Skip to main content
Springer logoLink to Springer
. 2020 Feb 14;31(3):2416–2438. doi: 10.1007/s12220-020-00358-6

Sharp Cheeger–Buser Type Inequalities in RCD(K,) Spaces

Nicolò De Ponti 1, Andrea Mondino 2,
PMCID: PMC7932992  PMID: 33746464

Abstract

The goal of the paper is to sharpen and generalise bounds involving Cheeger’s isoperimetric constant h and the first eigenvalue λ1 of the Laplacian. A celebrated lower bound of λ1 in terms of h, λ1h2/4, was proved by Cheeger in 1970 for smooth Riemannian manifolds. An upper bound on λ1 in terms of h was established by Buser in 1982 (with dimensional constants) and improved (to a dimension-free estimate) by Ledoux in 2004 for smooth Riemannian manifolds with Ricci curvature bounded below. The goal of the paper is twofold. First: we sharpen the inequalities obtained by Buser and Ledoux obtaining a dimension-free sharp Buser inequality for spaces with (Bakry–Émery weighted) Ricci curvature bounded below by KR (the inequality is sharp for K>0 as equality is obtained on the Gaussian space). Second: all of our results hold in the higher generality of (possibly non-smooth) metric measure spaces with Ricci curvature bounded below in synthetic sense, the so-called RCD(K,) spaces.

Keywords: Ricci curvature, Metric measure spaces, First eigenvalue laplace operator, Cheeger inequality, Buser inequality

Introduction

Throughout the paper (X,d) will be a complete metric space and m will be a non-negative Borel measure on X, finite on bounded subsets. The triple (X,d,m) is called metric measure space, m.m.s. for short. We denote by Lip(X) the space of real-valued Lipschitz functions over X and we write fLipb(X) if fLip(X) and f is bounded with bounded support. Given fLip(X) its slope |f|(x) at xX is defined by

|f|(x):=lim supyx|f(y)-f(x)|d(y,x), 1

with the convention |f|(x)=0 if x is an isolated point. The first non-trivial eigenvalue of the Laplacian is characterized as follows:

  • If m(X)<, the non-zero constant functions are in L2(X,m) and are eigenfunctions of the Laplacian with eigenvalue 0. In this case, we set
    λ1=inf{X|f|2dmX|f|2dm:0fLipb(X),Xfdm=0}. 2
  • When m(X)=, 0 may not be an eigenvalue of the Laplacian. Thus, we set
    λ0=inf{X|f|2dmX|f|2dm:0fLipb(X)}. 3

At this level of generality, the spectrum of the Laplacian may not be discrete (see Remark 1.3 for more details); in any case the definitions (2) and (3) make sense, and one can investigate bounds on λ1 and λ0.

Note that λ0 may be zero (for instance if m(X)< or if (X,d,m) is the Euclidean space Rd with the Lebesgue measure) but there are examples when λ0>0: for instance in the Hyperbolic plane λ0=1/4 and more generally on an n-dimensional simply connected Riemannian manifold with sectional curvatures bounded above by k<0 it holds λ0(n-1)2|k|/4 (see [27]).

Given a Borel subset AX with m(A)<, the perimeter Per(A) is defined as follows (see for instance [28]):

Per(A):=inf{lim infnX|fn|dm:fnLipb(X),fnχAinL1(X,m)}.

In 1970, Cheeger [17] introduced an isoperimetric constant, now known as Cheeger constant, to bound from below the first eigenvalue of the Laplacian. The Cheeger constant of the metric measure space (X,d,m) is defined by

h(X):=infPer(A)m(A):AXBorel subset withm(A)m(X)/2ifm(X)<infPer(A)m(A):AXBorel subset withm(A)<ifm(X)=. 4

The lower bound obtained in [17] for compact Riemannian manifolds, now known as Cheeger inequality, reads as

λ114h(X)2. 5

As proved by Buser [11], the constant 1/4 in (5) is optimal in the following sense: for any h>0 and ε>0, there exists a closed (i.e. compact without boundary) two-dimensional Riemannian manifold (Mg) with h(M)=h and such that λ114h(M)2+ε.

The paper [17] is in the framework of smooth Riemannian manifolds; however, the stream of arguments (with some care) extends to general metric measure spaces. For the reader’s convenience, we give a self-contained proof of (5) for m.m.s. in the Appendix (see Theorem 3.6).

Cheeger’s inequality (5) revealed to be extremely useful in proving lower bounds on the first eigenvalue of the Laplacian in terms of the isoperimetric constant h. It was thus an important discovery by Buser [12] that also an upper bound for λ1 in terms of h holds, where the inequality explicitly depends on the lower bound on the Ricci curvature of the smooth Riemannian manifold. More precisely, Buser [12] proved that for any compact Riemannian manifold of dimension n and RicK, K0 it holds

λ12-(n-1)Kh+10h2. 6

Note that the constant here is dimension-dependent. For a complete connected Riemannian manifold with RicK, K0, Ledoux [24] remarkably showed that the constant can be chosen to be independent of the dimension:

λ1max{6-Kh,36h2}. 7

The goal of the present work is twofold:

  1. The main results of the paper (Theorem 1.1 and Corollary 1.2) improve the constants in both the Buser-type inequalities (6)-(7) in a way that now the inequality is sharp for K>0 (as equality is attained on the Gaussian space).

  2. The inequalities are established in the higher generality of (possibly non-smooth) metric measure spaces satisfying Ricci curvature lower bounds in synthetic sense, the so-called RCD(K,) spaces.

For the precise definition of RCD(K,) space, we refer the reader to Section 2. Here, let us just recall that the RCD(K,) condition was introduced by Ambrosio-Gigli-Savaré [6] (see also [4]) as a refinement of the CD(K,) condition of Lott-Villani [26] and Sturm [33]. Roughly, a CD(K,) space is a (possibly infinite-dimensional, possibly non-smooth) metric measure space with Ricci curvature bounded from below by K, in a synthetic sense. While the CD(K,) condition allows Finsler structures, the main point of RCD is to reinforce the axiomatization (by asking linearity of the heat flow) to rule out Finsler structures and thus isolate the “possibly non-smooth Riemannian structures with Ricci curvature bounded below”. It is out of the scopes of this introduction to survey the long list of achievements and results proved for CD and RCD spaces (to this aim, see the Bourbaki seminar [34] and the recent ICM-Proceeding [1]). Let us just mention that a key property of both CD and RCD is the stability under measured Gromov–Hausdorff convergence (or more generally D-convergence of Sturm [6, 33], or even more generally pointed measured Gromov convergence [20]) of metric measure spaces. In particular pointed measured Gromov–Hausdorff limits of Riemannian manifolds with Ricci bounded below, the so-called Ricci limits, are examples of (possibly non-smooth) RCD spaces. Let us also recall that weighted Riemannian manifolds with Bakry-Émery Ricci tensor bounded below are also examples of RCD spaces; for instance the Gaussian space (Rd,|·|,(2π)-d/2e-|x|2/2dLd(x)), 1dN, satisfies RCD(1,). It is also worth recalling that if (X,d,m) is an RCD(K,) space for some K>0, then m(X)<; since scaling the measure by a constant does not affect the synthetic Ricci curvature lower bounds, when K>0, without loss of generality one can then assume m(X)=1.

To state our main result, it is convenient to set

JK(t)=2πKarctan(e2Kt-1)ifK>0,2πtifK=0,-2πKarctanh(1-e2Kt)ifK<0.t>0 8

The aim of the paper is to prove the following theorem.

Theorem 1.1

(Sharp implicit Buser-type inequality for RCD(K,) spaces) Let (X,d,m) be an RCD(K,) space, for some KR.

  • In case m(X)=1, then
    h(X)supt>01-e-λ1tJK(t). 9
    The inequality is sharp for K>0, as equality is achieved for the Gaussian space (Rd,|·|,(2π)-d/2e-|x|2/2dLd(x)), 1dN.
  • In case m(X)=, then
    h(X)2supt>01-e-λ0tJK(t). 10

Using the expression (8) of JK, in the next corollary we obtain more explicit bounds.

Corollary 1.2

(Explicit Buser inequality for RCD(K,) spaces) Let (X,d,m) be an RCD(K,) space, for some KR.

  • Case K>0. If Kλ1c>0, then
    λ1π2ch(X)2. 11
    The estimate is sharp, as equality is attained on the Gaussian space (Rd,|·|,(2π)-d/2e-|x|2/2dLd(x)), 1dN, for which K=1,λ1=1,h(X)=(2/π)1/2.
  • Case K=0, m(X)=1. It holds
    λ14πh(X)2infT>0T(1-e-T)2<πh(X)2. 12
    In case m(X)=, the estimate (12) holds replacing λ1 with λ0 and h(X) with h(X)/2.
  • Case K<0, m(X)=1. It holds
    λ1max{-K2log(e+e2-1)π(1-1e)h(X),2(log(e+e2-1))2π(1-1e)2h(X)2}<max2110-Kh(X),225h(X)2. 13
    In case m(X)=, the estimate (13) holds replacing λ1 with λ0 and h(X) with h(X)/2.

Remark 1.3

Even if the definitions of λ0 and λ1 as in (2) and (3) make sense regardless of the discreteness of the spectrum of the Laplacian (as well as the proofs of the above results), it is worth to mention some cases of interest where the Laplacian has discrete spectrum.

It was proved in [20] that an RCD(K,) space, with K>0 (or with finite diameter) has discrete spectrum (as the Sobolev imbedding V into L2 is compact). Even in case of infinite measure the embedding of V in L2 may be compact. An example is given by R with the Euclidean distance d(x,y)=|x-y| and the measure m:=12πex2/2dL1. It is a RCD(-1,) space and a result of Wang [35] ensures that the spectrum is discrete.

Comparison with Previous Results in the Literature

Theorem 1.1 and Corollary 1.2 improve the known results about Buser-type inequalities in several aspects. First of all, the best results obtained before this paper are the aforementioned estimates (6)-(7) due to Buser [12] and Ledoux [24] for smooth complete Riemannian manifolds satisfying RicK, K0. Let us stress that the constants in Corollary 1.2 improve the ones in both (6)-(7) and are dimension-free as well. In addition, the improvements of the present paper are:

  • In case K>0, the inequalities (9) and (11) are sharp (as equality is attained on the Gaussian space).

  • The results hold in the higher generality of (possibly non-smooth) RCD(K,) spaces.

The proof of Theorem 1.1 is inspired by the semi-group approach of Ledoux [23, 24], but it improves upon using Proposition 3.1 in place of:

  • A dimension-dependent Li-Yau inequality, in [23].

  • A weaker version of Proposition 3.1 (see [24, Lemma 5.1]) analyzed only in case K0, in [24].

Theorem 1.1 and Corollary 1.2 are also the first upper bounds in the literature of RCD spaces for the first eigenvalue of the Laplacian. On the other hand, lower bounds on the first eigenvalue of the Laplacian have been thoroughly analyzed in both CD and RCD spaces: the sharp Lichnerowitz spectral gap λ1KN/(N-1) was proved under the (non-branching) CD(K,N) condition by Lott-Villani [25], under the RCD(K,N) condition by Erbar-Kuwada-Sturm [18], and generalized by Cavalletti and Mondino [14] to a sharp spectral gap for the p-Laplacian for essentially non-branching CD(K,N) spaces involving also an upper bound on the diameter (together with rigidity and almost rigidity statements). Jiang-Zhang [21] independently showed, for p=2, that the improved version under an upper diameter bound holds for RCD(K,N). The rigidity of the Lichnerowitz spectral gap for RCD(K,N) spaces, K>0, N(1,), known as Obata’s Theorem was first proved by Ketterer [22]. The rigidity in the Lichnerowitz spectral gap for RCD(K,) spaces, K>0, was recently proved by Gigli-Ketterer-Kuwada-Ohta [19]. Local Poincaré inequalities in the framework of CD(K,N) and CD(K,) spaces were proved by Rajala [30]. Finally various lower bounds, together with rigidity and almost rigidity statements for the Dirichlet first eigenvalue of the Laplacian, have been proved by Mondino-Semola [29] in the framework of CD and RCD spaces. Lower bounds on Cheeger’s isoperimetric constant have been obtained for (essentially non-branching) CD(K,N) spaces by Cavalletti-Mondino [1315] and for RCD(K,) spaces (K>0) by Ambrosio-Mondino [7]. The local and global stability properties of eigenvalues and eigenfunctions in the framework of RCD spaces have been investigated by Gigli-Mondino-Savaré in [20], and by Ambrosio-Honda in [2, 3].

Preliminaries

Throughout the paper, unless otherwise stated, we assume (X,d) is a complete and separable metric space. We endow (X,d) with a reference σ-finite non-negative measure m over the Borel σ-algebra B, with supp(m)=X and satisfying an exponential growth condition, namely that there exist x0X, M>0 and c0 such that

m(Br(x0))Mexp(cr2)for everyr0.

Possibly enlarging B and extending m, we assume that B is m-complete. The triple (X,d,m) is called metric measure space, m.m.s for short.

We denote by P2(X) the space of probability measures on X with finite second moment and we endow this space with the Kantorovich–Wasserstein distance W2 defined as follows: for μ0,μ1P2(X) we set

W22(μ0,μ1):=infπX×Xd2(x,y)dπ, 14

where the infimum is taken over all πP(X×X) with μ0 and μ1 as the first and the second marginal.

The relative entropy functional Entm:P2(X)R{} is defined as

Entm(μ):=ρlogρdmifμ=ρm,otherwise. 15

A curve γ:[0,1]X is a geodesic if

d(γs,γt)=|t-s|d(γ0,γ1)s,t[0,1]. 16

In the sequel we use the notation:

D(Entm):={μP2(X):Entm(μ)R}.

We now define the CD(K,) condition, coming from the seminal works of Lott-Villani [26] and Sturm [33].

Definition 2.1

(CD(K,) condition) Let KR. We say that (X,d,m) is a CD(K,) space provided that for any μ0,μ1D(Entm) there exists a W2-geodesic (μt) such that μ0=μ0, μ1=μ1 and

Entm(μt)(1-t)Entm(μ0)+tEntm(μ1)-K2t(1-t)W22(μ0,μ1). 17

The space of continuous function f:XR is denoted by C(X) and the Lebesgue space by Lp(X,m), 1p.

The Cheeger energy (introduced in [16] and further studied in [5]) is defined as the L2-lower semicontinuous envelope of the functional f12X|f|2dm, i.e.:

Chm(f):=inf{lim infn12X|fn|2dm:fnLipb(X),fnfinL2(X,m)}. 18

If Chm(f)<, it was proved in [5, 16] that the set

G(f):={gL2(X,m):(fn)nLipb(X),fnf,|fn|hginL2(X,m)}

is closed and convex, therefore, it admits a unique element of minimal norm called minimal weak upper gradient and denoted by |Df|w. The Cheeger energy can be then represented by integration as

Chm(f)=12X|Df|w2dm.

We recall that the minimal weak upper gradient satisfies the following property (see e.g. [6, equation (2.18)]):

|Df|w=0m-a.e. on the set{f=0}. 19

One can show that Chm is a 2-homogeneous, lower semicontinuous, convex functional on L2(X,m) whose proper domain

V:={fL2(X,m):Chm(f)<}

is a dense linear subspace of L2(X,m). It then admits an L2 gradient flow which is a continuous semi-group of contractions (Ht)t0 in L2(X,m), whose continuous trajectories tHtf, for fL2(X,m), are locally Lipschitz curves from (0,) with values into L2(X,m) that satisfy

ddtHtf-Chm(Htf)for a.e.t(0,). 20

Here, denotes the subdifferential of convex analysis, namely for every fV we have Chm(f) if and only if

X(g-f)dmChm(g)-Chm(f),for everygL2(X,m). 21

We now define the RCD(K,) condition, introduced and throughly analyzed in [6] (see also [4] for the present simplified axiomatization and the extension to the σ-finite case).

Definition 2.2

(RCD(K,) condition) Let KR. We say that the metric measure space (X,d,m) is RCD(K,) if it satisfies the CD(K,) condition and moreover the Cheeger energy Chm is quadratic, i.e. it satisfies the parallelogram identity

Chm(f+g)+Chm(f-g)=2Chm(f)+2Chm(g),f,gV. 22

If (X,d,m) is an RCD(K,) space, then the Cheeger energy induces the Dirichlet form E(f):=2Chm(f) which is strongly local, symmetric and admits the Carré du Champ

Γ(f):=|Df|w2,fV.

The space V endowed with the norm fV2:=fL22+E(f) is Hilbert. Moreover, the sub-differential Chm is single-valued and coincides with the linear generator -Δ of the heat flow semi-group (Ht)t0 defined above. In other terms, the semigroup can be equivalently characterized by the fact that for any fL2(X,m) the curve tHtfL2(X,m) is locally Lipschitz from (0,) to L2(X,m) and satisfies

ddtHtf=ΔHtfforL1-a.et(0,),limt0Htf=f, 23

where the limit is in the strong L2(X,m)-topology.

The semigroup Ht extends uniquely to a strongly continuous semigroup of linear contractions in Lp(X,m),p[1,), for which we retain the same notation. Regarding the case p=, it was proved in [6, Theorem 6.1] that there exists a version of the semigroup such that Htf(x) belongs to CL((0,)×X) whenever fL(X,m). We will implicitly refer to this version of Htf when f is essentially bounded. Moreover, for any fL2L(X,m) and for every t>0 we have HtfVLip(X) with the explicit bound (see [6, Theorem 6.5] for a proof)

|DHtf|wKe2Kt-1f. 24

Two crucial properties of the heat flow are the preservation of mass and the maximum principle (see [5]):

XHtfdm=Xfdm,for anyfL1(X,m), 25
0HtfC,for any0fCm-a.e.,C>0. 26

A result of Savaré [31, Corollary 3.5] ensures that, in the RCD(K,) setting, for every fV and α[12,1] we have

|DHtf|w2αe-2αKtHt(|Df|w2α),m-a.e.. 27

In particular,

|DHtf|we-KtHt(|Df|w),m-a.e.. 28

Proof of Theorem 1.1

We denote by I:[0,1][0,12π] the Gaussian isoperimetric function defined by I:=φΦ-1 where

Φ(x):=12π-xe-u2/2du,xR,

and φ=Φ. The function I is concave, continuous, I(0)=I(1):=0 and 0I(x)I(12)=12π, for all x[0,1]. Moreover, IC((0,1)), it satisfies the identity

I(x)I(x)=-1,for everyx(0,1). 29

and (see [10])

limx0I(x)x2log1x=1. 30

Given KR, we define the function jK:(0,)(0,) as

jK(t):=Ke2Kt-1ifK0,12tifK=0. 31

Notice that jK is increasing as a function of K.

The next proposition was proved in the smooth setting by Bakry, Gentil and Ledoux (see [8, 10] and [9, Proposition 8.6.1]).

Proposition 3.1

(Bakry-Gentil-Ledoux Inequality in RCD(K,) spaces) Let (X,d,m) be an RCD(K,) space, for some KR. Then, for every function fL2(X,m), f:X[0,1] it holds

|D(Htf)|w2jK(t)([I(Htf)]2-[Ht(I(f))]2),m-a.e.,for everyt>0. 32

In particular, for every fL2L(X,m), it holds

|D(Htf)|w2πjK(t)f,m-a.e.,for everyt>0. 33

Proof

Given ε>0, η>2ε and δ>0 sufficiently small, consider fL2(X,m) with values in [0,1-η]. We define

ϕε(x):=I(x+ε)-I(ε), 34
Ψε(s):=[Hs(ϕε(Ht-sf))]2,for everys(0,t). 35

We notice that ϕε(0)=0 and ϕε(x)0 for every x[0,1-η]. Moreover, using the property (26), ϕε is Lipschitz in the range of Ht-sf. Since tHtf is a locally Lipschitz map with values in Lp(X,m) for 1<p< ( [32, Theorem 1, Section III]), we have that Ψε is a locally Lipschitz map with values in L1(X,m). Let ψL1L(X,m) be a non-negative function. By the chain rule for locally Lipschitz maps, the fundamental theorem of calculus for the Bochner integral and the properties of the semigroup Ht we have that for any ε>0 and 0<δ<t it holds

X([Hδ(ϕε(Ht-δf))]2-[Ht-δ(ϕε(Hδf))]2)ψdm=δt-δ(-ddsX[Hs(ϕε(Ht-sf))]2ψdm)ds=-2δt-δ(XHs(ϕε(Ht-sf))Hs(Δϕε(Ht-sf)-ϕε(Ht-sf)ΔHt-sf)ψdm)ds=2δt-δ(XHs(ϕε(Ht-sf))Hs(-ϕε(Ht-sf)|DHt-sf|w2)ψdm)ds. 36

Applying the Cauchy–Schwarz inequality

Hs(X)Hs(Y)[Hs(XY)]2,

and the identity I(x)I(x)=-1, for all x(0,1), we get that the right-hand side of (36) is bounded below by

2δt-δ(X[Hs((1-I(ε)I(Ht-sf+ε))|DHt-sf|w2)]2ψdm)ds. 37

Noticing that

X[Hs((1-I(ε)I(Ht-sf+ε))|DHt-sf|w2)]2ψdmX[Hs(|DHt-sf|w)]2ψdm

and that, for any fixed δ>0,

δt-δ(X[Hs(|DHt-sf|w)]2ψdm)ds<

thanks to the bound (24), we can pass to the limit as ε0 in (37) using Dominated Convergence Theorem.

Since I is continuous, I(0)=0 and I(x)>0 for every x(0,1), using the locality property (19), the Dominated Convergence Theorem yields

X([Hδ(I(Ht-δf))]2-[Ht-δ(I(Hδf))]2)ψdm2δt-δ(X[Hs(|DHt-sf|w)]2ψdm)ds, 38

for every δ(0,t). Now, we can bound the right-hand side of (38) using the inequality (28) to obtain

2δt-δ(X[Hs(|DHt-sf|w)]2ψdm)ds2X(δt-δe2Ksds)|DHtf|w2ψdm. 39

From (30) it follows that for every 0<a<1 there exists C=C(a)>0 and x¯=x¯(a)(0,1) such that I(x)Cxa for all x(0,x¯). In particular, if gL2(X,m), g:X[0,1-η], then I(g)Lp(X,m) for every p>2. We now apply this argument for p=4, so that we can take advantage of the continuity of I and the continuity of the semigroup and pass to the limit as δ0. We obtain

X([I(Htf)]2-[Ht(I(f))]2)ψdm1jK(t)X|DHtf|w2ψdm, 40

for every η>0 sufficiently small, every fL2(X,m), f:X[0,1-η].

Now, for fL2(X,m), f:X[0,1], consider the truncation fη:=min{f,1-η}. Applying (40) to fη, we have

X([I(Htfη)]2-[Ht(I(fη))]2)ψdm1jK(t)X|DHtfη|w2ψdm. 41

From fηf in L2L(X,m) as η0, we get that HtfηHtf in V for every t>0; we can then pass to the limit as η0 in (41) and obtain

X([I(Htf)]2-[Ht(I(f))]2)ψdm1jK(t)X|DHtf|w2ψdm.

Since ψL1L(X,m), ψ0 is arbitrary, the desired estimate (32) follows.

Recalling that 0I12π, the inequality (32) yields

|D(Htf)|wjK(t)2π,m-a.e.,for everyt>0, 42

for any fL2(X,m), f:X[0,1]. For any fL2L(X,m), write f=f+-f- with f+=max{f,0}, f-=max{-f,0}. Applying (42) to f+/f,f-/f and summing up we obtain

|DHtf|w|DHtf+|w+|DHtf-|w2πjK(t)f,m-a.e.,t>0.

We next recall the definition of the first non-trivial eigenvalue of the laplacian -Δ. First of all, if m(X)<, the non-zero constant functions are in L2(X,m) and are eigenfunctions of -Δ with eigenvalue 0. In this case, the first non-trivial eigenvalue is given by λ1

λ1=inf{X|Df|w2dmX|f|2dm:0fV,Xfdm=0}. 43

When m(X)=, 0 may not be an eigenvalue of -Δ and the first eigenvalue is characterized by

λ0=inf{X|Df|w2dmX|f|2dm:0fV}. 44

Observe that, by the very definition of Cheeger energy (18), the definition (2) of λ1 (resp. (3) of λ0) given in the Introduction in terms of slope of Lipschitz functions, is equivalent to (43) (resp. (44)).

It is also convenient to set

JK(t):=2π0tjK(s)ds, 45

where jK was defined in (31).

Proof of Theorem 1.1

Step 1:

Proof of (9), the case m(X)=1.

First of all, we claim that for any fL2(X,m) with zero mean it holds
Htf2e-λ1tf2. 46
To prove (46) let 0fL2(X,m) such that 0=Xfdm=XHtfdm. Then
2λ1X|Htf|2dm2X|D(Htf)|w2dm=-2XHtfΔ(Htf)dm=-ddtX|Htf|2dm, 47
and the Gronwall’s inequality yields (46).
Next, we claim that, by duality, the bound (33) implies
f-Htf1JK(t)|Df|w1,for allfLipb(X), 48
where JK(t) was defined in (45).
To prove (48) we take a function g, g1, and observe that
Xg(f-Htf)dm=-0t(XgΔHsfdm)ds=0t(XDHsg·Dfdm)ds|Df|w10t|D(Hsg)|wds.
Since g is arbitrary, the claimed (48) follows from the last estimate combined with (33).
We now combine the above claims to conclude the proof. Let AX be a Borel subset and let fnLipb(X), fnχA in L1(X,m), be a recovery sequence for the perimeter of the set A, i.e.:
Per(A)=limnX|fn|dmlim supnX|Dfn|wdm.
Inequality (48) passes to the limit since Ht is continuous in L1(X,m) [5, Theorem 4.16] and we can write
JK(t)Per(A)χA-Ht(χA)1=A[1-Ht(χA)]dm+AcHt(χA)dm=2(m(A)-AHt(χA)dm)=2(m(A)-XχAHt/2(Ht/2(χA))dm)=2(m(A)-XHt/2(χA)Ht/2(χA)dm)=2(m(A)-Ht/2(χA)22), 49
where we used properties (25), (26), together with the semigroup property and the self-adjointness of the semigroup. We observe that XHt/2(χA-m(A))dm=0 thanks to (25) and the fact that Ht1=1 when m(X)=1. We can thus apply (46) in order to bound Ht/2(χA)22 in the following way
Ht/2(χA)22=m(A)2+Ht/2(χA-m(A))22m(A)2+e-λ1tχA-m(A)22. 50
A direct computation gives χA-m(A)22=m(A)(1-m(A)), so that the combination of (49) and (50) yields
JK(t)Per(A)2m(A)(1-m(A))(1-e-λ1t),for everyt>0. 51
Recalling that in the definition of the Cheeger constant h(X) one considers only Borel subsets AX with m(A)1/2, the last inequality (51) gives (9).
Step 2:

Proof of (10), the case m(X)=.

Arguing as in (47) using Gronwall Lemma, for any fL2(X,m) it holds
Htf2e-λ0tf2. 52
Note that to establish (49), the finiteness of m(X) played no role. Now, we can directly use (52) to bound the right-hand side of the equation (49) in order to achieve
Per(A)m(A)2supt>0{1-e-λ0tJK(t)},
for any Borel subset AX with m(A)<. The estimate (10) follows.

From the implicit to explicit bounds (and sharpness in case K>0)

Proof of Corollary 1.2 In this section, we show how to derive explicit bounds for λ1 (resp. λ0) in term of the Cheeger constant h(X), starting from (9) (resp. (10)). We also show that (9) is sharp, since equality is achieved on the Gaussian space. First of all, the expression of the function JK defined in (45) can be explicitly computed as:

JK(t)=2πKarctan(e2Kt-1)ifK>0,2πtifK=0,-2πKarctanh(1-e2Kt)ifK<0.t>0 53

CaseK=0

When K=0, the estimate (9) combined with (53) gives

h(X)π2supt>01-e-λ1tt=πλ12supT>01-e-TT, 54

where we set T=λ1t in the last identity.

Let W-1:[-1/e,0)(-,-1] be the lower branch of the Lambert function, i.e. the inverse of the function xxex in the interval (-,-1]. An easy computation yields

M:=supT>01-e-TT=-4W-1(-12e)-22W-1(-12e),achievedatT=-W-1(-12e)-12. 55

A good lower estimate of M is given by 2/π. Using this bound, we obtain

λ1<πh2(X).

CaseK>0

We start with the following

Lemma 3.2

Let f1:(0,)(0,) be defined as

f1(x):=xarctan(eTx-1), 56

where T>0 is a fixed number. Then f1 is an increasing function and f1(x)1T.

Proof

The function f1 is differentiable and the derivative of f1 is non-negative if and only if

eTx-1arctan(eTx-1)-Tx0,x>0.

We put y:=eTx-1 so that we have to prove

yarctan(y)-log(y2+1)0,y>0. 57

Called g1(y) the function g1(y):=yarctan(y)-log(y2+1), we have that g1(0)=0 and

g1(y)=arctan(y)-y1+y20,

so that the inequality (57) is proved and f1 is increasing for any T>0. The proof is finished since

limx0f1(x)=1T.

Rewriting the estimate (9) using (53) in case K>0, we obtain

2πh(X)Ksupt>01-e-λ1tarctan(e2Kt-1)=λ1supT>0Kλ1arctan(e2Kλ1T-1)(1-e-T). 58

Thanks to the Lemma 3.2 it is clear that we can always obtain the same lower bound of the case K=0 (as expected), but this can be improved as soon as we have a positive lower bound of the quotient K/λ1. Indeed, let us suppose K/λ1c>0. Then, observing that

supT>01-e-Tarctan(e2cT-1)limT+1-e-Tarctan(e2cT-1)=2π,

from (58), we obtain

2cπh(X)λ1supT>01-e-Tarctan(e2cT-1)2πλ1. 59

When X=Rd endowed with the Euclidean distance d(x,y)=|x-y| and the Gaussian measure (2π)-d/2e-|x|2/2dLd, 1dN, we have that h(X)=2π, K=1 and λ1=1 (see [9, Section 4.1]). Thus, we can take c=1 and the equality in (59) is achieved, making sharp the lower bound.

CaseK<0

We begin by noticing that

JK(t)=-2πKarctanh(1-e2Kt)=-2πKlog(e-Kt+e-2Kt-1). 60

The following lemma holds:

Lemma 3.3

Let f2:(0,)(0,) be defined as

f2(x):=xlog(eTx+e2Tx-1), 61

where T>0 is a fixed number. Then, f2 is a decreasing function.

Proof

A direct computation shows that the derivative of f2 is non-positive if and only if

e2Tx-1log(eTx+e2Tx-1)2TxeTx,for allx>0,

which is equivalent to

1-e-2Txlog(1+1-e-2Tx)(2-1-e-2Tx)Tx,for allx>0. 62

We put y:=1-e-2Tx, and we write (62) as

ylog(1+y)+12(2-y)log(1-y2)0,for all0<y<1,

which in turn is equivalent to

1+y2log(1+y)+1-y2log(1-y)0,for all0<y<1. 63

Now define g2:(0,1)R as g2(y):=(1+y2)log(1+y)+(1-y2)log(1-y) and observe that g2 is concave with g2(0)=0, g2(0)=0. Thus g2 is non-positive on (0, 1) and the inequality (63) is proved.

The combination of (9), (53) and (60) implies that if (X,d,m) is an RCD(K,) space with K<0 and m(X)=1 then

h(X)-πK2supt>01-e-λ1tlog(e-Kt+e-2Kt-1). 64

We make two different choices:

  • When λ1-K, we choose t=-1K in (64) so that
    h(X)-πK21-eλ1Klog(e+e2-1)λ1-π2K1-1elog(e+e2-1), 65
    where we used the inequality
    1-e-x1-1ex,for all0x1.
  • When λ1>-K, we choose t=1λ1 in (64) so that
    h(X)π2λ11-1e-Kλ1log(e-Kλ1+e-2Kλ1-1).
    Applying now Lemma 3.3, we obtain
    λ12(log(e+e2-1))2π(1-1e)2h(X)2. 66

The combination of (65) and (66) gives that, if (X,d,m) is an RCD(K,) space with K<0 and m(X)=1

λ1max{-K2log(e+e2-1)π(1-1e)h(X),2(log(e+e2-1))2π(1-1e)2h(X)2}<max2110-Kh(X),225h(X)2. 67

In case (X,d,m) is an RCD(K,) space with K<0 and m(X)= then, using (10) instead of (9), the estimates (64) and (67) hold with λ1 replaced by λ0 and h(X) replaced by h(X)/2. Thus, in case m(X)=, we obtain:

λ0max{-Klog(e+e2-1)2π(1-1e)h(X),(log(e+e2-1))22π(1-1e)2h(X)2}<max2120-Kh(X),1110h(X)2. 68

Remark 3.4

Another bound, similar to the one obtained in the case K>0, can be achieved in the presence of a lower bound for K/λ1, if m(X)=1 (resp. a lower bound for K/λ0, if m(X)=). To see this, let us suppose K/λ1-c,c>0 (resp. K/λ0-c). Then, using (9) (resp. (10)), (53) and Lemma 3.3, we have that (resp. the left-hand side can be improved to h(X)/2π)

2πh(X)λ1supT>0-Kλ1log(e-Kλ1T+e-2Kλ1T-1)(1-e-T)cλ1supT>01-e-Tlog(ecT+e2cT-1). 69

Acknowledgements

The work has been developed when N. DP. was visiting the Mathematics Institute at the University of Warwick during fall term 2018. He wishes to thank the Institute for the excellent working conditions and the stimulating atmosphere. N.DP. is supported by the GNAMPA Project 2019 “Trasporto ottimo per dinamiche con interazione”. A.M. is supported by the EPSRC First Grant EP/R004730/1 “Optimal transport and Geometric Analysis” and by the ERC Starting Grant 802689 “CURVATURE”.

Appendix A: Cheeger’s Inequality in General Metric Measure Spaces

The Buser-type inequalities of Theorem 1.1 and Corollary 1.2 give an upper bound on λ1 (resp. on λ0, in case m(X)=) in terms of the Cheeger constant h(X). It is natural to ask if also a reverse inequality holds, namely if it possible to give a lower bound on λ1 (resp. on λ0, in case m(X)=) in terms of h(X). The answer is affirmative in the higher generality of metric measure spaces with a non-negative locally bounded measure without curvature conditions, see Theorem 3.6 below. This generalizes to the metric measure setting a celebrated result by Cheeger [17], known as Cheeger’s inequality. In contrast to the previous section, here we do not assume the separability of the space.

A key tool in the proof of Cheeger’s inequality is the co-area formula; more precisely, in the arguments it is enough to have an inequality in the co-area formula. For the reader’s convenience, we give below the statement and a self-contained proof.

Proposition 3.5

(Coarea inequality) Let (X,d) be a complete metric space and let m be a non-negative Borel measure finite on bounded subsets.

Let uLipb(X), u:X[0,) and set M=supXu. Then for L1-a.e. t>0 the set {u>t} has finite perimeter and

0MPer({u>t})dtX|u|dm. 70

Proof

The proof is quite standard, but since we did not find it in the literature stated at this level of generality (typically one assumes some extra condition like measure doubling and gets a stronger statement, namely equality in the co-area formula; see for instance [28]) we add it for the reader’s convenience.

Let Et:={u>t} and set V(t):=Et|u|dm. The function tV(t) is non-increasing and bounded, thus differentiable for L1-a.e. t>0.

Since Xudm<, we also have that m({u=t})=0 for L1-a.e. t>0.

Fix t>0 a differentiability point for V for which m({u=t})=0, and define ψ:(0,t)×(0,)[0,1] as

ψ(h,s):=0forst-h1h(s-t)+1fort-h<st1fors>t. 71

For h>0 define uh(x)=ψ(h,u(x)) and observe that the sequence (uh)hLipb(X).

We first claim that

uhχEtinL1(X,m)ash0. 72

Indeed

X|uh-χEt|dm={t-h<ut}ψ(h,u)dmmt-h<utm({u=t})=0ash0,

by Dominated Convergence Theorem, since by assumption u has bounded support, m is finite on bounded sets and χ{t-h<ut}χ{u=t} pointwise as h0.

In order to prove that Et is a set of finite perimeter it is then sufficient to show that lim suph0X|uh|dm<. To this aim observe that

X|uh|dm=1h{t-h<ut}|u|dm=V(t-h)-V(t)h.

Since by assumption t>0 is a differentiability point for V, we obtain that Et is a finite perimeter set satisfying

Per(Et)limh0X|uh|dm=-V(t). 73

Using that (73) holds for L1-a.e. t>0 and that V is non-increasing, we get

0MPer(Et)dt-0MV(t)dtV(0)-V(M)=X|u|dm. 74

Theorem 3.6

(Cheeger’s Inequality in metric measure spaces) Let (X,d) be a complete metric space and let m be a non-negative Borel measure finite on bounded subsets.

  1. If m(X)< then
    λ114h(X)2. 75
  2. If m(X)= then
    λ014h(X)2. 76

As proved by Buser [11], the constant 1/4 in (75) is optimal in the following sense: for any h>0 and ε>0, there exists a closed (i.e. compact without boundary) two-dimensional Riemannian manifold (Mg) with h(M)=h and such that λ114h(M)2+ε.

Proof

We give a proof of (75), the arguments for showing (76) being analogous (and even simpler).

By the very definition of λ1 as in (2), for every ε>0 there exists fLipb(X) with Xfdm=0, f0 such that

λ1X|f|2dmXf2dm-ε. 77

Let m be any median of the function f and set f+:=max{f-m,0}, f-:=-min{f-m,0}. Applying the co-area inequality (70) to u=(f+)2 (respectively (f-)2) and recalling the definition of Cheeger’s constant h(X) as in (4), we obtain

X|(f+)2|dm+X|(f-)2|dm0sup{(f+)2}Per({(f+)2>t})dt+0sup{(f-)2}Per({(f-)2>t})dth(X)0sup{(f+)2}m({(f+)2>t})dt+h(X)0sup{(f-)2}m({(f-)2>t})dt=h(X)X(f+)2dm+h(X)X(f-)2dm=h(X)X|f-m|2dm. 78

Since

|g2|2|g||g|,

and

|f+||f|,|f-||f|,

we can apply the Cauchy-Schwarz inequality and get

2(X|f|2dm)12(X|f-m|2dm)12X|(f+)2|dm+X|(f-)2|dm, 79

where we have used that |f+|+|f-|=|f-m|. It follows from (78) and (79) that for every median m of f it holds

X|f|2dmX|f-m|2dmh(X)24. 80

Finally, since Xfdm=0 and the mean minimises RcX|f-c|2dm, we have

X|f|2dmX|f|2dmX|f|2dmX|f-m|2dm

and we can conclude thanks to (77) and the fact that ε>0 is arbitrary.

Footnotes

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Contributor Information

Nicolò De Ponti, Email: nicolo.deponti01@universitadipavia.it.

Andrea Mondino, Email: Andrea.Mondino@maths.ox.ac.uk.

References

  • 1.Ambrosio, L.: Calculus, heat flow and curvature-dimension bounds in metric measure spaces, Proceedings of the ICM 2018, Rio de Janeiro, Vol. 1, 301–340
  • 2.Ambrosio, L., Honda, S.: New stability results for sequences of metric measure spaces with uniform Ricci bounds from below. In Measure Theory in Non-Smooth Spaces, pp. 1–51. De Gruyter Open, Warsaw, Poland (2017). 10.1515/9783110550832-001
  • 3.Ambrosio, L., Honda, S.: Local spectral convergence in RCD(K,N) spaces. Nonlinear Anal. 177(Part A), 1–23 (2018)
  • 4.Ambrosio L, Gigli N, Mondino A, Rajala T. Riemannian Ricci curvature lower bounds in metric measure spaces with σ-finite measure. Trans. Amer. Math. Soc. 2015;367(7):4661–4701. doi: 10.1090/S0002-9947-2015-06111-X. [DOI] [Google Scholar]
  • 5.Ambrosio L, Gigli N, Savaré G. Calculus and heat flow in metric measure spaces and applications to spaces with Ricci bounds from below. Invent. Math. 2014;195(2):289–391. doi: 10.1007/s00222-013-0456-1. [DOI] [Google Scholar]
  • 6.Ambrosio L, Gigli N, Savaré G. Metric measure spaces with Riemannian Ricci curvature bounded from below. Duke Math. J. 2014;163:1405–1490. doi: 10.1215/00127094-2681605. [DOI] [Google Scholar]
  • 7.Ambrosio L, Mondino A. Gaussian-type isoperimetric inequalities in RCD(K,) probability spaces for positive K, Atti Accad. Naz. Lincei Rend. Lincei Mat. Appl. 2016;27(4):497–514. doi: 10.4171/RLM/745. [DOI] [Google Scholar]
  • 8.Bakry D, Gentil I, Ledoux M. On Harnack inequalities and optimal transportation. Ann. Sc. Norm. Super. Pisa Cl. Sci., (5) 2015;14(3):705–727. [Google Scholar]
  • 9.Bakry, D., Gentil, I., Ledoux , M.: Analysis and geometry of Markov diffusion operators, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], 348. Springer, Cham. xx+552 (2014)
  • 10.Bakry D, Ledoux M. Lévy-Gromov’s isoperimetric inequality for an infinite-dimensional diffusion generator. Invent. Math. 1996;123(2):259–281. doi: 10.1007/s002220050026. [DOI] [Google Scholar]
  • 11.Buser, P.: Über eine Ungleichung von Cheeger [On an inequalityof Cheeger]. Math. Z. (in German). 158, no. 3, 245–252. (1978)
  • 12.Buser P. A note on the isoperimetric constant. Ann. Sci. Ecole Norm. Sup., (4) 1982;15(2):213–230. doi: 10.24033/asens.1426. [DOI] [Google Scholar]
  • 13.Cavalletti F, Mondino A. Sharp and rigid isoperimetric inequalities in metric-measure spaces with lower Ricci curvature bounds. Inventiones Math. 2017;208(3):803–849. doi: 10.1007/s00222-016-0700-6. [DOI] [Google Scholar]
  • 14.Cavalletti F, Mondino A. Sharp geometric and functional inequalities in metric measure spaces with lower Ricci curvature bounds. Geom. Topol. 2017;21(1):603–645. doi: 10.2140/gt.2017.21.603. [DOI] [Google Scholar]
  • 15.Cavalletti F, Mondino A. Isoperimetric inequalities for finite perimeter sets under lower Ricci curvature bounds. Atti Accad. Naz. Lincei Rend. Lincei Mat. Appl. 2018;29(3):413–430. doi: 10.4171/RLM/814. [DOI] [Google Scholar]
  • 16.Cheeger J. Differentiability of Lipschitz functions on metric measure spaces. Geom. Funct. Anal. 1999;9:428–517. doi: 10.1007/s000390050094. [DOI] [Google Scholar]
  • 17.Cheeger, J.: A lower bound for the smallest eigenvalue of the Laplacian. In Gunning, Robert C. Problems in analysis (Papers dedicated to Salomon Bochner: Princeton, pp. 195–199. Princeton Univ. Press, N. J. (1969)
  • 18.Erbar M, Kuwada K, Sturm KT. On the Equivalence of the Entropic Curvature-Dimension Condition and Bochner’s Inequality on Metric Measure Space. Invent. Math. 2015;201(3):993–1071. doi: 10.1007/s00222-014-0563-7. [DOI] [Google Scholar]
  • 19.Gigli, N., Ketterer, C., Kuwada, K., Ohta, S.I.: Rigidity for the spectral gap on RCD(K,)-spaces. preprint arXiv:1709.04017, to appear in Am. J. Math
  • 20.Gigli N, Mondino A, Savaré G. Convergence of pointed non-compact metric measure spaces and stability of Ricci curvature bounds and heat flows. Proc. Lond. Math. Soc., (3) 2015;111(5):1071–1129. [Google Scholar]
  • 21.Jiang Y, Zhang H-C. Sharp spectral gaps on metric measure spaces. Calc. Var. Partial Differential Equations. 2016;55:14. doi: 10.1007/s00526-016-0952-4. [DOI] [Google Scholar]
  • 22.Ketterer C. Obata’s rigidity theorem for metric measure spaces. Anal. Geom. Metr. Spaces. 2015;3:278–295. [Google Scholar]
  • 23.Ledoux M. A Simple Analytic Proof of an Inequality by P. Buser. Proceedings of the American Mathematical Society. 1994;121(3):951–959. doi: 10.1090/S0002-9939-1994-1186991-X. [DOI] [Google Scholar]
  • 24.Ledoux, M.: Spectral gap, logarithmic Sobolev constant, and geometric bounds, Surveys in differential geometry. Vol. IX, 219-240, Int. Press, Somerville, MA, (2004)
  • 25.Lott J, Villani C. Weak curvature conditions and functional inequalities. Journ. Funct. Analysis. 2007;245:311–333. doi: 10.1016/j.jfa.2006.10.018. [DOI] [Google Scholar]
  • 26.Lott J, Villani C. Ricci curvature for metric-measure spaces via optimal transport. Ann. of Math. 2009;169:903–991. doi: 10.4007/annals.2009.169.903. [DOI] [Google Scholar]
  • 27.McKean HP. An upper bound to the spectrum of Δ on a manifold of negative curvature. Journ. Diff. Geom. 1970;4:359–366. doi: 10.4310/jdg/1214429509. [DOI] [Google Scholar]
  • 28.Miranda M., Jr Functions of bounded variation on good metric spaces. J. Math. Pures Appl. 2003;82:975–1004. doi: 10.1016/S0021-7824(03)00036-9. [DOI] [Google Scholar]
  • 29.Mondino A, Semola D. Polya-Szego inequality and Dirichlet p-spectral gap for non-smooth spaces with Ricci curvature bounded below. J. Math. Pure. Appl. 2019 doi: 10.1016/j.matpur.2019.10.005. [DOI] [Google Scholar]
  • 30.Rajala T. Local Poincaré inequalities from stable curvature conditions on metric spaces. Calc. Var. Partial Differential Equations. 2012;44(3–4):477–494. doi: 10.1007/s00526-011-0442-7. [DOI] [Google Scholar]
  • 31.Savaré G. Self-improvement of the Bakry-Émery condition and Wasserstein contraction of the heat flow in RCD(K,) metric measure spaces. Discrete Contin. Dyn. Syst. 2014;34(4):1641–1661. doi: 10.3934/dcds.2014.34.1641. [DOI] [Google Scholar]
  • 32.Stein, E.M.: Topics in Harmonic Analysis related to the Littlewood-Paley Theory, Annals of Mathematics Studies, No. 63, Princeton University Press, Princeton, NJ, (1970)
  • 33.Sturm KT. On the geometry of metric measure spaces. Acta Math. 2006;196:65–131. doi: 10.1007/s11511-006-0002-8. [DOI] [Google Scholar]
  • 34.Villani, C.: Inégalités Isopérimétriques dans les espacesmétriques mesurés [d’après F. Cavalletti & A. Mondino] Séminaire BOURBAKI 69m`e année, 2016–2017, no. 1127. Availableat http://www.bourbaki.ens.fr/TEXTES/1127.pdf
  • 35.Wang F-Y. Functional inequalities and spectrum estimates: the infinite measure case. J. Funct. Anal. 2002;194(2):288–310. doi: 10.1006/jfan.2002.3968. [DOI] [Google Scholar]

Articles from Journal of Geometric Analysis are provided here courtesy of Springer

RESOURCES