Home Total mean curvatures of Riemannian hypersurfaces
Article Open Access

Total mean curvatures of Riemannian hypersurfaces

  • Mohammad Ghomi EMAIL logo and Joel Spruck
Published/Copyright: January 10, 2023

Abstract

We obtain a comparison formula for integrals of mean curvatures of Riemannian hypersurfaces via Reilly’s identities. As applications, we derive several geometric inequalities for a convex hypersurface Γ in a Cartan-Hadamard manifold M . In particular, we show that the first mean curvature integral of a convex hypersurface γ nested inside Γ cannot exceed that of Γ , which leads to a sharp lower bound for the total first mean curvature of Γ in terms of the volume it bounds in M in dimension 3. This monotonicity property is extended to all mean curvature integrals when γ is parallel to Γ , or M has constant curvature. We also characterize hyperbolic balls as minimizers of the mean curvature integrals among balls with equal radii in Cartan-Hadamard manifolds.

1 Introduction

Total mean curvatures of a hypersurface Γ in a Riemannian n -manifold M are integrals of symmetric functions of its principal curvatures. These quantities are known as quermassintegrals or mixed volumes when Γ is convex and M is the Euclidean space. They are fundamental in geometric variational problems, as they feature in Steiner’s polynomial, Brunn-Minkowski theory, and Alexandrov-Fenchel inequalities [11,14,19,20], which were all originally developed in Euclidean space. Extending these notions to Riemannian manifolds has been a major topic of investigation. In particular, total mean curvatures have been studied extensively in hyperbolic space in recent years [2,22, 23,24]. Here, we study these integrals in the broader setting of Cartan-Hadamard spaces, i.e., complete simply connected manifolds of nonpositive curvature and generalize a number of inequalities that had been established in Euclidean or hyperbolic space.

The main result of this article, Theorem 3.1, expresses the difference between the total r th mean curvatures of a pair of nested hypersurfaces Γ and γ in a Riemannian manifold M in terms of the sectional curvatures of M and the principal curvatures of a family of hypersurfaces that fibrate the region between Γ and γ . This formula simplifies when r = 1 , Γ and γ are parallel, or M has constant curvature, leading to a number of applications. In particular, we establish the monotonicity property of the total first mean curvature for nested convex hypersurfaces in Cartan-Hadamard manifolds (Corollary 4.1). This leads to a sharp lower bound in dimension 3 for the total first mean curvature in terms of the volume bounded by Γ (Corollary 4.3), which generalizes a result of Gallego-Solanes in hyperbolic 3-space [10, Cor. 3.2]. We also extend to all mean curvatures some monotonicity results of Schroeder-Strake [21] and Borbely [4] for total Gauss-Kronecker curvature (Corollaries 4.4 and 4.5). Finally, we include a characterization of hyperbolic balls as minimizers of total mean curvatures among balls of equal radii in Cartan-Hadamard manifolds (Corollary 4.7).

Theorem 3.1 is a generalization of the comparison result we had obtained earlier in [13] for the Gauss-Kronecker curvature, motivated by Kleiner’s approach to the Cartan-Hadamard conjecture on the isoperimetric inequality [16]. Similar to [13], our starting point here, in Section 2, will be an identity (Lemma 2.1) for the divergence of Newton operators, which were developed by Reilly [18,17] to study the invariants of Hessians of functions on Riemannian manifolds. This formula, together with Stokes’ theorem, leads to the proof of Theorem 3.1 in Section 3. Then, in Section 4, we develop the applications of that result.

2 Newton operators

Throughout this work, M denotes an n -dimensional Riemannian manifold with metric , and covariant derivative . Furthermore, u is a C 1 , 1 function on M . In particular, u is twice differentiable at almost every point p of M , and the computations below take place at such a point. The gradient of u is the tangent vector u T p M given by u ( p ) , X X u for all X T p M . The Hessian operator 2 u : T p M T p M is the self-adjoint linear map given by 2 u ( X ) X ( u ) . The symmetric elementary functions σ r : R k R , for 1 r k , and x = ( x 1 , , x k ) are defined by

σ r ( x ) i 1 < < i r x i 1 x i r .

We set σ 0 1 and σ r 0 for r k + 1 by convention. Let λ ( 2 u ) ( λ 1 , , λ n ) denote the eigenvalues of 2 u . Then, we set

σ r ( 2 u ) σ r ( λ ( 2 u ) ) .

These functions form the coefficients of the characteristic polynomial

P ( λ ) det ( λ I 2 u ) = i = 0 n ( 1 ) i σ i ( 2 u ) λ n i .

Let δ j 1 j m i 1 i m be the generalized Kronecker tensor, which is equal to 1 ( 1 ) if i 1 , , i m are distinct and ( j 1 , , j m ) is an even (odd) permutation of ( i 1 , , i m ) ; otherwise, it is equal to 0. Then [18, Prop. 1.2(a)],

(1) σ r ( 2 u ) = 1 r ! δ j 1 j r i 1 i r u i 1 j 1 u i r j r ,

where u i j i j u denote the second partial derivatives of u with respect to an orthonormal frame E i T p M , which we extend to an open neighborhood of p by parallel translation along geodesics. So E i E j = 0 at p . We call E i a local parallel frame centered at p and set i E i , i j i j . Each of the indices in (1) ranges from 1 to n , and we employ Einstein’s convention by summing over repeated indices throughout the article. The Newton operators T r u : T p M T p M [17,18] are defined recursively by setting T 0 u I , the identity map, and for r 1 ,

(2) T r u σ r ( 2 u ) I T r 1 u 2 u = i = 0 r ( 1 ) i σ i ( 2 u ) ( 2 u ) r i .

Thus, T r u is the truncation of the polynomial P ( 2 u ) obtained by removing the terms of order higher than r . In particular, T n u = P ( 2 u ) . So, by the Cayley-Hamilton theorem, T n u = 0 . Consequently, when 2 u is nondegenerate, (2) yields that

(3) T n 1 u = σ n ( 2 u ) ( 2 u ) 1 = det ( 2 u ) ( 2 u ) 1 = T u ,

where T u is the Hessian cofactor operator discussed in [13, Sec. 4]. See [18, Prop. 1.2] for other basic identities that relate σ and T . In particular, by [18, Prop. 1.2(c)], we have Trace ( T r u 2 u ) = ( r + 1 ) σ r + 1 ( 2 u ) . So, by Euler’s identity for homogeneous polynomials,

(4) ( T r u ) i j u i j = Trace ( T r u 2 u ) = ( r + 1 ) σ r + 1 ( 2 u ) = σ r + 1 ( 2 u ) u i j u i j .

Thus, it follows from (1) that

(5) ( T r u ) i j = σ r + 1 ( 2 u ) u i j = 1 r ! δ j j 1 j r i i 1 i r u i 1 j 1 u i r j r .

Furthermore, by [17, Prop. 1(11)] (note that the sign of the Riemann tensor R in [17] is opposite to the one in this article), we have

(6) ( div ( T r u ) ) j = 1 ( r 1 ) ! δ j j 1 j r i i 1 i r u i 1 j 1 u i r 1 j r 1 R i j r i r k u k ,

where R i j k l R ( E i , E j , E k , E ) = i j E k j i E k , E . Another useful identity [17, p. 462] is

(7) div ( T r u ( u ) ) = T r u , 2 u + div ( T r u ) , u ,

where , here indicates the Frobenius inner product (i.e., A , B A i j B i j for any pair of matrices of the same dimension). The divergence of T r u may be defined by virtually the same argument used for T u in [13, Sec. 4] to yield the following generalization of [13, (14)]:

(8) ( div ( T r u ) ) j = i ( T r u ) i j .

Recall that T u = T n 1 u by (3). Furthermore, T n u = 0 as we mentioned earlier. Thus, the following observation generalizes [13, Lem. 4.2].

Lemma 2.1

div T r 1 u u u r = div ( T r 1 u ) , u u r + r T r u ( u ) , u u r + 2 .

Proof

By Leibniz rule and (8), we have

div T r 1 u u u r = i ( T r 1 u ) i j u j u r = div ( T r 1 u ) , u u r + ( T r 1 u ) i j u i j u r r u j u u i u r + 2 ,

where the computation to obtain the second term on the right is identical to the one performed earlier in [13, Lem. 4.2]. To develop this term further, note that by (2)

( T r 1 u ) i j u i = σ r ( 2 u ) δ j ( T r u ) j ,

which in turn yields

( T r 1 u ) i j u i u j u u 2 = σ r ( 2 u ) ( T r u ) i j u i u j u 2 .

Hence,

( T r 1 u ) i j u i j u r r u j u u i u r + 2 = r σ r ( 2 u ) u r r u r σ r ( 2 u ) ( T r u ) i j u i u j u 2 = r ( T r u ) i j u i u j u r + 2 ,

which completes the proof.□

Below we assume, as was the case in [13, Sec. 4], that all local computations take place with respect to a principal curvature frame E i T p M of u , which is defined as follows. Assuming u ( p ) 0 , we set E n u ( p ) / u ( p ) , and let E 1 , , E n 1 be the principal directions of the level set of u passing through p . Then, we extend E i to a local parallel frame near p . The first partial derivatives of u with respect to E i , u i i u , satisfy

(9) u i = 0 ; for i n , and u n = u .

Furthermore, for the second partial derivatives, u i j = i j u , we have

(10) u i j = 0 , for i j n 1 , and u i i u κ i u , for i n ,

where κ 1 u , , κ n 1 u are the principal curvatures of level sets of u with respect to E n , i.e., they are eigenvalues corresponding to E 1 , , E n 1 of the shape operator X X ν on the tangent space of level sets of u , where ν u / u . We set κ u ( κ 1 u , , κ n 1 u ) . So σ r ( κ u ) is the r th mean curvature of the level set of u at p . In particular, σ n 1 ( κ u ) is the Gauss-Kronecker curvature of the level sets. The next observation generalizes [13, Lem. 4.1].

Lemma 2.2

σ r ( κ u ) = T r u ( u ) , u u r + 2 .

Proof

(5) together with (9) and (10) yields that

( T r u ) i j u i u j u r + 2 = 1 r ! δ j j 1 j r i i 1 i r u i 1 j 1 u i r j r u i u j u r + 2 = 1 r ! δ n j 1 j r n i 1 i r u i 1 j 1 u i r j r u r u n 2 u 2 = 1 r ! δ n i 1 i r n i 1 i r κ i 1 u κ i r u .

3 Comparison formula

Here, we establish the main result of this work. For a C 1 , 1 hypersurface Γ in a Riemannian n -manifold M , oriented by a choice of normal vector field ν , and 0 r n 1 , we let

r ( Γ ) Γ σ r ( κ )

be the total r th mean curvature of Γ , where κ ( κ 1 , , κ n 1 ) denotes principal curvatures of Γ with respect to ν . Note that 0 ( Γ ) = Γ , the volume of Γ , since σ 0 = 1 , and n 1 ( Γ ) is the total Gauss-Kronecker curvature of Γ (denoted by G ( Γ ) in [13]). A domain Ω M is an open set with a compact closure cl ( Ω ) . If Γ bounds a domain Ω , then by convention we set 1 ( Γ ) Ω , the volume of Ω . The following theorem generalizes [13, Thm. 4.7], where this result had been established for r = n 1 . It also uses less regularity than was required in [13, Thm. 4.7].

Theorem 3.1

Let Γ and γ be closed C 1 , 1 hypersurfaces in a Riemannian n-manifold M bounding domains Ω and D, respectively, with cl ( D ) Ω . Suppose there exists a C 1 , 1 function u on cl ( Ω D ) with u 0 , which is constant on Γ and γ . Let κ u ( κ 1 u , , κ n 1 u ) be the principal curvatures of level sets of u with respect to E n u / u , and let E 1 , , E n 1 be the corresponding principal directions. Then, for 0 r n 1 ,

r ( Γ ) r ( γ ) = ( r + 1 ) Ω D σ r + 1 ( κ u ) + Ω D κ i 1 u κ i r 1 u K i r n + 1 u κ i 1 u κ i r 2 u u i r 1 R i r i r 1 i r n ,

where u i E i u , R i j k l = R ( E i , E j , E k , E l ) are components of the Riemann curvature tensor of M , K i j = R i j i j is the sectional curvature, and the summations take place over distinct values of 1 i 1 , , i r n 1 , with i 1 < < i r 1 in the first sum and i 1 < < i r 2 in the second sum.

Proof

By Lemmas 2.1 and 2.2,

(11) div T r u u u r + 1 = ( r + 1 ) σ r + 1 ( κ u ) + div ( T r u ) , u u r + 1 .

By Stokes’ theorem and Lemma 2.2,

Ω D div T r u u u r + 1 = Γ γ T r u u u r + 1 , u u = r ( Γ ) r ( γ ) .

So integrating both sides of (11) yields

r ( Γ ) r ( γ ) = ( r + 1 ) Ω D σ r + 1 ( κ u ) + Ω D div ( T r u ) , u u r + 1 .

Using (6) and (9), we have

div ( T r u ) , u u r + 1 = 1 ( r 1 ) ! δ j j 1 j r i i 1 i r u i 1 j 1 u i r 1 j r 1 R i j r i r k u k u j u r + 1 = 1 ( r 1 ) ! δ n j 1 j r i i 1 i r u i 1 j 1 u u i r 1 j r 1 u R i j r i r n .

The last expression may be written as the sum of two components, A and B , which consist of terms with i = n and i n , respectively. Note that we may assume j 1 , , j r n , for otherwise δ n j 1 j r i i 1 i r = 0 . To compute A , note that if i = n , then for δ n j 1 j r i i 1 i r not to vanish, we must have i 1 , , i r n . Then, by (10), u i k j k = 0 unless i k = j k , which yields that

A = 1 ( r 1 ) ! δ n i 1 i r n i 1 i r u i 1 i 1 u u i r 1 i r 1 u R n i r i r n = κ i 1 u κ i r 1 u K i r n ,

where the sum ranges over all distinct values of 1 i 1 , , i r n 1 , with i 1 < < i r 1 as desired. To find B note that if i n , then for δ n j 1 j r i i 1 i r not to vanish, we must have i k = n for some 1 k r . If k = r , then R i j r i r n = R i j k n n = 0 . In particular, B = 0 when r = 1 . Now assume that r 2 . Then, we may assume that k r , or i r n . Then, by (10), u i r j r = 0 unless i r = j r . So, we may assume that i r = j r for r k , which in turn implies that j k = i . Thus, B = k = 1 r 1 B k , where

B k = 1 ( r 1 ) ! δ n i 1 i k 1 i i k + 1 i r i i 1 i k 1 n i k + 1 i r u i 1 i 1 u u i k 1 i k 1 u u n i u u i k + 1 i k + 1 u u i r 1 i r 1 u R i i r i r n = 1 ( r 1 ) κ i 1 u κ i k 1 u u i u κ i k + 1 u κ i r 1 u R i i r i r n ,

since u n = u . Here, the sum ranges over all distinct indices 1 i , i 1 , , i k 1 , i k , , i r n 1 , with i 1 < < i k 1 < i k + 1 < < i r 1 . Note that B 1 = = B r 1 . Thus,

B = ( r 1 ) B r 1 = 1 u κ i 1 u κ i r 2 u u i R i r i i r n ,

which completes the proof (after renaming i to i r 1 ).□

4 Applications

Here, we develop some consequences of Theorem 3.1. A subset of a Cartan-Hadamard manifold M is convex if it contains the (unique) geodesic segment connecting every pair of its points. A convex hypersurface Γ M is the boundary of a compact convex set with interior points. If Γ is of class C 1 , 1 , then its principal curvatures are nonnegative at all twice differentiable points with respect to the outward normal. Conversely, if the principal curvatures of a closed hypersurface Γ M are all nonnegative, then Γ is convex [1]. See [13, Sec. 2 and 3] for the basic properties of convex sets in Cartan-Hadamard manifolds. A set is nested inside Γ if it lies in the convex domain bounded by Γ .

Corollary 4.1

Let Γ and γ be C 1 , 1 convex hypersurfaces in a Cartan-Hadamard n -manifold. Suppose that γ is nested inside Γ . Then, 1 ( Γ ) 1 ( γ ) .

Proof

Setting r = 1 in the comparison formula of Theorem 3.1 yields

(12) 1 ( Γ ) 1 ( γ ) = 2 Ω D σ 2 ( κ u ) Ω D Ric u u ,

where Ric stands for Ricci curvature; more explicitly, in a principal curvature frame where E n u / u , Ric ( E n ) is the sum of sectional curvatures K in , for 1 i n 1 . So Ric ( E n ) 0 . If Γ and γ are smooth ( C ) and strictly convex, we may let u in Theorem 3.1 be a function with convex level sets [4, Lem. 1]. Then, σ 2 ( κ u ) 0 , which yields 1 ( Γ ) 1 ( γ ) as desired. This completes the proof since we may approximate Γ and γ by smooth strictly convex hypersurfaces, e.g., by applying the Greene-Wu convolution to their distance functions, see [12, Lem. 3.3]; furthermore, total mean curvatures will converge here since they constitute “valuations” in the sense of integral geometry, see [13, Note 3.7] or [3, Prop. 3.8].□

Dekster [9] constructed examples of nested convex hypersurfaces in Cartan-Hadamard manifolds where the monotonicity property in the last result does not hold for Gauss-Kronecker curvature. So the aforementioned corollary cannot be extended to all mean curvatures without further assumptions, which we will discuss below. First, we need to record the following observation.

Lemma 4.2

Let S ρ be a geodesic sphere of radius ρ centered at a point in a Riemannian manifold. As ρ 0 , r ( S ρ ) converges to 0 for r n 2 and to S n 1 for r = n 1 .

Proof

A power series expansion [6, Thm. 3.1] of the second fundamental form of S ρ in normal coordinates shows that the principal curvatures of S ρ are given by κ i ρ = ( 1 + O ( ρ 2 ) ) / ρ . So

σ r ( κ ρ ) = n 1 r 1 ρ r ( 1 + O ( ρ 2 ) ) .

Another power series expansion [15, Thm. 3.1] yields

S ρ = S n 1 ρ n 1 ( 1 + O ( ρ 2 ) ) .

So, it follows that

r ( S ρ ) = n 1 r S n 1 ρ n 1 r ( 1 + O ( ρ 2 ) ) ,

which completes the proof.□

Gallego and Solanes showed [10, Cor. 3.2] that if Γ is a convex hypersurface bounding a domain Ω in a hyperbolic n -space of constant curvature a < 0 , then

1 ( Γ ) > ( n 1 ) 2 a Ω .

When comparing formulas, note that in [10], mean curvature is defined as the average of κ i , as opposed to the sum of κ i , which is our convention. Large balls show that the above inequality is sharp. Here, we extend this inequality to Cartan-Hadamard 3-manifolds as follows:

Corollary 4.3

Let Γ be a C 1 , 1 convex hypersurface in a Cartan-Hadamard n-manifold M bounding a domain Ω . Suppose that curvature of M is bounded above by a 0 . Then,

1 ( Γ ) > ( n 1 ) a Ω .

Furthermore, if n = 3 , then

1 ( Γ ) > 4 a Ω .

Proof

Let γ = γ ρ in (12) be a geodesic sphere of radius ρ . By Lemma 4.2, 1 ( γ ρ ) 0 as ρ 0 , which yields

1 ( Γ ) = 2 Ω σ 2 ( κ u ) Ω Ric u u > ( n 1 ) a Ω ,

as desired. When n = 3 , Gauss’ equation states that

σ 2 ( κ u ) = K u K M u ,

where K u is the sectional curvature of level sets of u and K M u is the sectional curvature of M with respect to tangent planes to level sets of u . Thus,

1 ( Γ ) = 2 Ω K u 2 Ω K M u Ω Ric u u > 4 a Ω ,

which completes the proof.□

We say Γ is an outer parallel hypersurface of a convex hypersurface γ if all points of Γ are at a constant distance λ 0 from the convex domain bounded by γ . Since the distance function of a convex set in a Cartan-Hadamard manifold is convex [5, Prop. 2.4], Γ is convex. Furthermore, Γ is C 1 , 1 for λ > 0 [13, Lem. 2.6]. The following corollary generalizes [13, Cor. 5.3] and a theorem of Schroeder-Strake [21, Thm. 3], where this result was established for Gauss-Kronecker curvature; see also [13, Note 6.9].

Corollary 4.4

Let M be a Cartan-Hadamard n-manifold, and Γ and γ be C 1 , 1 convex hypersurfaces in M. Suppose that Γ is an outer parallel hypersurface of γ . Then, r ( Γ ) r ( γ ) for 1 r n 1 .

Proof

We may let u in Theorem 3.1 be the distance function of the convex domain bounded by Γ . Then, u is constant on level sets of u . So, u i = 0 for 1 i n 1 , which yields

r ( Γ ) r ( γ ) ( r + 1 ) Ω D σ r + 1 ( κ u ) a ( n r ) Ω D σ r 1 ( κ u ) ,

where a 0 is the upper bound for sectional curvatures of M . Since u is convex, σ r ( κ u ) 0 , which completes the proof.□

The next result generalizes [13, Cor. 5.2] and observation of Borbely [4, Thm. 1] for Gauss-Kronecker curvature.

Corollary 4.5

Let M be a Cartan-Hadamard n-manifold with constant curvature, and Γ and γ be C 1 , 1 convex hypersurfaces in M, with γ nested inside Γ . Then, r ( Γ ) r ( γ ) , for 1 r n 1 .

Proof

Again we may assume that the function u in Theorem 3.1 is convex [4, Lem. 1]. If M has constant curvature a , then R i j k = a ( δ i k δ j δ i δ j k ) . Thus, Theorem 3.1 yields

(13) r ( Γ ) r ( γ ) = ( r + 1 ) Ω D σ r + 1 ( κ u ) a ( n r ) Ω D σ r 1 ( κ u ) .

By assumption a 0 , and since u is convex, σ r ( κ u ) 0 , which completes the proof.□

The above result had been observed earlier by Solanes [22, Cor. 9]. It is due to the integral formula for quermassintegrals [22, Def. 2.1], which immediately yields that quermassintegrals of convex domains are increasing with respect to inclusion. Monotonicity of total mean curvatures follows due to a formula [22, Prop. 7] relating quermassintegrals to total mean curvatures. As an application of the last corollary, one may extend the definition of total mean curvatures to non-regular convex hypersurfaces as follows. If Γ is a convex hypersurface in a Cartan-Hadamard manifold, then its outer parallel hypersurface at distance ε , denoted by Γ ε , is C 1 , 1 for all ε > 0 [13, Lem. 2.6]. So r ( Γ ε ) is well defined. By Corollary 4.4, r ( Γ ε ) is decreasing in ε . Hence, its limit as ε 0 exists, and we may set r ( Γ ) lim ε 0 r ( Γ ε ) .

Next, we derive a formula that appears in Solanes [22, (1) and (2)] and follows from Gauss-Bonnet-Chern theorems [8,7]; see also [22, Cor. 8]. Here k !! , when k is a positive integer, stands for the product of all positive odd (even) integers up to k , when k is odd (even). For k 0 , we set k !! = 1 .

Corollary 4.6

Let Γ be a closed C 1 , 1 hypersurface in an n-manifold M bounding a domain Ω . Suppose that M has constant curvature a, and cl ( Ω ) is diffeomorphic to a ball. Then,

n 1 ( Γ ) = S n 1 i = 1 n ( n mod 2 ) 2 ( 2 i 1 ) !! ( n 2 i 2 ) !! ( n 2 ) !! a i n 2 i 1 ( Γ ) .

Proof

Let ϕ : cl ( Ω ) B n be a diffeomorphism to the unit ball in R n and set u ( x ) ϕ ( x ) 2 . All regular level sets γ of u satisfy (13). Furthermore, these level sets are convex near the minimum point x 0 of u , since u has positive definite Hessian at x 0 . So by Corollary 4.5, for these small level sets,

r ( S ) r ( γ ) r ( S ) ,

where S and S are geodesic spheres centered at x 0 such that S is nested inside γ and γ is nested inside S . Consequently, by Lemma 4.2, as γ shrinks to x 0 , n 1 ( γ ) converges to S n 1 , while r ( γ ) vanishes for r n 2 . Thus, since σ n ( κ u ) = 0 , (13) yields

n 1 ( Γ ) = S n 1 a Ω σ n 2 ( κ u )

and

Ω σ r ( κ u ) = 1 r r 1 ( Γ ) + a ( n r + 1 ) r Ω σ r 2 ( κ u )

for r n 2 . Using these expressions iteratively completes the proof.□

Finally, we include a characterization for hyperbolic balls, which extends to all mean curvatures a previous result of the authors on Gauss-Kronecker curvature [13, Cor. 5.5].

Corollary 4.7

Let M be a Cartan-Hadamard n-manifold with curvature a 0 , and B ρ be a ball of radius ρ in M . Then, for 1 r n 1 ,

r ( B ρ ) r ( B ρ a ) ,

where B ρ a denotes a ball of radius ρ in a manifold of constant curvature a. Equality holds only if B ρ is isometric to B ρ a .

Proof

For r = n 1 , the desired inequality has already been established [13, Cor. 5.5]. Suppose then that r n 2 . We will show that

(14) r ( B ρ ) ( r + 1 ) B ρ σ r + 1 ( κ u ) a ( n r ) B ρ σ r 1 ( κ u ) r ( B ρ a ) .

Letting u be the distance squared function from the center o of B ρ , and γ shrink to o in Theorem 3.1, yields the first inequality in (14) via Lemma 4.2. The principal curvatures of B ρ are bounded below by a coth ( a ρ ) [16, p. 184], which are the principal curvatures of B ρ a . Hence, the mean curvatures of B ρ satisfy

σ r ( κ u ) n 1 r ( a coth ( a ρ ) ) r = σ r a ( κ u ) ,

where σ r a ( κ u ) are the mean curvatures of B ρ a . Furthermore, if A ( ρ , θ ) d θ denotes the volume element of B ρ in geodesic spherical coordinates, then by [16, (1.5.4)],

A ( ρ , θ ) sinh ( a ρ ) a n 1 = A a ( ρ , θ ) ,

where A a ( ρ , θ ) d θ is the volume element of B ρ a ; see [13, Cor. 5.5]. Thus,

B ρ σ r ( κ u ) 0 ρ S n 1 σ r a ( κ u ) A a ( t , θ ) d θ d t = B ρ a σ r a ( κ u ) ,

which yields the second inequality in (14). If r ( B ρ ) = r ( B ρ a ) , then equality holds in the first inequality of (14). So K r n = a , i.e., the radial sectional curvatures of B ρ are constant, which forces B ρ to have constant curvature a [13, Lem. 5.4]. Hence, B ρ is isometric to B ρ a .□

  1. Funding information: The research of M.G. was supported by National Science Foundation grant DMS-2202337 and a Simons Fellowship. The research of J.S. was supported by a Simons Collaboration Grant.

  2. Conflict of interest: Authors state no conflict of interest.

References

[1] S. Alexander, Locally convex hypersurfaces of negatively curved spaces, Proc. Amer. Math. Soc. 64 (1977), no. 2, 321–325. 10.1090/S0002-9939-1977-0448262-6Search in Google Scholar

[2] B. Andrews, Harmonic mean curvature flow and geometric inequalities, Adv. Math. 375 (2020), 107393, 28. 10.1016/j.aim.2020.107393Search in Google Scholar

[3] A. Bernig, Courbures intrinsèques dans les catégories analytico-géométriques, Ann. Inst. Fourier (Grenoble) 53 (2003), no. 6, 1897–1924. 10.5802/aif.1995Search in Google Scholar

[4] Albert Borbély, On the total curvature of convex hypersurfaces in hyperbolic spaces, Proc. Amer. Math. Soc. 130 (2002), no. 3, 849–854. 10.1090/S0002-9939-01-06101-9Search in Google Scholar

[5] Martin R. Bridson, Metric spaces of non-positive curvature, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 319, Springer-Verlag, Berlin, 1999. 10.1007/978-3-662-12494-9Search in Google Scholar

[6] B.-Y. Chen, Differential geometry of geodesic spheres, J. Reine Angew. Math. 325 (1981), 28–67. 10.1515/crll.1981.325.28Search in Google Scholar

[7] S.-s. Chern, On the curvatura integra in a Riemannian manifold, Ann. Math. 46 (1945), no. 2, 674–684. 10.1142/9789812812834_0010Search in Google Scholar

[8] S.-s. Chern, A simple intrinsic proof of the Gauss-Bonnet formula for closed Riemannian manifolds, Ann. Math. 45 (1944), no. 2, 747–752. 10.2307/1969302Search in Google Scholar

[9] B. V. Dekster, Monotonicity of integral Gauss curvature, J. Differential Geom. 16 (1981), no. 2, 281–291. 10.4310/jdg/1214436104Search in Google Scholar

[10] E. Gallego, Integral geometry and geometric inequalities in hyperbolic space, Differential Geom. Appl. 22 (2005), no. 3, 315–325. 10.1016/j.difgeo.2005.01.006Search in Google Scholar

[11] R. J. Gardner, The Brunn-Minkowski inequality, Bull. Amer. Math. Soc. (N.S.) 39 (2002), no. 3, 355–405. 10.1090/S0273-0979-02-00941-2Search in Google Scholar

[12] M. Ghomi, Minkowski inequality in Cartan-Hadamard manifolds, 2022, arXiv: http://arXiv.org/abs/arXiv:2206.06554. 10.1093/imrn/rnad114Search in Google Scholar

[13] M. Ghomi, Total curvature and the isoperimetric inequality in Cartan-Hadamard manifolds, J. Geom. Anal. 32 (2022), no. 2, Paper No. 50, 54. 10.1007/s12220-021-00801-2Search in Google Scholar

[14] A. Gray, Tubes, Progress in Mathematics, vol. 221, 2nd ed., Birkhäuser, Verlag, Basel, 2004, With a preface by Vicente Miquel. 10.1007/978-3-0348-7966-8Search in Google Scholar

[15] A. Gray, The volume of a small geodesic ball of a Riemannian manifold, Michigan Math. J. 20 (1973), 329–344 (1974). 10.1307/mmj/1029001150Search in Google Scholar

[16] B. Kleiner, An isoperimetric comparison theorem, Invent. Math. 108 (1992), no. 1, 37–47. 10.1007/BF02100598Search in Google Scholar

[17] R. C. Reilly, Applications of the Hessian operator in a Riemannian manifold, Indiana Univ. Math. J. 26 (1977), no. 3, 459–472. 10.1512/iumj.1977.26.26036Search in Google Scholar

[18] R. C. Reilly, On the Hessian of a function and the curvatures of its graph, Michigan Math. J. 20 (1973), 373–383. 10.1090/pspum/027.1/9926Search in Google Scholar

[19] L. A. Santaló. Integral Geometry and Geometric Probability, Vol. 1, Addison-Wesley Publishing Co., Reading, Mass.-London-Amsterdam, 1976, With a foreword by Mark Kac, Encyclopedia of Mathematics and its Applications. Search in Google Scholar

[20] R. Schneider, Convex bodies: the Brunn-Minkowski theory, expanded ed., Encyclopedia of Mathematics and its Applications, vol. 151, Cambridge University Press, Cambridge, 2014. 10.1017/CBO9780511526282Search in Google Scholar

[21] V. Schroeder, Local rigidity of symmetric spaces of nonpositive curvature, Proc. Amer. Math. Soc. 106 (1989), no. 2, 481–487. 10.1090/S0002-9939-1989-0929404-0Search in Google Scholar

[22] G. Solanes, Integral geometry and the Gauss-Bonnet theorem in constant curvature spaces, Trans. Amer. Math. Soc. 358 (2006), no. 3, 1105–1115. 10.1090/S0002-9947-05-03828-6Search in Google Scholar

[23] G. Wang, Isoperimetric type problems and Alexandrov-Fenchel type inequalities in the hyperbolic space, Adv. Math. 259 (2014), 532–556. 10.1016/j.aim.2014.01.024Search in Google Scholar

[24] Y. Wei, Inequalities of Alexandrov-Fenchel type for convex hypersurfaces in hyperbolic space and in the sphere, Pacific J. Math. 277 (2015), no. 1, 219–239. 10.2140/pjm.2015.277.219Search in Google Scholar

Received: 2022-04-27
Revised: 2022-09-22
Accepted: 2022-09-22
Published Online: 2023-01-10

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Asymptotic properties of critical points for subcritical Trudinger-Moser functional
  3. The existence of positive solution for an elliptic problem with critical growth and logarithmic perturbation
  4. On some dense sets in the space of dynamical systems
  5. Sharp profiles for diffusive logistic equation with spatial heterogeneity
  6. Generic properties of the Rabinowitz unbounded continuum
  7. Global bifurcation of coexistence states for a prey-predator model with prey-taxis/predator-taxis
  8. Multiple solutions of p-fractional Schrödinger-Choquard-Kirchhoff equations with Hardy-Littlewood-Sobolev critical exponents
  9. Improved fractional Trudinger-Moser inequalities on bounded intervals and the existence of their extremals
  10. The existence of infinitely many boundary blow-up solutions to the p-k-Hessian equation
  11. A priori bounds, existence, and uniqueness of smooth solutions to an anisotropic Lp Minkowski problem for log-concave measure
  12. Existence of nonminimal solutions to an inhomogeneous elliptic equation with supercritical nonlinearity
  13. Non-degeneracy of multi-peak solutions for the Schrödinger-Poisson problem
  14. Gagliardo-Nirenberg-type inequalities using fractional Sobolev spaces and Besov spaces
  15. Ground states of Schrödinger systems with the Chern-Simons gauge fields
  16. Quasilinear problems with nonlinear boundary conditions in higher-dimensional thin domains with corrugated boundaries
  17. A system of equations involving the fractional p-Laplacian and doubly critical nonlinearities
  18. A modified Picone-type identity and the uniqueness of positive symmetric solutions for a prescribed mean curvature problem
  19. On a version of hybrid existence result for a system of nonlinear equations
  20. Special Issue: Geometric PDEs and applications
  21. Preface for the special issue on “Geometric Partial Differential Equations and Applications”
  22. Convex hypersurfaces with prescribed Musielak-Orlicz-Gauss image measure
  23. Total mean curvatures of Riemannian hypersurfaces
  24. On degenerate case of prescribed curvature measure problems
  25. A curvature flow to the Lp Minkowski-type problem of q-capacity
  26. Aleksandrov reflection for extrinsic geometric flows of Euclidean hypersurfaces
  27. A note on second derivative estimates for Monge-Ampère-type equations
  28. The Lp chord Minkowski problem
  29. Widths of balls and free boundary minimal submanifolds
  30. Smooth approximation of twisted Kähler-Einstein metrics
  31. The exterior Dirichlet problem for the homogeneous complex k-Hessian equation
  32. A Carleman inequality on product manifolds and applications to rigidity problems
  33. Asymptotic behavior of solutions to the Monge-Ampère equations with slow convergence rate at infinity
  34. Pinched hypersurfaces are compact
  35. The spinorial energy for asymptotically Euclidean Ricci flow
  36. Geometry of CMC surfaces of finite index
  37. Capillary Schwarz symmetrization in the half-space
  38. Regularity of optimal mapping between hypercubes
  39. Special Issue: In honor of David Jerison
  40. Preface for the special issue in honor of David Jerison
  41. Homogenization of oblique boundary value problems
  42. A proof of a trace formula by Richard Melrose
  43. Compactness estimates for minimizers of the Alt-Phillips functional of negative exponents
  44. Regularity properties of monotone measure-preserving maps
  45. Examples of non-Dini domains with large singular sets
  46. Sharp inequalities for coherent states and their optimizers
  47. Gradient estimates and the fundamental solution for higher-order elliptic systems with lower-order terms
  48. Propagation of symmetries for Ricci shrinkers
  49. Linear extension operators for Sobolev spaces on radially symmetric binary trees
  50. The Neumann problem on the domain in 𝕊3 bounded by the Clifford torus
  51. On an effective equation of the reduced Hartree-Fock theory
  52. Polynomial sequences in discrete nilpotent groups of step 2
  53. Integral inequalities with an extended Poisson kernel and the existence of the extremals
  54. On singular solutions of Lane-Emden equation on the Heisenberg group
Downloaded on 7.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2022-0029/html
Scroll to top button