The homotopy momentum map of
general relativity
Abstract.
We show that the action of spacetime vector fields on the variational bicomplex of general relativity has a homotopy momentum map that extends the map from vector fields to conserved currents given by Noether’s first theorem to a morphism of -algebras.
Key words and phrases:
general relativity, multisymplectic geometry, homotopy momentum map, lagrangian field theory, variational bicomplex, diffeomorphism symmetry2020 Mathematics Subject Classification:
53D20; 18N40, 37K06, 83C051. Introduction
1.1. Motivation
The diffeomorphism symmetry of general relativity, a mathematical implementation of the Einstein equivalence principle, is one of its defining features. In contrast to the internal symmetry of gauge theories, diffeomorphisms are external symmetries since they act not only on the fields (i.e. lorentzian metrics), but also on spacetime. The initial value problem, which yields the hamiltonian formulation of the field dynamics, and the presymplectic structure on the space of fields, which yields the Poisson bracket of observables, both depend on the choice of a codimension 1 submanifold as initial time-slice. But such a submanifold is not invariant under diffeomorphisms. In physics terminology: it breaks the symmetry. The consequence is that the basic ingredients of quantization, the hamiltonian and the Poisson bracket, are not compatible with the diffeomorphism symmetry. This issue lies at the heart of some of the fundamental open problems in general relativity and has captivated the interest of many authors since the 1960s.
One of its mathematical symptoms is that the action of the group of diffeomorphisms and the action of the Lie algebra of vector fields are not hamiltonian. More precisely, Noether’s first theorem, which associates to a symmetry a conserved momentum does not define a homomorphism of Lie algebras. (The Noether momenta are the components of the Einstein tensor integrated over the codimension 1 submanifold.) Worse, the space of Noether momenta is not even closed under the Poisson bracket.
In an earlier paper we could show that there is a natural diffeological groupoid describing the choices of initial submanifolds, which exhibits the Poisson brackets as the bracket of its Lie algebroid [BFW13]. Next, we have developed a notion of hamiltonian Lie algebroids, which generalizes the notion of hamiltonian Lie algebra action to the setting of Lie algebroids [BW]. We have conjectured that the Noether charges of general relativity are the components of the momentum section of a hamiltonian Lie algebroid, which would give a conceptual explanation of some of the intriguing features of the constraint functions. Finally, in [BSW] we have interpreted the momenta as elements of a generalized Lie-Rinehart algebra, which is connected to the BV-BFV approach to boundary conditions in classical field theories.
In this paper, we sidestep the choice of initial submanifolds altogether by using higher algebraic structures. We show that the map from vector fields to their Noether currents is part of a homotopy momentum map in the sense of multisymplectic geometry.
1.2. Content and main results
In Sec. 2 we study the premultisymplectic form of a lagrangian field theory (LFT), where is the Euler-Lagrange form and a boundary form. We prove in Prop. 2.4 that the obstruction of a premultisymplectic vector field to be hamiltonian lies in bidegree of the variational bicomplex, where is the dimension of the spacetime manifold. This shows that the -algebra of hamiltonian forms can be interpreted as generalized current algebra. We introduce the notion of manifest diffeomorphism symmetry (Def. 2.12) and observe that every such symmetry has a hamiltonian momentum map that is given explicitly in terms of the lagrangian and the boundary form (Prop. 2.15).
In Sec. 3 we consider the case of general relativity. Generalizing the concept of tensor fields, we introduce the notion of covariant and contravariant families of forms in the variational bicomplex (Def. 3.3). We then show that the product of families, the contraction of indices, the raising and lowering of indices, etc. have properties that are analogous to tensor fields. In Sec. 3.5 we introduce the notion of covariant derivative of families of forms. In Props. 3.13 and 3.14 we derive divergence formulas that show that horizontally exact forms can be expressed as the contraction of families of forms with the covariant derivative.
1.3. Relation to previous work
The study of multisymplectic forms in classical field theory goes back to at least the 1970s. Best known is perhaps the highly influential, but never finished GiMmsy project (named by the initials of the collaborators involved, with the main protagonists Gotay and Marsden capitalized). Their goal was “to explore some of the connections between initial value constraints and gauge transformations” in classical field theories with constraints, such as general relativity [GIMM04, p. 1]. Towards this end they introduced the notion of multimomentum maps [GIMM04, p. 46] (see also [CnCI91, Sec. 4.2]). Given the action of a Lie algebra on a manifold with a closed -form , a multimomentum map was defined as a smooth map or, equivalently, a linear map , such that for all .
Missing from this definition was a requirement of compatibility with the Lie bracket of , analogous to hamiltonian momentum maps in symplectic geometry. It seems natural to require to be -equivariant. Alternatively, the hamiltonian forms in can be equipped with a “Poisson bracket” and required to commute with the brackets. However, both conditions turn out to be too strong and rarely satisfied. Moreover, the “Poisson bracket” does not satisfy the Jacobi identity, so that it is not immediately clear for what kind of algebraic structure should be a homomorphism.
This lack of compatibility of algebraic structures leads to issues in the study of the constraints of classical field theories with diffeomorphism symmetries, of which general relativity is the theory “par excellence” [GIMM04, Interlude I, p. 52]. The constraint functions of general relativity are given by the values of the multimomentum map integrated over the Cauchy surface. The resulting map is called the energy-momentum map [GIM04, Sec. 7B]. While the energy-momentum map shows that the constraint functions arise from the multimomentum map, it does not explain the relation between the Lie brackets of the symmetry algebra of vector fields and the Poisson brackets of the constraints. (For the history of this often studied but elusive problem see Sec. 4 of [BFW13].) From the viewpoint of homotopical algebra, this was to be expected: Lie brackets that satisfy the Jacobi identity up to exact terms and maps that preserve the brackets up to exact terms are generally not compatible with homotopies of the underlying complexes. Instead we have to use the homotopy algebraic structure, i.e. the minimal extension of Lie algebras to differential complexes that is stable under quasi-isomorphisms. For a better behaved theory of multimomentum maps, we are thus led to -algebras.
In [Rog12, Thm. 5.2] it was shown that the bracket on hamiltonian forms in has a natural extension to an -algebra structure on a graded subspace of the de Rham complex, with the 1-bracket given by the de Rham differential.111In [BFLS98] it was shown that the bracket of integrated local functions on the jet bundle has extensions to alternative -algebra structures on cohomological resolutions. However, these -structures were not given by an explicit construction, depended on choices, and did not suggest a stronger notion of multimomentum maps. It was realized in [CFRZ16, Def./Prop. 5.1] that this is the natural setting for the generalization of hamiltonian momentum maps to the multisymplectic setting, defined as morphisms of -algebras. The component of every homotopy momentum map is a multimomentum map [CFRZ16, Sec. 12.1]. For the obstructions to lifting a multimomentum map to a homotopy momentum map see [CFRZ16, Sec. 9.2].
In local lagrangian field theories, a multimomentum map is given by Noether’s theorem [BHL10, Sec. 4.1]. Finding a homotopy momentum map, however, is a much more difficult problem, even more so in general relativity, where the Hilbert-Einstein lagrangian is non-polynomial in the fields and of second jet order. The situation simplifies greatly if the multi(pre)symplectic form has a primitive, , and if the action leaves invariant. Then a homotopy momentum map can be defined by inserting the fundamental vector fields in [CFRZ16, Sec. 8.1]. If we want to check whether this applies to a classical field theory, we have to identify the correct as well as the correct action of the diffeomorphism group. Many authors use for the boundary 1-form, so that is the universal current in the sense of [Zuc87], and assume that the action is vertical (e.g. [BHL10]). We will show that instead we have to take the sum of the lagrangian and the boundary 1-form for and the diagonal action (12) on fields and spacetime by vertical and horizontal vector fields.
1.4. Conventions
The spacetime manifold is assumed to be smooth, finite-dimensional, and second countable. The infinite jet manifold of a smooth fibre bundle is viewed as pro-manifold, so that its de Rham complex, i.e. the variational bicomplex, is an ind-differential complex. For the computations and proofs in this paper, however, this will not play much of a role. The same goes for the diffeological structure on the space of fields , of which we will only use the fact that the diffeological tangent bundle is given by the space of sections of the vertical tangent bundle . For the grading and differentials of the variational bicomplex we use the notation of [DF99]: A form in has vertical degree and horizontal degree . The vertical differential is denoted by and the horizontal differential by . Otherwise, we follow [And89], with specific references given in the text. We use the summation convention throughout the paper, so that all repeated indices are being summed over.
Remark 1.1.
Instead of “momentum map”, many authors use the term “moment map”, which derives from a mistranslation of the French term “moment” as in “moment cinétique” (angular momentum) or “application moment” [Sou70].
1.5. Brief review of homotopy momentum maps
For the reader’s convenience, we give a brief review of the main notions of multisymplectic geometry used in this paper. This is also necessary to fix the notation, the choice of gradings, and the signs.
Let be a manifold with a closed -form . A pair consisting of a vector field and a form is called hamiltonian if
A vector field or a form is called hamiltonian if it is part of a hamiltonian pair. We denote the space of hamiltonian vector fields by and the space of hamiltonian forms by .
For the pair we can construct an -algebra defined as follows [Rog12, Thm. 5.2]. The -graded vector space is
The brackets are defined by
for , by
for and where are hamiltonian pairs, and by zero in all other cases. With this degree convention, the degree of is .
Definition 1.2 (Def./Prop. 5.1 in [CFRZ16]).
Let be a manifold with a closed -form . Let be a homomorphism of Lie algebras. A homotopy momentum map of the action is a homomorphism of -algebras
such that
for all .
We recall that a morphism of -algebras is given by a family of linear maps , of degree , subject to relations that are best expressed either in terms of the -operad or in the language of formal pointed manifolds. If the domain is a Lie algebra, as is the case for a homotopy momentum map, the conditions for to be a morphism simplify greatly. They are best expressed in terms of the boundary operator of the Chevalley-Eilenberg complex for Lie homology,
A collection of linear maps is a homotopy momentum map if and only if [CFRZ16, Prop. 3.8]
for all , where we set and . This relation can be interpreted homotopically as follows. Shifting the degree of by and shifting the degree of the de Rham complex by , the right hand side can be expressed in terms of the degree 0 map
where we have used that . The maps have degree . The condition for to be a morphism of -algebras can be written succinctly as [CFRZ16, Sec. 6.2]
(1) |
that is, a homotopy momentum map is a null-homotopy of the map of cochain complexes .
In degree the condition (1) reads , that is is a hamiltonian pair. With this relation, can be expressed in terms of the -brackets as
For , Eq. (1) is spelled out as
(2) |
which shows that the failure of to be a homomorphism of Lie algebras is a -exact term. For we obtain
Proposition 1.3 (Sec. 8.1 in [CFRZ16]).
Let for some . If is invariant under the action , i.e.
for all , then it has a homotopy momentum map given by
Notation 1.4.
For shorter notation we will write the -bracket also as
Analogously, we will write the momentum map as
In this notation, Eq. (2) is written as
(3) |
2. Multisymplectic geometry of lagrangian field theories
The space of fields of a field theory is the set of sections of a fibre bundle over a manifold , naturally equipped with the functional diffeology. The lagrangian is a map , where is the dimension of . We will assume that the lagrangian is local, i.e. a differential operator, so that , where is the lagrangian form and is the infinite jet prolongation of the field .
If is compact, we can define the action by . Many interesting and important lorentzian spacetimes are not compact, however, so that the action is generally not well-defined. Therefore, we have to formulate the action principle, the derivation of the field equations, the notion of symmetries, etc. in a cohomological form within the variational bicomplex [Zuc87, DF99].
In Sec. 2.1 we state the action principle in its cohomological form, essentially replacing integration by taking cohomology classes with respect to the spacetime differential . By the cohomological version of partial integration the variation of the lagrangian can be written as , where is the Euler-Lagrange form and the the boundary form [Zuc87]. can be viewed as the differential operator of the field equations, so that it governs the dynamics of the field theory. The integration of over a codimension 1 submanifold of spacetime yields a presymplectic form on the space of fields, so that it describes the Poisson brackets of the field observables.
For the multisymplectic approach we will consider , which is an exact -form of degree . Its primitive is the Lepage form . In Prop. 2.4 we show that the bidegree -component of a hamiltonian form is a conserved current in the sense of [Zuc87]. The -algebra associated to as in [Rog12] can therefore be viewed as a higher version of the current algebra.
If is closed, so that the action is defined, a symmetry of the LFT is an automorphism of that leaves the action invariant, . Infinitesimally, a vector field on is a symmetry if . This is the case if and only if is -exact, which we take as the general definition of a symmetry. For a local lagrangian we have to require that the vector field , too, is local. In Sec. 2.3 we observe that a vector field on the diffeological space is local if and only if it descends to a vector field on , which is the infinite prolongation of an evolutionary “vector field”. Such vector fields are strictly vertical in the sense that their inner derivative commutes with the horizontal differential. The strictly horizontal vector fields whose inner derivative commutes with the vertical differential are the lifts of the spacetime vector fields by the Cartan connection.
In Def. 2.12 we introduce the notion of manifest diffeomorphism symmetry, which is an action of the Lie algebra of spacetime vector fields on the infinite jet bundle, such that is the sum of a strictly vertical vector field and the Cartan lift of that leaves the Lepage form invariant, . We point out in Prop. 2.15 that such a symmetry has a homotopy momentum map given by inserting the fundamental vector fields of the action into the Lepage form, which is a special case of the well-known Prop. 1.3.
2.1. The cohomological action principle
A variety of ingredients can play a constitutive role in the mathematical study of classical field theories. For the purpose of this paper the following minimal definition will suffice:
Definition 2.1.
A local lagrangian field theory (LFT) consists of a manifold of dimension , called the spacetime, a fibre bundle , called the configuration bundle, and a form in the variational bicomplex, called the lagrangian.
Let , where denotes the vertical and the horizontal degree. The vertical differential will be denoted by , the horizontal differential by . The form is represented by a form on a finite dimensional jet manifold , which is given by a map . In this way gives rise to a -th order differential operator
where is the -th jet prolongation of . In this notation, the integrand of the action can be written as .
The target of the differential operator is not a vector space, so it does not make sense to consider the equation “”, even though this is how the corresponding PDE is often written. And even if is a vector bundle so that is a vector bundle, the right hand side cannot be required to be the zero section, as this would imply that is the zero section of . Instead, we have to use that there is a zero form in every bidegree. The PDE can then be written more carefully as . If this equation holds, we will say that vanishes at .
Definition 2.2.
A form is said to be -exact at if there is a form such that vanishes at .
Definition 2.3.
A field is said to satisfy the cohomological action principle if is -exact at .
If a form is of top horizontal degree , then there is a unique representative of its -cohomology class, which has the following property: The form is -exact at if and only if vanishes at . The map is the cohomological version of partial integration and straightforward to compute. It is called the interior Euler operator. Forms in the image of are called “source” for and “functional” for [Tak79], [And89, Def. 2.5 and Ch. 3].
The map is called the Euler operator. The source form is called the Euler-Lagrange form. Since the interior Euler operator does not change the -cohomology class, and are in the same -cohomology class, i.e.
for some . The form is called a boundary form.
From the properties of source forms it follows that a field satisfies the cohomological action principle if and only if it satisfies the Euler-Lagrange equation
In physics terminology, such a field is called on shell.
2.2. Premultisymplectic structure and current -algebra
Let be a boundary form. The form
(4) |
of total degree will be called the Lepage form222For the terminology see [Kru83] or [And89, p. 199]. Deligne and Freed call the “total Lagrangian” [DF99, p. 161].. Let the total differential of be denoted by . The total differential of the Lepage form is
(5) |
which is the premultisymplectic structure we are interested in.
On we have the splitting of vector fields into a vertical and horizontal component which leads to the bigrading on the de Rham complex. Moreover, we have the acyclicity theorem of the variational bicomplex. This leads to the following description of hamiltonian vector fields.
Proposition 2.4.
Let be a vector field on with vertical component . Then is hamiltonian with respect to the premultisymplectic form if and only if
-
(i)
, and
-
(ii)
for some .
In the proof, we will use the following lemma [Del18, Thm. 11.1.6].
Lemma 2.5.
Let be a -closed form and
its decomposition into summands of bidegree . Then is -exact if and only if is -exact.
Proof.
A form can be decomposed as
into components of bidegree . The total differential decomposes as
into summands of homogeneous bidegree, where the first summand has bidegree and the last . Assume that . This condition must hold in each bidegree individually. In particular we have .
Conversely, assume that for some . The total differential of decomposes as
into summands of homogeneous bidegree, where the first summand has bidegree and the last . By assumption , which has to hold in each bidegree separately,
From the first equation we get
It follows from the acyclicity theorem for the variational bicomplex [Tak79, Thm. 4.6] that for some . The bidegree component of can now be written as
As before, it follows from the acyclicity theorem that for some . By induction, we obtain forms such that for . ∎
Proof of Prop. 2.4.
Let be the vertical and the horizontal component of . Assume that . The left hand side decomposes as
into summands of bidegree , , and . We conclude that the bidegree component of the hamiltonian condition is , which is condition (ii) for . Since is closed, we have which is condition (i).
Conversely, assume that (i) and (ii) hold. This means that is -closed and that , where . It now follows from Lem. 2.5 that . ∎
A form is also called a current. A current is called conserved if it is -closed on shell, i.e. if vanishes at every solution of the Euler-Lagrange equation. Prop. 2.4 shows that the degree component of a hamiltonian form is a conserved current. In this sense, can be viewed as higher current algebra.
The -bracket of hamiltonian forms is given by
where are the hamiltonian vector fields. The 2-bracket is given by
(6) |
where and are hamiltonian pairs.
2.3. Noether symmetries
The diffeological tangent space of is given by the space of sections of the vertical tangent bundle, , so that a vector field on is given by a map . This map is called local if it is a differential operator, i.e. if there is a commutative diagram
where is the vertical vector bundle. The map is often called an evolutionary “vector field”.
Remark 2.6.
We put quotes around evolutionary “vector field” because it cannot be naturally viewed as an actual vector field unless the configuration bundle is equipped with a flat connection. Readers who are used to this traditional (but abusive) terminology (e.g. Def. 1.15 in [And89]) are kindly asked to ignore the quotes.
The map can be prolonged to a vector field on , which is the unique vector field such that
(7) |
commutes. If such a commutative diagram exists, we will say that lifts to the vector field . The following proposition is a purely algebraic characterization of such vector fields.
Proposition 2.7.
A vector field lifts to a vector field on if and only if .
Proof.
The proof follows from a straightforward computation in jet coordinates. ∎
The kernel of defines an integrable distribution on , called the Cartan distribution, which can be interpreted as a flat Ehresmann connection on . The horizontal lift of a vector field is denoted by . Since the connection is flat, the map , is a homomorphism of Lie algebras. The following is a purely algebraic characterization of such lifts.
Proposition 2.8.
A vector field on is the horizontal lift of a vector field on by the Cartan connection if and only if .
Proof.
The proof follows from a straightforward computation in jet coordinates. ∎
The last two propositions can be understood geometrically as follows. Assume for the sake of argument that is a finite dimensional manifold. The de Rham complex of has a bigrading with vertical differential in the direction of and horizontal differential in the direction of . A vector field on is the lift of a vector field on if and only and a lift of a vector field on if and only if . Props. 2.7 and 2.8 show that this characterization is valid also in the variational bicomplex. In order to emphasize this geometric interpretation, we will use for the purpose of this paper the following terminology:
Definition 2.9.
A vector field on will be called strictly vertical if and strictly horizontal if .
The Lie derivatives of a strictly vertical vector field and of a strictly horizontal vector field are given by
Definition 2.10.
A strictly vertical vector field such that for some will be called a Noether symmetry of the LFT.
Remark 2.11.
Vector fields on the infinite jet bundle are sometimes called “generalized vector fields” and symmetries given by such vector fields “generalized symmetries” (e.g. in [DF99]). However, an analysis of Noether’s historic paper shows that this is Noether’s original notion of symmetry, which was only to be rediscovered later [KS11, Sec. 7.1].
Recall that a form is also called a current. If there is a strictly vertical vector field such that
then is called a Noether current and a Noether pair [DF99, Def. 2.97]. Noether currents are conserved. Noether’s first theorem states that if is a Noether symmetry, then
is a Noether current. The proof is a half-line calculation,
which highlights the advantage of working in the variational bicomplex.
2.4. Manifest diffeomorphism symmetries
In [DF99, p. 169], a manifest symmetry was defined to be a vector field such that:
-
(i)
is the sum of a strictly vertical vector field and a strictly horizontal vector field .
-
(ii)
.
This suggests the following terminology:
Definition 2.12.
Let be a LFT with boundary form . An action
by manifest symmetries will be called a manifest diffeomorphism symmetry.
Remark 2.13.
The Cartan lift of vector fields on is a homomorphism of Lie algebras. Since strictly vertical and strictly horizontal vector fields commute, it follows that the map is a homomorphism of Lie algebras, too.
Remark 2.14.
If is a natural bundle, i.e. diffeomorphisms between open subsets of lift functorially to diffeomorphisms between local sections, then it follows from [ET79] that we have an action of vector fields on . The diffeomorphism symmetries of LFTs often arise in this way.
Proposition 2.15.
Let be an LFT with boundary form . Then every manifest diffeomorphism symmetry has a homotopy momentum map
given by
Proof.
This is a special case of Prop. 1.3. ∎
The homotopy momentum map of a single vector field is split into a bidegree and a bidegree summand as
(8) |
where
(9) |
is the Noether current of . In general, the map splits into a and a component given by the two lines of the right hand side of the equation
For example, we have
Using Eq. (6), we can write the -bracket of the momenta as
(10) |
where the three lines of the right hand side are of bidegrees , , and . The right hand side of Eq. (3) is expressed in terms of the Noether current as
Remark 2.16.
If we integrate over a closed codimension 1 submanifold , we see from Eq. (8) that we obtain, up to a sign, the usual Noether charge . This is no longer true for the brackets. The integral of the first summand on the right hand side of Eq. (10) is the usual bracket of charges. The integral of the second summand, however, is an additional contribution, which is not present in the multimomentum map of [BHL10, Sec. 4.1]. The integrals of all other terms on the right hand side of Eq. (10) vanish for degree reasons.
Example 2.17 (Classical mechanics).
In classical mechanics spacetime is time and the configuration bundle is trivial, , so that is the space of smooth paths in . Let us consider the lagrangian of a particle of mass 1 in a potential ,
Here are coordinates on the infinite jet bundle, given by
for a path . Using the relations , , and , we find that with
For the presymplectic form we obtain
which is a form on . The Cartan lift of the infinitesimal generator of time translation, i.e of the coordinate vector field is
The time translation of paths descends to the strictly vertical vector field
The fundamental vector field of the diagonal action of time translation on is therefore given by
(11) |
This equation looks like a tautology, but the vector field on the right hand side is not horizontal and must not be identified with the vector field in the time direction. Moreover, is not -linear.
Eq. (11) implies that , so that time translation is a manifest symmetry. The corresponding momentum map is given by
since for degree reasons the term vanishes. The Noether momentum
is the energy.
3. The variational bicomplex of lorentzian metrics
We turn to general relativity. Here, the fields are lorentzian metrics on the spacetime manifold . Vector fields on act on metrics by the Lie derivative. This action is local, so that it descends to the infinite jet bundle, inducing an action on the variational bicomplex. In order to study this action, we introduce in Def. 3.3 the concept of covariant and contravariant families of forms in the variational bicomplex, which generalizes the concept of tensor fields. In Sec. 3.5 we generalize the notion of covariant derivative to such families of forms. In Sec. 3.6 we derive divergence formulas that express the horizontal differential of a form in terms of the covariant derivative and the metric volume form. While the computations are similar to those with tensor fields, there are also differences. For example, the metric volume form is invariant (Lem. 3.11), rather than transforming as a density.
3.1. The action of spacetime vector fields
Assume that is a manifold of finite dimension . The configuration bundle of general relativity is the bundle of fibre-wise lorentzian metrics on the tangent spaces of the spacetime manifold , which we denote by . We use the “east coast” sign convention in which the signature of the metric is . The diffeological space of lorentzian metrics on will be denoted by .
Remark 3.1.
In many papers on LFTs and the variational bicomplex one of the the following simplifying assumptions about the configuration bundle is made: is a vector bundle; the fibres of are connected; the space of sections is non-empty; the jet evaluations are surjective. All these assumptions generally fail for the bundle of lorentzian metrics.
The configuration bundle is natural, which means that local diffeomorphisms on lift functorially to the sheaf of sections. In particular, we have a left action of the diffeomorphism group on the space of fields by pushforward. Infinitesimally, we have a left action of the Lie algebra of vector fields,
where the symmetric 2-form represents a tangent vector in . This action is local, so that it descends to an action of on by strictly vertical vector fields,
Together with the Cartan lift of the vector field in , we obtain a homomorphism of Lie algebras
(12) | ||||
Our ultimate goal is to show that is a manifest symmetry of general relativity for a natural choice of boundary form. In this section we will gather the necessary tools.
3.2. Jet coordinates
Let be a system of local spacetime coordinates on an open subset . The coordinate vector fields will be denoted by , the coordinate 1-forms by . A lorentzian metric is written in local coordinates as , where are the matrix components of the metric. (Recall that we use the Einstein summation convention throughout the paper.)
The local coordinates on induce local jet coordinates given by
Since the partial derivatives commute, is invariant under permutations of the indices . To avoid overcounting in summation formulas it is convenient to use the multi-index notation of multi-variable analysis: A multi-index is a tuple of natural numbers . The number is called the length of the index. The concatenation of a multi-index with a single index is given by
The jet coordinate function labeled by a multi-index is given by
The collection of functions for and is a system of local coordinates on .
Remark 3.2.
In the physics literature, the same notation is usually used for both the jet coordinates and their evaluation on a field, which can be confusing. For example, if is non-compact, every -form is exact, in particular the integrand of the action. So for the step “discarding exact terms” during the derivation of the Euler-Lagrange equation to be meaningful, we must view the integrand as an element , i.e. as an expression of the jet coordinates like and not of the derivatives of a particular metric .
The variational bicomplex is generated as bigraded algebra by the coordinate functions, the vertical coordinate 1-forms in degree , and the horizontal coordinate 1-forms in degree . A -form is given in local coordinates by
where the coefficients are functions on . The other differentials of the jet coordinates are given by [And89, p. 18]
It follows that the differentials of the coordinate 1-forms are given by , , , and
Dually, the -module of vertical vector fields is spanned by the coordinate vector fields , which satisfy
The module of horizontal vector fields, called the Cartan distribution, is spanned by the vector fields
which satisfy
The Cartan distribution can be viewed as an Ehresmann connection on the bundle . The horizontal lift of a vector field on to is given by
The vertical and horizontal differentials of a function are given by [And89, pp. 18-19]
The horizontal differential of a form is given in local coordinates by
(13) |
A vector field is strictly horizontal if and only if it is the horizontal lift of a vector field on by the Cartan connection. A vector field is strictly vertical if and only if it is the infinite prolongation of an evolutionary “vector field”, i.e. of a map of bundles over , where is the vertical tangent bundle. In local coordinates it is of the form
(14) |
where the are functions on and where is the multi-index notation for the iterated application of the horizontal lifts of the coordinate vector fields.
3.3. Action of spacetime vector fields on infinite jets
The action of a vector field on a lorentzian metric by the negative Lie derivative, , is given in local coordinates by
We can view this as transformation of the coordinate functions
(15) |
which are the components of the evolutionary “vector field” . Its infinite prolongation is the strictly vertical vector field
which defines the action (12) of vector fields on the infinite jet bundle.
3.4. Covariant and contravariant families of forms
The Lie derivative of a coordinate function with respect to a strictly horizontal vector field is given by
In particular, we have
Note that this is the Lie derivative of a single function and must not be confused with the Lie derivative of a metric 2-form on . The formula (15) for the 0-jet component of can now be written as
This can be expressed in terms of the diagonal action as
(16) |
Since commutes with both and , it commutes with . This implies that
(17) |
Using , we get
(18) |
These calculations suggest the following definition.
Definition 3.3.
A family of forms , is called covariant in and contravariant in if
A form is called invariant if .
Def. 3.3 generalizes the notion of covariant and contravariant tensors to families of forms in . In this terminology Eqs. (16), (17), and (18) show that the indices of and are covariant, while those of and are contravariant. Covariant and contravariant families behave in many ways as tensors.
Lemma 3.4.
Let be a covariant and a contravariant family of forms. Then the family is covariant in and contravariant in .
Proof.
This follows immediately from the fact that is a degree 0 derivation of the algebra . ∎
Lemma 3.5.
Let be a family of forms that is covariant in and contravariant in , then the contracted form (summation over ) is invariant.
The last two lemmas generalize in an obvious way to families with several indices. An immediate consequence of Lem. 3.4 and Lem. 3.5 is that we can raise and lower indices with the metric coordinate functions in the usual way: If is covariant, then is contravariant. If is contravariant, then is covariant.
Lemma 3.6.
If the family is covariant, then the family is covariant in and .
Proof.
Let We have
which shows that is covariant in and . ∎
Lemma 3.7.
If the family is covariant, then the family is covariant.
Proof.
We have
which shows that is covariant. ∎
The last lemma generalizes in an obvious way to families of forms with covariant and contravariant indices. The analogous statement for the horizontal differential works only for invariant forms:
Lemma 3.8.
If is invariant, then is invariant.
Proof.
The differential commutes with , so that . ∎
Lemma 3.9.
If the form is invariant, then the family is covariant.
Proof.
We have
which shows that is a covariant family. ∎
Lemma 3.9 holds only for an invariant form . If is a covariant family, then is not covariant. In order to obtain a covariant family by differentiation we have to generalize the concept of covariant derivative to families of forms in the variational bicomplex.
3.5. Covariant derivative of families of forms
In the cohomological approach to general relativity, we have to interpret the connection coefficients, the covariant derivative, the curvature, the volume form, etc. as expressions in the variational bicomplex. The connection coefficients of the Levi-Civita connection have to be viewed as functions on that are given in local coordinates by the expression
(19) |
The covariant derivative has to be defined in the variational bicomplex as follows. For a family of forms that is covariant in the lower indices and contravariant in the upper indices we define
Using this definition, we can check by the usual calculation that the connection coefficients (19) of the Levi-Civita connection is the unique family of functions symmetric in and , such that . The Riemann curvature tensor is given by , which has now to be viewed as a family of functions on .
Lemma 3.10.
Let be a covariant family of vertical forms. Then the family is covariant in and .
Proof.
We have to compute the Lie derivative of with respect to . For the first summand we get
For the second summand we must compute the Lie derivative of the connection coefficients. For this we need the following formula.
With this relation, we can compute the action of vector fields on the connection coefficients, which yields
Putting everything together, we obtain
where the terms containing the second order derivatives of cancel. This finishes the proof. ∎
3.6. Divergence formulas
In the variational bicomplex, the metric volume form is the -form on defined as
(20) |
We recall that we have adopted the “east coast” sign convention for Lorentz metrics with 1 negative and positive signs, so that is negative. The partial derivative of the square root of the determinant with respect to the 0-jet coordinates is given by
The partial derivatives with respect to and all higher jet coordinates vanish. It follows that the vertical and the horizontal differentials are given by
For the vertical differential of the volume form we obtain
(21) |
Although is not a volume form on , every -form can be written as
for a unique function . Therefore, we can define the divergence of a vector field by the relation
For a strictly horizontal vector field we have
(22) |
We conclude that
While this looks like the usual expression, we point out that the divergence is now a function on .
Lemma 3.11.
The metric volume form is invariant.
Proof.
For the Lie derivative of the volume form with respect to the vertical vector field we obtain
where in the last step we have used Eq. (22). We conclude that for all . ∎
Remark 3.12.
Lem. 3.11 can be stated by saying that is divergence free.
From the formula for the divergence of a vector field we deduce
This formula generalizes to higher vertical forms, as we will show next.
Proposition 3.13.
Let be a family of -forms on . Then
(23) |
Proof.
For later use, we generalize the formula (23) further to families of -forms.
Proposition 3.14.
Let be a family of -forms on such that . Then
(24) |
Proof.
Consider the -form
We have the relation
Moreover, since , we have
Using these relations, we can compute the horizontal differential of as
which finishes the proof. ∎
4. The homotopy momentum map of general relativity
We now have all the tools needed for the multisymplectic interpretation of the diffeomorphism symmetry of general relativity. We start by recalling the Euler-Lagrange and the standard boundary form. Then we show in Thm. 4.1 that the Lepage form is invariant under the diagonal action of vector fields. In other words, the action of vector fields is a manifest diffeomorphism symmetry of general relativity in the sense of Def. 2.12. It follows from Prop. 2.15 that the symmetry has a homotopy momentum map, which is given explicitly in Thm. 4.2.
4.1. Euler-Lagrange and boundary form
The lagrangian form of the Hilbert-Einstein action is
(25) |
where is the scalar curvature, which has to be interpreted within the variational bicomplex as a function on as follows: The Riemann curvature tensor is given in local coordinates in terms of the connection coefficients (19) by
This is the usual formula [Wal84, Eq. (3.4.4)] with the partial coordinate derivatives replaced by the Cartan lifts and . The Ricci curvature is given by the contraction and the scalar curvature by the trace of the Ricci curvature .
The vertical differential of the scalar curvature is given by
The first term can be written as
The second term is given by [Wal84, Eq. (E.1.15)]
where
and where the covariant derivative is to be understood as
as explained in Sec. 3.5. The vertical differential of the volume form was computed in Eq. (21). Putting everything together, we get
The first term is the Euler-Lagrange form
where
is the Einstein tensor. The Einstein tensor is divergence-free, i.e.
Using Eq. (23), the second term can be written as a -exact term
where
(26) |
is the boundary form.
4.2. Invariance of the Lepage form
Theorem 4.1.
Proof.
The invariance must hold independently in every bidegree, so that we need to prove the two equations
We start by proving the invariance of . We have
(27) |
We will compute both summands separately. First we use (15) to compute
where we have used (19). With this formula we obtain
(28) |
where in the last step we have used the divergence formula (23). For the second term we compute
(29) |
where in the last step we have used the divergence formula (24). Inserting (28) and (29) into the right hand side of (27), we obtain
which finishes the proof of the invariance of .
It remains to prove the invariance of . The strategy of the proof is to show that all indices appearing in
are covariant or contravariant in the sense of Def. 3.3, so that their contraction is invariant by Lem. 3.5.
We have shown in Lem. 3.11 that the volume form is invariant. It follows from Lem. 3.6 that the index of is covariant. We have shown in Eq. (18) that the indices of and are contravariant. In Eq. (17) we have seen that the indices of are covariant. It follows from Lem. 3.10 that the indices of the covariant derivatives and are covariant. Lem. 3.4 shows that the wedge product is contravariant in all upper and covariant in all lower indices. With Lem. 3.5 we conclude that is invariant. ∎
Theorem 4.2.
The action of spacetime vector fields on the infinite jet bundle of Lorentz metrics defined in (12) has a homotopy momentum map
given by
The Noether current, which was given in (9) by the general formula , can be computed with (29) to
(30) |
The component of the homotopy momentum map, which was given in (8) by the general formula , is
Remark 4.3.
The Noether current of a symmetry is determined only up to a -closed form. Usually, the second summand of (30) is dropped, so that the Noether current is -linear in and can be interpreted as the energy-momentum tensor . Here, we must take (30) as Noether current so that is a homomorphism of -algebras.
Acknowledgements
This paper was branched out of a long and ongoing collaboration with Michele Schiavina and Alan Weinstein [BSW] on the constraint problem of general relativity. They have contributed with invaluable discussions and encouraged me to publish the homotopical approach separately. I am indebted to Yaël Frégier, Chris Rogers, and Marco Zambon for teaching me about homotopy momentum maps and for many illuminating discussions over the years. Finally, I would like to thank Janina Bernardy and Leonard Hofmann for valuable feedback on various versions of this paper.
References
- [And89] Ian M. Anderson. The variational bicomplex. Unpublished manuscript, available at https://ncatlab.org/nlab/files/AndersonVariationalBicomplex.pdf, downloaded on 2/5/2022, 1989.
- [BFLS98] G. Barnich, R. Fulp, T. Lada, and J. Stasheff. The sh Lie structure of Poisson brackets in field theory. Comm. Math. Phys., 191(3):585–601, 1998.
- [BFW13] Christian Blohmann, Marco Cezar Barbosa Fernandes, and Alan Weinstein. Groupoid symmetry and constraints in general relativity. Commun. Contemp. Math., 15(1):1250061, 25, 2013.
- [BHL10] Thomas J. Bridges, Peter E. Hydon, and Jeffrey K. Lawson. Multisymplectic structures and the variational bicomplex. Math. Proc. Cambridge Philos. Soc., 148(1):159–178, 2010.
- [BSW] Christian Blohmann, Michele Schiavina, and Alan Weinstein. A Lie-Rinehart algebra in general relativity. Preprint arxiv:2201.02883.
- [BW] Christian Blohmann and Alan Weinstein. Hamiltonian Lie algebroids. Preprint arxiv:1811.11109, 88 pages, to appear in Mem. Am. Math. Soc.
- [CFRZ16] Martin Callies, Yaël Frégier, Christopher L. Rogers, and Marco Zambon. Homotopy moment maps. Adv. Math., 303:954–1043, 2016.
- [CnCI91] J. F. Cariñena, M. Crampin, and L. A. Ibort. On the multisymplectic formalism for first order field theories. Differential Geom. Appl., 1(4):345–374, 1991.
- [Del18] Nestor Leon Delgado. Lagrangian field theories: ind/pro-approach and L-infinity algebra of local observables. PhD thesis, University of Bonn, 2018.
- [DF99] Pierre Deligne and Daniel S. Freed. Classical field theory. In Quantum fields and strings: a course for mathematicians, Vol. 1, 2 (Princeton, NJ, 1996/1997), pages 137–225. Amer. Math. Soc., Providence, RI, 1999.
- [ET79] D. B. A. Epstein and W. P. Thurston. Transformation groups and natural bundles. Proc. London Math. Soc. (3), 38(2):219–236, 1979.
- [GIM04] Mark J. Gotay, James Isenberg, and Jerrold E. Marsden. Momentum maps and classical relativistic fields. Part II: Canonical analysis of field theories. Preprint arXiv:math-ph/0411032, 2004.
- [GIMM04] Mark J. Gotay, James Isenberg, Jerrold E. Marsden, and Richard Montgomery. Momentum maps and classical relativistic fields. Part I: Covariant field theory. Preprint arXiv:physics/9801019, 2004.
- [Kru83] D. Krupka. Lepagean forms in higher order variational theory. In Proceedings of the IUTAM-ISIMM symposium on modern developments in analytical mechanics, Vol. I (Torino, 1982), volume 117, pages 197–238, 1983.
- [KS11] Yvette Kosmann-Schwarzbach. The Noether theorems. Sources and Studies in the History of Mathematics and Physical Sciences. Springer, New York, 2011. Invariance and conservation laws in the twentieth century, Translated, revised and augmented from the 2006 French edition by Bertram E. Schwarzbach.
- [Rog12] Christopher L. Rogers. -algebras from multisymplectic geometry. Lett. Math. Phys., 100(1):29–50, 2012.
- [Sou70] J.-M. Souriau. Structure des systèmes dynamiques. Dunod, Paris, 1970. Maîtrises de mathématiques.
- [Tak79] Floris Takens. A global version of the inverse problem of the calculus of variations. J. Differential Geom., 14(4):543–562 (1981), 1979.
- [Wal84] Robert M. Wald. General relativity. University of Chicago Press, Chicago, IL, 1984.
- [Zuc87] Gregg J. Zuckerman. Action principles and global geometry. In Mathematical aspects of string theory (San Diego, Calif., 1986), volume 1 of Adv. Ser. Math. Phys., pages 259–284. World Sci. Publishing, Singapore, 1987.