On the Lower Bound of the Principal Eigenvalue of a Nonlinear Operator
Abstract.
We prove sharp lower bound estimates for the first nonzero eigenvalue of the non-linear elliptic diffusion operator on a smooth metric measure space, without boundary or with a convex boundary and Neumann boundary condition, satisfying for . Our results extends the work of Koerber[12] for case and Naber-Valtorta[10] for the -Laplacian.
Key words and phrases:
Eigenvalue estimates, generalized -Laplacian, Bakry-Émery curvature dimension, gradient comparison, and half-linear differential equations2010 Mathematics Subject Classification:
35P15, 35P301. Introduction
Let be a compact manifold. The Laplacian operator on plays a key role in studying the geometry of , and one of the key quantity related to the Laplacian is its first nonzero eigenvalue , also called the principal eigenvalue. There have been a lot of works on the estimate of for the Neumann boundary value problem:
In [11] Payne and Weinberger showed for the Laplacian on the convex subset of with diameter . Later in [4] Cheeger gave a lower bound of in terms of the isoperimetric constant on compact Riemannian manifolds. In [7], given that has nonnegative Ricci curvature, P. Li and Yau proved the lower bound by using gradient estimate. Later Zhong and Yang [14] used a barrier argument to prove the sharp lower bound for compact Riemannian manifold with nonnegative Ricci curvature. Afterwards Kroger in [6] used a gradient comparison technique to recover the result of Zhong and Yang, and furthermore he was able to deal with a negative Ricci lower bound case. It was in Bakry and Emery’s work [2] that the situation is generalized into a manifold with weighted volume measure, and the Laplacian is replaced by a general elliptic diffusion operator . By defining curvature-dimension condition, we can make sense of Ricci lower bound in the senario of smooth measure spaces. Later Bakry and Qian [3] used gradient comparison technique similar to Kroger to prove the sharp lower bound for assuming to be for some . Later Andrews and Ni [1] recovered this result with a simple modulus of continuity method.
In recent years there is much attention to the nonlinear operator called Laplacian . In [12] he showed the sharp estimate for Riemannian manifolds with Ricci lower bound and , where is the half period of sine function which will be defined later. The method used is a gradient comparison via Bochner formula for -Laplacian, and a fine ODE analysis of the one dimensional model solution. Later Naber and Valtorta [10] extended the result to the case for . The key improvement is the better understanding of the one dimensional model equation in the case, which is considerably more complicated than non-negative case. Very recently Li-Wang in [8] and [9] used a modulus of continuity method to get the gradient comparison, in the case of drifted Laplacian, which fits in the setting of a Bakry-Emery manifold with weight , thus opening the possibility of studying the non-linear version of operator with drifted terms in metric measure spaces Satisfying . For case, [5] showed that is the sharp lower bound.
In this paper we follow the approaches of [3] and [10] to study the non-linear operator on a compact manifold with possibly convex boundary(to be defined later). We extend the Theorem 1.1 of [5] to the case, more precisely:
Theorem 1.1.
Let be compact and connected and be an elliptic diffusion operator with invariant measure . Assume that satisfies where . Let be diameter defined by the intrinsic distance metric on . Let be an eigenfunction associated with satisfying Neumann boundary condition if , where is the first nonzero eigenvalue of . Then denoting , we have
-
(1)
When , assuming further that , we have a sharp comparison:
where is the first nonzero eigenvalue of the following Neumann eigenvalue problem on :
-
(2)
When , we have a sharp comparison:
where is the first nonzero eigenvalue of the following Neumann eigenvalue problem on :
The rest of this paper is devoted to the proof of Theorem 1.1. The basic structure is the following. In section 2 we introduce the setting and definitions related to the linear elliptic diffusion operator . In section 3 we define the non-linear operator and its Neumann eigenvalue problem. In section 4 we use the Bochner formula to derive a useful estimate (Prop 4.2) which will be used in proving the gradient comparison theorem in section 5. In section 6 we study the associated three one-dimensional model equations and combined with section 7, we get maximum comparison between our model solutions and the solution to the Neumann eigenvalue problem. Finally combining the gradient, maximum and diameter comparison we prove the theorem in section 8.
Acknowledgment. The author would like to thank his advisor Professor Lei Ni for lots of encouragement and helpful suggestions, and Dr. Xiaolong Li for explaining his paper with Kui Wang [8] and [9] to him.
2. The geometry of elliptic diffusion operators
In this section we give some definitions which will be used later in our proof. First we introduce the elliptic diffusion operator, which is a natural generalization of a second order linear differential operator on a Riemannian manifold.
Definition 2.1.
A linear second order operator is called an elliptic diffusion operator if for any we have
and with equality if and only if . Here is defined as
Definition 2.2.
We say that a locally finite Borel measure is L-invariant if there is a generalized function such that
holds for all smooth . is called the outward normal function and is defined to be a set of pairs for a covering of such that and .
Definition 2.3.
We define the intrinsic distance as:
and the diameter of by .
Definition 2.4.
For any , , we define the Hessian by
and the -operator by
Definition 2.5.
We can define the -Ricci curvature as
and let .
Let and , we say that satisfies condition if and only if
for any .
If has a boundary, the geometry of also plays an important role in the eigenvalue estimate. We define the convexity of as follows.
Definition 2.6.
Let be the outward normal direction, and be an open set, , such that and on . We define the second fundamental form on by
If for any as above with on we have
Then we say is convex. If we have strict inequality then is called strictly convex.
3. The generalized -Laplacian and its eigenvalue problem
Now we are going to work on the eigenvalue problem of the non-linear operator derived from the previously defined . The generalized -Laplacian is defined by
We also define
which is the linearization of . Now we define the eigenvalue of . If and satisfies the Neumann boundary problem:
Then we call an eignevalue, and an eigenfunction of , however, we may not always find a classical solution. To define the eigenfunction in a weak sense, we first use the invariance of to deduce the following integration-by-parts formula:
Lemma 3.1.
Let and and on . Then we have
So we define the eigenvalue and eigenfunction by
Definition 3.1.
We say that is an eigenvalue of if there is a such that for any the following identity holds:
We have the following result concerning the regularity of principal eigenfunctions.
Lemma 3.2.
(Lemma 2.2 in [5]) If is a compact smooth Riemannian manifold with an elliptic diffusion operator and an -invariant measure . Then the principal eigenfunction is in for some , and is smooth near points such that and ; for , is , and for , is near where and .
4. Bochner formula
In this section, we will derive the Bochner formula and an estimate which is helpful to prove the gradient estimate in the next section.
Proposition 4.1 (Bochner formula).
Let be a first eigenfunction of , and be a point such that and . Then at we have the following formula:
Proof.
c.f.[5], Lemma 3.1.
Proposition 4.2.
Suppose satisfies for some and . Then for any , we have for ,
for ,
for ,
Proof.
Following [5] Lemma 3.3, we can scale on both sides so that . We can assume since implies for . When , by the curvature-dimension inequality and , we get
When , we have , therefore . Now if , for any , by the curvature-dimension inequality we have
Now we consider a quadratic form , which is non-negative for any . Let where . Then by standard computations, together with the assumption , we have
Then we get
Since for any , we have non-positive discriminant
Therefore we have
5. Gradient Comparison Theorem and Its Applications
In this section we prove the gradient comparison theorem of the eigenfunction with the solution to the one-dimensional model.
Theorem 5.1.
Let be a weak solution of
satisfying Neumann boundary condition if , where is the first nonzero eigenvalue of . Assume that satisfies . Let be a function that satisfies , and be a solution of the following ODE:
(5.1) |
such that is strictly increasing on and the range of is contained the range of . Then for all ,
Proof.
By scaling so that , we can assume that the range of is contained in the range of . By chain rule of what we need to show, equivalently, is
for all . Since depends smoothly on , we will first prove that for any , the gradient comparison holds when is replaced by in (5.1), and then we can take . This will give us a room to use proof by contradiction.
Now for we denote , and consider the function
Assume for contradiction that for some . Let
By our definition of , there is a such that is the maximum of . Now we denote as , as when there is no confusion. When is in the interior of , this clearly implies the following equations:
(5.2) | |||
(5.3) | |||
(5.4) |
If , since by the Neumann boundary condition, we have that at . Since achieves maximum at and is convex, we have
Therefore . This implies that the second derivative of along the normal direction is nonpositive. On the other hand, the second derivatives along tangential directions are nonpositive, hence the ellipticity of implies that . Hence we comfirmed the three equations above for all .
From (5.2) we get
which implies . Now by calculation we have
By chain rule we have , and , and by differentiating the ODE satisfied by we have
Therefore
Now we evaluate the above expression at . Since , and by (1) we have , we have
(5.5) |
By the ODE evaluated at , we have
Hence
Plugging the above equation into the third term of , we have
We can assume that has the same sign as since we can pick sufficiently close to . Now we have to rewrite the second term and finally get
By Proposition 4.1 and 4.2, together with the fact that we have
Hence in both cases, we have , which is a contradiction with the second derivative test. Therefore we conclude that on , which implies our gradient comparison result.
Remark 5.2.
When we know that near , hence the Bochner formula can not be directly applied to . In this case notice that does not vanish identically in a neighborhood of , we can choose with . As we apply the Bochner formula at , The first term since is a eigenfunction. Now this diverging term will cancel with in the expression of , which makes it possible to define to be the limit of as . Therefore the previous proof still works when .
6. One-dimensional Models
In this section we will study the one-dimensional comparison model ODEs and discuss some fine properties of their solutions. Let , be fixed.
6.1. The Model ODE
Let , be fixed, we will consider the following form of initial value problem:
(6.1) |
where is defined over a subset of , to be specified according to the cases or . To study this ODE we first define the - and - functions.
Definition 6.1.
For every , let be defined by:
The periodic function is defined via the integral on by
and we extend it to a periodic function on . Let , and we have the following identity which resembles the case of usual and :
Let us use the Prüfer transformation to study the model equation (6.1).
Definition 6.2 (Prüfer transformation).
Let , then for some solution of the ODE, we define functions and by
Standard calculation shows that and satisfies the following first order systems:
(6.2) | |||
(6.3) |
6.2. Choice of in the case
When , we define functions on , :
-
(1)
-
(2)
-
(3)
and let . Now we let and we get:
-
(1)
, defined on ;
-
(2)
, defined on ;
-
(3)
, defined on .
6.3. Choice of in the case
When , let and . Then is defined as
6.4. Fine analysis of the model equation (6.1)
The central question we need to address here is the existence of solution to (6.1) whose range matches the range of . Due to the normalization that , we will consider the maximum of . For this purpose we introduce some notations. For , let be the solution to the equation (6.1) with , and be the first critical point of after . If for , then we say . Also let and . We shall prove the following statement in the current and next section:
Theorem 6.3.
Under the same setting as Theorem 1.1, let be an eigenfunction of operator, normalized so that and . Then we have the following existence results:
-
(i)
() There is some and a solution such that .
-
(ii)
() There is some , and a solution such that .
To prove Theorem 6.3, first we establish the following existence and uniqueness of a solution to the model equation.
Proposition 6.1.
There is a unique solution to the initial value problem (6.1) with , in the following cases:(1) and and (2) and .
Proof of Proposition 6.1.
In the case for we obtain the existence and uniqueness result from the fact that is a Lipschitz continuous function starting with . Hence we need to confirm the boundary cases only. When (1) and and (2) and for model , we can use fixed-point theorem argument to prove the existence and uniqueness of the solution by slightly modifying the proof in [13], section 3.
Then we shall look at the case . In order to find such that matches the maximum of , we use the continuous dependence of on . We need to show that
Proposition 6.2.
Fix , and . Then there always exists a unique such that the solution is odd, and in particular, the maximum of restricted to is .
The proof of Proposition 6.2 requires certain weaker estimate of . We define the first Neumann eigenvalue of the equation (6.1) on to be
First we claim that
Lemma 6.1.
If , , then equation (6.1) admits a odd solution such that for all .
Proof.
Consider the initial value problem starting with :
(6.4) |
This problem admits a solution up to , the singularity of , which can be extended to an odd solution on . Now we claim that for all . Suppose for some , . This is an eigenfunction corresponding to , hence by the monotonicity of first eigenvalue, we get , which contradicts with our assumption that . Therefore, we have on .
From Lemma 6.1, we can get a weaker bound on :
Lemma 6.2.
When , we have .
Proof.
Suppose that on the contrary, , then by Lemma 6.1, we get an odd function such that on . Suppose the first eigenfuntion is scaled so that . When is bounded, we can scale so that . Pick such that and . Then by the gradient comparison theorem,
which is a contradiction. When is unbounded, we can choose and such that . Again using the gradient comparison theorem, we can prove , which again gives a contradiction. Hence .
Proof of Proposition 6.2.
Using the Prüfer Transformation, we consider the following initial value problem:
(6.5) |
Since , equation 6.5 has a solution such that for some . This implies that achieves maximum at from the equation satisfied by . Therefore we conclude that achieves maximum at , and can be extended to an odd function on such that , i.e. is a solution to the model equation (6.1).
By proposition 6.2, we know that . Now we show the continuous monotonicity of , and first we need the following lemma to confirm continuity at the left endpoint:
Lemma 6.3.
Proof of Lemma 6.3.
The idea of proof is from Proposition 1 of [3]. We will show that for any and , we have
We denote by . We consider the function , and know that the model equation 6.1 can be written as
where . Hence we have
Integrating the above equation over we get
Since , we get
(6.6) |
By another integration over , we have
(6.7) |
We know that as is fixed,
Since and as , the we have and as by equation (6.6) and (6.7).
Proposition 6.3.
is an continuous monotonic function of on .
Proof.
First we show that is an invertible function for . Suppose there are and such that . Then since and have same range and both are invertible functions on and respectively, by the gradient comparison theorem 5.1, we have
and hence , i.e. identical under a translation. However, by the model equation, this can only happen when . Therefore is invertible, and it is monotonic.
To see the continuity of , note that when , the continuous dependence of the solution on the initial value problem is automatic. When , Lemma 6.3 shows that is continuous.
Now let us turn to the case , which is more delicate. We will get a similar result as Proposition 6.2:
Proposition 6.4 ([10] Proposition 6.1).
Fix , and . Then there always exists a unique such that the solution is odd, and in particular, the maximum of restricted to is .
By studying the equation of one can show that there is a critical value at which the oscillatory behavior of changes. For the modeal , we have
Lemma 6.4 ([10] Proposition 6.4).
There exists a limiting value such that for we have for every . For , we have
for sufficiently large we have
When , we have
For model we get the following result:
Lemma 6.5 ([10] Proposition 6.5).
There exist such that when then for all . If then has finite limit at infinity and for all .
Both cases and need to be considered in proving the case (2) of Theorem 6.3. When we can always use model to produce the whole range comparison solutions , i.e. , and when we have restriction on the maximum value that can achieve. More precisely we have:
Lemma 6.6 ([10] Proposition 6.6).
Let . Then for each , there is an such that .
We can also see that model is translation invariant, hence for all , is a constant. For model and we have
Lemma 6.7 ([10] Proposition 6.7).
If , then is a decreasing function of , while is an increasing function of and
Combining the Proposition 6.4 and Lemmas above, we know that in the case , if , there is always a model solution to or such that .
6.5. Diameter Comparison
In order to get the eigenvalue comparison with one-dimensional moder of the same diameter bound, we still need to understand how varies with the diameter. Again we will follow [10].
Definition 6.4.
We define the minimum diameter of the one-dimesional model associated with to be
The following propositions deals with the lower bound of for :
Proposition 6.5.
For and any , we have , where is such that is odd.
Proof.
The proof here is based on the symmetry and convexity of the model . See [10] Proposition 8.4 for the proof.
For the case , we cite the following results from [10]:
Proposition 6.6 ([10], Proposition 8.2).
For and any , we have .
Model 3 needs a little bit careful attention. For this one we notice first that there is always with an odd solution for initial data at . Namely is odd function with min and max . This is a critical situation which minimizes the diameter given :
Proposition 6.7 ([10], Proposition 8.4).
For and , we have
and if , the inequality is strict.
It is also easy to see from the ODE for when that, . Therefore . Also from this we have is strictly decreasing function of , so as to . This means that is a strictly decreasing function. Thus if we see as a function of , we also have the monotonicity: if , we have
7. Maximum of Eigenfunctions
In this section we are going to compare the maximum of the eigenfunctions and the model functions . First, we define a measure on the interval which is essentially the pullback of the volume measure on by . By the ODE satisfied by , is positive before hits its first zero.
First we have a theorem which can be seen as a comparison between the model function and the eigenfunction.
Theorem 7.1.
This result is equivalent to the following statement:
Theorem 7.2.
(Theorem 35,[10]) Under the hypothesis of Theorem 6.1 the function
is increasing on and decreasing on .
To prove the maximum comparison we study the volume of a small ball around the minimum of . By the gradient comparison we have the following:
Lemma 7.1.
For sufficiently small, the set contains a ball of radius .
Now we can prove the maximum comparison, by combining Bishop-Gromov and the following:
Theorem 7.3.
Let and . If is an eigenfunction satisfying and , then there exists a constant such that for all sufficiently small, we have
Proof.
To keep notations short, let . Let be small such that . Then we have when . Let be the first zero of , then by Theorem 6.1 we have . Therefore by Theorem 6.2 we get
Since can be arbitrarily small, we have the claim holds for sufficiently small.
Corollary 7.4.
Let , and is an eigenfunction of with We have the following maximum comparison result:
-
(1)
If , and is the corresponding eigenfunction. Then .
-
(2)
If , and is the corresponding eigenfunction. Then .
Proof.
Let denotes or in either cases, and suppose that . Since is the least possible value among for all model solutions , by continuous dependence of the solution of model equation on , we can find so that is still less that the maximum of the correspoding model equation. Since is still satisfied, we have by Theorem 6.3, that for sufficiently small. However by Bishop-Gromov volume comparison we have . This is a contradiction since .
8. Proof of Theorem 1.1
Now we can combine the gradient and maximum comparison, together with properties of the model equation to show the Theorem 1.1.
Theorem 8.1.
Let be compact and connected and be an elliptic diffusion operator with invariant measure . Assume that satisfies where . Let be diameter defined by the intrinsic distance metric on . Let be an eigenfunction associated with satisfying Neumann boundary condition if , where is the first nonzero eigenvalue of . Then denoting , we have
-
(1)
When , assuming further that , we have a sharp comparison:
where is the first nonzero eigenvalue of the following Neumann eigenvalue problem on :
-
(2)
When , we have a sharp comparison:
where is the first nonzero eigenvalue of the following Neumann eigenvalue problem on :
Proof.
We scale so that and . By Proposition 6.3 we can find a model function such that . By the gradient comparison theorem, . Let and on be points where attains maximum and minimum, then we have
Therefore by the monotonicity of eigenvalue of the model equation, we have that
To check the sharpness of this result when , we have the following examples: let
be a warped product where is the standard unit sphere, and . If we consider being the classical Laplacian on , then standard computation shows that has and geodeiscially convex boundary. Hence it also satisfy the condition. If we take where is the solution to our one-dimensional model equation with . Since the diameter of tends to as , we see that the first eigenvalue on converges to , which shows the sharpness of our lower bound when . For , the round sphere serves as a model for sharp lower bound of .
References
- [1] Ben Andrews and Lei Ni. Eigenvalue comparison on Bakry-Emery manifolds. Comm. Partial Differential Equations, 37(11):2081–2092, 2012.
- [2] D Bakry and Michel Émery. Propaganda for 2. From local times to global geometry, control and physics (Coventry, 1984/85), 150:39–46, 1986.
- [3] Dominique Bakry and Zhongmin Qian. Some new results on eigenvectors via dimension, diameter, and Ricci curvature. Adv. Math., 155(1):98–153, 2000.
- [4] Jeff Cheeger. A lower bound for the smallest eigenvalue of the laplacian. In Proceedings of the Princeton conference in honor of Professor S. Bochner, pages 195–199, 1969.
- [5] Thomas Koerber. Sharp estimates for the principal eigenvalue of the p-operator. Calculus of Variations and Partial Differential Equations, 57(2):49, 2018.
- [6] Pawel Kröger et al. On the spectral gap for compact manifolds. Journal of Differential Geometry, 36(2):315–330, 1992.
- [7] Peter Li and Shing Tung Yau. Estimates of eigenvalues of a compact Riemannian manifold. In Geometry of the Laplace operator (Proc. Sympos. Pure Math., Univ. Hawaii, Honolulu, Hawaii, 1979), Proc. Sympos. Pure Math., XXXVI, pages 205–239. Amer. Math. Soc., Providence, R.I., 1980.
- [8] Xiaolong Li and Kui Wang. Sharp lower bound for the first eigenvalue of the weighted -laplacian. arXiv:1910.02295, 2019.
- [9] Xiaolong Li and Kui Wang. Sharp lower bound for the first eigenvalue of the weighted -laplacian ii. Math. Res. Lett, to appear, arXiv:1911.04596, 2019.
- [10] Aaron Naber and Daniele Valtorta. Sharp estimates on the first eigenvalue of the -Laplacian with negative Ricci lower bound. Math. Z., 277(3-4):867–891, 2014.
- [11] Lawrence E Payne and Hans F Weinberger. An optimal poincaré inequality for convex domains. Archive for Rational Mechanics and Analysis, 5(1):286–292, 1960.
- [12] Daniele Valtorta. Sharp estimate on the first eigenvalue of the -Laplacian. Nonlinear Anal., 75(13):4974–4994, 2012.
- [13] Wolfgang Walter. Sturm-liouville theory for the radial -operator. Mathematische Zeitschrift, 227(1):175–185, 1998.
- [14] Jia Qing Zhong and Hong Cang Yang. On the estimate of the first eigenvalue of a compact Riemannian manifold. Sci. Sinica Ser. A, 27(12):1265–1273, 1984.