Dissipative probability vector fields and generation of evolution semigroups in Wasserstein spaces
Abstract.
We introduce and investigate a notion of multivalued -dissipative probability vector field (MPVF) in the Wasserstein space of Borel probability measures on a Hilbert space . Taking inspiration from the theory of dissipative operators in Hilbert spaces and of Wasserstein gradient flows of geodesically convex functionals, we study local and global well posedness of evolution equations driven by dissipative MPVFs. Our approach is based on a measure-theoretic version of the Explicit Euler scheme, for which we prove novel convergence results with optimal error estimates under an abstract CFL stability condition, which do not rely on compactness arguments and also hold when has infinite dimension.
We characterize the limit solutions by a suitable Evolution Variational Inequality (EVI), inspired by the Bénilan notion of integral solutions to dissipative evolutions in Banach spaces. Existence, uniqueness and stability of EVI solutions are then obtained under quite general assumptions, leading to the generation of a semigroup of nonlinear contractions.
Key words and phrases:
Measure differential equations/inclusions in Wasserstein spaces, probability vector fields, dissipative operators, evolution variational inequality, explicit Euler scheme.1991 Mathematics Subject Classification:
Primary: 34A06, 34A45; Secondary: 34A12, 34A34, 34A60, 28A501. Introduction
The aim of this paper is to study the local and global well posedness of evolution equations for Borel probability measures driven by a suitable notion of probability vector fields in an Eulerian framework.
For the sake of simplicity, let us consider here a finite dimensional Euclidean space with scalar product and norm (our analysis however will not be confined to finite dimension and will be carried out in a separable Hilbert space) and the space (resp. ) of Borel probability measures in (resp. with bounded support).
A Cauchy-Lipschitz approach, via vector fields
A first notion of vector field can be described by maps , typically taking values in some subset of continuous vector fields in (as the locally Lipschitz ones of ), and satisfying suitable growth-continuity conditions. In this respect, the evolution driven by can be described by a continuous curve , , starting from an initial measure and satisfying the continuity equation
(1.1a) | |||||
(1.1b) |
in the distributional sense, i.e.
(1.2) |
If is sufficiently smooth, solutions to (1.1a,b) can be obtained by many techniques. Recent contributions in this direction are given by the papers [Pic19, Pic18, BF21, Cav+20], we also mention [PR14, PR19] for the analysis in presence of sources. In particular, in [BF21] the aim of the authors is to develop a suitable Cauchy-Lipschitz theory in Wasserstein spaces for differential inclusions which generalizes (1.1b) to multivalued maps and requires (1.1b), (1.2) to hold for a suitable measurable selection of . As it occurs in the classical finite-dimensional case, the differential-inclusion approach is suitable to describe the dynamics of control systems, when the velocity vector field involved in the continuity equation depends on a control parameter.
The Explicit Euler method
A natural approach, that is suitable for a great generalization, is to approximate (1.1a,b) by a measure-theoretic version of the Explicit Euler scheme. Choosing a step size and a partition of the interval , , we construct a sequence , by the algorithm
(1.3) |
where is the identity map and denotes the push forward of induced by a Borel map and defined by for every Borel set . If is the piecewise constant interpolation of the discrete values , one can then study the convergence of as , hoping to obtain a solution to (1.1a,b) in the limit.
It is then natural to investigate a few relevant questions:
-
E.1
what is the most general framework where the Explicit Euler scheme can be implemented,
-
E.2
what are the structural conditions ensuring its convergence,
-
E.3
how to characterize the limit solutions and their properties.
Concerning the first question E.1, one immediately realizes that each iteration of (1.3) actually depends on the probability distribution on the tangent bundle (which we may identify with , where the second component plays the role of velocity)
whose first marginal is . If we denote by the projections , and by the exponential map in the flat space , we recover by a single step of “free motion” driven by and given by
This operation does not depend on the fact that is concentrated on the graph of a map (in this case ): one can more generally assign a multivalued map such that for every every measure has first marginal . We call a multivalued probability vector field (MPVF in the following), which is in good analogy with a Riemannian interpretation of . The disintegration of with respect to provides a (unique up to -negligible sets) Borel family of probability measures on the space of velocities such that . is induced by a vector field only if is a Dirac mass for -a.e.. (1.3) now reads as
(1.4) |
Besides greater generality, this point of view has other advantages: working with the joint distribution instead of the disintegrated vector field potentially allows for the weakening of the continuity assumption with respect to . This relaxation corresponds to the introduction of Young’s measures to study the limit behaviour of weakly converging maps [CRV04]. Adopting this viewpoint, the classical discontinuous example in (see [Fil88]), where , admits a natural closed realization as MPVF given by
In particular, (see also [Cam+21, Example 6.2]).
The study of measure-driven differential equations/inclusions is not new in the literature \citesDalMasoRampazzo,SilvaVinter. However, these studies, devoted to the description of impulsive control systems [Bre96] and mainly motivated by applications in rational mechanics and engineering, have been used to describe evolutions in rather than in the space of measures.
A second advantage in considering a MPVF is the consistency with the theory of Wasserstein gradient flows generated by geodesically convex functionals introduced in [AGS08] (Wasserstein subdifferentials are particular examples of MPVFs) and with the multivalued version of the notion of probability vector fields introduced in [Pic19, Pic18], whose originating idea was indeed to describe the uncertainty affecting not only the state of the system, but possibly also the distribution of the vector field itself.
A third advantage is to allow for a more intrinsic geometric view, inspired by Otto’s non-smooth Riemannian interpretation of the Wasserstein space: probability vector fields provide an appropriate description of infinitesimal deformations of probability measures, which should be measured by, e.g., the -Kantorovich-Rubinstein-Wasserstein distance
(1.5) |
where is the set of couplings with marginals and respectively. It is well known [AGS08, Vil09, San15] that if belong to the space of Borel probability measures with finite second moment
then the minimum in (1.5) is attained in a compact convex set and is a complete and separable metric space. Adopting this viewpoint and proceeding by analogy with the theory of dissipative operators in Hilbert spaces, a natural class of MPVFs for evolutionary problems should at least satisfy a -dissipativity condition, , as
(1.6) |
Metric dissipativity
Condition (1.6) in the simple case has a clear interpretation in terms of one step of the Explicit Euler method: it is an asymptotic contraction as the time step goes to . By using the properties of the Wasserstein distance, we will first compute the right derivative of the (squared) Wasserstein distance along the deformation
(1.7) | ||||
and we will show that (1.6) admits the equivalent characterization
(1.8) |
If we interpret the left hand side of (1.8) as a sort of Wasserstein pseudo-scalar product of and along the direction of an optimal coupling between and , (1.8) is in perfect analogy with the canonical definition of -dissipativity (also called one-sided Lipschitz condition) for a multivalued map :
(1.9) |
It turns out that the (opposite of the) Wasserstein subdifferential [AGS08, Section 10.3] of a geodesically -convex functional is a MPVF and satisfies a condition equivalent to (1.6) and (1.8). We also notice that (1.8) reduces to (1.9) in the particular case when are Dirac masses in .
Conditional convergence of the Explicit Euler method
Differently from the Implicit Euler method, however, even if a MPVF satisfies (1.8), every step of the Explicit Euler scheme (1.4) affects the distance by a further quadratic correction according to the formula
which depends on the order of magnitude of and , and thus of , at and .
Our first main result (Theorems 7.5,7.7), which provides an answer to question E.2, states that if is a -dissipative MPVF according to (1.8) then every family of discrete solutions of (1.4) in an interval satisfying the abstract CFL condition
(1.10) |
is uniformly converging to a limit curve starting from , with a uniform error estimate
(1.11) |
and a universal constant . Apart from the precise value of , the estimate (1.11) is sharp [Rul96] and reproduces in the measure-theoretic framework the celebrated Crandall-Liggett [CL71] estimate for the generation of dissipative semigroups in Banach spaces. We derive it by adapting to the metric-Wasserstein setting the relaxation and doubling variable techniques of [NS06], strongly inspired by the ideas of Kružkov [Kru70] and Crandall-Evans [CE75].
This crucial result does not require any bound on the support of the measures neither local compactness of the underlying space , so that we will prove it in a general Hilbert space, possibly with infinite dimension. Moreover, if are two limit solutions starting from we show that
as it happens in the case of gradient flows of -convex functions. Once one has these building blocks, it is not too difficult to construct a local and global existence theory, mimicking the standard arguments for ODEs.
Metric characterization of the limit solution
As we stated in question E.3, a further important point is to get an effective characterization of the solution obtained as limit of the approximation scheme.
As a first property, considered in [Pic19, Pic18] in the case of a single-valued PVF, one could hope that satisfies the continuity equation (1.1a) coupled with the barycentric condition replacing (1.1b)
(1.12) |
This is in fact true, as shown in [Pic19, Pic18] in the finite dimensional case, if is single valued and satisfies a stronger Lipschitz dependence w.r.t. (see (H1) in Appendix A).
In the framework of dissipative MPVFs, we will replace (1.12) with its relaxation à la Filippov (see e.g. [Vin10, Chapter 2] and [AF09, Chapter 10])
where is the sequential closure of the graph of in the strong-weak topology of (see [NS21] and Section 2.2 for more details; in fact, a more restrictive “directional” closure could be considered, see (6.34)) and denotes the closed convex hull of the given section .
However, even in the case of a single valued map, (1.12) is not enough to characterize the limit solution, as it has been shown by an interesting example in [Pic19, Cam+21] (see also the gradient flow of Example 6.34).
Here we follow the metric viewpoint adopted in [AGS08] for gradient flows and we will characterize the limit solutions by a suitable Evolution Variational Inequality satisfied by the squared distance function from given test measures. This approach is also strongly influenced by the Bénilan notion of integral solutions to dissipative evolutions in Banach spaces [Bén72]. The main idea is that any differentiable solution to driven by a -dissipative operator in a Hilbert space as in (1.9) satisfies
In the framework of , we replace with and the scalar product with
as in (1.7). According to this formal heuristic, we obtain the -EVI characterization of a limit curve as
(-EVI) |
As for Bénilan integral solutions, we can considerably relax the apriori smoothness assumptions on , just imposing that is continuous and (-EVI) holds in the sense of distributions in . In this way, we obtain a robust characterization, which is stable under uniform convergence and also allows for solutions taking values in the closure of the domain of . This is particularly important when involves drift terms with superlinear growth (see Example 6.32).
The crucial point of this approach relies on a general error estimate, which extends the validity of (1.11) to a general -EVI solution and therefore guarantees its uniqueness, whenever the Explicit Euler method is solvable, at least locally in time.
Combining local in time existence with suitable global confinement conditions (see e.g. Theorem 6.31) we can eventually obtain a robust theory for the generation of a -flow, i.e. a semigroup in a suitable subset of such that is the unique -EVI solution starting from and for every
as in the case of Wasserstein gradient flows of geodesically -convex functionals.
Explicit vs Implicit Euler method
In the framework of contraction semigroups generated by -dissipative operators in Hilbert or Banach spaces, a crucial role is played by the Implicit Euler scheme, which has the advantage to be unconditionally stable, and thus avoids any apriori restriction on the local bound of the operator, as we did in (1.10). In Hilbert spaces, it is well known that the solvability of the Implicit Euler scheme is equivalent to the maximality of the graph of the operator.
In the case of a Wasserstein gradient flow of a geodesically convex , every step of the Implicit Euler method (also called JKO/Minimizing Movement scheme [JKO98, AGS08]) can be solved by a variational approach: has to be selected among the solutions of
(1.13) |
Notice, however, that in this case the MPVF is defined implicitely in terms of and each step of (1.13) provides a suitable variational selection in , leading in the limit to the minimal selection principle.
In the case of more general dissipative evolutions, it is not at all clear how to solve the Implicit Euler scheme, in particular when is not concentrated on a map, and to characterize the maximal extension of (in the Hilbertian case the maximal extension of a dissipative operator is explicitly computable at least when the domain of has not empty interior, see the Theorems of Robert and Bénilan in [Qi83]). Indeed, the analogy with the Hilbertian theory does not extend to some properties which play a crucial role. In particular, a dissipative MPVF in is not locally bounded in the interior of its domain (see Example 5.2) and maximality may fail also for single-valued continuous PVFs (see Example 5.3). Even more remarkably, in the Hilbertian case a crucial equivalent characterization of dissipativity reads as
which implies that the resolvent operators (and every single step of the Implicit Euler scheme) are contractions in . On the contrary, if we assume the forward characterizations (1.6) and (1.8) of dissipativity in (with ) we cannot conclude in general that
(1.14) |
since the squared distance map , , is not convex in general (see e.g. [AGS08, Example 9.1.5]) and the fact that its right derivative at (corresponding to ) is according to (1.8) does not imply that for (corresponding to (1.14) for ).
For these reasons, we decided to approach the investigation of dissipative evolutions in by the Explicit Euler method, and we defer the study of the implicit one to a forthcoming paper.
Plan of the paper
As we already mentioned, our theory works in a general separable Hilbert space : we collect some preliminary material concerning the Wasserstein distance in Hilbert spaces and the properties of strong-weak topology for in Section 2.
In Section 3, we will study the semi-concavity properties of along general deformations induced by the exponential map and we introduce and study the pairings , . We will apply such tools to derive the precise expressions of the left and right derivatives of along absolutely continuous curves in in Section 3.2.
In Section 4, we will introduce and study the notion of -dissipative MPVF, in particular its behaviour along geodesics (Section 4.2) and its extension properties (Section 4.3). A few examples are collected in Section 5.
The last two sections contain the core of our results. Section 6 is devoted to the notion of -EVI solutions and to their properties: local uniqueness, stability and regularity in Section 6.3, global existence in Section 6.4 and barycentric characterizations in Section 6.5. Section 7 contains the main estimates for the Explicit Euler scheme: the Cauchy estimates between two discrete solutions corresponding to different step sizes in Section 7.2 and the uniform error estimates between a discrete and a -EVI solution in Section 7.3.
Acknowledgments.
G.S. and G.E.S. gratefully acknowledge the support of the Institute of Advanced Study of the Technical University of Munich. The authors thank the Department of Mathematics of the University of Pavia where this project was partially carried out. G.S. also thanks IMATI-CNR, Pavia. G.C. and G.S. have been supported by the MIUR-PRIN 2017 project Gradient flows, Optimal Transport and Metric Measure Structures. G.C. also acknowledges the partial support of the funds FARB 2016 Politecnico di Milano Prog. TDG6ATEN04.
2. Preliminaries
In this section, we introduce the main concepts and results of Optimal Transport theory that will be extensively used in the rest of the paper. We start by listing the adopted notation.
the barycenter of as in Definition 3.1; | |
the open ball with radius centered at ; | |
the set of continuous functions from to ; | |
the set of bounded continuous real valued functions defined in ; | |
the set of continuous real valued functions with compact support; | |
the space of cylindrical functions on , see Definition 2.9; | |
the sequential closure and convexification of , see Section 4.3; | |
sequential closure of convexification and extension of , see Section 4.3; | |
the right upper/lower Dini derivatives of , see (6.3); | |
the proper domain of a set-valued function as in Definition 4.1 | |
the push-forward of through the map ; | |
the set of admissible couplings between , see (2.1); | |
the set of optimal couplings between , see Definition 2.5; | |
the set of optimal couplings conditioned to , see Definition 4.8; | |
an interval of ; | |
the identity function on a set ; | |
the set of time instants s.t. belongs to , see Definition 4.8; | |
the sets of couplings as in Definition 3.8 and Theorem 3.9; | |
the -nd moment of as in Definition 2.5; | |
the -nd moment of as in (3.2); | |
the -nd moment of at as in (6.17); | |
the metric derivative of a locally absolutely continuous curve ; | |
the set of Borel probability measures on the topological space ; | |
the set of Borel probability measures with bounded support; | |
the subset of measures in with finite quadratic moments; | |
the space endowed with a weaker topology as in Definition 2.14; | |
the subset of with fixed first marginal as in (3.3); | |
, | the pseudo scalar products as in Definition 3.5; |
, | the duality pairings as in Definition 3.17; |
, | the duality pairings as in Definition 4.9; |
the limiting duality pairings as in Definition 4.11; | |
the support of ; | |
the tangent space defined in Theorem 2.10; | |
the -Wasserstein distance between and , see Definition 2.5; | |
a separable Hilbert space; | |
the tangent bundle to , usually endowed with the strong-weak topology; | |
the projection and exponential maps defined in (3.1); | |
the evaluation map defined in (3.4). |
In the present paper we will mostly deal with Borel probability measures defined in (subsets of) some separable Hilbert space endowed with the strong or a weaker topology. The convenient setting is therefore provided by Polish/Lusin and completely regular topological spaces.
Recall that a topological space is Polish (resp. Lusin) if its topology is induced by a complete and separable metric (resp. is coarser than a Polish topology). We will denote by the set of Borel probability measures on . If is Lusin, every measure is also a Radon measure, i.e. it satisfies
is completely regular if it is Hausdorff and for every closed set and point there exists a continuous function s.t. and .
Given and Lusin spaces, and a Borel function , there is a canonical way to transfer the measure from to through . This is called the push forward of through , denoted by and defined by for every Borel set in , or equivalently
for every bounded (or nonnegative) real valued Borel function
on .
A particular case occurs if , and
is the projection on the -th component, .
In this case, is usually denoted with or , and is called the -th marginal of .
This notation is particularly useful when dealing with transport plans: given and completely regular spaces and , , we define
(2.1) |
i.e. the set of probability measures on the product space having and as marginals.
On we consider the so called narrow topology which is the coarsest topology on s.t. the maps are continuous for every , the space of real valued and bounded continuous functions on . In this way a net indexed by a directed set is said to converge narrowly to , and we write in , if
We recall the well known Prokhorov’s theorem in the context of completely regular topological spaces (see [Sch73, Appendix]).
Theorem 2.1 (Prokhorov).
Let be a completely regular topological space and let be a tight subset i.e.
Then is relatively compact in w.r.t. the narrow topology.
It is then relevant to know when a given is tight. If is a Lusin completely regular topological space, then the set is tight. Another trivial criterion for tightness is the following: if is s.t. are tight for , then also is tight. We also recall the following useful proposition (see [AGS08, Remark 5.1.5]).
Proposition 2.2.
Let be a Lusin completely regular topological space and let . Then is tight if and only if there exists with compact sublevels s.t.
We recall the so-called disintegration theorem (see e.g. [AGS08, Theorem 5.3.1]).
Theorem 2.3.
Let be Lusin completely regular topological spaces, and a Borel map. Denote with . Then there exists a -a.e. uniquely determined Borel family of probability measures such that for -a.e. , and
for every bounded Borel map .
Remark 2.4.
When and , we can canonically identify the disintegration of w.r.t. with a family of probability measures . We write .
2.1. Wasserstein distance in Hilbert spaces
Let be a separable (possibly infinite dimensional) Hilbert space. We will denote by (respt. ) the Hilbert space endowed with its strong (resp. weak) topology. Notice that is a Lusin completely regular space. and share the same class of Borel sets and therefore of Borel probability measures, which we will simply denote by , using and only when we will refer to the correspondent topology. Finally, if has finite dimension then the two topologies coincide.
We now list some properties of Wasserstein spaces and we refer to [AGS08, §7] for a complete account of this matter.
Definition 2.5.
Given we define
The -Wasserstein distance between is defined as
(2.2) |
The set of elements of realizing the infimum in (2.2) is denoted with . We say that a measure is optimal if .
We will denote by the open ball centered at with radius in . The metric space enjoys many interesting properties: here we only recall that it is a complete and separable metric space and that -convergence (sometimes denoted with ) is stronger than the narrow convergence. In particular, given and , we have [AGS08, Remark 7.1.11] that
(2.3) |
Finally, we recall that sequences converging in are tight. More precisely we have the following characterization of compactness in .
Lemma 2.6 (Relative compactness in ).
A subset is relatively compact w.r.t. the -topology if and only if
-
(1)
is tight w.r.t. ,
-
(2)
is uniformly -integrable, i.e.
(2.4)
Proof.
Tightness is clearly a necessary condition; concerning (2.4) let us notice that the maps , are upper semicontinuous, are decreasing w.r.t. , and converge pointwise to for every . Therefore, if is relatively compact, they converge uniformly to thanks to Dini’s Theorem.
In order to prove that (1) and (2) are also sufficient for relative compactness, it is sufficient to check that every sequence in has a convergent subsequence. Applying Prokhorov Theorem 2.1 we can find and a convergent subsequence such that in . Since is uniformly bounded, then . Applying [AGS08, Lemma 5.1.7], we also get so that by (2.3). ∎
Definition 2.7 (Geodesics).
A curve is said to be a (constant speed) geodesic if for all we have
We also say that is a geodesic from to . We say that is a geodesically convex set if for any pair there exists a geodesic from to such that .
We recall also the following useful properties of geodesics (see [AGS08, Theorem 7.2.1, Theorem 7.2.2]).
Theorem 2.8 (Properties of geodesics).
Let and . Then defined by
(2.5) |
is a (constant speed) geodesic from to , where is given by, . Conversely, any (constant speed) geodesic from to admits the representation (2.5) for a suitable plan .
Finally, if is a geodesic connecting to , then for every there exists a unique optimal plan between and (resp. between and ) and it is concentrated on a map.
We define moreover the analogous of when we have in place of .
Definition 2.9 ().
We denote by the space of linear maps of the form for an orthonormal set of . A function belongs to the space of cylindrical functions on , , if it is of the form
where and .
We recall the following result (see [AGS08, Theorem 8.3.1, Proposition 8.4.5 and Proposition 8.4.6]) characterizing locally absolutely continuous curves in defined in a (bounded or unbounded) open interval . We use the equivalent notation for the evaluation at time of a map .
Theorem 2.10 (Wasserstein velocity field).
Let be a locally absolutely continuous curve defined in an open interval . There exists a Borel vector field and a set with such that for every
and the continuity equation
holds in the sense of distributions in . Moreover, is uniquely determined in for and
(2.6) |
We conclude this section with a useful property concerning the upper derivative of the Wasserstein distance, which in fact holds in every metric space.
Lemma 2.11.
Let , , , , and consider the constant speed geodesic defined by for every . The upper right and left Dini derivatives defined by
are respectively decreasing and increasing in .
Proof.
Take . Since is a constant speed geodesic from to , we have
then, by triangular inequality
Dividing by and passing to the limit as we obtain that the function defined by
is decreasing. It is then sufficient to observe that for
The monotonicity property of follows by the same argument. ∎
2.2. A strong-weak topology on measures in product spaces
Let us consider the case when where are separable Hilbert spaces. is naturally endowed with the product Hilbert norm and with the corresponding topology induced by the -Wasserstein distance. However, it will be extremely useful to endow with a weaker topology which is related to the strong-weak topology on , i.e. the product topology of . We follow the approach of [NS21], to which we refer for the proofs of the results presented in this section.
In order to define the topology, we consider the space of test functions such that
Notice in particular that functions in have quadratic growth. We endow with the norm
Remark 2.12.
When is finite dimensional, (2.2) is equivalent to the continuity of .
Lemma 2.13.
is a Banach space.
Definition 2.14 (Topology of , [NS21]).
We denote by the space endowed with the coarsest topology which makes the following functions continuous
It is obvious that the topology of is finer than the topology of and the latter is finer than the topology of . It is worth noticing that
so that for every net indexed by a directed set , we have
(2.7) |
The following proposition justifies the interest in the -topology.
Proposition 2.15.
-
(1)
If is a net indexed by the directed set and satisfies
-
(a)
in ,
-
(b)
,
-
(c)
,
then in . The converse property holds for sequences: if and in as then properties (a), (b), (c) hold.
-
(a)
-
(2)
For every compact set and every constant the sets
are compact and metrizable in (in particular they are sequentially compact).
It is worth noticing that the topology is strictly weaker than even when is finite dimensional. In fact, does not contain the quadratic function , so that convergence of the quadratic moment w.r.t. is not guaranteed.
3. Directional derivatives and probability measures on the tangent bundle
From now on, we will denote by a separable Hilbert space with norm and scalar product . We denote by the tangent bundle to , which is identified with the set with the induced norm and the strong-weak topology of (i.e. the product of the strong topology on the first component and the weak topology on the second one). We will denote by the projection maps and by the exponential map defined by
(3.1) |
The set is defined thanks to the identification of with and it is endowed with the narrow topology induced by the strong-weak topology in . For we define
(3.2) |
We denote by the subset of of measures for which endowed with the topology of as in Section 2.2. If we will also consider
(3.3) |
We will also deal with the product space : we will use the notation
(3.4) |
If we can consider the probability
(3.5) |
In this case we will say that is concentrated on the graph of the map . More generally, given a Borel family of probability measures satisfying
(3.6) |
we can consider the probability
(3.7) |
Conversely, every can be disintegrated by a Borel family satisfying (3.6) and (3.7). can be associated to a vector field if and only if for -a.e. . Recalling the disintegration Theorem 2.3 and Remark 2.4, we give the following definition.
Definition 3.1.
Given , the barycenter of is the function defined by
where is the disintegration of w.r.t. .
Remark 3.2.
Notice that, by the linearity of the scalar product, we get the following identity which will be useful later
(3.8) |
3.1. Directional derivatives of the Wasserstein distance and duality pairings
Our starting point is a relevant semi-concavity property of the function
(3.9) |
with . We first state an auxiliary result, whose proof is based on [AGS08, Proposition 7.3.1].
Lemma 3.3.
Let , , and let . Then there exists such that .
Proof.
Define, for every ,
Consider the probabilities and . They are constructed in such a way that there exists s.t.
where we adopted the notation and , . We conclude by taking . ∎
Proposition 3.4.
Let with and , let be the function defined by (3.9) and let be defined by
(3.10) |
-
(1)
The function is concave, i.e. it holds
(3.11) for every and every .
-
(2)
The function is concave.
-
(3)
the function is concave.
Proof.
Semi-concavity is a useful tool to guarantee the existence of one-sided partial derivatives at : for every we have (see e.g. [HL93, Ch. VI, Prop. 1.1.2]) that
(resp. ) is a concave (resp. convex) and positively -homogeneous function, i.e. a superlinear (resp. sublinear) function. They satisfy
(3.12) | ||||
(3.13) | ||||
Notice moreover that
where is the function defined in (3.10); a similar representation holds for . We introduce the following notation for , , and .
Definition 3.5.
Remark 3.6.
Notice that and , where
Moreover, using the notation
(3.14) |
we have
In particular, the properties of (in or ) and the ones of in can be easily derived by the corresponding ones of in .
Corollary 3.7.
For every and for every , , it holds
Let us now show an important equivalent characterization of the quantities we have just introduced. As usual we will denote by the projection maps of a point in (and similarly for with ).
First of all we introduce the following sets.
Definition 3.8.
For every with and we set
Analogously, for every with and in we set
In the following proposition and subsequent corollary, we provide a useful characterization of the pairings and .
Theorem 3.9.
Proof.
First, we recall that the minima in the right hand side are attained since and are compact subsets of and respectively by Lemma 2.6 and the integrands are continuous functions with quadratic growth. Thanks to Remark 3.6, we only need to prove the second equality. For every and setting , , we have
and this immediately implies
In order to prove the converse inequality, thanks to Lemma 3.3, for every we can find s.t.
Then
(3.17) |
Since is compact in , there exists a vanishing sequence and s.t. in . Moreover it holds in so that , and therefore . The convergence in yields
so that, passing to the limit in (3.17) along the sequence , we obtain
Corollary 3.10.
Let and , then
(3.18) | ||||
3.2. Right and left derivatives of the Wasserstein distance along a.c. curves
Let us now discuss the differentiability of the map along a locally absolutely continuous curve , with an open interval of and .
Theorem 3.11.
Let be a locally absolutely continuous curve and let and be as in Theorem 2.10. Then, for every and every , it holds
(3.19) | ||||
so that the map is left and right differentiable at every . In particular,
-
(1)
if and are s.t. there exists a unique optimal transport plan between and , then the map is differentiable at ;
-
(2)
there exists a subset of full Lebesgue measure such that is differentiable in and
Proof.
Let and for every we set . By Theorem 3.9, we have
Since , then thanks to Theorem 2.10 we have that the above limits coincide respectively with the limits in the statement, for all .
Claim (1) comes by the characterizations given in Theorem 3.9 and Corollary 3.10. Indeed, if there exists a unique optimal transport plan between and , then .
Claim (2) is a simple consequence of the fact that is differentiable a.e. in . ∎
Remark 3.12.
Theorem 3.13.
Let be locally absolutely continuous curves and let be the corresponding Wasserstein velocity fields satisfying (2.6) in and respectively. Then, for every , it holds
In particular, there exists a subset of full Lebesgue measure such that is differentiable in and
(3.20) | ||||
3.3. Convexity and semicontinuity of duality parings
We want now to investigate the semicontinuity and convexity properties of the functionals and .
Lemma 3.14.
Let be converging to in , and let be converging to in . Then
(3.21) |
Finally, if , , are sequences converging to in then
(3.22) |
Proof.
We just consider the proof of the first inequality (3.21); the other statements follow by similar arguments and by Remark 3.6.
We can extract a subsequence of (not relabeled) s.t. the is achieved as a limit. We have to prove that
(3.23) |
For every take with , and observe that the family is relatively compact in (since the marginals of are converging w.r.t. ) so that is relatively compact in by Proposition 2.15 since the moments are uniformly bounded by assumption. Thus, possibly passing to a further subsequence, we have that converges to some in . In particular since optimality of the marginals is preserved by narrow convergence.
Remark 3.15.
Notice that in the special case in which is a singleton, then the limit exists and it holds
Lemma 3.16.
For every the maps and (resp. and ) are convex (resp. concave) in and .
Proof.
We prove the convexity of in ; the argument of the proofs of the other statements are completely analogous.
Let , , and let , with , . We set , , For every let us select such that
It is not difficult to check that so that
3.4. Behaviour of duality pairings along geodesics
We have seen that the duality pairings and may differ when the collection of optimal plans contains more than one element. It is natural to expect a simpler behaviour along geodesics. We will introduce the following definition, where we use the notation
Definition 3.17.
For , , and , we set
which is not empty since . We set
If moreover , , , we define | ||||
Notice that, if is the disintegration of with respect to , we can consider the barycentric coupling , i.e.
so that and
If we define by the map (with a similar definition for : ) it is easy to check that
so that
(3.24) |
(3.15) and (3.18) have simpler versions in two particular cases, which will be explained in the next remark.
Remark 3.18 (Particular cases).
Suppose that , , , and is -essentially injective so that is concentrated on a Borel map , i.e. . In this case contains a unique element given by and
(3.25) |
where in the last formula we have applied the barycentric reduction (3.8). When and is the unique element of then and we obtain
Another simple case is when for some vector field as in (3.5) (i.e. its disintegration w.r.t. takes the form and ). We have
In particular we get
An important case in which the previous Remark 3.18 applies is that of geodesics in .
Lemma 3.19.
Let , be a constant speed geodesic induced by an optimal plan by the relation
If , , , then
(3.26) | ||||||||||
Proof.
The crucial fact is that is injective on and thus a bijection on its image . Indeed, take , then
thanks to the cyclical monotonicity of (see [AGS08, Remark 7.1.2]).
Then, for every , there exists a unique couple s.t. , where we refer to Remark 3.18 for the definitions of (cf. also [San15, Theorem 5.29]). Hence, in the following diagram all maps are bijections:
4. Dissipative probability vector fields: the metric viewpoint
4.1. Multivalued Probability Vector Fields and -dissipativity
Definition 4.1 (Multivalued Probability Vector Field - MPVF).
A multivalued probability vector field is a nonempty subset of with domain . Given , we define the section of as
A selection of is a subset of such that . We call a probability vector field (PVF) if is injective in , i.e. contains a unique element for every . is a vector field if for every contains a unique element concentrated on a map, i.e. .
Remark 4.2.
We can equivalently formulate Definition 4.1 by considering as a multifunction, as in the case, e.g., of the Wasserstein subdifferential of a function , see [AGS08, Ch. 10] and the next Section 5.1. According to this viewpoint, a MPVF is a set-valued map such that for all . In this way, each section is nothing but the image of through . In this case, probability vector fields correspond to single valued maps: this notion has been used in [Pic19] with the aim of describing a sort of velocity field on , and later in [Pic18] dealing with Multivalued Probability Vector Fields (called Probability Multifunctions).
Definition 4.3 (Metrically -dissipative MPVF).
A MPVF is (metrically) -dissipative, , if
(4.1) |
We say that is (metrically) -accretive, if (recall (3.14)) is -dissipative, i.e.
Remark 4.4.
Notice that (4.1) is equivalent to ask for the existence of a coupling (thus is optimal between and ) such that
Recalling the discussion of the previous section, -dissipativity has a natural metric interpretation: for every with , we have the asymptotic expansion
Remark 4.5.
Thanks to Corollary 3.7, (4.1) implies the weaker condition
(4.2) |
It is clear that the inequality of (4.2) implies the inequality of (4.1) whenever contains only one element. More generally, we will see in Corollary 4.13 that (4.2) is in fact equivalent to (4.1) when is geodesically convex (according to Definition 2.7).
As in the standard Hilbert case, -dissipativity can be reduced to dissipativity (meaning -dissipativity) by a simple transformation. Let us introduce the map
observing that for every with , , the transformed plan satisfies
(4.3) |
Similarly, if with , the plan satisfies
(4.4) |
Lemma 4.6.
Proof.
Let us first check the case of (4.2). Since if and only if , (4.3) yields
and therefore
(4.5) |
Using the corresponding identity for we obtain that is dissipative.
A similar argument, using the identity (4.4), shows the equivalence between the -dissipativity of and the dissipativity of . ∎
Let us conclude this section by showing that -dissipativity can be deduced from a Lipschitz like condition similar to the one considered in [Pic19] (see Appendix A).
Lemma 4.7.
Suppose that the MPVF satisfies
where is defined by
with as in Definition 3.8. Then is -dissipative, for
Proof.
Let , then by Theorem 3.9 and Young’s inequality, we have
4.2. Behaviour of -dissipative MPVF along geodesics
Let us now study the behaviour of a MPVF along geodesics. Recall that in the case of a dissipative map in a Hilbert space , it is quite immediate to prove that the real function
is monotone increasing. This property has a natural counterpart in the case of measures.
Definition 4.8.
Let , , . We define the sets
(4.6) | ||||
Notice that these sets depend on just through . In particular, if and is open or geodesically convex according to Definition 2.7 then .
Definition 4.9.
Let be a MPVF. Let , and let , . For every we define
Theorem 4.10.
Let us suppose that the MPVF satisfies (4.2), let , and let with . Then the following properties hold
-
(1)
for every ;
-
(2)
for every , ;
-
(3)
and are increasing respectively in and in .
-
(4)
the right (resp. left) limits of and exist at every right (resp. left) accumulation point of , and in those points the right (resp. left) limits of coincide with the right (resp. left) limits of .
-
(5)
at every interior point of where one of them is continuous.
Proof.
Throughout all the proof we set and . Thanks to Lemma 4.6 and in particular to (4.5), it is easy to check that it is sufficient to consider the dissipative case .
-
(1)
It is a direct consequence of Lemma 3.19 and the definitions of and .
-
(2)
We prove that for every and it holds
(4.7) The thesis will follow immediately passing to the over in the LHS and to the over in the RHS. It is enough to prove (4.7) in case at least one between belongs to . Let us define the map as
Observe that, since it never happens that and at the same time, the map defined as
is a bijection s.t. for every . This immediately gives that
In the same way we can deduce that
Thanks to the dissipativity of we get
-
(3)
Combining (1) and (2) we have that for every with it holds
(4.8) This implies that both and are increasing in . Observe that, again combining (1) and (2), it also holds
for every , and then is increasing in and is increasing in .
-
(4)
It is an immediate consequence of (4.8).
-
(5)
It is a straightforward consequence of (4). ∎
Thanks to the previous Theorem 4.10 the next definition is well posed.
Definition 4.11.
Let us suppose that the MPVF satisfies (4.2), let .
Corollary 4.12.
Let us keep the same notation of Theorem 4.10 and let with .
-
(1)
If , we have that
(4.9) if moreover then
(4.10) -
(2)
If , we have that
if moreover then
(4.11) -
(3)
In particular, for every , and we obtain
(4.12)
(4.12) immediately yields the following property.
Corollary 4.13.
Proposition 4.14.
Let be a MPVF satisfying (4.2), let and let . Consider the following statements
-
for every with ;
-
for every there exists s.t. ;
-
for every , ;
-
for every , ;
-
for every , ;
-
for every , .
Then the following hold
- (1)
- (2)
- (3)
- (4)
4.3. Extensions of dissipative MPVF
Let us briefly study a few simple properties about extensions of -dissipative MPVFs. The first one concerns the sequential closure in (the sequential closure may be smaller than the topological closure, but see Proposition 2.15): given , we will denote by its sequential closure defined by
Proposition 4.15.
If is a -dissipative MPVF then its sequential closure is -dissipative as well.
Proof.
If , , belong to , we can find sequences such that in as , . It is then sufficient to pass to the limit in the inequality
using the lower semicontinuity property (3.22) and the fact that convergence in yields in as . ∎
A second result concerns the convexification of the sections of . For every we set
Notice that if is bounded in then coincides with the closed convex hull of .
Proposition 4.16.
If is -dissipative, then and are -dissipative as well.
Proof.
As a last step, we want to study the properties of the extended MPVF
(4.14) | ||||
It is obvious that ; if the domain of satisfies the geometric condition (4.16), the following result shows that provides the maximal -dissipative extension of .
Proposition 4.17.
Let be a -dissipative MPVF.
-
(a)
If is -dissipative with , then . In particular .
-
(b)
and .
-
(c)
is sequentially closed and is convex for every .
-
(d)
If satisfies (4.13), then the restriction of to is -dissipative and for every
(4.15) -
(e)
If and then
- (f)
Proof.
Claim (a) is obvious since every -dissipative extension of in satisfies .
(b) Let us prove that if then . If we can find a sequence converging to in as . We can then pass to the limit in the inequalities
using the lower semicontinuity results of Lemma 3.14. We conclude since .
In order to prove that we take ; for some , , and positive coefficients , , with . Taking a convex combination of the inequalities
and using Lemma 3.16 we obtain
The proof of claim (c) follows by a similar argument.
(d) Let , , , and . The implication of Proposition 4.14 applied to and to yields
so that (4.12) yields
In order to prove (4.15) we observe that so that, for every and every , we have and , hence (4.15) is a consequence of Definition 4.11 and Theorem 4.10.
The proof of claim (f) follows by the same argument.
In the case of claim (e), we use the implication of Proposition 4.14 applied to and the implication applied to , obtaining
and then
5. Examples of -dissipative MPVFs
In this section we present significant examples of -dissipative MPVFs which are interesting for applications.
5.1. Subdifferentials of -convex functionals
Recall that a functional is -(geodesically) convex on (see [AGS08, Definition 9.1.1]) if for any in the proper domain there exists such that
where is the constant speed geodesic induced by , i.e. .
The Fréchet subdifferential of [AGS08, Definition 10.3.1] is a MPVF which can be characterized [AGS08, Theorem 10.3.6] by
According to the notation introduced in (3.14), we set
(5.1) |
and we have the following result.
Theorem 5.1.
If is a proper, lower semicontinuous and -convex functional, then is a -dissipative MPVF.
Referring to [AGS08], here we list interesting and explicit examples of -dissipative MPVFs induced by proper, lower semicontinuous and -convex functionals, focusing on the cases when
-
(1)
Potential energy. Let be a l.s.c. and -convex functional satisfying
for some constant , where is the element of minimal norm in . By [AGS08, Proposition 10.4.2] the PVF
is a -dissipative selection of for the potential energy functional
-
(2)
Interaction energy. If is an even, differentiable, and -convex function for some , whose differential has a linear growth, then, by [AGS08, Theorem 10.4.11], the PVF
is a -dissipative selection of , the opposite of the Wasserstein subdifferential of the interaction energy functional
-
(3)
Opposite Wasserstein distance. Let be fixed and consider the functional defined as
which is geodesically -convex [AGS08, Proposition 9.3.12]. Setting
the PVF
is a selection of and it is therefore -dissipative.
5.2. MPVF concentrated on the graph of a multifunction
The previous example of Section 5.1 has a natural generalization in terms of dissipative graphs in [AC84, AF09, Bré73]. We consider a (not empty) -dissipative set , i.e. satisfying
The corresponding MPVF defined as
is -dissipative as well. In fact, if with , , and then -a.e., so that
since . Taking the supremum w.r.t. we obtain which is even stronger than -dissipativity. If then contains , the set of Borel probability measures with compact support. If has also a linear growth, then it is easy to check that as well.
Despite the analogy just shown with dissipative operators in Hilbert spaces, there are important differences with the Wasserstein framework, as highlighted in the following examples. The main point here is that the dissipativity property of Definition 4.3 does not force the sections to belong to the tangent space .
Example 5.2.
Let , let be the closed unit ball, let be the (normalized) Lebesgue measure on , and let , be the anti-clockwise rotation of degrees. We define the MPVF
Observe that and is obviously unbounded at . also satisfies (4.2) with (hence it is dissipative): it is enough to check that
(5.2) |
To prove (5.2), we notice that the optimal transport plan from to is concentrated on a map and optimal maps belong to the tangent space [AGS08, Prop. 8.5.2]; by Remark 3.18 we have just to check that
that is a consequence of the Divergence Theorem on . This example is in contrast with the Hilbertian theory of dissipative operators according to which an everywhere defined dissipative operator is locally bounded (see [Bré73, Proposition 2.9]).
Example 5.3.
In the same setting of the previous example, let us define the MPVF
It is easy to check that is dissipative and Lipschitz continuous (as a map from to ). Moreover, arguing as in Example 5.2, we can show that , where is defined in (4.14). This is again in contrast with the Hilbertian theory of dissipative operators, stating that a single valued, everywhere defined, and continuous dissipative operator coincides with its maximal extension (see [Bré73, Proposition 2.4]).
5.3. Interaction field induced by a dissipative map
Let us consider the Hilbert space , , endowed with the scalar product , for every . We identify with and we denote by the -th coordinate maps. Every permutation in operates on by the obvious formula , , .
Let be a Borel -dissipative map bounded on bounded sets (this property is always true if has finite dimension) and satisfying
(5.3) |
Denoting by the components of , by the projections from to and by , the MPVF
is -dissipative as well. In fact, if , and , and , we can consider the plan , where . Considering the map we have , so that
where we used (5.3) and the invariance of with respect to permutations. The -dissipativity of then yields
A typical example when is provided by
where is a Borel, locally bounded, dissipative and antisymmetric map satisfying . We easily get
In this case
6. Solutions to Measure Differential Inclusions
6.1. Metric characterization and EVI
Let denote an arbitrary (bounded or unbounded) interval in .
The aim of this section is to study a suitable notion of solution to the following differential inclusion in the -Wasserstein space of probability measures
(6.1) |
driven by a MPVF as in Definition 4.1. In particular, we will address the usual Cauchy problem when (6.1) is supplemented by a given initial condition.
Measure Differential Inclusions have been introduced in [Pic18] extending to the multi-valued framework the theory of Measure Differential Equations developed in [Pic19]. In these papers, the author aims to describe the evolution of curves in the space of probability measures under the action of a so called probability vector field (see Definition 4.1 and Remark 4.2). However, as exploited also in [Cam+21], the definition of solution to (6.1) given in \citesPiccoli_2019,Piccoli_MDI,Camilli_MDE is too weak and it does not enjoy uniqueness property which is recovered only at the level of the semigroup through an approximation procedure.
From the Wasserstein viewpoint, the simplest way to interpret (6.1) is to ask for a locally absolutely continuous curve to satisfy
(6.2) |
where is the Wasserstein metric velocity vector associated to (see Theorem 2.10). Even in the case of a regular PVF, however, (6.2) is too strong, since there is no reason why a given should be associated to a vector field of the tangent space . Starting from (6.2), we thus introduce a weaker definition of solution to (6.1), modeled on the so-called EVI formulation for gradient flows, which will eventually suggest, as a natural formulation of (6.1), the relaxed version of (6.2) as a differential inclusion with respect to the extension of introduced in (4.14).
We start from this simple remark: whenever is -dissipative, recalling Theorem 3.11 and Remark 4.5, one easily sees that every locally absolutely continuous solution according to the above definition (6.2) also satisfies the Evolution Variational Inequality (-EVI)
(-EVI) |
for every and every , where is the functional pairing in Definition 3.5 (in fact, (-EVI) holds a.e. in ). This provides a heuristic motivation for the following definition.
Definition 6.1 (-Evolution Variational Inequality).
In Example 6.32 we will clarify the interest in imposing no more than continuity in the above definition.
Recall that the right upper and lower Dini derivatives of a function are defined for every , by
(6.3) |
Remark 6.2.
Arguing as in [MS20, Lemma A.1] and using the lower semicontinuity of the map , the distributional inequality of (-EVI) can be equivalently reformulated in terms of the right upper or lower Dini derivatives of the squared distance function and requiring the condition to hold for every :
(-EVI1) | |||||
(-EVI2) |
A further equivalent formulation [MS20, Theorem 3.3] involves the difference quotients: for every ,
(-EVI3) |
Finally, if is also locally absolutely continuous, then (-EVI1) and (-EVI2) are also equivalent to
The following Lemma provides a further insight.
Lemma 6.3.
Proof.
In order to check (6.5a) it is sufficient to combine (3.19) of Theorem 3.11 with (-EVI1). (6.5b) and (6.5c) then follow applying Proposition 4.14. Let us now prove (6.4a): let us fix and . Take and define the constant speed geodesic by , thus in particular and . Then by Lemma 2.11, for every and we have
where the second inequality comes from (-EVI1). Taking and passing to the limit as we get (6.4a). Analogously for (6.4b). ∎
We can now give an interpretation of absolutely continuous -EVI solutions in terms of differential inclusions.
Theorem 6.4.
Let be a -dissipative MPVF and let be a locally absolutely continuous curve.
- (1)
-
(2)
is a -EVI solution of (6.1) for if and only if
(6.6) - (3)
- (4)
Proof.
(1) It is sufficient to apply Theorem 3.11 and the definition of -dissipativity.
Conversely, if satisfies (6.6), , , then Theorem 3.11 and the definition of yield
Claim (3) is an immediate consequence of Lemma 6.3, Proposition 4.17(d) and Proposition 4.14.
Claim (4) is a consequence of Proposition 4.17(f) and the -dissipativity of . ∎
Proposition 6.5.
Let be a proper, lower semicontinuous and -convex functional and let be a locally absolutely continuous curve. Then
- (1)
-
(2)
if is a -EVI solution of (6.1) for the MPVF and the domain of satisfies
then is a Gradient Flow for .
Proof.
The first assertion is a consequence Theorem 6.4(1). We prove the second claim; by (6.5b) we have that for a.e. it holds
We show that for every and every
(6.7) |
To prove that, we take and . By definition of subdifferential we have
where . Dividing by , using (3.26) and passing to the infimum w.r.t. we obtain
Passing to the limit as and using the lower semicontinuity of lead to the result. Once that (6.7) is established we have that for a.e. it holds
(6.8) |
To conclude it is enough to use the lower semicontinuity of the LHS (see Lemma 3.14) and the fact that is dense in in energy: indeed we can apply [NS21, Corollary 4.5] and [AGS08, Lemma 3.1.2] to the proper, lower semicontinuous and convex functional defined as
to get the existence, for every , of a family s.t.
Of course as and, applying [AGS08, Lemma 10.3.4], we see that . However (see (4.5)) so that . We can thus write (6.8) for in place of and pass to the limit as , obtaining that, by definition of subdifferential, for a.e. . ∎
We derive a further useful a priori bound for -EVI solutions.
Proposition 6.6.
Let be a -dissipative MPVF and let . Every -EVI solution with initial datum satisfies the a priori bound
(6.9) |
where
Proof.
We conclude this section with a result showing the robustness of the notion of -EVI solution.
Proposition 6.7.
If is a sequence of -EVI solutions locally uniformly converging to as , then is a -EVI solution.
6.2. Local existence of -EVI solutions by the Explicit Euler Scheme
In order to prove the existence of a -EVI solution to (6.1), our strategy is to employ an approximation argument through an Explicit Euler scheme as it occurs for ODEs.
In the following and denote the floor and the ceiling functions respectively.
Definition 6.8 (Explicit Euler Scheme).
Let be a MPVF and suppose we are given a step size , an initial datum , a bounded interval , corresponding to the final step and a stability bound . A sequence is a -stable solution to the Explicit Euler Scheme in starting from if
(EE) |
We define the following two different interpolations of the sequence :
-
•
the affine interpolation:
(6.10) -
•
the piecewise constant interpolation:
We will call (resp. ) the (possibly empty) set of all the curves (resp. ) arising from the solution of (EE).
The affine interpolation can be trivially written as
and satisfies the uniform Lipschitz bound
(6.11) |
Notice that, since in general is not reduced to a singleton, the sets and may contain more than one element (or may be empty). Stable solutions to the Explicit Euler scheme generated by a -dissipative MPVF exhibit a nice behaviour, which is clarified by the following important result, which will be proved in Section 7 (see Proposition 7.3 and Theorems 7.4, 7.5, 7.7), with a more accurate estimate of the error constants . We stress that in the next statement solely depend on (in particular, it is independent of ).
Theorem 6.9.
Let be a -dissipative MPVF.
-
(1)
For every , every , with we have
(6.12) -
(2)
For every there exists a constant such that if and with then
-
(3)
For every there exists a constant such that if is a -EVI solution and then
(6.13) -
(4)
If is a vanishing sequence of time steps, is a sequence in converging to in and , then is uniformly converging to a limit curve which is a -EVI solution starting from .
If we assume that the Explicit Euler scheme is locally solvable, Theorem 6.9 provides a crucial tool to obtain local existence and uniqueness of -EVI solutions.
Definition 6.10 (Local and global solvability of (EE)).
Let us now state the main existence result for -EVI solutions. Given and we denote by the right upper metric derivative
Theorem 6.11 (Local existence and uniqueness).
Let be a -dissipative MPVF.
-
(a)
If the Explicit Euler Scheme is locally solvable at , then there exists and a unique -EVI solution starting from , satisfying
(6.14) If is any other -EVI solution starting from then if .
-
(b)
If the Explicit Euler Scheme is locally solvable in and
(6.15) then for every there exist a unique maximal time and a unique strict -EVI solution starting from , which satisfies (6.14) and
(6.16) Any other -EVI solution starting from coincides with in .
Proof.
(a) Let positive constants such that is not empty for every . Thanks to Theorem 6.9(2), the family satisfies the Cauchy condition in so that there exists a unique limit curve which is also Lipschitz in time, thanks to the a-priori bound (6.11). Theorem 6.9(4) shows that is a -EVI solution starting from and the estimate (6.13) of Theorem 6.9(3) shows that any other -EVI solution in an interval starting from should coincide with in .
Let us now check (6.14): we fix such that and , and we set , . The curves , belong to and , so that (6.12) yields
for . Passing to the limit as we get
Dividing by and passing to the limit as we get (6.14).
(b) Let us call the collection of -EVI solutions starting from with values in and defined in some interval , . Thanks to (6.15) and the previous claim the set is not empty.
It is also easy to check that two curves coincide in the common domain with : in fact the set contains , is closed since are continuous, and it is also open since if in then the previous claim and the fact that show that also in a right neighborhood of . Since is connected, we conclude that in .
We can thus define obtaining that there exists a unique -EVI solution starting from and defined in with values in .
If , since is Lipschitz in thanks to (6.14), we know that there exists the limit in . If we can extend to a -EVI solution with values in and defined in an interval with , which contradicts the maximality of . ∎
Recall that a set in a metric space is locally closed if every point of has a neighborhood such that . Equivalently, is the intersection of an open and a closed subset of . In particular, open or closed sets are locally closed.
Corollary 6.12.
Let be a -dissipative MPVF for which the Explicit Euler Scheme is locally solvable in . If is locally closed then for every there exists a unique maximal strict -EVI solution , , satisfying (6.16).
Let us briefly discuss the question of local solvability of the Explicit Euler scheme. The main constraints of the Explicit Euler construction relies on the a priori stability bound and in the condition for every step . This constraint is feasible if at each measure , , the set defined by
is not empty. If is open and is locally bounded, then it is easy to check that the Explicit Euler scheme is locally solvable (see Lemma 6.13). We will adopt the following notation:
(6.17) |
and we will also introduce the upper semicontinuous envelope of the function : i.e.
Lemma 6.13.
If is a -dissipative MPVF, and is bounded in a neighborhood of , i.e. there exists such that is bounded in , then the Explicit Euler scheme is locally solvable at and the locally Lipschitz solution given by Theorem 6.11(a) satisfies
(6.18) |
In particular, if is open and is locally bounded, for every there exists a unique maximal -EVI solution satisfying (6.16) and (6.18).
Proof.
More refined estimates will be discussed in the next sections. Here we will show another example, tailored to the case of measures with bounded support.
Proposition 6.14.
Let be a -dissipative MPVF such that and for every there exist , such that for every
Then for every there exists and a unique maximal strict -EVI solution satisfying (6.16).
Proof.
Arguing as in the proof of Lemma 6.13, it is easy to check that setting , we can find a discrete solution satisfying the more restrictive condition , so that the Explicit Euler scheme is locally solvable and satisfies the uniform bound
(6.19) |
Theorem 6.11 then yields the existence of a local solution, and Theorem 6.9(3) shows that the local solution satisfies the same bound (6.19) on the support, so that (6.15) holds. ∎
6.3. Stability and uniqueness
In the following theorem we prove a stability result for -EVI solutions of (6.1), as it occurs in the classical Hilbertian case scenario. We distinguish three cases: the first one assumes that the Explicit Euler scheme is locally solvable in .
Theorem 6.15 (Uniqueness and Stability).
Let be a -dissipative MPVF such that the Explicit Euler scheme is locally solvable in , and let , , be -EVI solutions to (6.1). If is strict, then
(6.20) |
In particular, if then in .
If are both strict, then
(6.21) |
Proof.
In order to prove (6.20), let us fix . Since the Explicit Euler scheme is locally solvable and , there exist such that is not empty for every . If , then (6.13) yields
for Passing to the limit as we obtain
and a further limit as yields
which implies that the map is decreasing in . Since is arbitrary, we obtain (6.20).
It is possible to prove (6.21) by a direct argument depending on the definition of -EVI solution and a geometric condition on . The simplest situation deals with absolutely continuous curves.
Theorem 6.16 (Stability for absolutely continuous solutions).
Proof.
The last situation deals with comparison between an absolutely continuous and a merely continuous -EVI solution. The argument is technically more involved and takes inspiration from the proof of [NS06, Theorem 1.1]: we refer to the Introduction of [NS06] for an explanation of the heuristic idea. Since it is also at the core of the discrete estimates of Theorem 6.9, we present it here in the easier continuous setting.
Theorem 6.17 (Refined stability).
Let and let and be -EVI solutions for the -dissipative MPVF . If at least one of the following properties hold:
-
(1)
with ;
-
(2)
satisfies (6.2),
then
Proof.
We extend in with the constant value , we denote by the Wasserstein velocity field associated to (and extended to outside ) and we define the functions by
Theorem 3.11 yields
(6.22) |
In case (1) holds, writing (6.4b) for with with , then for every we obtain
(6.23) |
On the other hand (6.5b) yields
(6.24) | ||||
Combining (6.3) and (6.24) we obtain
Since , applying Lemma C.2 we get
(6.25) |
By multiplying both inequalities (6.22) and (6.25) by we get
We fix and and we apply the Divergence theorem in [NS06, Lemma 6.15] on the two-dimensional strip as in Figure 1,
and we get
Using
then, for every conjugate coefficients (), we get
(6.26) |
Integrating (6.26) w.r.t. in the interval , we obtain
(6.27) |
Finally, we have the following inequality
(6.28) |
where we have used the notation . Taking the limit as and , we obtain the thesis. ∎
Corollary 6.18 (Local Lipschitz estimate).
6.4. Global existence and generation of -flows
We collect here a few simple results on the existence of global solutions and the generation of a -flow. A first result can be deduced from the global solvability of the Explicit Euler scheme.
Theorem 6.19 (Global existence).
Let be a -dissipative MPVF. If the Explicit Euler Scheme is globally solvable at , then there exists a unique global -EVI solution starting from .
Proof.
Let us provide a simple condition ensuring global solvability, whose proof is deferred to Section 7.
Proposition 6.20.
Let be a -dissipative MPVF such that for every there exist and such that
(6.30) |
Then the Explicit Euler scheme is globally solvable in .
Global existence of -EVI solution is also related to the existence of a -flow.
Definition 6.21.
We say that the -dissipative MPVF generates a -flow if for every there exists a unique -EVI solution starting from and the maps induce a semigroup of Lipschitz transformations of satisfying
(6.31) |
Theorem 6.22 (Generation of a -flow).
Let be a -dissipative MPVF. If at least one of the following properties is satisfied:
-
(a)
the Explicit Euler Scheme is globally solvable for every in a dense subset of ;
-
(b)
the Explicit Euler Scheme is locally solvable in and, for every in a dense subset of , there exists a strict global -EVI solution starting from ;
-
(c)
the Explicit Euler Scheme is locally solvable in and is closed;
-
(d)
for every , and, for every in a dense subset of , there exists a locally absolutely continuous strict global -EVI solution starting from ;
-
(e)
for every in a dense subset of , there exists a locally absolutely continuous solution of (6.2) starting from ,
then generates a -flow.
Proof.
(a) Let be the dense subset of for which (EE) is globally solvable. For every we define , , as the value at time of the unique -EVI solution starting from , whose existence is guaranteed by Theorem 6.19.
If , , we can find such that and are not empty for every . We can then pass to the limit in the uniform estimate (6.12) for every choice of , , obtaining (6.31) for every .
We can then extend the map to still preserving the same property. Proposition 6.7 shows that for every the continuous curve is a -EVI solution starting from .
Finally, if is any -EVI solution starting from , we can apply (6.13) to get
(6.32) |
for every , , where is a suitable constant. Passing to the limit as in (6.32) we obtain
(6.33) |
Choosing now a sequence in converging to and observing that we can choose arbitrary , we eventually get for every .
(b) Let be the dense subset of such that there exists a global strict -EVI solution starting from . By Theorem 6.15 such a solution is unique and the corresponding family of solution maps satisfy (6.31). Arguing as in the previous claim, we can extend to still preserving (6.31) and the fact that is a -EVI solution.
If is -EVI solution starting from , Theorem 6.15 shows that (6.33) holds for every . By approximation we conclude that .
(c) Corollary 6.12 shows that for every initial datum there exists a global -EVI solution. We can then apply Claim (b).
(d) Let be the dense subset of such that there exists a locally absolutely continuous strict global -EVI solution starting from . By Theorem 6.16 such a solution is the unique locally absolutely continuous solution starting from and the corresponding family of solution maps satisfy (6.31). Arguing as in the previous claim (b), we can extend to still preserving (6.31) (again thanks to Theorem 6.16) and the fact that is a -EVI solution.
If is a -EVI solution starting from and is a sequence converging to , we can apply Theorem 6.17(1) and conclude that .
(e) The proof follows by the same argument of the previous claim, eventually applying Theorem 6.17(2). ∎
By Lemma 6.13 we immediately get the following result.
Corollary 6.23.
If is locally bounded -dissipative MPVF with then for every there exists a unique global -EVI solution starting from .
We conclude this section by showing a consistency result with the Hilbertian theory, related to the example of Section 5.2.
Corollary 6.24 (Consistency with the theory of contraction semigroups in Hilbert spaces).
Let be a dissipative maximal subset generating the semigroup of nonlinear contractions [Bré73, Theorem 3.1]. Let be the dissipative MPVF
The semigroup , , is the -flow generated by in .
Proof.
Let be the set of discrete measures with . Since every is supported in , is dense in . Our thesis follows by applying Theorem 6.22(e) if we show that for every there exists a locally absolutely continuous solution of (6.2) starting from .
It can be directly checked that
satisfies the continuity equation with Wasserstein velocity vector (defined on the finite support of ) satisfying
where is the minimal selection of . It follows that
so that is a Lipschitz EVI solution for starting from . We can thus conclude observing that the map are contractions in and the curve is continuous with values in . ∎
6.5. Barycentric property
If we assume that the MPVF is a sequentially closed subset of with convex sections, we are able to provide a stronger result showing a particular property satisfied by the solutions of (6.1) (see Theorem 6.27). This is called barycentric property and it is strictly connected with the weaker definition of solution discussed in \citesPiccoli_2019, Piccoli_MDI, Camilli_MDE.
We first introduce a directional closure of along smooth cylindrical deformations. We set
and
(6.34) | ||||
Definition 6.25 (Barycentric property).
Let be a MPVF. We say that a locally absolutely continuous curve satisfies the barycentric property (resp. the relaxed barycentric property) if for a.e. there exists (resp. ) s.t.
(6.35) |
Notice that and if is sequentially closed in . Recalling Proposition 4.17(a) we also get
so that the relaxed barycentric property implies the corresponding property for the extended MPVF .
Remark 6.26.
The aim is to prove that the -EVI solution of (6.1) enjoys the barycentric property of Definition 6.25, under suitable mild conditions on . This is strictly related to the behaviour of along the family of smooth deformations induced by cylindrical functions. Let us denote by the orthogonal projection in onto the tangent space and by the barycenter of as in Definition 3.1.
Theorem 6.27.
Let be a -dissipative MPVF such that for every there exist constants such that
(6.36) |
If is a locally absolutely continuous -EVI solution of (6.1) with Wasserstein velocity field satisfying (2.6) for every in the subset of full Lebesgue measure, then
(6.37) |
In particular, satisfies the relaxed barycentric property.
If moreover and for every is a convex subset of , then satisfies (6.35).
Proof.
In the following is a fixed element of and is the constant associated to the measure
in (6.36).
For every
there exists such that
and is the unique optimal transport plan between and .
Thanks to Theorem 3.11, the map is differentiable at , moreover by employing also (6.5b), it holds
(6.38) |
We can choose a decreasing vanishing sequence , measures and such that and in . Then, by (6.13), we get with and by (6.38) and the upper semicontinuity of (see Lemma 3.14) we get
(6.39) |
Indeed, notice that, by [AGS08, Lemma 5.3.2], we have with .
By means of the identity highlighted in Remark 3.2, the expression in (6.39) can be written as follows
so that
where
(6.40) |
Applying Lemma C.1 in we obtain that . In order to obtain (6.37) it is sufficient to prove that is the -projection of the barycenter of an element of .
Notice that an element coincides with for if and only if
(6.41) |
It is easy to check that
any element can be represented as
(and thus as in (6.41)) for some
.
If we can find
a sequence such that
and
in .
Since the sequence is relatively compact in
by Proposition 2.15(2),
we can extract a (not relabeled)
subsequence converging to
a limit in , as .
By definition with
.
We can eventually pass to the limit in (6.41)
written for and
thanks to convergence, obtaining
the corresponding identity for and in the limit.
Finally, being locally absolutely continuous, it satisfies the continuity equation driven by in the sense of distributions (see Theorem 2.10), so that
Remark 6.28.
We notice that it is always possible to estimate the value of in (6.40) by .
Remark 6.29.
As a complement to the studies investigated in this section, we prove the converse characterization of Theorem 6.27 in the particular case of regular measures or regular vector fields. We refer to [AGS08, Definitions 6.2.1, 6.2.2] for the definition of , that is the space of regular measures on . When has finite dimension, is just the subset of measures in which are absolutely continuous w.r.t. the Lebesgue measure .
Theorem 6.30.
Proof.
Take and observe that, since has the relaxed barycentric property, then for a.e. (recall Theorem 3.11) there exists such that
hence solves the continuity equation , with . By Theorem 3.11, we also know that
(6.42) |
Possibly disregarding a Lebesgue negligible set, we can decompose the set in the union , where correspond to the times for which the properties (1) and (2) hold.
If and , then by [AGS08, Theorem 6.2.10], since , there exists a unique and for some map s.t. (recall [AGS08, Proposition 8.5.2]), so that
(6.43) |
where we also applied Theorem 3.9 and Remark 3.18, recalling that in this case is a singleton.
If we can select the optimal plan along which
If is the barycenter of with respect to its first marginal , recalling that (see also the proof of [AGS08, Thm. 12.4.4]) we also get
(6.44) |
Thanks to Theorem 6.30, we can apply to barycentric solutions the uniqueness and approximation results of the previous Sections. We conclude this section with a general result on the existence of a -flow for -dissipative MPVFs, which is the natural refinement of Proposition 6.14
Theorem 6.31 (Generation of -flow).
Let be a -dissipative MPVF such that and for every there exist and such that
(6.45) |
Let . If there exists such that for every
(6.46) |
then generates a -flow.
Proof.
It is enough to prove that generates a -flow. Applying Proposition 6.14 to the MPVF , we know that for every there exists a unique maximal strict -EVI solution driven by and satisfying (6.16). We argue by contradiction, and we assume that . Notice that by (6.45) satisfies (6.36), so that is a relaxed barycentric solution for . Since , we know that for some .
It is easy to check that (6.46) holds also for every . Moreover, setting , condition (6.46) yields
(6.47) |
Let be any smooth increasing function such that if and if , and let . Clearly , with if , , and We thus have for a.e.
where in the last inequality we used (6.47) and the fact that the integrand vanishes if . We get
this implies that so that the limit measure belongs to as well, leading to a contradiction with (6.16) for .
We deduce that is a global strict -EVI solution for . We can then apply Theorem 6.22(b) to . ∎
6.6. A few borderline examples
We conclude this section with a few examples which reveal the importance of some of the technical tools we developed so far. First of all we exhibit an example of dissipative MPVF generating a -flow, for which solutions starting from initial data are merely continuous (in particular the nice regularizing effect of gradient flows does not hold for general dissipative evolutions). This clarifies the interest in a definition of continuous, not necessarily absolutely continuous, solution.
Example 6.32 (Lifting of dissipative evolutions and lack of regularizing effect).
Let us consider the situation of Corollary 6.24, choosing the Hilbert space . Following [Rul96, Example 3] we can easily find a maximal linear dissipative operator whose semigroup does not provide a regularizing effect.
The domain of is and is defined as
so that there is no regularizing effect for the semigroup generated by (the graph of) : evolutions starting outside the domain stay outside the domain and do not give raise to locally Lipschitz or a.e. differentiable curves. Corollary 6.24 shows that the -flow generated by on is given by
so that there is the same lack of regularizing effect on probability measures.
In the next example we show that a constant MPVF generates a barycentric solution.
Example 6.33 (Constant PVF and barycentric evolutions).
Given , we consider the constant PVF
is dissipative: in fact, if , , , and is defined by , then
so that (3.16) yields
Applying Proposition 6.20 and Theorem 6.19 we immediately see that generates a -flow in , obtained as a limit of the Explicit Euler scheme. It is also straightforward to notice that we can apply Theorem 6.27 to so that for every the unique EVI solution satisfies the continuity equation
Since is constant, we deduce that acts as a translation with constant velocity , i.e.
so that coincides with the semigroup generated by the PVF .
We conclude this section with a -dimensional example of a curve which satisfies the barycentric property but it is not an EVI solution.
Example 6.34.
Let . It is well known (see e.g. [NS09]) that is isometric to the closed convex subset of the (essentially) increasing maps and the isometry maps each measure into the pseudo inverse of its cumulative distribution function.
It follows that for every the functional defined as
is -convex, since it satisfies where is defined as
Thus generates a gradient flow which is a semigroup of contractions in ; for every is the unique -EVI solution for the MPVF starting from (see Proposition 6.5). Since the notion of gradient flow is purely metric, the gradient flow of starting from is just the image through of the gradient flow of starting from . It is easy to check that
is the gradient flow of starting from . Note that is the geodesic from to evaluated at the rescaled time , so that must coincide with the evaluation at time of the (unique) geodesic connecting to i.e.
where .
Let us now consider the particular case , where is a fixed parameter and . It is straightforward to see that
so that
where is the Wasserstein velocity field of . On the other hand, [AGS08, Lemma 10.3.8] shows that
so that the constant curve for has the barycentric property for the MPVF but it is not a EVI solution for , being different from .
7. Explicit Euler Scheme
In this section, we collect all the main estimates concerning the Explicit Euler scheme (EE).
7.1. The Explicit Euler Scheme: preliminary estimates
Our first step is to prove simple a priori estimates and a discrete version of (-EVI) as a consequence of Proposition 3.4.
Proposition 7.1.
Every solution of (EE) satisfies
(7.1) |
(7.2) |
(IEVI) |
with possibly countable exceptions. In particular
(7.3) |
Proof.
The second inequality of (7.1) is a trivial consequence of the definition of , the first inequality is a particular case of (7.2). The estimate (7.2) is immediate if since
This implies that the metric velocity of is bounded by in and therefore is -Lipschitz.
Let us recall that for every and the function satisfies
(7.4) |
by Definition 3.5 and Proposition 3.4. In particular, the concavity yields the differentiability of with at most countable exceptions. Thus, taking any , , and so that , (7.4) yields (IEVI). (7.3) follows by integration in each interval . ∎
In the following, we prove a uniform bound on curves which is useful to prove global solvability of the Explicit Euler scheme, as stated in Proposition 6.20. We will use the following discrete Gronwall estimate: if a sequence of positive real numbers satisfies
then
(7.5) |
Proposition 7.2.
Let be a -dissipative MPVF such that for every there exist and such that
(7.6) |
then the Explicit Euler scheme is globally solvable in . More precisely, if for a given with , and we set
(7.7) |
then for every the set is not empty.
Proof.
We want to prove by induction that for every integer , (EE) has a solution up to the index satisfying the upper bound
(7.8) |
corresponding to the constants given by (7.7). For the statement is trivially satisfied. Assuming that and elements , , , are given satisfying (EE) and (7.8), we want to show that we can perform a further step of the Euler Scheme so that (EE) is solvable up to the index and .
We conclude this section by proving the stability estimate (6.12) of Theorem 6.9. We introduce the notation
Notice that for every
(7.9) |
Proposition 7.3.
Let and . If then
for every .
7.2. Error estimates for the Explicit Euler scheme
Theorem 7.4.
Let be a -dissipative MPVF. If , with , then for every there exists a constant such that
for every .
Proof.
We argue as in the proof of Theorem 6.17 with the aim to gain a convenient order of convergence. Since -dissipativity implies -dissipativity for , it is not restrictive to assume . We set . We will extensively use the a priori bounds (7.1) and (7.2); in particular,
We will also extend and for negative times by setting
(7.10) |
The proof is divided into several steps.
1. Doubling variables.
We fix a final time and two variables together with the functions
(7.11) | ||||||
observing that
(7.12) |
By Proposition 7.1, we can write (IEVI) both for and and we obtain
() | ||||
() |
Apart from possible countable exceptions, () holds for and () for . Taking , , , summing the two inequalities , setting
using (7.1) and the -dissipativity of , we obtain
in (see also [NS06, Lemma 6.15]). By multiplying both sides by , we have
(7.13) |
Using (7.12), the inequality
and the elementary inequality we get
Thus (7.13) becomes
(7.14) |
2. Penalization.
We fix any and apply the Divergence Theorem to the inequality (7.14) in the two-dimensional strip as in Figure 1 and we get
(7.15) |
3. Estimates of the RHS.
We want to estimate the integrals (say of the right hand side of (7.15) in terms of
We easily get
(7.12) yields
and
after an integration,
Performing the same computations for the third integral term at the RHS of (7.15) we end up with
Eventually, using the elementary inequalities,
and for and for , we get
We eventually get
(7.16) |
4. LHS and penalization
We want to use the first integral term in (7.15) to derive a pointwise estimate for ;
7.3. Error estimates between discrete and EVI solutions
Theorem 7.5.
Let be a -dissipative MPVF. If is a -EVI solution and , then for every there exists a constant such that
Remark 7.6.
When and we obtain the optimal error estimate
Proof.
We repeat the same argument of the previous proof, still assuming , extending as in (7.10) and setting
We use (-EVI) for with and and (IEVI) for with obtaining
Using [NS06, Lemma 6.15] we can sum the two contributions obtaining
where
Let and . Applying the Divergence Theorem in (see Figure 1) we get
(7.19) |
Using
we get for every conjugate coefficients ()
(7.20) |
Similarly to (7.12) we have
and, after an integration,
(7.21) |
Performing the same computations for the third integral term at the RHS of (7.19) we end up with
(7.22) |
Finally, since if we have , then
(7.23) |
Using (7.21), (7.22), (7.23) in (7.19), we can rewrite the bound in (7.20) as
Choosing we get
A further application of (7.18) yields
As proved in the following, the limit curve of the interpolants of the Euler Scheme defined in (6.10) is actually a -EVI solution of (6.1).
Theorem 7.7.
Let be a -dissipative MPVF and let be a vanishing sequence of time steps, let be a sequence in converging to in and let . Then is uniformly converging to a limit curve which is a -EVI solution starting from .
Proof.
Theorem 7.4 shows that is a Cauchy sequence in , so that there exists a unique limit curve as . is also -Lipschitz; moreover we observe that
(7.24) |
so that is also the uniform limit of .
Let us fix a reference measure and .
(IEVI) and the -dissipativity of yield
for a.e. . Integrating the above inequality in an interval we get
(7.25) | ||||
Notice that as , by (7.24), we have
together with the uniform bound given by
Thanks to Fatou’s Lemma and the uniform convergence given by Theorem 7.4, we can pass to the limit as in (7.25) obtaining
A further limit as yields
which provides (-EVI). ∎
Appendix A Comparison with [Pic19]
In this section, we provide a brief comparison between the assumptions we required in order to develop a strong concept of solution to (6.1) and the hypotheses assumed in [Pic19]. We remind that the relation between our solution and the weaker notion studied in [Pic19] was exploited in Section 6.5. Here, we conclude with a further remark coming from the connections between our approximating scheme proposed in (EE) and the schemes proposed in [Cam+21] and [Pic19].
We consider a finite time horizon with , the space and we deal with measures in and in , i.e. compactly supported. We also deal with single-valued probability vector fields (PVF) for simplicity, which can be considered as everywhere defined maps such that . This is indeed the framework examined in [Pic19].
We start by recalling the assumptions required in [Pic19] for a PVF .
-
(H1)
there exists a constant such that for all ,
-
(H2)
satisfies the following Lipschitz condition: there exists a constant such that for every there exists satisfying
with as in Definition 3.8.
Remark A.1.
We stress that actually in [Pic19] condition (H2) is local, meaning that is allowed to depend on the radius of a ball centered at and containing the supports of and . Thanks to assumption (H1), it is easy to show that for every final time all the discrete solutions of the Explicit Euler scheme and of the scheme of [Pic19] starting from an initial measure with support in are supported in a ball where solely depends on and . We can thus restrict the PVF to the (geodesically convex) set of measures with support in and act as does not depend on the support of the measures.
Proposition A.2.
If is a PVF satisfying (H2), then is -dissipative for , the Explicit Euler scheme is globally solvable in , and generates a -flow, whose trajectories are the limit of the Explicit Euler scheme in each finite interval .
Proof.
The -dissipativity comes from Lemma 4.7. We prove that (6.30) holds. Let and take such that
Since by assumption, there exists such that . Hence, we have
where denotes the positive part. By the trivial estimate , we conclude
Hence (6.30) and thus the global solvability of the Explicit Euler scheme in by Proposition 6.20. To conclude it is enough to apply Theorem 6.22(a) and Theorem 7.7. ∎
It is immediate to notice that the semi-discrete Lagrangian scheme proposed in [Cam+21] coincides with the Explicit Euler Scheme given in Definition 6.8. In particular, we can state the following comparison between the limit obtained by the Explicit Euler scheme (EE) (leading to the -EVI solution of (6.1)) and that of the approximating LASs scheme proposed in [Pic19] (leading to a barycentric solution to (6.1) in the sense of Definition 6.25).
Corollary A.3.
Let be a PVF satisfying (H1)-(H2), and let . Let be a sequence such that the LASs scheme of [Pic19, Definition 3.1] converges uniformly-in-time and let be the affine interpolants of the Explicit Euler Scheme defined in (6.10), with . Then and converge to the same limit curve , which is the unique -EVI solution of (6.1) in .
Proof.
By Proposition A.2, is a -dissipative MPVF s.t. for every , where is a suitable constant depending on and . Thus by Theorem 7.7, uniformly converges to a -EVI solution which is unique since generates a -flow. Since we start from a compactly supported , the semi-discrete Lagrangian scheme of [Cam+21] and our Euler Scheme actually coincide. To conclude we apply [Cam+21, Theorem 4.1] obtaining that is also the limit of the LASs scheme. ∎
We conclude that among the possibly not-unique (see [Cam+21]) barycentric solutions to (6.1) - i.e. the solutions in the sense of [Pic19]/Definition 6.25 - we are selecting only one (the -EVI solution), which turns out to be the one associated with the LASs approximating scheme.
In light of this observation, we revisit an interesting example studied in [Pic19, Section 7.1] and [Cam+21, Section 6].
Example A.4 (Splitting particle).
For every define:
so that . We define the PVF , by
By [Pic19, Proposition 7.2], satisfies assumptions (H1)-(H2) with and the LASs scheme admits a unique limit. Moreover, the solution obtained as limit of LASs, is given by
(A.1) |
By Corollary A.3, (A.1) is the (unique) -EVI solution of (6.1). In particular:
-
i)
if , i.e. the normalized Lebesgue measure restricted to , we get ;
-
ii)
if , we get .
Notice that, in case (i), since for all , i.e. , we can also apply Theorem 6.30 to conclude that is the -EVI solution of (6.1) with . Moreover, take , and consider case (i) where we denote by the initial datum and by the corresponding -EVI solution to (6.1) with , . We can apply (6.31) with and in order to give another proof that, for all , the -limit of as , that is , is a -EVI solution starting from . Thus we end up with (ii).
Dealing with case (ii), we recall that, if then also the stationary curve , for all , satisfies the barycentric property of Definition 6.25 (see [Cam+21, Example 6.1]), thus it is a solution in the sense of [Pic19]. However, is not a -EVI solution since it does not coincide with the curve given by ii). This fact can also be checked by a direct calculation as follows: we find such that
(A.2) |
where is the dissipativity constant of the PVF coming from the proof of Proposition A.2. Notice that the LHS of (A.2) is always zero since is constant. Take so that we get , with if , if . Noting that , by using the characterization in Theorem 3.9 we compute
Since , we have
and thus we obtain the desired inequality (A.2) with .
Appendix B Wasserstein differentiability along curves
In general, if is a locally absolutely continuous curve and , then the map is locally absolutely continuous and thus differentiable in a set of full measure which, in principle, depends both on and . What Theorem 3.11 shows is that, independently of , there is a full measure set , depending only on , where this map is left and right differentiable. If moreover and are such that there is a unique optimal transport plan between them, we can actually conclude that such a map is differentiable at .
We want to highlight how this result is optimal giving an example of a locally absolutely continuous curve s.t. the full measure set of differentiability points of the map depends also on . To do that it is enough to show that
where is as in Theorem 2.10 and, for s.t. , we define
Indeed this will imply that , hence the non differentiability at .
Let us consider two regular functions and s.t. for every . Let be defined as the orthogonal direction to :
Being the norm of constant in time, there exists some regular s.t. for every . Finally we define
Observe that for every . Moreover, for every and , we have
where
Hence, the above defined vector field solves the continuity equation with . Let and let us define , and the plans by
Notice that they are optimal since any plan in has the same cost, being the points the vertexes of a rhombus. Finally, we compute and :
In this way, if and we have . A possible choice for and satisfying the assumptions is
so that for every .
Appendix C Support function and Dini derivatives
We recall the following characterization of the closed convex hull of a set (i.e. the intersection of all the closed convex sets containing ) in a Banach space.
Lemma C.1.
Let be a Banach space and let be nonempty. Then if and only if
(C.1) |
Moreover if is bounded, it is enough to have (C.1) holding for every , with a dense subset of .
Proof.
The result is a direct consequence of Hahn-Banach theorem.
Concerning the last assertion, observe that the function
is Lipschitz continuous if is bounded. Hence, if (C.1) holds only for some dense, then it holds for the whole . ∎
Let us state and prove a simple lemma that allows us to pass from a differential inequality for the right upper Dini derivative to the corresponding distributional inequality (see also [MS20, Lemma A.1] and [Gál57]).
Lemma C.2.
Let be an open interval (bounded or unbounded) and let be s.t. is continuous in and is measurable and locally bounded from above in . If
then the above inequality holds also in the sense of distributions, meaning that
Proof.
Let , then there exist s.t. the support of is contained in ; since is locally bounded from above, there exists a positive constant s.t. for every . Then the function is s.t.
so that it is decreasing in and hence a function of bounded variation in . Its distributional derivative is hence a non positive measure on whose absolutely continuous part (w.r.t. the -dimensional Lebesgue measure on ) coincides a.e. with the right upper Dini derivative. Then we have
where is the singular part of . This immediately gives the thesis. ∎
References
- [AC84] Jean-Pierre Aubin and Arrigo Cellina “Differential Inclusions” Springer-Verlag Berlin Heidelberg, 1984, pp. xx+342
- [AF09] Jean-Pierre Aubin and Hélène Frankowska “Set-valued analysis” Reprint of the 1990 edition [MR1048347], Modern Birkhäuser Classics Birkhäuser Boston Inc., 2009, pp. xx+461
- [AGS08] Luigi Ambrosio, Nicola Gigli and Giuseppe Savaré “Gradient Flows In Metric Spaces and in the Space of Probability Measures”, Lectures in Mathematics. ETH Zürich Birkhäuser Basel, 2008
- [Bén72] Philippe Bénilan “Solutions intégrales d’équations d’évolution dans un espace de Banach” In C. R. Acad. Sci. Paris Sér. A-B 274, 1972, pp. A47–A50
- [BF21] Benoı̂t Bonnet and Hélène Frankowska “Differential inclusions in Wasserstein spaces: the Cauchy-Lipschitz framework” In J. Differential Equations 271, 2021, pp. 594–637 DOI: 10.1016/j.jde.2020.08.031
- [Bré73] Haim Brézis “Opérateurs maximaux monotones et semi-groupes de contractions dans les espaces de Hilbert” North-Holland Mathematics Studies, No. 5. Notas de Matemática (50) North-Holland Publishing Co., Amsterdam-London; American Elsevier Publishing Co., Inc., New York, 1973, pp. vi+183
- [Bre96] Alberto Bressan “Impulsive control systems” In Nonsmooth analysis and geometric methods in deterministic optimal control (Minneapolis, MN, 1993) 78, IMA Vol. Math. Appl. Springer, New York, 1996, pp. 1–22 DOI: 10.1007/978-1-4613-8489-2“˙1
- [Cam+21] Fabio Camilli, Giulia Cavagnari, Raul De Maio and Benedetto Piccoli “Superposition principle and schemes for measure differential equations” In Kinet. Relat. Models 14.1, 2021, pp. 89–113 DOI: 10.3934/krm.2020050
- [Cav+20] Giulia Cavagnari, Stefano Lisini, Carlo Orrieri and Giuseppe Savaré “Lagrangian, Eulerian and Kantorovich formulations of multi-agent optimal control problems: Equivalence and Gamma-convergence”, 2020 arXiv:2011.07117 [math.OC]
- [CE75] Michael G. Crandall and Lawrence C. Evans “On the relation of the operator to evolution governed by accretive operators” In Israel J. Math. 21.4, 1975, pp. 261–278 DOI: 10.1007/BF02757989
- [CL71] Michael G. Crandall and Thomas M. Liggett “Generation of semi-groups of nonlinear transformations on general Banach spaces” In Amer. J. Math. 93, 1971, pp. 265–298 DOI: 10.2307/2373376
- [CRV04] Charles Castaing, Paul Raynaud de Fitte and Michel Valadier “Young measures on topological spaces” With applications in control theory and probability theory 571, Mathematics and its Applications Kluwer Academic Publishers, Dordrecht, 2004, pp. xii+320 DOI: 10.1007/1-4020-1964-5
- [DR91] Gianni Dal Maso and Franco Rampazzo “On systems of ordinary differential equations with measures as controls” In Differential Integral Equations 4.4, 1991, pp. 739–765
- [Fil88] A.. Filippov “Differential equations with discontinuous righthand sides” Translated from the Russian 18, Mathematics and its Applications (Soviet Series) Kluwer Academic Publishers Group, Dordrecht, 1988, pp. x+304 DOI: 10.1007/978-94-015-7793-9
- [Gál57] István S. Gál “On the fundamental theorems of the calculus” In Trans. Amer. Math. Soc. 86, 1957, pp. 309–320
- [HL93] Jean-Baptiste Hiriart-Urruty and Claude Lemaréchal “Convex analysis and minimization algorithms. I” Fundamentals 305, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] Springer-Verlag, Berlin, 1993, pp. xviii+417
- [JKO98] Richard Jordan, David Kinderlehrer and Felix Otto “The variational formulation of the Fokker-Planck equation” In SIAM J. Math. Anal. 29.1, 1998, pp. 1–17 DOI: 10.1137/S0036141096303359
- [Kru70] S.. Kružkov “First order quasilinear equations with several independent variables” In Mat. Sb. (N.S.) 81 (123), 1970, pp. 228–255
- [MS20] Matteo Muratori and Giuseppe Savaré “Gradient flows and evolution variational inequalities in metric spaces. I: Structural properties” In J. Funct. Anal. 278.4, 2020, pp. 108347\bibrangessep67 DOI: 10.1016/j.jfa.2019.108347
- [NS06] Ricardo H. Nochetto and Giuseppe Savaré “Nonlinear evolution governed by accretive operators in Banach spaces: error control and applications” In Math. Models Methods Appl. Sci. 16.3, 2006, pp. 439–477 DOI: 10.1142/S0218202506001224
- [NS09] Luca Natile and Giuseppe Savaré “A Wasserstein approach to the one-dimensional sticky particle system” In SIAM J. Math. Anal. 41.4, 2009, pp. 1340–1365 DOI: 10.1137/090750809
- [NS21] Emanuele Naldi and Giuseppe Savaré “Weak topology and Opial property in Wasserstein spaces, with applications to Gradient Flows and Proximal Point Algorithms of geodesically convex functionals”, 2021 arXiv:2104.06121 [math.OC]
- [Pic18] Benedetto Piccoli “Measure differential inclusions” In 2018 IEEE Conference on Decision and Control (CDC), 2018, pp. 1323–1328
- [Pic19] Benedetto Piccoli “Measure differential equations” In Arch. Ration. Mech. Anal. 233.3, 2019, pp. 1289–1317 DOI: 10.1007/s00205-019-01379-4
- [PR14] Benedetto Piccoli and Francesco Rossi “Generalized Wasserstein distance and its application to transport equations with source” In Arch. Ration. Mech. Anal. 211.1, 2014, pp. 335–358 DOI: 10.1007/s00205-013-0669-x
- [PR19] Benedetto Piccoli and Francesco Rossi “Measure dynamics with probability vector fields and sources” In Discrete Contin. Dyn. Syst. 39.11, 2019, pp. 6207–6230 DOI: 10.3934/dcds.2019270
- [Qi83] Li Qun Qi “Uniqueness of the maximal extension of a monotone operator” In Nonlinear Anal. 7.4, 1983, pp. 325–332 DOI: 10.1016/0362-546X(83)90086-X
- [Rul96] Jim Rulla “Error analysis for implicit approximations to solutions to Cauchy problems” In SIAM J. Numer. Anal. 33.1, 1996, pp. 68–87 DOI: 10.1137/0733005
- [San15] Filippo Santambrogio “Optimal transport for applied mathematicians” Calculus of variations, PDEs, and modeling 87, Progress in Nonlinear Differential Equations and their Applications Birkhäuser/Springer, Cham, 2015, pp. xxvii+353 DOI: 10.1007/978-3-319-20828-2
- [Sch73] Laurent Schwartz “Radon measures on arbitrary topological spaces and cylindrical measures” OUP, Tata Institute Monographs on Mathematics, 1973
- [SV96] G.. Silva and R.. Vinter “Measure driven differential inclusions” In J. Math. Anal. Appl. 202.3, 1996, pp. 727–746 DOI: 10.1006/jmaa.1996.0344
- [Vil09] Cédric Villani “Optimal transport” Old and new 338, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] Springer-Verlag, Berlin, 2009, pp. xxii+973 DOI: 10.1007/978-3-540-71050-9
- [Vin10] Richard Vinter “Optimal control” Paperback reprint of the 2000 edition, Modern Birkhäuser Classics Birkhäuser Boston, Ltd., Boston, MA, 2010, pp. xx+507 DOI: 10.1007/978-0-8176-8086-2