Abstract.
We prove the concavity of classical solutions to a wide class of degenerate elliptic differential equations on strictly convex domains of the unit sphere. The proof employs a suitable two-point maximum principle, a technique which originates in works of Korevaar, Kawohl and Kennington for equations on Euclidean domains. We emphasize that no differentiability of the differential operator is needed, but only some monotonicity and concavity properties.
The work of JS is funded by the ”Deutsche Forschungsgemeinschaft” (DFG, German research foundation); Project ”Quermassintegral preserving local curvature flows”; Grant number SCHE 1879/3-1.
1. Introduction
This paper is about classical solutions to fully nonlinear degenerate equations of the form
(1.1) |
|
|
|
on a domain , , of the unit sphere whose closure is geodesically convex. (We recall that a subset of is geodesically convex if no two points of are antipodal and the unique minimizing geodesic segment between any two points of lies entirely in .) Here
(1.2) |
|
|
|
where is the Levi-Civita connection of the round metric on and the -operator is the usual index raising operator with respect to , so that is the gradient vector field and the Hessian endomorphism field on .
We want to prove concavity of solutions to (1.1) satisfying suitable boundary conditions.
The question of concavity of solutions to elliptic equations on Euclidean domains is widely studied, in particular in connection with Laplacian eigenvalues. For example, Brascamp and Lieb [4] proved that the first Dirichlet eigenfunction of the Laplacian over a convex domain is log-concave. A recent highlight in this area was, with many partial results preceding it, the proof of the fundamental gap conjecture by Andrews and Clutterbuck [2].
Using a refined log-concavity estimate for the first eigenfunction, they obtained a sharp estimate for the difference between the first and second Dirichlet eigenvalues (the “fundamental gap”). An analogous result on the sphere was obtained by Seto, Wang and Wei [12].
Our main tool will be the “two-point maximum principle” approach introduced by Korevaar [9]. The idea is to estimate the function
(1.3) |
|
|
|
using a first and second derivative test from below by zero, where is the solution to a certain quasilinear equation with some structure properties. Refinements of this approach for -solutions are due to Kennington [8] and Kawohl [7]. Using regularization, Sakaguchi [11] has transferred this method to solutions of certain -Laplace equations, the solutions of which may be non-smooth.
Similar extensions to viscosity solutions are the content of [1] and [6]. The argument was used to obtain proofs of differential Harnack inequalities for hypersurface flows in [3].
To our knowledge, there is no such concavity maximum principle for non-Euclidean domains. The reason for this seems to arise from a major technical obstruction: in order to make the application of the maximum principle to (1.3) work, it is necessary to differentiate a family of geodesics with respect to their end-points. This leads to the consideration of a family of Jacobi fields and their variations. In order to exploit the first order conditions, one needs detailed knowledge about the nonlinear Jacobi fields along the -minimizing geodesic. Even worse, the second order condition contains a further space derivative of the Jacobi equation and in general it seems impossible to extract the correct sign, which is required to apply the maximum principle, from these conditions.
The purpose of this paper is to initiate the study of concavity maximum principles for degenerate elliptic equations on a Riemannian manifold and we start this journey on the unit sphere. Here we have managed to deal with the pretty well known Jacobi fields and their maybe not so well known derivatives.
One first key observation is that, for technical reasons, the full three-parameter function is no longer suitable. The concavity of a solution is equivalent to the concavity of
(1.4) |
|
|
|
for every geodesic
(1.5) |
|
|
|
cf. [13]. By a well known result due to Jensen [5], for continuous functions this is equivalent to asking this function to be midpoint-concave, i.e.
(1.6) |
|
|
|
Hence it will suffice to prove that the function
(1.7) |
|
|
|
|
|
|
|
|
is non-negative, where is the unique minimizing geodesic from to . The mid-point approach has also been taken for equations on Euclidean domains in [6, 7].
Before we state our main theorem, we need to introduce some more objects. We denote by the category of -dimensional real vector spaces with linear transformations as its objects. A function on that category is called isotropic if it acts on the subclass of endomorphisms in a -invariant way:
(1.8) |
|
|
|
where denotes the class of all invertible linear maps, understood to map to the correct space such that (1.8) makes sense.
From this property it is clear that the action of on real diagonalizable endomorphism is only determined through the action on the ordered set of eigenvalues ,
(1.9) |
|
|
|
The properties we will require from will in general only be satisfied if is restricted to a subclass of and the best way to describe the domain of definition of is via . Therefore we suppose the domain of definition for , , is a symmetric, open and convex cone which contains the positive cone
(1.10) |
|
|
|
Denote by the vector space of symmetric -matrices. Then there exists an open domain of definition for consisting of matrices with eigenvalues in , such that (1.9) holds for all .
Abusing notation, the orthogonal invariance of allows us to understand to be a subset of any other space of -self-adjoint endomorphisms of an arbitrary -dimensional inner product space .
Here is our main theorem:
1.1 Theorem.
Let , and suppose to be geodesically convex. Let be a symmetric, open and convex cone containing and let be a solution to
(1.11) |
|
|
|
Suppose the function has the following properties:
-
(i)
For every , is an isotropic function on
-
(ii)
is increasing in the first variable,
(1.12) |
|
|
|
-
(iii)
increasing in the second variable,
(1.13) |
|
|
|
-
(iv)
convex in the second variable,
(1.14) |
|
|
|
Suppose the function is
-
(a)
decreasing in the third variable,
-
(b)
jointly concave in the first two variables:
(1.15) |
|
|
|
-
(c)
strictly decreasing in the second variable.
If furthermore for all there holds
(1.16) |
|
|
|
then is concave.
1.2 Remark.
The boundary condition is slightly stronger than the ones in [8, 9]. This stems from our technical restriction that due to the nonlinear ambient geometry we have to use the notion of mid-point concavity. Hence we can not vary a boundary point without varying another point too. Geometrically it says that at every point , the totally geodesic hypersurface tangent to at must lie above and at every point the totally geodesic hypersurface tangent to at must lie above , and one of these relations must be strict.
1.3 Example.
Let us give a large class of examples of operators to which this theorem applies. Given a convex and increasing function , define
(1.17) |
|
|
|
Then the operator function
(1.18) |
|
|
|
associated to is convex and increasing. Here is also understood to be the operator function. Explicit examples are
(1.19) |
|
|
|
where the latter is the matrix exponential.
For the proof of Theorem 1.1 we need an easy result from linear algebra and some tedious calculations of Jacobi fields and their derivatives. We address these issues in the next two sections, before we can proceed to the proof of the theorem.
3. Jacobi fields
We have to differentiate (1.7) twice and in this section we collect the geometric ingredients of this process. The derivatives of will be Jacobi fields and variations of them and we will make extensive use of the Jacobi equation.
Define
(3.1) |
|
|
|
|
|
|
|
|
For the moment choose arbitrary coordinates in .
The vector fields
(3.2) |
|
|
|
are, for given , Jacobi fields along with
(3.3) |
|
|
|
where we identify with its pushforward under the identity map and likewise for . We will have to deal with the second derivatives of which are given by (neglecting the distinction between and , since later we will work in Riemannian normal coordinates)
(3.4) |
|
|
|
Those satisfy a differentiated Jacobi equation, which we derive in the following.
We use the following convention for the Riemannian curvature tensor.
(3.5) |
|
|
|
i.e. on we have
(3.6) |
|
|
|
where we recall that denotes the round metric.
With this convention the Jacobi equation, which for example satisfies for fixed , is
(3.7) |
|
|
|
see e.g. [10, Ch. 10]. The same holds for and now we have to differentiate this equation with respect to and .
We consider it to be easier to follow those calculations if we do not explicitly put in the specific form of the Riemann tensor just yet. However we will already use .
For we calculate:
3.1 Lemma.
There holds
(3.8) |
|
|
|
|
and similarly for , and .
Proof.
We only prove this for since the other identities are obtained in a similar manner. Differentiating the Jacobi equation and applying the first Bianchi identity (in the final step) yields
(3.9) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
as claimed.
∎
In order to get more information about these quantities, first we note that at all are zero. This is due to the fact that
(3.10) |
|
|
|
and the second covariant derivative of an isometry is always zero.
Also it is in order to pick the right coordinates. In the following, latin indices range from to and greek indices range from to .
At a given point pick Riemannian normal coordinates, where with a rotation we arrange
(3.11) |
|
|
|
while at we do the same. Since we have Riemannian normal coordinates at and , the vectors form an orthonormal basis at , while is the parallel transport of . Furthermore, the Christoffel symbols vanish at in the product space. After a possible orthogonal transformation at and reordering of basis vectors, there exists a set of vector fields parallel along , such that
(3.12) |
|
|
|
3.2 Lemma.
In the above constructed coordinates there hold:
-
(i)
-
(ii)
-
(iii)
The quantities
(3.13) |
|
|
|
and
(3.14) |
|
|
|
where is any parallel vector field, satisfy
(3.15) |
|
|
|
|
respectively
(3.16) |
|
|
|
|
Proof.
(i) Since and are not conjugate points, the Jacobi fields are uniquely determined by their boundary values. Hence, suppressing ,
(3.17) |
|
|
|
However, itself is the tangent to the normalized geodesic. Plugging this into (3.8) yields that those are Jacobi fields themselves, though with vanishing boundary values. Hence they are identically zero.
(ii) For any define
(3.18) |
|
|
|
Then solves
(3.19) |
|
|
|
|
|
|
|
|
|
|
|
|
where we have used
(3.20) |
|
|
|
Hence is, as a Jacobi field with vanishing boundary values, identically zero.
(iii) We add up all four equations of the form (3.8) with the appropriate sign. We treat and simultaneously by using in the appropriate places. We get
(3.21) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Applying this to gives
(3.22) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
A quick check of both cases reveals that and satisfy the desired equations.
∎
3.3 Lemma.
The quantities and from Lemma 3.2 both vanish at .
Proof.
From (3.7) we obtain
(3.23) |
|
|
|
with a function that satisfies
(3.24) |
|
|
|
Hence, from multiplying (3.15) and (3.16) with some , we see that is a Jacobi field with vanishing boundary values and hence zero.
Thus and are both multiples of and again from (3.15) and (3.16) we get
(3.25) |
|
|
|
|
|
|
|
|
|
|
|
|
In this sum, the contribution of terms involving is zero, because in this case the curvature term is zero. So we get
(3.26) |
|
|
|
|
|
|
|
|
We solve this equation by integration with respect to the boundary values and evaluate at .
In general we get with some constants and ,
(3.27) |
|
|
|
and
(3.28) |
|
|
|
Since , we find
(3.29) |
|
|
|
and hence we have
(3.30) |
|
|
|
Due to the symmetry of the function around we get and the proof is complete.
∎
To prove Theorem 1.1, we need to prove that the two-point function
(4.1) |
|
|
|
defined in is non-negative at minimal points. Therefore we may assume that . If one of the minimizing points, say , is at the boundary, then we prove that by moving and towards each other at the same speed (i.e. while fixing the mid-point ), we could decrease the value of . To see this, consider the function
(4.2) |
|
|
|
There holds
(4.3) |
|
|
|
Hence is strictly negative and must achieve its minimum at interior points.
Let .
The first order conditions on are (suppressing the spatial arguments for )
(4.4) |
|
|
|
and
(4.5) |
|
|
|
with
(4.6) |
|
|
|
and the latter solves (3.24) and hence
(4.7) |
|
|
|
We use the bases
(4.8) |
|
|
|
which are orthonormal at the respective points , to identify all of the three tangent spaces canonically with one single Euclidean . In these coordinates, define an endomorphism through its matrix representation,
(4.9) |
|
|
|
Then
(4.10) |
|
|
|
since
(4.11) |
|
|
|
The second order condition implies that for all and all
(4.12) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where
(4.13) |
|
|
|
We test this relation twice with suitable and to deduce a concavity relation. First let
Then we obtain in the sense of bilinear forms on ,
(4.14) |
|
|
|
since in this case and the term involving also vanishes in view of the case in Lemmata 3.2, LABEL: and 3.3. Due to the same reasoning using , from setting we obtain
(4.15) |
|
|
|
As we are working in Euclidean coordinates, we may view both sides to lie in . Hence we may apply after raising an index with respect to the Euclidean scalar product and use the naturality of to obtain
(4.16) |
|
|
|
|
where we have used the monotonicity of in the second variable and Lemma 2.1.
Using the convexity and the other monotonicity properties, we get
(4.17) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Hence, due to the strict monotonicity in the second variable,
(4.18) |
|
|
|
and the proof is complete.
4.1 Remark.
The method to introduce different variational direction in the and variables was already used in [7, Thm. 3.13], although in a simplified (quasilinear Euclidean) setting, where it gave a lot more information.