A Metric Stability Result for the Very Strict CD Condition
Abstract
In [15] Schultz generalized the work of Rajala and Sturm [13], proving that a weak non-branching condition holds in the more general setting of very strict CD spaces. Anyway, similar to what happens for the strong CD condition, the very strict CD condition seems not to be stable with respect to the measured Gromov Hausdorff convergence (cf. [11]).
In this article I prove a stability result for the very strict CD condition, assuming some metric requirements on the converging sequence and on the limit space. The proof relies on the notions of consistent geodesic flow and consistent plan selection, which allow to treat separately the static and the dynamic part of a Wasserstein geodesic. As an application, I prove that the metric measure space equipped with a crystalline norm and with the Lebesgue measure satisfies the very strict condition.
In their pivotal works Lott, Villani [10] and Sturm [18, 19] introduced a weak notion of curvature dimension bounds, which strongly relies on the theory of Optimal Transport. They noticed that, in a Riemannian manifold, a uniform bound on the Ricci tensor is equivalent to the uniform convexity of the Boltzmann-Shannon entropy functional in the Wasserstein space. This allowed them to define a consistent notion of curvature dimension bound for metric measure spaces, that is known as CD condition. The metric measure spaces satisfying the CD condition are called CD spaces and enjoy some remarkable analytic and geometric properties.
The validity of the CD condition in a metric measure space is strongly related to the metric structure of its Wasserstein space, which is in turn strictly dependent on the metric structure of . For this reason, it is not surprising that some properties of CD spaces hold under additional metric assumptions. Among them, one of the most important is the non-branching condition, which basically prevents two different geodesic to coincide in an interval of times. Since the first works on CD spaces, it has been clear that the non-branching assumption, associated with the CD condition, could confer some nice properties to a metric measure space. For example, Sturm [18] was able to prove the tensorization property and the local-to-global property, while Gigli [8] managed to solve the Monge problem. The relation between non-branching assumption and CD condition was made even more interesting by the work of Rajala and Sturm [13]. They proved that the strong CD condition implies a weak version of the non-branching one, that they called essentially non-branching. The work of Rajala and Sturm was then generalized to the wider context of very strict CD spaces by Schultz in [15] (see also [16] and [17], where he investigates some properties of very strict CD spaces). In particular, he also underlined that every very strict CD space satisfies a weak non-branching condition, that I will call weak essentially non-branching.
One of the most important properties of CD spaces is their stability with respect to the measured Gromov Hausdorff convergence. Unfortunately this rigidity result cannot hold for the strong CD condition and, accordingly to [11], it also does not hold for the so called strict CD condition, which is (a priori) weaker than the very strict one, but stronger than the weak one. In particular, as explained in [11], it is not possible to deduce in general any type non-branching condition for a measured Gromov Hausdorff limit space. This motivates to add some analytic or metric assumption on the converging spaces, in order to achieve non-branching at the limit. In this direction the most remarkable result is provided by the theory of RCD spaces, pioneered by Ambrosio Gigli and Savaré in [4] and [5]. In fact these spaces are stable with respect to the measured Gromov Hausdorff convergence and essentially non-branching. In this article I present a stability result for very strict CD spaces, assuming metric requirements on the converging sequence and on the limit space.
In particular, the first section is dedicated to introduce the necessary preliminary notions, related both to the Optimal Transport theory and to CD conditions. In this sense, this section should be understood as a list of prerequisites and not as a complete treatment of the basic theory. For a full and precise discussions about the Optimal Transport theory I refer the reader [1], [2], [20] and [21].
In the second section I prove a purely metric stability result, which assume some strong rigidity requirements, but nevertheless can be applied to a fair variety of metric measure spaces. This result relies on the notions of consistent geodesic flow and consistent plan selection, which, as it will be clear in the following, allow me to treat separately the dynamic and the static parts of Wasserstein geodesics. The proof of this result uses an approximation strategy, and it is completely different from the arguments used for the RCD spaces theory.
The result of the second section can be applied to the metric measure space equipped with a crystalline norm and with the Lebesgue measure, this is explained in the last section. In particular I will show how a secondary variational problem can provide a consistent plan selection, associated to the Euclidean consistent geodesic flow. This will allow to conclude that these metric measure spaces are very strict CD spaces, and therefore they are weakly essentially non-branching.
1 Preliminary Notions
This first section is aimed to state all the preliminary results I will need in the following.
1.1 The Optimal Transport Problem
The original formulation of the Optimal Transport problem, due to Monge, dates back to the XVIII century, and it is the following: given two topological spaces , two probability measures , and a non-negative Borel cost function , look for the maps that minimize the following quantity
(M) |
The minimizers of the Monge problem are called optimal transport maps and in general do not necessarily exist. Therefore for the development of a general theory, it is necessary to introduce a slight generalization, due to Kantorovich. Defining the set of transport plans from to :
the Kantorovich’s formulation of the optimal transport problem asks to find minima and minimizers of
(K) |
This new problem admits minimizers under weak assumptions, in fact the following theorem holds.
Theorem 1.1 (Kantorovich).
Let and be Polish spaces and a lower semicontinuous cost function, then the minimum in the Kantorovich’s formulation (K) is attained.
The minimizers of the Kantorovich problem are called optimal transport plans and the set of all of them will be denoted by . Notice that this set obviously depends on the cost function , anyway I will usually avoid to make this dependence explicit, since many times it will be clear from the context. A transport plan is said to be induced by a map if there exists a -measurable map such that . Notice that these transport plans are exactly the ones considered in the Monge’s minimization problem (M).
Remark 1.2.
Suppose that every minimizer of the Kantorovich problem between the measures is induced by a map, and thus is a minimizer for the Monge problem. Then the Kantorovich problem between and admit a unique minimizer, which is clearly induced by a map. In fact, given two distinct transport plans , their combination is itself an optimal plan and it is not induced by a map, contradicting the assumption.
A fundamental approach in facing the Optimal Transport problem is the one of -duality, which allows to prove some very interesting and useful results. Before stating them let me introduce the notions of -cyclical monotonicity, -conjugate function and -concave function.
Definition 1.3.
A set is said to be -cyclically monotone if
for every , every permutation of and every for .
Definition 1.4.
Given a function , define its -conjugate function as
Analogously, given , define its -conjugate function as
Notice that, by definition, given a function , for every .
Definition 1.5.
A function is said to be -concave if it is the infimum of a family of -affine functions . Analogously, is said to be -concave if it is the infimum of a family of -affine functions .
The first important result of the -duality theory is the following, which summarize the strict relation that there is between optimality and -cyclical monotonicity.
Proposition 1.6.
Let and be Polish spaces and a lower semicontinuous cost function. Then every optimal transport plan such that is concentrated in a -cyclically monotone set. Moreover, if there exist two functions and such that for every , a plan is optimal only if it is concentrated on a -cyclically monotone set.
The -duality theory also allows to state the following duality result.
Proposition 1.7.
Let and be Polish spaces and a lower semicontinuous cost function. If there exist two functions and such that for every , then there exists a -concave function satisfying
Such a function is called Kantorovich potential.
Remark 1.8.
Notice that, if the cost is continuous, every -concave function is upper semicontinuous, being the infimum of continuous functions. Therefore, according to Proposition 1.7, it is possible to find an upper semicontinuous Kantorovich potential and its -conjugate function will be itself upper semicontinuous.
1.2 The Wasserstein Space and the Entropy Functional
In this section I am going to consider the Optimal Transport problem in the special case in which , is a Polish metric space and the cost function is equal to the distance squared, that is . In this context the Kantorovich’s minimization problem induces the so called Wasserstein distance on the space of probabilities with finite second order moment. Let me now give the precise definitions.
Definition 1.9.
Define the set
Definition 1.10 (Wasserstein distance).
Given two measures define their Wasserstein distance as
Proposition 1.11.
is actually a distance on and is a Polish metric space.
The convergence of measures in with respect to the Wasserstein distance can be easily characterized and this is very useful in many situation.
Proposition 1.12.
Let be a sequence of measures and let , then if and only if and
In particular, if is a compact metric space, the Wasserstein convergence is equivalent to weak convergence.
Let me now deal with the geodesic structure of , which, as the following statement shows, is heavily related to the one of the base space . This fact makes the Wasserstein space very important, and allows to prove many remarkable facts. First of all, notice that every measure induces a curve , therefore in the following I will consider measures in in order to consider curves in the Wasserstein space.
Proposition 1.13.
If is a geodesic space then is geodesic as well. More precisely, given two measures , the measure is a constant speed Wassertein geodesic connecting and if and only if it is concentrated in (that is the space of constant speed geodesics in ) and . In this case it is said that is an optimal geodesic plan between and and this will be denoted as .
Remark 1.14.
I will say that an optimal geodesic plan is induced by a map if there exists a -measurable map , such that . Following the argument explained in Remark 1.2, it is possible to conclude that, if any optimal geodesic plan between two given measures is induced by a map, then there exists a unique and it is obviously induced by a map.
Let me now introduce the entropy functional that will be the main character in defining the notion of weak curvature dimension bounds. As it will be soon clear, the most appropriate framework in which deal with the entropy functional, is the one of metric measure spaces.
Definition 1.15.
A metric measure space is a triple , where is a Polish metric space and is a non-negative and non-null Borel measure on , finite on bounded sets. Moreover, a quadruple is called pointed metric measure space if is a metric measure space and is a point in .
Remark 1.16.
In this article I assume that every metric measure space I am going to consider satisfies the following estimate
(1) |
for some (and thus all) and a suitable constant . This is essentially a technical assumption, but it is useful to ensure the lower semicontinuity of the entropy functional (see Proposition 1.19).
Let me now properly define the entropy functional.
Definition 1.17.
In a metric measure space , given a measure define the relative entropy functional with respect to as
The entropy functional relative to the reference measure will be simply denoted by .
The most important property of the entropy functional is its lower semicontinuity with respect to the different notions of convergence in spaces of probabilities. Some results in this direction are collected in this proposition.
Proposition 1.19.
If the functional is lower semicontinuous with respect to the weak topology of , while if (but (1) holds) is lower semicontinuous with respect to the Wasserstein convergence.
I conclude this subsection introducing two more spaces of probabilities, that will play an important role in the following.
Definition 1.20.
Introduce the space of probabilities absolutely continuous with respect to , with finite second order moments. Define also the space as
1.3 Curvature Dimension Bounds
In this section I introduce the notions of curvature dimension bound and CD space, presenting also the results which are the starting point of this work. Let me begin with the definition of weak and strong CD condition.
Definition 1.21.
A metric measure space is said to satisfy the (weak) condition and to be a (weak) space, if for every there exists a Wasserstein geodesic with constant speed connecting them, along which the relative entropy functional is -convex, that is
Moreover is said to satisfy a strong condition and to be a strong space if, for every , the relative entropy functional is -convex along every Wasserstein geodesic with constant speed connecting them.
The following proposition due to Villani [21] ensures the validity of CD condition in some familiar metric measure spaces and it will be fundamental in the last section.
Proposition 1.22.
Let be a norm on and let be the associated distance, then the metric measure space is a (weak) space.
The next result states the stability of CD condition with respect to the (pointed) measured Gromov Hausdorff convergence. I am not interested in giving a precise the definition of this notion of convergence, because in this article I will only deal with a different and stronger convergence for metric measure spaces. For a precise definition I refer the reader to [21], where also the next theorem is proven. Let me also point out that the measured Gromov Hausdorff convergence can be in some sense metrized by the distance, introduced by Sturm in [19]. Moreover in [9] Gigli, Mondino and Savaré showed that some different notion of convergence for pointed metric measure spaces are equivalent to the pointed measured Gromov Hausdorff convergence.
Theorem 1.23.
Let be a sequence of locally compact pointed metric measure spaces converging in the pointed measured Gromov Hausdorff sense to a locally compact pointed metric measure space . Given , if each satisfies the weak curvature dimension condition , then also satisfies .
I am now going to present the Rajala-Sturm theorem, which is the starting point of this work. In order to do this I have to preliminary introduce the notion of essentially non-branching metric measure space.
Definition 1.24.
A metric measure space is said to be essentially non-branching if for every absolutely continuous measures , every optimal geodesic plan connecting them is concentrated on a non-branching set of geodesics.
Theorem 1.25.
Every strong metric measure space is essentially non-branching.
The work of Rajala and Sturm was then generalized by Schultz [15] and applied to the context of very strict CD spaces.
Definition 1.26.
A metric measure space is called a very strict space if for every absolutely continuous measures there exists an optimal geodesic plan , so that the entropy functional satisfies the K-convexity inequality along for every , and for all bounded Borel functions with .
This CD condition is intermediate between the weak and the strong one and it easy to realize that it cannot imply the essentially non-branching property. Anyway, as pointed out by Schultz, it is possible to prove a weaker version of the non-branching condition.
Definition 1.27 (Weak Essentially Non-Branching).
A metric measure space is said to be weakly essentially non-branching if for every absolutely continuous measures , there exists an optimal geodesic plan connecting them, that is concentrated on a non-branching set of geodesics.
Theorem 1.28.
Every very strict space is weakly essentially non-branching.
Unfortunately, as the reader can easily notice, the strong CD condition is not stable with respect to the measured Gromov Hausdorff convergence. Moreover, the results in [11] suggest that it is not possible to prove a general stability result also for the very strict CD condition. These observations motivate this article, where I assume some metric requirements on the converging sequence and on the limit space, in order to prove the very strict CD condition for suitable measured Gromov Hausdorff limit spaces.
2 A Metric Stability Result
In this section I state and prove some results that allow to prove very strict condition, and thus weak essentially non-branching, for some special measured Gromov Hausdorff limit spaces. These results do not assume any analytic requirement and are purely metric, therefore they can be applied to a wide variety of metric measure spaces. The way to prove non-branching at the limit in this case is very different from the one used by Ambrosio, Gigli and Savaré in [5] and it is actually more straightforward.
First of all, let me introduce two notions which provide a nice strategy to prove the very strict CD condition, they are called consistent geodesic flow and consistent plan selection. As it will be clear in the proof of Theorem 2.4, these two concepts allow to decouple the static part from the dynamic one, taking full advantage of Proposition 1.13. This, associated with suitable assumption, makes easier to deal with restriction of optimal geodesic plans and thus to prove the very strict CD condition.
Definition 2.1.
Let be a metric space. A measurable map is called consistent geodesic flow if the following properties hold:
-
1)
for every , is a constant speed geodesic connecting and , that is with and ,
-
2)
for every and every .
A consistent geodesic flow is said to be -Lipschitz if
i.e. if it is an -Lipschitz map considered as
where .
Definition 2.2.
Let be a metric measure space and assume there exists a consistent geodesic selection for the metric space . A map is called consistent plan selection, associated to the flow if
-
1)
for every
-
2)
For every , every pair of times and every bounded Borel function with , if
where denotes the map for every , then it holds
Before going on I present the following lemma, that provides a useful equivalent characterization for condition 2 in the last definition.
Lemma 2.3.
Condition 2 in Definition 2.2 is equivalent to the combination of the following two requirements
-
2.1)
for every and every bounded Borel function with .
-
2.2)
For every and every , if
then it holds
Proof.
First of all notice that, putting in condition 2, one obtains condition 2.2. Moreover, also condition 2.1 can be deduced by condition 2, considering only the case where and . Therefore condition 2 implies the combination of 2.1 and 2.2.
On the other hand, assume that both 2.1 and 2.2 hold, then for every , every and every bounded Borel function with , if
it holds that
where I have denoted by the plan , in order to ease the notation. This last relation is exactly the requirement of condition 2. ∎
I have introduced everything I need to prove one of the crucial results of this section. It shows how the existence of a consistent geodesic flow and a consistent plan selection, satisfying suitable assumptions, ensures the validity of the very strict CD condition.
Theorem 2.4.
Given a metric measure space , assume there exist a consistent geodesic flow for and a consistent plan selection associated to . Suppose also that for every pair of measures , the -convexity inequality of the entropy is satisfied along the Wasserstein geodesic for a suitable , that is
for every . Then is a very strict space.
Proof.
Fix two measures and call . Then I need to prove that the -convexity inequality of the entropy holds along the optimal geodesic plan , for every and every bounded Borel function with . This is obviously true when at least one of its marginals at time and is not absolutely continuous, therefore I can assume that
(2) |
In particular this allows me to apply condition 2 in Definition 2.2. Now notice that, since is obviously injective, if one calls it holds
Observe now that the definition of consistent geodesic flow ensures that , thus
On the other hand it is obvious that
and similarly , so I can conclude that
At this point the fact that the entropy functional is -convex along is an easy consequence of the assumption of the theorem, associated with (2). ∎
In the remaining part of the section I show how Theorem 2.4 can be applied in order to prove the very strict CD condition for some suitable measured Gromov Hausdorff limit spaces. The first result I want to present provides sufficient conditions to ensure the existence of a consistent geodesic flow for a limit space. The reader must notice that I am considering a notion of convergence, that is much stronger than the measured Gromov Hausdorff convergence. This choice allows me to avoid some tedious technical details, but it is easy to notice that this result can be somehow extended to measured Gromov Hausdorff limit spaces. Anyway, as the next section confirms, in many easy applications this stronger hypothesis is sufficient.
Proposition 2.5.
Let be a compact metric measure space and let be a sequence of distances on (inducing the same topology) such that there exist a sequence converging to zero satisfying
in particular the sequence measured Gromov Hausdorff converges to by means of the identity map. Given , assume that for every there exists an -Lipschitz consistent geodesic flow for the metric measure space . Then there exists an -Lipschitz consistent geodesic flow for the metric measure space and, up to subsequences, converges uniformly to .
Proof.
The space is compact and thus separable, therefore there exists a countable dense set . Notice that for every and every , the sequence is contained in the compact set . Then the diagonal argument ensures that, up to taking a suitable subsequence, there exists a function
such that for every and every it holds
Now for every the function is a -Lipschitz function on , in fact for every it holds
This allows to extend to a -Lipschitz function on the whole interval , moreover, since clearly and , I can infer that . Then for every it is possible to extend the pointwise convergence of to to the interval . In fact, given , for every there exists with that allows to perform the following estimate
the claim follows from the arbitrariness of . I end up with the map
that is the pointwise limit of the -Lipschitz maps , thus also is an -Lipschitz map. Therefore it can be extended to an -Lipschitz function on the whole space , furthermore, since is closed with respect to the sup norm, I obtain
Then the equicontinuity of the maps ensures that the sequence uniformly converges to .
In order to conclude the proof I only need to show that is a consistent geodesic flow, proving property 3 of Definition 2.1. To this aim fix , and a small , subsequently take such that . Then it holds that
thesis follows from the arbitrariness of . ∎
Once Proposition 2.5 has provided a consistent geodesic flow for the limit space, the next result shows how, under suitable assumptions, it is possible to prove the very strict CD condition for the metric measure space .
Proposition 2.6.
Under the same assumptions of the last proposition, suppose that there exists a consistent plan selection on , associated to , such that for every there exists a sequence satisfying
-
1.
(up to the extraction of a subsequence),
-
2.
the -convexity of the entropy functional holds along the -Wasserstein geodesic , with respect to the distance .
Then the metric measure space is a very strict space.
Proof.
Fix a time and notice that the assumption 2 ensures that
(3) |
Now, since in compact space weak convergence and Wasserstein convergence coincide, it holds that . Then taking an optimal transport plan between and and having in mind that is -Lipschitz, it is possible to do the following estimate
Consequently, I am able to infer that
and thus that . Finally, since obviously , it is possible to pass to the limit in (3) using the lower semicontinuity of the entropy and obtain
which, associated to Theorem 2.4, allows to conclude the proof because is arbitrary. ∎
Following verbatim the proof of Proposition 2.6 it is easy to deduce the following slight generalization.
Corollary 2.7.
Under the same assumptions of Proposition 2.5, suppose that there exists a consistent plan selection on , associated to . Moreover assume that for every there exist three sequences and satisfying
-
1.
, and , ,
-
2.
(up to the extraction of a subsequence),
-
3.
the -convexity of the entropy functional holds along the -Wasserstein geodesic .
Then the metric measure space is a very strict space.
As already anticipated before, similar results can be proven for suitable measured Gromov Hausdorff limit spaces, also in the non-compact case. These generalizations require some technical assumption but their proof basically follow the proofs I have just presented. Anyway, in order to be concise, I prefer not to present the most general statements, except for the following proposition, which will be fundamental in the next section. The reader can easily notice that it can be proven following the proof of Proposition 2.6, except for two technical details that I will fix below.
Proposition 2.8.
Let be a locally compact metric measure space and let be a sequence of distances on (inducing the same topology), locally uniformly convergent to as , such that there exists a constant satisfying
(4) |
for every . Assume that there exists a map which is a Lipschitz consistent geodesic flow for and a consistent geodesic flow for every distance . Moreover, suppose that there is a consistent plan selection on , associated to , such that for every there exists a sequence , satisfying
-
1.
(up to the extraction of a subsequence),
-
2.
the -convexity of the entropy functional holds along the -Wasserstein geodesic , with respect to the distance .
Then the metric measure space is a very strict space.
Remark 2.9.
Notice that condition (4) ensures that for every .
Proof.
In order to repeat the same strategy used for Proposition 2.6 I only need to prove that and that . For the first condition, according to Proposition 1.12, it is sufficient to prove that
for every fixed . But this can be easily shown, in fact for every it holds
On the other hand, taking , condition (4) allows to use the dominated convergence theorem and deduce
Moreover for every compact set there exists a continuous function such that outside a compact set and on . Then uniformly, therefore
Since is arbitrary it is possible to conclude that
and consequently that . Having that and that , the proof of Proposition 2.6 can be repeated step by step and gives the thesis. ∎
Remark 2.10.
This section has shown how the existence of a consistent geodesic flow and a consistent plan selection associated to it, can help in proving the very strict CD condition. However, I have not stated any results (except for Proposition 2.5) that would guarantee the existence of these two objects in a metric measure space. To this aim, it would be very interesting to investigate under which assumptions on a given consistent geodesic flow (or on the metric measure space), there exists a consistent plan selection associated to . In the next section I will show how a (double) minimization procedure allows to identify a consistent plan selection in a particular metric measure space. It is possible that these arguments can also apply to a more general context.
3 Application to Crystalline Norms in
The aim of this section is to prove the very strict condition for equipped with a crystalline norm and with the Lebesgue measure, using the theory developed in the last section and in particular Proposition 2.8. Let me point out that the Optimal Transport problem in these particular metric spaces has been already studied by Ambrosio Kirchheim and Pratelli in [6]. They were able to solve the -Monge problem using a secondary variational minimization in order to suitably decompose the space in transport rays. Despite the problem I want to face and the way I will do it are different from the theory developed in [6], I will in turn use a secondary variational problem to select a suitable transport plan connecting two given measures, obtaining, as a byproduct, the existence of optimal transport map between them.
Before going on, I fix the notation I will use in this section. Given a finite set of vectors such that , introduce the associate crystalline norm, which is defined as follows
and the corresponding distance
For sake of exposition, from now on I am going to use the following equivalent formulations for the norm and the distance:
where denotes the set .
As the reader can easily guess, in this framework the choice of a consistent geodesic flow is not really problematic, in fact it is sufficient to consider the Euclidean one, that is
The rest of the chapter will be then dedicated to the choice of a suitable plan selection, associated to , satisfying the requirements of Proposition 2.8. It will be identified via a secondary variational minimization. This type of procedure turns out to be useful in many situation (see for example Chapter 2 and 3 in [14]) and in this specific case is inspired by the work of Rajala [12]. Let me now go into the details. Given two measures , consider the usual Kantorovich problem with cost , that is
calling the set of its minimizers. Consequently consider the secondary variational problem
(5) |
where I denote by the Euclidean distance, and denote by the set of minimizers, which can be easily seen to be not empty. In Theorem 3.2 I will show that, if is absolutely continuous, consists of a single element, but, in order to do this I have to preliminarily exploit the cyclical monotonicity properties of the plans in .
Proposition 3.1.
Every is concentrated in a set , such that for every it holds that
(6) |
moreover, if , then
(7) |
Proof.
Fix and notice that, since in particular , Proposition 1.6 yields that is concentrated in a set satisfying (6). Furthermore, according to Proposition 1.7 and Remark 1.8, fix an upper semicontinuous Kantorovich potential for the cost , such that also is upper semicontinuous. In particular for every , it holds
As a consequence, notice that being a minimizer of the secondary variational problem (5) is equivalent to realize the minimum of
where the cost is defined as
Observe that, since and are upper semicontinuous, the cost is lower semicontinuous. Thus Proposition 1.6 ensures that is concentrated in a set which is -cyclically monotone. Moreover, up to modify in a -null set, it is possible to assume that for every
Now take with and deduce that
On the other hand and , therefore I obtain
Finally the -cyclical monotonicity allows to conclude that
which is exactly (7). Summing up, it is easy to check that the set satisfies the requirements of Proposition 3.1. ∎
I can now go into the proof of one of the main results of this work.
Theorem 3.2.
Given two measures with absolutely continuous with respect to , there exists a unique and it is induced by a map.
Proof.
Reasoning as in Remark 1.2, it is sufficient to prove that every plan in is induced by a map. So take , applying Proposition 3.1 it is possible to find a full -measure set , satisfying the monotonicity requirements (6) and (7). Assume by contradiction that is not induced by a map, calling the disintegration with respect to the projection map , then is not a delta measure for a -positive set. Moreover, given a non-empty set , define the sets
Notice that, for every fixed , the sets constitute a partition of as varies. Consequently, I divide the proof in three steps, whose combination will allow me to conclude by contradiction.
Step 1: Given two nonempty sets such that for every and (that is ), the set
has zero -measure.
First of all, notice that if is non-empty, then for every fixed there exist and such that and in particular
and
Therefore, calling , it holds that
(8) |
Now, assume by contradiction that has positive -measure, in particular it is non-empty and there exists satisfying (8). Moreover, notice that, since is -measurable and has full measure, then is -measurable with for -almost every . In particular for small enough the set
has positive -measure, and thus it also has positive -measure. Take a Lebesgue density point of , then in a neighborhood of there exist such that for a suitable . Now, there exist and such that . Notice that for every , it holds
(9) |
while for every it is possible to perform the following estimate:
(10) |
The combination of (9) and (10) yields
(11) |
Similarly, it holds
for every , and
for every , which together show that
(12) |
Now, the inequalities (11) and (12) allow to infer that
Step 2: Given two nonempty sets such that and , the set
has zero -measure.
Call , and . Assume by contradiction that has positive -measure, then for sufficiently small the set
has positive -measure too. As a consequence
is a strictly positive measure on with . Thus there exists with and then
for every . In particular, proceeding as in the first step, it is possible to conclude that for every the set
has positive -measure, and thus it also has positive -measure. Now, I divide the proof in two cases, depending on the vector :
-
•
Case 1: for every .
Since , for every there exist such that and , . Then, given , for every it holds thatfor small enough. Thus, if is sufficiently small, follows that
and similarly
Taking the limit as , clearly and , therefore I conclude that
(13) and
(14) Now, fix sufficiently small such that
(15) As already emphasized, the set has positive Lebesgue measure, then take one of its density points . In a neighborhood of there exists , such that for a suitable , subsequently take with , and with . Notice that for every it holds
moreover (13) ensures that for every
while for every the following estimate can be performed
This last three relations show that
(16) and analogously using (14) it can be proven that
(17) On the other hand, the choice of I made (see (15)) guarantees that
which, together with (16) and (17), contradicts the condition (7) of Proposition 3.1.
-
•
Case 2: there exists such that .
Without losing generality I can assume , then it is possible to fix a sufficiently small such that, for a suitable , it holdsFix a vector . Repeating the argument used in Case 1 it is possible to find a point , such that for a suitable . Then take and with , and notice that for every it holds that
while for every I have
therefore follows that
(18) On the other hand, observe that
(19) It is then possible to conclude that
where I used both (18) and (19). This last inequality contradicts condition (6) of Proposition 3.1.
Step 3: Given a nonempty set , the set
has zero -measure.
The proof of this step is very similar to the one of Step 2, nevertheless I decided to present it anyway, but avoiding all the details which can be easily fixed following the proof of Step 2.
Assume by contradiction that has positive -measure, then for sufficiently small the set
has positive -measure too. As a consequence
is a strictly positive measure on that is not concentrated on . Thus there exists with and then
for every . In particular, proceeding as in the first step, it is possible to conclude that for every the set
has positive -measure, and thus it also has positive -measure. Now, as I did in Step 2, I divide the proof in two cases:
-
•
Case 1: for every .
First of all, fix sufficiently small such thatProceeding as in Step 2, I can find , such that for a positive, suitably small . Subsequently take with , and with . Following the proof of Step 2, it is easy to realize that
(20) and
(21) On the other hand, the choice of I made guarantees that
which, together with (20) and (21), contradicts the condition (7) of Proposition 3.1.
-
•
Case 2: there exists such that .
Without losing generality I can assume , then it is possible to fix a sufficiently small such that, for a suitable , it holds thatOnce fixed a vector , it is possible to find a point , such that for a positive, suitably small . Then take and with . Proceeding as I did in Step 2, it is easy to notice that
(22) and
(23) Then, combining (22) and (23), I can conclude that
As anticipated before, it is easy to realize that the combination of the three steps allows to conclude the proof. ∎
At this point it is clear that Theorem 3.2 provides a plan selection on , simply imposing to be equal to the only optimal transport plan in . The following proposition ensures that is a consistent plan selection.
Proposition 3.3.
The map is a consistent plan selection, associated to .
Proof.
Considering how has been defined, in order to conclude the proof, is sufficient to prove conditions 2.1 and 2.2 of Lemma 2.3. It is easy to realize that condition 2.1 is satisfied since with bounded density, for every suitable . Condition 2.2 is a little bit trickier and I am going to prove it with full details.
Assume by contradiction that, for some , is not a minimizer for the secondary variational problem (5), with absolutely continuous marginals and . Since is clearly an optimal transport plan, this means that there exists such that
Then Dudley’s gluing lemma ensures the existence of a probability measure such that
where and . Defining it is possible to perform the following estimate
Moreover, this last three integrals can be further estimated, inferring that
and similarly
Putting together this last three inequalities, it is possible to deduce that
where I used the fact that is an optimal geodesic plan. In particular this shows that . Furthermore, performing the same computation as before, one can infer that
where this last equality holds because is concentrated in Euclidean geodesic. Notice that I have found such that
this contradicts the definition of . ∎
In order to deduce the main result of this section I only have to prove the approximation property stated in Proposition 2.8, and to this aim I need to preliminary state and prove the following proposition. Let me also point out that this result can be proven using general theorems (see for example Theorem 10.27 in [21] or Theorem 1.3.1 in [7]), anyway I prefer to present a proof that uses only cyclical monotonicity arguments, similar to the ones explained previously.
Proposition 3.4.
Let be a smooth norm, such that is -convex for some . Calling the associated distance and given with , there exists a unique and it is induced by a map.
Proof.
According to Remark 1.2, it is sufficient to prove that every it is induced by a map. To this aim, fix and call the -full measure, -cyclically monotone set, provided by Proposition 1.6. Assume by contradiction that is not induced by a map, denote by the disintegration kernel with respect to the projection map , then is not a delta measure for a -positive set of . Therefore there exists a compact set with , such that is not a delta measure for every . Consequently consider
which is a positive measure on . Moreover is not concentrated on , thus there exists with and in particular for every . Now call and notice that, since is smooth there exists such that for every it holds that
for every , and
for every . Moreover, since is -convex, for every it holds that
and consequently
(24) |
for every and every . On the other hand, since , the set
has positive -measure and thus it has positive -measure. Let be the density point of , then in a neighborhood there exists such that for some . Consequently, it is possible to find and , such that
Then it holds that
where the last passage follows from (24). This last inequality contradicts the -cyclical monotonicity of , concluding the proof. ∎
Having a consistent geodesic flow and an associated plan selection, it only remains to apply Proposition 2.8 and deduce the main result. In order to do so, I introduce a sequence of distances on by requiring the following three properties:
-
•
for every n, is induced by a smooth norm , such that is -convex for some and satisfies condition (4),
-
•
converges to uniformly on compact sets,
-
•
converges to uniformly on compact sets, and for every .
It is easy to see that such a sequence exists. Now, fixed a pair of absolutely continuous measures , Proposition 3.4 ensures that for every there exists a unique transport plan between and , with respect to the cost . Let me now prove that it is possible to apply Proposition 2.8.
Proposition 3.5.
The maps and and the sequences and I introduced satisfy the assumptions of Proposition 2.8 with .
Proof.
Condition 2 is easily satisfied, in fact since is induced by a strictly convex norm the only geodesics in are the Euclidean ones. Then, because is unique and Proposition 1.22 holds, it is clear that the entropy functional is convex along , with respect to the distance . Let me now prove condition 1. Notice that for every , therefore the sequence is tight and Prokhorov theorem ensures the existence of such that, up to the extraction of a subsequence, . I am now going to prove that . Observe that is an optimal transport plan for the distance and thus
therefore for every compact set it holds
It is then possible to pass to the limit as , using the uniform convergence for the left hand side and the dominated convergence (ensured by (4)) at the right hand side, obtaining
Since this last equation holds for every compact set , it is possible to conclude that
in particular . Using once more the minimizing property of , follows that
consequently it holds that
and proceeding as before I can infer that
In particular and this concludes the proof, considering the definition of the map . ∎
Finally, the combination of this last result with Proposition 2.8 allows me to conclude the final result of this article.
Corollary 3.6.
The metric measure space is a very strict space and consequently it is weakly essentially non-branching.
Aknowlegments : This article contains part of the work I did for my master thesis, that was supervised by Luigi Ambrosio and Karl-Theodor Sturm.
References
- [1] L. Ambrosio. Lecture notes on optimal transport problem. Euro Summer School ”Mathematical aspects of evolving interfaces”, 2000.
- [2] L. Ambrosio and N. Gigli. A user’s guide to optimal transport. In Modelling and optimisation of flows on networks, pages 1–155. Springer, 2013.
- [3] L. Ambrosio, N. Gigli, A. Mondino, and T. Rajala. Riemannian Ricci curvature lower bounds in metric measure spaces with -finite measure. Transactions of the American Mathematical Society, 367(7):4661–4701, 2015.
- [4] L. Ambrosio, N. Gigli, and G. Savaré. Calculus and heat flow in metric measure spaces and applications to spaces with Ricci bounds from below. Inventiones mathematicae, 195(2):289–391, 2013.
- [5] L. Ambrosio, N. Gigli, and G. Savaré. Metric measure spaces with Riemannian Ricci curvature bounded from below. Duke Mathematical Journal, 163(7):1405–1490, 2014.
- [6] L. Ambrosio, B. Kirchheim, and A. Pratelli. Existence of optimal transport maps for crystalline norms. Duke Mathematical Journal, 125(2):207–241, 2004.
- [7] A. Figalli. Optimal Transportation and Action-Minimizing Measures. Edizioni della Normale. SNS, 2010.
- [8] N. Gigli. Optimal maps in non branching spaces with Ricci curvature bounded from below. Geometric and Functional Analysis, 22:990–999, 2011.
- [9] N. Gigli, A. Mondino, and G. Savaré. Convergence of pointed non-compact metric measure spaces and stability of Ricci curvature bounds and heat flows. Proceedings of the London Mathematical Society, 111:1071–1129, 2015.
- [10] J. Lott and C. Villani. Ricci curvature for metric-measure spaces via optimal transport. Annals of Mathematics, 169:903–991, 2009.
- [11] M. Magnabosco. Example of an highly branching CD space. arXiv preprint, 2021.
- [12] T. Rajala. Failure of the local-to-global property for spaces. Ann. Sc. Norm. Super. Pisa Cl. Sci., 15:45–68, 2016.
- [13] T. Rajala and K.-T. Sturm. Non-branching geodesics and optimal maps in strong spaces. Calculus of Variations and Partial Differential Equations, 50:831–846, 2014.
- [14] F. Santambrogio. Optimal Transport for Applied Mathematicians: Calculus of Variations, PDEs, and Modeling. Progress in Nonlinear Differential Equations and Their Applications. Springer International Publishing, 2015.
- [15] T. Schultz. Existence of optimal transport maps in very strict spaces. Calculus of Variations and Partial Differential Equations, 57, 2018.
- [16] T. Schultz. Equivalent definitions of very strict spaces. arXiv preprint, 2019.
- [17] T. Schultz. On one-dimensionality of metric measure spaces. Proc. Amer. Math. Soc., 149:383–396, 2020.
- [18] K.-T. Sturm. On the geometry of metric measure spaces. Acta Math., 196(1):65–131, 2006.
- [19] K.-T. Sturm. On the geometry of metric measure spaces. II. Acta Math., 196(1):133–177, 2006.
- [20] C. Villani. Topics in Optimal Transportation. Graduate studies in mathematics. American Mathematical Society, 2003.
- [21] C. Villani. Optimal transport – Old and new. Grundlehren der mathematischen Wissenschaften. Springer, 2008.