We prove that the first nonzero eigenvalue of the Laplace-Beltrami operator of equator-like minimal submanifold embedded in the sphere Sn+1 is equal to n. The proof uses the spectral properties of the heat kernel operator corresponding to the submanifold.
Citation: Ibrahim Aldayel. Value of first eigenvalue of some minimal hypersurfaces embedded in the unit sphere[J]. AIMS Mathematics, 2023, 8(11): 26532-26542. doi: 10.3934/math.20231355
[1] | Meraj Ali Khan, Ali H. Alkhaldi, Mohd. Aquib . Estimation of eigenvalues for the $ \alpha $-Laplace operator on pseudo-slant submanifolds of generalized Sasakian space forms. AIMS Mathematics, 2022, 7(9): 16054-16066. doi: 10.3934/math.2022879 |
[2] | Yanlin Li, Mehraj Ahmad Lone, Umair Ali Wani . Biharmonic submanifolds of Kaehler product manifolds. AIMS Mathematics, 2021, 6(9): 9309-9321. doi: 10.3934/math.2021541 |
[3] | Noura Alhouiti, Fatemah Mofarreh, Fatemah Abdullah Alghamdi, Akram Ali, Piscoran-Ioan Laurian . Geometric topology of CR-warped products in six-dimensional sphere. AIMS Mathematics, 2024, 9(9): 25114-25126. doi: 10.3934/math.20241224 |
[4] | Yanlin Li, Nasser Bin Turki, Sharief Deshmukh, Olga Belova . Euclidean hypersurfaces isometric to spheres. AIMS Mathematics, 2024, 9(10): 28306-28319. doi: 10.3934/math.20241373 |
[5] | Bangchao Yin, Shujie Zhai . Classification of Möbius minimal and Möbius isotropic hypersurfaces in $ \mathbb{S}^{5} $. AIMS Mathematics, 2021, 6(8): 8426-8452. doi: 10.3934/math.2021489 |
[6] | Shahroud Azami . Monotonicity of eigenvalues of Witten-Laplace operator along the Ricci-Bourguignon flow. AIMS Mathematics, 2017, 2(2): 230-243. doi: 10.3934/Math.2017.2.230 |
[7] | Ibrahim Al-dayel . Investigating the characteristics of Clifford hypersurfaces and the unit sphere via a minimal immersion in $ S^{n+1} $. AIMS Mathematics, 2024, 9(10): 26951-26960. doi: 10.3934/math.20241311 |
[8] | Lamia Saeed Alqahtani, Akram Ali . The eigenvalues of $ \beta $-Laplacian of slant submanifolds in complex space forms. AIMS Mathematics, 2024, 9(2): 3426-3439. doi: 10.3934/math.2024168 |
[9] | Yanlin Li, Erhan Güler, Magdalena Toda . Family of right conoid hypersurfaces with light-like axis in Minkowski four-space. AIMS Mathematics, 2024, 9(7): 18732-18745. doi: 10.3934/math.2024911 |
[10] | Zhanbing Bai, Wen Lian, Yongfang Wei, Sujing Sun . Solvability for some fourth order two-point boundary value problems. AIMS Mathematics, 2020, 5(5): 4983-4994. doi: 10.3934/math.2020319 |
We prove that the first nonzero eigenvalue of the Laplace-Beltrami operator of equator-like minimal submanifold embedded in the sphere Sn+1 is equal to n. The proof uses the spectral properties of the heat kernel operator corresponding to the submanifold.
In the theory of minimal submanifolds in a sphere, an interesting question asks about the value of first nonzero eigenvalue of the Laplacian for a minimal hypersurface Σ embedded in (n+2)-unit sphere Sn+1 in Rn+2. In its list of famous problems, the following question has been raised by S. T. Yau (problem 100, [17]).
Conjecture: [17] Let Σ be a minimal hypersurface embedded in the n+1-unit sphere Sn+1. Then, λ1(Σ)=n.
The upper bound λ1(Σ)≤n is not obvious, and was obtained before the statement of the conjecture due to Takahashi [15]. Just after the conjecture was published, Choi and Wang proved that λ1(Σ)≥n/2. In fact, they proved a more general statement based on Reilly's formula, see [9]. Until this day, it was the best known lower bound in the general case. Many important steps towards this conjecture has been done by proving the conjecture for some minimal homogenous hypersurfaces due to Muto-Ohnita-Urakawa [12], Kotani [10] and Solomon [13,14]. Recently, Z. Tang and W. Yan proved that the conjecture is valid for closed minimal isoparametric hypersurfaces [16]. In a recent work, S Deshmukh has proved some results related to the conjecture in [6]. For the case when λ1(Σ)<n, it is shown that one has the following alternative, either λ1(Σ)≤(1+k0)n or λ1(Σ)≥n+(nk0−n+1)n2. In the opposite case, when λ1(Σ)=n, either Σ is isometric to the unit sphere Sn or otherwise k0≤n−1/n. A generalization of this work for pseudo-umbilical hypersurface in the unit sphere has been proved by M. A. Choudhary in [2].
The method we are going to use in the paper are very different to the previous works, which have studied in this topic. Indeed, we are going to focus on the spectral properties of the Laplacian of a special type of immersed minimal submanifolds in the unit sphere. One of the most important objects in spectral geometry is the heat kernel operator associated with a given Riemmanian manifold, which corresponds to the solution of the heat equation on the manifold. The first nonzero eigenvalue controls the rate of growth of the heat kernel when time tends to infinity.
Let Σ be the hypersurface given by the locus of vanishing of some smooth function ψ on the unit sphere Sn+1 i.e.,
Σ={x∈Sn+1|ψ(x)=0}. |
We assume that Σ is embedded in Sn+1, which amounts to say that the gradient of ψ never vanishes on Σ. Thus, Σ is Riemannian submanifold on Sn+1, in particular it has an induced metric which gives rise to the corresponding Laplace-Beltrami operator ΔΣ which is also self-adjoint. The spectrum of ΔΣ which is discrete, has a least nonzero eigenvalue, λ1(Σ). Let us consider the polar coordinates parametrization, (σ,θ) coming from the stereographic projection. Suppose ψ is chosen so that Σ is a minimal embedded hypersurface in the n+1-unit sphere Sn+1. Then, one has the following positive answer to the conjecture for minimal submanifolds satisfying some tranversality conditions.
Theorem 1.1. Let Σ a minimal embedded hypersurface in the n+1-unit sphere Sn+1 and assume that the normal bundle of Σ is a one-dimensional subspace of T(Sn+1) generated by the vector field ∂θ coming from the stereographic projection. Then,
λ1(Σ)=n. |
We begin by some classical results that we are going to need. These can be found in many places (see e.g., [1,8]). Given an n-dimensional Riemannian manifold (M,g), one can define the Laplace-Beltrami operator which acts on smooth functions over M. In the local coordinate around the point x=(x1,…,xn) with associated frame (∂1,…,∂n) which forms a basis of the tangent space Tx(M), the Laplace-Beltrami operator takes the following form
ΔM.g=1√gn∑i=1∂i(√gn∑j=1gij∂j), | (2.1) |
where g=det(gij) and (gij)=(gij)−1. Assuming that M is a compact makes the operator −ΔM being a self-adjoint operators in L2(M). In particular, it has a discrete spectrum given by
0=λ0(M)<λ1(M)≤λ2(M)≤… |
The spectral decompositon of the Hilbert space L2(M) with respect to ΔM allows to write any function f∈L2(M) as
f=∑k≥0⟨f,Φk⟩L2(M)Φk, |
where (Φk)k≥0 is a basis of eigenfunctions of L2(M). Associated to this, one can strongly continous operator in L2(M), Pt=e−tΔM satisfying the property Ps+t=Ps∘Pt such that ‖Pt‖≤1. It can be proved see e.g., [3] that the operator Pt has a kernel Kt:M×M→R for all t≥0. This means that, for any function L2(M)
Ptf(x)=∫MKt(x,y)f(y)volg(dy). |
The heat kernel characterizes the heat operator, and it can be obtained to perform the following evaluation
Pt(δy)(x)=Kt(x,y). |
The latter evaluation is allowed since one can identify the regular distribution with the function whenever it is continous, which is the case for the kernel operator. In can be shown that the function u(t,x):=Ptf(x) satisfies the following equation with initial Dirichlet boundary condition
{ΔMu(t,x)+∂∂tu(t,x)=0,u(0,x)=f(x) on ∂M. | (2.2) |
The spectral decomposition of the heat kernel is given by
Kt(x,y)=∑k≥0eλk(M)tΦk(x)Φk(y), | (2.3) |
for every x,y∈M. The exponential growth of Kt(x,y) is controlled by the first eigenvalue λ1(M), which is nonzero for compact Riemannian manifolds with Dirichlet initial value condition. A fundamental result for long time behavior of the heat kernel it that for every x,y∈M one has (see e.g., [11])
limt→∞logKt(x,y)t=λ1(M). | (2.4) |
The non-nullity of the first eigenvalue is granted by the fact that we consider the heat equation on a bounded domain of the sphere with Dirichlet boundary conditions. In fact, we can say much more about it since we are going to work with embedded closed minimal hypersurfaces in the unit sphere Sn+1. For this class of domains, the first eigenvalue is quite large in a certain sense, since it was proved by Choi and Wang that λ1(M)≥n2 [9]. In particular, the first eigenvalue is not zero.
Yau's conjecture predicts that this value is maximal, in that λ1(M)=n for such hypersurfaces. Thus, the minimality condition for an hypersurface on the unit sphere implies maximality of the first eigenvalue.
For the sphere M=Sn+1, the eigenvalues of the operator (−ΔSn+1) acting on L2(Sn+1) are given by
λl(Sn+1)=l(n+l), | (2.5) |
in particular the first eigenvalue is given by
λ1(Sn+1)=n+1. | (2.6) |
In particular, using 2.4 and 2.6 for the unit sphere one has asymptotic estimate as t tends to infinity
KSn+1t(x,x)∼e−(n+1)tϕ1(x)2. | (2.7) |
Let us consider the stereographic projection π of the sphere Sn+1 on Rn+1 relatively to the north pole N=en+2=(0,…,0,1)∈Sn+1. It is given by the rule
π(x)=11−xn+2(x1,…,xn+1), |
provided x=(x1,…,xn+2) is not N. Let us set σ(x)=π(x)‖π(x)‖, this defines an element of Sn. Also one defines a map θ:Sn+1→[0,π] by assigning to each x∈Sn+1, the angle θ(x)=2^(→NO,→Nx). One can explicit a formula for θ, indeed one has
‖→NO‖‖→Nx‖cos(θ2)=⟨→N0,→Nx⟩. |
In terms of the coordinates x=(x1,…,xn+2) and →NO=−en+2=(0,…,−1), the previous equality gives
√x21+…+x2n+1+(xn+2−1)2cos(θ(x)2)=1−xn+2. |
Thus,
θ(x)=2cos−1(1−xn+2√x21+…+x2n+1+(1−xn+2)2)∈[0,π). |
We are able to product a diffeomorphism ψ:Sn+1−{N}→Sn×[0,π] by setting
ψ(x)=(π(x)‖π(x)‖,θ(x)). |
This gives the well-known realization of the unit sphere Sn+1 minus N as the product Sn×[0,π] with the polar coordinates x=(σ,θ).
By now on, we use the parametrization of the unit sphere minus the north pole using the change of coordinates (x1,…,xn+1)↔(σ1,…,σn+1) where σn+1=θ. The length element is given by
dx2|Sn+1=sin2θdσ2|Sn+dθ2. |
The metric g|Sn+1 in the local coordinates (σ1,…,σn+1,θ) is given by the diagonal matrix
(g|Sn+1)ij=[sin2θ0…00⋱⋮⋮sin2θ00…01]. |
The metric g|Sn+1 gives rise to the Christoffel symbols given by
Γkij=12∑1⩽l⩽n+2gkl{∂j(gil)+∂i(gjl)−∂l(gij)} (1⩽i,j,k⩽n+1). |
We can use this coefficients to define a connection on the tangent bundle of Sn+1. Let x∈Sn+1−N with local coordinates x=(σ1,…,σn,θ) and corresponding orthonormal frame {∂1,…,∂n,∂θ} with respect to the metric g|Sn+1 that is, for every i,j=1,…,n+1
g|Sn+1(∂i,∂j)=δji. |
Using the basis {∂1,…,∂n,∂θ} of Tx(Sn+1) we are able to define a bilinear map
∇=∇Sn+1 : T(Sn+1)×T(Sn+1)→T(Sn+1), |
by assigning the values taken by this form at the elements of the basis of T(Sn+1) by introducing the coefficients,
∇∂i∂j=n+1∑k=1Γkij∂k, |
where we have denoted ∂n+1=∂θ. The operator ∇ therefore defines a connection on the tangent bundle T(Sn−1). Basic computations show that the Christoffel symbols relative to the metric g are symmetric in the sense Γkij=Γkji for all 1⩽i,j,k⩽n+1. This implies that ∇ is torsion-free i.e., ∇XY−∇YX=[X,Y] and by uniqueness, ∇ is the Levi-Civita connection on T(Sn+1).
We consider the unit sphere equipped with the polar coordinates system (σ,θ) introduced in the previous paragraph. Thus, the hypersurface Σ in the coordinate system (σ,θ) is defined as follows
Σ={(σ1,…,σn,θ)∈Sn×[0,π] | ψ(σ1,…,σn,θ)=0}. |
Since Σ is embedded in Sn+1, the chain rule implies that gradient of ψ satisfies ∇ψ(x)≠0 for any x∈Σ in the polar coordinates. The hypersurface Σ inherits a structure of Riemannian manifold given by a metric gΣ which the one induced by g|Sn+1 with associated volume Riemmanian form dvolΣ=√gΣ(σ,θ)dσ∧dθ which we do not need to explicit. In the local coordinates, we can assume that {∂∂σ1,…,∂∂σn} is an orthonormal frame of T(Σ) whereas ξ=∂∂θ generates the normal bundle N(Σ). The vector fields (∂∂σi) are simply denoted ∂i for 1⩽i⩽n and sometimes we will denote either ∂n+1 or ∂θ the vector field ∂∂θ. With these notations, we have the orthonormal frame for T(Sn+1)={∂1,…,∂n,∂θ} which extends the tangent bundle T(Σ)={∂1,…,∂n}. The tangent space of T(Sn+1) can be splited as folllows:
T(Sn+1)=T(Σ)⊕N(Σ). |
Since Σ is a smooth hypersurface, namely of codimension one, N(Σ) is a line bundle over Σ which is generated by the normal vector field ξ=∂θ. One can give an explicit expression for ξ, the metric gΣ and the mean curvature of Σ in function of the derivatives of u with respect to the frame {∂1,…,∂n}, but we will not need it. Instead, we will use the general expression of the mean curvature in terms of the connections on Σ and Sn+1. Let us denote ∇Sn+1 (resp. ∇Σ) the Levi-Civita connection of Sn+1 (resp. Σ) relative to the metric g and the induced metric g|Σ given in polar coordinates. The second fundamental form ⅡΣ of Σ in Sn+1 is defined by
∇Sn+1XY=∇ΣXY+ⅡΣ(X,Y), |
for any two vectors fields in X,Y∈T(Sn+1). In particular taking {∂1,…,∂n} as an orthonormal basis for T(Σ), the previous relation applied to X=Y=∂i gives us
∇Sn+1∂i∂i=∇Σ∂i∂i+ⅡΣ(∂i,∂i). |
Taking the sum we obtain the fundamental relation
n∑i=1∇Sn+1∂i∂i=n∑i=1∇Σ∂i∂i+HΣ, | (2.8) |
where HΣ=∑ni=1ⅡΣ(∂i,∂i) is the mean curvature vector of Σ.
By translating Σ using a rotation k∈SO(n+1), one can sufficiently rotate the hypersurface Σ so that N∉Σ, that is, Σ⊂Sn+1−{N}. The hypothesis of Theorem 1.1 tells us that T(Σ) ⊥ ∂θ and Σ is the graph of a smooth real valued function u:Sn→[0,π],
Σ={(σ,θ)∈Sn×[0,π]: θ=u(σ)}. | (3.1) |
Since Σ is embedded in Sn+1, the gradient of ψ does not vanish, the implicit function theorem shows that
(σ,θ)∈Σ−C if and only if θ=u(σ), | (3.2) |
in the coordinates (σ,θ). In particular, for any x=(σ,θ)∈V∩Σ−C, one has ψ(σ,u(σ))=0 and at such point x, the hypersurface Σ only depends on the coordinates σ1,…,σn. Thus, Tx(Σ) is a hyperplane in Tx(Sn+1) with orthonormal basis {∂1,…,∂n}. This basis extends to a local frame {∂1,…,∂n,∂θ} of Tx(Sn+1) where ∂n=∂θ. Thus, the normal direction is given by the line Nx(Σ) generated by ∂θ.
We give an explicit expression for the Laplacian of Σ in our setting which can also be found in [4,5].
Lemma 3.1. For every (σ,θ)↦f(σ,θ) smooth function on Sn+1, one has
ΔΣf|Σ=(ΔSn+1f)|Σ−∂2θf. |
Proof. The local orthonormal frame {∂1,…,∂n} of Σ gives the following expression for the Laplacian of Σ
ΔΣ=n∑i=1(∂2i−∇Σ∂i∂i). |
By (2.8), one has
n∑i=1∇Σ∂i∂i=n∑i=1∇Sn+1∂i∂i−HΣ. |
Therefore,
ΔΣ=n∑i=1(∂2i−∇Sn+1∂i∂i)+HΣ. |
Let us write,
ΔΣ=n+1∑i=1(∂2i−∇Sn+1∂i∂i)+HΣ−(∂2∂−∇Sn+1∂θ∂θ). |
The choice of the system of coordinates tells us that ∂n+1=∂θ. We claim that ∇Sn+1∂θ∂θ=0. Indeed one has,
∇Sn+1∂θ∂θ=n+1∑j=1Γjθθ∂j, |
where
Γjθθ=12n∑k=1gkj{∂θ(gθk)+∂θ(gθk)−∂k(gθθ)}. |
Since the inverse metric tensor (gSn+1)kj of Sn+1 is diagonal with gθθ=1 and gkk=1/sin2θ for k=1,…,n. Thus one has, for all 1⩽j⩽n+1
Γjθθ=12gjj{∂θ(gθj)+∂θ(gθj)−∂j(gθθ)}=0. |
Hence, as expected one has
∇Sn−1∂θ∂θ=0. |
Finally, one obtains the Laplace operator on the hypersurface Σ expressed in the local coordinates (σ1,…,σn,θ) in Sn+1
ΔΣ=ΔSn+1+HΣ−∂2θ. | (3.3) |
Since Σ is minimal (i.e., HΣ=0), the previous equality gives us the expected decompostion of the Laplacian
ΔΣ=ΔSn+1−∂2θ. |
Now, let us restrict our attention to the spectrum of ΔΣ. The main task is to find an explicit form of the heat kernel KΣt,(t>0) of Σ. The previous lemma gives rise to the following relation between the heat operators of Σ, Sn+1 and [0,π]
PΣt=e−tΔΣ=e−tΔSn−1+t∂2θ. | (3.4) |
The fact that the two operators ΔSn+1 and ∂2θ commutes, i.e., [ΔSn+1,∂2θ]=0 implies the following:
PΣt=e−tΔSn+1+t∂2θ=e−tΔSn+1et∂2θ=PSn+1tPS1t. | (3.5) |
Note that we have denoted PS1t=et∂2θ the heat operator acting on L2(0,π) in order to emphasis with the fact that the operator ∂2θ acts isopectrally either on L2(S1) and L2(0,π) meaning that their eigenvalues are the same, λl=−l2 for l=0,1,2,… We arrive to the key lemma which gives a formula of the heat kernel of the hypersurface Σ in the (σ,θ)-coordinates of the unit sphere Sn+1.
Lemma 3.2. For any x=(τ,α) and (σ,θ) in Σ, one has
KΣt(x;(σ,θ))=KSn+1t(x;(σ,θ))(∫πβ=0KS1t(θ,β)dβ). |
Proof. Let us denote by (KSn+1t)t>0 (resp. (KSn+1t)t>0) the heat kernel of PSn+1t (resp. PS1t).
Let us consider f∈L2(Sn+1) with support in Σ, then the factorization (3.5) yields
PΣtf(x)=∫(σ,θ)∈ΣKSn+1t(x;(σ,θ))(PS1tf)(σ,θ)dvolΣ(σ,θ). |
Therefore,
PΣtf(x)=∫(σ,θ)∈ΣKSn+1t(x;(σ,θ))(∫πβ=0KS1t(θ,β)f(σ,β)dβ)dvolΣ(σ,θ). |
Sard's Theorem [7] tells us that volΣ(C)=0, in other words, one can restrict the integral to Σ−C
PΣtf(x)=∫(σ,θ)∈Σ−CKSn+1t(x;(σ,θ))(∫πβ=0KS1t(θ,β)f(σ,β)dβ)dvolΣ(σ,θ). |
Using the implicit function u as in (3.2), locally, (σ,θ)∈Σ−C means that u(σ)=θ or equivalently σ∈u−1(θ). In other words, (σ,θ)∈Σ implies that for every β∈[0,π]
f(σ,β)=f(σ,θ). |
The latter fact comes from the fact u is a function, in that, it takes an unique value at each element of Sn.
In particular, provided (σ,θ)∈Σ−C we infer that
∫πβ=0KS1t(θ,β)f(σ,β)dβ=(∫πβ=0KS1t(θ,β)dβ)f(σ,θ). |
Therefore, we obtain the following form of the heat operator on Σ
PΣtf(x)=∫(σ,θ)∈Σ−CKSn+1t(x;(σ,θ))(∫πβ=0KS1t(θ,β)dβ)f(σ,θ)dvolΣ(σ,θ). |
Hence, using duality, we obtain the heat kernel for the hypersurface Σ
KΣt(x;(σ,θ))=KSn+1t(x;(σ,θ))(∫πβ=0KS1t(θ,β)dβ). | (3.6) |
The lemma is proved.
We are ready to prove the theorem, using Lemma 3.2 we get
limt→0+logKΣt(x;(σ,θ))t=limt→0+logKSn+1t(x;(σ,θ))t+limt→0+1tlog(∫πβ=0KS1t(θ,β)dβ), |
for every x=(τ,α) and (σ,θ) in Σ. In view of (2.4), we obtain the equality
λ1(Σ)=λ1(Sn+1)+limt→0+1tlog(∫πβ=0KS1t(θ,β)dβ). | (3.7) |
The spectral expansion of ∂2θ with respect to the orthonormal eigenfunctions in L2(0,π) given by (Φk(x))n≥1=(1√2πsin(kx))k≥1 is given by the uniformly convergent series
KS1t(θ,β)=2π∑k≥1e−k2tsin(kθ)sin(kβ). |
Therefore, using Fubini-Tonelli we can write
∫πβ=0KS1t(θ,β)dβ=2π∑k≥1e−k2tsin(kθ)∫πβ=0sin(kβ)dβ. |
Since ∫πβ=0sin(kβ)dβ=1k(1−(−1)k), we get
∫πβ=0KS1t(θ,β)dβ=2π∑m≥0e−(2m+1)2tsin((2m+1)θ)2m+1. |
The long time asymptotic behaviour of the previous series is controlled by its first term, namely m=0. Thus, we obtain the estimate as t→∞,
∫πβ=0KS1t(θ,β)dβ∼2πe−tsinθ. |
Note that since θ∈(0,π), sin(θ)>0 we are allowed to take the logarithm in order to get
limt→0+1tlog(∫πβ=0KS1t(θ,β)dβ)=−1. |
Thus, (3.7) gives the equality
λ1(Σ)=λ1(Sn+1)−1. | (3.8) |
The fact that λ1(Sn+1)=n+1 gives the required and finishes the proof of Theorem 1.1.
The author declares they have not used Artificial Intelligence (AI) tools in the creation of this article.
The author declares no conflict of interest.
[1] | I. Chavel, Eigenvalues in Riemannian geometry, Orlando: Academic Press, 1984. |
[2] |
M. A. Choudhary, First nonzero eigenvalue of a pseudo-umbilical hypersurface in the unit sphere, Russ. Math., 58 (2014), 56–64. https://doi.org/10.3103/S1066369X14080076 doi: 10.3103/S1066369X14080076
![]() |
[3] |
J. Cheeger, S. T. Yau, A lower bound for the heat kernel, Commun. Pur. Appl. Math., 34 (1981), 465–480. https://doi.org/10.1002/cpa.3160340404 doi: 10.1002/cpa.3160340404
![]() |
[4] |
L. F. Cheung, P. F. Leung, Eigenvalues estimates for submanifolds with bounded mean curvature in the hyperbolic space, Math. Z., 236 (2001), 525–530. https://doi.org/10.1007/PL00004840 doi: 10.1007/PL00004840
![]() |
[5] |
J. Choe, R. Gulliver, Isoperimetric inequalities on minimal submanifolds of space forms, Manuscripta Math., 77 (1992), 169–189. https://doi.org/10.1007/BF02567052 doi: 10.1007/BF02567052
![]() |
[6] |
S. Deshmukh, First nonzero eigenvalue of a minimal hypersurface in the unit sphere, Annali di Mathematica, 191 (2012), 529–537. https://doi.org/10.1007/s10231-011-0194-1 doi: 10.1007/s10231-011-0194-1
![]() |
[7] | H. Federer, Geometric measure theory, Berlin Heidelberg: Springer-Verlag, 1996. https://doi.org/10.1007/978-3-642-62010-2 |
[8] | A. Grigor'yan, Heat kernel and analysis on manifolds, Washington: American Mathematical Society/International Press, 2009. |
[9] | H. I. Choi, A. N. Wang, A first eigenvalue estimate for minimal hypersurfaces, J. Differ. Geom., 18 (1983), 559–562. |
[10] |
M. Kotani, The first eigenvalue of homogeneous minimal hypersurfaces in a unit sphere Sn+1(1), Tohoku Math. J., 37 (1985), 523–532. https://doi.org/10.2748/tmj/1178228592 doi: 10.2748/tmj/1178228592
![]() |
[11] |
P. Li, Large time behaviour of the heat equation on complete manifolds with non-negative Ricci curvature, Ann. Math., 124 (1986), 1–21. https://doi.org/10.2307/1971385 doi: 10.2307/1971385
![]() |
[12] |
H. Muto, Y. Ohnita, H. Urakawa, Homogeneous minimal hypersurfaces in a unit sphere and the first eigenvalue of the Laplacian, Tohoku Math. J., 36 (1984), 253–267. https://doi.org/10.2748/tmj/1178228851 doi: 10.2748/tmj/1178228851
![]() |
[13] |
B. Solomon, The harmonic analysis of cubic isoparametric minimal hypersurfaces Ⅰ: Dimensions 3 and 6, Am. J. Math., 112 (1990), 157–203. https://doi.org/10.2307/2374713 doi: 10.2307/2374713
![]() |
[14] |
B. Solomon, The harmonic analysis of cubic isoparametric minimal hypersurfaces Ⅱ: Dimensions 12 and 24, Am. J. Math., 112 (1990), 205–241. https://doi.org/10.2307/2374714 doi: 10.2307/2374714
![]() |
[15] |
T. Takahashi, Minimal immersions of Riemannian manifolds, J. Math. Soc. Japan, 18 (1966), 380–385. https://doi.org/10.2969/jmsj/01840380 doi: 10.2969/jmsj/01840380
![]() |
[16] | Z. Tang, W. Yan, Isoparametric foliation and Yau conjecture on the first eigenvalue, J. Differ. Geom., 94 (2013), 521–540. |
[17] | S. T. Yau, Seminar on differential geometry, Princeton: Princeton University Press, 1982. https://doi.org/10.1515/9781400881918 |