1. Introduction
This article has a double aim in Lorentz–Finsler geometry. The first one is to revisit the physical grounds of the stress–energy tensor
T,
Section 3. The possible extensions of the relativistic
T are discussed from the viewpoint of both fluids mechanics and Lagrangian systems. The second one is to revise geometrically the notion of divergence,
Section 4, yielding consequences about the conservation of
T,
Section 5. With this aim, we introduce new notions of the Lie bracket and the derivative associated with a nonlinear connection and applicable to anisotropic tensors fields, which appear naturally in Finsler geometry.
Finslerian modifications of General Relativity aim to find a tensor
T collecting the possible anisotropies in the distribution of energy, momentum, and stress, which will serve as a source for the (now Lorentz–Finsler) geometry of the spacetime [
1,
2,
3,
4,
5]. Some of these proposals may be waiting for experimental evidence, postponing then how the basic relativistic notions would be affected. However, such a discussion is relevant to understand the scope and implications of the introduced Finslerian elements. In a previous reference [
6], the fundamentals of observers in the Finslerian setting were extensively studied, including its compatibility with the Ehlers–Pirani–Schild approach. Now we focus on the stress–energy tensor
T.
The difficulty to study such a
T is apparent. Recall that, using the principle of equivalence, General Relativity is reduced infinitesimally into the Special one, which provides a background for interpretations. However, in the Lorentz–Finsler case, the infinitesimal model is changed into a Lorentz norm (instead of scalar product), implying a breaking of Lorentz invariance. This is a substantial issue in its own right which has been studied in the context of
Very Special Relativity and others [
7,
8,
9,
10,
11]. As an additional difficulty, the infinitesimal model changes with the point
1.
Two noticeable pre-requisites are the following: (a) only the value of the Lorentz–Finsler metric on causal directions is relevant [
6,
14] (this is briefly commented in the setup
Section 2.3), and (b) there is a significant variety of possible extensions of the relativistic kinematic objects to the Finsler case, at least from the geometric viewpoint (see
Appendix A). Taking into account these issues, the extension of the notion of the stress–energy tensor to the Finslerian setting is discussed in
Section 3.
We start at the fluids approach. As a preliminary question, energy-momentum is discussed, in
Section 3.1. We emphasize that, even though this is well-defined as a tangent vector in each tangent space
,
, different observers
u,
at
p will use coordinates related by non-trivial linear transformations. Indeed, the latter will depend on both
L and the chosen method to measure relative velocities. Moreover, when the stress–energy
T is considered in
Section 3.2, the arguments in Classical Mechanics and Relativity, which support its status as a tensor, hold only partially in the Lorentz–Finsler setting. Indeed,
T acquires a nonlinear nature that is codified in an (observer-dependent) anisotropic tensor, rather than in a tensor on
M.
The Lagrangian approach is discussed in
Section 3.3. This approach has been developed recently by Hohmann, Pfeifer, and Voicu [
15,
16], who introduced an
energy-momentum scalar function. Here, we discuss the analogies and differences of this function with the canonical relativistic stress–energy tensor
and the 2-tensor
T obtained from the fluids approach above. Relevant issues are the existence of different methods to obtain a 2-tensor starting at a scalar function, the recovery of this function from a matter Lagrangian, and the possibility to consider the Palatini Lagrangian as the background one (rather than Einstein–Hilbert-type Lagrangians used by the cited authors; recall that Palatini’s becomes especially meaningful in the Finslerian case [
17]). The important case of kinetic gases is considered explicitly (Example 2).
Once the definition of
T has been discussed, we focus on its conservation,
Section 5, revisiting first the divergence theorem,
Section 4. This is crucial in the Finslerian setting because, as discussed before, the Lagrangian approach above does not guarantee a conservation law as the relativistic
.
Section 4 analyzes the divergence from a purely mathematical viewpoint. Now,
L is regarded as pseudo-Finsler (the results will be useful not only in any indefinite signature but also in the classical positive definite case), and
T will not be assumed to be symmetric a priori. Classically, the divergence of a vector field
Z is defined with the derivation associated with the Lie bracket
, applied to the volume element. In the Finslerian case, however, the Lie derivative and bracket do not make sense for arbitrary anisotropic vector fields. This difficulty was circumvented by Rund [
18], who redefined
in such a way that a type of divergence theorem held. However, the Lie viewpoint is restored here.
Section 4.1. Once a nonlinear connection
(seen as a horizontal distribution on
A) is prescribed, we can define a Lie bracket
and, then, a Lie derivative
( Definitions 1 and 2; Theorem 1 (C)). Noticeably, the former
is expressible in terms of the infinitesimal flow of
Z (Proposition 1).
Section 4.2. The divergence of
Z is naturally defined by using this Lie bracket (Definition 3). For the computation of
, however, one can use an
anisotropic connection ∇ (this can be seen as a Finsler connection dropping its vertical part, see
Section 2) and a priori Chern’s one is not especially privileged (Proposition 2).
Section 4.3. We give a general Finslerian version of the divergence theorem for any anisotropic vector field
Z, emphasizing the role of the choice of an (admissible) vector field
, which in the Lorentzian case can be interpreted as an observer field; this is expressed in terms of integration of forms in the spirit of Cartan’s formula (Theorem 2, Remark 5). We also explain how the boundary term can be expressed in different ways by using a normal either with respect to the pseudo-Riemannian metric
or to the fundamental tensor, which were the choices of Rund [
18] and Minguzzi [
19] resp.
Section 5 gives some applications to conservation laws.
Section 5.1. First, we discuss the definition of divergence for the case of
T. Our definition for vector fields was not biased to the Chern anisotropic connection, but this will be used for
(Definition 4). The reason is that
should behave under contraction in a similar way as in the isotropic case (namely, as in Formula (
11)), which privileges Chern’s connection (Proposition 3).
Section 5.2. As an interlude about the appearance of Chern’s ∇, a comparison with the possible use of Berwald’s and previous approaches in the literature is done.
Section 5.3. A conservation law for the flow of
is obtained (Corollary 2), stressing three hypotheses on the vanishing for
V of elements related to the stress–energy
T (
), the anisotropic vector
X (
, generalizing the isotropic case) and a derivative of
V. The latter hypothesis is genuinely Finslerian, and it means that some terms related to the nonlinear covariant derivative
must vanish globally (
V can always be chosen such that they vanish at some point). It is worth pointing out that our general formula for the integral of the divergence (
36) recovers the classical interpretation of the divergence as an infinitesimal growth of the flow (now observer-dependent). So,
is equivalent to the conservation of energy-momentum in the instantaneous rest-space of each observer—see Remark 10.
We finish by applying this general result to two examples. First, we apply it to Lorentz norms, showing that the conservation laws of Special Relativity still hold even though, now, the conserved quantity may be different for different observers. As a second example, we give natural conditions so that the flow of
(whenever it exists as a Lebesgue integral, eventually equal to
) is equal in two Cauchy hypersurfaces of a globally hyperbolic Finsler spacetime. Indeed, we refine a previous result by Minguzzi [
19], who assumed that
L was defined on the whole
and
was compactly supported. We show that a combination of Rund’s and Minguzzi’s methods to compute the boundary terms allows one to obtain appropriate decay rates (namely, the properly Finslerian hypothesis (
49)), which ensure the conservation.
3. Basic Interpretations on the Stress–Energy Tensor
Let us start with a discussion at each event of a Finsler spacetime . We can consider endowed with the Lorentz norm . In most of this section, the discussion relies essentially on the particular case when M is a real affine n-space with associated vector space V (which plays the role of in the general case) and L is a Lorentz–Finsler norm on V with indicatrix and cone included in V. Given , consider the corresponding fundamental tensors and and take orthonormal bases , , obtained extending u, . In a natural way, these bases live in , and they can be identified with bases in V itself. Assuming this, the change in coordinates between , is linear but not a Lorentz transformation, in general.
Extending the interpretations in relativity,
is an
event; the affine simplification includes the case of Very Special Relativity [
7,
8,
10];
can be regarded as an
observer; the tangent space to the indicatrix
(i.e., the subspace
-orthogonal to
u in
) becomes the
rest-space of the observer
u; and
is an
inertial reference frame for this observer. The
Lorentz invariance breaking corresponds to the fact that the bases
and
are orthonormal for the different metrics
, and, thus, the linear transformation between the coordinates of
and
(when regarded as elements of the same vector space
) is not a Lorentz one. If the affine simplification is dropped, such elements (observers and rest-spaces) must be regarded as instantaneous at
.
It is worth emphasizing that, according to the viewpoint introduced in [
14] and discussed extensively in [
6], the space-like directions are not physically relevant for the Lorentz–Finsler metric. However, each (instantaneous) observer does have a restspace with a Euclidean scalar product. In the case of classical relativity, Lorentz-invariance permits natural identifications between these rest-spaces, and they become consistent with the value of the scalar product on space-like directions. Certainly, a Lorentz norm
L could be extended outside these directions (maintaining the Lorentz signature for its fundamental tensor), but this can be done in many different ways, and no relation with the scalar products
would hold.
The dropping of natural identifications associated with the Lorentz invariance implies that many notions that are unambiguously defined in classical relativity admit many different alternatives now. In the
Appendix A, we analyze some of them for the
relative velocity between observers as well as other kinematical concepts. This is taken into account in the following discussion about how the Finslerian setting affects the notion of the
energy–momentum–stress tensor.
3.1. Particles and Dusts: Anisotropic Picture of Isotropic Elements
In principle, there is no reason to modify the classical relativistic interpretation of as the (energy-) momentum vector of a particle of (rest) mass moving in the observer’s direction . Moreover, if the particle moves in such a way that m is constant, it will be represented by a unit time-like curve such that will be its instantaneous momentum at each proper time . The (covariant) derivative would be the force F acting on the particle, which is necessarily -orthogonal to (i.e., the force lies in the instantaneous rest-space of the particle). Then, the relativistic conservation of the momentum in the absence of external forces would retain its natural meaning, namely, if the particle represented by splits into two and at some then .
The
Appendix A suggests that the way how an observer
u may measure the energy-momentum and conservation may be non-trivial. In particular, if one assumes that an observer
u measures
by using a
-orthonormal basis
in general,
. Moreover, as we have already commented, the coordinates for other observer
will not transform by means of Lorentz transformation. However, as the transformation of their coordinates is still linear, and both of them will write consistently
in their coordinates.
Particles are also the basis to model dusts, which constitute the simplest class of relativistic fluids. A dust is represented by a number-flux vector field
, where
U represents the intrinsic velocity of the particle in the dust, i.e., a comoving observer, and
n is the density of the dust for each momentaneously comoving reference frame. Comparing with the case of energy momentum,
N is also an intrinsic object that lives at the tangent space of each point, and
U gives the privileged observer who measures
n. However, the measures of
n by different observers involve different measures of the volume. As explained in the
Appendix A, the length contraction may be fairly unrelated to the relative velocities of the observers. This implies a more complicated transformation of the coordinates by different observers. Anyway, the transformations between these coordinates would remain linear, and, so, they could still agree in the fact that they are measuring the same intrinsic vector field.
Summing up, in the case of both particles and dusts, one assumes that the physical property lives in V (or, more properly, in each tangent space of the affine space), and there is a privileged (comoving) observer u. The transformation of coordinates for another observer may be complicated, but, at the end, it is a linear transformation that can be determined by specifying the geometric quantities that are being measured as well as the geometry of . Thus, by using the coordinates measured by each observer one could construct and anisotropic vector field at each , which will fulfill some constraints, as the measurement by one of the observers (in particular, the privileged one) would determine the measurements by all the others.
3.2. Emergence of an Anisotropic Stress–Energy Tensor
The situation, however, is subtler for more general fluids, which are modeled classically by a 2-tensor on the underlying manifold.
Let us start by recalling the Newtonian and Lorentzian cases. In Classical Mechanics, one starts working in an orthonormal basis of Euclidean space to obtain the components
of the Cauchy stress tensor, which give the flux of
i-momentum (or force) across the
j-surface in the background
2. The laws of conservation of linear momentum and static equilibrium of forces imply that these components give truly a 2-tensor (linear in each variable), and the conservation of linear momentum implies that this tensor is symmetric.
In the relativistic setting, each observer will determine some symmetric components
in its rest-space by essentially the same procedure as above. Additionally, it constructs
,
, and
as the density energy and the energy flux across
i-surface and
i-momentum density, resp. The interpretation of these magnitudes completes the symmetry
3 as well as the linearity in the 0-component. However, the bilinearity in the components
has been only ensured for vectors in the rest-space of the observer. In relativity, one can claim Lorentz invariance in order to complete the reasons justifying that, finally, the components
will transform as a tensor
4.
Nevertheless, it is not clear in Lorentz–Finsler geometry why the transformation of the components from an observer u to a second one must be linear, taking into account that they apply to space-like coordinates in distinct Euclidean subspaces and no Lorentz-invariance is assumed. Indeed, the following simple academic example shows that this is not the case.
Example 1. Assume that is an affine space with a Lorentz norm with domain A and consider the anisotropic tensor , where ℂ is the canonical (Liouville) vector field, and is a smooth function, which is extended as a 0-homogeneous function on A. Then, for each and , one has , , and . In this case, each is a symmetric 2-tensor, but the information on requires the knowledge of for all possible . Recall that this example holds even if is the Lorentz–Minkowski spacetime regarded as a Finsler spacetime (but no Lorentz-invariance is assumed for ).
Therefore, the following issues about T appear:
- (a)
Observer dependence: even if we assume that the components measured by any observer u are bilinear, and, then, it is a standard tensor, the components measured by a second observer may transform by a linear map, which depends on as well as the experimental method of measuring (as in the case of the energy–momentum vector).
- (b)
Nonlinearity: it is not clear even why such a linear transformation must exist, as bilinearity is only ensured in the direction of u and of its rest-space. Thus, the tensor measured by a single observer u would not be enough to grasp the physics of the fluid at each event , as in the example above.
- (c)
Contribution of the anisotropies of : as an additional possibility, the local geometry of at u underlies the measurements of this observer and might provide a contribution for the stress–energy tensor itself.
Summing up, Lorentz–Finsler geometry leads to assume that the measurements by u are not enough to determine the state of the fluid, and the stress-energy tensor should be regarded as a non-isotropic tensor field, determined by the measurements of all the observers.
Formally, this means an
anisotropic tensor (see [
20] for a summary of the formal approach), which can be expressed locally as
where
for all
(i.e.,
depends only on the direction of
v). As a first approach, we can assume
(recall
Appendix A.5). Consistently, we will assume that there exists a Lorentz–Finsler metric
L on
M with indicatrix
, and, so, indexes can be raised and lowered by using its fundamental tensor
g. The fact that
T has order 2 is important to establish classical analogies. However, other tensors might appear as more fundamental energy-momentum tensors, and, then, one would try to derive a semi-classical 2-tensor as in
Section 3.3.
In principle, the intuitive relativistic interpretations would be transplanted directly to each
v, whenever
. That is, given two
-unit vectors
, the value
of the 2-covariant stress–energy tensor perceived by the observer
v (at
) is obtained as the flux of
w-energy-momentum per unit of
-volume orthogonal to
u. More precisely, let
be a small coordinate 3-cube in a hypersurface
-orthogonal to
u, and
is the total flux of the energy-momentum of particles crossing
(being positive from the
side to the
u side and negative the opposite direction), then the
w-energy-momentum per unit of
-volume is
where
. As a Finslerian subtlety, recall that
is only defined in
and then in
(i.e., it is trivially extended to
in a coordinate-depending way), but the above limit depends only on the value of
. Namely, if one considers two semi-Riemannian metrics
g and
in a neighborhood of
p such that
and
are open subsets with
p in the interior of
for all
and
, then
In particular, we have the interpretations (recall signature ):
is the energy density measured by
,
being
the measured energy.
If
w is
-orthogonal to
v and
-unit,
measures the flow of energy per unit of
-volume in a surface
-orthogonal to
v and
w (i.e., some small surface of area
A flowing a lapse
), while
measures the
w-momentum density,
If
are
-orthogonal to
v and
-unit,
measures the flow of
w-momentum per unit of
-volume in a surface
-orthogonal to
v and
z,
3.3. Lagrangian Viewpoint
In the Lagrangian approach for Special Relativity, the background spacetime is assumed to be endowed with a flat metric
. So, the Lagrangian
is constructed by using the prescribed
and some matter fields
. The stress–energy tensor coincides with the
canonical energy–momentum tensor associated with the Lagrangian, in most cases (the exceptions include theories involving spin). This canonical tensor appears as the Noether current associated with the invariance by spacetime translations (i.e., when
), namely
5 In principle, these interpretations would hold unaltered for the case of an affine space with a Lorentz norm, including the case of Very Special Relativity.
In General Relativity, however, the Lagrangian formulation introduces a background Lagrangian independent of matter fields (the Einstein–Hilbert one, eventually with a cosmological constant) and, then, a matter Lagrangian
, which includes a constant of coupling with the background. Then, the safest way to define the stress–energy tensor is the canonical one obtained as the corresponding action term
in the Euler–Lagrange equations
6,
Any tensor obtained in this way will have some advantages to play the role of a stress–energy tensor, because it will be automatically symmetric (in contrast to (
8)) and will have vanishing divergence.
In the Finslerian setting, the variational viewpoint has been systematically studied in a very recent study by Hohmann, Pfeifer, and Voicu [
16]. Previously, the background Lagrangian closest to the Einstein–Hilbert functional in the Finslerian setting had been studied in [
15,
29]. Such a functional is obtained as the integral of the Ricci scalar function on the indicatrix of the Lorentz–Finsler metric
7 L. Taking into account this background functional, they define the energy-momentum scalar function by taking the corresponding variational action term (Formula (84) in [
16]),
Notice that, here, the functional coordinate for the Lagrangian is
L, and, thus, an (anisotropic) function rather than a 2-tensor is obtained. However, starting at this function some tensors become useful (Formulas (88) and (91) in [
16]), in particular a canonically associated (anisotropic Liouville) 2-tensor
as in Example 1. Notice that, essentially, the information of these tensors is codified in
. Even though such a tensor is justified by the procedure of Gotay–Mardsen in [
30], some issues as the following ones might deserve interest for a further discussion:
Finally, let us discuss an example analyzed from the Lagrangian viewpoint in [
1,
16] taking into account also the observers’ one in
Section 3.2.
Example 2. The gravitational field sourced by a kinetic gas has been deeply studied in [1,16]. In the relativistic setting, this is derived from the Einstein–Vlasov equations in terms of a 1-particle distribution function (1PDF) , which encodes how many gas particles at a given spacetime point x propagate on worldlines with a normalized 4-velocity . Specifically, the stress energy tensor is:being the indicatrix (future-directed unit vectors of the Lorentz metric) and the volume element induced by the scalar product at each x. In [1], they propose to derive the gravitational field of a kinetic gas directly from the 1PDF without averaging, i.e., taking into account the full information on the velocity distribution. This leads to consider the function , as an energy–momentum function, which plays the role of a stress–energy tensor (even though it is a scalar rather than a 2-tensor). Moreover, the original Lorentz metric is naturally allowed to be Lorentz–Finsler, which permits to obtain more general cosmological models (Section III in [1]). Indeed, up to a coupling constant, ϕ is regarded directly as the matter source in the Finslerian Einstein–Hilbert equation (i.e., it is placed at the right-hand side of this equation, (Equation (7) in [1])). It is worth pointing out: ϕ can be reobtained as a Lagrangian energy-momentum by inserting it directly as a term in the background Lagrangian (Equation (75) in [16]). However, the Lagrangian is not natural then, as ϕ is written in terms of (recall (Appendix 3, Section (a) in [16])). As discussed above, such a function allows one to construct several tensors, in particular, the vertical Hessian (as in (10)), which also might play a role to compare with the relativistic .
Anyway, starting at the 1PDF ϕ, another Finslerian interpretation would be possible. In particular, one can define the energy–momentum distribution.
Then, given an observer and a -unit vector, the w-energy–momentum might be defined asIn particular, when , this would be the energy perceived by v, and when w is unit and -orthogonal to v would be (minus) the momentum in the direction w(compare with the discussion at the end of Section 3.2). So, an alternative stress–energy tensor perceived by each observermight be defined as the anisotropic tensor:where the integration in u is carried out with the volume form of , denoted by . 4. Divergence of Anisotropic Vector Fields
After studying the basic properties of the Finslerian stress–energy tensor
T, our next aim is to analyze the meaning and significance of the
infinitesimal conservation law . Along this and the next section, we will always consider an anisotropic tensor
interpreted as an endomorphism of anisotropic vector fields.
and
will be defined on vectors and 1-forms by
and
resp., where
is the inverse fundamental tensor, and
is the transpose of
T. They will have components
and
, and in principle we will not even assume that these are symmetric. We will be assuming that
M is orientable and oriented. This is not restrictive: one could always reduce the theory to this case by pulling back all the objects (the fibered manifold
included) to the oriented double cover of
M (Chapter 15 in [
31]).
Let us briefly recall the mathematically precise meaning of the conservation laws in classical General Relativity (
g,
T, and
X isotropic). One has
with ∇ the Levi–Civita connection. The first contribution vanishes due to
, and there are different situations in which the second one vanishes as well. For instance, if
is antisymmetric, then
and if
is symmetric and
is antisymmetric (equiv.,
X is a Killing vector field), then also
Anyway, whenever
, one can integrate (
11) and apply the pseudo-Riemannian divergence theorem to get the
integral conservation law
where
is a domain of appropriate regularity, ı is the interior product operator, and
is the metric volume form. In a sense that will be made more precise in §5, this is expressing that the total amount of
X-momentum in a space region only changes along time as much as it flows across the spatial boundary of the region. In this way, there is no “creation” nor “destruction” of
X-momentum in any space region.
Extending the infinitesimal or the integral conservation laws poses, first and foremost, the problem of appropriately defining the divergence of an anisotropic
T. Observe that a priori it is not clear even how to define the divergence of a vector field
Z, isotropic or not, as one could consider
for different anisotropic connections ∇, mainly Chern’s and Berwald’s. An alternative is to seek for a more geometric, hence,
unbiased, definition. For instance, the
metric (anisotropic) volume form of L,
for
positively oriented, is well-defined, and when
(i.e.,
Z is isotropic), so is the Lie derivative
(see § 5 in [
21]). So, by analogy with the classical case, one could think of
for defining
.
It turns out that the unbiased definition, including all
, is achieved with a modification of this Lie derivative that we will regard as an extension of the classical Lie bracket. We devote the next subsection to the technical mathematical foundations of such an
anisotropic Lie bracket, which needs of a nonlinear connection on
to be well-defined. All the maps
that will appear in
Section 4.1 will be
(anisotropic) tensor derivations in the sense of (Definition 2.6 in [
21]), and their local nature will be apparent, so we will not explicitly discuss it. For example, the Lie derivative along
is the only tensor derivation such that for
and
,
4.1. Mathematical Formalism of the Anisotropic Lie Bracket
During this subsection, we fix an arbitrary nonlinear connection given by
or by the nonlinear covariant derivative
(keep in mind (
1) and (
2)), and also an anisotropic vector field
.
For
, it is very natural to consider the commutator of the horizontal lifts of
Z and
X:
We recall that
is always vertical. Indeed,
, where
is the curvature tensor of the nonlinear connection (see [
17], where this curvature is regarded as an anisotropic tensor and the homogeneity of the connection is not really required). This means that the horizontal part of
has coordinates
, and this corresponds to a globally well-defined
A-anisotropic vector field:
Definition 1. is the anisotropic Lie bracket of Z and X with respect to the nonlinear connection.
Remark 1. The word “anisotropic” could be omitted in the previous definition, in the sense that for , there is no other Lie bracket, isotropic or not, defined in general. Nonetheless, (17) makes apparent that when (i.e., when Z and X are isotropic), coincides with the standard Lie bracket regardless of the connection. Lemma 1. Given a nonlinear connection , , and anisotropic vector fields , it holds that Proof. Observe that
which concludes (
18). In particular,
, and using this in (
17), (
19) follows. □
We also recall that the
torsion of an
A-anisotropic connection ∇ (18 in [
21]), (Definition 5 in [
20]) is the anisotropic tensor
defined first on isotropic fields
by
and then extended by
-bilinearity. Therefore, it can be regarded as and
-bilinear map
and it has coordinates
where the
’s are the Christoffel symbols of ∇
8.
Theorem 1. Let a nonlinear connection and an anisotropic vector field be fixed.
- (A)
If ∇
is any A-anisotropic connection whose underlying nonlinear connection is , then for any , (where is the torsion of ∇
). - (B)
By imposing the Leibniz rule with respect to tensor products and the commutativity with contractions, the map extends unequivocally to an (anisotropic) tensor derivation given byfor and . In coordinates, ifthen - (C)
The mapis also a tensor derivation. When ,for all , where is the Lie derivative (16), regardless of the nonlinear connection. - (D)
Given and (), it holds that
Proof. (A) It is straightforward to compute that the right-hand side of (
21) is
-multilinear. Moreover, the identity is trivial on isotropic vector fields
, as
in this case, which concludes.
(B) Given
, for
it follows from (
17) that
Thus, in order to respect the Leibniz rule, the only possibility is to define
Now, given
, in order to respect again the Leibniz rule and the commutativity with contractions, the only possibility is to define
on every
by
(
26), (
17), and (
27) make apparent that
is already local on functions, vector fields, and 1-forms, and they allow to compute
Finally, given
, one is led to define
by (
22). Clearly, this indeed provides a tensor derivation and (
23) follows from the evaluation of (
22) at
together with (
26) and (
28).
(C)
is a tensor derivation for any
, in particular for
(see (
17)). Thus, the difference
is again a derivation. As for the last assertion, where
, we are going to use (Proposition 2.7 in [
21]). For
, we have
(recall Remark 1). For
, we have
(see (
26), (
29), (
1), and (
16)). As
and
act the same on isotropic vector field and anisotropic functions, they are equal.
(D) Observe that for
, the term
vanishes in (
19). Moreover, if
and
, then
. Given a local reference frame
, and taking into account the last two identities and the definitions of
and
, it follows that
As
, (
25) follows. □
Definition 2. The tensor derivation defined in Theorem 1 (B) is the (anisotropic) Lie bracket with Z, while is the (anisotropic) Lie derivative along Z, both of them with respect to the connection.
Remark 2 (Anisotropic Lie bracket and Lie derivative).
The derivation defined in Theorem 1 (C) would be the Lie derivative along
Z with respect to
. Analogously to the discussion of Remark 1, what makes this name consistent is (24): whenever the Lie derivative along Z was already defined, coincides with it. Even though the Lie bracket and the Lie derivative are equal in the classical regime, it is heuristically useful to regard as the anisotropic generalization of the former and as that of the latter, in order to distinguish them. It is actually , and not , which will be relevant for the definition of divergence. The reason is that the former, as we will see below, has a clear geometric interpretation in terms of flows, while the latter would just add the term to that interpretation. Moreover, Theorem 1 (D) actually corresponds to a Cartan formula for whose full development we postpone for a future work. Thus, can be regarded as an initial guess for the divergence of Z, but we will not employ from now on. Let us observe that given a diffeomorphism that is the flow of an isotropic vector field Z, we can define the pullback of an anisotropic differential form as the anisotropic form given by , where is the -parallel transport of v along the integral curve of Z and .
Proposition 1. If and , thenwhere is the (possibly local) flow of Z. Proof. Observe that
can be obtained as
with
V an extension of
v such that
. Then (
25) and the classical formula for the Lie derivative in terms of the flow imply (
31). □
Remark 3. Even though, for convenience, we stated the previous geometrical interpretation for an s-form ω, it should be clear that it holds true for any r-contravariant s-covariant A-anisotropic tensor.
4.2. Lie Bracket Definition of Divergence
Finally, in this and the next subsections a pseudo-Finsler metric L defined on A is fixed again. In its presence, and in view of the Riemannian case and Proposition 1, the most natural way of defining the divergence of an anisotropic vector field Z is by . Here there is a canonical choice for : the metric nonlinear connection of L. The definition obtained this way is unbiased, in that one does not choose any anisotropic connection a priori. Notwithstanding, it will turn out to be most conveniently expressed in terms of the Chern connection.
Definition 3. For , its divergence with respect to the pseudo-Finsler metric L is the anisotropic function defined bywhere and are the metric nonlinear connection (4) and the metric volume form (15) of L, resp. Remark 4. Even though we will keep assuming it for simplicity, the hypothesis of M being orientable is not really needed for this definition. As in pseudo-Riemannian geometry, on small enough open sets it is always possible to choose an orientation, define with respect to it and put . The different definitions will be coherent because when the orientation changes, changes to and In particular, when M is orientable, is independent of the orientation choice.
Proposition 2. Let L be a fixed pseudo-Finsler metric defined on A, and let . If ∇
is any symmetric A-anisotropic connection such that its underlying nonlinear connection is the metric one and , then or in coordinates, This, in particular, is true for the (Levi-Civita)–Chern anisotropic connection of L, so one can take the Christoffel symbols to be those of (5). Proof. One expresses the
Z-Lie bracket of the volume form in terms of the anisotropic connection, analogously to the isotropic case. From (
15) and the fact that
is a tensor derivation, we obtain
(
26) and the fact that
is the underlying nonlinear connection of ∇ give
(
21) and
From these and
,
where the last equality is reasoned analogously as in the proof of (
25).
For the Chern connection, it can be checked that
by considering a parallel orthonormal basis with respect to a parallel observer
V along the integral curves of any vector field. The coordinate expression of
in this case concludes (
33). □
4.3. Divergence Theorem and Boundary Term Representations
Our Lie bracket derivation allows us to obtain a statement of the Finslerian divergence theorem that subsumes both Rund’s (3.17 in [
18]) and Minguzzi’s (Theorem 2 in [
19]). This way, it does not need of computations in coordinates from the beginning nor of the “pullback metric” (
in our notation). Naturally, our statement does not include Shen’s (Theorem 2.4.2 in [
32]), as this one is an independent generalization of the Riemannian theorem not dealing with anisotropic differential forms nor vector fields.
Lemma 2. For , the vertical derivative of is given bywhere is the mean Cartan tensor of L (see (3)). Proof. Let
, …,
be a positively oriented
-orthonormal basis for every
for a certain
. Then
for all
. This implies that
Moreover, as
,
Using this relation above, we conclude (
35). □
In the present article, by a
domain we understand a nonempty connected set that coincides with the closure of its interior
D; then, its boundary is
. Physically, it is very important to include examples in which different parts of
have different causal characters, and this tipically leads to the boundary not being totally smooth. Hence, we will make a weaker regularity assumption that still allows one to apply Stokes’ theorem on
. A subset of
M has
0 m-dimensional measure if its intersection with any embedded
m-dimensional submanifold
is of 0 measure in the smooth manifold
. Finally, the interior product of an
s-form
with a vector field
X will be
Theorem 2. Let L be a fixed pseudo-Finsler metric defined on A. If
- (i)
is an anisotropic vector field,
- (ii)
is an A-admissible field with open, and
- (iii)
is a domain with smooth up to subset of 0 -dimensional measure on M and compact,
thenwhere is the mean Cartan tensor, and is computed with the metric nonlinear connection (4). Proof. The idea is to apply Stokes’ theorem to
. However, taking into account (
25) and Lemma 2, it follows that
concluding (
36). □
Remark 5 (Riemannian and Finslerian unit normals). Let be the inclusion of a smooth open subset .
- (i)
Even though we do not use the pseudo-Riemannian metric to derive Theorem 2, from our physical viewpoint it is natural to use it to re-express the boundary term. If Γ
is non--lightlike, then for a -normal field and a transverse field X along i, the form is nonvanishing and independent of X. In particular, is independent of the scale of , which we will always assume to be -unitary and D-salient, so coincides with the hypersurface -volume form of Γ.
Taking into account that vanishes wherever is tangent to Γ
and that , (37) allows us to represent and the right-hand side of (36) as In fact, this is how Rund’s divergence theorem follows from Theorem 2.
- (ii)
There is another way that one can try to represent the boundary term. Namely, assume that there exists a smooth with and (in the Lorentz–Finsler case, it will necessarily be ). This is called a Finslerian unit normal along
. Analogously as in (i), one can put here, due to the possible orientation difference between both sides, In fact, this is how Minguzzi deduces his divergence theorem (Theorem 2 in [19]). Note, however, that he does it under the hypothesis of vanishing mean Cartan tensor (), which implies that is independent of V. As we do not require this, Theorem 2 is more general statement than Minguzzi’s. - (iii)
The Finslerian unit normal presents some issues in the general case, as we are not taking . In our physical interpretation, with L Lorentz–Finsler, A consists of timelike vectors, so asking for a Finslerian unit normal is only reasonable when Γ is L-spacelike, that is, for . In such a case, the strong concavity of the indicatrix guarantees the existence and uniqueness of ξ: one defines to be the unique vector such that and the indicatrix are tangent at .
- (iv)
Of course, if L comes from a pseudo-Riemannian metric on M, then and .
- (v)
It should be clear from this discussion that the form that one integrates on the right-hand side of (36) is always the same and that the only difference between Rund’s and Minguzzi’s divergence theorems is how each of them represents it. Notwithstanding, this is an important difference, for the boundary terms (38) and (39) could potentially have different physical interpretations.