Home On interpolation and curvature via Wasserstein geodesics
Article
Licensed
Unlicensed Requires Authentication

On interpolation and curvature via Wasserstein geodesics

  • Martin Kell EMAIL logo
Published/Copyright: December 9, 2015

Abstract

In this article, a proof of the interpolation inequality along geodesics in p-Wasserstein spaces is given. This interpolation inequality was the main ingredient to prove the Borel–Brascamp–Lieb inequality for general Riemannian and Finsler manifolds and led Lott–Villani and Sturm to define an abstract Ricci curvature condition. Following their ideas, a similar condition can be defined and for positively curved spaces one can prove a Poincaré inequality. Using Gigli’s recently developed calculus on metric measure spaces, even a q-Laplacian comparison theorem holds on q-infinitesimal convex spaces. In the appendix, the theory of Orlicz–Wasserstein spaces is developed and necessary adjustments to prove the interpolation inequality along geodesics in those spaces are given.

MSC 2010: 58E35; 28C15

Communicated by Frank Duzaar


Funding statement: The research was funded by the IMPRS “Mathematics in the Sciences”.

A Curvature conditions using Orlicz–Wasserstein spaces

In this appendix we show that the interpolation inequality can be proven also for Orlicz–Wasserstein spaces using similar arguments. Before that we will define and investigate Orlicz–Wasserstein spaces. The main difference between a general convex and increasing function L and a homogeneous function is that there is no well-defined dual problem. However, one can use a cL-concave function and the geodesic structure to determine the interpolation potentials.

A.1 Orlicz–Wasserstein spaces

Let L:[0,)[0,) be a strictly convex increasing function with L(0)=0. Assume further there is an increasing function l:(0,)(0,) with limr0l(r)=0 and

L(r)=0rl(s)𝑑s

and hence L(s)=l(s). Define Lλ(r)=L(rλ). Note

Lλ(r)=0rlλ(s)𝑑s=0rλl(s)𝑑s

and thus

lλ(s)=1λl(sλ),lλ-1(t)=λl-1(λt).

We denote by cL the cost function given by cL(x,y)=L(d(x,y)) and as an abbreviation cλ=cLλ.

The cL-transform of a function ϕ:X¯ relative to (X,Y) is defined as

ϕcL(y):=infxXcL(x,y)-ϕ(x)

and similarly the c¯L-transform.

Definition A.1

Definition A.1 (Orlicz–Wasserstein space)

Let μi be two probability measures on M and define

wL(μ0,μ1):=inf{λ>0:infπΠ(μ0,μ1)Lλ(d(x,y))𝑑π(x,y)1}

with the convention inf=.

According to Sturm [35, Proposition 3.2], wL is a complete metric on

𝒫L(M):={μ1𝒫(M):wL(μ1,δx0)<},

where x0 is some fixed point.

Even though the following lemma is not needed, it makes many proofs below easier.

Lemma A.2

Lemma A.2 ([35, Proposition 3.1])

For every μiPL(M) there is an optimal coupling πopt of (μ0,μ1) such that

λmin=wL(μ0,μ1)Lλmin(d(x,y))𝑑πopt(x,y)=1.

Actually, the lemma shows that the whole theory of Kantorovich potentials will depend on the distance. Furthermore, the cL-concave functions are not necessarily star-shaped. Nevertheless, we will show that 𝒫L(M) is a geodesic space if and only if M is and that a similar property to the star-shapedness holds.

Proposition A.3

Let Φ be a convex increasing function with Φ(1)=1. Then

wLwΦL.
Remark

This just uses Sturm’s idea to show the same inequality for Luxemburg norm of Orlicz spaces. Compare this also to [37, Remark 6.6].

Proof.

This follows easily from Jensen’s inequality. Let μ0,μ1 be two measures and λ>0 and π be a coupling such that (ΦL)λ(d(x,y))𝑑π(x,y)1. Then since (ΦL)λ=ΦLλ,

Φ(Lλ(d(x,y))𝑑π(x,y))ΦLλ(d(x,y))𝑑π(x,y)1.

Since Φ(1)1 and Φ is increasing, we see that

Lλ(d(x,y))𝑑π(x,y)1,

which implies wL(μ0,μ1)wΦL(μ0,μ1).∎

Proposition A.4

Assume for all λ>0,

supRL(λR)L(R)<.

If μn,μPL(M) and μn converges weakly to μ, then

wL(μn,μ)0limRlim supnM\BR(x0)Lλ(d(x,x0))𝑑μn=0

for all 0<λ<λ0.

Remark

This generalizes [36, Theorem 7.12]. The other equivalences in Villani’s theorem can be proven similarly. However, only the stated condition is needed below.

Proof.

Fix some x0M. It is not difficult to see that for any λ>0 and any μ𝒫L(M),

limRM\BR(x0)Lλ(d(x,x0))𝑑μ(x)=0limRM\BR(x0)L(d(x,x0))𝑑μ(x)=0.

First assume wL(μn,μ)0 and let πn be the optimal plans with ln=wL(μn,μ) and

Lln(d(x,y))𝑑πn(x,y)=1.

For n large, for any λ>0 choose a sequence rn12 such that ln=rnλ. Then using the triangle inequality and convexity of L we get

Lλ(d(x,x0))𝑑μn(x)=Lλ(d(x,x0))𝑑πn(x,y)
rnLrnλ(d(x,y))𝑑πn(x,y)+(1-rn)L(1-rn)λ(d(y,x0))𝑑πn(x,y)
rn+(1-rn)L12λ(d(y,x0))𝑑μ(y)

since L(1-rn)λL12λ. Therefore,

limRlim supnM\BR(x0)Lλ(d(x,x0))𝑑μn(x)limRM\BR(x0)L12λ(d(x,x0))𝑑μ(x)=0.

Now assume that

limRlim supnM\BR(x0)Lλ(d(x,x0))𝑑μn(x)=0

for any 0<λ<λ0 and μn converges weakly to μ. This bound ensures that μ is in 𝒫L(M).

Take any λ>0 and an optimal coupling πn of (μn,μ) with respect to Lλ. For R>0 and AB=min{A,B} we have

d(x,y)d(x,y)R+2d(x,x0)χM\BR2(x0)(x)+2d(x0,y)χM\BR2(x0)(y)

and thus by convexity of L and L(0)=0

Lλ(d(x,y))13Lλ3(d(x,y)R)+13Lλ6(d(x,x0)χM\BR2(x0)(x))+13Lλ6(d(x0,y)χM\BR2(x0)(y)).

Thus integrating over πn we get

3Lλ(d(x,y))𝑑πn(x,y)Lλ3(d(x,y)R)𝑑πn(x,y)+M\BR2(x0)Lλ6(d(x,x0))𝑑μn(x)+M\BR2(x0)Lλ6(d(x0,y))𝑑μ(y).

We first take the lim sup with n and then R and conclude that the last two terms converges to zero by our assumption and since Lλ3(d(x,y)R) is a bounded continuous function and πn converges weakly to the trivial coupling (Id×id)*μ, the first term converges to zero as well. In particular, for nN(λ) we have

Lλ(d(x,y))𝑑πn(x,y)1

and thus

wL(μn,μ)λ.

Since λ was arbitrary, we conclude wL(μn,μ)0. ∎

Proposition A.5

Assume M is a proper metric space and Φ is convex, increasing, Φ(1)=1 and rΦ(r)0 as r. In addition, assume for all λ>0,

supRL(λR)L(R)<.

Suppose A is closed subset of PL(M) such that wL~ is bounded, where L~=ΦL. Then A is precompact in PL(M).

Remark

Compare this to [17, Theorem 6] for the case L(t)=tp, Φ(t)=tr for p1 and r>1.

Proof.

It suffices to show that each wL~-ball is compact in 𝒫L(M). So for some r>0 and μ0𝒫L~(M)𝒫L(M) let

B~:=B~r(μ0)={μ1𝒫L(M):wL~(μ0,μ1)r},

and let (μn)n be a sequence in B~. Then there are (optimal) couplings πn such that

L~r(d(x,y))𝑑πn(x,y)1.

Using the proposition above, we see

Lr(d(x,y))𝑑πn(x,y)1.

Because of the stability of optimal couplings and lower semicontinuity of the cost [37, Theorem 5.20, Lemma 4.3], we only need to show that (μn)n is weakly precompact and

limRlim supnM\BR(x0)Lλ(d(x,x0))𝑑μn=0,

i.e. it is precompact in 𝒫L(M) by the lemma above.

Since B~ is bounded with respect to wL~, we can assume that for some R>0,

wL~(μn,δx0)λ0

for some λ0>0. For cλ=λ0 and c(0,1) we have

M\BR(x0)Lλ(d(x,x0))𝑑μn(x)Lλ(R)Φ(Lλ0(R))M\BR(x0)L~λ0(d(x,x0))𝑑μn(x)Lλ0(R)Φ(Lλ0(R))Lλ0(c-1R)Lλ0(R)CLλ0(R)Φ(Lλ0(R))

for some C>0 depending only on λ0,c and L. Hence by the fact that L(R),Φ(R) as R we conclude

limRlim supnM\BR(x0)Lλ(d(x,x0))𝑑μn=0.

In order to show weak precompactness notice that L(R)1 for Rr0=r0(L) implies tightness, which is equivalent to precompactness by the classical Prokhorov theorem. Indeed, BR(x0) is compact and for r0R,

M\BR(x0)𝑑μnCLλ0(R)Φ(Lλ0(R))0

uniformly in n. ∎

Proposition A.6

Assume M is a geodesic space. Let πopt be the optimal coupling of (μ0,μ1). Then there is a Π supported on the geodesics such that

(ei)*Π=μi,i=0,1.

Furthermore, let μt=(et)*Π. Then

wL(μs,μt)=|s-t|wL(μ0,μ1).

In particular, PL(M) is a geodesic space.

Proof.

The first part follows from using the measurable selection theorem for

(x,y){γ:[0,1]Mγ is a geodesic from x to y}

similar to [19] in case of p-Wasserstein spaces.

For the second part note for λmin=wL(μ0,μ1) that

L(d(γs,γt)|s-t|λmin)𝑑Π(γ)=Lλmin(d(γ0,γ1))𝑑Π(γ)=1.

Hence

wL(μt,μs)|s-t|λmin.

So tμt is absolutely continuous in 𝒫L(M) and |μ˙t|λmin. But we also have

λmin=wL(μ0,μ1)=01|μ˙t|𝑑t.

Therefore, |μt|=λmin and

wL(μs,μt)=|st|μ˙r|dr|=|s-t|wL(μ0,μ1).

It is also possible to define a dual problem by

sup{λ>0:supϕL1(μ0){ϕ𝑑μ0+ϕcλ𝑑μ1}1}.

However, we will not go into this dual problem and directly deal with the cλ-transform whenever Kantorovich potentials are needed. Main “problem”: the restriction property does not hold for wL and many results depend on (the number) wL(μ0,μ1).

The following inequality will help to show that cL-concave functionals enjoy a similar property to star-shapedness. It will also show that the Jacobians of the interpolation measures are positive semidefinite.

Lemma A.7

If x,yM and zZt(x,y) for some t[0,1], then for all mM,

t-1L(d(m,y))Lt(d(m,z))+t-1(1-t)L(d(x,y)).

Furthermore, choosing x=m, this becomes an equality.

Remark

This extends Lemma 2.7.

Proof.

Since L is convex and increasing,

L(d(m,y))L(tt-1d(m,z)+(1-t)d(x,y))tLt(d(m,z))+(1-t)L(d(x,y)).

Dividing by t we get the inequality, and choosing x=m we see that all inequalities are actually equalities. ∎

Lemma A.8

Let η:[0,1]M be a geodesic between two distinct points x and y. For t(0,1] define

ft(m):=ct(m,ηt).

Then for some fixed t[0,1] the function h(m):=ft(m)-t-1f1(m) has a minimum at x.

Proof.

Using Lemma 2.7 above for t(0,1) we have for z=ηtZt(x,y),

-h(m)=t-1L(d(m,y))-Lt(d(m,z))t-1(1-t)L(d(x,y))=t-1L(d(x,y))-Lt(d(x,ηt))=-h(x).
Lemma A.9

Let X,Y be compact subsets of M and let t(0,1]. If ϕIcL(X,Y), then t-1ϕIct(X,Zt(X,Y)).

Proof.

For t=1 there is nothing to prove. For the rest we follow the strategy of [9, Lemma 5.1].

Set Ly(x):=L(d(x,y)), let t(0,1] and yY, and define ϕ(x):=cL(x,y)=Ly(x). We claim that the following representation holds:

t-1Ly(m)=infzZt(X,y){(Lt)z(m)+inf{xX:zZt(x,y)}t-1(1-t)Ly(x)}.

Indeed, by Lemma A.7 the left-hand side is less than or equal to the right-hand side for any zZt(X,y). Furthermore, choosing x=m we get an equality and thus showing the representation.

Now note that the claim implies that t-1ϕ is the c¯t-transform of the function

ψ(z)=-inf{xX:zZt(x,y)}t-1(1-t)Ly(x)

and therefore t-1ϕ is ct-concave relative to (X,Zt(X,y)). Since ct(X,Zt(X,y))ct(X,Zt(X,Y)), we see that each t-1Ly is in ct(X,Zt(X,Y)).

It remains to show that for an arbitrary cL-concave function ϕ and t(0,1] the function t-1ϕ is ct-concave relative to (X,Zt(X,Y)). Since ϕ=ϕcLc¯L, we have

t-1ϕ(x)=infyt-1L(d(x,y))-t-1ϕcL(y).

But each function

ψy(x)=t-1Ly(x)-t-1ϕcL(y)

is ct-concave relative to (X,Zt(X,Y)) and ϕ is proper, thus also the infimum is ct-concave relative to (X,Zt(X,Y)), i.e. t-1ϕct(X,Zt(X,Y)). ∎

A.2 Orlicz–Wasserstein spaces on Finsler manifolds

A.2.1 Technical ingredients

For simplicity, assume throughout the section that L is smooth away from 0.

For Lx:=L(d(x,)) and xy,

Lx(y)=l(d(x,y))dx(y).

Define

Lϕ:=l-1(|ϕ|)|ϕ|ϕ.

Note that for vTxM with |v|=1 and r0,

ϕ(x)=l(r)v

if and only if

Lϕ=rv.

We also use the abbreviation

λϕ=Lλϕ.

It is easy to see that under our assumptions that ϕLϕ is continuous and (as) smooth (as L) wherever ϕ(x)0.

Similar to the cp-case, we will use the abbreviation 𝒦xLdϕx (resp. 𝒦xλdϕx) for Lϕ(x) (resp. λϕ(x)). As mentioned above, this can also be seen as a Legendre transform from T*M to TM.

Lemma A.10

Lemma A.10 (Cut locus characterization)

If yx is a cut point of x, then f(z):=L(d(z,y)) satisfies

lim infv0TxMf(ξv(1))+f(ξv(-1))-2f(x)F(v)2=-,

where ξv:[-1,1]M is the geodesic with ξ˙v(0)=v.

Proof.

The proof follows in the same fashion as Lemma 3.1. We will show the necessary adjustments.

As above, let us first assume there are two distinct unit speed geodesics η,ζ:[0,d(x,y)]M from x to y and let v=ζ˙(0) and w=η˙(0). For fixed small ϵ>0 set yϵ:=η(d(x,y)-ϵ). Then yϵCut(x){x} and using the first variation formula we get for t>0,

f(ξv(-t))-f(x)L(d(ξv(-t),yϵ)+ϵ)-L(d(x,yϵ)+ϵ)
=tl(d(x,yϵ)+ϵ)gη˙(0)(v,η˙(0))+𝒪(t2)
=tl(d(x,y))gη˙(0)(v,η˙(0))+𝒪(t2).

The term 𝒪(t2) is ensured by smoothness of ξv and by the facts that xyϵ and that L(d(,)) is bounded in a neighborhood of (x,y). We also get by the Taylor formula

f(ξv(t))-f(x)=L(d(x,y)-t)-L(d(x,y))=-tl(d(x,y))+𝒪(t2).

Combining these two facts with gw(v,w)<1 (η and ξ are distinct), we get

f(ξv(-t))+f(ξv(t))-2f(x)t2gw(v,w)-1tl(d(x,y))+t-2𝒪(t2)-as t0.

For the conjugate point case, we use the same construction and notation as in the proof of Lemma 3.1. Note that

lims0L((σs))+L((σ-s))-2L((σ0))s2=(l((σ0))2s2(σs)|s=0+l((σ0))((σs)s|s=0)2)
l(d(x,y))(-2ϵgη˙(v,v)d(x,y)+ϵ2{𝒯η˙(0)(v)d(x,y)+I(V,V)})
+l(d(x,y))ϵ2F(v)2.

Using the fact that f(ξv(ϵs))L((σs)) we obtain

lim infs0f(ξv(ϵs))+f(ξv(-ϵs))-2f(x)ϵ2s2lim infs0L((σs))+L((σ-s))-2L((σ0))ϵ2s2
l(d(x,y))(-2ϵ-1gη˙(v,v)d(x,y)+𝒯(v)d(x,y)+I(V,V))+l(d(x,y))F(v)2.

Letting ϵ tend to zero completes the proof. ∎

A.2.2 The Brenier–McCann–Ohta solution

Lemma A.11

Let ϕ:MR be a cL-concave function. If ϕ is differentiable at x, then

cLϕ(x)={expx(L(-ϕ)(x))}.

Moreover, the curve η(t):=expx(tL(-ϕ)(x)) is a unique minimal geodesic from x to expx(L(-ϕ)(x)).

Remark

See also [23, Theorem 13] for the Riemannian case.

Proof.

Let ycLϕ(x) be arbitrary and define f(z):=cL(z,y)=L(d(z,y)). By the definition of cLϕ(x) we have for any vTxM,

f(expxv)ϕcL(y)+ϕ(expxv)=f(x)-ϕ(x)+ϕ(expxv)=f(x)+dϕx(v)+o(F(v)).

Let η:[0,d(x,y)]M be a minimal unit speed geodesic from x to y. Given ϵ>0, set yϵ:=η(d(x,y)-ϵ) and note that η|[0,d(x,y)-ϵ] does not cross the cut locus of x. By the first variation formula we have

f(expxv)-f(x)L(d(expxv,yϵ)+ϵ)-L(d(x,yϵ)+ϵ)=-l(d(x,yϵ)+ϵ)gη˙(0)(v,η˙(0))+o(F(v))=-l(d(x,y))x-1(η˙(0))(v)+o(F(v)).

Therefore, dϕx(v)-l(d(x,y))x-1(η˙(0))(v) for all vTxM and thus (-ϕ)=l(d(x,y))η˙(0), i.e.

L(-ϕ)=d(x,y)η˙(0).

In addition, note that η(t)=expx(tL(-ϕ)(x)/d(x,y)) which is uniquely defined. ∎

Lemma A.12

Let tμt be a geodesic between μ0 and μ1, i.e. wL(μ0,μt)=tλ. Then ϕt=t-1ϕ is a Kantorovich potential of (μ0,μt) with respect to Ltλ.

Proof.

For xycλϕ1(x) define xt:=expx(tL(-ϕ)(x)). Since xtctλϕt(x), we have for t(0,1],

xt=expx(tLλ(-ϕ)(x))
=expx(tlλ-1(tt-1|(-ϕ)|(x))|(-ϕ)|(x)(-ϕ)(x))
=expx(ltλ-1(t-1|(-ϕ)|(x))|(-ϕ)|(x)(-ϕ)(x))
=expx(ltλ-1(|(-t-1ϕ)|(x))|(-t-1ϕ)|(x)(-t-1ϕ)(x)).

Since t-1ϕ is ct-concave and t-1ϕ(x0)=0, uniqueness implies ϕt=t-1ϕ. ∎

Remark

Note that this agrees with the cases L(r)=rp/p: Assume for simplicity that wp(μ0,μ1)=1. Then ϕL=ϕcp and Lt=t-pdp/p. Hence

ϕtct(y)=inft-pdp(x,y)p-t-1ϕ(x)=t-pinfdp(x,y)p-tp-1ϕ(x)=t-p(tp-1ϕ)cp(y).

Thus up to a factor the interpolation potentials are the same (recall that tp-1ϕ gives the potential of (μ0,μt) with respect to cp).

The next results follow using exactly the same arguments as for cp.

Lemma A.13

Let μ0 and μ1 be two probability measures on M. Then there exists a cL-concave function ϕ that solves the Monge–Kantorovich problem with respect to L. Moreover, if μ0 is absolutely continuous, then the vector field L(-ϕ) is unique among such minimizers.

Remark

At this point we do not work with 𝒫L(M) directly. However, all statements make sense also for Lλ and any λ>0 and we will see later that Lemma A.9 can be used to show that the interpolation inequality in Theorem A.21 is actually an interpolation inequality with respect to the geodesic tμt in 𝒫L(M) if the function Lλ is used with λ=wL(μ0,μ1).

Theorem A.14

Let μ0 and μ1 be two probability measures on M and assume μ0 is absolutely continuous with respect to μ. Then there is a cL-concave function ϕ such that π=(Id×F)*μ0 is the unique optimal coupling of (μ0,μ1) with respect to L, where F(x):=expx(L(-ϕ)). Moreover, F is the unique optimal transport map from μ0 to μ1.

Corollary A.15

If ϕ is cL-concave and μ0 is absolutely continuous, then the map F(x):=expx(L(-ϕ)) is the unique optimal transport map from μ0 to F*μ0 with respect to the cost function cL(x,y)=L(d(x,y)).

Corollary A.16

Assume μ0 is absolutely continuous and ϕ is cλ-concave with λ=wL(μ0,(F1)*μ0), where Ft(x):=expx(tλ(-t-1ϕ)). Then Ft is the unique optimal transport map from μ0 to μt=(Ft)*μ0 with respect to Lλ and tμt is a constant speed geodesic from μ0 to μ1 in PL(M).

Remark

We will see in Lemma A.22 below that the interpolation measures are absolutely continuous if μ0 and ()t*μ0 are.

Proof.

We only need to show that wL(μs,μt)|s-t|wL(μ0,μ1).

Let π be the plan on Geo(M)={γ:[0,1]Mγ is a geodesic in M} given by μ0, the map 1 and the unique geodesic connecting μ-almost every xM to a point 1(x) (existence follows from [19, proof of Proposition 4.1], see also [37, Chapter 7]); in particular, μt=(t)*μ0. We also have

L(d(γ0,γ1λ)𝑑π(γ)=1

for λ=wL(μ0,μ1) by definition wL. Since (es,et)*π is a plan between μs and μt for s,t[0,1], we have

L(d(γs,γt)|t-s|λ)𝑑π(γ)=L(d(γ0,γ1)λ)𝑑π(γ)=1.

Therefore, wL(μs,μt)|t-s|λ. ∎

A.2.3 Almost semiconcavity of Orlicz-concave functions

The proof of almost semiconcavity of cL-concave functions follows along the lines of the proof of Theorem 3.8 by noticing that ϕs=s-1ϕ will be cs-concave instead of cL-concave, i.e. the type of concavity changes since the “distance changes”.

Theorem A.17

Let ϕ be a cL-concave function. Let Ωid be the points xM where ϕ is differentiable and dϕx=0, or equivalently cLϕ(x)={x}. Then ϕ is locally semiconcave on an open subset UM\Ωid of full measure (relative to M\Ωid). In particular, it is twice differentiable almost everywhere in U.

A.3 Proof of the interpolation inequality in the Orlicz case

Theorem A.18

Theorem A.18 (Volume distortion for L)

Let xy with yCut(x) and let η be the unique minimal geodesic from x to y. For t(0,1] define ft(z):=-Lt(d(z,η(t)). Then we have

𝔳t<(x,y)=𝐃[d(expx)Ltft(x)[d(expx)Lf1(x)]-1],
𝔳t>(x,y)=(1-t)-n𝐃[d(expx𝒦xt)d(t-1f1)x[d(d(t-1f1))x-d(dft)x]].

Remark

The statements hold equally if one takes Lλ and Ltλ, they only depend on the smoothness of L.

Proof.

Recall Theorem 3.9 and the function gt(z)=-d2(x,η(t))/2.

We have for Lη(t)=Lt(d(,η(t))) that

Lη(t)(x)=lt(d(x,η(t)))d(x,η(t))

and thus

tft(z)=lt-1(lt(d(z,η(t))))(-d(z,η(t)))=gt(z),

which implies the first equation.

For the second part note that (see calculations in the proof of Lemma A.12)

𝒦zt(d(t-1f1)z)=lt-1(t-1|f1|(z))|f1|(z)f1(z)=tl-1(|f1|(z))|f1|(z)f1(z)=tLf1(z)=z(d(tg1)z)

and hence

𝔳t>(x,y)=(1-t)-n𝐃[d(exp(d(tg1)z))]=(1-t)-n𝐃[d(exp𝒦t)d(t-1f1)xd(d(t-1f1))x].

We have d(ft)x=d(t-1f1)x. Indeed, since lt(r)=t-1l(t-1r) and d(d(,η(t))x=d(d(,y))x,

-d(ft)x=d(Lt(d(,η(t))))x=lt(d(x,η(t)))d(d(,η(t)))x=t-1l(t-1td(x,y))d(d(,y))x=-d(t-1f1)x.

Similar to [25, proof of Lemma 3.2] it suffices to show that

d(expx𝒦xt)d(ft)xd(t-1df1)x=0.

Now since ft(z)=-lt(d(z,η(t)))dη(t)(z), we get in a neighborhood U of x not containing η(t)

𝒦zt(d(ft)z)=Lt(t-1ft)(z)=lt-1(lt(d(z,η(t))))(-dη(t))(z)=z(d(gt)z)

and thus the function D:UM defined as

D(z):=expz𝒦z(d(ft)z)=expzz(d(gt)z)=η(t)

is constant in a neighborhood of x. This immediately implies dDx=0.∎

Proposition A.19

Let ϕ:MR be a cL-concave function and define F(z):=expz(L(-ϕ)(z)) at all point of differentiability of ϕ. Fix some xM such that ϕ is twice differentiable at x and dϕx0. Then the following holds:

  1. y=(x) is not a cut point of x.

  2. The function h(z)=cL(z,y)-ϕ(z) satisfies dhx=0 and

    (2hxixj(x))0

    in any local coordinate system (xi)i=1n around x.

  3. Define fy(z):=-cL(z,y) and

    dx:=d(expx𝒦xL)d(-ϕ)x[d(d(-ϕ))x-d(dfy)x]:TxMTyM,

    where the vertical part of Td(-ϕ)x(T*M) and Td(-ϕ)x(Tx*M) are identified. Then the following holds:

    sup{|u-dx(v)|:expyucLϕ(expxv),|u|=d(y,expyu)}=o(|v|)for all vTxM.

Proof.

The proof follows without any change from the proof of Proposition 3.11 but using Lemma A.10 instead and the fact that yCut(x){x} implies that fy is C at x and Lfy(x)=L(-ϕ)(x). ∎

Similarly the Jacobian equation holds.

Proposition A.20

Let μ0 and μ1 be absolutely continuous measures with respective densities f0 and f1, and let λ=wL(μ0,μ1). Also assume that there are an open set Ui with compact closure X=U¯0 and Y=U¯1 such that suppμiUi. Let ϕ be the unique cλ-concave Kantorovich potential and define F(z):=expz(λ(-ϕ)(z)). Then F is injective μ0-almost everywhere, and for μ0-almost every xM\Ωid the following hold:

  1. The function h(z)=cλ(z,(z))-ϕ(z) satisfies

    (2hxixj(x))>0

    in any local coordinate system (xi)i=0n around x.

  2. In particular, 𝐃[dx]>0 holds for the map dx:TxMT(x)M defined as above and

    limr0μ(cλϕ(Br+(x)))μ(Br+(x))=𝐃[dx]

    and

    f(x)=g((x))𝐃[dx].

Remark

Defining dx=Id for points x of differentiability of ϕ with dϕx=0 we see that the second statement above holds μ0-almost everywhere.

Proof.

Similar to Proposition 3.12, the proof follows without any change from [25, Theorem 5.2], see also [37, Chapter 11].∎

Theorem A.21

Let ϕ:MR be a cL-concave function and let xM be a point such that ϕ is twice differentiable with dϕx0. For t(0,1], define yt:=expx(t(-t-1ϕ)), ft(z):=-ct(z,yt) and Jt(x):=D[d(Ft)x], where

d(t)x:=d(expx𝒦xt)d(-t-1ϕ)x[d(d(-t-1ϕ))x-d(d(ft))x]:TxMTytM.

Then for any t(0,1),

𝐉t(x)1n(1-t)𝔳t>(x,y1)1n+t𝔳t<(x,y1)1n𝐉1(x)1n.
Remark

The proof is based on the proof of [25, Proposition 5.3] but is notationally slightly more involved than the proof of Theorem 3.13.

Proof.

Note first that

d(d(-t-1ϕ))x-d(dft)x={d(d(-t-1ϕ))x-d(d(t-1f1))x}+{d(d(t-1f1))x-d(dft)x}

and

d(ft)x=d(-t-1ϕ)x=d(t-1f1)x.

Now define τs:T*MT*M as τs(v)=s-1v and note for ϕ(x)0,

𝒦xt(t-1dϕx)=ltλ-1(|t-1ϕ(x)|)|t-1ϕ(x)|t-1ϕ(x)=tlλ-1(|ϕ(x)|)|ϕ(x)|ϕ(x)=t𝒦xL(dϕx)

and thus

𝒦xtτt(𝒦xL)-1=tIdTxM,

which implies

d(expx𝒦xt)d(-t-1ϕ)x(d(d(-t-1ϕ))x-d(d(t-1f1))x)
=d(expx𝒦xt)d(-t-1ϕ)xd(τt)d(-ϕ)x[d(d(-ϕ))x-d(d(f1))x]
=d(expx𝒦xt)d(-t-1ϕ)xd(τt)d(-ϕ)x[d(expx𝒦xL)d(-ϕ)x]-1d(1)x
=d(expx)t(-t-1ϕ)xd(𝒦xτt𝒦x-1)L(-ϕ)x[d(expx)L(-ϕ)x]-1d(1)x
=td(expx)t(-t-1ϕ)x[d(expx)L(-ϕ)x]-1d(1)x,

where we have identified Tt(-t-1ϕ)(x)(TM) with TL(-ϕ)(x)(TxM) to get the last inequality (remember that tt(-t-1ϕ)=tL(-ϕ)).

Because 𝐃 is concave, we get

𝐉t(x)1n=𝐃[d(t)x]1n
=𝐃[d(expx𝒦xt)d(-t-1ϕ)x[d(d(t-1f1))x-d(dft)x]
+d(expx𝒦xt)d(-t-1ϕ)x(d(d(-t-1ϕ))x-d(d(t-1f1))x)]1n
=𝐃[d(expx𝒦xt)d(-t-1ϕ)x(d(d(t-1f1))x-d(dft)x)
+td(expx)t(-t-1ϕ)x[d(expx)L(-ϕ)x]-1d(1)x]1n
(1-t)𝐃[(1-t)-1d(expx𝒦xt)d(-t-1ϕ)x(d(d(t-1f1))x-d(dft)x)]1n
+t𝐃[d(expx)t(-t-1ϕ)x[d(expx)L(-ϕ)x]-1d(1)x]1n
=(1-t)𝔳t>(x,y1)1n+t𝔳t<(x,y1)1n𝐉1(x)1n.

Combing this with Lemma 2.11 (see the remark after that lemma) and Lemma A.23 below we get similar to Lemma 3.14 and [25, Theorem 6.2]:

Lemma A.22

Given two absolutely continuous measures μi=ρiμ on M, let ϕ be the unique cλ-concave optimal Kantorovich potential with λ=wL(μ0,μ1). Define Ft(x):=expx(tλ(-t-1ϕ)) for t(0,1]. Then μt=ρtdμ is absolutely continuous for any t[0,1].

Proof.

By Lemma 2.11 the map t is injective μ0-almost everywhere. Let Ωid be the points xM of differentiability of ϕ with dϕx=0. Then

μt|Ωid=(t)*(μ0|Ωid)=μ0|Ωid.

By Theorem A.17 the potential ϕ is twice differentiable in a subset ΩM\Ωid of full measure. In addition, 𝐃[d(1)x]>0 for all xΩ (see Proposition A.20) and t is continuous in Ω for any t[0,1]. The map d(t)x:TxMTt(x)M is defined in Proposition A.19 as

d(t)x:=d(expx𝒦xt)d(-t-1ϕ)[d(d(-t-1ϕ))x-d(dft)x],

where ft(z):=-ctλ(z,t(x)) for t(0,1]. Also note that for xΩ,

d(d(-t-1ϕ))x-d(dft)x={d(d(-t-1ϕ))x-d(d(t-1f1))x}+{d(d(t-1f1))x-d(dft)x},

which implies 𝐃[d(t)x]>0 because 𝐃[d(1)x]>0 and the lemma below.

The result then immediately follows by [9, Claim 5.6].∎

Lemma A.23

Let yCut(x){x} and let η:[0,1]M be the unique minimal geodesic from x to y. Define

ft(z):=-ct(z,η(t)).

Then the function h(z):=t-1f1(z)-ft(z) satisfies

(2hxixj(x))0

in any local coordinate system around x.

Proof.

This follows directly from Lemma A.8. ∎

Using this interpolation inequality, one can show that a curvature dimension condition CDL(K,N) holds on any n-dimensional (n<N) Finsler manifold M with (weighted) Ricci curvature bounded from below by K. The condition CDL(K,N) is nothing but a convexity property of functionals in 𝒟𝒞N along geodesics in 𝒫L(M). Most geometric properties (Brunn–Minkowski, Bishop–Gromov, local Poincaré and doubling) also hold under such a condition. However, the lack of an “easy-to-understand” dual theory makes it difficult to prove statements involving (weak) upper gradients.

Corollary A.24

Any n-dimensional Finsler manifold with N-Ricci curvature bounded from below by K and N>n satisfies the very strong CDL(K,N) condition for any strictly convex, increasing functional L:[0,)[0,) which is smooth away from zero.

Acknowledgements

The author wants to thank Professor Jürgen Jost and the MPI MiS for providing a stimulating research environment. Thanks also to the anonymous referee for helpful remarks which improved readability of the manuscript.

References

[1] Ambrosio L., Gigli N., Mondino A. and Rajala T., Riemannian Ricci curvature lower bounds in metric measure spaces with σ-finite measure, Trans. Amer. Math. Soc. 367 (2015), 4661–4701. 10.1090/S0002-9947-2015-06111-XSearch in Google Scholar

[2] Ambrosio L., Gigli N. and Savaré G., Gradient Flows in Metric Spaces and in the Space of Probability Measures, 2nd ed., Lectures in Mathematics ETH Zürich, Birkhäuser, Basel, 2008. Search in Google Scholar

[3] Ambrosio L., Gigli N. and Savaré G., Density of Lipschitz functions and equivalence of weak gradients in metric measure spaces, Rev. Mat. Iberoam. 29 (2013), no. 3, 969–996. 10.4171/RMI/746Search in Google Scholar

[4] Ambrosio L., Gigli N. and Savaré G., Calculus and heat flow in metric measure spaces and applications to spaces with Ricci bounds from below, Invent. Math. 195 (2014), no. 2, 289–391. 10.1007/s00222-013-0456-1Search in Google Scholar

[5] Ambrosio L., Gigli N. and Savaré G., Metric measure spaces with Riemannian Ricci curvature bounded from below, Duke Math. J. 163 (2014), no. 7, 1405–1490. 10.1215/00127094-2681605Search in Google Scholar

[6] Bacher K. and Sturm K.-T., Localization and tensorization properties of the curvature-dimension condition for metric measure spaces, J. Funct. Anal. 259 (2010), no. 1, 28–56. 10.1016/j.jfa.2010.03.024Search in Google Scholar

[7] Bao D., Chern S. and Shen Z., An Introduction to Riemann–Finsler Geometry, Springer, Berlin, 2000. 10.1007/978-1-4612-1268-3Search in Google Scholar

[8] Brenier Y., Polar factorization and monotone rearrangement of vector-valued functions, Comm. Pure Appl. Math. 44 (1991), no. 4, 375–417. 10.1002/cpa.3160440402Search in Google Scholar

[9] Cordero-Erausquin D., McCann R. J. and Schmuckenschläger M., A Riemannian interpolation inequality à la Borell, Brascamp and Lieb, Invent. Math. 146 (2001), no. 2, 219–257. 10.1007/s002220100160Search in Google Scholar

[10] Erbar M., Kuwada K. and Sturm K.-T., On the equivalence of the entropic curvature-dimension condition and Bochner’s inequality on metric measure spaces, Invent. Math. 201 (2015), no. 3, 993–1071. 10.1007/s00222-014-0563-7Search in Google Scholar

[11] Garofalo N. and Mondino A., Li–Yau and Harnack type inequalities in metric measure spaces, Nonlinear Anal. 95 (2014), 721–734. 10.1016/j.na.2013.10.002Search in Google Scholar

[12] Gigli N., The splitting theorem in non-smooth context, preprint 2013, http://arxiv.org/abs/1302.5555. Search in Google Scholar

[13] Gigli N., On the differential structure of metric measure spaces and applications, Mem. Amer. Math. Soc. 236 (2015), No. 1113. 10.1090/memo/1113Search in Google Scholar

[14] Gigli N., Rajala T. and Sturm K.-T., Optimal maps and exponentiation on finite dimensional spaces with Ricci curvature bounded from below, preprint 2013, http://arxiv.org/abs/1305.4849. 10.1007/s12220-015-9654-ySearch in Google Scholar

[15] Gozlan N., Roberto C. and Samson P.-M., Hamilton Jacobi equations on metric spaces and transport-entropy inequalities, Rev. Mat. Iberoam. 30 (2014), no. 1, 133–163. 10.4171/RMI/772Search in Google Scholar

[16] Hua B., Kell M. and Xia C., Harmonic functions on metric measure spaces, preprint 2013, http://arxiv.org/abs/1308.3607. Search in Google Scholar

[17] Kell M., Stability of the global attractor under Markov–Wasserstein noise, preprint 2011, http://arxiv.org/abs/1103.3401. Search in Google Scholar

[18] Kell M., q-heat flow and the gradient flow of the Renyi entropy in the p-Wasserstein space, preprint 2014, http://arxiv.org/abs/1401.0840. 10.1016/j.jfa.2016.06.016Search in Google Scholar

[19] Lisini S., Characterization of absolutely continuous curves in Wasserstein spaces, Calc. Var. Partial Differential Equations 28 (2006), no. 1, 85–120. 10.1007/s00526-006-0032-2Search in Google Scholar

[20] Lott J. and Villani C., Weak curvature conditions and functional inequalities, J. Funct. Anal. 245 (2007), no. 1, 311–333. 10.1016/j.jfa.2006.10.018Search in Google Scholar

[21] Lott J. and Villani C., Ricci curvature for metric-measure spaces via optimal transport, Ann. of Math. (2) 169 (2009), no. 3, 903–991. 10.4007/annals.2009.169.903Search in Google Scholar

[22] McCann R. J., A convexity principle for interacting gases, Adv. Math. 128 (1997), no. 1, 153–179. 10.1006/aima.1997.1634Search in Google Scholar

[23] McCann R. J., Polar factorization of maps on Riemannian manifolds, Geom. Funct. Anal. 11 (2001), no. 3, 589–608. 10.1007/PL00001679Search in Google Scholar

[24] Ohta S., Uniform convexity and smoothness, and their applications in Finsler geometry, Math. Ann. 343 (2008), no. 3, 669–699. 10.1007/s00208-008-0286-4Search in Google Scholar

[25] Ohta S., Finsler interpolation inequalities, Calc. Var. Partial Differential Equations 36 (2009), no. 2, 211–249. 10.1007/s00526-009-0227-4Search in Google Scholar

[26] Ohta S., On the curvature and heat flow on Hamiltonian systems, Anal. Geom. Metr. Spaces 2 (2014), no. 1, 81–114. 10.2478/agms-2014-0003Search in Google Scholar

[27] Ohta S., (K,N)-convexity and the curvature-dimension condition for negative N, J. Geom. Anal. 1 (2015), 10.1007/s12220-015-9619-1. 10.1007/s12220-015-9619-1Search in Google Scholar

[28] Ohta S., Splitting theorems for Finsler manifolds of nonnegative Ricci curvature, J. Reine Angew. Math. 700 (2015), 155–174. 10.1515/crelle-2013-0011Search in Google Scholar

[29] Rajala T., Local Poincaré inequalities from stable curvature conditions on metric spaces, Calc. Var. Partial Differential Equations 44 (2011), no. 3–4, 477–494. 10.1007/s00526-011-0442-7Search in Google Scholar

[30] Rajala T., Interpolated measures with bounded density in metric spaces satisfying the curvature-dimension conditions of Sturm, J. Funct. Anal. 263 (2012), no. 4, 896–924. 10.1016/j.jfa.2012.05.006Search in Google Scholar

[31] Shen Z., Volume comparison and its applications in Riemann–Finsler geometry, Adv. Differential Equations 128 (1997), no. 2, 306–328. 10.1006/aima.1997.1630Search in Google Scholar

[32] Shen Z., Lectures on Finsler Geometry, World Scientific, Singapore, 2001. 10.1142/4619Search in Google Scholar

[33] Sturm K.-T., On the geometry of metric measure spaces, Acta Math. 196 (2006), no. 1, 65–131. 10.1007/s11511-006-0002-8Search in Google Scholar

[34] Sturm K.-T., On the geometry of metric measure spaces. II, Acta Math. 196 (2006), no. 1, 133–177. 10.1007/s11511-006-0003-7Search in Google Scholar

[35] Sturm K.-T., Generalized Orlicz spaces and Wasserstein distances for convex-concave scale functions, Bull. Sci. Math. 135 (2011), no. 6–7, 795–802. 10.1016/j.bulsci.2011.07.013Search in Google Scholar

[36] Villani C., Topics in Optimal Transportation, Grad. Stud. Math. 58, American Mathematical Society, Providence, 2003. Search in Google Scholar

[37] Villani C., Optimal Transport: Old and New, Springer, Berlin, 2009. 10.1007/978-3-540-71050-9Search in Google Scholar

Received: 2014-12-1
Revised: 2015-5-29
Accepted: 2015-10-20
Published Online: 2015-12-9
Published in Print: 2017-4-1

© 2017 by De Gruyter

Downloaded on 5.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/acv-2014-0040/html
Scroll to top button