1. Introduction
A fundamental question in the study of PDEs is the existence of solutions. It is well understood for linear PDEs; for nonlinear equations there have been numerous studies of specific equations. In this paper we prove the existence of local solutions of general second order quasi-linear elliptic systems. This result can be applied to many well known geometric equations and be used to define a new Kobayashi metric on Riemannian manifolds.
Let be a vector function from an open set in to . We use to denote an elliptic quasi-linear operator on , and write
|
|
|
where
|
|
|
(), and there is a such that
|
|
|
for all .
We will prove that any quasi-linear system defined by such an operator is always locally solvable, given the value and the first derivatives of the solution at any one point in . Without loss of generality we can assume the domain of is a ball centered at the origin.
Denote a ball centered at the origin with radius in as
|
|
|
and denote a ball centered at the origin in as
|
|
|
The following is our main result.
Theorem 1.1.
Let be of .
For any given and , the elliptic quasi-linear system
(1) |
|
|
|
has solutions from to when is sufficiently small.
As will be shown in the proof, there are in fact infinitely many such solutions. An interesting feature of our theorem is that the only assumptions about the coefficient functions are ellipticity and sufficient regularity, therefore it is as general as we can hope for. It is intriguing to know if such a general existence theorem holds for fully nonlinear elliptic systems. The equation , although being non-elliptic, seems to suggest that general local existence may not be true in the fully nonlinear case.
Theorem 1.1 has a wide range of applications in differential geometry, since many of the well known geometric equations are semilinear or quasilnear.
Example 1.
(The Minimal Surface Equation) For a domain and a smooth function on , the area of its graph where , is Its Euler-Lagrange equation is the minimal surface equation
|
|
|
which is an elliptic quasilinear equation. The graph of is called a minimal surface, on which the mean curvature is identically 0. Without loss of generality we can assume . By Theorem 1.1, given and , this equation always has local solutions near . Therefore, given any point and any normal vector , there are infinitely many local minimal surfaces through with as the normal vector.
Similar existence also holds for the prescribed mean curvature equation
|
|
|
where is a given function on .
Example 2.
(The Harmonic Map System) Let and be two Riemannian manifolds of dimensions and with metrics and , respectively. A map is called a harmonic map if it is a critical point of the energy functional. Let be a local coordinate system on . In a local coordinate system on , we may write , and it satisfies the quasi-linear equations
|
|
|
where are the Christoffel symbols on .
By Theorem 1.1, given any , , and a subspace of the tangent space , there exist infinitely many local harmonic maps which map to and to . The theorem also implies the existence of any local objects to be defined by the Laplace-Beltrami operator.
A novel application of this existence of harmonic maps is a definition of Kobayashi metric on Riemmannian manifolds. The Kobayashi metric was first introduced as a pseudometric on complex manifolds by Kobayashi [3]. Let be the unit disc in the complex plane, and let be a complex manifold. We denote the set of all holomorphic functions from to as . For any and , the infinitesimal Kobayashi metric of at is defined by
|
|
|
|
|
|
|
|
|
|
It can be proved that is upper semicontinuous on , and therefore it is a Finsler metric on .
The Kobayashi metric is a well studied, important tool in several complex variables and complex geometry, so a natural question is whether a similar concept can be introduced on Riemannian manifolds. Theorem 1.1 paves the way for answering this question by establishing a well-defined, real version of Kobayashi metric as follows.
Again we use to denote the unit disk in , and let be a disk of radius in centered at the origin with coordinates ). Let be a Riemannian manifold with metric . We consider harmonic maps that are conformal at , so
|
|
|
Definition 1.2.
Let ba a Riemannian manifold, and let be a non-zero tangent vector at a point . The Kobayashi metric of is defined by
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
If , we define .
For any non-zero vector , we can find another vector satisfying and . Then by Theorem 1.1, there is a harmonic map on a local disk , such that , , and . This function is conformal at by the choice of and . Therefore in (1.2) the infimum of is always a finite number, and consequently the Kobayashi metric is well defined on Riemannian manifolds:
Theorem 1.3.
Let be any Riemannian manifold. The Kobayashi metric in (1.2) is well defined. Namely, for all .
We would like to point out that the condition of conformality at 0 excludes geodesics from the admissible harmonic maps. For any geodesic on such that and , the function defines a harmonic map from a neighborhood of the origin to , with and . However, since too, is not conformal at .
There are a lot of questions we can ask about . The first one is whether it is a Finsler metric, which needs to be upper-semicontinuous and positive definite. We hope to explore this in future work.
The rest of the paper is organized as follows. The main strategy for proving Theorem 1.1 is similar to that in [4]. The solution is found by applying the Fixed Point Theorem to an appropriately defined Banach space, but the quasilinear term introduces additional subtlety that needs to be handled carefully. In Section 2 we define a Banach space of functions with vanishing order from which we will seek possible solutions. Then, we study the Newtonian potential as an operator on this Banach space in Section 3. Next, in Section 4 we show that proving Theorem 1.1 is equivalent to proving Lemma 4.1, which is the local existence of solutions of a Poisson type system, and we define a map for the Poisson system. Finally, in Section 5 and Section 6 we prove this map is a contraction if some parameters are appropriately chosen, then the Fixed Point Theorem can be invoked to find a solution, proving Lemma 4.1. Thus, Theorem 1.1 is also true.
2. Defining the Function Spaces and Some Preliminary Estimates
First we define some notations we will be using throughout the proof.
Let be a function defined on .
-
•
The Hölder seminorm of is
|
|
|
-
•
is the set of all functions on such that is finite.
-
•
The weighted Hölder norm of is
(3) |
|
|
|
It is well known (see [2]) that is a norm on with which is a Banach algebra, i.e.
(4) |
|
|
|
for any .
-
•
The Hölder space consists of all functions on whose second order partial derivatives exist and belong to .
-
•
We use to denote the set of all functions in whose first order derivatives all vanish at the origin:
|
|
|
-
•
For any integer , define
(5) |
|
|
|
|
|
|
|
|
|
|
where we have used the notation , , and
|
|
|
Note that
We can also extend the definition to vector functions and define
(6) |
|
|
|
In our proof we will only use (5) and (6) for . A key fact to be used in our proof is that is a norm on , and becomes a Banach space under this norm. We will establish this fact in Lemma 2.4, which is based on the estimates in the following Lemmas 2.1 to 2.3.
Lemma 2.1.
If , then for any ,
|
|
|
Proof: Define
|
|
|
Then
|
|
|
|
|
|
|
|
|
|
After subtracting the second derivatives term, we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
which can be written as
|
|
|
Lemma 2.2.
If , then
|
|
|
Proof: Let , by definition and for all .
For any , define
|
|
|
then and . Therefore,
(7) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
It follows easily from (3) that for all . Then for any with , by (4) we have
|
|
|
Now applying (4) to (7), we obtain
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Lemma 2.3.
If , then for any ,
|
|
|
Proof: Let . If , then .
Similar to the proof of Lemma 2.2, we can show that
|
|
|
|
|
|
|
|
|
|
Therefore
|
|
|
Next, we prove the main result of this section.
Lemma 2.4.
The function space equipped with the norm is a Banach space.
Proof: By definition is a semi-norm on . If , then for all on , which implies is a constant or a linear function. If in addition , then and its first derivatives all vanish at , so must be identically . Thus is a norm on .
Since is a closed subspace of and is a Banach space with the norm, we know is a Banach space with the norm. Then by Lemma 2.3, is also complete under the norm. Therefore, equipped with the norm is a Banach space.
3. A Hölder Estimate for the Newtonian Potential
Recall that the fundamental solution of the Laplace’s equation is given by
|
|
|
For an integrable function on , the Newtonian potential of is defined on by
|
|
|
it solves the Poisson’s Equation . In our proof we will consider as an operator acting on a function space. The following result is well-known and a proof is given in [1].
Lemma 3.1.
Let . For any ,
|
|
|
where is the interior of .
Now we discuss a technical result in order to study the Hölder estimate for functions under the operator . Let be any interior point of , and let be the open ball centered at with radius . It was proved in [1] that
Lemma 3.2.
There is a constant depending only on , such that for any and ,
|
|
|
Although in [1] Lemma 3.2 was proved in the case , the proof actually holds for all without restrictions.
It is well known (see [2]) that maps to continuously. Next we will prove a stronger estimate which is essential to our construction of the contraction map.
Theorem 3.3.
If , then and there is a constant , independent of , such that
|
|
|
We would like to point out that although the constant is independent of , there is an term in the definition of the weighted norm , so the norm actually does depend on . This shows the advantage of choosing the weighted norm over unweighted norm.
Proof: Recall that by (5),
|
|
|
Let
|
|
|
By Lemma 3.1, , and therefore
(8) |
|
|
|
So we only need to bound in terms of .
First, for ,
(9) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the polar coordinates are centered at .
To compute the Hölder constant of , let be two (distinct) points in . Let be the open ball of radius and centered at .
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Next we will estimate each of , , , and .
|
|
|
|
|
|
|
|
|
|
where is a point on the line segment between .
Since ,
|
|
|
then
|
|
|
and
(11) |
|
|
|
Therefore,
(12) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the polar coordinates are centered at and we have used .
By Lemma 3.2, we have
(13) |
|
|
|
|
|
|
|
|
|
|
The next term
(14) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the polar coordinates are also centered at .
The last term
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
For any ,
|
|
|
Then
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the polar coordinates are centered at .
Therefore,
(15) |
|
|
|
Combining (3) to (15), we have
|
|
|
This and (9) now imply
|
|
|
and thus by (8) the proof is completed.
4. The Integral System and the Map
A crucial observation in the proof of Theorem 1.1 is that we only need to prove it in the case
|
|
|
For arbitrary and , we may choose small enough so that for all . We first solve the elliptic quasilinear system
|
|
|
for . Then,
|
|
|
will be a solution to (1).
Thus in the rest of the proof we will assume
|
|
|
Consider the elliptic system (1), for any , we can write
|
|
|
Therefore, (1) can be written in vector form as
(16) |
|
|
|
|
|
|
|
|
|
|
Since the constant matrix is positive definite, after a change of the coordinates , where is an invertible matrix, we can write
|
|
|
The right-hand side of (16) becomes
|
|
|
|
|
|
|
|
|
|
where denotes differentiation with respect to the variable.
After swapping the indices with , it can be expressed as
|
|
|
|
|
|
|
|
|
|
Let
|
|
|
and let
|
|
|
Then we can denote the right-hand side of (16) as
|
|
|
where
Thus in the coordinates (16) becomes
|
|
|
Since the two coordinate systems have the same origin, in the coordinates we still have
|
|
|
and thus
|
|
|
Therefore, we only need to prove Theorem 1.1 under the coordinate system , but to simplify notations we will drop the in all the subsequent notations.
Our goal now is to prove the following result about a Poisson type system.
Lemma 4.1.
Assume the functions and are in . If then the quasi-linear system
(17) |
|
|
|
has solutions from to when is sufficiently small.
Denote
(18) |
|
|
|
so the equation in (17) can be written as
(19) |
|
|
|
A key observation is that if we can find a function satisfying
(20) |
|
|
|
where and each is a harmonic function, then will be a solution of (19). Furthermore, because is arbitrary, it allows us to construct infinitely many such solutions.
Written as a system of equations, (20) is
|
|
|
Since our strategy is to construct a contraction map on functions in and apply the Banach Fixed Point Theorem, it seems natural to define the map as the integral on the right hand side of (20). However, the function may not remain in under such a map, so we need to modify it slightly.
For any function , define
|
|
|
|
|
|
|
|
|
|
and
(22) |
|
|
|
Since is , by Theorem 3.3 we know is and hence is
Direct calculations of the first and second derivatives show that
|
|
|
and
|
|
|
Thus, if a function satisfies
|
|
|
for all then it will be a solution to (19). Therefore, we define the main operator
|
|
|
by
|
|
|
We will find a solution to the system (19) by applying the Banach Fixed Point Theorem to the map on a closed subset of , defined by
|
|
|
where is a parameter to be determined.
In order to show that maps into itself, we need to estimate . To show is a contraction, we need to estimate . The essential step in the estimates is to choose an appropriate value for and then adjust accordingly.
In the rest of the paper, we will use to denote all constants that depend only on and .
5. The Estimate for
First, we show that maps into itself when is appropriately chosen. Specifically, we will show there is a such that for all in with , we have . Then, by the definition of the norm of vector functions in (6) we will have
From (22) we have
(23) |
|
|
|
It follows from (4) and Theorem 3.3 that
(24) |
|
|
|
Also by Theorem 3.3, we know
(25) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Therefore we only need to estimate
We use the coordinates and to denote the components in .
Then
(26) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where
(27) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
By Lemma 2.2, Lemma 2.3, and the fact that ,
(28) |
|
|
|
|
|
|
|
|
|
|
|
|
Therefore, by (26), (4), and the fact , we have
(29) |
|
|
|
|
|
|
|
|
|
|
Next, we will estimate , , , and .
By (28) we only need to estimate on the domain
(30) |
|
|
|
Denote
(31) |
|
|
|
|
|
|
|
|
(32) |
|
|
|
|
|
|
|
|
(33) |
|
|
|
|
|
|
|
|
(34) |
|
|
|
|
|
|
|
|
We use to denote the Lipschitz constant in the variable and define
(35) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Note that by the definition of in (18), there is no variable in , and therefore
(36) |
|
|
|
5.1. Estimates for , , , and
By (27) and (31),
|
|
|
Now we will estimate . Let ,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
It then follows that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Note that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Similarly,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
In addition,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Therefore
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
which implies
|
|
|
|
|
Thus,
|
|
|
|
|
|
|
|
|
|
Denote
|
|
|
|
|
|
|
|
|
|
then we can write
(38) |
|
|
|
(39) |
|
|
|
where
|
|
|
|
|
|
|
|
|
|
and
(41) |
|
|
|
where
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
and we also have
(43) |
|
|
|
where
|
|
|
|
|
|
|
|
|
|
5.2. The Estimates for
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
By the expression of in (5.1), it can be written as
|
|
|
|
|
|
|
|
|
|
Denote
|
|
|
|
|
|
|
|
|
|
Then
(46) |
|
|
|
and
(47) |
|
|
|
Next, we will derive an upper bound for .
By (18)
|
|
|
So as defined in (33),
|
|
|
Since for any as defined in (30),
|
|
|
|
|
|
|
|
|
|
Therefore
|
|
|
Thus we know
(48) |
|
|
|
Now, we choose the parameter such that the term in (46) satisfies
(49) |
|
|
|
Then, by (47) and (48) we can choose sufficiently small such that the sum of the remaining terms in (46) is also less than . Therefore, we have proved that if then for sufficiently small ,
|
|
|
and this proves maps into itself.
6. The Estimates for
Now it remains to show that is a contraction on . The estimates are similar to those for , so we will only point out the main steps without repeating all the calculations.
For any , from (22) we have
(50) |
|
|
|
(51) |
|
|
|
|
|
|
|
|
|
|
and
(52) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Therefore we only need to estimate
|
|
|
Note that
(53) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where , , and are as defined in (27).
|
|
|
|
|
(54) |
|
|
|
|
|
|
|
|
|
|
Then by (53), (4), (6), (38), (39), and (41) we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Note that by (5.1) and (48),
|
|
|
Therefore
|
|
|
Then by (50)-(52), when is sufficiently small, we have
|
|
|
which implies is a contraction.
Finally, by the Fixed Point Theorem there is a function such that
|
|
|
Then
|
|
|
with and .
Thus Lemma 4.1 is proved, and this completes the proof of Theorem 1.1.