Compactifications of Complete Riemannian manifolds and Their ...

76 downloads 83 Views 174KB Size Report
Dec 14, 2010 - DG] 14 Dec 2010 ... consider certain invariants defined on the Martin boundary and prove .... reference on Martin compactification is Ancona [A].
arXiv:1012.3131v1 [math.DG] 14 Dec 2010

COMPACTIFICATIONS OF COMPLETE RIEMANNIAN MANIFOLDS AND THEIR APPLICATIONS XIAODONG WANG Dedicated to Professor Richard Schoen in honor of his 60th birthday

1. Introduction To study a noncompact Riemannian manifold, it is often useful to find a compactification or attach a boundary. For example, in hyperbolic geometry a lot of investigation is carried out on the sphere at infinity. An eminent illustration is Mostow’s proof of his rigidity theorem for hyperbolic manifolds [Mo]. More genf is simply connected and nonpositively curved, one can compactify it erally, if M by equivalent geodesic rays and the boundary is a topological sphere, called the geometric boundary. This compactification was first introduced in [EO] and has f is not nonposbeen indispensable in the study of negatively curved manifolds. If M itively curved, then the geometric compactification does not work in general. But there are other compactifications which are useful for various studies. In this short survey, we will discuss some of these compactifcations and the relationships among them. Our discussion will focus on general Riemannian manifolds and therefore we ignore the large literature on compactifications of symmetric spaces (see the book [GJL]). We first discuss the geometric compactification for Cartan-Hadamard manifolds and Gromov hyperbolic spaces in Section 2. In Section 3 we discuss the Martin compactification. In Section 4 we discuss the Busemann compactification. In the last section, we discuss how these compactifications are used. In particular, we consider certain invariants defined on the Martin boundary and prove a comparison inequality using a method of Besson, Courtois and Gallot. It should be noted that when the author showed this inequality to Fran¸cois Ledrappier he was informed that it had been known to Besson, Courtois and Gallot (unpublished). Acknowledgement: It is my great pleasure to dedicate this article to Professor Rick Schoen on the occasion of his 60th birthday. Professor Schoen’s many fundamental contributions to geometry are well-known and greatly admired. I had the good fortune to be one of his students and his influence on my mathematical career, through his teaching, his work and his example, has been tremendous. I wish him good health and more great theorems in the years to come. In preparing this article, I have benefited from several stimulating discussions with Fran¸cois Ledrappier. I want to thank him warmly for teaching me a lot of things. I would also like to thank Jianguo Cao for helpful discussion on the work of Ancona [A] and for bringing to my attention his paper [Cao]. The author was partially supported by NSF grant DMS-0905904. 1

2

XIAODONG WANG

2. The geometric compactification The most familiar compactification is the geometric compactification first introfn be duced by Eberlein and O’Neill [EO] for Cartan-Hadamard manifolds. Let M f a Cartan-Hadamard manifold. We can compactify M using geodesic rays. More precisely, two geodesic rays γ1 and γ2 are said to be equivalent, if d (γ1 (t) , γ2 (t)) f (∞), is is bounded for t ∈ [0, ∞). The set of equivalence classes, denoted by M called the geometric boundary and can be naturally identified with the unit sphere f∗ = M f⊔M f (∞) Sn−1 if we fix a base point. We then obtain a compactification M that is homeomorphic to the closed unit n-ball with the natural ”cone” topology. 2 If the sectional curvature satisfies −b2 ≤ KM f ≤ −a , where a, b > 0, it is proved ∗ α f has a C -structure, where α = a/b. For details, by Anderson and Schoen that M cf. [E, SY]. The same compactification works for the so called Gromov hyperbolic spaces. We first recall one of several equivalent definitions of Gromov hyperbolic spaces. Definition 1. A complete geodesic metric space (X, d) is called Gromov hyperbolic if for some δ > 0 s.t. for all points o, x, y, z ∈ X (x · y)o ≥ min {(x · z)o , (y · z)o } − δ. where we use Gromov products, e.g. (x · y)o =

1 2

(d (o, x) + d (o, y) − d (x, y)).

We make the following remarks Remark 1. In the definition one can take o to be fixed. f is Gromov hyperbolic if the sectional Remark 2. A Cartan-Hadamard manifold M curvature has a negative upper bound.

This concept was introduced by Gromov [G]. For detailed study of Gromov hyperbolic spaces, see the excellent books [BH, GH, O]. It suffices to say that the definition captures the global features of the geometry of a complete simply connected manifold of negative curvature. It is very robust as illustrated by the following remarkable fact. Theorem 1. Let X and Y be geodesic spaces. Suppose f : X → Y is a quasiisometry, i.e. there are L, ε > 0 s.t. for any x1 , x2 ∈ X L−1 d (x1 , x2 ) − ε ≤ d (f (x1 ) , f (x1 )) ≤ Ld (x1 , x2 ) . If Y is Gromov hyperbolic, then so is X.

We will further assume that X is proper. Then we can define the geometric boundary X (∞) for a Gromov hyperbolic spaces in the same way s.t. X = X ⊔ X (∞) with a natural topology is a compact metrizable space. Moreover X (∞) has a canonical quasi-conformal structure. Theorem 2. Let X and Y be proper Gromov hyperbolic spaces. If f : X → Y is a quasi-isometry, then f extends to a homeomorphism f : X (∞) → Y (∞). In fact, the boundary map is furthermore a quasi-conformal map. The boundary map can be described as follows: given the equivalence class ξ ∈ ∂X of a geodesic ray γ : [0, ∞) → X, f ◦ γ is a quasi-geodesic in Y and hence has a well defined end point f ◦ γ (∞) ∈ Y (∞) which is defined to be f (ξ).

COMPACTIFICATIONS OF RIEMANNIAN MANIFOLDS

3

3. The Martin compactification f We can  also  compactify M using all positive harmonic functions. The vector f space H M of harmonic functions with seminorms f compact kukK = sup |u (x)| , K ⊂ M K

  f : u (o) = 1}. It is a convex and is a Frechet space. Let Ko = {u ∈ H+ M   f . We assume that M f is nonparabolic and G(x, y) is the compact set in H M minimal positive Green’s function. Define the Martin kernel k (x, y) =

G(x, y) . G(o, y)

A sequence yi → ∞ is called a Martin sequence if limi→∞ k (x, yi ) converges to a harmonic function. By Harnack inequality and the elliptic theory every sequence yi → ∞ has a Martin subsequence. Two Martin sequences are called equivalent if they have the same harmonic function as limit. The collection of all such equivalence f. It is easy to classes is called the Martin boundary and will be denoted by ∂∆ M f ⊂ Ko is a compact set. The Martin compactification is defined to be see that ∂∆ M c=M f ⊔ ∂M f M

with a natural topology that makes it a compact metrizable space. An excellent reference on Martin compactification is Ancona [A]. f is called minimal if any nonnegDefinition 2. A harmonic function h > 0 on M ative harmonic function ≤ h is proportional to h. Remark 3. If h (o) = 1, then h is minimal iff h is an extremal point of Ko .

f. It is proved that all minimal harmonic function h with h (o) = 1 belong to ∂∆ M Therefore we can introduce the following Definition 3. The minimal Martin boundary of M is

f = {h ∈ Ko : h is minimal}. ∂∗M

f ⊂ ∂M f is at least a Borel subset (cf. [A]). According to a Moreover ∂ ∗ M theorem of Choquet ([A]), for any positive harmonic function h there is a unique f such that Borel measure µh on ∂ ∗ M Z ξ (x) dµh (ξ) . h (x) = f ∂∗M

Let ν be the measure corresponding to the harmonic function 1. Thus Z (3.1) 1= ξ (x) dν (ξ) . f ∂∗M

n o f with ν x = ξ (x) ν are called the The family of probability measures ν x : x ∈ M   f we get a bounded harmonic function harmonic measures. For f ∈ L∞ ∂ ∗ M Z f (ξ) ξ (x) dν (ξ) . Hf (x) = f ∂∗M

4

XIAODONG WANG

  f and the space of bounded harThis defines an isomorphism between L∞ ∂ ∗ M f. monic functions on M The study of the Martin compactification is closely related to the study of Browf. For simplicity we further add a mild condition that the Ricci nian motion on M curvature is bounded from toensure stochastic completeness. Therefore we  below  f f with a family of probability measure have the sample space Ω M = C R+ , M n o f s.t. for any 0 < t1 < · · · < tk and open sets U1 , · · · , Uk Px : x ∈ M Px {ω ∈ Ω (M ) : ω (t1 ) ∈ U1 , · · · ω (tk ) ∈ Uk } Z pt1 (x, y1 ) pt2 −t1 (y1 , y2 ) · · · ptk −tk−1 (yk−1 , yk ) dy1 × · · · dyk . = U1 ×···×Uk

f. For each t ≥ 0 we have a random variable Here pt(x,  y) is the heat kernel on M f →M f which is simply the position at t, i.e. Xt (ω) = ω (t). It is an Xt : Ω M intriguing and important problem to understand the asymptotic behavior for ω (t) as t → ∞. The answer is closely related to the Martin boundary. f and for Px -a.e. ω ∈ Ω (M ), Xt (ω) admits a Theorem 3. (1) For any x ∈ M ∗f limit X∞ (ω) ∈ ∂ M as t → ∞, i.e. lim k (x, ω (t))

t→∞

exists and is a minimal harmonic function. (2) Under Px , the distribution of X∞ is ν x , i.e. (X∞ )∗ Px = ν x . For detailed discussion, see Ancona [A]. f with sectional curvature bounded between For a Cartan-Hadamard manifold M two negative constants, Anderson and Schoen [AS] proved that the Martin boundary is homeomorphic to the geometric boundary. f is a Cartan-Hadamard manifold with whose sectional Theorem 4. Suppose that M 2 curvature satisfies −b ≤ K ≤ −a2 < 0. Then there exists a natural homeomorf→M f (∞) between the Martin boundary and the geometric boundary. phism Φ : ∂ M −1 Moreover, Φ is H¨ older continuous. f = ∂∗M f in this case. From the proof, it is also clear that ∂ M This theorem was generalized by Ancona who proved

f Theorem  5. (Ancona [A, Theorem 6.2]) Suppose that M is Gromov hyperbolic f > 0. Then the Martin boundary is homeomorphic to the geometric and λ0 M f = ∂∗M f. boundary. Moreover ∂ M   f is the bottom of the L2 spectrum of M f, i.e. In the statement, λ0 M R 2   f |∇u| f R λ0 M = inf M , 2 fu M

where the infimum is taken over all smooth functions with compact support. It 2 fn is easy  to see that for a Cartan-Hadamard manifold M with K ≤ −a , we have 2 2 f ≥ (n − 1) a /4. λ0 M

COMPACTIFICATIONS OF RIEMANNIAN MANIFOLDS

5

4. The Busemann boundary f. Instead of harmonic functions, one can use distance functions to compactify M This leads to the Busemann compactification, first introduced by Gromov in [BGS]. f and define, for x ∈ M f the function ξx (z) on M f by: Fix a point o ∈ M ξx (z) = d(x, z) − d(x, o).

The assignment x 7→ ξx is continuous, one-to-one and takes values in a relatively compact set of functions for the topology of uniform convergence on compact subsets f. The Busemann compactification M c of M f is the closure of M f for that of M c topology. The space M is a compact separable space. The Busemann boundary c := M c\M f is made of 1-Lipschitz continuous functions ξ on M f such that ξ(o) = 0 ∂M f and there exists a sequence {ak } ⊂ M s.t. d (o.ak ) → ∞ and ξ (x) = lim d(ak , x) − d(ak , o), k→∞

c are called where the convergence is uniform over compact sets. Elements of ∂ M horofunctions. We note that this compactification works for any proper metric space X (cf. [KL]) But in general, X may fail to be open in its Busemann compactification. This pathology does not happen for Riemannian manifolds, i.e. we have f is open in its Busemann compactification M c. Hence the BuseProposition 1. M c mann boundary ∂ M is compact.

For proof see [LW1]. For a Cartan-Hadamard manifold, the Busemann compactification coincides with the geometric compactification. More precisely, f be a Cartan-Hadamard manifold and {ak } a sequence in Proposition 2. Let M f M s.t. d (o.ak ) → ∞. Let σk be the unique geodesic ray from o to ak . Then ξak converges to a horofunction ξ iff σk converges to a ray σ. Furthermore, we have ξ (x) = lim d (σ (t) , x) − t. t→∞

For proof see Ballmann [B] (p30). Recently, the Busemann compactification has found to be very useful in various questions, cf. [KL, L1, LW1]. We first describe the application in [LW1]. Suppose f its universal covering (noncompact). M is a compact Riemannian manifold and M f isometrically. Observe that we Let G be the fundamental group of M acting on M f c, in such a way that for ξ may extend by continuity the action of G from M to M c in M and g in G, g.ξ(z) = ξ(g −1 z) − ξ(g −1 o).

The volume entropy of M is defined to be the limit

ln volBM f (x, r) , r where BM f (x, r) is the ball of radius r centered at x in the universal covering space f M . This important invariant was introduced by Manning [Ma] who proved f, (1) the limit exists and is independent of the center x ∈ M v (g) = lim

r→∞

6

XIAODONG WANG

(2) v ≤ H, the topological entropy of the geodesic flow on the unit tangent bundle of M , (3) v = H if M is nonpositively curved. In [LW1], we extended the classical theory of Patterson-Sullivan measure by c. constructing a family of measures on the Busemann boundary ∂ M n o f of finite measures on the BuseTheorem 6. There exists a family νx : x ∈ M c s.t. mann boundary ∂ M f the two measures νx and νy are equivalent with (1) for any pair x, y ∈ M dνx (ξ) = e−v(ξ(x)−ξ(y)) , dνy

f (2) for any g ∈ G and x ∈ M

g∗ νx = νgx .

This family of measures plays a crucial role in the proofs of the following rigidity results involving the volume entropy. Theorem 7. Let M n be a compact Riemannian manifold with Ric ≥ − (n − 1). Then the volume entropy satisfies v ≤ n − 1 and equality holds iff M is hyperbolic. Remark 4. This result was proved by Knieper [Kn] under the additional assumption that M is negatively curved.   f ≤ v 2 /4, we deduce As a corollary, in view of the well-known inequality λ0 M the following result which was previously proved in [W] by a different method.

n Theorem  8. Let M be a compact Riemannian manifold with Ric ≥ − (n − 1). f ≤ (n − 1)2 /4 and equality holds iff M is hyperbolic. Then λ0 M

Theorem 9. Let M be a compact K¨ ahler manifold with dimC M = m. If the bisectional curvature KC ≥ −2, then the volume entropy satisfies v ≤ 2m. Moreover equality holds iff M is complex hyperbolic (normalized to have constant holomorphic sectional curvature −4).

Theorem 10. Let M be a compact quaternionic K¨ ahler manifold of dim = 4m with m ≥ 2 and scalar curvature −16m (m + 2). Then the volume entropy satisfies v ≤ 2 (2m + 1). Moreover equality holds iff M is quaternionic hyperbolic. We refer to the original paper [LW1] for details. More recently, we can prove some pinching theorems using our method. The first step is the following rigidity result for C 1,α metrics. Theorem 11. Let M n be a (smooth) compact smooth manifold and g a C 1,α Riemannian metric. Suppose that gi is a sequence of smooth Riemannian metrics on M s.t. (1) Ric (gi ) ≥ − (n − 1) for each i, (2) gi → g in C 1,α norm as i → ∞, (3) the volume entropy v (gi ) → n − 1 as i → ∞. Then g is hyperbolic. From this result, we then deduce the following pinching theorem.

COMPACTIFICATIONS OF RIEMANNIAN MANIFOLDS

7

Theorem 12. There exists a positive constant ε = ε (n, D) s.t. if (M n , g) is a compact Riemannian manifold of dimension n satisfying the following conditions • g has negative sectional curvature, • Ric (g) ≥ − (n − 1), • diam (M, g) ≤ D, • the volume entropy v (g) ≥ n − 1 − ε, then M is diffeomorphic to a hyperbolic manifold (X, g0 ). Moreover, the GromovHausdorff distance dGH (M, X) ≤ α (ε) → 0 as ε → 0. We note that this theorem was established by Courtois [C] (unpublished) in 2000 using the Cheeger-Colding theory. Our proof is different and simpler. The details will appear in [LW2]. In [L1] Ledrappier studied another fundamental invariant: the linear drift (introduced by Guivarc’h [Gu]) which is the following limit for almost every path ω f of the Brownian motion on M

1 d (ω (0) , ω (t)) . t If M is negatively curved, Kaimanovich [K1] established a remarkable integral forf be the geometric boundary of M f. As usual we fix a base point mula for l. Let ∂ M f. Recall that there is a homeomorphism Φ from ∂ M f to the Martin boundary o∈M f by the theorem of Anderson-Schoen. For each ξ ∈ ∂ M f ∂∆ M  , hξ =  Φ (ξ) is the ∗ f s.t. hξ (o) = 1 and hξ ∈ C M f \ {ξ} with boundunique harmonic function on M l = lim

t→∞

ary value zero. With these notations, the Kaimanovich formula can be written as  Z Z h∇Bξ , ∇ ln hξ i (x) hξ (x) dν (ξ) dm (x) , l=− M

f ∂M

f defined by (3.1) and m is the normalized where ν is the harmonic measure on ∂ M Lebesgue measure on M . The main result in [L1] is a similar integral formula for l in the general case. The key step is to construct certain measures on the Busemann c. boundary ∂ M 5. A comparison theorem

In this section we discuss why compactifications are useful. The basic principle is that often times a geometric object is much simpler near infinity. When we look at it on the boundary, we capture its essential features while all the background noise dies off. The first illustration of this principle is perhaps Mostow’s rigidity theorem [Mo]: If f : M → N is a (smooth) homotopy equivalence between two compact hyperbolic manifolds then f is homotopic to an isometry. In the proof Mostow f→N e between the universal coverings (which are both considers the lifting fe : M Hn in this case) which is a quasi-isometry. Then feextends to a homeomorphism f → ∂N e between the boundaries. Using the theory of quasi-conformal maps f : ∂M f ergodically, Mostow shows and the fact that the fundamental group acts on ∂ M that f is in fact a Mobius transformation. More recently, in a seminal paper [BCG1], Besson, Courtois and Gallot proved the following theorem which implies the Mostow rigidity theorem in the rank one cases.

8

XIAODONG WANG

Theorem 13. Let (N n , g0 ) be a compact locally symmetric space of negative curvature. Let M n be another compact manifold and f : M → N is a continuous map of nonzero degree. Then for any metric g on M n n (1) v (g) vol (M, g) ≥ |deg f | v (g0 ) vol (N, g0 ); (2) the equality holds iff f is homotopic to an covering map. f into a Hilbert space and a calibration argument. The proof involves embedding M In a later paper [BCG2], the same authors gave a very elegant and simpler proof of their theorem under the additional assumption that g is also negatively curved. In this second approach, the Patterson-Sullivan measure on the geometric boundary f plays a fundamental role. Their method is geometric and flexible and we will ∂M apply it in a slight different situation. f → M its universal Let (M n , g) be a compact Riemannian manifold and π : M f. Let ∂ M f be the Martin boundary. covering. We pick a base point o ∈ M

Definition 4. For any p > 0 let p Z Z 2 |∇ log ξ (x)| ξ (x) dν (ξ) dm (x) . βp (g) = M

Similarly we can consider Z e βp (g) =

M

f ∂M

Z

f ∂M

|∇ log ξ (x)|

2p

ξ (x) dν (ξ) dm (x) .

We have βp (g) ≤ βep (g) by the H¨ older inequality. When p = 1, βp (g) = βep (g) is the Kaimanovich entropy, an invariant of fundamental importance. This was introduced by Kaimanovich [K1]. We summarize its main properties: R f (1) [K1] β1 = limt→∞ − 1t M f pt (x, y) log pt (x, y) dy for any x ∈ M ; f (2) [K1] β1 > 0 iff M has nonconstant bounded harmonic functions. (3) [L1, L2] 4λ0 ≤ β1 ≤ v 2 , where v is the volume entropy and λ0 is the bottom f. of the L2 spectrum of M

Let (N n , g0 ) be a compact locally symmetric space of negative curvature. Theorem 13 says that among all metrics g on N with vol (N, g) = vol (N, g0 ) the metric g0 has the smallest volume entropy. A natural question is whether the same is true for the Kaimonovich entropy. Problem 1. Let (N n , g0 ) be a compact locally symmetric space of negative curvature. Is it true that for any metric g n/2

β1 (g)

n/2

vol (N, g) ≥ β1 (g0 )

vol (N, g0 )?

We do not know the answer to this question. What we can prove is the following result which gives an affirmative answer to the same question for βn/2 . Theorem 14. Let (N n , g0 ) be a compact locally symmetric space of negative curvature. Let M n be another compact manifold and f : M → N a (smooth) homotopy equivalence. Then for any metric g on M (1) βn/2 (g) vol (M, g) ≥ βn/2 (g0 ) vol (N, g0 ); (2) the equality holds iff f is homotopic to an isometry.

COMPACTIFICATIONS OF RIEMANNIAN MANIFOLDS

9

As a consequence we have the following result which is also an easy corollary of Theorem 13. Corollary 1. If g0 is real hyperbolic and g satisfies Ric ≥ − (n − 1), then vol (M, g) ≥ vol (N, g0 ). Moreover, equality holds iff g is also hyperbolic. This follows from the sharp gradient estimate. Proposition 3. (Li-J. Wang [LiW, Lemma 2.1]) Let N n be a complete manifold with Ric ≥ − (n − 1). If u is a positive harmonic function on N , then |∇ log u| ≤ n − 1. We now prove Theorem 14. The homotopy equivalence f : (M, g) → (N, g0 ) induces an isomorphism ρ : Γ := π1 (M ) → π1 (N ). We view the fundamental groups as groups of deck transformations acting on the universal covering manifolds. f→N e which is Γ-equivariant , i.e. for any Lifting f we obtain a smooth map fe : M γ∈Γ fe(γ · x) = ρ (γ) · fe(x) .

f is Gromov hyperbolic as N e is. Hence fe This is a quasi-isometry and hence M f e extends to a homeomorphism f : ∂ M →  ∂ N between the boundaries. By a theorem   e > 0 and the two fundamental groups are f > 0 since λ0 N of Brooks [Br], λ0 M f is also the Martin boundary of M f and let isomorphic. Therefore, by Theorem 5 ∂ M n o f be the harmonic measures. We now define a new map Fe : M f→N e νx : x ∈ M applying the construction in [BCG2]: Fe (x) is the barcenter of the measure f ∗ ν x e , i.e. Fe (x) is the unique minimum point of the following function on N e on ∂ N Z  Bθ (y) d f ∗ ν x (θ) , y→ e ∂N

e associated to θ ∈ ∂ N e . For detailed where Bθ is the Busemann function on N discussion of the barcenter see [BCG2]. Note that this map is well defined as the support ν x always has more that two points. By the implicit function theorem, it is easy to show that Fe is smooth. Moreover it is Γ-equivariant and hence yields a smooth map F : M → N . What remains is to estimate the Jacobian of this map. By the definition of F we have Z

e ∂N

=

Z

   dBθ Fe (x) (·) d f ∗ νx (θ)

f ∂M

= 0.

  dBf (ξ) Fe (x) (·) ξ (x) dν (ξ)

Differentiating in x we get Z    D2 Bf (ξ) Fe (x) Fe∗ (x) (·) , · ξ (x) dν (ξ) f ∂M Z   dBf (ξ) Fe (x) (·) dξ (·) dν (ξ) . =− f ∂M

10

XIAODONG WANG

e We define the following quadratic form K and H on TF (x) N Z     D2 Bθ Fe (x) (u, u) ξ (x) dν (ξ) , g0 KFe(x) (u) , u = f ∂M    Z  dBf (ξ) Fe (x) (u)2 ξ (x) dν (ξ) . g0 HFe(x) (u) , u = f ∂M

f, u ∈ T e N e Then for any v ∈ Tx M F (x)   g0 KFe(x) (F∗ (x) (v)) , u 

≤ g0 HFe(x) (u) , u

1/2

Z

 1/2 Z = g0 HFe(x) (u) , u

f ∂M

Therefore

|det K| JacFe (x) ≤

1 nn/2

2

|h∇ξ (x) , vi| dν (ξ) ξ (x) f ∂M

|h∇ log ξ (x) , vi| ξ (x) dν (ξ)

|det H|1/2

By [BCG1, Appendix B] we obtain

2

!1/2

Z

f ∂M

1/2

.

n/2 |∇ log ξ (x)|2 ξ (x) dν (ξ) .

Z n/2 1 2 |∇ log ξ (x)| ξ (x) dν (ξ) , n (n + d − 2) f ∂M   e , g0 is the real, complex, quaternionic hyperbolic where d = 1, 2, 4 or 8 when N space or the Cayley hyperbolic plane, respectively. Integrating over M yields n/2 Z Z vol (M, g) 2 vol (N, g0 ) ≤ |∇ log ξ (x)| ξ (x) dν (ξ) dm (x) . n (n + d − 2) M f ∂M JacFe (x) ≤

n

This proves the inequality as βn/2 (g0 ) = (n + d − 2) . If equality holds, it is easy to see that F has to be an isometry up to a scaling. References ´ A. Ancona, Th´ eorie du potentiel sur les graphes et les vari´ et´ es. Ecole d’´ et´ e de Probabilit´ es de Saint-Flour XVIII—1988, 1–112, Lecture Notes in Math., 1427, Springer, Berlin, 1990. [AS] M. Anderson and R. Schoen, Positive harmonic functions on complete manifolds of negative curvature. Ann. of Math. (2) 121 (1985), no. 3, 429–461. [B] W. Ballmann, Lectures on spaces of Nonpositive Curvature. Birkh¨ auser, Boston, 1985. [BGS] W. Ballmann, M. Gromov and V. Schroeder, Manifolds of Nonpositive Curvature. With an appendix by Misha Brin. Birkh¨ auser Verlag, Basel, 1995. [Br] R. Brooks, The fundamental group and the spectrum of the Laplacian. Comment. Math. Helv. 56 (1981), no. 4, 581–598. [BH] M. Bridson and A. Haefliger, Metric spaces of non-positive curvature. Grundlehren der Mathematischen Wissenschaften, 319. Springer-Verlag, Berlin, 1999. [BCG1] G. Besson, G. Courtois and S. Gallot, Entropies et rigidit´ es des espaces localement sym´ etriques de courbure strictement n´ egative. Geom. Funct. Anal. 5 (1995), no. 5, 731–799. [BCG2] G. Besson, G. Courtois and S. Gallot, Minimal entropy and Mostow’s rigidity theorems. Ergodic Theory Dynam. Systems 16 (1996), no. 4, 623–649. [Cao] J. Cao, Cheeger isoperimetric constants of Gromov-hyperbolic spaces with quasi-poles. Commun. Contemp. Math. 2 (2000), no. 4, 511–533.

[A]

COMPACTIFICATIONS OF RIEMANNIAN MANIFOLDS

[C] [E] [EO] [GH] [G] [Gu] [GJL] [K1] [K2] [KL] [Kn] [L1] [L2] [LW1] [LW2] [LiW] [Ma] [Mo] [O] [SY] [V] [W]

11

G. Courtois, Pincement spectral des vari´ et´ es hyperboliques. Preprint, 2000. P. Eberlein, Geometry of nonpostively curved manifolds. Chicago Lectures in Mathematics. University of Chicago Press, Chicago, IL, 1996. P. Eberlein and B. O’Neill, Visibility manifolds. Pacific J. Math. 46 (1973), 45–109. ´ Ghys and P. de la Harpe (ed), Sur les groupes hyperboliques d’apr` E. es Mikhael Gromov. Progress in Mathematics, 83. Birkh¨ auser, Boston, MA, 1990. M. Gromov, Hyperbolic groups. Essays in group theory, 75–263, Math. Sci. Res. Inst. Publ., 8, Springer, New York, 1987. Y. Guivarc’h, Sur la loi des grands nombres et le rayon spectral d’une marche al´ eatoire. Ast´ erisque, 74 (1980) 47–98. Y. Guivarc’h, L. Ji and J. C. Taylor, Compactifications of Symmetric Spaces. Birkhauser, Bostorn, MA, 1998. V. Kaimanovich, Brownian motion and harmonic functions on covering manifolds: an entropy approach. Soviet Math. Dokl. 33 (1986) 812-816. V. Kaimanovich, The Poisson formula for groups with hyperbolic properties. Ann. of Math. 152 (1990), 659-692. A. Karlsson and F. Ledrappier, On laws of large numbers for random walks. Ann. Probab. 34 (2006), no. 5, 1693–1706. G. Knieper, Spherical means on compact Riemannian manifolds of negative curvature. Differential Geom. Appl. 4 (1994), no. 4, 361–390. F. Ledrappier, Linear drift and entropy for regular covers. arXiv math.ds 0910.1425, to appear in Geom. Func. Anal. F. Ledrappier, Harmonic measures and Bowen-Margulis measures. Israel J. Math. 71 (1990) 275-287. F. Ledrappier and X. Wang, An integral formula for the volume entropy with applications to rigidity. arXiv math.dg 0911.0370, to appear in J. Diff. Geom. F. Ledrappier and X. Wang, Pinching theorems for the volume entropy. Preprint. P. Li and J. Wang, Complete manifolds with positive spectrum. II. J. Differential Geom. 62 (2002), no. 1, 143–162. A. Manning, Topological entropy for geodesic flows. Ann. of Math. (2) 110 (1979), no. 3, 567–573. G. D. Mostow, Quasi-conformal mappings in n-space and the rigidity of hyperbolic space ´ forms. Inst. Hautes Etudes Sci. Publ. Math. No. 34 1968 53–104. K. Ohshika, Discrete groups, Translations of Mathematical Monographs, 207. American Mathematical Society, Providence, RI, 2002. R. Scheon and S.-T. Yau, Lectures on Differential Geometry. International Press, 1994. N. Varopoulos, Information theory and harmonic functions. Bull. Sci. Math. (2) 110 (1986), no. 4, 347–389. X. Wang, Harmonic functions, entropy, and a characterization of the hyperbolic space. J. Geom. Anal. 18 (2008), no. 1, 272–284.

Department of Mathematics, Michigan State University, East Lansing, MI 48824 E-mail address: [email protected]