf. Section 5.1(ii)). Moreover, let be the family of discrete functions on defined by for .
Then in , and
Proof.
We first observe that in . This follows directly from estimate
Now, we show that realises the upper bound for the -limit proven in Lemma 5.13, i.e.
Since is smooth, the discrete gradient for can be approximated by
Moreover, the difference between and is of a small order:
which implies that
Substituting into yields:
where we used (UB). Applying Jensen’s inequality we get:
By Lemma 5.12(iii) and Proposition 5.1 one can pass on the right-hand side to obtain:
The last step we need to take is to show that . We consider the expansion of :
where
Since and recalling (UB) once again, we conclude the proof. ∎
5.3. Fisher information
In this section, we prove the -convergence for the family of discrete Fisher information defined as
(5.7) |
We state the main result of this section.
Theorem 5.15.
Up to passing to a subsequence, the family of functionals has a -limit w.r.t. the -topology taking the form
(5.8) |
where defined in Lemma 5.12.
Remark 5.16.
Since we assume that the densities are uniformly bounded from above and away from 0, and in , then is bounded in the same way. Consequently, the norms in and are equivalent.
In Theorem 5.15 we implicitly consider to depend on for all and take the -limit in the corresponding topology. To simplify the notation we set and consider again the localized functional:
Notice that . In what follows we set -, -, and -.
Proposition 5.10 provides the existence of an integral representation:
Unlike in Section 5.2, we are interested in the exact -limit for . 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 and it holds that:
Proof.
Let be a sequence with in such that
In particular, for all .
To prove the lower bound, one can repeat some elements of the proof of Lemma 4.7. We fix an arbitrary and denote . Let be such that , then
where we used Proposition 4.6 to assert that for some constant , independent of . Now one can repeat the transformations from the proof Lemma 4.7 to obtain:
Passing to the limit superior as then yields
For , passing yields
Since is inner regular (Proposition 5.9), then
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 be a functional satisfying:
-
(i)
is the restriction to of a Radon measure;
-
(ii)
is lower semicontinuous;
-
(iii)
is local, which means if -a.e. on ;
-
(iv)
for every ;
-
(v)
satisfies the growth condition:
for some .
Then for every and
where
for all , , , and where, for ,
Applying Proposition 5.18 to our setting, we notice that does not depend explicitly on (since for every Proposition 5.10). This gives
(5.9) |
The identification formula (5.9) suggests looking for the minimizer of 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
We also make use of the following definition of M
that was proven to be equivalent to the one from Proposition 5.18 in [21, Remark 7.4].
Note that the -convergence of to does not suffice to conclude the convergence of to . Hence, we define
and the corresponding discrete counterpart of is
Similarly to [21, Lemma 7.9], the next proposition claims that for any with Lipschitz boundary. For completeness, we include the proof in Appendix A.
Proposition 5.19.
Let and . For any sequence , there exists a subsequence that -converges in the -topology to .
Now we comment on the assumptions on tessellations and kernels needed for proving the representation result. It suffices to use (B) and (A). The last assumption (A) was not necessary for any preceding statements and its role here is to ensure that the discrete functions
are minimizers for , where . To relax (A) we can introduce an asymptotic assumption involving almost minimizers. Namely, we assume that
(AMin) |
Finally, we state the representation result.
Theorem 5.20.
Proof.
Proposition 5.19 and the theorem fundamental on convergence of minimizers (see, for instance, [9, Theorem 1.21]) together with (AMin) provides:
Substituting into yields
with
Since defined in Lemma 5.12 is piecewise constant on the tessellation, we can rewrite as
Using that in and Lemma 5.12(iii), we then obtain
Therefore,
Finally, we substitute M into the expression (5.9) for , and obtain for almost every :
thereby concluding the proof. ∎
Corollary 5.21.
Proof.
Computing the first variation for gives
where satisfies the boundary condition on .
Substituting and then using the symmetry, we have
Notice that the boundary condition implies that the summation goes over cells strictly contained within . This means that the inside sum goes over all the neighbors of the cell . This allows us to apply assumption (A) to obtain
and conclude that for all with on . Since is convex, this implies that is the minimizer. ∎
Lemma 5.22.
The diffusion tensor is uniformly elliptic and uniformly bounded:
with some .
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 inequalities for the functionals , and , therewith establishing the inequality for the energy-dissipation functional (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 . Finally, we conclude with the proof of Theorem A.
6.1. Lim inf inequalities
Given a density-flux pair we define
(6.1) |
where is defined in the way that (we constructed explicitly in Lemma 4.1).
Definition 6.1 (Density-flux convergence).
A discrete density-flux pair is said to converge to a density-flux pair if the pair of reconstructions defined as in (6.1) converges in the following sense
-
(1)
in for almost every ,
-
(2)
in .
We now summarize the lower bounds for all components of the energy-dissipation functional defined in Section 3.1. The form of the lower bounds is already suggested by Lemma 5.14 for the dissipation potential and by Theorem 5.15 for the Fisher information . Let us first give the definitions of , , , and and then summarize the corresponding inequalities in Theorem 6.2.
The dual dissipation potential takes the form
The dissipation potential is
The Fisher information is defined as
The energy functional is given by .
Theorem 6.2.
Let converge to in the sense of Definition 6.1. Then the following lower bounds hold for
-
(i)
the dissipation potential:
-
(ii)
the Fisher information:
-
(iii)
the energy functional:
Proof.
(i) The dissipation potential. We employ the dual formulation of . Let and be arbitrary. Then from the weak∗-convergence of and Lemma 5.14 we obtain
where for all . For the first term on the right-hand side, we have
owing to the regularity of and Lemma 4.4, and therefore
Consequently, we obtain
To conclude, we will make use of Legendre–Fenchel’s duality. In what follows, we set , , and as the closure in of the subspace , where denotes the square root of the positive definite matrix . Writing the term on the left in the previous inequality as
the Fenchel–Moreau duality theorem then gives
where the last equality follows from the fact that for -almost every .
(ii) The Fisher information. Since strongly in for almost every , we have by Theorem 5.15 that
Applying Fatou’s lemma then yields
(iii) The energy functional. As the following calculations hold for any , we drop the subscript . We recall that
where . Since and are probability measures,
Our piecewise constant reconstruction provides that
The narrow convergence of and in , along with the joint lower semicontinuity of the relative entropy [3, Lemma 9.4.3] then gives
as required. ∎
6.2. Chain rule
In this section we aim to establish the chain rule inequality:
from which we establish the nonnegativity of the limit energy-dissipation functional, i.e.
We will show that this inequality can be obtained from the chain rule for the relative entropy along -absolutely continuous curves.
We begin by rewriting in a more convenient form for the purpose of this section. We denote . If measures and have Lebesgue densities, then it holds that
which can be justified by monotone convergence. Recall that by the assumptions on (Section 2.2) , and, therefore, is finite whenever is finite.
Now we extend the definition of the energy for all measures with Lebesgue densities. First, we define an extended potential by
Since , is a lower semicontinuous on . Then for any we consider the extended energy functional defined by
We remark that the functionals and coincide on their sublevel sets. We also mention that if [3].
The following lemma results from a minor modification of [3, Theorem 10.4.13]. In our case, is not -convex, but the result remains true due to the regularity assumed on , i.e. .
Lemma 6.3.
A measure belongs to if and only if and
(6.2) |
In this case, is the minimal selection in .
Theorem 6.4 (Chain rule).
Let be such that
Then the map is absolutely continuous, and
In particular, this implies
Proof.
From the continuity equation and finiteness of , we deduce from [3, Theorem 8.3.1] that is -absolutely continuous (cf. Remark 3.6).
Furthermore, it is not difficult to show that the extended functional 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
where 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 is -negligible.
Due to the -convexity of w.r.t. the -metric [3] (see also [33]), we have that
On the other hand, the Lipschitz continuity of gives
Altogether, we find that
In particular, the map is differentiable for almost every .
We now show that and that (6.2) holds. Notice that if , then
thus implying that and . Here we used the fact that
We now define
Then, with a similar computation as above, we obtain
Hence, Lemma 6.3 implies that and is a minimal selection.
We then conclude that is -negligible and for all almost every :
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.
Appendix A Properties of Gamma-limits as set functions
Proposition 5.9.
The functional defined in (5.1) has the following properties:
-
(i)
Inner regularity:For any and for any it holds that
-
(ii)
Subadditivity: For any and for any such that and it holds that:
-
(iii)
Locality: For any and any such that -a.e. on there holds
Proof.
(i) Inner regularity. It is enough to prove that
because the opposite inequality holds since is an increasing set function. Note that is finite for any (Lemma 5.8).
First we need to choose two sequences and which converge to in . To define the first sequence we fix some and choose such that
Then by definition of there exists a sequence such that
where the upper bound was shown in Lemma 5.8. To define the second sequence let be such that . Again by definition, we find a sequence such that in and
Notice that both and are finite for sufficiently small, which necessarily implies that and are piecewise constant functions on . Everywhere in this proof, we use notation with ”hats” and superscript (for example, ) for functions in and superscript (for example, ) for the corresponding discrete functions on .
Next, we construct a sequence that ”interpolates” between and . Set and define sets for . Note that the following inclusions hold . Denote by a cut-off function between and , i.e. , on , and in a neighborhood of . with . It has a piecewise constant approximation , which satisfies on and on for . Now define
Observe that since converges pointwisely uniformly to as , the sequence still converges to in as for any .
For sufficiently small, the following holds:
where
We are now left to estimate the last term in the previous inequality.
We begin by bounding the discrete gradient of by
By Lemma 4.1 and since , then , therefore
where we used the upper bound assumption (UB). On the other hand, for any , we can choose such that
In particular, we can choose depending on such that as .
Making use of these estimates gives
with . Choosing such that
we then obtain
Combining the estimates together, we have
Taking the limit superior gives
By sending and , we eventually conclude
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 and converging to in such that
Set and define sets for . Note that the following inclusions hold . Let be a cut-off function between and with . We use the piecewise constant approximation to define the sequence:
For sufficiently small, it holds that
with as in (1). The last term on the right-hand side may be estimated as in (1) to obtain
Choosing such that
we then obtain
where the last term may be estimated by
Taking the limit superior as gives
By sending and applying the inner regularity property, we conclude
(iii) Locality. We first prove . The argument is similar to the previous points. For a fixed there exists such that . We choose two sequences and such that , in satisfying
Set and define sets for . Denote by a cut-off function between and with . It has a piecewise constant approximation . Then, we define
with still converging to in as for any . Similar to the proof of inner regularity, we obtain the existence of some such that
Passing to the limits and then yields
The assertion follows as we can swap the roles of and . ∎
We recall that
and the corresponding discrete counterpart of is
Proposition 5.19.
Let be arbitrary with Lipschitz boundary and . For any sequence there exists a subsequence that -converges in the -topology to .
Proof.
Let us first prove the - inequality. We consider a sequence in such that . This implies that on and . Consequently, we also have that . By the same argument as in Lemma 5.17, we deduce that . Since - it remains to prove that .
Notice that for all . Since on , their piecewise reconstructions satisfy on , and hence also on for all . Using the fact that and in , we easily deduce that in . The deduced regularity and assumed regularity then allows to conclude that .
Thus, the inequality follows:
Now we show the approximate inequality. Let such that . There exists a recovery sequence in such that .
Set and define sets for . Denote by a cut-off function between and with . It has a piecewise constant approximation . Then, we define
with still converging to in as for any . Similar to the proof of inner regularity, we obtain the existence of some such that
Passing yields
∎
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 and in . Archive for Rational Mechanics and Analysis, 165(3):187–242, 2002.
- [9] A. Braides. -Convergence for Beginners, volume 22. Clarendon Press, 2002.
- [10] A. Braides. A handbook of -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 -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 -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 -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. -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. -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.