1. Introduction
The existence of harmonic coordinates is well-known both on Riemannian and Lorentzian manifolds. Actually, apart from isothermal coordinates on a surface, the problem of existence of harmonic coordinates on a Lorentzian manifold (
wave harmonic) was considered before the Riemannian case. Indeed, A. Einstein, T. De Donder, and C. Lanczos considered harmonic coordinates in the study of the Cauchy problem for the Einstein field equations, cf. [
1]. In a Riemannian manifold with a smooth metric, a proof of the existence of harmonic coordinates is given in [
2] (Lemma 1.2) but actually it can be found in the work of other authors such as [
3] or [
4] (p. 231). The motivation to consider harmonic coordinates comes from the fact that the expression of the Ricci tensor in such coordinates simplifies highly (see the introduction in [
2]).
Different from the Riemannian case, the Finsler Laplacian is a quasi-linear operator and, although it is uniformly elliptic with smooth coefficients where
, the lack of definition of the coefficients on the set where
makes the analogous Finslerian problem not completely similar to the Riemannian one. In a recent paper, T. Liimatainen and M. Salo [
5] considered the problem of existence of “harmonic” coordinates for nonlinear degenerate elliptic operators on Riemannian manifold, including the
p-Laplace operator. We will show that this result extends to the Finslerian Laplace operator as well.
Theorem 1. Let be a smooth Finsler manifold of dimension m, endowed with a smooth volume form μ, such that ; (respectively, —M is endowed with an analytic structure, μ is also analytic, and ). Let , then, there exists a neighborhood V of p and a map such that is a ; depending on V (respectively, ; analytic) diffeomorphism; and , for all , where Δ is the nonlinear Laplacian operator associated to F.
Theorem 1 is proved in
Section 3, where we also show (Proposition 3) that harmonic (for the Finsler nonlinear Laplacian) coordinates can be used to prove the Myers–Steenrod theorem about regularity of distance-preserving bijection between Finsler manifolds. In the Riemannian case, this was established first by M. Taylor in [
6].
For a semi-Riemannian metric
h, the expression in local coordinates of the Ricci tensor is given by
(see Lemma 4.1 [
2]) where
and
are the Christoffel symbols of
h. As recalled above, in harmonic coordinates, the higher order terms in this expression simplify to
because
in such coordinates. When
h is Riemannian, by regularity theory for elliptic PDEs system [
7,
8], this observation leads to optimal regularity results for the components of the metric
h once a certain level of regularity of the Ricci tensor is known [
2]. Roughly speaking, the fact that the nonlinear Laplacian is a differential operator on
M while the fundamental tensor and the components of any Finslerian connection are objects defined on
—harmonic coordinates for the nonlinear Finsler Laplacian do not give such information (see Remark 5). Nevertheless, in
Section 4, we will consider these types of problems for Berwald metrics and we will obtain some partial result in this direction.
2. The Nonlinear Finsler Laplacian
Let M be a smooth (i.e., ), oriented manifold of dimension m, and let us denote by and , respectively, the tangent bundle and the slit tangent bundle of M, i.e., . A Finsler metric on M is a non-negative function on such that for any , is a strongly convex Minkowski norm on , i.e.,
for all and , if and only if ;
the bilinear symmetric form on
, depending on
,
is positive definite for all
and it is called the
fundamental tensor of
F.
Let us recall that a function defined in some open subset U of is of class , for and , (respectively, and ), if all its derivative up to order k exist and are continuous in U and its k-th derivatives are Hölder continuous in U with exponent (respectively, if the derivatives of any order exists and are continuous in U; and if it is real analytic in U).
We assume that , where , (respectively, ; , provided that M is endowed with an analytic structure) in the natural charts of associated to an atlas of M—it is easy to prove, by using 2-homogeneity of the function , that is on with Lipschitz derivatives on subsets of of the type , with K compact in M.
Let us consider the function
which is a
co-Finsler metric, i.e., for any
,
is a Minkowski norm on
. It is well-known (see, e.g., [
9] (p. 308)) that
, where
is the Legendre map:
and
is the vertical derivative of
evaluated at
, i.e.,
. Thus,
, and
(respectively,
;
). Its fundamental tensor, obtained as in (
2), will be denoted by
. The components
of
,
, in natural local coordinate of
define a square matrix which is the inverse of the one defined by the components
of
.
Henceforth, we will often omit the dependence on x in , , , , etc. (which is implicitly carried on by vectors or covectors), writing simply , , , , etc.
For a differentiable function , the gradient of f is defined as . Hence, and, wherever , .
Given a smooth volume form
on
M,
locally given as
, the
divergence of a vector field
is defined as the function
such that
, where
is the Lie derivative—in local coordinates this is the function
. The Finslerian Laplacian of a smooth function on
M is then defined as
, thus, in local coordinates it is given by
where
are natural local coordinates of
and the Einstein summation convention has been used. Notice that, wherever
,
is equal to
thus,
is a quasi-linear operator, and when
F is the norm of a Riemannian metric
h and
—it becomes linear and equal to the Laplace–Beltrami operator of
h.
Let
be an open relatively compact subset with smooth a boundary, and let
,
be the Sobolev space of functions defined on
that are of
class on the open subsets
with compact closure contained in
.
can be defined only in terms of the differentiable structure of
M, see [
10] (Section 4.7). Let us also denote by
and
the usual Sobolev spaces on a smooth compact manifold with boundary (see [
10] (Sections 4.4 and 4.5)).
Let
be the Dirichlet functional of
. Let
, then, the critical points
u of
E on
are the weak solutions of
i.e., for all
it holds:
Let
be a coordinate system in
M, with
,
, and
compact. In the coordinates
, (up to the factor
), the equation
corresponds to
, where
is the map whose components are given by
and
is the vector whose components are
. Observe that
(respectively,
;
). Thus, Finslerian harmonic functions are locally
-harmonic in the sense of [
5]. We notice that
satisfies the following properties: There exists
such that
for all
:
for all
and all
:
for all
and
:
in particular, (
6) comes from strong convexity of
on
(i.e., by (
5)) and by a continuity argument when
, for some
.
By the theory of monotone operators or by a minimization argument based on the fact that
E satisfies the Palais–Smale condition (see [
11] (pp. 729–730)) we have that for all
, there exists a minimum of
E on
, which is then a weak solution of (
3).
Now, as in [
11,
12], the following proposition holds:
Proposition 1. Any weak solution of (
3)
belongs to , for some . Remark 1. The Hölder constant α in the above proposition depends on the open relatively compact subset , where u is seen as a local weak solution of , i.e., From the above proposition and classical results for uniformly elliptic operators, we can obtain higher regularity, where
. In fact, if
on an open subset
, then, being
, the equation
is equivalent to
This can be interpreted as a linear elliptic equation:
where
, and
. Thus, when
, the coefficients
and
f are at least
-Hölder continuous, then,
. By a bootstrap argument, we then get the following proposition (see, e.g., [
13] (Appendix J, Theorem 40)):
Proposition 2. Let be an open subset such that on U, then, any weak solution of (
3)
belongs to , for some α depending on U. Moreover, it is (respectively, ) if , (respectively, if M is endowed with an analytic structure, , and σ is also analytic). 3. Harmonic Coordinates in Finsler Manifolds
Let us consider a smooth atlas of the manifold M and a point . Let be a chart of the atlas, with components , such that , , also let us assume that the open ball , for some , is contained in .
Proof of Theorem 1. Let
be weak solutions of the
m Dirichlet problems:
Following [
14] (Section 3.9) and [
5] (Theorem 2.4), we can rescale the above problems by considering
and
, so that the problems are transferred on
with Dirichlet data
. Notice that
satisfies (
4)–(
6) uniformly with respect to
. Hence, there exists a solution
of
Moreover, there exists such that , for all , uniformly with respect to (notice that, since depends only on , and , which is uniformly bounded with respect to to , is independent of as well).
Let us denote
by
and let
. From (
6), we have
Recalling that
solves (
8) and using
as a test function, we obtain
where the last equality is a consequence of being
,
-harmonic. As
is locally Lipschitz and
is a constant vector, also using Hölder’s inequality, we then get
thus,
, i.e.,
. Since there exists also a constant
such that
by [
5] (Lemma A.1)) we get that there exists
, such that
as
, for all
. Therefore, if
, we have
as
, and then
(
) is invertible. Moreover, up to considering a smaller
, we have also that
for all
and all
. Thus, from Proposition 2
is a
(respectively,
; analytic) diffeomorphism on
and for
, we have that
is a
(respectively,
; analytic) diffeomorphism whose components
are harmonic. □
Harmonic coordinates were successfully used by M. Taylor [
6] in a new proof of the Myers–Steenrod theorem about regularity of isometries between Riemannian manifolds (including the case when the metrics are only Hölder continuous). In the Finsler setting, Myers–Steenrod theorem has been obtained with different methods, in [
15,
16] for smooth (and strongly convex) Finsler metrics; and in [
17] for Hölder continuos Finsler metrics. We show here that harmonic coordinates can be used to prove the Myers–Steenrod in the Finsler setting too, provided that the Finsler metrics are smooth enough.
Let and be two oriented Finsler manifolds of the same dimension m, endowed with the volume forms and . Let us assume that and are locally Lipschitz with locally bounded differential, meaning that in the local expressions of and , , , and are Lipschitz functions and their derivatives (which are defined a.e. by Rademacher’s theorem); and and , for each , are functions. Let , , be the (nonsymmetric) distances associated to . Let be a distance-preserving bijection. Clearly, is an isometry of the symmetric distance , thus, in particular, it is a bi-Lipschitz map (i.e., it is Lipschitz with Lipschitz inverse) with respect to the distances ; and it is locally Lipschitz with respect to the distances associated to any Riemannian metric on and . Hence, and its inverse are differentiable a.e. on and , respectively. In the next lemma we deal with the relations existing between the Finsler metrics and , the inverse maps of their Legendre maps and , and their co-Finsler metrics and , in presence of an isometry . For a fixed x in or , let us denote by , , the diffeomorphisms between and , given by .
Lemma 1. Let be a distance preserving bijection between and . Then, for a.e. , we have:
- (a)
, (i.e., );
- (b)
;
- (c)
, (i.e., ).
Proof. It is well-known that any Finsler metric
F on a manifold
M can be computed by using the associated distance
d as
, where
is a smooth curve on
M such that
and
—hence,
immediately follows from this property and the fact that
is a distance preserving map. From
, we get
then, we deduce
. Finally,
follows from
□
Proposition 3. Let and be two oriented Finsler manifolds, where and are locally Lipschitz with locally bounded differential (in the sense specified above) volume forms. Let be a distance preserving bijective map, if (i.e., locally , where is the Jacobian of ) and , are at least on , then is a diffeomorphism.
Proof. Given that
a bi-Lipshitz map, it is enough to prove that
is locally
with locally
inverse. Under the assumptions on
and
,
, we have that (
4)–(
6) hold, then, given an open relatively compact subset
with Lipschitz boundary, we have the existence of a minimum of the functional
on
, for any
—clearly, the same existence of minima holds for
on analogous subsets
. Moreover, the same assumptions on
and
ensure that such minima are locally weak harmonic and, arguing as in [
12] (Theorem 4.6 and Theorem 4.9), they are
and
on subsets
(respectively,
). Therefore, Theorem 1 still holds under the assumptions of Proposition 3, and it provides
diffeomorphisms
,
, whose components are harmonic. Now, let
and
be a chart of harmonic coordinates of
centred at
,
. Let us show that, for each
,
is weakly harmonic on
. First, we notice that for any function
,
, because
is Lipschitz. Moreover, from
of Lemma 1 and the change of variable formula for integrals under bi-Lipschitz transformations, we have
Hence, given that is a minimum of on , we deduce that is a minimum of on , and then it is a weakly harmonic function. Therefore, is a diffeomorphism, and and its inverse are both map as well. □
Remark 2. If, for each , is of class on , with ; and is of class , as in Theorem 1 (recall, in particular, (7)), we can deduce that is a diffeomorphism provided that and are related by . 4. Regularity Results for Berwald Metrics
Let us recall the following result from [
2], which gives optimal regularity of a Riemannian metric in harmonic coordinates, in connection with the regularity of the Ricci tensor (the meaning of “optimal” here is illustrated in all its facets in [
2]).
Theorem 2 (Deturck–Kazdan). Let h be a Riemannian metric, and be its Ricci tensor. If, in harmonic coordinates of h, is of class , for , (respectively, and ), then, in these coordinates, h is of class (respectively, and ).
Let us consider now a Finsler manifold
such that
F is
on
. A role similar to the one of the Ricci tensor in the result above will be played by the
Riemann curvature of
F. This is a family of linear transformations of the tangent spaces defined in the following way (see [
18] (p. 97)): let
,
, and
be the spray coefficients of
F:
where
are the components of the inverse of the matrix representing the fundamental tensor
g at the point
.
As above, we will omit the explicit dependence on
x, by writing simply
. The Riemann curvature of
F at
is then the linear map
, given by
. It can be shown (see [
18] (Equations (8.11)–(8.12)) that
where
are the components of the
part of the curvature 2-forms of the Chern connection, which are equal, in natural local coordinate on
, to
where
are the components of the Chern connection and
is the vector field on
defined by
, where
.
Finally, let us introduce the
Finsler Ricci scalar as the contraction of the Riemann curvature
(see [
18] (Equation (6.10))).
We recall that, if F is the norm of a Riemannian metric h (i.e., ), then, the components of the Chern connection coincide with those of the Levi–Civita connection, so they do not depend on y but only on , and the functions are then equal to the components of the standard Riemannian curvature tensor of h.
Let us also recall (see [
18] (p. 85)) that, to any Finsler manifold
, we can associate a canonical
covariant derivative of a vector field
V on
M in the direction
, defined in local coordinates as
and extending it as 0 if
.
There are several equivalent ways to introduce Berwald metrics (see e.g., [
19]), we will say that a Finsler metric is said
Berwald if the nonlinear connection
is actually a linear connection on
M—thus,
(see [
20] (prop. 10.2.1)), so that the components of the Chern connection do not depend on
y. From (
9), the same holds for the components of the Riemannian curvature tensor
.
From a result by Z. Szabó [
21], we know that there exists a Riemannian metric
h such that its Levi–Civita connection is equal to the Chern connection of
. Actually, such a Riemannian metric is not unique, and different ways to construct one do exist—in particular, a branch of these methods is based on averaging over the indicatrixes (or, equivalently, on the unit balls) of the Finsler metric
,
(see the nice review [
22]). Moreover, the fundamental tensor of
F can be used in this averaging procedure as shown first by C. Vincze [
23]; then, in a slight different way, in [
24] (based on [
25]); and in other manners also in [
22]. In particular, as described in [
22], the Riemannian metric obtained in [
25] is given, up to a constant conformal factor in the Berwald case, by
where
denotes the measure induced on
(seen as an hypersurface on
) by the Lebesgue measure on
.
Proposition 4. Let be a Berwald manifold such that F is a function on . Assume that the Finsler Ricci scalar R of F is of class , , (respectively, and ) on , for some open set . Then, the Riemannian metric h in (10) is of class (respectively, and ) in a system of harmonic coordinates , , of the same metric. Proof. Given that
F is of class
on
, the partial derivatives of
, up to the second order, exist on
and are continuous—for each
,
, thus, 1 is a regular value of the function
, and the indicatrix bundle
is a
embedded hypersurface in
. Thus, both the area of
and the numerator in (
10) are
in
x, then,
h is a
Riemannian metric on
M. From (
9), and the fact that
F is Berwald, the components
are equal to the ones of the Riemannian curvature tensor of
h, then we have
Moreover,
R is quadratic in the
variables, i.e.,
and its second vertical derivatives
, being independent of
, are
(respectively,
and
) functions on
W. Thus, the result follows from (
11) and Theorem 2. □
Remark 3. Clearly, an analogous result holds for any Riemannian metric, such that its Levi-Civita connection is equal to the canonical connection of the Berwald metric as the Binet–Legendre metric in [26]. Remark 4. Under the assumptions of Proposition 4, we get that components of the Chern connection of the Berwald metric F are (respectively, and ) in harmonic coordinates of the metric h. In particular the geodesic vector field is (respectively, and ) in the corresponding natural coordinate system of .
Remark 5. Other notions of Finslerian Laplacian, which take into account the geometry of the tangent bundle more than the nonlinear Finsler Laplacian, could be considered in trying to obtain some regularity results for the fundamental tensor without averaging. Natural candidates are the horizontal Laplacians studied in [27] which, in the Berwald case, are equal (up to a minus sign) towhere f is a smooth function defined on some open subset of and . We notice that this is also the definition of the horizontal Laplacian of any Finsler metric given [28]. Moreover, for a function , is equal to the g-trace of the Finslerian Hessian of f, , where ∇ is the Chern connection. In natural local coordinates of , for each , and all , is given by (see e.g., [29])—hence, the g-trace of is equal to , where . This expression is equivalent tobecause for a function defined on M. For a function f on , taking into account that and , we havewhich coincides with (
12)
when f is a function defined on M. In particular, if f is a -harmonic coordinate on M, then,which is analogous to (1). Anyway, ellipticity (and quasi-diagonality)—that hold in the Riemannian case for the system , where are (respectively, and ) functions—would be spoiled, in -harmonic coordinates of a Berwald metric, by the presence of second order terms of the type .