Deformations of Lie groupoid morphisms
Abstract.
We establish the deformation theory of Lie groupoid morphisms, describe the corresponding deformation cohomology of morphisms, and show the properties of the cohomology. We prove its invariance under isomorphisms of morphisms. Additionally, we establish stability properties of the morphisms using Moser-type arguments. Furthermore, we demonstrate the Morita invariance of the deformation cohomology and consider simultaneous deformations of the morphism, its domain, and codomain. These simultaneous deformations are utilized to define cohomology for generalized morphisms and to study deformations of multiplicative forms on Lie groupoids.
1. Introduction
Deformation theory constitutes a distinct research area in mathematics, with numerous noteworthy works emerging in connection with various branches of mathematics and physics such as algebraic geometry, quantum mechanics, complex geometry, and algebra. As defined in Kontsevich’s lectures, deformation theory is the infinitesimal study of moduli spaces. Specifically, it involves the infinitesimal study of families of structures, which we call deformations, around a given structure. This infinitesimal study yields tangent vectors to the moduli spaces, measuring the direction of each family. Typically, these vectors are elements of low degree (1 or 2) for a specific cohomology, known as the deformation cohomology of that structure. Almost every mathematical object possesses its own deformation theory. Examples include Lie algebras, Lie subalgebras, associative algebras, and algebra homomorphisms, each exhibiting a rich deformation theory [38], [21], [35]. Some works in the realm of differential geometry span topics ranging from complex manifolds, foliations, G-structures, and pseudogroup structures to Lie groups, Lie algebroids, and Lie groupoids, among others [28], [1], [22], [26], [25], [3], [2]. Notably, deformations of the latter two Lie objects enable us to address the deformations of many well-known geometric structures, including foliations, Lie groups, Lie group actions, and Poisson structures, as evidenced in [16], [15] and [17]. This is not coincidental; indeed, Lie algebroids and Lie groupoids have recently attracted significant attention for their role in codifying various geometric structures. Moreover, they boast numerous connections with physics-related topics such as symplectic foliations, Poisson structures, Dirac structures, quantization, non-commutative geometry, and more [32], [34], [5], [6], [13], [12].
In general, the philosophy of Deformation Theory asserts that deformation cohomology arises from either a DGLA structure or a -algebra structure, as exemplified in [39], [36]. In this context, their Maurer-Cartan elements correspond to structures nearby to the initial structure undergoing deformation. The study of deformations, therefore, allows us to understand the behaviour of structures around a given one.
However, obtaining the algebraic structure on the deformation complex is not always straightforward. This is exemplified well in the case of Lie groups or Lie group homomorphisms [10], where the usual Lie group cohomology is employed to study deformations, but an algebraic structure on this complex remains unknown. The deformation complex of morphisms of Lie groupoids, which we will work in this paper, also falls into this category of having (so far) an unknown algebraic structure on the complex. Nevertheless, we can still use the deformation cohomology to approach the study of nearby structures in an alternative manner. Indeed, in Section 10 we approach the problem of stability under deformations of morphisms which give conditions to understand when any smooth family of Lie groupoid morphisms represents a constant path in the corresponding moduli space. Stable morphisms under deformations are closely related to representing the isolated points on the moduli space of morphisms. For example, one can consider the compact-open topology in the space of morphisms between two Lie groupoids. If this space is locally path-connected around a fixed morphism then the stability of under deformations amounts to representing an isolated point in the moduli space. However, despite the known fact that the space of smooth maps between manifolds is locally path-connected under compactness of the domain (See [23], Theorem 1.11, p. 76), the question of whether or not the space of morphisms between two Lie groupoids is locally path-connected is rather subtle, and we leave it to be explored elsewhere. The stability under deformations is then equivalent to stating that can be deformed only in some trivial ways determined by the conjugation by either bisections or gauge-maps; we will detail these trivial ways and relations between them in Section 3. For instance, we employ the properness of Lie groupoids to verify that the stability under deformations property holds for morphisms as stated in the following result.
Theorem 1.1.
Let be any morphism of Lie groupoids. If is proper and is a transitive Lie groupoid then any deformation of is trivial.
In Section 4 we introduce the deformation complex of morphisms which, together with deformations of morphisms, are the main notions we will work in this paper. As in the statement below, we show that naturally isomorphic Lie groupoid morphisms have isomorphic deformation cohomologies. This result establishes the deformation cohomology of morphisms as a key concept within the category of Lie groupoids and equivalence classes of morphisms.
Theorem 1.2.
If and are equivalent morphisms then their deformation cohomologies and are isomorphic.
We also provide an alternative description of the complex using the theory of -groupoids and their cohomology [24]. In Section 5 we illustrate the deformation complex with examples and explore its relation to other complexes. In Section 8 we use the exactness of 1-cocycles associated to a deformation of morphisms to characterize the trivial deformations of morphisms:
Theorem 1.3.
A deformation by morphisms is trivial if and only if the family of associated cocycles is smoothly exact.
Such deformation 1-cocycles should be conceived as the velocity vectors of the deformation and their exactness as the vanishing of their velocities. Still in Section 8, we conduct a similar analysis by considering other types of relations between the morphisms that determine the respective coarse moduli spaces of morphisms. We use the associated 1-cocycles to characterize the trivial deformations in each of these moduli spaces and that arise naturally when observing the deformation complex. For instance, we consider the relation induced by the composition action of the full group of automorphisms of the codomain Lie groupoid. A morphism can then be deformed by composing it with family of such automorphisms, and the family that it produces is said to be a weakly trivial deformation.
In Section 9 we extend results from Section 8 to -deformations of morphisms. As an application, we obtain a more geometric proof of the Thom-Levine Theorem, which characterizes the triviality of -deformations of smooth maps between manifolds ([23], Theorem 3.3, p. 124).
The Thom-Levine Theorem is crucial in the study of stability of smooth maps and provides an intermediate step in establishing the equivalence between stability and infinitesimal stability of smooth maps. It plays a key role in verifying that infinitesimal stability implies stability under deformations. This, in turn, is used to prove stability under the compactness hypothesis on the domain, ensuring that any map close to a fixed one can be reached by a deformation (path) which, at each time , is locally trivial.
In Sections 6 and 7, using the -cohomology description of the deformation complex, we establish additional results to enhance our understanding of the behaviour of the deformation cohomology. These results are aplicable more generally to any -groupoid over a base . For instance, if is the complex of vector bundles over induced by (where is the core and the side bundles of ) with and then we show that there exists an exact sequence for the low-degree cohomology as follows:
Proposition 1.4.
Additionally, we verify that the previous sequence can be extended in the particular case of regular groupoids, indeed if is a regular -groupoid
Theorem 1.5.
There exists a map such that the cohomology associated to fits into a long exact sequence
This -cohomology approach gives us a way to make explicit the deformation cohomology groups of a morphism under the properness condition of its domain.
Proposition 1.6.
If is proper, then , and for every , where and are the isotropy bundle of and the normal bundle to the orbits of , respectively.
We consider more properties of the complex in Section 12 where we establish a Morita invariance of the complex. We use this invariance, combined with simultaneous deformations of morphisms, their domains, and codomains (detailed in Section 11) to derive as an application a deformation cohomology for generalized morphisms. In essence, we introduce a deformation complex for fractions between morphisms of groupoids, and verify that such a cohomology is invariant under equivalence of fractions:
Theorem 1.7.
If and are equivalent fractions, then their deformation cohomologies and are isomorphic.
Thus, this result yields a well-defined algebraic object associated to generalized morphisms or maps between differentiable stacks. The use of this complex in the study of deformations of fractions will be explored in future work. The study of deformations of Lie subgroupoids, which is explored in detail in [9], is also motivated by the work on simultaneous deformations, where the deformation complexes are obtained through the cohomological construction of the mapping-cone complex. As a final application, in Section 13 we introduce the study of deformation of multiplicative forms on Lie groupoids. There, we define the deformation complex for multiplicative forms and characterize the trivial deformations in terms of the exactness of the associated cocycles in the complex, which can be thought of as a Moser’s type Theorem for multiplicative forms. This topic is deeply explored in [8] for studying deformations of symplectic groupoids.
Acknowledgements
I would like to thank to Ivan Struchiner for many valuable discussions, important advice as well as for suggestions, and comments on initial drafts of this paper. Special thanks to Joao Nuno Mestre for valuable suggestions on the first versions of this paper. This work was also benefited by conversations with Cristian Ortiz and Matias del Hoyo, which improved the final version of this paper. This research received support from a PNPD postdoctoral fellowship at UFF.
2. Background: Deformation Theory of Lie Groupoids and VB-Groupoids
In this section we will recall some preliminary content that will be used throughout the paper. Its purpose is mostly to establish notations and to maintain this paper as self-contained as possible. For further details on Lie groupoids, Lie algebroids, deformation theory of Lie groupoids and VB-groupoids, we refer the reader to [32], [15] and [24].
Let be a Lie groupoid, denote by , , , and the source, target, multiplication, inversion and unit of , respectively. We will write and when the context is clear, and identify with the corresponding unity . For , we write meaning and .
If , there is a right translation
and analogously a left-translation , between the -fibers. We denote their differentials, respectively, by and . With this, is a right-invariant vector field if and for all . Observe that the subset of right-invariant vector fields has a -module structure given by . Further, is a Lie subalgebra with the usual Lie bracket on vector fields.
Just as a Lie group has an associated Lie algebra , Lie groupoids also can be studied infinitesimally giving rise to the notion of Lie algebroid. The Lie algebroid of is determined (like ) by the Lie algebra of right-invariant vector fields on . More precisely, is the vector bundle over , where and is viewed inside as the units. In this way, there is a map which is easily seen to be an isomorphism of -modules inducing then a Lie bracket on . The vector field is called the right-invariant vector field associated to . The vector bundle is also equipped with a vector bundle map given by the restriction of to . The map is called the anchor map of .
In other words, the Lie algebroid associated to consists of the pair together with the Lie bracket on sections of , induced from that of . With this point of view, we can abstract such a notion of Lie algebroid and to say that a vector bundle over is a Lie algebroid if there exist a vector-bundle map together with a Lie bracket on the sections of in such a way that a Leibniz rule is satisfied:
for every and . As the reader may expect, defined as above is an example of a Lie algebroid in this more abstract context. More examples can be found in [34] and [32].
2.1. Deformation theory of Lie Groupoids
The deformation theory of Lie groupoids was recently introduced in [15]. In there, the authors developed the main aspects of the theory; among other things, they exhibite the corresponding cohomology attached to deformations of Lie groupoids and use it to prove the stability of compact Lie groupoids. We recall here some key facts of the constructions in [15].
Deformations of Lie groupoids
A deformation of a manifold is roughtly understood in terms of a smooth family of manifolds. A smooth family of manifolds is viewed as a manifold together with a submersion , such that every is the fiber over . One also says that the family is smoothly parametrized by . This notion is the central idea to define deformations of Lie groupoids. Explicitly,
Definition 2.1.
(Smooth family of Lie groupoids)
A smooth family of Lie groupoids parametrized by a manifold is given by a Lie groupoid and a surjective submersion such that
In this way, determines the family of Lie groupoids , where denotes the restricted groupoid over . One says that the family is proper if is proper, i.e., if is a proper map.
Two familes and are isomorphic if there exists an isomorphism of groupoids compatible with the submersions and in the sense that . This isomorphism can be thought of as a smooth family of isomorphisms parametrized by .
Definition 2.2.
(Deformation of Lie groupoids)
Let be a Lie groupoid with structural maps A deformation of is a smooth family of Lie groupoids parametrized by an open interval containing zero,
such that . We denote the structural maps of by
The deformation of is called strict if the all fibers and are diffeomorphic to and in a smooth way, i.e., if there exist two diffeomorphisms and such that and . In other words, essentially we only deform the structural maps of : . In such a case, we can assume and the deformation is said to be -constant if does not depend on . The (strict) deformation such that as groupoids is called the constant deformation of .
Two deformations and are locally equivalent if there exist a family of isomorphisms of groupoids , smoothly parametrized by in a open interval containing zero (contained in ), such that .
Remark 2.1.
Consider two locally equivalent deformations and . For simplicity and because around the families and are isomorphic, we will just say that and are equivalent deformations of (even if ).
With the convention of the last remark, the deformation is called trivial if it is equivalent to the constant deformation.
Remark 2.2 (Fibrations).
Recall that a fibration between two Lie groupoids is a Lie groupoid morphism such that the map , , is a surjective submersion. As pointed out in [18], a deformation of also can be regarded in terms of fibrations of Lie groupoids. The data involved in the definition of a deformation of can be expressed in the form
where is a fibration of Lie groupoids. The family of Lie groupoids corresponds to the fibers of the fibration. In this sense, a strict deformation can be thought of as a fibration where the maps between the arrows and the objects are locally trivial. In fact, two trivializations and induce a family of Lie groupoid structures on the manifold . For instance, the deformation of the source map is determined by and so on.
Examples of deformations of Lie groupoids are considered in ([18], p. 16). As a manner of illustration we sketch here some of them.
Examples 2.3.
-
(1)
Let . Consider the family of Lie groups given by
Due to the fact that for the multiplication is non-abelian, this is a non-trivial deformation of .
-
(2)
Consider the family of Lie group actions of on , given by:
Thus, if then such a family of actions can be seen as a family of action groupoids . This is of course a non-trivial deformation of since the topology of the orbits varies with .
Deformation cohomology of Lie groupoids
The fundamental fact of the deformation complex of a Lie groupoid is that it governs deformations of . Concretely, to every deformation of one associates a cohomology class in , and this correspondence also shows a relation between the equivalence classes of deformations of and the classes of . The deformation complex of is defined as follows.
For any , consider the manifold of -strings of composable arrows, and define . The space of -cochains is given by
where -projectable means that does not depend on . The differential of is defined by
where , is the division map of .
For , with differential defined by
where is the left-invariant vector field on associated to defined by . Note that a section of can be viewed as a map , with such that .
This data in fact defines a cohomology () and denotes the deformation cohomology of .
In this way, one can describe explicitly the cohomology class associated to an -constant deformation of by
where denotes the multiplication of . The fact that is a cocycle is implied from applying to the associativity property of . The element is called the deformation cocycle of the deformation of . For deformations which are not necessarily -constant, a slightly different approach needs to be used yielding a non canonical 2-cocycle, however one does gets a canonical a 2-cohomology class for any deformation in the same equivalence class (see Section 5.4 in [15]).
Of remarkable importance is the transgression of the 2-cocycle ; when it exists, it plays a key role in the stability under deformations problem of Lie groupoids, as we explain below.
Moser’s argument (towards stability under deformations)
One fundamental step to study the stability question for Lie groupoids is given by the following proposition, which uses the deformation complex to state a result in the same spirit as that of the classical Moser’s theorem of symplectic geometry (see e.g. [33] p. 93).
Proposition 2.4.
[15] Let be an -constant deformation of . Consider the induced cocycles , at each time , defined in analogous way to above . Assume that for every small enough there exists such that
(1) |
and that the resulting time-dependent vector field on is smooth. Then, for and close to 0, the time-dependent flow is a locally defined morphism from to covering the time-dependent flow of on .
Additionally, if is proper, is defined if and only if and are defined.
This proposition tells us the conditions under which one finds a flow compatible with the variations of the structural maps of . However, by considering the structural maps of the total groupoid , one can express the following equivalent version of the proposition.
Proposition 2.5.
In this way, one knows that the flow of (when uniformly defined) is given by automorphisms of ([31], Prop. 3.5). And the stability under deformations question of Lie groupoids is essentially solved by finding a complete vector field like (or ) above for every deformation of ([15], Theorem. 7.1).
Analogous concepts to those described in the three steps above (deformations, cohomology and Moser’s trick) will be developed when working with the deformation theory of Lie groupoid morphisms in this paper.
We introduce now some notions of the theory of VB-groupoids which will serve us to obtain alternative descriptions and give a treatment of the deformation complex of morphisms which we will work with.
2.2. Interlude on VB-groupoids
VB-groupoids
A VB-groupoid can be thought of as a groupoid object in the category of vector bundles. They provide alternative ways to look at the representation theory and the deformation theory of Lie groupoids. For instance, the deformation complex of Lie groupoids (Subsection 2.1) can be seen as the VB-complex which is naturally associated to the cotangent groupoid when regarded as a VB-groupoid (See [15], [20] and Remark 2.11 below). This point of view will be useful in the study of the deformation complexes defined in this paper. A more detailed description of the theory of VB-groupoids and VB-complexes can be found in [32], [24] and [4].
Definition 2.3.
A -groupoid is a structure of two Lie groupoids and two vector bundles as in the diagram below
(2) |
where the vertical directions are vector bundle structures and the horizontal ones are Lie groupoids, such that the structure maps of the groupoid (source, target, identity, multiplication, inversion) are vector bundle morphisms over the corresponding structure maps of the groupoid .
Remark 2.6.
Note that the multiplication makes sense as a vector bundle morphism when one considers the induced vector bundle structure of over (guaranteed from the fact that the ‘double source map’ is a surjective submersion (appendix A in [30]).
In this setting a morphism of VB-groupoids is a morphism between the Lie groupoids and preserving the vector bundle structures, i.e, such that and are vector bundle morphisms covering the maps and , respectively. Observe that, by restricting to the zero section, turns out to be a Lie groupoid morphism.
Example 2.7.
(Tangent VB-groupoid) Given a Lie groupoid with source, target and multiplication maps , and , by applying the tangent functor one gets the tangent groupoid with structure maps , , and so on. This tangent groupoid is further a VB-groupoid over (with respect to the tangent projections).
Remark 2.8.
Note that in the previous example one has the following short exact sequences of vector bundles over ,
(3) |
and
(4) |
where and are the right and left multiplication on vectors tangent to the -fibers and -fibers of , respectively; and and are the maps induced by and with image on the corresponding pullback bundles.
Example 2.9.
(Cotangent groupoid) As noticed in [11], given a Lie groupoid its cotangent bundle inherits a groupoid structure over the dual of the Lie algebroid of ,
with source and target maps induced, respectively, from the dual of the exact sequences (3) and (4). Explicitly, for and ,
and
With multiplication determined by
for .
There are two canonical bundles over associated to a VB-groupoid , they are called the side and core bundles of . The side bundle is just the vector bundle over . The core bundle is determined by the restriction to of the kernel of the surjective map
explicitly, one defines as the restriction of to the units of . The core bundle then can be thought of as the bundle in the complementar direction to that of the side bundle. These two bundles gives rise to the short exact sequence
(5) |
which is called the core sequence of , where is the right multiplication by zero elements of on the vectors of . Observe thus that the sequence (4) above is a particular case of this sequence.
The target map of determines a vector bundle map between the core and side bundles of . This map, which is called the core-anchor map, coincides with the known anchor of when .
An splitting of the exact sequence (5) is called unitary if over the units of it coincides with the unit map of . In the particular case of the tangent groupoid (i.e., ), such an unitary splitting is called an Ehresmann connection of for the source .
These horizontal lifts of the core sequence have a relevant role in the representation theory of Lie groupoids. They allow to define quasi-actions and of on the side and core bundles of which are key elements in the notion of 2-terms representations up to homotopy of Lie groupoids [24]. Explicitly, they are given by
Remark 2.10.
Notice that an interesting fact of these quasi-actions and is that they restrict to canonical actions of on the spaces and . Explicitly, the action on is given by
In particular, in the case of the tangent VB-groupoid of a groupoid , the action is called the adjoint action on the isotropy (possibly singular) bundle of ; and the action turns out to be the so-called action of on the normal (possibly singular) bundle whose fibers are the normal spaces to the orbits of .
VB-groupoid cohomology
VB-groupoids have a special cohomology induced from their own groupoid structure which, additionally, takes into account the linear structure of the vector bundle and aims to give a geometric interpretation of the 2-terms representations up to homotopy of Lie groupoids. Such a complex, which is called the VB-groupoid complex, was defined by Gracia-Saz and Mehta in [24] and it turns out to be (canonically) isomorphic to the deformation complex of Lie groupoids when considering the cotangent groupoid (see Remark 2.11), providing another interpretation of the deformation complex of a Lie groupoid. Its definition is as follows.
Let be a VB-groupoid. The differentiable complex of (as Lie groupoid) has a natural subcomplex given by the fiberwise linear cochains of . The VB-groupoid complex of is the subcomplex of determined by the left-projectable elements of , that is, the elements satisfying the following two conditions
-
(1)
,
-
(2)
.
This VB-complex turns out to be isomorphic to the one of 2-terms representation up to homotopy of over the side and core bundles of the dual VB-groupoid, and thus allows us to think about the 2-terms representation theory in the geometric terms of VB-groupoids ([24]). In particular, it yields an interpretation of the adjoint representation of a Lie groupoid in terms of the VB-complex of the cotangent groupoid. In that way, having in mind the relation of the deformation complex of groupoids with the adjoint representation, then an expected but relevant fact of the VB-complex concerns its relation with the deformation complex of Lie groupoids:
Remark 2.11.
A straightforward computation shows that the deformation complex of a Lie groupoid is isomorphic to the VB-groupoid complex of its cotangent VB-groupoid ([20], Prop. 4.5). The isomorphism is given by , with
VB-Morita maps
A VB-Morita map takes the so important notion of Morita maps of Lie groupoids to the level of VB-groupoids. A morphism of VB-groupoids is a VB-Morita map if is a Morita morphism [20]. In that sense, VB-Morita maps are supposed to play the same role as Morita maps for Lie groupoids. For instance, in [20] the authors prove the VB-Morita invariance of the VB-cohomology. That is, if is a VB-Morita map then the VB-cohomologies and are isomorphic. In particular, since the tangent lift of a Morita morphism is a VB-Morita map then, by using Remark 2.11 and the fact that VB-Morita maps are preserved by dualization, the authors give an alternative proof of the Morita invariance of the deformation cohomology of Lie groupoids, first proven in [15]. We will also use the notion of VB-Morita morphisms to obtain a Morita invariance of the deformation cohomology of Lie groupoid morphisms. Such a result will make possible to connect the deformation cohomology of morphisms with maps of differentiable stacks (see Section 12).
3. Gauge maps and Deformation of morphisms
In this section we will define deformations of Lie groupoid morphisms. We also explain two ways in which two such deformations can be considered as equivalent deformations: by using either gauge maps or bisections. Each of these two ways turns out to be convenient depending on the examples and/or applications one has in mind, as we will see in the examples below. Let us first introduce the notion of gauge map.
A gauge map arises naturally from the concept of natural transformations between functors and allows us to relate two Lie groupoid morphisms when we regard them as functors between the groupoids. In that case, a natural transformation yields an obvious relation between two functors, and a gauge map is just an abstraction of that idea in which we define it without making reference to the functors (as it is the case of natural transformations). Formally: let be a Lie groupoid, a manifold and be a smooth map. We will call a gauge map covering . Thus indeed, given a gauge map covering and a Lie groupoid morphism with base map , then relates the morphism to the morphism defined by
Given a bisection of and , one obtains a gauge map over by considering . But, there are in general more gauge maps than those obtained from bisections of . Hence, a bisection yields another relation between two morphisms: it relates the morphism to the morphism defined by
Thus, the set of morphisms related by bisections to is smaller than the set of morphisms related to by gauge maps. These two manners of acting on a Lie groupoid morphism (either by gauge maps or bisections) are natural ways of generalizing the adjoint action of a Lie group on a Lie group homomorphism considered in [10] when working with deformations of Lie group homomorphisms.
We will consider below these two relations between morphisms at the level of deformations of morphisms: Definitions 3.2 and 3.3. However, as we will show in Proposition 3.9, under some conditions, a gauge map can be obtained by considering local bisections, and that fact can be used to prove that, at the level of deformations, such two relations can eventually be the same (see Proposition 3.11 and Theorem 3.13).
Deformations of Morphisms
Let and be two Lie groupoids and let
be a Lie groupoid morphism. Let be an open interval containing .
Definition 3.1.
A deformation of is a pair of smooth maps , and such that , , and for each the map
is a Lie groupoid morphism, where and similarly .
Remark 3.1 (Fibrations).
A deformation of can be equivalently described by a morphism between the trivial fibrations and covering the identity such that restricted to the fiber over 0 is . The corresponding base-map of between the units and will be .
In what follows we will denote a deformation of by .
We will consider smooth families of gauge maps covering smooth families of maps . By a smooth family of maps from to we mean a smooth map . Similarly, a smooth family of gauge maps over is a smooth map such that is a gauge map over .
Definition 3.2.
Two deformations and of are equivalent if there exists a smooth family of gauge maps with , where denotes the identity bisection of , and such that
(6) |
for all in some open interval containing , and all .
A deformation will be called trivial if it is gauge equivalent to the constant deformation.
Remark 3.2.
The set of gauge maps is naturally a groupoid over with structure determined pointwise by the groupoid structure of . Moreover there is a natural (left) action of this groupoid on the set of Lie groupoid morphisms from to , where acts on if the base map of is equal to , i.e., the moment map of the action is the map which associates to , its base map . The action is given by
With this notation, expression (6) becomes , for all in some open interval containing .
Remark 3.3.
We remark that expression (6) only makes sense if is a family of gauge maps over . Observe also that if we regard the maps and as smooth families of functors, then the gauge maps are simply a smooth family of natural isomorphisms between these functors.
Example 3.4 (Non-trivial Deformation).
Let be the family of morphisms given by
where we view the Lie groups and as Lie groupoids over a point. This is a nontrivial deformation because is not injective, but is injective for any .
Example 3.5.
Let be the trivial principal -bundle over . There is a one-to-one correspondence between flat principal connections on and Lie groupoid morphisms
The correspondence is obtained by using parallel transport along curves on and the canonical identification of the fibers of with , i.e., if is a path, and is such that , and is horizontal with respect to , then . Under this correspondence, gauge equivalence of morphisms translates to gauge equivalence of the connections. In particular, a deformation of a morphism is trivial if and only if the corresponding deformation of is gauge trivial.
In example 3.8 below, we deal also with flat vector bundle connections in the context of Lie groupoid morphisms.
In some situations the notion of equivalence between morphisms we have defined may be too wide. The following is an example of such situation.
Example 3.6.
Let and be pair groupoids. Any morphism from to is of the form
and this determines a one to one correspondence between morphism for pair groupoids and smooth maps from to . Note that any two morphisms are equivalent by a gauge transformation. In fact, if and are morphisms from to , then is a gauge transformation such that .
The example above suggests looking at deformations of morphisms up to stronger equivalences. An instance of that is obtained by taking bisections of groupoids (instead of gauge maps) which, in the previous example, corresponds to diffeomorphisms of the base manifold.
A smooth family of bisections of is a smooth map such that is a bisection for all . Such a family of bisections will be denoted be .
Definition 3.3.
Let be a morphism of Lie groupoids. Two deformations and of are strongly equivalent if there exist an open interval containing , and a smooth family of bisections of such that is the identity bisection, and
for all and .
In many instances, we will denote the conjugation by by , i.e.,
so that is strongly equivalent to if . A deformation will be called strongly trivial if is strongly equivalent to the constant deformation for all , or in other words, if there exists a smooth family of bisections such that , and .
Remark 3.7.
Note that if is a strongly trivial deformation of , then also can be regarded as a strongly trivial deformation of for any . Indeed, , where denotes the product in the group of bisections of , which is given by ; for and bisections of and . Notice that an analogous observation also holds for trivial deformations.
Example 3.8.
Let be a vector bundle over . There is a one-to-one correspondence between flat connections on and Lie groupoid morphisms
This correspondence is obtained by using the parallel transport along curves on . Under this correspondence, two connections and related by a gauge transformation induce strongly equivalent Lie groupoid morphisms and . Explicitly, if then , where we see as a bisection of the groupoid . We just remark that for a general bisection , however, the morphism does not correspond to some connection of due to the fact that its base map can be different from the identity , but it does corresponds to a connection of the pullback bundle .
Equivalent vs Strongly Equivalent deformations
For each a bisection of induces a gauge transformation covering . It then follows that strongly equivalent deformations are equivalent. Theorem 3.13 below will deal with the converse and more subtle statement due to the fact that not all the gauge maps are obtained from bisections of . Still, Proposition 3.9 below shows that there are some cases where the gauge map can be induced by a local bisection. Recall that a local bisection of is a local section of the source map such that is a diffeomorphism onto its image.
Proposition 3.9.
Let be a Hausdorff Lie groupoid and be a gauge map covering an embedding, i.e., such that is an embedding. Then there exists a local bisection defined on a neighbourhood of such that if, and only if, is also an embedding.
To prove the proposition we will use an Ehresmann connection on which is (also) transversal to the -fibers, that is the reason of assuming to be Hausdorff since, under non-Hausdorffness, an Ehresmann connection may not exist (see Example 13.93 in [14]). The existence of such a and -transversal connection is the content of Lemma 3.10 which we state next.
Lemma 3.10.
Let be the source and target maps of a Lie groupoid . If admits an Ehresmann connection, then there exists an Ehresmann connection for the source which is also an Ehresmann connection for the target .
Proof.
The strategy to obtain this connection will be to vary locally an Ehresmann connection of in such a way that the intersections with the tangent spaces to the -fibers are , and then glue these local connections together by using a partition of the unity.
Let denote by the intersection of with the tangent space at to the -fibers. Then, for every , , where means the orthogonal to inside for some Riemannian metric on such that is the orthogonal space to the -fibers.
Recall that an Ehresmann connection can be equivalently described by a morphism of vector bundles which is a left-inverse of the right-multiplication . This map is regarded as the projection of the tangent vectors of to the -vertical bundle . The horizontal distribution corresponds with the kernel of the morphism .
Thus, for any let be an open coordinate neighbourhood around such that the -fiber through corresponds to a slice on . We will define a new (local) connection on by modifying the values of at the part to be non-zero and extending, by translation, to the other points of . In that way, by shrinking if needed, will be the local connection on which is complementar to the -fibers and -fibers.
Explicitly, if then we choose small enough such that, for all points in , . Otherwise, if for some , then, in order to modify , we just need to do it in a linear way. For that it suffices to take basis for and and make an injective correspondence of the basic elements of with the elements of the basis of . We then define by extending to the modified by translation, say, as a constant-coefficient linear map on the coordinate neighborhood . Shrinking if necessary it follows that the intersection is trivial for any .
Thus, consider the open cover of given by the sets , and let be a partition of 1 subordinated to , therefore
determines an Ehresmann connection on for the source map which is also complementar to the -vertical bundle. That is, is an Ehresmann connection for the source and the target maps of .
∎
Proof of Proposition 3.9.
Let be a gauge transformation such that is an embedding. It is clear that if for some local bisection of , then is an embedding.
For the converse statement, we first use the horizontal curves of an Ehresmann connection as in the previous Lemma to construct a local section of the source map which contains the image of the gauge map. Then we check that it is also a local bisection of .
Use the Ehresmann connection to make the source map a Riemannian submersion where the horizontal spaces correspond to the orthogonal spaces to the -fibers. Then, by using the horizontal geodesics starting at , we obtain a submanifold such that its projection by the source map yields a tubular neighborhood of . This fact defines uniquely the local section by making the correspondence of curves under the -projection.
Moreover, the local section is in fact a local bisection of : since the horizontal spaces are also complementar to the -fibers then, by shrinking the radius of if necessary, the submanifold -projects injectively to a tubular neighborhood of in . Therefore, is a local diffeormorphism of . ∎
Hence, in some cases the gauge maps are induced by (local) bisections. We can then wonder if that also happens when considering a smooth family of gauge maps as in the case of trivial deformations. It turns out that, under a compactness condition we can avoid the Hausdorffness of and check that there exists a smoooth family of (global) bisections inducing the gauge maps. That is the content of the Proposition below.
Proposition 3.11.
Let and be Lie groupoids over and , respectively, and be a morphism covering an injective immersion . Assume that is compact. Then, any trivial deformation of is indeed strongly trivial.
Proof.
Let be the smooth family of gauge maps associated to the trivial deformation . We will prove that there exists a smooth family of bisections of which induces the gauge maps in the sense that .
We will do that by considering the smooth family of local sections of induced by the family , and then extending them to global sections to obtain the global bisections we want by using the exponential map.
On the one hand, let denote the deformation , as in Remark (3.1), covering the map . Then, after shrinking if necessary, the map is an injective imersion, and also an embedding due to the fact that each is an embedding. Moreover, the compactness of also implies that is closed inside .
On the other hand, define the local sections , induced by the family of gauge maps , by
(7) |
Notice that this family of sections can be regarded as the smooth section of the pullback by of the Lie algebroid given by . Thus, since the map is injective, the section can be seen as a section of the restriction Lie algebroid . Therefore, we can extend the section to a section of all which we extend in such a way that its support is contained in an open subset , where is an open subset containing with compact closure .
This extended section has the form , with extending . Let denote the bisection of induced by the exponential flow of the section . Such a bisection amounts to having a smooth family of bisections of which can be alternatively obtained by using the flow of the time-dependent vector field on as follows. Let denote the time-dependent flow of . Due to the compactness of and the vanishing of the sections outside , the flow is defined on all for small enough. Thus, the family of bisections of is given by , for .
In that way, the curve is an integral curve of the time-dependent vector field starting at . Also, by equation (7) the curve is an integral curve starting at the same point. Hence the family induces the family of gauge maps; that is,
Therefore, the strong triviality of the deformation follows. ∎
Remark 3.12.
In the previous proof, notices that the family of bisections can be different depending on the choice of the open set and on the extension of . However, if is another such an extension with induced family of bisections , then these two families of bisections agree when restricted to , which is the key fact for the proof. That follows from the last equation in the proof where one gets
The next theorem tells us that under certain conditions, the relations of equivalence and strongly equivalence of deformations are the same.
Theorem 3.13.
Let and be Lie groupoids over and , respectively, and be a Lie groupoid morphism covering an injective immersion . Assume that is compact and is Hausdorff. Then, any two deformations of are equivalent if, and only if, are strongly equivalent.
Proof.
Since a bisection induce a gauge map by composition with the base map, then two strongly equivalent deformations are clearly equivalent. For the converse question, assume that and are two equivalent deformations of . Then,
(8) |
where is a smooth family of gauge maps covering the family .
Let denote by the morphisms covering and which encode the deformations and , respectively. That is, one has and . Then, the equation (8) can also be expressed by
(9) |
where is the gauge map covering the map . Observe then that, shrinking if necessary, since and are embedding maps, it follows from Proposition 3.9 that there exists a local bisection of the groupoid such that .
Next step now consists in obtaining a global bisection inducing the gauge map . Let be the smooth family of local bisections of induced by . We will obtain the global bisection by considering the infinitesimal side of the local bisections and then using the exponential map. Indeed, let and consider the vectors tangent to
These vectors are not necessarily tangent to the -fibers of , for that we use an Ehresmann connection on to project the vectors to the -fibers. Let define by the family of sections obtained by right translation of the projection to the -fibers of the vectors , for all . Thus, since is an embedding map (shrinking if necessary), it follows that the family can be regarded as a section over the closed submanifold of the Lie algebroid of .
Therefore, we can extend the section to a global section of . Then finally, by using the exponential flow of , as in the proof of Proposition 3.11, we obtain a smooth family of bisections of which induces the family of gauge maps . ∎
4. Deformation complex of a morphism and properties
In this section we introduce the deformation complex of morphisms and study some of its properties.
Deformation complex of morphisms
The deformation complex of a morphism was briefly discussed in [15], it turns out to be the appropriate complex to deal with trivial deformations of morphisms (Section 8). We will first recall its definition here. Let and be two Lie groupoids and be a morphism between them.
For any , consider the manifold of -strings of composable arrows of with . The space of -cochains is given by
where -projectable means that the -projection of , , does not depend on . The differential of is defined by
where , is the division map of .
For , and the differential is
where and .
The fact that follows in a similar way to the proof that in the deformation complex of Lie groupoids; so in fact defines a cohomology and denotes the deformation cohomology of . Observe that .
Remark 4.1.
Note that, given a morphism , there are two natural cochain maps between the deformation complexes (of Lie groupoids and morphisms):
defined by
Observe that in the case the cochain map is just the pullback of sections (recall that is the induced map on the units). Similarly, if is another Lie groupoid morphism, one can define a cochain-map .
Remark 4.2.
As a special case, if above is bijective, there is an inverse of ,
Similarly, if is the map between deformation complexes of morphisms (see Remark 4.1), then also admits an inverse map analogously defined.
The deformation complex has a canonical subcomplex which controls deformations of that keeps the base map fixed (see remark 4.3). It is defined by,
(10) |
It is straightforward to check that preserves , so is indeed a complex.
Remark 4.3.
Any deformation which keeps the base map fixed determines a 1-cocycle
which explains the choice of the name for this complex. This fact will be fully detailed for the more general case of arbitrary deformations of morphisms in Section 8.
Example 4.4.
By using right multiplication to translate the tangent vectors over the unit elements of , the complex takes a simpler description as the differentiable complex with values in the pullback by of the (maybe singular) isotropy bundle of . The action of on is given by the pullback by of the adjoint action of on , as defined in remark 2.10. We observe that even in the case of a singular isotropy bundle one can still make sense of the differentable complex by defining smooth functions with values in as those smooth functions with image in which take values in .
4.1. Alternative description of the deformation complex
The deformation complex of morphisms can be expressed in a different way by using the theory of VB-groupoids. This interpretation will be useful to get a better understanding of the properties of this cohomology; in particular, it will allow us to obtain the vanishing results in a simple way. Note first that, if is a VB-groupoid over with source and target and and with core-bundle then, as pointed out in [24], the VB-complex of can be identified with a new complex which we denote by , where and
(11) |
for ; and whose differential is the one induced by the differential of under the identification. Here, one says that is -projectable if does not depend on .
In fact, the identification is given by
(12) | |||||||
where is in the fiber over .
With this at hand, it is straightforward to check that the deformation complex of a morphism is the complex , where is the pullback by of the tangent VB-groupoid . Thus we have,
Proposition 4.5.
The map (12) above induces an isomorphism between the complexes and . Moreover, after the choice of a connection on , is isomorphic to , which is the complex with the structure of pullback adjoint representation (up to homotopy) of induced by . So, in particular, .
Sketch of proof. The proof of the second claim follows from the interpretation of VB-complexes as 2-terms representations up to homotopy of groupoids after the choice of a unitary splitting of the core sequence (5) (see [24]). In particular, under this philosophy, the VB-complex of induces on the structure of the adjoint representation of ; in this case an unitary splitting of the core sequence of amounts to a connection on . Similarly, by using the connection , one can see the VB-complex as the pullback by of the adjoint representation of . Thus, by the isomorphism (12) above, the deformation complex of morphisms is isomorphic to the complex representing the pullback of the adjoint representation of .
Remark 4.6.
One can check that a proof of the second claim in the previous Proposition can also be made by some explicit computations after the choice of an Ehresmann connection of , in the same way as the proof of Lemma 9.1 in [15].
4.2. Properties and Variation of the complex
According to the terminology of equivalence of deformations of morphisms we can also say that two morphisms of Lie groupoids are equivalent if they are isomorphic when viewed as functors. The following result then tell us that the deformation cohomology of morphisms is an object associated to the equivalence classes of morphisms.
Theorem 4.8.
Let , be two Lie groupoid morphisms. If and are equivalent then their deformation cohomologies and are isomorphic.
For the proof of this Theorem we first recall the groupoid of arrows of , where the objects of are the arrows of and the arrows of are given by commutative squares of arrows in or, in other words, by three arrows with common sources.
where the source, target and multiplication maps are respectively given by the right and left arrows of the commutative square, and by the concatenation of the diagramas. That is, , and . With these notations, it is straightforward to see that the maps which take the upper and lower arrows of the square are Morita morphisms. Explicitly,
Lemma 4.9.
Let and denote the maps given by
which take the upper and lower arrows of a commutative square in . Then, the maps and are Morita morphisms covering the source and target maps of , respectively.
Proof of Theorem 4.8.
Let be the gauge map which relates the morphism to .
Notice that the gauge map can be seen as the Lie groupoid morphism
Such a morphism encodes the isomorphic morphisms and by composing with the upper and lower morphisms and of , indeed, and .
We will now prove that the cohomologies and are isomorphic to . Indeed, these isomorphisms follow from observing that if is any morphism, then a Morita map induces the quasi-isomorphism . This is the content of Proposition 12.2 in Section 12 where we prove the isomorphism between deformation cohomologies by using the notion of VB-Morita maps. Therefore, this fact along with the previous Lemma tell us that the upper and lower morphisms induce isomorphisms in the cohomologies, as we wanted. ∎
We describe now a variation of the deformation complex of morphisms which should be though of as the tangent space to the Moduli space of morphisms with the relation induced by bisections. Indeed, such a variation will be relevant for us in order to deal with the characterization of strongly trivial deformations. It just consists of changing the space of sections of the pullback algebroid by the space of pullback sections . We denote this complex by . Explicitly,
and
Observe that the cohomology is larger than the usual deformation cohomology . This fact is not a surprise in view that the space of trivial deformations is larger than the space of strongly trivial deformations.
5. Examples
Representations of Lie groupoids and flat connections can be viewed as morphisms of Lie groupoids. We study below their deformation cohomologies when regarded as morphisms.
Recall that a representation of a Lie groupoid on a vector bundle is a Lie groupoid morphism , covering the identity, from to the General Linear groupoid. One can naturally consider deformations of representations of Lie groupoids as a special instance of deformations of morphisms of Lie groupoids. We define a deformation of a representation of a Lie groupoid on a vector bundle as a deformation of the morphism which keeps the base map fixed (which is the identity). Therefore, the complex (see expression (10)) should control the deformations of the representation .
The representation induces a canonical action of on the vector bundle of endomorphisms of as follows:
(13) |
where is an element of and denotes an element in the fiber of over . Thus, since the complex can be described in terms of the isotropy bundle of and , it follows that the complex controlling the deformations of the representation of is , where the differential is induced from the canonical action (13) defined above. By other side, one can check that the deformation cohomology agrees with such a .
Proposition 5.1 (Representations of Lie groupoids).
Let be a representation of on the vector bundle . Then
Proof.
This isomorphism in cohomology can be checked directly as the induced by the right translation after notice two things: (i) that is the isotropy bundle of Lie algebras of the Lie algebroid and (ii) that the action of on defined above agrees with the canonical adjoint action described in Remark 2.10. In Section 7 we give an alternative way to check this isomorphism (see example 7.3).
∎
The previous Proposition tell us then that the usual deformation cohomology of is the one that controls deformations of the representation of .
Remark 5.2.
Similarly one can see that if the target groupoid, for any morphism , is transitive then all the information of a deformation of is concentrated just on the cohomology of with values in the isotropy bundle of . In some sense that means that the non-trivial deformations of will be determined by the isotropy directions of .
Remark 5.3.
In the special case of a representation of on , we can get an alternative view-point in terms of 1-jets: the groupoid can be though as the Lie groupoid of 1-jets of local diffeomorphisms of , and its Lie algebroid corresponds to the set of 1-jets of vector fields on . The vector bundle translates to the set of 1-jets of vector fields on its singular points. The action of on will be obtained by pulling back the canonical action of on given by
Example 5.4 (Connections on vector bundles).
Example 3.8 allows us to see any flat connection on a vector bundle as a Lie groupoid morphism covering the identity map . Equivalently, is regarded as a representation of .
Thus, we define a deformation of by flat connections as a deformation of the associated morphism which keeps the base map fixed (which is the identity).
Hence, by Remark 5.2, the deformation cohomology controlling deformations of by flat connections should be
In particular, when is a simply connected manifold, since will be proper, it follows that
In that case, the Theorem 8.5 tells us that the flat connections can be deformed just in a trivial manner.
Example 5.5 (Flat principal connections).
Example 3.5 regards any flat principal connection on the trivial -principal bundle over as a Lie groupoid morphism uniquely.
Therefore, by Remark 5.2, the cohomology controlling deformations of by flat principal connections is
where is the trivial vector bundle with fiber over .
Example 5.6 (Morse Lie groupoid morphisms).
The notion of Morse-Lie groupoid morphism has been defined recently in [37] as a morphism towards the unital groupoid over such that every critical orbit of is non-degenerate. Indeed, the critical points of a Lie groupoid morphism towards the unital groupoid come in saturated subspaces, in that sense the non-degeneracy of a critical orbit is determined in terms of the non-degeneracy of the normal Hessian of . Even though, a Morse-Lie groupoid morphism is in particular a morphism towards the unital groupoid over , which is a regular groupoid with no isotropy . Therefore its deformation cohomology groups can be computed according the following sequence induced from (20) for every
The degree zero cohomology vanishes and is computed according Example 6.3. More explicitly, the cohomology groups are given in terms of the differentiable cohomology
Therefore, it turns out that the cohomology does not depend on the morphism but only on the groupoid .
6. Low degrees and vanishing
In this section we will describe the deformation cohomology groups in low degrees. A central point here will be the description of the deformation complex in terms of VB-groupoids. The content of this section will be key to show the stability properties of morphisms.
Let be a VB-groupoid over , with core and core-anchor map . We consider the following (possibly singular) vector bundles
over . It is a known fact that, even though and may be singular, induces canonical actions of on them (see Section 2.2 or [24]). In particular they turn out to be actual representations if is a regular VB-groupoid (see Section 7). Nevertheless, we can make sense of the cohomologies with values in and in the singular case, as we show below. We are mainly interested in the low degree cohomologies. Define the ‘smooth sections’ of and by
and
where we are looking at as the induced map on sections .
Remark 6.1.
Note that the definition of as the subspace of sections in with values in has the direct generalization to define as the subspace of sections in which take values in ; where is given by . Also, the canonical action of on allows us define a differential with the same formula as that of the differentiable cohomology of with values in a representation. Such a differential makes a cochain complex, whose cohomology we denote by . Moreover, this complex with values in can be viewed as a subcomplex of the VB-complex through the right multiplication by zero elements:
(14) |
Following the idea of Lemma 4.5 and Definition 4.6 of [15], we can make sense of as being the invariant sections of .
Definition 6.1.
Let . We say that is invariant if there exists a section which is both -projectable and -projectable to . In other words, is an -lift of . We denote the space of invariant elements by
Observe that when and are vector bundles, the previous definitions agree with the usual ones of cochains with values in a representation of . With the general setting above, we obtain the following two propositions related to the low degree cohomologies.
Proposition 6.2.
If is a VB-groupoid over one has .
Proof.
It is clear that the differential of on a 0-cochain is given by . Thus, since then is a 0-cocycle if and only if and . That is, if and only if . ∎
Thus, in particular we obtain,
Example 6.3.
For a morphism of Lie groupoids , .
Proposition 6.4.
Let be a VB-groupoid as above. Then we have the following exact sequence
(15) |
where the maps and are determined as follows:
the map is induced by the cochain map (denoted again by) defined in (14); is induced by the -projection of the elements of to the sections of the side bundle; and takes an invariant element to , where is any -lift of .
Proof.
As particular cases of the previous sequence we obtain some key sequences in the context of deformation cohomologies.
Examples 6.5.
-
(1)
It is straightforward to see that if then the sequence above reproduces the sequence in Proposition 4.11 of [15] for the deformation complex of Lie groupoids.
-
(2)
Let be a morphism of Lie groupoids. By taking the VB-groupoid we obtain a sequence for the deformation complex of the morphism :
(16) where is the isotropy bundle of and is the normal bundle to the orbits of .
Vanishing of cohomologies
Here we state the vanishing results for the deformation cohomology. The proofs are straightforward applications of the VB-groupoid interpretation of the deformation complex, the vanishing result of the VB-cohomology [7] and of the sequence (16).
Proposition 6.6.
Let be a morphism of Lie groupoids. If is proper, then , and for every , where is the normal bundle to the orbits of .
Remark 6.7.
Alternatively, a direct proof of the vanishing of cohomology for morphisms also can be made in an analogous way to that of the vanishing of the deformation cohomology of proper Lie groupoids in [15].
7. Regular setting
In this section we show that the sequence (15) is just part of a long exact sequence when we impose some regularity conditions on the groupoids. Later we illustrate the long exact sequences which can be deduced from it. We say that a VB-groupoid is regular if its core-anchor map has constant rank.
Theorem 7.1.
Let be a Lie groupoid and be a regular VB-groupoid over . Then there exists a map such that the cohomology associated to fits into the long exact sequence
(17) |
where and are the maps induced by the right multiplication by zero elements of and the -projection of elements in .
Proof.
The proof of this theorem is an adaptation of the one of Proposition 8.1 in [15] for the deformation cohomology of regular Lie groupoids. Note that the regularity condition on the VB-groupoid tells us that the associated cohomology induced by the complex associated to is given by the cohomology bundles and over , where and are the core and side bundles of . With this setting we construct the complexes and which fit into the following exact sequences
(18) |
(19) |
where is acyclic. Namely, is defined by
with differential , where is expressed by
It is straightforward to check that is acyclic: it is the mapping cone of the identity .
The map is given by
where denotes the class of in and ‘’ the action of induced by on . This map is compatible with the differentials. The complex is taken as being the kernel of , and takes a cochain to the pair
where and are the and -projection of , respectively. Observe that the definition of the quasi-action of on , induced by , implies that in fact takes values in and hence . Again, is compatible with the differentials. By choosing a splitting of the core-sequence
it is possible to show, analogous to the proof of the Proposition 8.1 in [15], that . It is also clear that . The surjectivity of follows from the surjectivity of the projection ; in particular, if with then .
In that way, the long exact sequence induced by the sequence (18) is exactly the long sequence to be proved up to an isomorphism
induced by the sequence (19). Hence to complete the proof it suffices to show that in cohomology. Indeed, take a cocycle in , thus . Therefore, ; that is, represents de cohomology class . On the other hand, since is a cocycle, then , which implies . Hence, represents the cohomology class , which completes the proof. ∎
Remark 7.2.
The a priori arbitrary complex defined in the previous proof is a key element of the argument. Nevertheless one can give a more geometric meaning of it in terms of VB-groupoids as follows. Consider the anchor morphism
between VB-groupoids. Then, the complex is just the complex associated to the pullback VB-groupoid over . And thus the acyclicity of follows directly from the vanishing of the VB-cohomology of the proper VB-groupoid and from the acyclicity of its associated 2-term complex .
We can thus use the previous theorem and obtain a long exact sequence for the deformation cohomology.
Example 7.3.
Let be a morphism of Lie groupoids and assume that is regular. Then there exists a map such that fits into the long exact sequence:
(20) |
where is the normal bundle to the orbits of .
8. Triviality of deformations of morphisms
We discuss here characterizations of the several types of triviality of deformations of morphisms in terms of the deformation cohomology. The main results in this section can be regarded as a Moser type theorem in the context of morphisms of Lie groupoids.
Proposition 8.1.
Let be a deformation of . Then, for each we obtain a 1-cocycle
in the deformation complex of . Moreover, the corresponding cohomology class at time in depends only on the equivalence class of the deformation.
Proof.
The first part follows from taking derivative at of the morphism condition satisfied by every . In fact, we get
which says that is a 1-cocycle in .
Now suppose that is a deformation of which is equivalent to . Then for a smooth family of gauge maps over with . Denote by the associated 1-cocycle at time zero. Heuristically, the exactness of comes from taking derivatives at of the equivalence condition . However, notice that since acts merely on the elements of then we can not use directly the chain rule to differentiate the expression with respect to . Therefore, consider the maps , , and , , which codify the families of gauge maps and morphisms . Thus the map contains all the information of the expression we want to differentiate. In order to differentiate this map, we will write it now in an equivalent way. Indeed, let be the Lie groupoid which is the cartesian product of with , , and denote the maps
for . Thus, we get that
Therefore, we get the derivative with respect to by applying the differential of this map to the vector field . That is,
which, after a straightforward computation (see Theorem 1.4.14 in [32]), can be written as
(21) |
where and , and are local bisections of through , and .
Therefore, for we can choose , and the previous equation becomes
(22) |
where is given by
That is, and are in the same cohomology class, as we wanted to prove. ∎
The 1-cocycle is also called the infinitesimal deformation associated to the deformation , and will be called the deformation cocycle.
Remark 8.2.
Notice that if is a trivial deformation of , then equation (21) shows that every 1-cocycle is exact. Alternatively, by using the rule chain, this can be checked by a direct computation as below.
Define the family of sections by
(23) |
Therefore we get the exactness of the cocycles,
Remark 8.3.
Since two deformations which are strongly equivalent are, in particular, equivalent then they determine the same cohomology class in . Moreover, in a totally analogous way, one can prove that they determine the same cohomology class in the variation cohomology . In fact, it suffices considering a smooth family of bisections of instead of the family of gauge maps of the previous proof and make . Furthermore, for a smooth family of bisections one can define the family of sections of
for each , where , and obtain, for a strongly trivial deformation, that
We remark this result in the following proposition.
Proposition 8.4.
Let be a deformation of . Then, the corresponding cohomology class at time in depends only on the strong equivalence class of the deformation. Moreover, a strongly trivial deformation has exact cocycles in the variation complex .
In order to prove the triviality of a deformation of by means of a Moser type argument, we will need not only that the cohomology class of the deformation cocycle vanishes, but that it vanishes in a smooth manner as we now explain.
Recall that a deformation of is a smooth map
Definition 8.1.
A family of cocycles is smoothly exact if there exists a smooth section such that for each , and
A family will be smooth if it can be encoded in a smooth section as above.
Equivalently, defining the morphism , which is the projection to of , the family is smooth if the section given by is smooth.
Theorem 8.5.
Let be a deformation of the morphism . Then, is trivial if and only if the family of cocycles is smoothly exact in .
Remark 8.6.
The smooth exactness condition of the family is just another way to say that each 1-cocycle is equal to , where is a smooth family of 0-cochains in the sense of Definition 8.1.
Proof.
The smooth exactness of the cocycles was already verified in Remark 8.2. We prove now the converse statement, where the goal is finding a smooth family of gauge-maps which verifies the triviality of the deformation . Assume that, for every , , for such that is a smooth section in (see Definition 8.1 for the notations). We will define the family of gauge maps in terms of the flow of an appropriate vector field determined by the sections .
Consider the vector field
(24) |
defined on the fibered-product of the target map with the base map , as in the diagram below,
(25) |
This vector field is indeed well-defined since , and therefore .
Lemma 8.7 below guarantees that the flow of is defined for all over the points of Thus defines the family of gauge maps (which projects to )
where is the natural projection to as in diagram (25) above. That is, the flow of over can be written as .
Claim: The family proves the triviality of , i.e., it holds .
We will prove this claim by showing that and determine the same integral curves of a vector field defined below. For that, notice first that since , then the curve belongs to a unique orbit of . Thus, for , the curve lies inside the restriction groupoid . That is, for any , both curves and are inside the restriction ; where . Therefore, we can consider the following fibered product
(26) |
which, by transversality of the maps involved in the diagram, will be well-defined. Let denote the vector field on defined by
Then, for every , and . It follows that and are integral curves of starting at the same point. Therefore, , for every . Since is arbitrary, it follows that .
∎
Lemma 8.7.
Let be the vector field defined by equation (24) above. The flow of is defined for all over
Proof.
Note that the vector field projects by the target map to a vector field given by
It follows from the fact that , that the integral curves of are determined by the smooth family of base maps . Therefore they are defined for all when starting at points of . Thus, the proof now follows by an argument completely analogous to that of Theorem 3.6.4 in [32], which allows us to check that the flow over is defined for the same time as the flow of ; i.e., for all .
∎
Remark 8.8.
It is straightforward to see that the previous result generalizes that of [10] concerning the triviality of deformations of Lie group homomorphisms.
Notice that one can use Theorem 8.5 and Proposition 3.11 to obtain as a direct consequence the following kind of characterization of strongly trivial deformations.
Theorem 8.9.
Let be a Lie groupoid morphism and be a deformation of . Assume that is an injective immersion and is compact. Then, the deformation is strongly trivial if and only if the family of 1-cocycles is smoothly exact in .
The following theorem shows that with the help of the subcomplex is posible to obtain a cleaner characterization of the strongly trivial deformations.
Theorem 8.10.
Let be a Lie groupoid morphism and be a deformation of . Assume that is compact. Then, the deformation is strongly trivial if and only if the family of 1-cocycles is smoothly exact in the subcomplex .
Proof.
If is strongly trivial, the smooth exactness of the family of 1-cocycles was already proved in Remark 8.3. Conversely, let , where is a family of sections of , such that is a smooth family of sections in the sense of Definition 8.1.
Since is compact, is a closed subset inside . Shrinking the interval if necessary, let be open subsets with compact closure such that . Then we can extend the smooth family of restriction sections to a smooth family of sections supported on such that .
Then, on the one hand, since over it follows that every cocycle is the pullback by of the vector field . Indeed,
(27) |
On the other hand, let be the smooth family of bisections of induced, by the exponential flow, by the family of sections as in the proof of Proposition 3.11. Note that such a family is defined for all small enough. Thus, by Remark 8.3,
(28) |
Therefore, in other words, by equations (27) and (28) one has that and are integral curves of the time-dependent vector field passing through at time . That is, for all small enough.
∎
The triviality of the following special type of deformations is not hard to prove directly without using the cohomological tools of this section, nevertheless we will verify it as a consequence of the Theorem 8.5 above.
Example 8.11.
Let be a surjective submersion. Assume that is a Lie groupoid morphism between and the submersion groupoid and that is a deformation of . Then, is trivial if and only if .
In fact, on the one hand, notice first that, since the Lie algebroid of the submersion groupoid consists of the vertical vectors of , then the family of elements given by is a family of 0-cochains in if and only if for all .
On the other hand, it is straightforward to check that every morphism is of the form . Therefore, if for every and all , and are in the same -fiber, then the family of 1-cocycles is smoothly transgressed by the family of 0-cochains . And conversely, if is smoothly transgressed, then must be a vector tangent to the -fibers, for all .
The previous example tells us that the family is trivial if and only if it preserves the correspondence, determined by , between the leaves of the foliation by orbit-groupoids of and the leaves of . That is, if for every , the curve lies inside the restriction groupoid . The following proposition explores this idea for any Lie groupoid (not only the submersion groupoid).
Proposition 8.12.
Let be a deformation of and assume that is proper. The deformation is trivial if, and only if, the curves , , determined by lie inside the leaves of the foliation by orbit-groupoids of .
Proof.
If is a trivial deformation, then obviously the curve lies inside a unique leaf because , for a smooth family of gauge maps covering . Conversely, let be the morphism such that restricted to is , as in Definition 8.1. Then, the family of 1-cocycles can be encoded in a unique 1-cocycle given by
And, turns out to be exact if and only if the family is smoothly exact.
Now, since is a proper groupoid, it follows that is trivial and the sequence (16) above for the morphism becomes
(29) |
Moreover, due to the fact that is always tangent to the orbit groupoids of then lies in the kernel of . Therefore is an exact cocycle which, by Theorem 8.5, says that is a trivial deformation. ∎
We can also apply our methods to study deformations which are trivial up to automorphisms of , that is, those which consider also the group of outer automorphisms of a Lie groupoid. Additionally, in Subsection 8.1 below, we sketch other types of equivalences between deformations of morphisms which arise naturally.
Definition 8.2.
A deformation of a Lie groupoid morphism is said to be trivial up to automorphisms of if there exist an open interval containing 0 and smooth families of Lie groupoid automorphisms and of gauge maps over such that , and for all . Analogously we say that is strongly trivial up to automorphisms of if there exists a smooth family of bisections of such that and for all .
Remark 8.13.
Note that considering only outer automorphisms produces weakly trivial deformations whose study is already include in the strongly trivial up to automorphisms deformations.
Recall that a Lie groupoid morphism induces a pull-back map . The key to characterizing the previous types of deformations lies in studying the pre-image of the deformation cocycle of a deformation through the pull-back map , when it exists.
Definition 8.3.
We will say that a family has a smooth pre-image in if there exist smooth families of cocycles and of cochains such that
Analogously, considering the variated cohomology, we say that has a smooth pre-image in if the sections are of the form , for .
The statements of the following two theorems concerning the two types of deformations in definition 8.2 are analogous to the statements of the Theorems 8.9 and 8.5 concerning strongly trivial and trivial deformations.
Theorem 8.14.
Let be a deformation of the morphism . Assume that is compact. Then, is trivial up to automorphisms of if and only if has a smooth preimage by in for all , where is some interval containing the zero.
Proof.
Let be a gauge trivial deformation up to automorphisms of , that is, , for smooth families of automorphisms of , with , and of gauge maps with base such that . By applying to both sides of the equation we obtain
where and is a smooth family of 1-cocycles in . It follows that for all .
Conversely, assume that
(30) |
for and smooth families of elements in and , respectively. Let be the flow from time to of the time dependent vector field on . Recall that every will be an automorphism of . We claim that is a trivial deformation of . That is, we will obtain a smooth family of gauge maps with base , starting at , such that for small enough, concluding the proof.
On the one hand, taking to be such that , equation (30) becomes
(31) |
On the other hand, we set to be the family of deformation cocycles associated to the deformation of . We will check that , i.e., is smoothly exact. In fact,
from where it follows that
where the last equality follows from equation (31). It then follows from Theorem 8.5 that is trivial concluding the proof of the theorem. ∎
Analogously, adding the condition of to be an injective inmersion, the previous proof and Theorem 8.9 prove the following.
Theorem 8.15.
Let be a deformation of the morphism . Assume that is an injective immersion and that and are compact. Then, is strongly trivial up to automorphisms of if and only if the family of cohomology classes has a smooth preimage by in .
Remark 8.16.
In particular, if the family of the previous Theorem is smoothly exact, then Theorem 8.9 shows that the family of automorphisms of can be taken as a family of inner automorphisms.
With a very similar proof to that of Theorem 8.14, we can consider the variated complex and prove
Theorem 8.17.
Let be a deformation of the morphism . Assume that is compact. Then, is strongly trivial up to automorphisms of if and only if has a smooth preimage by in for all , where is some interval containing the zero.
As a final result of this subsection we sketch an alternative characterization of these deformations under the weaker condition of smooth exactness on the cokernel complexes of the pullback maps . Given a deformation , let denote the image of the cocycles in the cokernel complex. The smooth exactness of the cocycles is defined following the philosophy that all the elements involved in the transgression of the family form smooth families.
Theorem 8.18.
Let be a deformation of the morphism . Assume that is a surjective submersion and that and are compact and connected. Then, the deformation is trivial up to automorphisms of if and only if the family of cocycles in is smoothly exact.
Sketch of proof.
The exactness condition amounts to have
(32) |
for and smooth families of elements in and , respectively. Let denote the flow from time to of the time dependent vector field on . A priori, every will be a diffeomorphism of . However we claim that every is a morphism for all small enough.
Indeed, in order to prove the claim note that by applying to equation (32) we get
which evaluating in arrows shows us that is multiplicative on the image of . Thus, since the morphism is assumed to be a surjective submersion then, by the compactness of and the connectedness of the groupoids, will be a surjective submersion for all small enough. That is, every is multiplicative on . Therefore one can follows now the proof of Theorem 8.14 to check that the deformation is trivial up to automorphisms of . The converse statement follows as in Theorem 8.14.
∎
8.1. Additional remarks on triviality
The previous results take account of four notions of equivalences defined for deformations of morphisms. However, there are some other very natural types of deformations that can be considered. Here we briefly explain these notions and show how they fit in the framework of deformation complexes.
Definition 8.4.
We say that a deformation of is strongly trivial up to automorphisms of or strongly trivial up to automorphisms on the left if there exist an open interval around 0 and a smooth family of automorphisms of with and a smooth family of bisections of with such that , for all . Analogously a deformation is called trivial up to automorphisms of if there exists a smooth family of automorphisms of with and a smooth family of gauge maps over such that and , for all .
Remark 8.19.
Deformations which are strongly trivial up to automorphisms on left are very much related to what should be called trivial deformations of Lie subgroupoids where it is required to deform the groupoid on left as well, and the family defines the type of deformations allowed. These deformations are studied in detail in [9].
The arguments used in the proof of Theorem 8.5 can also be applied to get results concerning the two types of deformations defined above. The following two results give conditions to characterize deformations which are strongly trivial up to automorphisms on left by using the usual deformation complex and its variation . In order to illustrate that, recall first that a Lie groupoid morphism induces a push-forward map .
Definition 8.5.
We will say that a family has a smooth pre-image in if there exist smooth families of cocycles and such that
Analogously, considering the variation , since , we say that a family has a smooth pre-image in if there exist smooth families of cocycles and sections such that
(33) |
Theorem 8.20.
Let be a deformation of . Assume that is compact and is an injective immersion. Then is strongly trivial up to automorphisms of if and only if has a smooth pre-image in by for , where is an open interval containing 0.
Proof.
Since the proof of this theorem is completely analogous to the proof of the previous theorems, we will only sketch the main ingredients. Assume that
(34) |
for smooth families and . Let denote the flow from time 0 to of the time dependent vector field on . Recall that since every is a 1-cocycle it follows that is a family of automorphisms of . Next we enunciate the key steps which prove the statement.
-
•
Shrinking if necessary, use the injective immersive condition of to extend the sections to a smooth time dependent section in which vanishes outside a compact set containing . That is, satisfies .
-
•
Consider the smooth family of bisections of induced by the flow of the time dependent right-invariant vector field on , and check that , for all . This last part can be checked by showing that both sides are integral curves of the time dependent vector field on starting at the same point at the same time. Indeed, on the one hand, the derivative follows from Proposition 8.1; and on the other hand, the derivative is
The proof of the converse statement is a direct computation after applying to the left strongly trivial expression . ∎
Analogously, we can avoid the injective immersive condition in the previous Theorem by considering the variation complex .
Theorem 8.21.
Let be a deformation of . Assume that is compact. Then is strongly trivial up to automorphisms of if and only if has a smooth pre-image in by for all , where is an open interval containing 0.
Proof.
The proof of this theorem follows the idea similar to that of the previous Theorem, however we now use the argument of extension of sections of the Theorem 8.10 to obtain the smooth family of global sections which give rise to the family of bisections of . ∎
We remark that considering the cokernel complex we get an equivalent result to that of Theorem 8.20. Indeed, let denote the image in the cokernel complex of the cocycles associated to the deformation .
Theorem 8.22.
Let be a deformation of . Assume that is compact and is an injective immersion. Then is strongly trivial up to automorphisms of if and only if is smoothly exact for , where is an open interval containing 0.
Proof.
The proof of this theorem follows directly from the fact that, under the injectivity of the cochain maps , the exactnees condition of the family of cocycles is equivalent to the existence of a smooth pre-image in by of the classes of . Thus, we will be able to use the same proof of Theorem 8.20. ∎
Remark 8.23.
In contrast, we can modify the cokernel complex in zero degree changing it by the 0-cochains of the variation complex, and the exactness of the deformation cocycles in this modified cokernel complex will allow us to define a trivial deformation different from that of Theorem 8.21. Indeed, in this case, the exactness of the deformation cocycles in the complex is equivalent to get a strongly trivial up to diffeomorphisms on the left deformation. These deformations are totally analogous to the ones which are strongly trivial up to automorphisms on the left, the only difference lies in taking a family of diffeomorphisms instead of a family of automorphisms of the Lie groupoid as in Definition 8.4.
As an application of the Theorem 8.21 we deduce now the particular case of 1-deformations of smooth maps of the Thom-Levine’s Theorem (see [23], p. 124) regarding the characterization of trivial -deformations of smooth functions between manifolds. In Section 9 we will get the full Thom-Levine’s Theorem as an instance of the deformation theory developed in this Section.
Example 8.24.
[Thom-Levine’s Theorem for 1-deformations] Note that smooth functions between manifolds are in 1-1 correspondence with morphisms between the associated pair groupoids. Indeed, these morphisms are of the form , where is a smooth function between manifolds. From this one also gets that an automorphism of a pair groupoid is totally determined by the corresponding diffeomorphism on the base manifold.
With this setting, one checks that the characterizing equation (33) of strongly trivial up to automorphisms deformations of morphisms between pair groupoids translates exactly to the condition which characterizes trivial deformations of smooth functions from the Thom-Levine Theorem. Indeed, if is a deformation of the smooth function , the characterizing condition given by Thom-Levine’s Theorem is described by the equation
(35) |
where and are time-dependent vector fields on the source and target manifolds, respectively. The equivalence between equations (33) and (35) follows directly from the fact that 1-cocycles (i.e. multiplicative vector fields) on a pair groupoid are of the form
where is an usual vector field on the base of the pair groupoid.
Theorem 8.25.
Let be a deformation of . Assume that is compact. Then is trivial up to automorphisms of if and only if has a smooth pre-image in by for all , where is an open interval containing 0.
Proof.
Assume that has smooth preimage by for all . That is, equation (34) is satisfied. Then, following the notations of the proof of the Theorem above, define the deformation of . The proof is completed after checking that is a trivial deformation. In fact, observe first that the cocycles are smoothly exact, and this smooth exactness follows from equation (34). Therefore, the triviality of is a consequence of Theorem 8.5. ∎
Remark 8.26.
Notice that, by considering the cokernel complex , one similarly checks that the smooth exactness of the cocycles, for all small enough, is equivalent to get a deformation which is trivial up to diffeomorphisms of .
As a final remark of this section, observe that all the notions of triviality defined here arise from the complexes and maps involved in the following diagram of exact sequences.
Indeed we can summarizes the results as follows. On the vertical direction: under the injective immersive condition of the morphisms, the requirements on the deformation cocycles, regarding either existence of smooth pre-images or exactness on the cokernel complexes, turn out to be equivalent to characterize strongly trivial u.t.a.l. (up to automorphisms on the left) deformations (Theorems 8.20, 8.22). Under the non-injective immersive condition, the existence of smooth pre-image requirement gives us triviality up to automorphisms on the left (triviality u.t.a.l.), see Theorem 8.25. However the exactness of the cocycles on the cokernel complex, gives us triviality u.t.d.l. (up to diffeomorphisms on the left), see Remark 8.26.
Analogously, considering the variated complex we can avoid the injective imersive requirement on the morphism and show that the existence of smooth pre-images condition gives us deformations which are strongly trivial u.t.a.l. (Theorem 8.21) and the exactness on a variation of the cokernel complex (using the 0-degree cochains ) gives us strongly trivial u.t.d.l. deformations (Remark 8.23).
On the horizontal direction, pre-image by the pullback gives deformations which are trivial up to automorphisms on the right (u.t.a.r.), see Theorem 8.14. By adding the injective immersive condition on the morphism, we obtain strongly trivial u.t.a.r. (see Theorem 8.15). Similarly, the weaker condition of exactness on the cokernel complex gives us deformations which are trivial u.t.a.r. after assuming surjectivity and submersion conditions on the morphism (Theorem 8.18).
Finally, regarding the variated complexes , assuming the existence of smooth pre-images for the cocycles, we do not need to require the injective immersive condition on the morphism to prove the strong triviality u.t.a.r. of the deformation (see Theorem 8.17).
9. -deformations and Thom-Levine’s Theorem
In the previous sections we have studied 1-parameter deformations of morphisms or, in other words, paths of morphisms. The main point of this section is sketching the behaviour of deformations with parameters, that is, deformations depending on instead of just , and to get as an application the Thom-Levine’s Theorem regarding triviality of -deformations of differentiable maps.
Definition 9.1.
Given a morphism and an interval containing zero, a Lie groupoid morphism such that will be called a -deformation of . For every we denote by the morphism .
The notions of triviality, strong triviality and all the other equivalences between deformations of the previous section are defined by taking -parameter families instead of 1-parameter families of the elements involved in the definitions. For instance, a -deformation of is said to be trivial if there exists a smooth -parameters family of gauge maps covering such that and
Similarly a -deformation is called strongly trivial up to automorphisms on the left if there exist smooth -parameters families and of automorphisms of and bisections of such that , and
for all small enough.
By fixing the first components of , say in , in a -deformation , one gets a 1-parameter family which we call a canonical 1-deformation in along the (canonical) direction of . Similarly one defines the canonical 1-deformations in along the other canonical directions of , for and all small enough.
Remark 9.1.
Notice that all the canonical 1-deformations in a trivial -deformation are trivial but, on the other hand, all the canonical 1-deformations in along a specific direction, say , might be trivial without implying that the whole -deformation is trivial. In fact, as Theorem below shows, a necessary and sufficient condition for the triviality of is that all the canonical 1-deformations in are trivial. Note also that, by Theorem 8.5, the triviality of all the canonical 1-deformations (in ) along amounts to the existence of a section (i.e., a -parameter family of sections ) such that
or equivalently
for all .
Theorem 9.2.
Let be a -deformation of . Then the deformation is trivial if and only if the canonical 1-deformations in are trivial. That is, if and only if there exist -families of sections, such that
(36) |
for and all small enough.
Proof.
Assume that the canonical 1-deformations are trivial, and that ,…, are the families of gauge maps, induced by the families of sections ,…, , which make the triviality along the canonical directions, that is, they hold
for every . Then, it follows that
(37) |
proving that the -deformation is trivial. The converse statement follows easily. ∎
Analogous to Theorem 8.21, to characterize -deformations which are strongly trivial up to automorphisms on the left we consider the pre-images by of the deformation classes in of the canonical 1-deformations in .
Theorem 9.3.
Let be a -deformation of , and assume that is compact. Then the deformation is strongly trivial up to automorphisms on the left if and only if the canonical 1-deformations in are strongly trivial up to automorphisms on the left. That is, if and only if there exist -families and ,…, of sections of and multiplicative vector fields on , such that
(38) |
for , and small enough.
Proof.
Assume that equations (38) are satisfied by the -deformation then, by Theorem 8.21, every canonical 1-deformation in is strongly trivial up to automorphisms of . Let
denote the families of bisections of and automorphisms of induced, respectively, by the families of sections ,…, and vector fields ,…, . Then, they hold
for every . Hence, it follows that
(39) |
which proves the strong triviality u.t.a.l. of . The converse statement follows directly in a similar way to that of the triviality case in the previous theorem. ∎
Remark 9.4.
Note that the proofs of the previous two theorems can be seen as an application of a zig-zag principle where we go through each of the canonical directions checking the triviality one by one.
Remark 9.5.
Notice that, using this zig-zag principle, we can state analogous results for -deformations considering all the other type of "trivial" deformations defined in the previous section. In fact, as in the previous two theorems, which imitate the corresponding comological equations of Theorems 8.5 and 8.21 in the Section above, we get then a set of cohomological equations imitating the respective 1-parameter ones which characterize each type of deformation.
We next consider the notion of -deformations of smooth maps necessary for the statement of the Thom-Levine Theorem which we establish below as well. After that we will see that the Thom-Levine Theorem is just a particular case of the Theorem 9.3 above.
Let be a smooth map between the manifolds and . The smooth -family of maps between the manifolds is called a -deformation of if . We will also refer to the -family by its restrictions . The deformation is said to be trivial if there exist -deformations and of the identity maps and such that
(40) |
for all small enough. The following version of Thom-Levine’s Theorem we take from ([23], p.124).
Theorem 9.6 (Thom-Levine’s Theorem).
Let be a -deformation of , and assume that is compact. Then is trivial if and only if there exist smooth families and of vector fields on and (for ) such that
(41) |
for .
Remark 9.7.
Remark 9.8.
We remark that Theorem 9.3 turns out to be a generalization to Lie groupoids of the Thom-Levine Theorem, more precisely, one checks that the Thom-Levine Theorem identifies exactly with the Theorem 9.3 applied to morphisms between pair groupoids. Indeed, similar to Example 8.24, which considers 1-deformations, a -deformation of a morphism between two pair groupoids amounts to a -deformation of smooth maps and, also, a strongly trivial up to automorphisms on the left deformation of a morphism between pair groupoids is equivalent to a trivial -deformation of the corresponding smooth map between the base manifolds of the pair groupoids. And additionally, the equations (38) translate exactly to the equations (41), where this matching follows directly from the fact that 1-cocycles (i.e. multiplicative vector fields) on a pair groupoid are of the form
where is an usual vector field on the base of the pair groupoid.
Remark 9.9.
The Thom-Levine Theorem is a supporting step in order to prove the equivalence between the stability and the infinitesimal stability of smooth maps ([23], p. 127). Heuristically, the original proof of Thom-Levine Theorem and ours are very similar, consisting in integrate certain vector fields to get the triviality of the deformation out of these integrations, however our proof using Lie groupoids turns out to be more straightforward and geometrical, no needing so many technical details and also obtaining the necessity condition for triviality directly (and with a fast computation) as a velocity interpretation of the deformation. In summary, to get the Thom-Levine Theorem, we have directly used Theorems 9.3, 8.21, 8.10 and Proposition 8.4.
10. Stability of morphisms
In this section we apply the results obtained in the Section 8 in order to obtain stability properties of Lie groupoid morphisms under deformations. The key fact in the proof of the results below will be to combine the vanishing results (to get smooth transgressions of the deformation cocycles) with the Moser type arguments explored in Section 8.
Theorem 10.1.
Let be a Lie groupoid morphism. Assume that is transitive. If either the groupoid is proper or has trivial isotropy (i.e. is a pair groupoid!), then any deformation of is trivial.
Proof.
Following the notations of Definition 8.1, let be a deformation of . On the one hand, since the groupoid is transitive, it follows that the normal bundle becomes the null bundle over . That fact implies that the 1-cocycle (see the proof of Proposition 8.12, where this cocycle is defined), corresponding to the deformation , lies in the kernel of the map of the sequence (16) in Section 6. On the other hand, the condition of either trivial isotropy of or properness of implies that vanishes. Hence, the 1-cocycle is exact, which amounts to the smooth exactness of the family of 1-cocycles associated to the deformation. Thus, the Theorem 8.5 concludes the proof. ∎
Corollary 10.2.
Let be a Lie groupoid morphism whose base map is an injective immersion. Assume that is transitive. If the groupoid is compact, then any deformation of is strongly trivial.
Corollary 10.3.
Let be a Lie groupoid morphism whose base map is an injective immersion. Assume that is transitive and that the base of is compact. If the groupoid has trivial isotropy, then any deformation of is strongly trivial.
Thus, Theorem 10.1 says that under properness and transitivity of and , respectively, any curve , with , is constant when viewed in the category of Lie groupoids and isomorphism classes of morphisms. Moreover, such conditions of compactness and transitivity assumed in the results are fundamental: for instance, any curve passing through more than one orbit of determines a non-trivial deformation for a constant morphism . It also follows that if we additionally take as being a compact Lie group, then the base map of is an injective immersion but the deformation will not be strongly trivial; thus the transitivity of is also necessary in the corollaries. Also for non-compact in Corollary 10.2, one can check that if and viewed as Lie groups, then the inclusions of the linear spaces in yield a non-strongly trivial deformation even if is transitive. Considering now the last corollary, this latter counterexample also verifies that the trivial isotropy condition can not be removed in the statement. Additionally, assuming that is non-compact, take the pair groupoids over , since the identity map over can be deformed, by using a bump function, to not be a diffeomorphism for any small, it follows that it induces a non-strongly trivial deformation of the identity morphism.
11. Simultaneous deformations
In this section we put together the deformation theory of both Lie groupoids and Lie groupoid morphisms in order to study the most general deformation problem: the simultaneous deformation of the triple given by , where is a morphism of Lie groupoids.
Definition 11.1.
Let be a Lie groupoid morphism. A deformation of the triple is a deformation of and a deformation of over a common open interval containing 0, and a morphism which deforms in a compatible manner, in the sense that and, for each , is a morphism of Lie groupoids. We will denote a deformation of the triple by .
Definition 11.2.
Let and be two deformations of the triple . We will say that both deformations are equivalent if there exist an open interval containing 0 and isomorphisms of Lie groupoids , and a gauge map covering defined over such that
(42) |
We will also say that a deformation of is trivial if it is equivalent to the constant deformation .
Remark 11.1.
On rigidity of triples it is straighforward checking, from the rigidity of compact Lie groupoids and Lie groupoid morphisms from Section 10, that if and are compact Lie groupoids, and moreover is transitive then any deformation of the triple is trivial.
Remark 11.2.
For deformations of the triple we can also define strongly equivalent deformations. This equivalence relation corresponds to the special case in which the gauge map of equation (42) is induced by a bisection of . That is, when , for . A deformation will be called strongly trivial if it is strongly equivalent to the constant deformation.
11.1. Deformation complex and triviality of simultaneous deformations
Consider the diagram of cochain maps explained in Remark 4.1
We construct the complex which controls the deformations of the triple by taking the cone of this diagram as follows. Take the mapping-cone complex associated to the cochain map ,
(43) |
with differential . Notice that the map above induces a cochain map putting zero on the first component, . Take now the mapping-cone associated to , getting the deformation complex of the triple,
with differential .
In this way, given a -constant deformation of (i.e., those where and are -constant deformations), by computations similar to those in Proposition 8.1, we get that is a 2-cocycle in , where and are the respective deformation cocycles for and , and is the usual cochain (Section 8) associated to a deformation of morphisms. Indeed, the fact that is a cocycle follows by applying to the compatibility of the deformation :
The fact that the corresponding cohomology class of only depends on the equivalence class of the deformation is also an analogous computation.
Remark 11.3.
One can also consider a non -constant deformation of the triple and obtain its associated deformation cohomology class. Indeed, if and are transversal vector fields (i.e., vector fields which project to ) with and being the corresponding deformation cocycles, then is the associated deformation cocycle, where .
We pass now to establish the main results concerning simultaneous -constant deformations.
Theorem 11.4.
Let be a deformation of , with and compact. If the family of associated cocycles is transgressed by a smooth family of cochains , then the deformation is trivial.
Proof.
Exactness of the family of cocycles amounts to
(44) |
By the first and third equation, if and are the time-dependent flows starting at zero of the vector fields and respectively, then they define the equivalences with the corresponding constant deformations of Lie groupoids. We claim that these equivalences can be used to prove the triviality of . In fact, by using Theorem 8.5, we will show that the family of morphisms is a trivial deformation in the sense of the definition at the beginning of Section 3. Such a family satisfies,
where the second equality follows from the fact that
which is obtained by applying to . Therefore, by Theorem 8.5, , as claimed. ∎
Analogously, one can check the following
Theorem 11.5.
Let be a deformation of , with and compact, and an injective immersion. If the family of associated cocycles is transgressed by a smooth family of cochains , then the deformation is strongly trivial.
11.2. Particular cases and relations between (sub)complexes
In view that the complex controls the most general type of deformations of the three structures , in this section we consider particular cases of deformations of the triple and their relation with some subcomplexes of . We begin with the simplest case.
and are fixed
In this case, we get a deformation of a Lie groupoid morphism. This fact is expressed, in cohomological terms, by the injection
Moreover, this map takes the infinitesimal cocycle of to the infinitesimal cocycle of . Therefore, in this case, the relevant subcomplex controlling deformations of this type is given by as expected.
is fixed
In this case, the relevant subcomplex is . In fact, it is not hard to see that a deformation of the form is governed by the mapping-cone complex (see (43)), where one associates the cocycle to the deformation. Thus the (injective) chain map
(45) |
shows that the subcomplex controls the deformations of the triple when we fix the groupoid .
This kind of -fixed deformations is quite related to what is called deformations of Lie subgroupoids. Indeed, it can be checked that the subcomplex can be viewed as the complex which controls such a deformations. The details of that will be developed in the future work [9].
12. Morita invariance and Deformation cohomology of generalized morphisms
We now investigate the behaviour of the deformation cohomology under Morita maps of Lie groupoids, show its invariance by Morita morphisms and use it to define a deformation cohomology for generalized morphisms between Lie groupoids. The proof of the invariance results here are just applications of the recent developed concept of VB-Morita maps between VB-groupoids [20].
Proposition 12.1.
Let be a Lie groupoid morphism. Assume is a Morita map. Then .
Proof.
Recall that the deformation complexes of and are respectively isomorphic to the VB-complexes of the pullback VB-groupoids and . Then, since is a Morita map, then the canonical vector bundle map insures that such VB-groupoids are VB-Morita equivalent (see Corollary 3.7 of [20]) and thus, by the VB-Morita invariance of the VB-cohomology of [20], we have that , as claimed.
∎
Proposition 12.2.
Let be a Lie groupoid morphism. If is a Morita map, then .
Proof.
Recall that the complexes computing the deformation cohomology of and are isomorphic, respectively, to the VB-complexes of the two VB-groupoids and , thus we shall prove that the cohomologies of these VB-complexes are isomorphic.
Since is a Morita map, it follows that the differential and the canonical bundle map are VB-Morita maps. Thus, the induced map of the VB-groupoids (over ) turns out to be also a VB-Morita map.
Therefore, Corollary 3.9 in [20] ensures that its dual map
is a VB-Morita map. Hence, finally by taking the pullback by of these VB-groupoids, one gets the induced VB-Morita map which, by the VB-Morita invariance of the VB-cohomology [20], then induces the isomorphism of the indicated VB-cohomologies. ∎
We can use now the results in this Section to define a deformation cohomology for generalized maps which are regarded as the morphisms in the category of differentiable stacks.
12.1. A deformation complex for fractions
In the setting of the theory of localization of categories and calculus of fractions [27], given two Lie groupoids and , a fraction is defined by two maps and where is a third Lie groupoid and is a Morita morphism. That fraction is often also denoted by .
(46) |
Two fraction are said to be equivalent if there exist a third fraction and Morita maps and making the below diagram commutative up to isomorphisms of morphisms.
This is indeed an equivalence relation on fractions as can be proved by using weak fibred products ([34], p. 124). The equivalence class of the fraction determines a generalized map , also known as generalized morphism or stacky map, which can be viewed as a smooth map between the differentiable stacks presented by and (see [19], Section 6.2).
Given the fraction , there is a map of complexes induced by the morphisms and . We define the deformation complex of the fraction by the mapping-cone complex of the map . That is,
with differential .
Remark 12.3.
Notice that this complex can also be defined for any pair of morphisms set as in diagram (46). However the fact that in a fraction the left leg is a Morita map can be used to get an alternative expression of the deformation complex useful for computations.
Equivalent fractions have isomorphic deformation cohomology as can be proven by using the quasi-isomorphisms and of deformation complexes (Propositions 12.1 and 12.2), induced by a Morita map which relates two fractions, and by the isomorphism of deformation complexes of isomorphic morphisms (Theorem 4.8). Hence, the deformation complex of a fraction induces a well-defined deformation cohomology for generalized morphisms. We register that fact in the following theorem.
Theorem 12.4.
If and are equivalent fractions from to , then their deformation cohomologies and are isomorphic.
Thus, the deformation cohomology of a fraction is an algebraic object associated to the stacky map it represents. The deformation cohomology also turns out to be very involved in the infinitesimal study of the space of generalized maps. For instance, every deformation of a fraction has a corresponding 1-cocycle whose cohomology class should be regarded as the velocity vector at (the class of) of the associated path of generalized morphisms. More extended and detailed results will lie on future work.
13. Application: some remarks on deformations of multiplicative forms
In this section we use the deformation complex of morphisms to study deformations of multiplicative forms on Lie groupoids. Also, we characterize trivial defomations of multiplicative forms in cohomological terms. The content of this section is also relevant to develope the theory of deformations of symplectic groupoids as in [8].
Recall that a -form is called a multiplicative -form if it satisfies the multiplicativity condition
(47) |
where are the canonical projections and multiplication of . A map
is called a smooth family of multiplicative -forms if, for every , is a multiplicative -form. We say that a smooth family of multiplicative -forms is a deformation of if .
We consider first the particular case of multiplicative symplectic 2-forms on (Proposition 13.1 below), then we will generalize the situation to multiplicative -forms. A more advanced study of the case of multiplicative symplectic 2-forms is made in [8] where we consider also a simultaneous deformation of the underlying Lie groupoid and a relation of its deformation cohomology to the Bott-Shulmann-Stasheff complex. Recall that the classical Moser’s theorem of symplectic geometry deals with symplectic 2-forms on a differentiable manifold. This theorem says that a smooth family of symplectic forms on a manifold is recovered as the pullback by a family of diffeomorphisms of the symplectic form at time zero if, and only if, there exists a smooth family of 1-forms such that
(48) |
The following proposition formulates an analogous result in the context of Lie groupoids, where instead we consider the multiplicative de Rham complex of , whose elements are multiplicative forms of .
Proposition 13.1.
Let be a compact symplectic groupoid, and assume that is a deformation of . Then, for a smooth family of groupoid automorphisms of , with , if and only if the family of cocycles is smoothly exact in .
Proof.
This proof is just a multiplicative version of the proof of the classical Moser theorem. In order to prove it, we notice that the multiplicative symplectic 2-form yields an isomorphism between the space of multiplicative 1-forms and the space of multiplicative vector fields . Therefore the time dependent flow of the transgressing family of vector fields will be given by a family of automorphisms of starting at the identity . ∎
Remark 13.2.
We can express equivalently the smooth exactness of by saying that has a smooth preimage by in . That is, saying that there exists a smooth family such that for every , which is the same condition of equation (48).
The following example shows that the classical Moser theorem is obtained as an application of the previous proposition to the pair groupoid.
Example 13.3.
Let be a smooth family of symplectic structures on a compact manifold , and consider the induced family of symplectic groupoids , where . Thus, since is smoothly exact in if and only if is smoothly exact in , and the automorphisms of are of the form , where , it follows that the previous proposition reduces to the classical Moser’s theorem in this case.
Next we pass to the general case of multiplicative -forms on the groupoid . A multiplicative -form can also be viewed as the morphism of Lie groupoids
which is -linear and skewsymmetric in the sense that is -linear with respect to the linear structure of over . With this viewpoint, a smooth family of skew-symmetric and -linear Lie groupoid morphisms is called a deformation of the multiplicative -form .
From the cohomological perspective, the skew-symmetry and -linearity of the morphism translate to skew-symmetric and -linear deformation cochains in as described below.
Definition 13.1.
The deformation complex of a multiplicative -form consists of the skew-symmetric and -linear cochains of . We denote by such a subcomplex of deformation cochains.
Explicitly, the deformation complex can be described as follows. Consider the natural identification , where and are the projections of the respective tangent bundles. Thus, the elements of are given by those elements in which turn the composition
(49) |
a -linear and skew-symmetric map. Along with the restriction of the deformation differential of , becomes indeed a subcomplex: the deformation complex of .
Remark 13.4.
Considering the identification of with and the composition
the elements of the complex can be regarded as the elements of the subcomplex of consisting of fiberwise -linear and skew-symmetric differentiable cochains of . Moreover, from a straightforward computation, one also gets the correspondence between the differentials of these complexes.
The following proposition shows that this deformation complex is isomorphic to one that only depends on the simplicial structure of and not of or explicitly.
Proposition 13.5.
Given any multiplicative -form , the deformation complex is isomorphic to ; where is the differential induced from the simplicial structure of .
Proof.
The correspondence between the cochains is straightforward from the description of the elements of in expression (49). And the correspondence between the differentials follows directly from the identification of the deformation differential with the simplicial differential of the differentiable subcomplex of Remark 13.4 above, which in turn identifies with the simplicial differential of the simplicial complex of -forms over the nerve of .
∎
With this setting, a deformation of by multiplicative forms on determines:
-
(1)
a smooth family of 1-cocycles in (Proposition 8.1);
- (2)
Notice, however, that the cocycle condition also follows by taking directly the derivative of the multiplicativity condition of (equation (47)), obtaining that .
We define now a map of complexes which, composing with the map above, is a relevant element in the statement of the theorem below that generalizes Proposition 13.1 to -forms. Consider the tangent lift of deformation cochains which, as checked in [29], turns out to be a cochain map. For it is defined by
where is the involution map of the double tangent bundle of . We thus define as the map . Explicitly,
Theorem 13.6.
Let be a deformation of the multiplicative -form . Assume that is compact. Then, for a smooth family of groupoid automorphisms of , with , if and only if the family of multiplicative -forms has a smooth pre-image in by the map .
Proof.
The smoothness of the pre-images of implies
for some smooth family of deformation 1-cocycles (i.e. multiplicative vector fields) of . Thus, let denote by the flow of the time-dependent vector field (starting at time zero) covering , then we have
Equivalently,
In other words,
This says that
where is a smooth family of automorphisms of due to the multiplicativity of the time-dependent vector field .
∎
Remark 13.7.
The previous Theorem reduces to the Proposition 13.1 when is taken as a family of multiplicative and symplectic 2-forms on .
The following theorem now tells us about the meaning of the smooth cohomological triviality of the family of 1-cocycles in terms of the deformation . Its proof is similar to the previous one. This fact will be approached a bit more for the particular case of symplectic groupoids in [8].
Theorem 13.8.
Let be a deformation of the multiplicative -form . Assume that the groupoid is compact. Then, smooth exactness of the familiy of cocycles amounts to the fact that the deformation is of the form ; where is a smooth family of groupoid automorphisms of , with and is a smooth family of 2-forms on with .
References
- [1] Alaoui, A. E. K., Guasp, G., and Nicolau, M. On deformations of transversely homogeneous foliations. Topology 40, 6 (2001), 1363–1393.
- [2] Blaom, A. Geometric structures as deformed infinitesimal symmetries. Transactions of the American Mathematical Society 358, 8 (2006), 3651–3671.
- [3] Bunk, S. A method of deforming G-structures. Journal of Geometry and Physics 96 (2015), 72–80.
- [4] Bursztyn, H., Cabrera, A., and del Hoyo, M. Vector bundles over Lie groupoids and algebroids. Advances in Mathematics 290 (2016), 163–207.
- [5] Bursztyn, H., Crainic, M., Weinstein, A., Zhu, C., et al. Integration of twisted Dirac brackets. Duke Mathematical Journal 123, 3 (2004), 549–607.
- [6] Cabrera, A. Generating functions for local symplectic groupoids and non-perturbative semiclassical quantization. Communications in Mathematical Physics 395, 3 (2022), 1243–1296.
- [7] Cabrera, A., and Drummond, T. Van est isomorphism for homogeneous cochains. Pacific Journal of Mathematics 287, 2 (2017), 297–336.
- [8] Cárdenas, C. C., Mestre, J. N., and Struchiner, I. Deformations of symplectic groupoids. arXiv preprint arXiv:2103.14008 (2021).
- [9] Cárdenas, C. C., and Struchiner, I. On deformations of Lie subgroupoids. Work in progress.
- [10] Cárdenas, C. C., and Struchiner, I. Stability of Lie group homomorphisms and Lie subgroups. Journal of Pure and Applied Algebra 224, 3 (2020), 1280–1296.
- [11] Coste, A., Dazord, P., and Weinstein, A. Groupoϊdes symplectiques. Publ. Dept. Math. Lyon 2, A (1987), 1–62.
- [12] Crainic, M., and Fernandes, R. L. Integrability of Poisson brackets. Journal of Differential Geometry 66, 1 (2004), 71–137.
- [13] Crainic, M., and Fernandes, R. L. Lectures on integrability of Lie brackets. Geom. Topol. Monogr 17 (2011), 1–107.
- [14] Crainic, M., Fernandes, R. L., and Mărcuţ, I. Lectures on Poisson geometry, vol. 217. American Mathematical Soc., 2021.
- [15] Crainic, M., Mestre, J. N., and Struchiner, I. Deformations of lie groupoids. International Mathematics Research Notices 2020, 21 (2020), 7662–7746.
- [16] Crainic, M., and Moerdijk, I. Deformations of Lie brackets: cohomological aspects. Journal of the European Mathematical Society 10, 4 (2008), 1037–1059.
- [17] Del Hoyo, M., and Fernandes, R. L. On deformations of compact foliations. Proceedings of the American Mathematical Society 147, 10 (2019), 4555–4561.
- [18] del Hoyo, M., and Fernandes, R. L. Riemannian metrics on differentiable stacks. Mathematische Zeitschrift 292, 1-2 (2019), 103–132.
- [19] del Hoyo, M., and Fernandes, R. L. Riemannian metrics on differentiable stacks. Mathematische Zeitschrift 292, 1-2 (2019), 103–132.
- [20] del Hoyo, M., and Ortiz, C. Morita equivalences of vector bundles. International Mathematics Research Notices 2020, 14 (2020), 4395–4432.
- [21] Gerstenhaber, M. On the deformation of rings and algebras. Annals of Mathematics (1964), 59–103.
- [22] Geudens, S., Tortorella, A. G., and Zambon, M. Deformations of symplectic foliations. Advances in Mathematics 404 (2022), 108445.
- [23] Golubitsky, M., and Guillemin, V. Stable mappings and their singularities, vol. 14. Springer Science & Business Media, 2012.
- [24] Gracia-Saz, A., and Mehta, R. A. VB-groupoids and representation theory of Lie groupoids. Journal of Symplectic Geometry 15, 3 (2017), 741–783.
- [25] Griffiths, P. A. Deformations of G-structures. Mathematische Annalen 155, 4 (1964), 292–315.
- [26] Guillemin, V., and Sternberg, S. Deformation theory of pseudogroup structures. No. 64. American Mathematical Soc., 1966.
- [27] Kashiwara, M., and Schapira, P. Categories and Sheaves. Springer Berlin Heidelberg, 2006.
- [28] Kodaira, K., and Spencer, D. C. On deformations of complex analytic structures, i. Annals of Mathematics (1958), 328–401.
- [29] La Pastina, P. P., and Vitagliano, L. Deformations of vector bundles over Lie groupoids. Revista Matemática Complutense (2022), 1–39.
- [30] Li-Bland, D., and Ševera, P. Quasi-Hamiltonian groupoids and multiplicative manin pairs. International Mathematics Research Notices 2011, 10 (2011), 2295–2350.
- [31] Mackenzie, K., and Xu, P. Classical lifting processes and multiplicative vector fields. arXiv preprint dg-ga/9710030 (1997).
- [32] Mackenzie, K. C. General theory of Lie groupoids and Lie algebroids, vol. 213. Cambridge University Press, 2005.
- [33] McDuff, D., and Salamon, D. Introduction to symplectic topology.
- [34] Moerdijk, I., and Mrcun, J. Introduction to foliations and Lie groupoids, vol. 91. Cambridge University Press, 2003.
- [35] Nijenhuis, A., and Richardson Jr, R. W. Deformations of homomorphisms of Lie groups and Lie algebras. Bulletin of the American Mathematical Society 73, 1 (1967), 175–179.
- [36] Oh, Y.-G., and Park, J.-S. Deformations of coisotropic submanifolds and strong homotopy Lie algebroids. Inventiones mathematicae 161, 2 (2005), 287–360.
- [37] Ortiz, C., and Valencia, F. Morse theory on Lie groupoids. arXiv preprint arXiv:2207.07594 (2022).
- [38] Richardson Jr, R., et al. Deformations of subalgebras of Lie algebras. Journal of Differential Geometry 3, 3-4 (1969), 289–308.
- [39] Schätz, F., and Zambon, M. Deformations of coisotropic submanifolds for fibrewise entire Poisson structures. Letters in Mathematical Physics 103, 7 (2013), 777–791.
Universidade Federal Fluminense, Instituto de Matemática e Estatίstica, Rua Prof. Marcos Waldemar de Freitas Reis, S/n, 24210-201, Niterói, RJ, Brazil
E-mail address, C. C. Cárdenas: ccardenascrist@gmail.com