A lower bound for the first eigenvalue of a minimal hypersurface in the sphere
Abstract.
Let be a closed embedded minimal hypersurface in the unit sphere and let be the norm of its second fundamental form. In this work we prove that the first eigenvalue of the Laplacian of satisfies
and , when . In particular, this estimate improves the one obtained recently in [5]. The proof of our main result is based on a Rayleigh quotient estimate for a harmonic extension of an eigenfunction of the Laplacian of in the spirit of [3].
Key words and phrases:
minimal hypersurfaces in , first eigenvalue of the laplacian, Yau’s conjecture2020 Mathematics Subject Classification:
MSC 53C42 and MSC 53C441. Introduction.
Let denote the unit sphere in . and a closed embedded hypersurface within . The eigenvalues of the Laplacian operator on form a discrete set of non-negative real numbers. We denote by the first nonzero eigenvalue. It is well known that is minimal if and only if all coordinate functions in restricted on are eigenfunctions corresponding to eigenvalues . This implies that .
On the other hand, it is interesting and important to find the sharp lower bound of for minimal hypersurfaces in . Yau [11] conjectured that . Choi and Wang [3] proved that . This estimate was later refined in [1] by Barros-Bessa who gave the lower bound
(1.1) |
Many progress has been made towards proving Yau’s conjecture after Choi-Wang’s paper (see for instance [1], [5], [9], [4], and [12]). Despite an extensive literature relating to the study of under additional assumptions on , (1.1) has remained the strongest explicit lower bound that is known to hold for a general embedded minimal hypersurface in . The new estimate that we obtain in this work depends on the geometry of as we explain in detail next.
Given , let where are the principal curvatures of in with respect to . We call the norm of the second fundamental form and we define . It is known (see [8]) that in the case , then . Therefore our contribution concerns only the case .
Precisely, we prove
Theorem 1.1.
Let be a closed embedded minimal hypersurface in the unit sphere and let be the norm of its second fundamental form. Then, the first eigenvalue of the Laplacian of satisfies
(1.2) |
and , when .
Remark 1.2.
A recent improvement to (1.1) is given by Duncan-Sire-Spruck in [5], where they proved that
for specific functions and (see (3.28)). By simple comparison of the order of growth it is easy to see that our estimate is bigger when is big enough. Indeed a computation at the end of this paper shows that it is bigger for every and .
Remark 1.3.
Our estimate depends on the norm of second fundamental form and is not sharp. It would be interesting to find a lower bound which is bigger than and depends only on .
We can combine (1.2) with the Yang-Yau inequality [10] is an area bound for embedded minimal surfaces in in terms of their genus. This plays a crucial role in the compactness theory of Choi-Schoen in [2]. They find a constant which is an upper bound for the norm of the second fundamental form of any compact minimal embedded minimal surface in with Euler number .
Corollary 1.4.
Let be the constant in Choi-Schoen’s theorem. Then the first nonzero eigenvalue of the Laplacian of a compact embedded minimal surface in with Euler number satisfies
The rest of the paper is divided into two sections.
In Section 2, we recall the Reilly’s formula and reformulate a result in [1]. We also give a lower bound for in terms of the Rayleigh quotient of its harmonic extension. Namely
where
and is a solution to
where the bars in the expressions above refer to operations in the ambient sphere. Here, is an eigenfunction of the Laplacian associated to the eigenvalue and is a component of which is chosen appropriately. We call the harmonic extension of to . Then we prove Theorem 1.1 assuming the validity of an appropriate estimate for .
In Section 3 we make a quick review of the normal exponential map and we prove the estimate
(1.3) |
which may be considered as an inverse Poincaré type inequality, i.e. is an upper estimate for . We finish the paper by comparing our estimate with the one in [5].
It is important to note that, as in [5], we also use Reilly’s formula and an upper bound of the mean curvature of the parallel surfaces to . However, we provide an upper bound of that depends only on the geometry of , via an elementary result on harmonic extensions.
We need to point out that our techniques can be improved and generalized in order to obtain an estimate for the first eigenvalue of a minimal surface embedded in an ambient of bounded sectional curvature. However, the sharp bound can only be achieved by proving that . Further work will be part of the PhD thesis of the second author. In fact, this work was in preparation before we had access to [5].
2. A first eigenvalue estimate via Rayleigh quotient
In this section we will review Reilly’s formula and give a lower bound for the nonzero first eigenvalue in terms of the Rayleigh quotient of the harmonic extension of the corresponding eigenfunction. As a consequence, we will prove our main result, Theorem 1.1, assuming the inequality (1.3) which will be explicitly stated in Theorem 3.10.
From now on, will denote a closed embedded hypersurface of . It follows that divides the sphere into two components and , where (see [3]). Set as the unit normal of pointing outward to ( as the unit normal of and pointing outward to ) and the second fundamental form of with respect to .
Let . We can assume, without loss of generality, it satisfies the property
and denote . Otherwise we can choose Let us denote all the functions of class that extend the function over as . The following equation is known as Reilly’s formula (see [7]).
Lemma 2.1.
For all we have
(2.1) |
where and denote the Laplacian, gradient and Hessian of in , while and denote the Laplacian and the gradient of in with respect to the induced metric of . On the other hand, denotes the outward normal derivative of in , is the Ricci tensor of and is the mean curvature of .
The following corollary follows from (2.1) by using that .
Corollary 2.2.
Let be a closed embedded minimal hypersurface in the unit sphere . For any we assume is chosen so that . Then
(2.2) |
The following result is proven by Barros and Bessa in [1]. We include here a simpler proof.
Lemma 2.3.
Let be a closed embedded minimal hypersurface in the unit sphere . Assume that is a eigenfunction of on corresponding to the eigenvalue and that is its harmonic extension to , i.e.
(2.3) |
Then for all ,
(2.4) |
where is defined by
(2.5) |
Proof.
Next we obtain a first estimate for which is a corollary of Barros-Bessa’s theorem.
Theorem 2.4.
Let be a closed embedded minimal hypersurface in the unit sphere . Then
(2.11) |
Proof.
First note that, choosing any in (2.4) such that , then we trivially have . On the other hand, by choosing in (2.4) we have that
Again from (2.4) we have
Then
(2.12) |
where . Note that . We have that at the points where , if and only if
Note that, in particular, has no critical points in the interval . It follows that
(2.13) |
The above equality is valid from the fact that , i.e. .
∎
Proof of Theorem 1.1.
It is well known that when , is either a great sphere or a Clifford torus and so . Therefore, we can now assume that . Let be the eigenfunction corresponding to the first nonzero eigenvalue and its harmonic extension to as before. It follows from Theorem 3.10 that Therefore we have from (2.11) that
(2.14) |
On the other hand, since for all ,
we can consider in (2.14) and deduce from Theorem 3.10 that
(2.15) |
The proof is then complete.
3. Gradient estimate via an inverse Poincaré - type inequality.
The aim of this section is to prove an inverse Poincaré-type inequality in Theorem 3.10. To do that we recall first some preliminary results concerning the normal exponential map.
In what follows will denote the normal bundle of , will denote the normal unit bundle of and defined by will denote the normal exponential map on . Such map is well defined in since is embedded with compact closure on the sphere (see for instance [6] ). Let denote the Jacobian determinant of the normal exponential map . On the other hand, let be defined by and
From Proposition 2.1 in [5], if we define and
(3.1) |
we have
Similarly, we have
Defining , we have that
The following lemma correspond to Lemma 10.9 in [6].
Lemma 3.1.
For all
where are the principal curvatures of in with respect to . In addition
(3.2) |
The proof of the following Lemma is straightforward from Lemma 8.1 in [6] and the definition of .
Lemma 3.2.
For all , the application defined by is a diffeomorphism. In particular for any continuous function over we have
(3.3) |
where .
Lemma 3.3.
For each and we have that
where .
Proof.
Given , we define . The function is called the normal extension of and it is well defined in the set
Lemma 3.4.
For each and we have
where is the normal extension of into , denotes the gradient of in and .
Proof.
For , let be an orthonormal basis of where is an orthonormal basis of such that and the parallel transport of along the geodesic from to . It follows that
(3.4) |
Next we are going to construct a transition function wich will be a key technical tool in the proof of our main result. For any we define
(3.6) |
where the function is defined by
It follows that
-
(1)
-
(2)
-
(3)
and .
-
(4)
The function is such that , for all and

In particular, if we denote for the special choices for some , and for some , we have that
(3.7) |
Lemma 3.5.
For any and , the function defined by
(3.8) |
satisfies
(3.9) |
Proof.
Let be the function defined by . It follows that and
(3.10) |
For and , consider be as in the proof of Lemma 3.4. Then
where the last equality is a consequence of (3.4). It follows that
In order to get an upper estimate for , we will use the fact that the harmonic extension minimizes the Dirichlet energy in .
Lemma 3.6.
Let be the harmonic extension of . For all we have
Proof.
Since is harmonic, by Stokes’ theorem
And so
and so we are done.
∎
Proposition 3.7.
Let be the harmonic extension of . Then, if is a first eigenfunction of the laplacian satisfying , we have
(3.13) |
where .
Proof.
Let , and let be the function defined by (3.8).
where the last inequality is a consequence of the condition and (3.2).
On the other hand, note that . Then it follows that
where in the last equality we have used that .
Then (3.15) stands as
Making , choosing and from the definition of in (3.1), it follows that
(3.16) |
∎
Lemma 3.8.
For all and we have
where is the signed distance to in , i.e.
and the mean curvature of the hypersurface .
Proof.
Making the change of variable ,
Here we have used that (see Lemma 10.9 in [6]), where denotes the derivative of with respect to the first variable.
∎
The following result is a consequence of Lemma 3.5 in [5].
Lemma 3.9.
Let . Then for ,
Proof.
Let and . From Lemma 3.5 in [5] it follows that
On the other hand, since and we have that
We conclude that . ∎
Theorem 3.10.
Let be a closed embedded minimal hypersurface in the unit sphere and let be the norm of its second fundamental form. Assume that and is a first eigenfunction of the laplacian satisfying . Then the harmonic extension of satisfies
Proof.
For , let and . It follows that . Moreover, from the coarea formula
(analogously ) and so .
From Lemma 3.8 we have
where in the last two lines we have used Lemma 3.9 and Stokes’ formula respectively. Therefore, by Proposition 3.7
(3.17) |
Multiplying by both sides of (3.17) and integrating from to we have
(3.18) |
Now we can integrate from to in (3.18) and deduce that
(3.19) |
Considering that in (3.19) it follows that
(3.20) |
(3.22) |
From (3.22), it follows that for any ,
In particular, for we have
Finally, this last inequality joint with Proposition 3.7 lead us to
(3.23) |
On the other hand, note that . Then, using the fact that and since for , we have
(3.24) |
We now define
(3.25) |
From (3.24) and (3.25) we have
(3.26) |
Combining (3.23) and (3.26) we have
This completes the proof the theorem.
∎
We conclude the paper by comparing our estimate and the lower bound for obtained by Duncan, Sire and Spruck in [5]. In their work, it is established that given a closed and embedded minimal hypersurface in with , then
(3.27) |
where
(3.28) |
Since when , we have , then from (3.28) we deduce
(3.29) |
Then, since and we trivially obtain from (3.29) that
References
- [1] Abdênago Barros and G. Pacelli Bessa, Estimates of the first eigenvalue of minimal hypersurfaces of , Matemática Contemporânea, 17 (1999), 71–76.
- [2] Hyeong In Choi and Richard Schoen, The space of minimal embeddings of a surface into a threedimensional manifold of positive Ricci curvature, Invent. Math., 81 (1985), pp. 387–394.
- [3] Hyeong In Choi and Ai Nung Wang, A first eigenvalue estimate for minimal hypersurfaces, Journal of Differential Geometry, 18 (1983), no. 3, 559–562.
- [4] Jaigyoung Choe and Marc Soret, First eigenvalue of symmetric minimal surfaces in , Indiana University Mathematics Journal, 58 (2009), 269–281.
- [5] Jonah A.J. Duncan, Yannick Sire, and Joel Spruck, An improved eigenvalue estimate for embedded minimal hypersurfaces in the sphere, arXiv:2308.12235 (2023).
- [6] Alfred Gray, Tubes, vol. 221, Springer Science & Business Media, 2003.
- [7] Robert C. Reilly, Applications of the hessian operator in a riemannian manifold, Indiana University Mathematics Journal, 26 (1977), no. 3, 459–472.
- [8] James Simons, Minimal varieties in riemannian manifolds, Annals of Mathematics, 88 (1968), 62–105.
- [9] Zizhou Tang, Yuquan Xie, and Wenjiao Yan, Isoparametric foliation and yau conjecture on the first eigenvalue, II, Journal of Functional Analysis, 266 (2014), no. 10, 6174–6199.
- [10] Paul C. Yang and Shing-Tung Yau, Eigenvalues of the Laplacian of compact Riemann surfaces and minimal submanifolds, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 7 (1980), pp. 55–63.
- [11] Shing-Tung Yau, Seminar on Differential Geometry, Annals of Mathematics Studies, No. 102, Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1982.
- [12] Yuhang Zhao, First eigenvalue of embedded minimal surfaces in , arXiv:2304.06524 (2023).