Gradient flow formulation of diffusion equations in the Wasserstein space over a Metric graph

  • Received: 01 May 2021 Revised: 01 April 2022 Published: 16 June 2022
  • Primary: 35R02; Secondary: 49Q22, 60B05

  • This paper contains two contributions in the study of optimal transport on metric graphs. Firstly, we prove a Benamou–Brenier formula for the Wasserstein distance, which establishes the equivalence of static and dynamical optimal transport. Secondly, in the spirit of Jordan–Kinderlehrer–Otto, we show that McKean–Vlasov equations can be formulated as gradient flow of the free energy in the Wasserstein space of probability measures. The proofs of these results are based on careful regularisation arguments to circumvent some of the difficulties arising in metric graphs, namely, branching of geodesics and the failure of semi-convexity of entropy functionals in the Wasserstein space.

    Citation: Matthias Erbar, Dominik Forkert, Jan Maas, Delio Mugnolo. Gradient flow formulation of diffusion equations in the Wasserstein space over a Metric graph[J]. Networks and Heterogeneous Media, 2022, 17(5): 687-717. doi: 10.3934/nhm.2022023

    Related Papers:

    [1] Karoline Disser, Matthias Liero . On gradient structures for Markov chains and the passage to Wasserstein gradient flows. Networks and Heterogeneous Media, 2015, 10(2): 233-253. doi: 10.3934/nhm.2015.10.233
    [2] Steinar Evje, Kenneth H. Karlsen . Hyperbolic-elliptic models for well-reservoir flow. Networks and Heterogeneous Media, 2006, 1(4): 639-673. doi: 10.3934/nhm.2006.1.639
    [3] Fabian Rüffler, Volker Mehrmann, Falk M. Hante . Optimal model switching for gas flow in pipe networks. Networks and Heterogeneous Media, 2018, 13(4): 641-661. doi: 10.3934/nhm.2018029
    [4] Joachim von Below, José A. Lubary . Stability implies constancy for fully autonomous reaction-diffusion-equations on finite metric graphs. Networks and Heterogeneous Media, 2018, 13(4): 691-717. doi: 10.3934/nhm.2018031
    [5] Andrea Corli, Lorenzo di Ruvo, Luisa Malaguti, Massimiliano D. Rosini . Traveling waves for degenerate diffusive equations on networks. Networks and Heterogeneous Media, 2017, 12(3): 339-370. doi: 10.3934/nhm.2017015
    [6] Pedro Aceves-Sanchez, Benjamin Aymard, Diane Peurichard, Pol Kennel, Anne Lorsignol, Franck Plouraboué, Louis Casteilla, Pierre Degond . A new model for the emergence of blood capillary networks. Networks and Heterogeneous Media, 2021, 16(1): 91-138. doi: 10.3934/nhm.2021001
    [7] Michael Helmers, Barbara Niethammer, Xiaofeng Ren . Evolution in off-critical diblock copolymer melts. Networks and Heterogeneous Media, 2008, 3(3): 615-632. doi: 10.3934/nhm.2008.3.615
    [8] Seung-Yeal Ha, Jeongho Kim, Jinyeong Park, Xiongtao Zhang . Uniform stability and mean-field limit for the augmented Kuramoto model. Networks and Heterogeneous Media, 2018, 13(2): 297-322. doi: 10.3934/nhm.2018013
    [9] Dimitra Antonopoulou, Georgia Karali . A nonlinear partial differential equation for the volume preserving mean curvature flow. Networks and Heterogeneous Media, 2013, 8(1): 9-22. doi: 10.3934/nhm.2013.8.9
    [10] Leandro M. Del Pezzo, Nicolás Frevenza, Julio D. Rossi . Convex and quasiconvex functions in metric graphs. Networks and Heterogeneous Media, 2021, 16(4): 591-607. doi: 10.3934/nhm.2021019
  • This paper contains two contributions in the study of optimal transport on metric graphs. Firstly, we prove a Benamou–Brenier formula for the Wasserstein distance, which establishes the equivalence of static and dynamical optimal transport. Secondly, in the spirit of Jordan–Kinderlehrer–Otto, we show that McKean–Vlasov equations can be formulated as gradient flow of the free energy in the Wasserstein space of probability measures. The proofs of these results are based on careful regularisation arguments to circumvent some of the difficulties arising in metric graphs, namely, branching of geodesics and the failure of semi-convexity of entropy functionals in the Wasserstein space.



    This article deals with the equivalence of static and dynamical optimal transport on metric graphs, and with a gradient flow formulation of McKean–Vlasov equations in the Wasserstein space of probability measures.

    Let (V,E) be a finite (undirected) connected graph and let :E(0,) be a given weight function. Loosely speaking, the associated metric graph is the geodesic metric space (G,d) obtained by identifying the edges eE with intervals of length e, and gluing the intervals together at the nodes. In other words, a metric graph describes a continuous "cable system", rather than a discrete set of nodes; see Section 2.2 for a more formal definition. Metric graphs arise in many applications in chemistry, physics, or engineering, describing quasi-one-dimensional systems such as carbon nano-structures, quantum wires, transport networks, or thin waveguides. They are also widely studied in mathematics; see [6,23] for an overview.

    For 1p<, the Lp-Wasserstein distance Wp between Borel probability measures μ and ν on G is defined by the Monge–Kantorovich transport problem

    Wp(μ,ν):=minσΠ(μ,ν){G×Gdp(x,y)dσ(x,y)}1/p,

    where Π(μ,ν) denotes the set of transport plans; i.e., all Borel probability measures σ on G×G with respective marginals μ and ν. Since G is compact, the Wasserstein distance Wp metrises the topology of weak convergence of Borel probability measures on G, for any p1. The resulting metric space of probability measures is the Lp-Wasserstein space over G.

    Metric graphs (G,d) are prototypical examples of metric spaces that exhibit branching of geodesics: it is (typically) possible to find two distinct constant speed geodesics (γt)t[0,1] and (˜γt)t[0,1], taking the same values for all times t[0,t0] up to some t0(0,1). For this reason, several key results from optimal transport are not directly applicable to metric graphs. This paper contains two of these results.

    The Benamou–Brenier formula. On Euclidean space Rd, a dynamical characterisation of the Wasserstein distance has been obtained in celebrated works of Benamou and Brenier [4,5]. The Benamou–Brenier formula asserts that

    W22(μ,ν)=inf(μt,vt){10vt2L2(μt)dt} (1)

    where the infimum runs over all 2-absolutely continuous curves (μt)t[0,1] in the L2-Wasserstein space over Rd, satisfying the continuity equation

    ddtμt+(vtμt)=0 (2)

    with boundary conditions μ0=μ and μ1=ν.

    Here we are interested in obtaining an analogous result in the setting of metric graphs. However, such an extension is not straightforward, since standard proofs in the Euclidean setting (see [2,26]) make use of the flow map T:[0,1]×RdRd associated to a (sufficiently regular) vector field (vt)t[0,1], which satisfies

    ddtT(t,x)=vt(T(t,x)),T(0,x)=x,T(t,)#μ0=μt,

    see, e.g., [2,Proposition 8.1.8]. On a metric graph such a flow map T typically fails to exist, since solutions (μt)t[0,1] to the continuity equation (2) are usually not uniquely determined by an initial condition μ0 and a given vector field (vt)t[0,1].

    Circumventing this difficulty, Gigli and Han obtained a version of the Benamou–Brenier formula in very general metric measure spaces [13]. However, this paper requires a strong assumption on the measures (namely, a uniform bound on the density with respect to the reference measure). While this assumption is natural in the general setting of [13], it is unnecessarily restrictive in the particular setting of metric graphs.

    In this paper, we prove a Benamou–Brenier formula that applies to arbitrary Borel probability measures on metric graphs. The key ingredient in the proof is a careful regularisation step for solutions to the continuity equation.

    Gradient flow structure of diffusion equations. As an application of the Benamou–Brenier formula, we prove another central result from optimal transport: the identification of diffusion equations as gradient flow of the free energy in the Wasserstein space (P(G),W2). In the Euclidean setting, results of this type go back to the seminal work of Jordan, Kinderlehrer, and Otto [14].

    Here we consider diffusion equations of the form

    tη=Δη+(η(V+W[μ])), (3)

    for suitable potentials V:GR and W:G×GR. In analogy with the Euclidean setting, we show that this equation arises as the gradient flow equation for a free energy F:P(G)(,+] composed as the sum of entropy, potential, and interaction energies:

    F(μ):=Gη(x)logη(x)dλ(x)+GV(x)η(x)dλ(x)+12G×GW(x,y)η(x)η(y)dλ(x)dλ(y),

    if μ=ηλ, where λ denotes the Lebesgue measure on G.

    A key difference compared to the Euclidean setting is that the entropy is not semi-convex along W2-geodesics; see Section 4. This prevents us from applying "off-the-shelf" results from the theory of metric measure spaces. Instead, we present a self-contained proof of the gradient flow result. Its main ingredient is a chain rule for the entropy along absolutely continuous curves, which we prove using a regularisation argument.

    Interestingly, we do not need to assume continuity of the potential V at the vertices. Therefore our setting includes diffusion with possibly singular drift at the vertices.

    The Wasserstein distance over metric graphs for p=1 has been studied in [20]. The focus is on the approximation of Kantorovich potentials using p-Laplacian type problems.

    The recent paper [10] deals with dynamical optimal transport metrics on metric graphs. The authors start with the dynamical definition à la Benamou–Brenier and consider links to several other interesting dynamical transport distances. The current paper is complementary, as it shows the equivalence of static and dynamical optimal transport, and a gradient flow formulation for diffusion equations.

    Various different research directions involving optimal transport and graphs exist. In particular, dynamical optimal transport on discrete graphs have been intensively studied in recent years following the papers [11,19,21]. The underlying state space in these papers is a discrete set of nodes rather than a gluing of one-dimensional intervals.

    Another line of research deals with branched optimal transport, which is used to model phenomena such as road systems, communication networks, river basins, and blood flow. Here one starts with atomic measures in the continuum, and graphs emerge to describe paths of optimal transport [28,7].

    Organisation of the paper. In Section 2 we collect preliminaries on optimal transport and metric graphs. Section 3 is devoted to the continuity equation and the Benamou–Brenier formula on metric graphs. In particular, we present a careful regularisation procedure for solutions to the continuity equation. Section 4 contains an example which demonstrates the lack convexity of the entropy along W2-geodesics in the setting of metric graphs. Section 5 deals with the gradient flow formulation of diffusion equations.

    In this section we collect some basic definitions and results from optimal transport and metric graphs.

    In this section we collect some basic facts on the family of Lp-Wasserstein distances on spaces of probability measures. We refer to [26,Chapter 5], [1,Chapter 2] or [27,Chapter 6] for more details.

    Let (X,d) be a compact metric space. The space of Borel probability measures on X is denoted by P(X). The pushforward measure T#μ induced by a Borel map T:XY between two Polish spaces is defined by (T#μ)(A):=μ(T1(A)) for all Borel sets AY.

    Definition 2.1 (Transport plans and maps).

    1. A (transport) plan between probability measures μ,νP(X) is a probability measure σP(X×X) with respective marginals μ and ν, i.e.,

    (proj1)#σ=μand(proj2)#σ=ν,

    where proji(x1,x2):=xi for i=1,2. The set of all transport plans between μ and ν is denoted by Π(μ,ν).

    2. A transport plan σΠ(μ,ν) is said to be induced by a Borel measurable transport map T:XX if σ=(Id,T)#μ, where (Id,T):XX×X denotes the mapping x(x,T(x)).

    Definition 2.2 (Kantorovich–Rubinstein–Wasserstein distance). For p1, the Lp-Kantorovich–Rubinstein–Wasserstein distance between probability measures μ,νP(X) is defined by

    Wp(μ,ν):=inf{(X×Xdp(x,y)dσ(x,y))1/p:σΠ(μ,ν)}. (4)

    For any μ,νP(X) the infimum above is attained by some σminΠ(μ,ν); any such σmin is called an optimal (transport) plan between μ and ν. If a transport map T induces an optimal transport plan, we call T optimal as well.

    By compactness of (X,d), the Lp-Wasserstein distance metrises the weak convergence in P(X) for any p1; see, e.g., [27,Corollary 6.13]. Moreover, (P(X),Wp) is a compact metric space as well.

    We conclude this section with a dual formula for the Wasserstein distance (see, e.g., [26,Section 1.6.2]). To this aim, we recall that for c:X×XR, the c-transform of a function φ:XR{+} is defined by φc(y):=infxXc(x,y)φ(x). A function ψ:XR{+} is called c-concave if there exists a function φ:XR{+} such that ψ=φc.

    Proposition 2.3 (Kantorovich duality). For any μ,νP(X) we have

    Wpp(μ,ν)=supφ,ψC(X){Xφdμ+Xψdν : φ(x)+ψ(y)dp(x,y)x,yX}.

    Moreover, the supremum is attained by a maximising pair of the form (φ,ψ)=(φ,φc), where φ is a c-concave function, for c(x,y):=dp(x,y).

    Any maximiser φ is called a Kantorovich potential.

    In this subsection we state some basic facts on metric graphs; see e.g., [20], [6] or [16] for more details.

    Definition 2.4 (Metric graph). Let G=(V,E,) be an oriented, weighted graph, which is finite and connected. We identify each edge e=(einit,eterm)E with an interval (0,e) and the corresponding nodes einit,etermV with the endpoints of the interval, (0 and e respectively). The spaces of open and closed metric edges over G are defined as the respective topological disjoint unions

    E:=eE(0,e)and¯E:=eE[0,e].

    The metric graph over G is the topological quotient space

    G:=¯E/,

    where points in ¯E corresponding to the same vertex are identified.

    Note that the orientation of e determines the parametrisation of the edges, but does not otherwise play a role. To distinguish ingoing and outgoing edges at a given node, we introduce the signed incidence matrix I=(ιev) whose entries are given by

    ιev:={+1if v=einit,1if v=eterm,0otherwise.

    As a quotient space, any metric graph naturally inherits the structure of a metric space from the Euclidean distance on its metric edges [9,Chapter 3]: indeed, under our standing assumption that G is connected the quotient semi-metric d becomes a metric.

    The distance d:G×G[0,) on G can be more explicitly described as follows: For x,yG, let ˜G=(˜V,˜E) be the underlying discrete graph obtained by adding new vertices at x and y, and let d(x,y) be the weighted graph distance between x and y in ˜G, i.e.,

    d(x,y):=minni=1ei,

    where the minimum is taken over all sequences of vertices x=x0,,xn=y in the extended graph ˜G such that xi1 and xi are vertices of an edge ei˜E for all i=1,,n. In particular, if x and y are vertices in G, no new nodes are added, and we recover the graph distance in the original graph. We refer to [23,Chapter 3] for details.

    By construction, the distance function d metrises the topology of G. It is readily checked that d is a geodesic distance, i.e., each pair of points x,yG can be joined by a curve of minimal length d(x,y). Consequently, also the Wasserstein space (P(G),W2) is a geodesic space.

    In a metric space (X,d), recall that the local Lipschitz constant of a function f:XR is defined by

    lip(f)(x):=lim supyx|f(y)f(x)|d(x,y),

    whenever xX is not isolated, and 0 otherwise. The (global) Lipschitz constant is defined by

    Lip(f):=supyx|f(y)f(x)|d(x,y).

    If the underlying space X is a geodesic space, we have Lip(f)=supxlip(f)(x).

    At the risk of being redundant, we explicitly introduce a few relevant function spaces, although they are actually already fully determined by the metric measure structure of the metric graph G.

    (ⅰ) C(G) denotes the space of continuous real-valued functions on G, endowed with the uniform norm .

    (ⅱ) Ck(¯E) is the space of all functions φ on ¯E such that the restriction to each closed edge has continuous derivatives up to order kN.

    (ⅲ) Lp(G), for p[1,], is the p-Lebesgue space over the measure space (G,λ), where λ denotes the image of the 1-dimensional Lebesgue measure on ¯E under the quotient map.

    (ⅳ) Likewise, we consider the Sobolev spaces W1,p(¯E) of Lp(G)-functions whose restriction on each edge is weakly differentiable with weak derivatives in Lp(G).

    In this section we fix a metric graph G and perform a study of the continuity equation

    tμt+Jt=0 (5)

    in this context.

    In this work we mainly deal with weak solutions to the continuity equation, which will be introduced in Definition 3.2. To motivate this definition, we first introduce the following notion of strong solution.

    Definition 3.1 (Strong solutions to the continuity equation). A pair of measurable functions (ρ,U) with ρ:(0,T)×GR+ and U:(0,T)ׯER is said to be a strong solution to (5) if

    (i) tρ(t,x) is continuously differentiable for every xG;

    (ii) xUt(x) belongs to C1(¯E) for every t(0,T);

    (iii) the continuity equation ddtρt(x)+Ut(x)=0 holds for every t(0,T) and xE;

    (iv) for every t(0,T) and wV we have eEwιewUt(we)=0.

    Here, we write ρt:=ρ(t,) and Ut:=U(t,) and denote by the spatial derivative. Moreover, Ew denotes the set of all edges adjacent to the node wV, and we¯E denotes the corresponding endpoint of the metric edge e which corresponds to wG.

    To motivate the definition of a weak solution, suppose that we have a strong solution (ρt,Ut)t(0,T) to the continuity equation (5). Let ψC1(¯E)C(G) be a test function. Integration by parts on every metric edge e gives

    ddte0ψρtdx=e0ψUtdx+ψUt|e0,

    and summation over eE yields

    ddtGψρtdx=¯EψUtdx+wVψ(w)eEwιewUt(we)=¯EψUtdx,

    where we use the continuity of ψ on G as well as the node condition (ⅳ) above in the last step. This ensures that the net ingoing momentum vanishes at every node in V. In particular, choosing ψ1 yields

    Gρsdλ=Gρtdλ,

    for all s,t(0,T), i.e. solutions to the continuity equation are mass-preserving. Here condition (ⅳ) is crucial to ensure that no creation or annihilation of mass occurs at the nodes.

    Definition 3.2 (Weak solution). A pair (μt,Jt)t(0,T) consisting of probability measures μt on G and signed measures Jt on ¯E, such that t(μt(A),Jt(A)) is measurable for all Borel sets A, is said to be a weak solution to (5) if

    (i) tGψdμt is absolutely continuous for every ψC1(¯E)C(G);

    (ii) T0|Jt|(¯E)dt<;

    (iii) for every ψC1(¯E)C(G) and a.e. t(0,T), we have

    ddtGψdμt=¯EψdJt. (6)

    Remark 1. Proposition 3.9 below shows that continuous functions on G can be uniformly approximated by C1 functions. By standard arguments it then follows that (μt,Jt)t(0,T) is a weak solution if and only if the following conditions hold: (i) tμt is weakly continuous; (ii) from Definition 3.2 holds; and

    (iii) for every φC1c((0,T)ׯE) such that φt,tφtC1(¯E)C(G) for all t(0,T) we have

    T0(Gtφdμt+¯EφdJt)dt=0. (7)

    See Lemma 3.10 below for the weak continuity in (i).

    The next result asserts that the momentum field does not give mass to vertices for a.e. time point. Hence, we can equivalently restrict the integrals over ¯E in (6) and (7) to the space of open edges E.

    Lemma 3.3. Let B:=¯EE denote the set of all boundary points of edges. For any weak solution to the continuity equation (μt,Jt)t(0,T), we have

    T0|Jt|(B)dt=0.

    Proof. Fix a metric edge e in ¯E and take w{einit,eterm}. Without loss of generality we take w=einit. Then we can construct a family of functions φε for ε(0,e) with the following properties:

    (a) φεC1(¯E)C(G);

    (b) φε0 on ¯Ee and φε0 uniformly on G as ε0;

    (c) |φε|=1 on (0,ε)e and |φε|0 uniformly on compact subsets of ¯E{w}.

    For instance, we could set φε(x):=1e(x)ηε(x), where ηε:RR is a C1 approximation of the function x(xε(ex)eε)0. Choosing φ(t,x)=φε(x) in (7), we obtain by passing to the limit ε0 that T0|Jt|({w})dt=0.

    Lemma 3.4 (Weak and strong solutions). The following assertions hold:

    (i) If (ρt,Ut)t(0,T) is a strong solution to the continuity equation, then the pair (μt,Jt)t(0,T) defined by μt=ρtλ and Jt=Utλ is a weak solution to the continuity equation.

    (ii) If (μt,Jt)t(0,T) is a weak solution to the continuity equation (6) such that the densities ρt:=dμtdλ and Ut:=dJtdλ exist for all times t(0,T) and satisfy the regularity conditions (ⅰ) and (ⅱ) of Definition 3.1, then (ρt,Ut)t(0,T) is a strong solution to the continuity equation.

    Proof. Both claims are straightforward consequences of integration by parts on each metric edge in E.

    Let (X,d) be a metric space and let T>0.

    Definition 3.5. For p1, we say that a curve γ:(0,T)X is p-absolutely continuous if there exists a function gLp(0,T) such that

    d(γs,γt)tsgrdrs,t(0,T):st. (8)

    The class of p-absolutely continuous curves is denoted by ACp((a,b);(X,d)). For p=1 we simply drop p in the notation. The notion of locally p-absolutely continuous curve is defined analogously.

    Proposition 3.6. Let p1.For every p-absolutely continuous curve γ:(0,T)X, the metric derivative defined by

    |˙γ|(t):=limstd(γs,γt)|st|

    exists for a.e. t(0,T) and t|˙γ|(t) belongs to Lp(0,T).The metric derivative |˙γ|(t) is an admissible integrand in the right-hand side of (8). Moreover, any other admissible integrand gLp(0,T) satisfies|˙γ|(t)g(t) for a.e. t(0,T).

    Proof. See, e.g., [2,Theorem 1.1.2].

    The next result relates the metric derivative of tμt to the L2(μt)-norm of the corresponding vector fields vt.

    Theorem 3.7 (Absolutely continuity curves). The following statements hold:

    (i) If (μt)t(0,T) is absolutely continuous in (P(G),W2), then there exists, for a.e. t(0,T), a vector field vtL2(μt) such that vtL2(μt)|˙μ|(t) and (μt,vtμt)t(0,T) is a weak solution to the continuity equation (6).

    (ii) Conversely, if (μt,vtμt)t(0,T) is a weak solution to the continuity equation (6) satisfying 10vtL2(μt)dt<+, then (μt)t(0,T) is an absolutely continuous curve in (P(G),W2) and |˙μ|(t)vtL2(μt) for a.e. t(0,T).

    Proof of (ⅰ). We adapt the proof of [2,Theorem 8.3.1] to the setting of metric graphs.

    The idea of the proof is as follows: On the space-time domain Q:=(0,T)×G we consider the Borel measure μ:=T0δtμtdt whose disintegration with respect to the Lebesgue measure on (0,T) is given by (μt)t(0,T). To deal with the fact that gradients of smooth functions are multi-valued at the nodes, we define ¯μtM+(¯E) by ¯μt(A):=eEμt(A¯e) for every Borel set A¯E. We then set ¯Q:=(0,T)ׯE and define ¯μM+(¯Q) by ¯μ:=T0δt¯μtdt. (Note that mass at the nodes is counted multiple times). Consider the linear spaces of functions T and V given by

    T:=span{(0,T)×G(t,x)a(t)φ(x) : aC1c(0,T),φC1(¯E)C(G)},V:={(0,T)ׯE(t,x)xΦ(t,x) : ΦT}.

    The strategy is to show that the linear functional L:VR given by

    L(aφ):=Q˙a(t)φ(x)dμ(x,t),

    is well-defined and L2(¯Q,¯μ)-bounded with L2T0|˙μ|2(t)dt. Once this is proved, the Riesz Representation Theorem yields the existence of a vector field vv in ¯VL2(¯Q,¯μ) such that vv2L2(¯μ)T0|˙μ|2(t)dt and

    T0˙a(t)Gφ(x)dμt(x)dt=L(aφ)=T0a(t)¯Eφ(x)vt(x)d¯μt(x)dt (9)

    for vt:=vv(t,) and all aC1c(0,T) and φC1(¯E)C(G).

    Once this is done, we show that the momentum vector field JJ:=vvμ does not assign mass to boundary points in ¯E, so that vv can be interpreted as an element in L2(Q,μ) and the integration over vector fields can be restricted to E.

    Step 1. Fix a test function φC1(¯E)C(G) and consider the bounded and upper semicontinuous function H:G×GR given by

    H(x,y):={lip(φ)(x)if x=y,|φ(x)φ(y)|d(x,y)if xy,

    for x,yG. For s,t(0,T), let σstΠ(μs,μt) be an optimal plan. The Cauchy–Schwarz inequality yields

    |GφdμsGφdμt|G×Gd(x,y)H(x,y)dσst(x,y)W2(μs,μt)(G×GH2(x,y)dσst(x,y))1/2. (10)

    As φ is globally Lipschitz on G, we obtain

    |GφdμsGφdμt|Lip(φ)W2(μs,μt)

    and infer that the mapping tGφdμt is absolutely continuous, hence, differentiable outside of a null set Nφ(0,T).

    Fix t(0,T) and take a sequence {sn}nN converging to t. Since {μsn} is weakly convergent, this sequence is tight. Consequently, {σsnt}nN is tight as well, and we may extract a subsequence converging weakly to some ˆσP(G×G). It readily follows that ˆσΠ(μt,μt). Moreover, along the convergent subsequence, we have

    G×Gd2(x,y)dˆσ(x,y)lim infnG×Gd2(x,y)dσsnt(x,y)=lim infnW22(μsn,μt)=0,

    which implies that ˆσ=(Id,Id)#μt.

    Using this result and the upper-semicontinuity of H, it follows from (10) that

    lim supst|GφdμsGφdμtst||˙μ|(t)lim supst(G×GH2(x,y)dσst(x,y))1/2|˙μ|(t)lip(φ)L2(μt). (11)

    Step 2. Take ΦT. Using dominated convergence, Fatou's Lemma, and (11), we obtain

    |QddtΦ(x,t)dμ(x,t)|=limh0|1hQΦ(x,th)Φ(x,t)dμ(x,t)|=limh0|1hT0(GΦ(x,t)dμt+h(x)GΦ(x,t)dμt(x))dt|T0|˙μ|(t)lipx(Φ)(,t)L2(μt)dt(T0|˙μ|2(t)dt)1/2(Q|lipx(Φ)(x,t)|2dμ(x,t))1/2. (12)

    Since Q|lipx(Φ)(x,t)|2dμ(x,t)¯Q|Φ(x,t)|2d¯μ(x,t), we infer that L is well-defined and extends to a bounded linear functional on the closure of V in L2(¯Q,¯μ) with L2T0|˙μ|2(t)dt, which allows us to apply the Riesz Representation Theorem, as announced above.

    In particular, (9) implies that tEφvtdμt is a distributional derivative for tGφdμt. Since the latter function is absolutely continuous and, therefore, belongs to the Sobolev space W1,1(0,T), we obtain

    ddtGφdμt=¯Eφvtdμtfor a.e. t(0,T). (13)

    We conclude that (μt,vt)t(0,T) solves the continuity equation in the weak sense. Lemma 3.3 implies that for a.e. t the momentum field Jt:=vt¯μt does not give mass to any boundary point in ¯E. Consequently, the spatial domain of integration on the right-hand side of (9) may be restricted to E.

    Step 3. It remains to verify (by a standard argument) the inequality relating the L2(μt)-norm of the vector field vt to the metric derivative of μt.

    For this purpose, fix a sequence (ϖϖi)iN of functions ϖϖiV converging to vv in L2(¯μ) as i. For every compact interval I(0,T) and aC1(0,T) satisfying 0a1 and suppa=I, we then obtain

    Qa(t)|vv(x,t)|2dμ(x,t)=limiQa(t)ϖϖi(x,t)vv(x,t)dμ(x,t)=limiL(aϖϖi)(T01I|˙μt|2dt)1/2limi(Q1I|ϖϖi|2dμ)1/2=(T01I|˙μ|2(t)dt)1/2(Q1I|v|2dμ)1/2.

    Letting a1I0, this inequality implies

    IE|vt|2dμtdtI|˙μ|2(t)dt.

    Since I(0,T) is arbitrary, this implies that vtL2(μt)|˙μ|(t) for a.e. t(0,T).

    Next we introduce a suitable spatial regularisation procedure for solutions to the continuity equation. This will be crucial in the proof of the second part of Theorem 3.7.

    Let ε>0 be sufficiently small, i.e., such that 2ε is strictly smaller than the length of every edge in E. We then consider the supergraph GextG defined by adjoining an auxiliary edge eextv of length 2ε to each node vV (see Figure 1). The corresponding set of metric edges will be denoted by EextE.

    Figure 1. 

    The supergraph Gext is constructed by adjoining an additional leaf at every node in V

    .

    We next define a regularisation procedure for functions based on averaging. The crucial feature here is that non-centred averages are used, to ensure that the regularised function is continuous.

    In the definition below, we parametrise each edge e=(einit,eterm)E using the interval (e2,e2) instead of (0,e). The auxiliary edges eexteinit and eexteterm will then be identified with the intervals (e22ε,e2) and (e2,e2+2ε), respectively. We stress that for each vertex v, there is only one additional edge, but we use several different parametrisations for it – one for each edge incident in v.

    Definition 3.8 (Regularisation of functions). For φL1(Gext), we define φε:GR by

    φε(x):=12εαεex+εαεexεφ(y)dyfor xe=[e2,e2], (14)

    where αεe:=(e+2ε)/e. We write αε(x):=αεe, whenever xE is a point on the metric edge e.

    Note that the value of φε in each of the nodes depends only on data on the corresponding auxiliary edge. In particular, the value at the nodes does not depend on the choice of the edge, so that φε indeed defines a function on G. We collect some basic properties of this regularisation in the following result.

    Proposition 3.9. The following properties hold for every ε>0 sufficiently small:

    (i) Regularising effect: For any φC(Gext) we have φεC1(¯E)C(G) and

    φε(y)=αεe2ε(φ(αεey+ε)φ(αεeyε)) (15)

    for y[e/2,e/2].

    (ii) If φ belongs to C(Gext), then φε converges uniformly to φ|G as ε0.

    Proof. (ⅰ) follows by direct computation; (ⅱ) follows using the uniform continuity of φ on Gext.

    Lemma 3.10 (Weak continuity). Let (ρt,Jt)t(0,T) be a weak solution to the continuity equation on G.Then tGφdμt is continuous for every φC(G).

    Proof. Fix φC(G), take a continuous extension to Gext, and define φε accordingly. Proposition 3.9.ii then implies that φε converges uniformly to φ on G as ε0. As a result, the function tGφεdμt converges uniformly to tGφdμt. Since φε belongs to C(G)C1(¯E) by Proposition 3.9.i, we conclude that the mapping tGφdμt is continuous, being a uniform limit of continuous functions.

    By duality, we obtain a natural regularisation for measures.

    Definition 3.11 (Regularisation of measures). For μM(G) we define μεM(Gext) by

    Gextφdμε:=Gφεdμ. (16)

    for all φC(Gext).

    Analogously, for JM(E) we define JεM(Eext) as follows: first we extend J to a measure on G giving no mass to GE. Then we define ˜JεM(Gext) by the formula above. Finally, we define JεM(Eext) by restriction of ˜Jε to Eext.

    It is readily checked that the right-hand side defines a positive linear functional on C(Gext), so that με is indeed a well-defined measure.

    Proposition 3.12. The following properties hold for any ε>0:

    (i) Mass preservation: με(Gext)=μ(G) for any μM(G).

    (ii) Regularising effect:For any μP(G), the measure με is absolutely continuous with respect to λ with density

    ρε(x)={12εμ(eIe(x)),for x on e in E12ε(1{d(x,w)2ε}μ({w})+eE:weμ(eIe(x))),for x on eextwwV,

    where

    Ie(x):=(xεaεe(e2),x+εaεee2).

    In particular, ρε(x)12ε for all xGext.

    (iii) Kinetic energy bound: For μP(G) and vL2(μ), define J=vμ|EM(E).Consider the regularised measures μεP(Gext)and JεM(Eext).Then we have Jε=vεμε for some vεL2(με) and

    Eext|vε|2dμεE|v|2dμ. (17)

    (iv) For any μP(G) we have weak convergence μεμ in P(Gext) as ε0.

    (v) Let (μt,Jt)t(0,T) be a weak solution to the to the continuity equation (6). Then the regularised pair (μεt,Jεt)t(0,T) is a weak solution to a modified continuity equation on Gext in the following sense:

    For every absolutely continuous function φ on Gext, the function tGextφdμεt is absolutely continuous and for a.e. t(0,T) we have

    ddtGextφdμεt=EextαεφdJεt, (18)

    with αε as in Definition 3.8.

    In order to prove (ⅲ), we will make use of the so-called Benamou–Brenier functional (see, e.g., [26,Section 5.3.1] for corresponding results in the Euclidean setting).

    Define K2:={(a,b)R×R:a+b2/20}. By a slight abuse of notation, C(G,K2) (resp. L(G,K2)) denotes the set of all continuous (resp. bounded and measurable) functions a,b:GR such that a+b2/20.

    Definition 3.13. The Benamou–Brenier functional B2:M(G)×M(E)R{+} is defined by

    B2(μ,J):=sup(a,b)C(G,K2){Gadμ+EbdJ}.

    Some basic properties of this functional are collected in the following lemma.

    Lemma 3.14. The following statements hold:

    (i) For y,zR we have

    α(z,y):=sup(a,b)K2{az+by}={|y|22z if z>0,0 if z=0 and y=0,+otherwise. (19)

    (ii) For μM(G)and JM(E)we have

    B2(μ,J)=sup(a,b)L(G,K2){Gadμ+EbdJ}. (20)

    (iii) The functional B2 is convex and lower semicontinuous with respect to the topology of weak convergence on M(G)×M(E).

    (iv) If μM(G)is nonnegative and JM(E)satisfies Jμ|Ewith J=vμ|E, then we have

    B2(μ,J)=12E|v|2dμ. (21)

    Otherwise, we have B2(μ,J)=+.

    Proof. (ⅰ): see [26,Lemma 5.17].

    (ⅱ): Clearly, the right-hand side of (20) is bounded from below by B2(μ,J). To prove the reverse inequality, let a,b:GR be measurable functions satisfying a+b2/20. By Lusin's theorem (see, e.g., [8,Theorem 7.1.13]) there exist functions aδ,bδC(G,K2) satisfying

    μ({aaδ})δ2,sup|aδ|sup|a|and|J|({bbδ})δ2,sup|bδ|sup|b|.

    Define ˜aδ:=min{aδ,|bδ|2/2}, so that the inequality ˜aδ+b2δ/20 is satisfied. Hence, the pair (˜aδ,bδ) is admissible for the supremum on the right-hand side of (20). Since G˜aδdμ+EbδdJ converges to Gadμ+EbdJ as δ0, we obtain (20).

    (ⅲ): This follows from the definition of B2 as a supremum of linear functionals.

    (ⅳ): Let μ be nonnegative and Jμ with J=vμ. Setting v=0 on GE, (20) and (19) yield

    B2(μ,J)=sup(a,b)L(G,K2){Ga+bvdμ}=12E|v|2dμ.

    To prove the converse, suppose first that there exists a Borel set AG with μ(A)<0. Pick a=k1A and b0 with k0, so that B2(μ,J)kμ(A). Since k can be taken arbitrarily large, we infer that B2(μ,J)=+. Now suppose μ is non-negative, but the signed measure J is not absolutely continuous with respect to μ|E, i.e., there exists a μ-null set AE such that J(A)0. For a=k221A and b=k1A with kR, we have B2(μ,J)kJ(A), which implies the result.

    Proof of Proposition 3.12. (ⅰ): The claim follows readily from the definitions.

    (ⅱ): For φC(Gext) we have

    Gextφ(y)dμε(y)=Gφε(x)dμ(x)=wVφε(w)μ({w})+eEeφε(x)dμ(x).

    For wV, we note that φε(w) is obtained by averaging φ on a subset of the auxiliary edge eextw:

    φε(w)=12εeextw1{d(w,y)2ε}φ(y)dy.

    For eE we obtain, interchanging the order of integration,

    eφε(x)dμ(x)=12ε(e/2,e/2)(αεex+εαεexεφ(y)dy)dμ(x)=12ε(e/22ε,e/2+2ε)φ(y)μ(Iε(y))dy

    Combining these three identities, the desired result follows.

    (ⅲ): Take bounded measurable functions a,b:GR satisfying a+b2/20, extend them by 0 to Gext, and define regularised functions aε,bε:GR as done for φ in (14). By Jensen's inequality and the fact that the regularisation is linear and positivity-preserving, we obtain

    aε(x)+12|bε(x)|2(a+12|b|2)ε(x)0xG,

    i.e., aε and bε are admissible for the supremum in (20). Therefore,

    Gextadμε+EextbdJε=Gaεdμ+EbεdJ12E|v|2dμ. (22)

    The result follows by taking the supremum over all admissible functions a and b.

    (ⅳ): This follows from the uniform convergence of φε to φ; see Proposition 3.9(ⅱ).

    (ⅴ): The function φ is absolutely continuous, hence a.e. differentiable on Eext. Proposition 3.9(ⅰ) yields φεC1(¯E)C(G) and

    φε(x)=αε(x)(φ)ε(x)xE. (23)

    Since αε is constant on each metric edge in E, we obtain, using the continuity equation and the definition of the regularisation,

    ddtGextφdμεt=ddtGφεdμt=Eφεvtdμt=E(φ)εαεvtdμt=Eextφ(αεvεt)dμεt.

    Now we are ready to prove the second part of Theorem 3.7: we adapt the proof of [13], where (much) more general metric measure spaces are treated, but stronger assumptions on the measures are imposed (namely, uniform bounds on the density with respect to the reference measure). Here we consider more general measures using the above regularisation procedure.

    In the proof of the second part of Theorem 3.7, we make use of the Hopf–Lax formula in metric spaces and its relation to the dual problem of optimal transport.

    Definition 3.15 (Hopf–Lax formula). For a real-valued function f on a geodesic Polish space (X,d), we define Qtf:XR{} by

    Qtf(x):=infyX{f(y)+12td2(x,y)}

    for all t>0, and Q0f:=f.

    The operators (Qt)t0 form a semigroup of nonlinear operators with the following well-known properties; see [3] for a systematic study.

    Proposition 3.16 (Hopf–Lax semigroup). Let (X,d) be a geodesic Polish space. For any Lipschitz function f:XR the following statements hold:

    (i) For every t0 we have Lip(Qtf)2Lip(f).

    (ii) For every xX, the map tQtf(x) is continuous on R+, locally semi-concave on (0,), and the inequality

    ddtQtf(x)+12lip(Qtf)2(x)0 (24)

    holds for all t0 up to a countable number of exceptions.

    (iii) The mapping (t,x)lip(Qtf)(x) is upper semicontinuous on (0,)×X.

    Proof. (i): This statement can be derived from [3,Proposition 3.4]. For the convenience of the reader we provide the complete argument here.

    Fix t>0 and xX. For zX we write F(t,x,z):=f(z)+12td2(x,z). We claim that

    Qtf(x)=infzXd(x,z)2tLF(t,x,z),

    where L denotes the Lipschitz constant of f. Indeed, if d(x,z)>2tL, we have

    F(t,x,z)=f(z)+12td2(x,z)f(x)Ld(x,z)+12td2(x,z)>f(x)=F(t,x,x),

    which implies the claim.

    Fix now yX and ε>0. Using the claim, we may pick zX such that d(y,z)2tL and F(t,y,z)Qtf(y)+ε. Then:

    Qtf(x)Qtf(y)F(t,x,z)F(t,y,z)+ε=12t(d2(x,z)d2(y,z))+εd(x,y)2t(d(x,z)+d(y,z))+εd(x,y)2t(d(x,y)+4tL)+ε.

    Reversing the roles of x and y, this estimate readily yields

    lip(Qtf)(x)2L.

    Since X is assumed to be a geodesic space, we have Lip(Qtf)=supxlip(Qtf)(x) and the result follows.

    (ii): See [3,Theorem 3.5].

    (iii): See [3,Propositions 3.2 and 3.6].

    We can now conclude the proof of Theorem 3.7 on the characterisation of absolutely continuous curves in the Wasserstein space over a metric graph.

    Proof of (ⅱ) in Theorem 3.7. Without loss of generality, we set T=1. The main step of the proof is to show that

    W22(μ0,μ1)10vr2L2(μr)dr. (25)

    From there, a simple reparametrisation argument (see also [2,Lemma 1.1.4 & 8.1.3]) yields

    W22(μt,μs)1|st|tsvr2L2(μr)dr

    for all 0s<t1, which implies the absolute continuity of the curve (μt)t(0,1) in W2(G) as well as the desired bound |˙μ|(t)vtL2(μt) for every Lebesgue point t(0,1) of the map tvt2L2(μt).

    Thus, we have to show (25). To this aim, we will work on the supergraph GextG.

    By Kantorovich duality (Proposition 2.3), there exists φC(G) satisfying

    12W22(μ0,μ1)=GQ1φdμ1Gφdμ0. (26)

    Moreover, φ is Lipschitz, which follows from the fact that φ is c-concave with c(x,y)=12d(x,y)2 and (G,d) is compact. We consider Lipschitz continuous extensions of φ and Q1φ to Gext, both constant on each auxiliary metric edge in ¯Eext. In particular, φ and Q1φ are Lipschitz on Gext.

    Set Jt:=vtμt and consider a regularised pair (μεt,Jεt)t(0,1) as defined by (16). We write

    GextQ1φdμε1Gextφdμε0=n1i=0(Gext(Q(i+1)/nφQi/nφ)dμε(i+1)/n+GextQi/nφd(με(i+1)/nμεi/n)), (27)

    and bound the two terms on the right-hand side separately.

    Bound 1. To estimate the first term on the right-hand side of (27), we use (24) to obtain

    n1i=0Gext(Q(i+1)/nφQi/nφ)dμε(i+1)/n12n1i=0Gext(i+1)/ni/nlip2(Qtφ)dtdμε(i+1)/n=12Gext×(0,1)lip2(Qtφ)(x)dμμεn(x,t), (28)

    where the measures μμεn:=n1i=0με(i+1)/nL1|(i/n,(i+1)/n) are defined on Gext×(0,1).

    To show weak convergence of the sequence (μμεn)nN, we take ψC(Gext×[0,1]). Note that tμt is weakly continuous by Lemma 3.10, hence tμεt is weakly continuous as well. Consequently, Gextψ(,t)dμεtn/nGextψ(,t)dμεt for every t. Integrating in time over (0,1), we infer, using dominated convergence, that μμεn converges weakly to μμε:=10μεtδtdt as n.

    As lip2(Qtφ) is not necessarily continuous, an additional argument is required to pass to the limit in (28). For this purpose, we observe that Proposition 3.12.ii yields μμεnλL1 with a density ρεn(x,t)1/(2ε) for xGext and t(0,1). In particular, the family (ρεn)nN is uniformly integrable with respect to λL1. Consequently, the Dunford–Pettis Theorem (see, e.g., [8,Theorem 4.7.18]) implies that (ρεn)nN has weakly-compact closure in L1(Gext×(0,1)). Since lip2(Qtφ) is bounded, we may pass to the limit n in (28) and infer that

    lim supnn1i=0Gext(Q(i+1)/nφQi/nφ)dμε(i+1)/n12Gext×(0,1)lip2(Qtφ)(x)dμμε(x,t)=1210Eextlip2(Qtφ)dμεtdt, (29)

    where we use that μεtλ on Gext to remove the set of nodes V from the domain of integration.

    Bound 2. We now treat the second term in (27). As (μt)t(0,1) belongs to a weak solution to the continuity equation and we know from Proposition 3.9.(ⅰ)) that (Qi/nφ)ε belongs to C1(¯E)C(G), we infer that the mapping tG(Qi/nφ)εdμt is absolutely continuous. Therefore,

    n1i=0GextQi/nφd(με(i+1)/nμεi/n)=n1i=0G(Qi/nφ)εd(μ(i+1)/nμi/n)=n1i=0(i+1)/ni/n(E(Qi/nφ)εdJt)dt=n1i=0(i+1)/ni/n(eEαεee(Qi/nφ)εdJt)dt=n1i=0(i+1)/ni/n(eEαεeeQi/nφdJεt)dtαεmax2n1i=0(i+1)/ni/nEext|Qi/nφ|2dμεtdt+αεmax210Eext|vεt|2dμεtdt, (30)

    where αεmax:=maxeEαεe and Jεt=vεtμεt.

    Proposition 3.16.iiiyields the bound lim supnlip2(Qnt/nφ)lip2(Qtφ). As φ is Lipschitz continuous, (ⅰ) in the same proposition shows that supt,xlip(Qtφ)(x)<. Thus, we may invoke Fatou's lemma to obtain

    lim supnn1i=0(i+1)/ni/nEext|Qi/nφ|2dμεtdt10Eextlip2(Qtφ)dμεtdt.

    Using this estimate together with Proposition 3.12.iii, we obtain

    lim supnn1i=0GextQi/nφd(με(i+1)/nμεi/n)αεmax210Gextlip2(Qtφ)dμεtdt+αεmax210E|vt|2dμtdt. (31)

    Combination of both bounds. Recalling (27), we use (29) and (31) to obtain

    GextQ1φdμε1Gextφdμε0αεmax210E|vt|2dμtdt+αεmax12Gext10lip2(Qtφ)dμεtdt.

    Using Proposition 3.12.iv, the fact that αεmax1, and the bound suptLip(Qtφ)<, we may pass to the limit (ε0) to obtain

    GQ1φdμ1Gφdμ01210E|vt|2dμtdt. (32)

    In view of (26), this yields the result.

    Corollary 1 (Benamou–Brenier formula).For any μ,νP(G), we have

    W22(μ,ν)=min{10E|vt|2dμtdt}, (33)

    where the minimum is taken among all weak solutions to the continuity equation (μt,vtμt)t[0,1] satisfying μ0=μ and μ1=ν.

    Proof. As (P(G),W2) is a geodesic space, we may write

    W22(μ,ν)=min{10|˙μ|2(t)dt},

    where the minimum runs over all absolutely continuous curves (μt)t[0,1] connecting μ and ν. Therefore, the result follows from Theorem 3.7.

    In this section we consider the entropy functional Ent:P(G)(+] defined by

    Ent(μ):={Gρlogρdλif μ=ρλ,+otherwise. (34)

    As is well known, this functional is lower semicontinuous on (P(G),W2); see, e.g., [17,Corollary 2.9].

    A celebrated result by McCann asserts that Ent is geodesically convex on (P(Rd),W2), the Wasserstein space over the Euclidean space Rd. More generally, on a Riemannian manifold M, the relative entropy (with respect to the volume measure) is geodesically κ-convex on (P(M),W2) for κR, if and only if the Ricci curvature is bounded below by κ, everywhere on M [24,12,25].

    Metric graphs are prototypical examples in which such bounds fail to hold. Here we present an explicit example, which shows that the functional Ent on the metric space (P(G),W2) over a metric graph G induced by a graph with maximum degree larger than 2 is not geodesically κ-convex for any κR.

    Example 4.1. Consider a metric graph induced by a graph with 3 leaves as shown in Figure 2.

    Figure 2. 

    The support of probability measures μ and ν on a metric graph induced by a oriented star with 3 leaves

    .

    We impose an edge weight 1 on each of the edges e1,e2,f. Consider the probability measures μ,νP(G) with respective densities ρ,η:GR+ given by

    ρ(x):={12ε1[0,ε](x),xe1 or xe2,0,xf,, η(x):={0,xe1 or xe2,1ε1[1ε,1](x),xf.

    Lemma 4.2. The unique optimal coupling of μ and ν is given by monotone rearrangement from each of the edges e1 and e2 to f, i.e., by the map T with

    T(x)=1ε+xf for xe1 or xe2.

    Proof. Let π be an optimal coupling of μ and ν and decompose it as π=π1+π2, where πi, i=1,2, are couplings of μi, the restriction of μ to ei, and some νi a measure on f such that ν1+ν2=ν. Necessarily π1,π2 are also optimal. By standard optimal transport theory on the interval eifR, πi is given by a map Ti, the monotone rearrangement from μi to νi. If we show that ν1=ν2, then T1=T2=T with T as above and the claim is proven. To this end, assume by contradiction that ν1ν2 and hence T1T2. We consider the coupling π=π1+π2 where π1=π2=(π1+π2)/2 (here, we identify e1f and e2f in the obvious way). Since T1T2, the support of π1 is not contained in the graph of a function. Since π is optimal, the couplings πi are also optimal between their marginals μi and 12ν, and thus have to be induced by the monotone rearrangement map T, a contradiction.

    Consequently, the constant speed-geodesic from μ to ν is given by (Tt#μ)t[0,1] where Tt is the linear interpolation of T and the identity on e1f and e2f respectively, more precisely

    Tt(x):={x+(2ε)tei,if x1(2ε)t,x+(2ε)t1f,if x>1(2ε)t,

    for xei, i{1,2}.

    Set tε0:=1ε2ε and tε1:=12ε. The relative entropy of μt is given by:

    Ent(μt)={log(12ε),t[0,tε0],1ε(1(2ε)t)log(12)+log(1ε),t[tε0,tε1],log(1ε),if t[tε1,1].

    Thus, tEnt(μt) is piecewise affine; see Figure 3. It follows that tEnt(μt) is not κ-convex for any κR.

    Figure 3. 

    Plot of the entropy along the geodesic interpolation

    .

    In this section we study gradient flows in the Wasserstein space over a metric graph. Namely, we consider diffusion equations on metric graphs arising as the gradient flow of free energy functionals composed as the sum of entropy, potential, and interaction energies. We give a variational characterisation of these diffusion equations via energy-dissipation identities and we discuss the approximation of solutions via the Jordan–Kinderlehrer–Otto scheme (minimizing movement scheme). This provides natural analogues on metric graphs of the corresponding classical results in Euclidean space. We follow the approach from the Euclidean case; see in particular [2,Section 10.4], and adapt it to the current setting.

    Let V:¯ER be Lipschitz continuous and define the weighted volume measure m:=eVλ on G (since λ gives no mass to vertices, the potential ambiguity of V there does not matter). We consider the following functionals on P(G):

    (i) the relative entropy EV:P(G)(,] defined by

    EV(μ):=Ent[μ|m]={Gρ(x)logρ(x)dm(x) if μ=ρm,+ otherwise. (35)

    (ii) the interaction energy W:P(G)R defined by

    W(μ):=12G×GW(x,y)dμ(x)dμ(y),

    where W:G×GR is symmetric and Lipschitz continuous.

    Moreover, we define F:P(G)(,+] as the sum of the previous quantities:

    F:=EV+W.

    Note that EV is bounded from below (by Jensen's inequality and finiteness of m) and lower semicontinuous with respect to weak convergence. Moreover, W is bounded and continuous with respect to weak convergence.

    Further note that for μP(G) with μλ we can write

    EV(μ)=E(μ)+V(μ),

    where E(μ)=Ent[μ|λ] is the Boltzmann entropy on G and V:P(G)R defined by

    V(μ):=GV(x)dμ(x)

    is the potential energy. The latter is well defined for μλ despite the potential discontinuity of V at the vertices.

    We consider the following diffusion equation on the metric graph G given by

    tη=Δη+(η(V+W[μ])). (36)

    Here μ=ηλ is a probability on G and we set W[μ](x):=GW(x,y)dμ(y). In analogy with the classical Euclidean setting we will show below that this PDE is the Wasserstein gradient flow equation of the free energy F. Though the setting of metric graphs is one-dimensional, we prefer to use multi-dimensional notation such as Δ and for the sake of clarity.

    We consider the following notion of weak solution for (36).

    Definition 5.1. We say that a curve (ηt)t[0,T] of probability densities w.r.t. λ on G is a weak solution to (36) if for ρt:=ηteV we have ρtW1,1(¯E)C(G) for a.e. t, and the pair (μ,J) given by

    μt=ρtmandJt=(ρt+ρtW[μt])m

    is a weak solution to the continuity equation in the sense of Definition 3.2, i.e., we ask that tμt is weakly continuous and for every φC1c((0,T)ׯE) such that φt,tφtC1(¯E)C(G) for all t(0,T), we have

    T0(Gtφρtdm¯Eφ(ρt+ρtW[μt])dm)dt=0. (37)

    Remark 2. Let us briefly consider the special case where V=W=0. Then (36) is simply the heat equation on a metric graph, which has been introduced already in [18]. It is known since [15,22] that the Laplacian with natural vertex conditions (continuity across the vertices along with a Kirchhoff-type condition on the fluxes) is associated with a Dirichlet form, hence the Cauchy problem for this PDE is well-posed on L2(G): more precisely, it is governed by an ultracontractive, Markovian C0-semigroup (etΔ)t0 that extrapolates to Lp(G) for all p1, as well as to C(G) and, by duality, to the space M(G) of Radon measures on G. For any initial value μ0M(G), (t,x)ρ(t,x)=(etΔμ0)(x) defines a classical solution to the Cauchy problem for the heat equation with initial value μ0. It is easy to see that this solution is a weak solution in the sense of Definition 5.1 as well.

    The dissipation of the free energy along solutions to (36) at μ=ρm is formally given by

    ddtF(μ)=E|ρρ+W[μ]|2dμ.

    This motivates the following definition.

    Definition 5.2 (Energy dissipation functional). The energy dissipation functional I:P(G)[0,+] is defined as follows. If μ=ρm with ρW1,1(¯E)C(G) and ρ+ρW[μ]=wρ for some wL2(μ) we set

    I(μ):=E|w|2dμ.

    Otherwise, we set I(μ)=+.

    Remark 3. We emphasize that continuity of ρ on G is required for finiteness of I(μ) in this definition. It is not sufficient that ρ belongs to W1,1(¯E). The continuity is important in the proof of Theorem 5.5 below, as it ensures spatial continuity for gradient flow curves (i.e., curves of maximal slope with respect to the upper gradient I) and it allows us to identify them with weak solutions to the diffusion equation (36). The requirement of spatial continuity for weak solutions to (36) in Definition 5.1 is essential in order to couple the dynamics on different edges across common vertices.

    We collect the following properties of the dissipation functional.

    Lemma 5.3. Let {μn}nP(G) be a sequence weakly converging to μP(G) such that

    supnF(μn)< and supnI(μn)<. (38)

    Then we have

    I(μ)lim infnI(μn). (39)

    Proof. First note that we can rewrite I(μ) as the integral functional

    I(μ)=Gα(μ,J)=Eα(dμdσ,dJdσ)dσ, (40)

    where α:[0,)×R[0,] is the lower semicontinuous and convex function defined by

    α(s,u)={|u|2/sif s>0,0if s=0 and u=0,+otherwise, (41)

    and J=wwμ=(ρ+ρW[μ])m. Here σ is any measure such that μ,Jσ; its choice is irrelevant by 1-homogeneity of α.

    Now, let μn=ρnm. Lower semicontinuity of F and (38) imply that F(μ)<. Therefore we can write μ=ρm for a suitable density ρ. Superlinearity of rrlogr implies that ρn converges weakly to ρ in L1(G,m).

    Recall that I(μn)=|wwn|2dμn with Un:=ρnwwn=ρn+ρnW[μn]. Hölder's inequality and the bound (38) yield that the measures Jn=Unm have uniformly bounded total variation. Hence up to extracting a subsequence, we have that Jn converges weakly* to a measure J on E. Lower semicontinuity of the integral functional Gα yields

    Gα(μ,J)lim infnI(μn)<. (42)

    This allows us to write J=wwρm for a wwL2(μ) and Gα(μ,J)=|ww|2dμ. Since ρnW1,1(¯E)C(G), we have for any function sC1c(E) by integration by parts,

    Gρnsdλ=Gsρndλ=GseV(ρn+ρnW[μn])dmGseVW[μn]dμn=GseVdJnGseVW[μn]dμn. (43)

    The convergence of Jn to J weakly and ρn to ρ in L1 together with the fact that J=wwρm gives no mass to V and the boundedness of W allows us to pass to the limit and obtain

    Gρsdλ=GseVdJGseVW[μ]dμ=GswwρdλGsρW[μ]dλsC1c(E).

    We infer that ρW1,1(¯E) and that ρww=ρ+ρW[μ]. In particular, ρC(¯E) by the Sobolev embedding theorem.

    Let us now show that ρC(G). For this purpose, note that each pair of adjacent edges (e,f) can be identified with the interval [e,f]. Consider sC1(E) such that s=0 for all ge,f and s=r on e,f for some rC1c(e,f). Repeating the argument above, we infer that ρW1,1([e,f]) and in particular that ρ is continuous at 0. We thus obtain that ρC(G). Hence from (42) we obtain the claim.

    Let us denote by I0 the energy dissipation functional with W=0, more precisely, if μ=ρm with ρW1,1(¯E)C(G) with ρ=wρ for some wL2(μ) we set

    I0(μ):=E|w|2dμ.

    Otherwise, we set I0(μ)=+. As already observed for the functional I, we can write I0(μ)=α(ρ,ρ)dm, with μ=ρm where α is the function in (41). Then I0 is a convex and lower semi-continuous functional by the previous lemma. We note that under the assumption that W is Lipschitz, I(μ) is finite if and only if I0(μ) is finite.

    Next, we observe that finiteness of the I0 implies a quantitative L bound on the density.

    Lemma 5.4. For μP(G) with I0(μ)< and μ=ρm we have ρC(G) with

    ρAI0(μ),

    where the constant A>0 depends on V.

    Proof. The continuity of ρ follows from the definition of I0(μ). Using Hölder's we obtain with ρ=ρw,

    |ρ|dλeV|ρ|dmeV(|w|2dμ)12.

    The claim then follows immediately from the Sobolev embedding theorem applied to each of the finitely many edges.

    The main result of this section is the following gradient flow characterisation of the diffusion equation (36).

    Theorem 5.5. For any 2-absolutely continuous curve (μt)t[0,T] in (P(G),W2) with F(μ0)< we have

    LT(μ):=F(μT)F(μ0)+12T0|˙μ|2(r)+I(μr)dr0.

    Moreover, we have LT(μ)=0 if and only if μt=ρtm is a weak solution to (36) in the sense of Definition 5.1.

    The main step in proving this result is to establish a chain rule for the free energy F along absolutely continuous curves in (P(G),W2). Recall the definition of α from (41).

    Proposition 5.6 (Chain rule). Let (μt)t[0,T] be a 2-absolutely continuous curve in (P(G),W2) satisfying T0I(μt)dt<+.Write μt=ρtm and let Jt=Utm be an optimal family of momentum vector fields, i.e., (μt,Jt)t[0,T] solves the continuity equation and|˙μ|2(t)=α(ρt,Ut)dm.Let ρtwt=ρt+ρtW[μt] as in Definition 5.2. Then, tF(μt)is absolutely continuous and we have

    ddtF(μt)=Ewt,dJtfor a.e.t[0,T]. (44)

    Proof. We first note that the assumptions ensure that also T0I0(μt)dt<. Hence, we have

    T0Eα(ρt,ρt)dmdt<,T0Eα(ρt,Ut)dmdt<. (45)

    We proceed by a twofold regularisation. Using a family (ηδ)δ>0 of even and smooth approximation kernels with compact support in [δ,δ], we regularise in time via

    ρδt:=δδηδ(s)ρtsds,

    and set μδt=ρδtm. Here we extend ρ to a curve on the time-interval [δ,T+δ] which is constant on [δ,0] and [T,T+δ]. Similarly defining Jδt as the time-regularisation of Jt we obtain that (μδ,Jδ) is a solution to the continuity equation. Convexity of I0 and the Benamou–Brenier functional yield that T0I0(μδt)dt and T0|˙μδ|2(t)dt<. Hence, (45) also holds with ρδt,Uδt in place of ρt,Ut. Moreover, we must have Uδ=0 on {ρδ=0}.

    Further, we regularise the logarithm and define for ε>0 the function Fε:[0,)R by setting

    Fε(r)=(r+ε)log(r+ε)εlogε.

    Then we define the regularized free energy Fε of μ=ρm via

    Fε(μ)=GFε(ρ)dm+12G×GW(x,y)dμ(x)dμ(y).

    Let us set gε,δ=Fε(ρδ)=1+log(ρδ+ε). Note that by Lemma 5.4, ρδ is bounded and thus also gε,δ is bounded. We will first show that

    Fε(μδT)Fε(μδ0)=T0Ggε,δt+W[μδt],dJδtdt. (46)

    Then passing to the limit δ,ε0 will yield the claim.

    While establishing (46) we write g instead of gε,δ for simplicity. We have

    Fε(μδT)Fε(μδ0)=T0G(Fε(ρδ)+W[μδt])tρδtdmdt.

    Differentiation under the integral sign is justified by boundedness of Fε(ρδ) and W and the regularity in time of ρδ. Note that W[μδt] is an admissible test function in the continuity equation for (μδ,Jδ). Thus, to establish (46) it remains to show that the continuity equation can also be used on the function g=Fε(ρδ). Recall that g is bounded with g=ρδ/(ρδ+ε). We can approximate g uniformly by functions gαC1(¯E×[0,T])C(G×[0,T]) as follows. First extend g to Gext×[0,T] with constant values on the additional edge incident to vertex v equal to the value of g at v. Then we apply the regularising procedure (14). The continuity equation for (μδ,Jδ) yields

    T0Ggαttρδtdmdt=T0Egαt,dJδt.

    Passing to the limit as α0 then will finish the proof of (46). Convergence of the left hand side is immediate since g is bounded and so gα is bounded uniformly in α. For the right hand side we estimate with Jδ=Uδm:

    |T0Egαtgt,dJδtdt|(T0E|gαtgt|2ρδtdmdt)12×(T0Eα(ρδt,Uδt)dmdt)12.

    The second factor is finite as noted above. To estimate the first factor, we recall that ρδ is bounded. Hence for a suitable constant C<

    T0E|gαtgt|2ρδtdmdtCT0E|gαtgt|2dλdt

    Using (15) and dominated convergence, the latter term goes to zero as α0 if we show that

    T0E|g|2dλdt<.

    But using g=ρδ/(ρδ+ε) we can estimate

    T0E|g|2dλdteVT0I0(μδt)dt<.

    Thus, (46) is established.

    We will now pass to the limits δ,ε0 in (46), starting with the right-hand side.

    For the limit δ0, note that UδU a.e. and wwε,δwwε a.e. as δ0 with wwε=ρ/(ρ+ε)+W[μ]. Dominated convergence then yields that as δ0:

    T0Ewwε,δt,UδtdmdtT0Ewwεt,Utdmdt.

    Indeed, we have the majorant

    |wwε,δ,Uδ|12|wwε,δ|2(ρδ+ε)+12|Uδ|2ρδ+εα(ρδ,ρδ)+C(ρδ+ε)+12α(ρδ,Uδ)[α(ρ,ρ)]δ+C(ρδ+ε)+12[α(ρ,U)]δ

    for a suitable constant C, using the fact W[ρδ] is uniformly bounded. Here, []δ denotes the time-regularisation of the function in brackets and we have used Jensen's inequality in the last step to interchange the regularisation operation with the convex function α. Since by assumption α(ρ,ρ) and α(ρ,U) are integrable on [0,T]×E, the majorant above indeed converges in L1([0,T]×E).

    To further pass to the limit ε0, we note that wwεww a.e. on the set {ρ>0}. Since U=0 a.e. on the set {ρ=0}, we conclude that wwε,Uww,U a.e. Using similar as before the majorant |wwε,U|α(ρ,ρ)+Cρ+α(ρ,U)/2, we conclude by dominated convergence.

    It remains to pass to the limit on the left-hand side in (46). The weak continuity of tμt implies that μδtμt weakly as δ0. This is sufficient to conclude that W(μδt)W(μt). Note that for any μ=ρmP(G) we have

    Fε(ρ)dm=EV(με)Mεlogε,

    with με=(1+M)1(μ+εm) and M=m(G). Convexity of rrlogr and Jensen's inequality yield that EV(μδ,εt)EV(μεt). Thus, lower semicontinuity of EV under weak convergence shows that EV(μδ,εt)EV(μεt) and hence Fε(ρδt)dmFε(ρt)dm as δ0. The limit ε0 is then easily achieved by monotone convergence. From the convergence of the right-hand side of (46) and the assumption that F(μ0)< we finally conclude that F(μt)< for all t>0 and that

    F(μT)F(μ0)=T0wwt,dJtdt.

    Hence tF(μt) is absolutely continuous and (44) follows.

    We can now prove Theorem 5.5.

    Proof of Theorem 5.5. Note that the right-hand side of (44) may be estimated by means of Hölder's and Young's inequality as

    Ewt,UtdmE|wt|2ρtdmE|Ut|2ρtdm12E|wt|2ρtdm12E|Ut|2ρtdm. (47)

    Hence, by integrating both sides of (44) from 0 to T we obtain that LT(μ)0. Moreover, we have equality if and only if for a.e. t and μt-a.e. we have Ut=ρtwt. Now the continuity equation with Jt=Utm=ρtwtm=ρt+ρtW[μt]m becomes the weak formulation of (36)

    Here, we recast the variational characterisation of McKean–Vlasov equations on metric graphs from the previous section in the language of the theory of gradient flows in metric spaces. Let us briefly recall the basic objects. For a detailed account we refer the reader to [2].

    Let (X,d) be a complete metric space and let E:X(,] be a function with proper domain, i.e., the set D(E):={x:E(x)<} is non-empty.

    The following notion plays the role of the modulus of the gradient in a metric setting.

    Definition 5.7 (Strong upper gradient). A function g:X[0,] is called a strong upper gradient of E if for any xAC((a,b);(X,d)) the function gx is Borel and

    |E(xs)E(xt)|tsg(xr)|˙x|(r)drastb .

    Note that by the definition of strong upper gradient, and Young's inequality ab12(a2+b2), we have that for all st:

    E(xt)E(xs)+12tsg(xr)2+|˙x|2(r)dr0.

    Definition 5.8 (Curve of maximal slope). A locally 2-absolutely continuous curve (xt)t(0,) is called a curve of maximal slope of E with respect to its strong upper gradient g if tE(xt) is non-increasing and

    E(xt)E(xs)+12tsg(xr)2+|˙x|2(r)dr00<st. (48)

    We say that a curve of maximal slope starts from x0X if limt0xt=x0.

    Equivalently, we can require equality in (48). If a strong upper gradient g of E is fixed we also call a curve of maximal slope of E (relative to g) a gradient flow curve.

    Finally, we define the (descending) metric slope of E as the function |E|:D(E)[0,] given by

    |E|(x)=lim supyxmax{E(x)E(y),0}d(x,y). (49)

    The metric slope is in general only a weak upper gradient for E, see [2,Theorem 1.2.5]. In our application to metric graphs we will show that the square root of the dissipation I provides a strong upper gradient for the free energy F.

    Corollary 2. The functional I is a strong upper gradient of F on (P(G),W2). The curves of maximal slope for F with respect to this strong upper gradient coincide with weak solutions to (36) satisfying T0I(μt)dt<.

    Proof. For a 2-absolutely continuous curve with (μt)t with T0I(μt)dt<, the chain rule (44) together with the estimate (47) yields

    |F(μt)F(μs)|tsI(μr)|˙μr|drs,t[0,T]:st, (50)

    i.e., I is a strong upper gradient. Theorem 5.5 yields the identification of curves of maximal slope.

    The dissipation functional I(μ) can be related to the metric slope of the free energy F under suitable conditions on μP(G).

    Lemma 5.9. For any μP(G) we have

    I(μ)|F|2(μ).

    Proof. We assume that μ=ρm=˜ρλP(G) satisfies |F|(μ)<, since otherwise there is nothing to prove.

    Step 1. We show first that ρW1,1(¯E) and ρ+ρW[μ]=ρww with wwL2(μ).

    For this purpose, take any ssC(¯E) which vanishes in a neighbourhood of every node and put ~ss=eVss. As a consequence, the mapping rt:GG defined by rt:=Id+t~ss maps each edge into itself for t>0 sufficiently small. We claim that:

    limt0F((rrt)#μ)F(μ)t=G[ρss+ssρW[μ]]dλ. (51)

    To show this, note that rrt is injective for every t small enough. Therefore, the change of variables formula yields that the density ˜ρt of (rrt)#μ with respect to λ satisfies

    ˜ρ(x)=˜ρt(rrt(x))rrt(x);.

    Consequently, with F(r)=rlogr we have

    F((rrt)#μ)F(μ)=GF(˜ρrrt)rrtF(˜ρ)dλ+G[VrrtV]dμ+G×G[W(rrtrrt)W]dμμ.

    Note that rrt=1+t~ss. Dividing by t and letting t0 and noting that ~ss=eV(ss+ssV) and ˜ρ~ss=ρss we deduce (51).

    For t>0 sufficiently small, we then have

    W2(μ,(rrt)#μ)t~ssL2(μ).

    This estimate, together with (51), implies

    G[ρss+ssρW[μ]]dλ=limt0F((rrt)#μ)F(μ)t|F|(μ)~ssL2(μ).

    Hence, the left-hand side defines an L2(μ)-bounded linear functional on a subspace of L2(μ). Using the Hahn–Banach theorem we may extend this functional to L2(μ). The Riesz representation theorem then yields a unique element wL2(μ) such that wL2(μ)|F|(μ) and

    G[ρss+ssρ(V+W[μ])]dλ=Gww~ssdμ=Gwwρsdλ (52)

    for all ss as above.

    Considering in particular functions ss supported on a single edge, we infer that ρW1,1(¯E) and ρ+ρW[μ]=ρww with wwL2(μ). In particular, we have ρC(¯E) by Sobolev embedding.

    Step 2. Next we show that ρ belongs to C(G). For this purpose we repeat the argument above, for a different class of functions ss.

    Consider a pair of adjacent edges e,fE with common vertex v. For the moment, we identify the concatenation of the two edges with the interval [e,f] such that v corresponds to 0. Let s:[e,f]R be a C-function vanishing in a neighbourhood of e and f. In particular, rt:=Id+t˜s maps [e,f] into itself for t>0 sufficiently small, where ˜s=eVs. Let the maps ss:ER and rt:EG be defined on ef by s and rt through the identification with [e,f] and by setting ss=0 and rt=id on all other edges. Repeating the argument from Step 1, we infer that ρW1,1([e,f]) with the identification of e,f with [e,f] as above. In particular, ρ is continuous at 0. Since the choice of the pair e,f is arbitrary, we conclude that ρC(G).

    Combining both steps, we infer that I(μ)< and

    I(μ)=ww2L2(μ)|F|2(μ).

    From Lemma 5.9 and the lower semicontinuity of I we immediately infer that I lies below the relaxed metric slope, i.e., the lower semicontinuous relaxation of |F| given by

    |F|(μ):=inf{lim infn|F|(μn):μnμ}.

    Corollary 3. For all μP(G) we have I(μ)|F|(μ).

    In this section, we consider the time-discrete variational approximation scheme of Jordan–Kinderlehrer–Otto for the gradient flow [14].

    Given a time step τ>0 and an initial datum μ0P(G) with F(μ0)<, we consider a sequence (μτn)n in P(G) defined recursively via

    μτ0=μ0,μτnargminν{F(ν)+12τW2(ν,μτn1)2}. (53)

    Then we build a discrete gradient flow trajectory as the piecewise constant interpolation (ˉμτt)t0 given by

    ˉμτ0=μ0,ˉμτt=μτn if t((n1)τ,nτ]. (54)

    Then we have the following result.

    Theorem 5.10. For any τ>0 and μ0P(G) with F(μ0)< the variational scheme (53) admits a solution (μτn)n. As τ0, for any family of discrete solutions there exists a sequence τk0 and a locally 2-absolutely continuous curve (μt)t0 such that

    ˉμτktμtt[0,). (55)

    Moreover, any such limit curve is a gradient flow curve of F, i.e., a weak solution to the diffusion equation (36).

    Proof of Theorem 5.10. The result basically follows from general results for metric gradient flows, where the scheme is known as the minimizing movement scheme; see [2,Section 2.3]. We consider the metric space (P(G),W2) and endow it with the weak topology σ. It follows that [2,Assumptions 2.1 (a,b,c)] are satisfied. Existence of a solution to the variational scheme (53) and of a subsequential limit curve (μt)t now follows from [2,Corollary 2.2.2,Proposition 2.2.3]. Moreover, [2,Theorem 2.3.2] gives that the limit curve is a curve of maximal slope for the strong upper gradient |F|, i.e.,

    12t0|˙μ|2(r)+|F(μr)|2dr+F(μt)F(μ0).

    Thus, by Corollary 3, it is also a curve of maximal slope for the strong upper gradient I. Theorem 5.5 yields the identification with weak solutions to (36).

    ME acknowledges funding by the Deutsche Forschungsgemeinschaft (DFG), Grant SFB 1283/2 2021 – 317210226. DF and JM were supported by the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (grant agreement No 716117). JM also acknowledges support by the Austrian Science Fund (FWF), Project SFB F65. The work of DM was partially supported by the Deutsche Forschungsgemeinschaft (DFG), Grant 397230547. This article is based upon work from COST Action 18232 MAT-DYN-NET, supported by COST (European Cooperation in Science and Technology), www.cost.eu. We wish to thank Martin Burger and Jan-Frederik Pietschmann for useful discussions. We are grateful to the anonymous referees for their careful reading and useful suggestions.



    [1] A user's guide to optimal transport. Modelling and Optimisation of Flows on Networks (2013) 2062: 1-155.
    [2]

    L. Ambrosio, N. Gigli and G. Savaré, Gradient Flows: In Metric Spaces and in the Space of Probability Measures, 2nd edition, Springer Science & Business Media, 2008.

    [3] Calculus and heat flow in metric measure spaces and applications to spaces with Ricci bounds from below. Invent. Math. (2014) 195: 289-391.
    [4] A numerical method for the optimal time-continuous mass transport problem and related problems. Monge Ampére Equation: Applications to Geometry and Optimization (1999) 226: 1-12.
    [5] A computational fluid mechanics solution to the Monge-Kantorovich mass transfer problem. Numer. Math. (2000) 84: 375-393.
    [6]

    G. Berkolaiko and P. Kuchment, Introduction to Quantum Graphs, Mathematical Surveys and Monographs, 186. American Mathematical Society, Providence, RI, 2013.

    [7]

    M. Bernot, V. Caselles and J.-M. Morel, Optimal Transportation Networks, Models and theory. Lecture Notes in Mathematics, 1955. Springer-Verlag, Berlin, 2009.

    [8]

    V. I. Bogachev, Measure Theory, Springer-Verlag, Berlin, 2007.

    [9]

    D. Burago, Y. Burago and S. Ivanov, A Course in Metric Geometry, Studies in Mathematics, 33. American Mathematical Society, Providence, RI, 2001.

    [10]

    M. Burger, I. Humpert and J.-F. Pietschmann, Dynamic optimal transport on networks, arXiv: 2101.03415, 2021.

    [11] Fokker-Planck equations for a free energy functional or Markov process on a graph. Arch. Ration. Mech. Anal. (2012) 203: 969-1008.
    [12] A Riemannian interpolation inequality à la Borell, Brascamp and Lieb. Invent. Math. (2001) 146: 219-257.
    [13] The continuity equation on metric measure spaces. Calc. Var. Partial Differential Equations (2015) 53: 149-177.
    [14] The variational formulation of the Fokker–Planck equation. SIAM J. Math. Anal. (1998) 29: 1-17.
    [15] Variational and semigroup methods for waves and diffusion in networks. Appl. Math. Optim. (2007) 55: 219-240.
    [16] Quantum graphs: An introduction and a brief survey. Analysis on Graphs and Its Applications (2008) 77: 291-312.
    [17] Optimal entropy-transport problems and a new Hellinger-Kantorovich distance between positive measures. Invent. Math. (2018) 211: 969-1117.
    [18] Connecting of local operators and evolution equations on networks. Potential Theory (Proc. Copenhagen 1979) (1980) 234: 230-243.
    [19] Gradient flows of the entropy for finite Markov chains. J. Funct. Anal. (2011) 261: 2250-2292.
    [20] Optimal mass transport on metric graphs. SIAM J. Optim. (2015) 25: 1609-1632.
    [21] A gradient structure for reaction-diffusion systems and for energy-drift-diffusion systems. Nonlinearity (2011) 24: 1329-1346.
    [22] Gaussian estimates for a heat equation on a network. Networks Het. Media (2007) 2: 55-79.
    [23]

    D. Mugnolo, Semigroup Methods for Evolution Equations on Networks, Understanding Complex Systems. Springer, Cham, 2014.

    [24] Generalization of an inequality by Talagrand and links with the logarithmic Sobolev inequality. J. Funct. Anal. (2000) 173: 361-400.
    [25] Transport inequalities, gradient estimates, entropy, and Ricci curvature. Comm. Pure Appl. Math. (2005) 58: 923-940.
    [26]

    F. Santambrogio, Optimal Transport for Applied Mathematicians, Progress in Nonlinear Differential Equations and their Applications, 87. Birkhäuser/Springer, Cham, 2015.

    [27]

    C. Villani, Optimal Transport: Old and New, volume 338., Springer-Verlag, Berlin, 2009.

    [28]

    Q. Xia, Optimal paths related to transport problems, Commun. Contemp. Math., 5 2003,251–279.

  • This article has been cited by:

    1. Martin Burger, Antonio Esposito, Porous medium equation and cross-diffusion systems as limit of nonlocal interaction, 2023, 235, 0362546X, 113347, 10.1016/j.na.2023.113347
    2. Qiang Du, Amir Sagiv, Minimizing Optimal Transport for Functions with Fixed-Size Nodal Sets, 2023, 33, 0938-8974, 10.1007/s00332-023-09952-8
    3. Martin Burger, Ina Humpert, Jan-Frederik Pietschmann, Dynamic Optimal Transport on Networks, 2023, 29, 1292-8119, 54, 10.1051/cocv/2023027
    4. Ariane Fazeny, Martin Burger, Jan-F. Pietschmann, Optimal transport on gas networks, 2025, 0956-7925, 1, 10.1017/S0956792525000051
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1769) PDF downloads(246) Cited by(4)

Figures and Tables

Figures(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog