Acessibilidade / Reportar erro

Integral inequalities for closed linear Weingarten submanifolds in the product spaces

Abstract

An integral inequality for closed linear Weingarten 𝑚-submanifolds with parallel normalized mean curvature vector field (pnmc lw-submanifolds) in the product spaces 𝑀𝑛(𝑐) × ℝ, 𝑛 > 𝑚 ≥ 4, where 𝑀𝑛(𝑐) is a space form of constant sectional curvature 𝑐 ∈ {−1, 1}, is proved. As an application is shown that the sharpness in this inequality is attained in the totally umbilical hypersurfaces, and in a certain family of standard product of the form 𝕊1(√1 − 𝑟2) × 𝕊𝑚−1(𝑟) with 0 < 𝑟 < 1 when 𝑐 = 1. In the case where 𝑐 = −1, is obtained an integral inequality whose sharpness is attained only in the totally umbilical hypersurfaces. When 𝑚 = 2 and 𝑚 = 3, an integral inequality is also obtained with equality happening in the totally umbilical hypersurfaces.

Key words
Closed pnmc lw-submanifolds; product spaces; totally umbilical hypersurfaces; standard product

Introduction

Within the theory of isometric immersions, the characterization of closed submanifolds (compact with empty boundaries) with one of their constant curvatures using integral inequalities constitutes a classical research topic. Notable among these is Simons’ integral inequality (see Simons 1968SIMONS J. 1968. Minimal varieties in Riemannian manifolds. Ann Math 88: 62-105.), which establishes a relationship between the squared norm of the second fundamental form and the dimension and codimension of the minimal submanifold in the unit sphere. It is worth highlighting that Simons’ tool has proven effective not only in the study of minimal closed submanifolds in the sphere but also in the investigation of submanifolds with other constant curvatures, as well as in more general ambient spaces (see, for example, Chern et al. 1970CHERN SS, DO CARMO MP & KOBAYASHI S. 1970. Minimal Submanifolds of a Sphere with Second Fundamental Form of Constant Length. Springer Berlin Heidelberg, 59-75 p., Lawson 1969LAWSON HB. 1969. Local Rigidity Theorems for Minimal Hypersurfaces. Ann of Math 89(1): 187-197., Ôtsuki 1970, dos Santos & da Silva 2021, and references therein).

In the context of hypersurfaces, Cheng & Yau (1977)CHENG SY & YAU ST. 1977. Hypersurfaces with constant scalar curvature. Math Ann 225: 195-204. investigated the rigidity of hypersurfaces with constant scalar curvature in a space form. They introduced a new second-order differential operator known as the square operator. Building upon Cheng-Yau’s technique, Li (1996)LI H. 1996. Hypersurfaces with constant scalar curvature in space forms. Math Ann 305: 665-672. studied the pinching problem concerning the square norm of the second fundamental form for complete hypersurfaces with constant scalar curvature. Later, Wei (2008)WEI G. 2008. Simons’ type integral formula for hypersurfaces in a unit sphere. J Math Anal Appl 340: 1371-179. derived a Simons’ type integral inequality for closed k-minimal rotational hypersurfaces immersed in 𝕊m+1, characterizing the equality through the standard product 𝕊1(1r2)×𝕊m1(r). In higher codimension, Guo & Li (2013)GUO X & LI H. 2013. Submanifolds with constant scalar curvature in a unit sphere. Tohoku Math J 65(3): 331-339. extended the results of Li (1996)LI H. 1996. Hypersurfaces with constant scalar curvature in space forms. Math Ann 305: 665-672. and showed that the only closed submanifolds with parallel normalized mean curvature (pnmc) in the unit sphere 𝕊m+p with constant scalar curvature, and whose second fundamental form satisfies appropriate boundedness, are the totally umbilical sphere 𝕊m(r) and the standard product 𝕊1(1r2)×𝕊m1(r).

Recently, Alías & Meléndez (2020) studied the rigidity of closed hypersurfaces with constant scalar curvature isometrically immersed in 𝕊m+1. In particular, they established a sharp Simons-type integral inequality for the squared norm of the traceless second fundamental form, with equality characterizing the totally umbilical hypersurfaces and the standard product 𝕊1(1r2)×𝕊m1(r). More recently, by using the approach developed by Alías & Meléndez (2020), dos Santos & da Silva (2021) generalized the sharp Simons-type integral inequality of Alías & Meléndez (2020) for pnmc submanifolds immersed in the Riemannian product space 𝕊n× having constant second mean curvature. As an application, they showed that the sharpness in this inequality is attained in the totally umbilical hypersurfaces, and in a certain family of standard product of the form 𝕊1(1r2)×𝕊m1(r)𝕊m+1×{t0}𝕊n×, for some t0 with n>m4.

On the other hand, a natural extension of the submanifolds with constant second mean curvature is the linear Weingarten submanifolds. A submanifold is said to be linear Weingarten (here we will denote by lw-submanifolds) when the first and the second mean curvatures satisfy a certain linear relation. Here, we deal with m-dimensional closed pnmc lw-submanifolds immersed in a Riemannian product space Mn(c)×, where Mn(c) is a space form of constant sectional curvature c=1,1 with n>m4. In this setting, we extend the technique developed by the first two authors in dos Santos & da Silva (2021, 2022) in order to prove a sharp integral inequality for pnmc lw-submanifolds obtaining natural generalizations of the main results of Alías & Meléndez (2020) and dos Santos & da Silva (2021). Furthermore, we also obtain integral inequalities when c=1, which is not contemplated in dos Santos & da Silva (2021).

This manuscript is organized as follows: In Section 1, we provide a brief review of fundamental concepts related to submanifolds immersed in a Riemannian product space Mn(c)×. Subsequently, we establish a Simons’ type formula for pnmc lw-submanifolds in Mn(c)× (see Proposition 2). In Section 2, we present auxiliary lemmas concerning pnmc lw-submanifolds in Mn(c)×. Moving on to Section 3, we provide a lower estimate for a Cheng-Yau modified operator acting on the square norm of the traceless second fundamental form of such submanifolds (see Proposition 9). We then apply this result to establish our characterization theorems for closed pnmc lw-submanifolds in Mn(c)× with a constant angle between the normalized mean curvature and the unit vector field tangent to (see Theorems 3.3 and 3.4). Finally, in the last section, we examine the cases of two and three dimensions (see Theorems 4.1 and 4.2)..

A Simons type formula for submanifolds in Mn(c)×

Along this manuscript, we will always deal with an m-dimensional connected submanifold Σm immersed in a Riemannian manifold M¯n+1 with nm. We choose a local field of orthonormal frames e1,,en+1 in M¯n+1, with dual coframes ω1,,ωn+1, such that, at each point of Σm, e1,,em are tangent to Σm and em+1,,en+1 are normal to Σm. We will use the following convention of indices:

1i,j,k,mandm+1α,β,γ,n+1.
Now, restricting all the tensors to Σm, ωα=0 on Σm. Hence, iωαiωi=dωα=0 and as it is well known we get
ωαi=jhijαωjandhijα=hjiα.(1)
This gives
A=α,i,jhijαωiωjeα,hijα=Aα(ei),ej=A(ei,ej),eα(2)
with A denoting the second fundamental form of Σm in M¯n+1. The square length of the shape operator is
|A|2=α|Aα|2=α,i,j(hijα)2.(3)
Furthermore, we define the mean curvature vector h and the mean curvature function H of Σm in M¯n+1, respectively by
h=1mαtr(Aα)eαandH=|h|=1mαtr(Aα)2,(4)
where tr(Aα)=ihiiα.

As it is well known, the basic equations of the submanifolds are the Gauss equation

Rijkl=R¯ijkl+β(hikβhjlβhilβhjkβ),(5)
where R¯ijkl and Rijkl are the components of the curvature tensor of M¯n+1 and Σm, respectively, the Ricci equation
Rαβij=R¯αβij+k(hikαhkjβhkjαhikβ),(6)
where Rαβij are the components of the normal curvature tensor of Σm, and the Codazzi equation
hijkαhikjα=R¯αijk.(7)
where hijkα denote the first covariant derivatives of hijα. Additionally,
|A|2=α,i,j,k(hijkα)2,(8)
where denotes the covariant derivative of the second fundamental form A. In particular, we say that Σm is a parallel submanifold of M¯n+1 when A=0 (see van der Veken & Vrancken 2008).

In this setting, the following Simons-type formula is well-known (see dos Santos & da Silva 2021, 2022):

Proposition 1.1.. Let Σm a submanifold immersed isometrically in a Riemannian manifold M¯n+1, nm. Then, we have

1 2 Δ | A | 2 = | A | 2 + α , i , j , k h i j α ( h k k i j α R ¯ α i k j k R ¯ α k k i j ) + α , β , i , j , k h i j α ( h k k β R ¯ α i j β + 2 h j k β R ¯ α β k i h i j β R ¯ α k β k + 2 h k i β R ¯ α β k j ) α , β ( N ( A α A β A β A α ) + [ tr ( A α A β ) ] 2 tr ( A β ) tr ( A α 2 A β ) ) + 2 α , i , j , k , p h p j α ( h p k α R ¯ p i j k + h p j α R ¯ p k i k ) , (9)

where N(A)=tr(AAt) for all matrix A=(aij).

From now on, let us consider the case where the ambient space is a product space. Let M¯n+1=Mn(c)× be a product space, where Mn(c) be a connected Riemannian manifold endowed with metric tensor ,M and of constant sectional curvature c=1,1 and is the real line. Thus, the product space Mn(c)× is the differential manifold Mn(c)× endowed with the Riemannian metric

v,w=(πM)*v,(πM)*wM+(π)*v,(π)*w,(10)
with (p,t)Mn(c)× and v,wT(p,t)(Mn(c)×), where π and πM denote the projections onto the corresponding factor. Associated with the product space, we know that, the vector field
t:=(/t)|(p,t),(p,t)Mn(c)×(11)
is parallel and unitary, that is,
¯t=0andt,t=1,(12)
where ¯ is the Levi-Civita connection of the Riemannian metric of Mn(c)×. Using the notations established in Fetcu & Rosenberg (2013)FETCU D & ROSENBERG H. 2013. On complete submanifolds with parallel mean curvature in product spaces. Rev Mat Iberoam 29: 1283-1304., we write the decomposition
t=T+N(13)
where T:=t and N:=t denotes, respectively, the tangent and normal parts of the vector field t on the tangent and normal bundle of the submanifold Σm in Mn(c)×. Moreover, from (12) and (13), we get the relation
1=t,t=|T|2+|N|2.(14)
It is clear that, if T=0 then, t is normal to Σm and, hence Σm lies in Mn(c).

Moreover, let us recall that the curvature tensor2 2 We adopt for the (1,3)-curvature tensor the following definition of Chapter 3 of O’Neill (1983): R¯(X,Y)Z=∇¯[X,Y]Z−[∇¯X,∇¯Y]Z. of Mn(c)× satisfies, (see Daniel 2007DANIEL B. 2007. Isometric immersions into 3-dimensional homogeneous manifolds. Commentarii Math Helv 82: 87-131.),

R¯(X,Y)Z=c(X,ZYY,ZX)+cZ,t(Y,tXX,tY)+c(Y,ZX,tX,ZY,t)t,(15)
where X,Y,Z𝔛(Mn(c)×).

In what follows, we will denote by and , respectively, the tangent and normal Levi-Civita connections along the tangent and normal bundle of Σm, a direct computation by (11) give us

XT=AN(X)andXN=A(T,X),for allX𝔛(M),(16)
where AN=αN,eαAα denotes the Weingarten operator in the N direction.

By this digression, our aim now is to get a Simons-type formula for a pnmc lw-submanifold Σm in Mn(c)×. Firstly, since Mn(c)× locally symmetric, we have R¯αikjk=R¯αkkij=0. On the other hand, a direct computation from (15), gives R¯αβkj=0, for all α,β,j,k. Moreover,

α,β,i,j,khijα(hkkβR¯αijβ+hijβR¯αkβk)=cm2H2+cmA(T,T),hcm|AN|2+cm2h,N2+c(m|T|2)|A|2(17)
and
α,i,j,k,phpjα(hpkαR¯pijk+hpjαR¯pkik)=cmα|Aα(T)|2+c(m|T|2)|A|2cm2H2+2cmA(T,T),h.(18)
Next, we will also consider the traceless second fundamental form
ϕ=α,i,jϕijαωiωjeα,ϕijα=ϕα(ei),ej=hijαh,eαδij.(19)
It is easy to check that each ϕα=Aαh,eαI is traceless and that
|ϕ|2=α|ϕα|2=α,i,j(ϕijα)2=|A|2mH2.(20)
Observe that |ϕ|2=0 if and only if Σm is a totally umbilical submanifold of Mn(c)×. Within this context, a standard computation give us
m|AN|2=m|ϕN|2+m2h,N2(21)
and
α|Aα(T)|2=α|ϕα(T)|2+2ϕh(T),T+H2|T|2.(22)

Now, let Σm be a pnmc submanifolds immersed in product space Mn(c)×. This means that H>0 and the normalized mean curvature vector field η=h/H is parallel as a section of the normal bundle. In this setting, we will consider {em+1,,en+1} be a local orthonormal frame field in the normal bundle such that em+1=η. By this,

tr(Aη)=mHandtr(Aα)=mh,eα=0,for allαm+2,(23)
and by (19)
ϕijm+1=hijm+1Hδijandϕijα=hijα,for allαm+2.(24)
Since η parallel, the Ricci equation (6) guarantees that AαAη=AηAα for all αm+2. Using this, (20) and (24),
α,βtr(Aβ)tr(Aα2Aβ)α,β(N(AαAβAβAα)+[tr(AαAβ)]2)=mH2|ϕ|2+mHαtr(ϕα2ϕη)α,β>m+1N(ϕαϕβϕβϕα)α,β[tr(ϕαϕβ)]2.(25)
Therefore, inserting (17), (18), (21), (22) and (25) in Proposition 1.1 we get
12Δ|A|2=|A|2+mi,jhijm+1Hij+cm|ϕN|22cmα|ϕα(T)|2+(c(m|T|2)+mH2)|ϕ|2cmϕh(T),T+mHαtr(ϕα2ϕη)α,β>m+1N(ϕαϕβϕβϕα)α,β[tr(ϕαϕβ)]2.(26)

According to Grosjean (2002)GROSJEAN JF. 2002. Upper bounds for the first eigenvalue of the Laplacian on closed submanifolds. Pacific J Math 206: 93-112. and Cao & Li (2007)CAO L & LI H. 2007. r-Minimal submanifolds in space forms. Ann Glob Anal Geom 32: 311-341., we define the r-th mean curvature function Hr of an m-dimensional submanifold immersed in a Riemannian space, as follows: for any even integer r{0,1,,m1}, the r-th are given by

(nr)Hr:=Sr=1r!i1irj1jrδj1jri1irBi1j1,Bi2j2Bir1jr1,Birjr,(27)
where (nr) is the binomial coefficient, δj1jri1ir is the generalized Kronecker symbol and Bij=α,i,jhijαeα with {em+1,,en+1} an orthonormal frame on the normal bundle. By convention, H0=S0=1. For our study on submanifolds Σm in the product space Mn(c)×, we will consider the second mean curvature function H2, which is given by
m(m1)H2=2S2=m2H2|A|2.(28)

On the other hand, a natural extension of submanifolds having constant second mean curvature is the so-called linear Weingarten, in short, lw-submanifolds. A submanifold is said to be linear Weingarten red if its first and second mean curvatures are linearly related, that is,

H2=aH+b(29)
for constants a,b. Observe that when a=0, (29) reduces to H2 constant.

For the study of the lw-submanifolds, we will consider the following Cheng-Yau’s modified differential operator given by

L(u)=i,j[(mHm12a)δijhijm+1]uij=(mHm12a)Δui,jhijm+1uij,(30)
where uij stands for a component of the Hessian of u𝒞2(M). From the tensorial point of view, (30) can be written as
L(u)=tr(PHessu),(31)
with
P=(mHm12a)Ihm+1(32)
where I is the identity in the algebra of smooth vector fields on Σm and hm+1=(hijm+1) denotes the second fundamental form of Σm in the direction em+1. By (31), it is not difficult to see that
L(uv)=uL(v)+vL(u)+2P(u),v(33)
for every u,v𝒞2(M) and
L(f(u))=f(u)L(u)+f(u)P(u),u(34)
for every smooth function f:.

Hence, taking u=mH in (30), by (28) and (29), we obtain

L(mH)=i,j[(mHm12a)δijhijm+1](mH)ij=mHΔ(mH)m(m1)2Δ(aH)mi,jhijm+1Hij=12Δ(m2H2m(m1)H2)m2|H|2mi,jhijm+1Hij=12Δ|A|2m2|H|2mi,jhijm+1Hij.

From all these results we have the following Simons-type formula for Cheng-Yau’s modified operator acting on the mean curvature function of Σm in Mn(c)× which generalizes Proposition 2 of dos Santos & da Silva (2022):

Proposition 1.2.. If Σm is a pnmc lw-submanifold of Mn(c)×, then we have

L ( m H ) = | A | 2 m 2 | H | 2 + c m | ϕ N | 2 2 c m α | ϕ α ( T ) | 2 + ( c ( m | T | 2 ) + m H 2 ) | ϕ | 2 c m H ϕ η ( T ) , T + m H α tr ( ϕ α 2 ϕ η ) α , β ( N ( ϕ α ϕ β ϕ β ϕ α ) + [ tr ( ϕ α ϕ β ) ] 2 ) .

Key lemmas

In this section, we will present some necessary results for the proof of our results. The first ones are extensions of the Lemmas 1 and 2 of dos Santos & da Silva (2022) (see also Lemma 2.3 of dos Santos & da Silva (2021) and Lemmas 4.1 and 4.3 of dos Santos (2021)) to lw-submanifolds.

Lemma 2.1.. Let Σm be an lw-submanifold in the product space Mn(c)×, such that H2=aH+b with

( m 1 ) a 2 + 4 m b 0 . (35)

Then

| A | 2 m 2 | H | 2 . (36)

Moreover, if the inequality (35) is strict and the equality occurs in (36), then Σm is an open piece of a parallel submanifold of Mn(c)×.

Proof. Inserting H2=aH+b in (28) we have

m2H2=|A|2+m(m1)(aH+b).(37)
By taking the derivative in (37),
2|A||A|=(2m2Hm(m1)a)H(38)
and consequently
4|A|2||A||2=(2m2Hm(m1)a)2|H|2.(39)
It is not difficult to check that
(2m2Hm(m1)a)2=4m2|A|2+m2(m1)(4mb+(m1)a2).(40)
Thus by using (35),
4|A|2||A||2=[4m2|A|2+m2(m1)(4mb+(m1)a2)]|H|24m2|A|2|H|2.(41)
Now, from Kato’s inequality
||A||2|A|2(42)
we obtain
m2|A|2|H|2|A|2||A||2|A|2|A|2.(43)
Therefore, we have either
|A|2=0andm2|H|2=|A|2=0(44)
or
|A|2m2|H|2.(45)
If the inequality (35) is strict, from (41) we get
(2m2Hm(m1)a)2>4m2|A|2.(46)
Now, let us assume in addition that the equality holds in (36) on Σm. In this case, we wish to show that H is constant on Σm. Suppose, by contradiction, that it does not occur. Consequently, there exists a point pΣm such that |H(p)|>0. So, one deduces from (39) that
4|A|2(p)|A|2(p)>4m2|A|2(p)|H(p)|2(47)
and, since |A|2(p)=m2|H(p)|2>0, we arrive at a contradiction. Hence, in this case, we conclude that H must be constant on Σm. ◻

Lemma 2.2.. Let Σm be a pnmc lw-submanifold in the product space Mn(c)×, such that H2=aH+b with b0. Then the operator P defined in (32) is positive semidefinite. In the case where b>0, we have that P is positive definite.

Proof. Let us consider {e1,,em} an orthonormal frame on Σm such that hijm+1=λim+1δij. Since b0, from (37), we have

m2H2=|A|2+m(m1)(aH+b)(λim+1)2+m(m1)aH,(48)
for each principal curvature λim+1 of Σm, i=1,,m.

On the other hand, with a straightforward computation, we verify that

(λim+1)2m2H2m(m1)aH=(mHm12a)2(m1)24a2(mHm12a)2.(49)
Now, we claim that mHm12a0. For this, let us consider two cases. When a0, our assertion is immediate. Otherwise, if a>0, from (37) we see that
mH(mH(m1)a)=|A|2+m(m1)b>0,(50)
since Σm is a pnmc submanifold. Thus, mH(m1)a>0 and consequently, mHm12a0 as claimed.

So, from (49) we obtain

mH+m12aλim+1mHm12a,i=1,,m,(51)
and hence, for each i{1,,m}
0mHm12aλi2mH(m1)a.(52)
Since mHm12aλi are the eigenvalues of P, follows that P is positive semidefinite. Similarly if b>0. ◻

Given a unit normal vector field ξ𝔛(Σ), we say that a submanifold Σm of Mn(c)× has constant ξ-angle if the angle between ξ and t is constant, that is, the function ξ,t is constant along of Σm. We should notice that constant η-angle submanifolds, where η=h/H, corresponds to a natural extension of hypersurfaces with constant angle in a product space, which was widely studied by Dillen and many other authors (see, for instance, Dillen et al. 2007DILLEN F, FASTENAKELS J, VAN DER VEKEN J & VRANCKEN L. 2007. Constant angle surfaces in 𝕊 2 × ℝ . Monatsh Math 152: 89-96., Dillen & Munteanu 2009DILLEN F & MUNTEANU MI. 2009. Constant angle surfaces in ℍ 2 × ℝ . Bull Braz Math Soc 40: 85-97., Navarro et al. 2016NAVARRO M, RUIZ-HERNÁNDEZ G & SOLIS DA. 2016. Constant mean curvature hypersurfaces with constant angle in semi-Riemannian space forms. Differ Geo and its Appl 49: 473-495., Nistor 2017NISTOR AI. 2017. New developments on constant angle property in 𝕊 2 × ℝ . Ann Mat Pura Appl 196: 863-875.). By using this context, the next result is a suitable adaptation of Lemma 2.1 of dos Santos & da Silva (2021) which assures that the integral of the L operator acting on any nonnegative function is equal to zero.

Lemma 2.3.. Let Σm be a closed pnmc lw-submanifold in Mn(c)× such that H2=aH+b. If Σm has constant η-angle, then this angle is always zero. Moreover

ΣL(u)dΣ=0,(53)
for all nonnegative functions u𝒞2(Σ).

Proof. By a standard tensorial computation, it is not difficult to see that

L(u)=div(P(u))div(P),u,(54)
for every u𝒞2(M), where
div(P)=i(eiP)ei=tr(P),(55)
with P is defined as
P(X,Y)=(XP)Y=XP(Y)P(XY),X,YTM.(56)
From this and (32), we write
P(ei,ei)=mH,eieihm+1(ei,ei),(57)
where {e1,,en+1} an orthonormal frame on Mn× adapted to Σm, that is, {e1,,em} are tangent to Σm and choose em+1=η. By Codazzi equation (7),
hm+1(ei,ei),X=(eihm+1)ei,X=ei,(eihm+1)X=(Xhm+1)ei,eijX,ejR¯(m+1)iji,(58)
for all XTM. By using (15), a direct computation give us
jX,ejR¯(m+1)iji=cη,t(ei,TX,eiT,Xei,ei).(59)
Hence,
div(P)=mHmHc(m1)η,tT=c(m1)η,tT.(60)
On the other hand, we take the vector field X=uT. Computing its divergence, we obtain
div(X)=udiv(T)+T(u)=udiv(T)+T,u.(61)
By (16),div(T)=mh,t. So,
div(X)=umh,t+T,u.(62)
Since Σm has constant η-angle, we get
div(η,tX)=umη,th,t+η,tT,u.(63)
Therefore, as h,t=Hη,t, from (54), (62) and (63),
div(P(u))=L(u)(m1)div(η,tX)+m(m1)uHη,t2.(64)
Taking into account Stokes’ Theorem,
ΣL(u)dΣ=cm(m1)η,t2ΣuHdΣ.(65)
Finally, let us choose u a positive constant function. Since H>0, from (65) we must have η,t=0. Therefore, inserting this in (65) we obtain the result. ◻

The following two results are fundamental to our study and can be found in Li & Li (1992)LI AM & LI JM. 1992. An intrinsic rigidity theorem for minimal submanifolds in a sphere. Arch Math 58: 582-594. and Santos (1994)SANTOS W. 1994. Submanifolds with parallel mean curvature vector in spheres. Tohoku Math J 46: 403-415., respectively.

Lemma 2.4.. Let B1,,Bp, where p2, be symmetric m×m matrices. Then

α , β = 1 p ( N ( B α B β B β B α ) + [ tr ( B α B β ) ] 2 ) 3 2 ( α = 1 p N ( B α ) ) 2 .

Lemma 2.5.. Let B,C:mm be symmetric linear maps that [B,C]=0 and tr(B)=tr(C)=0, then

m2m(m1)|B|2|C|tr(B2C)m2m(m1)|B|2|C|.
Moreover, the equality holds if and only if (m1) of the eigenvalues xi of B and corresponding eigenvalues yi of C satisfy

| x i | = N ( B ) m ( m 1 ) , x i y i 0 and y i = N ( C ) m ( m 1 ) ( r e s p . N ( C ) m ( m 1 ) ) .

We will conclude this section by quoting the following codimension reduction result for submanifolds in the product space Mn(c)×, see Lemma 1.6 of Mendonça & Tojeiro (2013).

Lemma 2.6.. Let Σm be a submanifold of Mn(c)× and let N be the normal vector field defined by (13). Assume that L:=N1+span{N} is a subbundle of TΣ with rank q<nm+1 and that N1L, where N1 denotes the first normal subspace of Σm. Then the codimension of Σm reduces to q, that is, Σm is contained in a totally geodesic submanifold Mm+q1(c)× of Mn(c)×.

Main results

In our first result, we obtain a suitable lower estimate for the operator L applied on the squared norm of the traceless operator of a lw-submanifold, which will be also essential to the proofs of our main results.

Proposition 3.1.. Let Σm be a pnmc lw-submanifold in a product space Mn(c)×, n>m4, such that H2=aH+b with a,b0. Then

L(|ϕ|2)2(m1)(|ϕ|2φa,b,c(|ϕ|,|T|)𝒬c)|ϕ|2m(m1)+b+a24,(66)
where
𝒬c=cm|ϕN|22cmα|ϕα(T)|2cmHϕη(T),T(67)
and
φa,b,c(x,y)=m2m1x2+cy2m(am2m(m1)x)x2m(m1)+b+a24+m(m2)a2m(m1)xm(a22+b+c).(68)
In particular, if b>0 and equality holds in (66), then Σm is a part of a parallel submanifold in Mn(c)× with two distinct principal curvatures, one of which is simple.

Proof. From Cauchy Schwarz’s inequality and Lemma 2.5, we get

α,β[tr(ϕαϕβ)]2|ϕη|4+2|ϕη|2(|ϕ|2|ϕη|2)+α,β>m+1[tr(ϕαϕβ)]2,(69)
and
mHαtr(ϕα2ϕη)m(m2)m(m1)H|ϕ|2|ϕη|.(70)
Besides these, by Lemma 2.4, we also can estimate
α,β>m+1(N(ϕαϕβϕβϕα)+[tr(ϕαϕβ)]2)32(α>m+1|ϕα|2)2=32(|ϕ|2|ϕη|2)2.(71)
Then, inequalities (69), (70) and (71), becomes in
mHαtr(ϕα2ϕη)α,βm+1N(ϕαϕβϕβϕα)α,β[tr(ϕαϕβ)]2m(m2)m(m1)H|ϕ|2|ϕη||ϕη|42|ϕη|2(|ϕ|2|ϕη|2)32(|ϕ|2|ϕη|2)2.(72)
After some standard computations, we can express (72) as follows:
mH2|ϕ|2m(m2)m(m1)H|ϕ|2|ϕη|32|ϕ|4+|ϕ|2|ϕη|212|ϕη|4=(|ϕ||ϕη|)(m(m2)m(m1)H|ϕ|212(|ϕ||ϕη|)(|ϕ|+|ϕη|)2)+|ϕ|2(|ϕ|2m(m2)m(m1)H|ϕ|+mH2).(73)
Hence, by replacing (73) into Proposition 1.2,
L(mH)|A|2m2|H|2+cm|ϕN|22cmα|ϕα(T)|2cmHϕη(T),T+(|ϕ||ϕη|)(m(m2)m(m1)H|ϕ|212(|ϕ||ϕη|)(|ϕ|+|ϕη|)2)+|ϕ|2(|ϕ|2m(m2)m(m1)H|ϕ|+c(m|T|2)+mH2).(74)
On the other hand, from (20) and (37), we write
H2=1m(m1)|ϕ|2+aH+b,(75)
and since a,b0 we obtain
H1m(m1)|ϕ|.(76)
Moreover, the following inequality is well known (see Equation 3.5 of Guo & Li 2013GUO X & LI H. 2013. Submanifolds with constant scalar curvature in a unit sphere. Tohoku Math J 65(3): 331-339.)
(|ϕ||ϕη|)(|ϕ|+|ϕη|)23227|ϕ|3.(77)
Thus, from (76) and (77) we conclude that
m(m2)m(m1)H|ϕ|212(|ϕ||ϕη|)(|ϕ|+|ϕη|)2(m2m11627)|ϕ|3.(78)
Assuming that m4,
m2m11627>0.(79)
Therefore, from (78) and (79), (74) becomes
L(mH)|A|2m2|H|2+cm|ϕN|22cmα|ϕα(T)|2cmHϕη(T),T+|ϕ|2(|ϕ|2m(m2)m(m1)H|ϕ|+c(m|T|2)+mH2).(80)
On the other hand, from (50) we have mH(m1)a>0. Since m4, it follows that Ha21m(mH(m1)a)>0. Consequently, by making a direct computation, (75) can be written as follows:
Ha2=|ϕ|2m(m1)+b+a24.(81)
By using this and (75) we can write
|ϕ|2m(m2)m(m1)H|ϕ|+c(m|T|2)+mH2=m2m1|ϕ|2m(m2)m(m1)|ϕ|(|ϕ|2m(m1)+b+a24+a2)+ma|ϕ|2m(m1)+b+a24+m(a22+b)+c(m|T|2)=|ϕ|2φa,b,c(|ϕ|,|T|),(82)
where φa,b,c is a real function defined in (68). Since b0, Lemma 2.1 assures that
|A|2m2|H|20.(83)
Therefore, inserting (83) and (82) into (80), we obtain.
L(mH)cm|ϕN|2cmHϕη(T),T2mcα|ϕα(T)|2|ϕ|2φa,b,c(|ϕ|,|T|),(84)
where φa,b,c is defined in (68).

Now, Lemma 2.2 guarantees that the operator P is positive definite since b0. So, by (33) and (75), we can write , we can write

1m1L(|ϕ|2)=2HL(mH)+2mP(H),HaL(mH)2(Ha2)L(mH).(85)
Hence, by inserting (84) in (85), we get (66).

Finally, if equality holds in (66), considering that b>0 and P is positive definite, we can deduce from (85) that H is constant. Moreover, (83) must also be satisfied as an equality. Since we already established that H is constant, this implies A=0, indicating that the second fundamental form is parallel. Additionally, in order to achieve equality in Lemma 2.5, (70) must also be an equality. Consequently, we conclude that Σm is a parallel submanifold of Mn(c)× with exactly two distinct principal curvatures, one of which is simple. ◻

Remark 3.2.. Since the mean curvature vector field is normalized, it follows that H>0. By using (75),

|ϕ|2=m(m1)(H2aHb).(86)
If a=b=0 and there exists a point pΣm such that |ϕ|(p)=0, then H must vanish, which contradicts the fact that H>0. Therefore, we conclude that |ϕ|, a, and b cannot vanish simultaneously.

Now, we are ready to give proof of our first result.

Theorem 3.3.. Let Σm be a closed pnmc lw-submanifold in 𝕊n×, n>m4, such that H2=aH+b with a,b0. If Σm has constant η-angle, then

Σ|ϕ|p+2a,b(|ϕ|,|T|)dΣ0,(87)
for every real number p>2, where a,b is the real function given by

a , b ( x , y ) = m 2 m 1 x 2 + ( 2 m + 1 ) y 2 m ( a m 2 m ( m 1 ) x ) x 2 m ( m 1 ) + b + a 2 4 + m ( m 2 ) a 2 m ( m 1 ) x m ( a 2 2 + b + 1 ) . (88)

Moreover, if b>0 the equality holds in (87) if and only if:

  • either Σm is a totally umbilical hypersurface in 𝕊m+1×{t0}𝕊n× for some t0;

  • or |ϕ|2=γ(m,a,b), where γ(m,a,b) is a positive constant depending only on m,a,b and Σm is isometric to a standard product 𝕊1(1r2)×𝕊m1(r)𝕊m+1×{t0}𝕊n× for some t0, with r=(m1)/m(H2+1)>0.

Proof. Firstly, let us take c=1 in Proposition 3.1. By using Cauchy-Schwarz inequality, we get

2mα|ϕα(T)|22mα|ϕα|2|T|2=2m|ϕ|2|T|2.(89)
On the other hand, since If Σm has constant η-angle and η=em+1 is parallel, by (16),
0=Xη,t=Xη,N+η,XN=A(T,X),η=Aη(T),X,(90)
for all X𝔛(M). So, from (24), ϕη(T)=HT. Thus, from (89),
𝒬1=m|ϕN|22mα|ϕα(T)|2mHϕη(T),Tm|ϕN|2+mH2|T|22m|ϕ|2|T|22m|ϕ|2|T|2,(91)
with equality holding if and only if ϕN=T=0. Thus, inserting (91) in (66), we obtain
L(|ϕ|2)2(m1)|ϕ|2a,b(|ϕ|,|T|)|ϕ|2m(m1)+b+a24,(92)
where a,b(x,y) is given in (88).

From now on, for simplicity, we will denote u=|ϕ|2. So, (85) can be rewritten as follows

L(u)m1mua,b(u,|T|)4u+m(m1)(a2+4b).(93)
Taking into account that u0 and b0, from Remark 3.2 and (93) we get
up+22a,b(u,|T|)mm1up24u+m(m1)(a2+4b)L(u),(94)
for every real number p. By closedness of Σm, we can integrate both sides of (94) in order to obtain
Σup+22a,b(u,|T|)dΣmm1Σup24u+m(m1)(a2+4b)L(u)dΣ.(95)

Now, we will define the function

f(t)=t0tg(s)ds,(96)
where g(s) is given by
g(s)=sp/24s+m(m1)(a2+4b),s0.(97)
Since p>2, b0 and g is a smooth function, we have that f is well defined (see Remark 3.2) and f0. Hence, taking into the integral, from (34) and Lemma 2.3, we have
0=ΣL(f(u))dΣ=Σf(u)L(u)dΣ+Σf(u)P(u),udΣ,(98)
that is,
Σf(u)L(u)dΣ=Σf(u)P(u),udΣ.(99)
Taking the first and second derivatives of (96), we have
f(t)=tp/24t+m(m1)(a2+4b)0(100)
and
f(t)=4(p1)tp/2+pm(m1)(a2+4b)tp222(4t+m(m1)(a2+4b))3/20.(101)
Lemma 2.2 assures that the operator P is positive semidefinite, using (99), (100) and (101) in (95), we can estimate
Σup+22a,b(u,|T|)dΣmm1Σf(u)P(u),udΣ0.(102)
Therefore, we conclude
Σup+22a,b(u,|T|)dΣ0.(103)
This proves inequality (87).

We assume that the equality holds in (103) and b>0. By (102), we get

Σf(u)P(u),udΣ=0,(104)
where
f(u)=4(p1)up/2+pm(m1)(a2+4b)up222(4u+m(m1)(a2+4b))3/20,(105)
with equality holding if and only if p>2 and u=0. Since b>0, from Lemma 2.2, P is positive definite, consequently
P(u),u0(106)
with equality if and only if u=0. Therefore, it follows from (104) that:
f(u)P(u),u=0onΣm,(107)
which implies that the function u=|ϕ|2 must be constant, either u0 or uu0>0.

If |ϕ|2=u0, then Σm is a totally umbilical submanifold. Hence, by (91) we get T=0. Otherwise, if |ϕ|2=uu0>0. From this, the equality in (103) implies in

Σa,b(|ϕ|,|T|)dΣ=0.(108)
Hence, from (91) we also must have ϕN=0 and T=0 in the non-totally umbilical case. Thus, by this, (88) can be written as follows, a,b=const. and follows by (108) that a,b=0. Consequently, we must have that |ϕ|2=γ(m,a,b), where γ(m,a,b) is the only positive root of φa,b,1. From this, we are able to see that all inequalities obtained along of the proof become equalities. In particular, the equality holds in (70) and (83). So, from Lemmas 2.1 and 2.5 we must have that Σm is a parallel submanifold in 𝕊n with two distinct principal curvatures one of which is simple. Besides this, also occurs the equality in (80), which implies |ϕ|=|ϕη|. In both cases, |ϕ|=0 or |ϕ|=|ϕη|, we can always get that ϕα=0 for all α>m+1. By using this, since n>m, if n=m+1, then Σm𝕊m+1× and as T=0, we obtain that Σm is a hypersurface of 𝕊m+1×{t0} for some t0. So, let us assume then n>m+1. Once Aα=0 for all αm+2, we observe that the first normal subspace
N1={ξTM;Aξ=0}=span{η},(109)
has dimension 1 and N1L=N1+span{N}. Since η is orthogonal to t we have that rank(L)=q=2. Finally, we observe that the condition nm+1>2=q is satisfied. Therefore we can apply Lemma 2.6 in order to obtain that Σm lies in a totally geodesic submanifold 𝕊m+1× of 𝕊n×. So, we can conclude that Σm is an isoparametric hypersurface in 𝕊m+1×{t0}, for some t0, with at most two distinct principal curvatures. Therefore, we can use Theorem 1.1 of dos Santos & da Silva 2021 (see also Theorem 1 of Alías et al. 2012) in order to conclude that Σm must be isometric to the following standard product 𝕊1(1r2)×𝕊m1(r) with r=(m2)/m(H2+1). ◻

In the case c=1, we have:

Theorem 3.4.. Let Σm be a closed pnmc lw-submanifold in n×, n>m4, such that H2=aH+b with a,b0. If Σm has constant η-angle, then

Σ|ϕ|p+2𝒢a,b(|ϕ|,|T|)dΣ0,(110)
for every real number p>2, where 𝒢a,b is given by
𝒢a,b(x,y)=m2m1x2(m+1)y2m(am2m(m1)x)x2m(m1)+b+a24+m(m2)a2m(m1)xm(a22+b2).(111)
Moreover, if b>0 the equality holds in (110) if and only if Σm is a totally umbilical hypersurface in m+1×{t0}n× for some t0.

Proof. Let us consider a local orthonormal frame field {em+1,,en+1} in the normal bundle such that em+1=η. Then, from (19), it is easy to see that

ϕN=α=m+1n+1N,eαϕα.(112)
From Cauchy-Schwarz’s inequality and Hilbert-Schmidt’s norm definition, we have
|ϕN|2=α,iN,eα2ϕα(ei),ϕα(ei)α,i|N|2|eα|2ϕα(ei),ϕα(ei)=|N|2|ϕ|2.(113)
Hence, from (14), (90) and (113),
𝒬1=m|ϕN|2+mHϕη(T),T+2mα|ϕα(T)|2m|N|2|ϕ|2+mHϕη(T),T+2m|ϕη(T)|2+2mα>m+1|ϕα(T)|2m(1|T|2)|ϕ|2+mHϕη(T),T+2m|ϕη(T)|2=m(1|T|2)|ϕ|2mH2|T|2+2mH2|T|2=m(1|T|2)|ϕ|2+mH2|T|2m(1|T|2)|ϕ|2,(114)
with equality holding if and only if T=0. Thus, inserting (114) in (66),
L(|ϕ|2)2(m1)|ϕ|2𝒢a,b(|ϕ|,|T|)|ϕ|2m(m1)+b+a24,(115)
where 𝒢a,b(x,y)is given in (111).

At this point the proof follows as the one of Theorem 3.3 until reaching inequality (110). If the equality holds, from (104) we have

P(|ϕ|2),|ϕ|2=0,(116)
since
f(|ϕ|)=4(p1)|ϕ|p+m(m1)(4b+a2)p|ϕ|p22(4|ϕ|2+m(m1)(4b+a2))3/2>0,
where it was used that p>2 and b>0. Hence, being P positive definite, from (116) it follows that |ϕ| is constant along of Mn. If |ϕ|=0, then Mn is a totally umbilical, and as before, Σm hypersurface in m+1n×{t0}, for some t0. Otherwise, |ϕ| is a positive constant. So, from the equality (110),
Σ𝒢a,b(|ϕ|,|T|)dΣ=0.(117)
Therefore, reasoning as in the last part of Theorem 3.3, we must have that Σm is an isoparametric hypersurface of m+1×{t0}n×, for some t0. So, taking into account Theorem 2 of Alías et al. (2012), we conclude that Mn should be isometric to a hyperbolic cylinder 1(1+r2)×𝕊n1(r), which is not closed manifold. Therefore, the equality holds in (110) if, and only, Σm is a totally umbilical hypersurface in m+1. ◻

Remark 3.5.. Let us recall that a submanifold Σm of Mn(c)× is said to be a vertical cylinder over Mm1 if Σm=πM1(Mm1) where Mm1 is a submanifold of Mn(c). It is not difficult to check that Σm is a non-minimal parallel vertical cylinder in Mn(c)× if, and only if, Mm1 is a non-minimal parallel submanifold in Mn(c). Moreover, its mean curvature vector field h is given by h=m1mh0, where h0 denotes the mean curvature vector field of Mm1. Hence, Σm is a pnmc lw-submanifold of Mn(c)× having constant η-angle and that is not lies in a slice provided vertical cylinders are characterized by the fact that t is always tangent to Σm (see Fetcu & Rosenberg 2013FETCU D & ROSENBERG H. 2013. On complete submanifolds with parallel mean curvature in product spaces. Rev Mat Iberoam 29: 1283-1304.). Therefore, we conclude that the hypothesis of the submanifold to be closed in Theorems 3.3 and 3.4 is, indeed, necessary.

Further results for m=2 and m=3

We should notice that when m=2 and m=3, the integral inequalities obtained in Theorems 3.3 and 3.4 is, indeed, necessary. holds. To see this, it is sufficient to do a rereading on the first inequality of (72). In fact, from (72),

mHαtr(ϕα2ϕm)α,βm+1N(ϕαϕβϕβϕα)α[tr(ϕαϕβ)]2m(m2)m(m1)H|ϕ|2|ϕη|12|ϕη|4+|ϕ|2|ϕη|232|ϕ|4.(118)
A straightforward computation, gives
|ϕ|2|ϕη|212|ϕη|412|ϕ|4,(119)
with equality holding if and only if |ϕ|=|ϕη|=0, that is, if and only if Σm is totally umbilical submanifold. Besides this,
|ϕ|2=|ϕη|2+α>m+1|ϕα|2|ϕη|2,(120)
with equality holding if and only if |ϕα|=0 for α>m+1. Hence, inserting (119) and (120) in (118), we get
mHαtr(ϕα2ϕη)α,βm+1N(ϕαϕβϕβϕα)α[tr(ϕαϕβ)]2|ϕ|2(m(m2)m(m1)H|ϕ|+3|ϕ|2).
By using this in Proposition 1.2,
L(mH)|A|2m2|H|2+cm|ϕN|22cmα|ϕα(T)|2cmHϕη(T),T+(c(m|T|2)+mH2m(m2)m(m1)H|ϕ|3|ϕ|2)|ϕ|2.(121)
Hence, by replacing (75) and (81) in (121), we have
L(mH)|A|2m2|H|2+𝒬c|ϕ|2φ¯a,b,c(|ϕ|,|T|),(122)
where
φ¯a,b,c(x,y)=2x2+φa,b,c(x,y)(123)
with 𝒬c and φa,b,c(x,y) defined in (67) and (68), respectively.

Therefore, since b0, we can apply Lemma 2.1 together with inequality (85) in (122) in order to obtain:

L(|ϕ|2)2(m1)(|ϕ|2φ¯a,b,c(|ϕ|,|T|)𝒬c)|ϕ|2m(m1)+b+a24.(124)

By this previous digression, we obtain:

Theorem 4.1.. Let Σm be a closed pnmc lw-submanifold in 𝕊n×, n>m, such that H2=aH+b with a,b0. If Σm has constant η-angle, then

Σ|ϕ|p+2¯a,b(|ϕ|,|T|)dΣ0,(125)
for every real number p>2, where ¯a,b is the real function given by
¯a,b(x,y)=2x2+a,b(x,y),
with a,b(x,y) defined in (88). Moreover, the equality holds in (125) if and only if Σm is a totally umbilical hypersurface in 𝕊m+1×{t0}𝕊n× for some t0.

Proof. TThe proof follows the same steps as the proof of Theorem 3.3 until we reach inequality (103), changing the function φa,b,c by φ¯a,b,c along of the computations. If the equality in (125) holds, then also occurs equality in (119) and hence, Σm is a totally umbilical. Besides this, the equality also occurs in (91), from where we conclude that T=0. Therefore, Σm is a totally umbilical hypersurface in 𝕊m+1×{t0}𝕊n× for some t0. ◻

Following the same steps of the proof of Theorems 3.4 and 4.1, we have:

Theorem 4.2.. Let Σm be a closed pnmc lw-submanifold in n×, n>m, such that H2=aH+b with a,b0. If Σm has constant η-angle, then

Σ|ϕ|p+2𝒢¯a,b(|ϕ|,|T|)dΣ0,(126)
for every real number p>2, where 𝒢¯a,b is given by
𝒢¯a,b(x,y)=2x2+𝒢a,b(x,y),
with 𝒢a,b(x,y) defined in (111). Moreover, the equality holds in (126) if and only if Σm is a totally umbilical hypersurface in m+1×{t0}n× for some t0.

Remark 4.3.. The approach developed here is not effective to find parallel submanifolds with two distinct principal curvatures as in Proposition 3.1, because of this, we use it only in the cases where m=2 and m=3. So, following Remark 3.2 of Guo & Li (2013)GUO X & LI H. 2013. Submanifolds with constant scalar curvature in a unit sphere. Tohoku Math J 65(3): 331-339., it is an interesting question is to know if Proposition 3.1 holds or not for m=2 and m=3.

ACKNOWLEDGMENTS

The authors would like to express their thanks to the referees for reading the manuscript in great detail and for the valuable suggestions and comments that helped to improve the paper. The first author is partially supported by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Brazil, grants 431976/2018-0 and 311124/2021-6 and Propesqi (UFPE).

  • 1
    2020 Mathematics Subject Classification: Primary 53C42; Secondary 53A10, 53C20.
  • 2
    We adopt for the (1,3)-curvature tensor the following definition of Chapter 3 of O’Neill (1983)O’NEILL B. 1983. Semi-Riemannian Geometry, with Applications to Relativity. New York: Academic Press.: R¯(X,Y)Z=¯[X,Y]Z[¯X,¯Y]Z.
  • ALÍAS LJ, GARCÍA-MARTÍNEZ SC & RIGOLI M. 2012. A maximum principle for hypersurfaces with constant scalar curvature and applications. Ann Glob Anal Geom 41: 307-320.
  • ALÍAS LJ & MELÉNDEZ. 2020. Integral Inequalities for Compact Hypersurfaces with Constant Scalar Curvature in the Euclidean Sphere. Mediterr J Math 17: 61.
  • CAO L & LI H. 2007. r-Minimal submanifolds in space forms. Ann Glob Anal Geom 32: 311-341.
  • CHENG SY & YAU ST. 1977. Hypersurfaces with constant scalar curvature. Math Ann 225: 195-204.
  • CHERN SS, DO CARMO MP & KOBAYASHI S. 1970. Minimal Submanifolds of a Sphere with Second Fundamental Form of Constant Length. Springer Berlin Heidelberg, 59-75 p.
  • DANIEL B. 2007. Isometric immersions into 3-dimensional homogeneous manifolds. Commentarii Math Helv 82: 87-131.
  • DILLEN F, FASTENAKELS J, VAN DER VEKEN J & VRANCKEN L. 2007. Constant angle surfaces in 𝕊 2 × ℝ . Monatsh Math 152: 89-96.
  • DILLEN F & MUNTEANU MI. 2009. Constant angle surfaces in ℍ 2 × ℝ . Bull Braz Math Soc 40: 85-97.
  • DOS SANTOS FR. 2021. Rigidity of Surfaces with Constant Extrinsic Curvature in Riemannian Product Spaces. Bull Braz Math Soc, New Series 52: 307-326.
  • DOS SANTOS FR & DA SILVA SF. 2021. A Simons type integral inequality for closed submanifolds in the product space 𝕊 n × ℝ . Nonlinear Anal 209: 112366.
  • DOS SANTOS FR & DA SILVA SF. 2022. On complete submanifolds with parallel normalized mean curvature in product spaces. Proc Roy Soc Edinburgh Sect A 152(2): 331–355.
  • FETCU D & ROSENBERG H. 2013. On complete submanifolds with parallel mean curvature in product spaces. Rev Mat Iberoam 29: 1283-1304.
  • GROSJEAN JF. 2002. Upper bounds for the first eigenvalue of the Laplacian on closed submanifolds. Pacific J Math 206: 93-112.
  • GUO X & LI H. 2013. Submanifolds with constant scalar curvature in a unit sphere. Tohoku Math J 65(3): 331-339.
  • LAWSON HB. 1969. Local Rigidity Theorems for Minimal Hypersurfaces. Ann of Math 89(1): 187-197.
  • LI AM & LI JM. 1992. An intrinsic rigidity theorem for minimal submanifolds in a sphere. Arch Math 58: 582-594.
  • LI H. 1996. Hypersurfaces with constant scalar curvature in space forms. Math Ann 305: 665-672.
  • MENDONÇA B & TOJEIRO R. 2014. Umbilical submanifolds of 𝕊 n × ℝ . Canad J Math 66: 400-428.
  • NAVARRO M, RUIZ-HERNÁNDEZ G & SOLIS DA. 2016. Constant mean curvature hypersurfaces with constant angle in semi-Riemannian space forms. Differ Geo and its Appl 49: 473-495.
  • NISTOR AI. 2017. New developments on constant angle property in 𝕊 2 × ℝ . Ann Mat Pura Appl 196: 863-875.
  • O’NEILL B. 1983. Semi-Riemannian Geometry, with Applications to Relativity. New York: Academic Press.
  • ÔTSUKI T. 1970. Minimal hypersurfaces in a Riemannian manifold of constant curvature. Amer Jour Math 92: 145-173.
  • SANTOS W. 1994. Submanifolds with parallel mean curvature vector in spheres. Tohoku Math J 46: 403-415.
  • SIMONS J. 1968. Minimal varieties in Riemannian manifolds. Ann Math 88: 62-105.
  • VAN DER VEKEN J & VRANCKEN L. 2008. Parallel and semi-parallel hypersurfaces of 𝕊 n × ℝ . Bull Braz Math Soc, New Series 39: 335-370.
  • WEI G. 2008. Simons’ type integral formula for hypersurfaces in a unit sphere. J Math Anal Appl 340: 1371-179.

Publication Dates

  • Publication in this collection
    30 Oct 2023
  • Date of issue
    2023

History

  • Received
    01 Apr 2023
  • Accepted
    24 May 2023
Academia Brasileira de Ciências Rua Anfilófio de Carvalho, 29, 3º andar, 20030-060 Rio de Janeiro RJ Brasil, Tel: +55 21 3907-8100 - Rio de Janeiro - RJ - Brazil
E-mail: aabc@abc.org.br