Next Article in Journal
A Broad TSK Fuzzy Classifier with a Simplified Set of Fuzzy Rules for Class-Imbalanced Learning
Previous Article in Journal
Reliability Optimization of Hybrid Systems Driven by Constraint Importance Measure Considering Different Cost Functions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Generalized Minkowski Type Integral Formulas for Compact Hypersurfaces in Pseudo-Riemannian Manifolds

Department of Mathematics, College of Sciences, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(20), 4281; https://doi.org/10.3390/math11204281
Submission received: 24 August 2023 / Revised: 10 October 2023 / Accepted: 12 October 2023 / Published: 13 October 2023
(This article belongs to the Section Algebra, Geometry and Topology)

Abstract

:
We obtain some generalized Minkowski type integral formulas for compact Riemannian (resp., spacelike) hypersurfaces in Riemannian (resp., Lorentzian) manifolds in the presence of an arbitrary vector field that we assume to be timelike in the case where the ambient space is Lorentzian. Some of these formulas generalize existing formulas in the case of conformal and Killing vector fields. We apply these integral formulas to obtain interesting results concerning the characterization of such hypersurfaces in some particular cases such as when the ambient space is Einstein admitting an arbitrary (in particular, conformal or Killing) vector field, and when the hypersurface has a constant mean curvature.

1. Introduction

In 1903, H. Minkowski published in [1] his two famous integral formulas for compact surfaces in three-dimensional Euclidean space. After that, many authors obtained integral formulas that generalized the two Minkowski formulas to hypersurfaces in Euclidean space and then in a general Riemannian manifold that admits a Killing or conformal vector field. For instance, in [2,3], C. C. Hsiung obtained generalized integral formulas of Minkowsi type for embedded hypersurfaces in Riemannian manifolds (see also [4]). In [5,6], Y. Katsurada generalized the work of Hsiung and derived some integral formulas of Minkowski type that were valid for Einstein manifolds and used them to prove that given a hypersurface M , g with constant mean curvature in an Einstein Riemannian manifold M ¯ , g ¯ , and given a homothetic vector field ξ of M ¯ , g ¯ such that the inner product of ξ and the normal to M does not change sign and does not vanish on M, then M is necessarily umbilical. In [7], K. Yano obtained three integral formulas of Minkowski type for hypersurfaces with constant mean curvature in a Riemannian manifold admitting a homothetic vector field. Then, over time, several integral formulas of Minkowski type appeared in the literature that were used to obtain rigidity results for isometrically immersed hypersurfaces in pseudo-Riemannian manifolds admitting a conformal vector filed. In [8,9,10], L. J. Alias, A. Romero, and M. Sanchez obtained the first and second integral formulas of Minkowski type for compact spacelike hypersurfaces in a generalized Robertson–Walker spacetime (resp., conformally stationary spacetime), and applied them to the study of compact spacelike hypersurfaces with constant mean curvature. Two years later, in [11], S. Montiel provided another proof of the first and second Minkowski formulas in the case where the ambient spacetime is equipped with a conformal timelike vector field. In 2003, L. J. Alias, A. Brasil JR, and A. G. Colares generalized in [12] the integral formulas obtained in [8,9,10] for spacelike hypersurfaces in conformally stationary spacetimes. See also [13,14].
The assumption that the ambient space admits a conformal vector field is inspired by the fact that the position vector field in Euclidean space is a closed conformal vector field (which in some references is called a concircular vector field). The importance of conformal vector fields comes from the use of conformal mappings as a mathematical tool in general relativity. In fact, although a conformal vector field does not leave the Einstein tensor invariant, its existence in a pseudo-Riemannian manifold M , g is a symmetry assumption for g that can be used (for example) to obtain exact solutions of Einstein’s equation.
Consider now an n + 1 -dimensional either Riemannian or Lorentzian manifold M ¯ , g ¯ admitting a conformal vector field ξ ¯ that we assume to be timelike in the case where M ¯ , g ¯ is Lorentzian. Let M , g be a connected n -dimensional Riemannian manifold that is isometrically immersed as a hypersurface into M ¯ , g ¯ , and let ξ denote the restriction of ξ ¯ to M . Consider the function θ = g ¯ ξ , N , where ξ ¯ is an arbitrary vector field and N is a globally defined unit vector field normal to M . In the case where it is Riemannian, L. J. Alias, M. Dajczer, and J. Ripoll gave in [15] an expression for the Laplacian Δ θ in terms of the Ricci curvature of M ¯ , g ¯ and the norm of the shape operator of M , g . One year later, in 2008, A. Barros, A. Brasil, and A. Caminha obtained in [16] the analogous expression when M ¯ , g ¯ is Lorentzian.
In 2010, A. L. Albujer, J. A. Aledo, and L. J. Alias gave in [17] an expression for Δ θ in a slightly different way than given in [15,16]. Then, they used that expression to obtain a Minkowski type integral formula for compact Riemannian and spacelike hypersurfaces, and applied this to deduce some interesting results concerning the characterization of compact Riemannian and spacelike hypersurfaces under certain hypotheses such as the constancy of the mean curvature or the assumption that the ambient space is Einstein or a product space. For more recent references pertaining to this work, we may cite [18,19,20,21].
In this paper, we mainly wish to generalize previous results concerning Minkowski type integral formulas for Riemannian (resp., spacelike) hypersurfaces in Riemannian (resp., Lorentzian) manifolds in the presence of an arbitrary vector field that we assume to be timelike in the case where it is Lorentzian, and apply these integral forms to compact Riemannian and spacelike hypersurfaces in order to obtain interesting results concerning the characterization of such hypersurfaces in some particular cases, such as the ambient space being Einstein and admitting an arbitrary (and in particular, a conformal Killing) vector field, or the hypersurface being minimal (resp., maximal) or having a constant mean curvature.
In particular, we generalize the results in [7,17] for any arbitrary Riemannian or spacelike hypersurface in any arbitrary ambient space with an arbitrary vector field. More precisely, given an n + 1 -dimensional either Riemannian or Lorentzian manifold M ¯ , g ¯ admitting an arbitrary vector field ξ ¯ that we assume to be timelike in the case where M ¯ , g ¯ is Lorentzian, and given a connected n -dimensional Riemannian manifold M , g that is isometrically immersed as a hypersurface into M ¯ , g ¯ , let ξ denote the restriction of ξ ¯ to M, and let N be a globally defined unit vector field normal to M . Of course, N is supposed to be timelike in the case where M ¯ , g ¯ is Lorentzian. Our first main goal in this paper is to give a useful expression for the Laplacian Δ θ of the function θ = g ¯ ξ , N in terms of the Ricci and scalar curvatures of the ambient space, the mean curvature of the hypersurface, and the tangent part of the restriction of the vector field ξ to M. In the particular case where ξ ¯ is a conformal (resp., Killing) vector field, our expression reduces to that obtained in [17] (resp. [7]). We deduce from the generalized expression for Δ θ different generalized Minkowski type integral formulas valid for any Riemannian or spacelike hypersurface in any arbitrary Riemannian or Lorentzian manifold admitting an arbitrary vector field. In particular, we generalize an integral formula obtained in [17] in the case where ξ ¯ is conformal to the case of an arbitrary vector field. We also apply the obtained generalized Minkowski type formulas to deduce interesting results concerning the characterization of Riemannian and spacelike hypersurfaces in some particular cases, such as the ambient space being Einstein and admitting an arbitrary (and in particular, a conformal Killing) vector field, or the hypersurface having a constant mean curvature.

2. Preliminaries

Let n 2 , and let M , g be a connected n -dimensional pseudo-Riemannian manifold. In this paper, we adopt the opposite convention of that in [22] to define the Riemannian tensor. That is, the Riemannian tensor is defined here to be the 1 , 3 tensor field given by
R ( X , Y ) Z = [ X , Y ] Z X , Y Z ,
for all X , Y , Z X M .
For every p M and every orthonormal basis e 1 , , e n of T p M , the Ricci curvature tensor R i c and the scalar curvature S c a l are, respectively, defined to be
R i c ( X , Y ) = t r a c e ( Z R ( Z , X ) Y ) = i = 1 n ϵ i g ( R ( X , e i ) e i , Y ) ,
for all X , Y T p M .
S c a l p = t r a c e R i c = i = 1 n ϵ i R i c e i , e i ,
where ϵ i = g e i , e i .
Throughout this paper, we assume that M , g is Riemannian (i.e., the metric g has index 0) which is isometrically immersed as a hypersurface into an n + 1 -dimensional pseudo-Riemannian manifold M ¯ , g ¯ that we assume to be Riemannian or Lorentzian (i.e., the metric g ¯ has index 0 or 1). Let   and   ¯ denote the Levi-Civita connections on M and M ¯ , respectively. Let X M and X ( M ¯ ) denote, respectively, the sets of all tangent vector fields on M and M , ¯ and let X ¯ M denote the set of all M ¯ vector fields on M . We use the two notations X · f or X f to denote the value of a vector field X on a function f.
Let ξ ¯ X ( M ¯ ) , which we assume to be timelike in the case where M ¯ is Lorentzian, and let θ ξ ¯ denote its dual one-form, that is, the one-form given by θ ξ ¯ Y = g ¯ ξ ¯ , Y , for every Y X ( M ¯ ) . Let A ξ ¯ be the 1 , 1 -tensor (viewed as an endomorphism) defined by
A ξ ¯ Y = ¯ Y ξ ¯
We write as usual
L ξ ¯ g ¯ Y , Z + d θ ξ ¯ Y , Z = 2 g ¯ A ξ ¯ Y , Z ,
for all Y , Z X ( M ¯ ) , where L is the Lie derivative of the metric g ¯ with respect to ξ ¯ .
Let B and ϕ be the symmetric and skew-symmetric parts of A ξ ¯ . In other words, we have
L ξ ¯ g ¯ Y , Z = 2 g ¯ B Y , Z
d θ ξ ¯ Y , Z = 2 g ¯ ϕ Y , Z ,
Now, in the case where M ¯ is Riemannian, we assume that there exists a globally defined unit vector field N normal to M . In this case, M is said to be a two-sided hypersurface. In the case where M ¯ is Lorentzian, since M is a spacelike hypersurface in M ¯ , and ξ ¯ is assumed to be timelike, then we can choose a (globally defined) timelike unit vector field N normal to M and in the same time orientation of ξ ¯ , that is, we have g ¯ ξ ¯ , N < 0 on M . In both cases, if ξ is the restriction of ξ ¯ to M, then we denote by θ the smooth function on M, called the support function, which is defined by θ = g ¯ ξ , N . It is clear that in the case where M ¯ is Lorentzian, we have θ g ¯ ξ , ξ < 0 . If T is the tangential component of ξ to M, then we have
ξ = T + ϵ θ N ,
where ϵ = g ¯ N , N = ± 1 , according to whether M ¯ is Riemannian or Lorentzian, respectively.
Since ξ X ¯ M , then the operator A ξ : X M X ¯ M given by A ξ X = ¯ X ξ is well defined (see for instance [22], pp. 97–99). Then, we have
A ξ X = ψ X + ϵ α X N ,
where ψ X = ¯ X ξ is the tangential component of ¯ X ξ to M, and α is a one-form on M .
Let η X M be the vector field associated to α . Therefore, for all X X M , we have
g ( η , X ) = α X = g ¯ A ξ X , N = g ¯ ϕ X + B X , N = g ¯ X , ϕ N + g ¯ X , B N = g X , ϕ N + g X , B N
Since ϕ is skew-symmetric, we have g ¯ ϕ N , N = 0 , that is, ϕ N X M . Therefore, (5) implies that
η = B N ϕ N
On the other hand, the Gauss and Weingarten formulae for M as a hypersurface of M ¯ are given by
¯ X Y = X Y + ϵ g A X , Y N A X = ¯ X N ,
for all X , Y X M , where A is the shape operator of M with respect to N .
Therefore, for all X X M , we have
¯ X ξ = X T ϵ θ A X + ϵ g A T + θ , X N
From (4) and (7), we deduce that
X T = ψ X + ϵ θ A X
θ = η A T
We need to start by revisiting some key definitions. In general, recall that for a 1 , 1 -tensor S, the covariant derivative S of S is defined as follows
S ( X , Y ) = X S Y = X S Y S ( X Y )
The divergence of a vector field X X M is defined as the function
d i v X = i = 1 n g e i X , e i ,
where e 1 , , e n is a local orthonormal frame of vector fields.
The divergence of a 1 , 1 -tensor S on M is defined as the vector field
d i v S = t r a c e S = i = 1 n ( S ) ( e i , e i ) ,
where, as above, e 1 , , e n is a local orthonormal frame of vector fields.
We observe that without loss of generality, we may assume e 1 , , e n to be parallel. In this case, we see that
d i v S = i = 1 n e i S e i
We also recall that the curvature tensor R of M is given in terms of the curvature tensor R ¯ of M ¯ and the shape operator by the so-called Gauss equation
R ( X , Y ) Z = R ¯ ( X , Y ) Z + ϵ ( g ( A Y , Z ) A X g ( A X , Z ) A Y ) ,
for all X , Y , Z X M .
Recalling that the mean curvature of M is defined to be
H = ϵ n t r a c e ( A ) ,
it follows from (11) that the Ricci curvatures R i c and R i c ¯ of M and M ¯ are related as follows
R i c ( X , Y ) = R i c ¯ ( X , Y ) ϵ g ¯ ( R ¯ ( N , X ) Y , N ) + g ( A X , n H Y ϵ A Y ) ,
for all X , Y X M .
Moreover, by tracing (13), we see that the scalar curvatures S c a l and S c a l ¯ of M and M ¯ are related as follows
S c a l = S c a l ¯ 2 ϵ R i c ¯ ( N , N ) + ϵ n 2 H 2 A 2

3. Some Useful Tensor Formulas

With the notations above, let M ¯ , g ¯ be an n + 1 -dimensional either Riemannian or Lorentzian manifold, and let ξ ¯ X ( M ¯ ) be an arbitrary vector field that we assume to be timelike in the case where M ¯ , g ¯ is Lorentzian. Let M , g be a connected n -dimensional Riemannian manifold that is isometrically immersed as a hypersurface into M ¯ , g ¯ , and let ξ denote the restriction of ξ ¯ to M .
Our main goal in this section is to give a useful expression for the Laplacian Δ θ of the function θ = g ¯ ξ , N , where ξ ¯ is an arbitrary vector field and N is a globally defined unit vector field normal to M . In the case where M ¯ , g ¯ is Riemannian and ξ ¯ is a Killing (resp., conformal) vector field, an expression for Δ θ has been given in [23] (resp. [15]) in terms of the Ricci curvature of M ¯ , g ¯ and the norm of the shape operator. An analogous formula has been obtained in [16] in the case where M ¯ , g ¯ is Lorentzian and ξ ¯ is a timelike conformal vector field. As we have mentioned in the introduction, in [17], a formula for Δ θ was obtained in a slightly different way as given in [15,16].
Let us denote by B the restriction of B to T M , and let f = g ¯ ¯ N ξ , N . It is clear that f is a smooth function on M . In fact, from (1), we see that
f = 1 2 L ξ g ¯ ( N , N ) = g ¯ ( B N , N )
To calculate Δ θ , we use (9). Thus, we start by computing the divergences of T and A T .
Proposition 1. 
Let the notation and assumptions be as above. Then, we have
d i v T = d i v ξ ϵ f + n H θ
d i v A T = g T , d i v A + t r a c e A B T + ϵ θ A 2
Proof. 
Let e 1 , , e n be a local orthonormal frame of vector fields. When e 1 , , e n and N are extended arbitrarily to vector fields on M ¯ , then according to (10), and making use of (3), we have
d i v T = i = 1 n g ( e i T , e i ) = ( d i v ξ ϵ f ) + n H θ .
Using the same formula (10) and making use of (8) and of the fact that if S is a self-adjoint operator, then so is X S , we obtain
d i v A T = i = 1 n g ( e i A T , e i ) = i = 1 n g ( e i A T , e i ) + i = 1 n g ( A e i T , e i ) = g ( T , i = 1 n e i A e i ) + i = 1 n g ( ψ e i + ϵ θ A e i , A e i ) = g ( T , d i v A ) + i = 1 n g ¯ ( A ξ e i , A e i ) + ϵ θ A 2 = g ( T , d i v A ) + i = 1 n g ¯ B e i , A e i + ϵ θ A 2 ,
where we have also used here (at the last step) the fact that since A is self-adjoint and ϕ is skew-symmetric, g ¯ ϕ e i , A e i = 0 for all i. □
In the following proposition, we give an explicit useful formula for d i v A T in terms of the Ricci curvature (compare to formula (14) in [17]).
Proposition 2. 
Let the notation and assumptions be as above. Then, we have
d i v A T = ϵ n T H + t r a c e ( A B T ) + ϵ θ A 2 R i c ¯ N , T
Proof. 
Let e 1 , , e n be a local orthonormal frame of vector fields that we assume to be parallel. As we have noticed above, when e 1 , , e n and N are extended arbitrarily to vector fields on M ¯ , then, using symmetric properties of the curvature tensor R ¯ of M ¯ , we have
R i c ¯ ( N , T ) = i = 1 n g ¯ ( R ¯ ( e i , N ) T , e i ) + ϵ g ¯ ( R ¯ ( N , N ) T , N ) = i = 1 n g ¯ ( R ¯ ( e i , T ) N , e i ) = i = 1 n g ( ( A ) ( T , e i ) ( A ) ( e i , T ) , e i ) = i = 1 n g ( T A e i , e i ) i = 1 n g ( e i A T , e i ) = i = 1 n T · g ( A e i , e i ) g ( T , i = 1 n e i A e i ) = T · t r a c e A g ( T , i = 1 n ( A ) ( e i , e i ) ) = ϵ n T H g ( T , d i v A )
Now, from this last expression and (17), we obtain (18).
We now give an expression for d i v ( ¯ N ξ ) . For this purpose, we note that
¯ N ξ = ( ¯ N ξ ) + ϵ f N ,
from which we have
¯ X ¯ N ξ = X ( ¯ N ξ ) + g ¯ A X , ( ¯ N ξ ) N + ϵ X f N + ϵ f ¯ X N = X ( ¯ N ξ ) + g ¯ A X , ( ¯ N ξ ) N + ϵ X f N ϵ f A X
Since ¯ N ξ = B N + ϕ N and ϕ N X M , it follows that
( ¯ N ξ ) = B N + ϕ N
( ¯ N ξ ) = B N
Proposition 3. 
Let the notation and assumptions be as above. Then, we have
d i v ( ¯ N ξ ) = R i c ¯ ( N , ξ ) + N · d i v ξ ϵ f t r a c e A B + n f H
Proof. 
Let e 1 , , e n be a parallel local orthonormal frame of vector fields in X M . Then, by using (19), we deduce that
d i v ( ¯ N ξ ) = i = 1 n g ( e i ( ¯ N ξ ) , e i ) = i = 1 n g ¯ ( ¯ e i ¯ N ξ , e i ) + i = 1 n g ( ϵ f A e i , e i ) = i = 1 n g ¯ ( ¯ e i ¯ N ξ , e i ) + ϵ f t r a c e A = i = 1 n g ¯ ( ¯ e i ¯ N ξ , e i ) + n f H
On the other hand, by extending e 1 , , e n so that ¯ N e i = 0 for all i = 1 , , n , we have
R ¯ ( e i , N ) ξ = ¯ e i ¯ N ξ ¯ N ¯ e i ξ + ¯ A ( e i ) ξ
By using (24), we see that (23) becomes
d i v ( ¯ N ξ ) = R i c ¯ ( N , ξ ) + i = 1 n g ¯ ( ¯ N ¯ e i ξ , e i ) i = 1 n g ¯ ( ¯ A ( e i ) ξ , e i ) + n f H = R i c ¯ ( N , ξ ) + i = 1 n N · g ¯ ( ¯ e i ξ , e i ) i = 1 n g ¯ ( ¯ A ( e i ) ξ , e i ) + n f H = R i c ¯ ( N , ξ ) + N · d i v ξ ϵ f t r a c e A B + n f H ,
where we have used the fact that
g ¯ ( ¯ A ( e i ) ξ , e i ) = g ¯ B A e i + ϕ A e i , e i = t r a c e A B .
Remark 1. 
Note that since
N · d i v ξ ϵ f = N · t r a c e B = t r a c e N B
and
g ¯ ( ϕ A e i , e i ) = 0 , 1 i n ,
we can express (22) as follows
d i v ( ¯ N ξ ) = R i c ¯ ( N , ξ ) + t r a c e N B A B + n f H
We are now ready to give the desired expression for Δ θ .
Theorem 1. 
Let M , g be a connected n -dimensional Riemannian manifold that is isometrically immersed as a hypersurface into an n + 1 -dimensional either Riemannian or Lorentzian manifold M ¯ , g ¯ . Let ξ ¯ X ( M ¯ ) be an arbitrary vector field that we assume to be timelike in the case where ( M ¯ , g ¯ ) is Lorentzian, and let ξ denote the restriction of ξ ¯ to M . Let N be a globally defined unit vector field normal to M , and let the notation used here be as above. Then, the Laplacian Δ θ of the function θ = g ¯ ξ , N is given by
Δ θ = 2 d i v B N ϵ θ ( R i c ¯ ( N , N ) + A 2 ) N · d i v ξ ϵ f ϵ n ( T H + ϵ f H )
Proof. 
By (6) and (9), we have
θ = B N ϕ N A T
By using (18) and (22), it follows that
Δ θ = d i v ( B N ϕ N A T ) = 2 d i v B N R i c ¯ ( N , ξ ) N · ( d i v ξ ϵ f ) + t r a c e ( A B ) n f H ϵ n T H t r a c e ( A B T ) ϵ θ A 2 + R i c ¯ ( N , T ) = 2 d i v B N ϵ θ ( R i c ¯ ( N , N ) + A 2 ) N · d i v ξ ϵ f ϵ n ( T H + f H ) ,
as desired. □
As a straightforward consequence of Theorem 1, we obtain an interesting expression for Δ θ in the particular case where ξ ¯ is a conformal Killing vector field on M ¯ , that is, a vector field satisfying
L ξ ¯ g ¯ = 2 ψ g ¯ ,
for some smooth function ψ on M ¯ , called the conformal factor (or potential function) of ξ ¯ .
Corollary 1. 
Let the notation and assumptions be as in Theorem 1, and assume in addition that the vector field ξ ¯ is conformal. Then, we have
Δ θ = ϵ θ ( R i c ¯ ( N , N ) + A 2 ) ϵ n ψ H + ϵ N ψ + T H
Proof. 
From (1) and (28), we deduce that B = ψ I , where I is the identity. It follows that B N = 0 and B N = ψ N . Consequently, by using (15), we obtain
f = 1 2 L ξ g ¯ ( N , N ) = g ¯ ( B N , N ) = ϵ ψ ,
where we have identified here the function ψ with its restriction to M . We also have d i v ξ ¯ = t r a c e A ξ ¯ = t r a c e B = n + 1 ψ , from which we deduce that d i v ξ ϵ f = n ψ . Now, by substituting these into (26), we obtain (29). □
It would be of some use to express Δ θ in terms of the scalar curvatures of M and M ¯ . This can be done by combining the two formulas (14) and (26), so that we obtain formula (31) in the following theorem.
Theorem 2. 
Let the notation and assumptions be as in Theorem 1. Then, we have
Δ θ = 2 d i v B N θ ( S c a l ¯ S c a l ϵ R i c ¯ ( N , N ) + ϵ n 2 H 2 ) N · d i v ξ ϵ f ϵ n ( T H + f H )
Note that (31) is a generalization to the case of an arbitrary vector field on M ¯ of formula (9) in [17], which was given in the case where ξ ¯ is a conformal Killing vector field.
Theorem 3 
([17]). Let the notation and assumptions be as in Theorem 1, and assume in addition that the vector field ξ ¯ is conformal. Then, we have
Δ θ = θ ( S c a l ¯ S c a l ϵ R i c ¯ ( N , N ) + ϵ n 2 H 2 ) ϵ n ψ H + ϵ N ψ + T H
On the other hand, we give an expression for d i v ( ¯ N ξ ) .
Proposition 4. 
Let the notation and assumptions be as above. Then, we have
d i v ( ¯ N ξ ) = N · f ϵ n f H
Proof. 
Let e 1 , , e n be a parallel local orthonormal frame of vector fields. Note first that
¯ e i e i = g ( A e i , e i ) N
We also note that since N · g ¯ ( N , N ) = 0 , we have ¯ N N X M . With these in hand, we can calculate
d i v ( ¯ N ξ ) = i = 1 n g ¯ ( ¯ e i ( ¯ N ξ ) , e i ) + g ¯ ( ¯ N ( ¯ N ξ ) , N )
= i = 1 n ( e i · g ¯ ( ( ¯ N ξ ) , e i ) g ¯ ( ( ¯ N ξ ) , g ( A e i , e i ) N )
+ N · g ¯ ( ( ¯ N ξ ) , N ) g ¯ ( ( ¯ N ξ ) , ¯ N N )
= t r a c e ( A ) g ¯ ( ¯ N ξ ) , N ) + N · g ¯ ( ¯ N ξ ) , N )
= ϵ n f H + N · f ,
as desired. □
Remark 2. 
By combining the two formulas (22) and (33), we deduce that
d i v ¯ N ξ = R i c ¯ ( N , ξ ) + N · d i v ξ t r a c e A B
Note that (40) is similar to the following general formula where ξ is a projective vector field
d i v A ξ Y = R i c Y , ξ + Y · d i v ξ + t r a c e A ξ A Y
To obtain (40), it suffices to take Y = N in (41) and remember that d i v ¯ N ξ = d i v A ξ N and A N = A .

4. Integral Formulas for Compact Riemannian Hypersurfaces in Pseudo-Riemannian Manifolds

In this section, we assume that M , g is an n-dimensional compact Riemannian manifold that is isometrically immersed as a hypersurface in an n + 1 -dimensional either Riemannian or Lorentzian manifold M ¯ , g ¯ with all the assumptions stated at the beginning of the above section. The first integral formula that we can display here results directly from the integration of the simple formula (16).
Proposition 5. 
Let M , g be as above. Then, we have
M d i v ξ ϵ f + n H θ d V = 0
In particular, if ξ ¯ is a conformal Killing vector field with conformal factor ψ, then
M ψ + H θ d V = 0
By using formula (42) of the previous proposition, the following results can be easily deduced.
Proposition 6. 
In a Lorentzian manifold with an arbitrary timelike vector field ξ (resp., a conformal vector field with a potential function ψ), there does not exist any compact spacelike hypersurface for which the mean curvature function H satisfies the condition d i v ξ + f H < 0 , where f is the function defined by (15) (resp., H ψ < 0 ).
Proof. 
It is clear that if d i v ξ + f H < 0 , then either d i v ξ + f < 0 and H > 0 or d i v ξ + f > 0 and H < 0 . Since θ < 0 and H does not change sign, we deduce that either d i v ξ + f + n H θ < 0 or d i v ξ + f + n H θ > 0 . If M is compact, then formula (42) implies in both cases that d i v ξ + f = 0 and H = 0 , which is absurd. □
Proposition 7. 
With the notations and assumptions previously stated at the beginning of the above section, assume that M , g is either a compact Riemannian manifold that is either minimal or maximal according to whether M ¯ , g ¯ is Riemannian or Lorentzian, respectively. Then, there exists a point p M such that t r a c e B p = 0 , that is, d i v ξ p = g ¯ ¯ N ξ , N p or equivalently ( ¯ N ξ ) p = d i v ξ p · N p . In particular, if ξ is affine, then ( ¯ N ξ ) p = 0 . If ξ is conformal with conformal factor ψ , then ψ p = 0 .
As an immediate consequence of Proposition 7, we have the following corollary.
Corollary 2. 
Given the notations and assumptions outlined earlier in this section, consider that ξ ¯ is a homothetic vector field. Then, when it is Riemannian (resp., Lorentzian), M ¯ , g ¯ contains no compact minimal (resp., maximal) Riemannian hypersurface.
A more general result than Proposition 7 is the following proposition.
Proposition 8. 
With the notations and assumptions stated at the beginning of the above section, let ξ ¯ be an arbitrary vector field, and assume that M is compact with constant mean curvature. Assume in addition, in the case where M ¯ , g ¯ is Riemannian, that the function θ is not constant and does not change sign.
(a) 
If H θ > 0 , then there exists a point p M such that d i v ξ p < g ¯ ¯ N ξ , N p .
(b) 
If H θ < 0 , then there exists a point p M such that d i v ξ p > g ¯ ¯ N ξ , N p .
(c) 
If H = 0 , then there exists a point p M such that d i v ξ p = g ¯ ¯ N ξ , N p .
Remark 3. 
On the other hand, we easily deduce from (43) that if ξ ¯ is a Killing vector field (i.e., ψ = 0 ) and M has constant mean curvature, then either θ vanishes somewhere or H = 0 (i.e., M is minimal in the case where M ¯ is Riemannian and maximal in the case where M ¯ is Lorentzian). Conversely, if ξ ¯ is a homothetic vector field (i.e., ψ is constant) and H = 0 , then ξ ¯ is necessarily a Killing vector field. We also deduce from (43) that if ξ ¯ is a homothetic vector field and θ = 0 , then ξ ¯ is necessarily a Killing vector field. This is exactly what Theorem 5.3 states in [7].
Our second integral formula involves R i c ¯ N , T and is as follows.
Theorem 4. 
Let M , g be an n-dimensional compact Riemannian manifold that is isometrically immersed as a hypersurface in an n + 1 -dimensional either Riemannian or Lorentzian manifold M ¯ , g ¯ . Then, with the assumptions stated in Theorem 1, we have
M ϵ n 1 T H + t r a c e A ϵ H I B + ϵ θ A 2 n H 2 R i c ¯ N , T d V = 0
In particular, when ξ ¯ is a conformal Killing vector field with conformal factor ψ, then
M ϵ n 1 T H + ϵ θ A 2 n H 2 R i c ¯ N , T d V = 0
Proof. 
Using (16), we obtain
d i v H T = H d i v T + T H = H ( d i v ξ ϵ f + n H θ ) + T H = H d i v ξ ϵ f + n H 2 θ + T H
Using (18), and recalling that d i v ξ ϵ f = t r a c e B , we obtain
d i v A T ϵ d i v H T = ϵ n 1 T H + t r a c e A B ϵ H d i v ξ ϵ f + ϵ θ A 2 n H 2 R i c ¯ N , T = ϵ n 1 T H + t r a c e A ϵ H I B + ϵ θ A 2 n H 2 R i c ¯ N , T
Now, by integrating both sides of the above equation, we obtain formula (44). When ξ ¯ is a conformal Killing vector field with conformal factor ψ , we have B = ψ I . Substituting this into (44), we obtain (45). □
Our third integral formula involves the scalar curvatures of both M and M ¯ , as well as R i c ¯ N , N , and is stated as follows.
Theorem 5. 
Let M , g be an n-dimensional compact Riemannian manifold that is isometrically immersed as a hypersurface in an n + 1 -dimensional either Riemannian or Lorentzian manifold M ¯ , g ¯ with the assumptions stated in Theorem 1. Then, we have
M θ ( S c a l S c a l ¯ + ϵ R i c ¯ ( N , N ) ) d V = M N · d i v ξ ϵ f d V ϵ n M H d i v ξ 2 ϵ f d V
In particular, when ξ ¯ is a conformal Killing vector field with conformal factor ψ, we meet formula (18) in [17], that is,
M θ ( S c a l S c a l ¯ + ϵ R i c ¯ ( N , N ) ) d V = n M N ψ d V ϵ n n 1 M H ψ d V
Proof. 
By using (16), we have
T H = d i v H T H d i v T = d i v H T H d i v ξ ϵ f + n H θ ,
and by substituting this into (31), we obtain (46). To obtain (47), it suffices to substitute the values f = ϵ ψ and d i v ξ = n + 1 ψ into (46). □
Remark 4. 
We notice that in the above result, formula (47) is nothing but formula (18) in [17], so that (46) can be considered as a generalization of that formula to the case of a general vector field ξ ¯ .

5. Integral Formulas for CMC Compact Riemannian Hypersurfaces in Pseudo-Riemannian Manifolds

In this section, we focus on the case where M has a constant mean curvature H . The first result gives an integral formula for a hypersurface with constant mean curvature without any assumption on the ambient space M ¯ , g ¯ or on the vector field ξ ¯ .
Theorem 6. 
Under the notations and assumptions stated in Theorem 1, let M ¯ , g ¯ be an n + 1 -dimensional either Riemannian or Lorentzian manifold, and M , g a compact Riemannian manifold that is isometrically immersed as a hypersurface with constant mean curvature H in M ¯ , g ¯ . Then, we have
M θ ( R i c ¯ ( N , N ) + A 2 n H 2 ) d V + ϵ M N · d i v ξ ϵ f d V H M d i v ξ ϵ n + 1 f d V = 0
In particular, when ξ ¯ is a conformal Killing vector field with conformal factor ψ, we have
M θ ( R i c ¯ ( N , N ) + A 2 n H 2 ) d V + ϵ n M N · ψ d V = 0 ,
and when ξ ¯ is homothetic, we have
M θ ( R i c ¯ ( N , N ) + A 2 n H 2 ) d V = 0
Proof. 
First, since we know that
S c a l S c a l ¯ + ϵ R i c ¯ ( N , N ) = ϵ ( R i c ¯ ( N , N ) + A 2 n 2 H 2 ) ,
then (46) yields
M θ ( R i c ¯ ( N , N ) + A 2 n 2 H 2 ) d V + ϵ M N · d i v ξ ϵ f d V n M H d i v ξ 2 ϵ f d V = 0 ,
or equivalently
M θ ( R i c ¯ ( N , N ) + A 2 n H 2 ) d V n n 1 M H 2 θ d V + ϵ M N · d i v ξ ϵ f d V n M H d i v ξ 2 ϵ f d V = 0
Since H is constant, (42) yields
n n 1 M H 2 θ d V = ( n 1 ) H M n H θ d V = n 1 H M d i v ξ ϵ f d V
Now, if we substitute (54) into (53), we obtain (48). Formulas (49) and (50) follow easily from (48) using the facts that f = ϵ ψ and d i v ξ = n + 1 ψ when ξ ¯ is a conformal Killing vector field with conformal factor ψ , and the fact that ψ is constant when ξ ¯ is homothetic, respectively. □
Since A 2 n H 2 0 , (50) can be used to deduce the following result which generalizes Theorem 5.1 in [7] to the case of a spacelike hypersurface.
Corollary 3. 
Let M ¯ , g ¯ be an n + 1 -dimensional either Riemannian or Lorentzian manifold which admits a homothetic vector field ξ ¯ , and let M , g be an n-dimensional compact Riemannian manifold that is isometrically immersed in M ¯ , g ¯ as a hypersurface with constant mean curvature. Let N and ξ denote, respectively, the normal to M and the restriction of ξ ¯ to M . Assume that R i c ¯ N , N 0 on M and assume (in the case where M ¯ , g ¯ is Riemannian) that the function θ = g ¯ N , ξ does not change sign and is not identically zero. Then, M , g is totally umbilical and R i c ¯ N , N = 0 on M .
The second result is a direct consequence of Theorem 4 under the assumptions that M ¯ , g ¯ is Einstein and M has a constant mean curvature H. This has been proved in [6] in the case where M ¯ , g ¯ is Riemannian.
Theorem 7. 
Let M ¯ , g ¯ be an n + 1 -dimensional either Riemannian or Lorentzian Einstein manifold with a conformal Killing vector field ξ ¯ , and let M , g be an n-dimensional compact Riemannian manifold that is isometrically immersed as a hypersurface in M ¯ , g ¯ , with constant mean curvature H . With all the notations and assumptions stated at the beginning of the above section, assume in addition (in the case where M ¯ , g ¯ is Riemannian) that the function θ does not change sign and is not identically zero. Then, M , g is necessarily totally umbilical.
Proof. 
Under the assumptions of the proposition, formula (45) becomes
M θ A 2 n H 2 = 0
Since θ does not change sign and is not identically zero, and since A 2 n H 2 0 , we should obtain from the integral above that A 2 = n H 2 . We deduce that A = ϵ H I , that is, M , g is totally umbilical. □

6. Conclusions

In conclusion, the study presented in this work has yielded a set of generalized Minkowski type integral formulas applicable to compact Riemannian and spacelike hypersurfaces within Riemannian and Lorentzian manifolds, respectively, especially when an arbitrary vector field is present, assuming it to be timelike in the Lorentzian case. These formulas extend and build upon existing results, particularly in scenarios involving conformal and Killing vector fields. Moreover, the practical application of these integral formulas has yielded valuable insights into the characterizations of such hypersurfaces, especially in cases where the ambient space is an Einstein manifold and accommodates various vector field types, including conformal and Killing vector fields, as well as scenarios where the hypersurfaces exhibit a constant mean curvature. We believe that this research contributes to a deeper understanding of geometric properties in these specific contexts, providing valuable tools for further exploration in differential geometry and related fields. These tools also extend their utility to the study of Ricci solitons [24,25,26,27,28,29].

Author Contributions

Conceptualization , M.G.; investigation, N.A. and M.G.; writing—original draft, N.A. and M.G.; writing—review & editing, M.G.; supervision, M.G. All authors have read and agreed to the published version of the manuscript.

Funding

The second author was supported by Researchers Supporting Project number (RSPD2023R1053), King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Minkowski, H. Volumen und Oberflache. Math. Ann. 1903, 57, 447–495. [Google Scholar] [CrossRef]
  2. Hsiung, C.C. Some integral formulas for closed hypersurfaces. Math. Scand. 1954, 2, 286–294. [Google Scholar] [CrossRef]
  3. Hsiung, C.C. Some integral formulas for closed hypersurfaces in Riemann space. Pac. J. Math. 1956, 6, 291–299. [Google Scholar] [CrossRef]
  4. Hsiung, C.C.; Liu, J.; Mittra, S. Integral formulas for closed submanifolds of a Riemannian manifold. J. Differ. Geom. 1977, 12, 133–151. [Google Scholar] [CrossRef]
  5. Katsurada, Y. Generalised Minkowski formula for closed hypersurfaces in Riemann space. Ann. Mat. Pura Appl. 1962, 57, 283–294. [Google Scholar] [CrossRef]
  6. Katsurada, Y. On a certain property of closed hypersurfaces in an Einstein space. Comment. Math. Helv. 1964, 38, 165–171. [Google Scholar] [CrossRef]
  7. Yano, K. Closed hypersurfaces with constant mean curvature in a Riemannian manifold. J. Math. Soc. Jpn. 1965, 17, 330–340. [Google Scholar] [CrossRef]
  8. Alias, L.J.; Romero, A.; Sanchez, M. Uniqueness of complete spacelike hypersurfaces of constant mean curvature in Generalized Robertson-Walker spacetimes. Gen. Relativ. Grav. 1995, 27, 71–84. [Google Scholar] [CrossRef]
  9. Alias, L.J.; Romero, A.; Sanchez, M. Spacelike hypersurfaces of constant mean curvature and Calabi-Bernstein type problems. Tohoku Math J. 1997, 49, 337–345. [Google Scholar] [CrossRef]
  10. Alias, L.J.; Romero, A.; Sanchez, M. Spacelike hypersurfaces of constant mean curvature in certain spacetimes. Nonlinear Anal. TMA 1997, 30, 655–661. [Google Scholar] [CrossRef]
  11. Montiel, S. Unicity of constant mean curvature hypersurfaces in some Riemannian manifolds. Indiana Univ. Math. J. 1999, 48, 711–748. [Google Scholar] [CrossRef]
  12. Alias, L.J.; Brasil, A.; Colares, A.G. Integral formulae for spacelike hyper-surfaces in conformally stationary spacetimes and applications. Proc. Edinb. Math. Soc. 2003, 46, 465–488. [Google Scholar] [CrossRef]
  13. Alias, L.J.; García-Martínez, S.C. On the scalar curvature of constant mean curvature hypersurfaces in space forms. J. Math. Anal. Appl. 2010, 363, 579–587. [Google Scholar] [CrossRef]
  14. Alias, L.J.; Dajczer, M. Constant mean curvature hypersurfaces in warped product spaces. Proc. Edinb. Math. Soc. 2007, 50, 511–526. [Google Scholar] [CrossRef]
  15. Alias, L.J.; Dajczer, M.; Ripoll, J. A Bernstein-type theorem for Riemannian manifolds with a Killing field. Ann. Glob. Anal. Geom. 2007, 31, 363–373. [Google Scholar] [CrossRef]
  16. Barros, A.; Brasil, A.; Caminha, A. Stability of spacelike hypersurfaces in foliated spacetimes. Differ. Geom. Appl. 2008, 26, 357–365. [Google Scholar] [CrossRef]
  17. Albujer, A.L.; Aledo, J.A.; Alias, L.J. On the scalar curvature of hypersur-faces in spaces with a Killing field. Adv. Geom. 2010, 10, 487–503. [Google Scholar] [CrossRef]
  18. Alohali, H.; Alodan, H.; Deshmukh, S. Conformal vector fields sub-manifolds of a Euclidean space. Publ. Math. Debr. 2017, 91, 217–233. [Google Scholar] [CrossRef]
  19. Chen, B.Y.; Wei, S.W. Riemannian submanifolds with concircular canonical field. Bull. Korean Math. Soc. 2019, 56, 1525–1537. [Google Scholar]
  20. Evangelista, I.; Viana, E. Conformal gradient vector fields on Riemannian manifolds with boundary. Colloq. Math. 2020, 159, 231–241. [Google Scholar] [CrossRef]
  21. Liu, J.; Xie, X. Complete spacelike hypersurfaces with CMC in Lorentz Einstein mani-folds. Bull. Korean Math. Soc. 2021, 58, 1053–1068. [Google Scholar]
  22. O’Neill, B. Semi-Reimannian Geometry with Applications to Relativity; Academic Press: New York, NY, USA, 1983. [Google Scholar]
  23. Furnari, S.; Ripoll, J. Killing fields, mean curvature, translation maps. Ill. J. Math. 2004, 48, 1385–1403. [Google Scholar] [CrossRef]
  24. Li, Y.; Patra, D.; Alluhaibi, N.; Mofarreh, F.; Ali, A. Geometric classifications of k-almost Ricci solitons admitting paracontact metrices. Open Math. 2023, 21, 20220610. [Google Scholar] [CrossRef]
  25. Diógenesa, R.; Ribeiro, E., Jr.; Filhoa, J.S. Gradient Ricci solitons admitting a closed conformal Vector field. J. Math. Anal. Appl. 2017, 455, 1975–1983. [Google Scholar] [CrossRef]
  26. Li, Y.; Kumara, H.A.; Siddesha, M.S.; Naik, D.M. Characterization of Ricci Almost Soliton on Lorentzian Manifolds. Symmetry 2023, 15, 1175. [Google Scholar] [CrossRef]
  27. Li, Y.; Srivastava, S.K.; Mofarreh, F.; Kumar, A.; Ali, A. Ricci Soliton of CR-Warped Prod-uct Manifolds and Their Classifications. Symmetry 2023, 15, 976. [Google Scholar] [CrossRef]
  28. Li, Y.; Gupta, M.K.; Sharma, S.; Chaubey, S.K. On Ricci Curvature of a Homogeneous Gen-eralized Matsumoto Finsler Space. Mathematics 2023, 11, 3365. [Google Scholar] [CrossRef]
  29. Li, Y.; Bhattacharyya, S.; Azami, S.; Saha, A.; Hui, S.K. Harnack Estimation for Nonlinear, Weighted, Heat-Type Equation along Geometric Flow and Applications. Mathematics 2023, 11, 2516. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alessa, N.; Guediri, M. Generalized Minkowski Type Integral Formulas for Compact Hypersurfaces in Pseudo-Riemannian Manifolds. Mathematics 2023, 11, 4281. https://doi.org/10.3390/math11204281

AMA Style

Alessa N, Guediri M. Generalized Minkowski Type Integral Formulas for Compact Hypersurfaces in Pseudo-Riemannian Manifolds. Mathematics. 2023; 11(20):4281. https://doi.org/10.3390/math11204281

Chicago/Turabian Style

Alessa, Norah, and Mohammed Guediri. 2023. "Generalized Minkowski Type Integral Formulas for Compact Hypersurfaces in Pseudo-Riemannian Manifolds" Mathematics 11, no. 20: 4281. https://doi.org/10.3390/math11204281

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop