Differential models for the Anderson dual to bordism theories and invertible QFT’s, II
Abstract.
This is the second part of the work on differential models of the Anderson duals to the stable tangential -bordism theories , motivated by classifications of invertible QFT’s. Using the model constructed in the first part [YY21], in this paper we show that pushforwards in generalized differential cohomology theories induces transformations between differential cohomology theories which refine the Anderson duals to multiplicative genera. This gives us a unified understanding of an important class of elements in the Anderson duals with physical origins.
1. Introduction
This is the second part of the work on differential models of the Anderson duals to the stable tangential -bordism theories motivated by classifications of invertible QFT’s. The generalized cohomology theory is conjectured by Freed and Hopkins [FH21, Conjecture 8.37] to classify deformation classes of possibly non-topological invertible quantum field theories (QFT’s) on stable tangential -manifolds. Motivated by this conjecture, in the previous paper [YY21] by Yonekura and the author, we constructed a model of and its differential extension by abstractizing properties of partition functions for invertible QFT’s. This paper is devoted to their relations with multiplicative genera. We show that pushforwards (also called integrations) in generalized differential cohomology theories allow us to construct differential refinements of certain cohomology transformations which arise from the Anderson dual to multiplicative genera and the module structures of the Anderson duals. This gives us a unified understanding of an important class of elements in the Anderson duals with physical origins. Moreover, having transformations in the differential level gives a more direct connection to physical picture, since they can be regarded as transformations of QFTs without taking deformation classes, recovering the information of partition functions.
First, we explain the motivations of the previous paper. As we recall in Section 2, the differential group consists of pairs , where where is the total degree, and is a map which assigns -values to differential stable tangential -cycles of dimension over , which satisfy a compatibility condition with respect to bordisms. The physical interpretation is that is the complex phase of the partition function of an invertible QFT. For example, given a hermitian line bundle with unitary connection over , the pair of the first Chern form and the holonomy function gives an element for and (Subsection 3.4.1). Similarly, given a hermitian vector bundle with unitary connection, we can construct even-degree elements for using the reduced eta invariants of twisted Dirac operators (Subsection 3.4.3): this theory is related to anomaly of spinor field theory. Then, a natural mathematical question arises: what are these elements mathematically? It is natural to expect a topological characterization of these elements. Questions of this kind also appears in [FH21, Conjecture 9.70] (also recalled in Conjecture 3.50 below). This paper is devoted to this question. Actually, these examples are special cases of the general construction in this paper which we now explain.
Now we explain the general settings. In this paper, the tangential structure groups (see Section 2) is assumed to be multiplicative, i.e., the corresponding Madsen-Tillmann spectrum is equipped with a structure of a ring spectrum. Assume we are given a ring spectrum with a homomorphism of ring spectra,
such is also called a multiplicative genus, and examples include the usual orientation and the Atiyah-Bott-Shapiro orientations and .
On the topological level, a ring homomorphism gives pushforwards in for proper -oriented smooth maps. Pushforwards in differential cohomology, or differential pushforwards, are certain differential refinements of topological pushforwards. Basically, they consist of corresponding maps in for each proper map with a “differential -orientation”. The formulations depend on the context. To clarify this point, in this paper we use the differential extension of constructed by Hopkins-Singer [HS05], and use the formulation of differential pushforwards in that paper. Throughout this paper, we assume that is rationally even, i.e., for any integer . In this case, by [Upm15] there exists a canonical multiplicative structure on the Hopkins-Singer’s differential extension associated to a fundamental cycle . The theory of differential pushforwards gets simple in this case. This point is explained in Subsection 3.1 and Appendix A. Of course our result applies to any model of differential extension of which is isomorphic to the Hopkins-Singer’s model. Practically, most known examples of differential extensions are isomorphic to Hopkins-Singer’s model (see Footnote 2). The holonomy functions are examples of differential pushforwards in the case , and the reduced eta invariants are those for by the result of Freed and Lott [FL10] and Klonoff [Klo08].
Let be an integer such that . As we show in Subsection 3.3, the above data defines the following natural transformation,
(1.1) |
on (Definition 3.36).
The main result of this paper is the following topological characterization of the transformation (1.1).
Theorem 1.2 (=Theorem 3.26).
In the above settings, let be a manifold and be an integer. For and , the element coincides with the following composition,
(1.3) |
Here we denoted .
As we will see in Subsection 3.4, this result applies to the above mentioned examples as follows.
Example 1.4 (Subsection 3.4.1).
Set with . We have the Anderson self-duality element . The transformation
(1.5) |
sends the class of a hermitian line bundle with connection to the element . Applying Theorem 1.2, we see that its deformation class coincides with the following composition,
Example 1.6 (Subsections 3.4.3 and 3.4.4).
Set with the Atiyah-Bott-Shapiro orientations . We have the Anderson self-duality element . The transformation
(1.7) |
maps the class of hermitian vector bundle with unitary connection to the element . Applying Theorem 1.2, we see that its deformation class coincides with the following composition,
In Subsection 3.4.4, we explain that this transformation can be interpreted as taking anomaly theories of complex spinor field theories.
The paper is organized as follows. In Section 2 we recall the definition of the differential models in [YY21]. Section 3 is the main part of this paper. We construct the natural transformation (1.1) and prove Theorem 1.2 in Subsection 3.3. We explain some examples, as well as relation with anomalies, in Subsection 3.4. As we explain in Subsection 3.1, there are certain subtleties regarding the formulations of differential pushforwards. In Appendix A, we collect the necessary results concerning differential pushforwards for submersions when is rationally even.
1.1. Notations and Conventions
-
•
By manifolds, we mean smooth manifolds with corners. We use the conventions explained in [YY21, Subection 2.3].
-
•
The space of -valued differential forms on a manifold is denoted by .
-
•
We deal with differential forms with values in a graded real vector space . In the notation , means the total degree. In the case if is infinite-dimensional, we topologize it as the colimit of all its finite-dimensional subspaces with the caonical topology, and set . This means that, any element in can locally be written as a finite sum with and for some for each . The space of closed forms are denoted by .
-
•
For a manifold and a real vector space , we denote by the trivial bundle over .
-
•
For a topological space , we denote by the map to . We set .
-
•
For two topological spaces and , we denote by the projection to .
-
•
We set .
-
•
For a real vector bundle over a topological space, we denote its orientation line bundle (rank- real vector bundle) by . For a manifold , we set .
-
•
For a spectrum , we require the adjoints of the structure homomorphisms are homeomorphisms. For a sequence of pointed spaces with maps , we define its spectrification to be the spectrum given by
-
•
For a generalized cohomology theory , we set .
-
•
The Chern-Dold homomorphism [Rud98, Chapter II, 7.13] for a generalized cohomology theory is denoted by
(1.8) - •
2. Preliminaries from [YY21]
In this section we recall necessary parts of the previous paper [YY21]. In [YY21], we constructed a model of a differential extension of the Anderson dual to the tangential -bordism homology theory .
First we recall the definition of the Anderson duals to spectra (see [YY21, Section 2.1], [HS05, Appendix B] and [FMS07, Appendix B]). The functor on the stable homotopy category is represented by a spectrum denoted by . The Anderson dual to the sphere spectrum, denoted by , is defined as the homotopy fiber of the morphism representing the transformation . For any spectra , the Anderson dual to , denoted by , is defined to be the function spectrum from to , . This implies that we have the following exact sequence for any spectra .
(2.1) | ||||
In [YY21] and the current paper, we are particularly interested in the Anderson dual to the -bordism homology theory. Here, is a sequence of compact Lie groups equipped with homomorphisms and for each , such that the following diagram commutes.
Here we use the inclusion defined by
throughout this paper. Given such , the stable tangential -bordism homology theory assigns the stable tangential bordism group . It is represented by the Madsen-Tillmann spectrum , which is a variant of the Thom spectrum . For details see for example [Fre19, Section 6.6]. In this paper we take and to be a spectrum as in [HS05, (4.60)]. The Anderson dual fits into the following exact sequence.
(2.2) | ||||
Now we proceed to recall the definition of . In [YY21], we defined the relative groups . But since we only deal with the absolute case in this paper, we concentrate on this case. To recall the definition of , we first recall the differential stable -structures on vector bundles.
Definition 2.3 (Differential stable -structures on vector bundles, [YY21, Definition 3.1]).
Let be a real vector bundle of rank over a manifold .
-
(1)
A representative of differential stable -structure on is a quadruple , where is an integer, is a principal -bundle with connection over and is an isomorphism of vector bundles over .
-
(2)
We define the stabilization of such by , where and are naturally induced on from and , respectively.
-
(3)
A differential stable -structure on is a class of representatives under the relation .
-
(4)
Suppose we have two representatives of the forms and , such that and are homotopic. In this case, the resulting differential stable -structures and are called homotopic.
If we forget the information of the connection , we get the corresponding notion of (topological) differential stable -structures. For a differential stable -structure , we denote the underlying topological structure by . Similar remarks apply to the various definitions below.
We also recall the normal variant, which we use in Appendix A.
Definition 2.4 (Differential stable normal -structures on vector bundles, [YY21, Definition 4.75]).
Let be a real vector bundle of rank over a manifold .
-
(1)
A representative of differential stable normal -structure on is a quadruple , where is an integer, is a principal -bundle with connection over and is an isomorphism of vector bundles over .
-
(2), (3)
We define the stabilization of such in the same way as Definition 2.3, and a differential stable normal -structure on is defined to be a class of representatives under the stabilization relation.
-
(4)
We define the homotopy relation between two such ’s also in the same way.
A differential stable tangential -structure on a manifold is a differential stable -structure on . Given a manifold , a differential stable tangential -cycle of dimension over is a triple , where is an -dimensional closed manifold, is a differential stable tangential -structure on and . Using these cycles, we defined
-
•
The abelian group , consisting of the equivalence classes of differential stable tangential -cycles of dimension over , with the equivalence relation generated by isomorphisms, opposite relation and homotopy relation [YY21, Definition 3.5].
-
•
The Picard groupoid , whose objects are differential stable tangential -cycles of dimension over , and morphisms are the bordism classes of bordisms [YY21, Definition 3.8].
By the theorem of Pontryagin-Thom we have an equivalence of Picard groupoids [YY21, Lemma 3.10],
(2.5) |
Here the right hand side denotes the fundamental Picard groupoid.
Let denote the module with the underlying vector space and the -action by . By the Thom isomorphism, we have
The fourth arrow in (2.2) gives the homomorphism
(2.6) |
By [YY21, Proposition 4.9] we have a canonical homomorphism
(2.7) |
The map is isomorphism if is finitely generated for all .
Let be any -graded vector space over . Given a vector bundle with a differential stable -structure , we get a homomorphism of -graded real vector spaces by the Chern-Weil construction [YY21, Definition 4.4 and Remark 4.10],
(2.8) |
Since the orientation bundle of a vector bundle and its normal bundle are canonically identified, for equipped with a differential stable normal -structure , we also get
(2.9) |
For we get a homomorphism
(2.10) | ||||
for each pair of objects in . In the following definition, we use (2.10) in the case and .
Definition 2.11 ( and , [YY21, Definition 4.15]).
Let be a manifold and .
-
(1)
Define to be an abelian group consisting of pairs , such that
-
(a)
is a closed -form111Recall that is the total degree. .
-
(b)
is a group homomorphism .
-
(c)
and satisfy the following compatibility condition. Assume that we are given two objects and in and a morphism from the former to the latter. Then we have
Abelian group structure on is defined in the obvious way.
-
(a)
-
(2)
We define a homomorphsim of abelian groups,
Here the homomorphism is defined by
We set
For we set and .
We defined the structure homomorphisms , and along with the -integration map for . One of the main results of [YY21] is the following.
3. Pushforwards in differential cohomologies and the Anderson duality
This is the main section of this paper. The main part is Subsection 3.3, where we construct the natural transformation (1.1) and prove Theorem 1.2. Subsections 3.1 and 3.2 are preparation for the construction and proof. We explain some examples, as well as relations with anomaly, in Subsection 3.4.
3.1. Preliminary–Differential pushforwards in the Hopkins-Singer model
In this subsection, we briefly explain the differential extensions of generalized cohomology theories constructed by Hopkins-Singer and the differential pushforwards (called integration in [HS05]) in that model. We explain it in more detail in Appendix A.
On the topological level, a ring homomorphism gives pushforwards in for -oriented proper smooth maps. For proper smooth maps of relative dimension with (topological) stable relative tangential -structures , we get the corresponding pushforward map,
(3.1) |
In particular in the case , for a closed manifold of dimension with a stable tangential -structure ([YY21, Definition 3.2]), we get
There are notions of differential refinements of the pushforward maps in . For example see [HS05, Section 4.10], [BSSW09, Section 2] and [Bun12, Section 4.8 – 4.10]. Basically, they consist of corresponding maps in for each proper map with a “differential -orientation”. The formulations depend on the context. In this paper, we adopt the one by Hopkins-Singer222 In particular we use the differential extension . Practically this is not restrictive. We are assuming is rationally even and multiplicative, so is equipped with a canonical multiplicative structure by [Upm15]. Thus, when the coefficients of are countably generated, we can apply the uniqueness result in [BS10, Theorem 1.7] to conclude that any other multiplicative differential extension (defined on the category of all smooth manifolds) is isomorphic to . .
Hopkins and Singer gave a model of differential extensions, which we denote by , for any spectrum , in terms of differential function complexes. In general we choose a -graded vector space , and a singular cocycle . Then for each and for each manifold , we get a simplicial complex called differential function complex,
consisting of differential functions . This complex has a filtration , . The differential cohomology group is defined as (it is denoted by in [HS05]),
In particular this means that an element in is represented by a differential function , consisting of a continuous map , a closed form and a singular cochain such that as smooth singular cocycles.
A particularly important case is when and is the fundamental cocycle, i.e., a singular cocycle representing the Chern-Dold character of . In this case the associated differential cohomology groups satisfies the axioms of differential cohomology theory in [BS10]. The isomorphism class of the resulting group is independent of the choice of the fundamental cocycle , with an isomorphism given by a cochain cobounding the difference.
In [HS05, Section 4.10], a differential pushforward is defined simply as maps of differential function spaces333 This point is important in the proof of Proposition 3.21, which is the main ingredient of the proof of the main result (Theorem 3.26). This is the reason why we want to use the Hopkins-Singer’s formulation. ,
(3.2) |
refining the map . Here we are taking , and the cocycle is obtained by applying on the coefficient of . Then444 As we explain in Appendix A.2, this process needs some additional choices of cochains. By the assumption that is rationally even, the resulting map on the differential cohomology level does not depend on the choices. , the map associates to every proper neat map of of relative dimension with a differential (tangential) -orientation with a map
(3.3) |
called the differential pushforward map.
Remark 3.4.
As we explain in Appendix A.2 and A.3, the definition of (the tangential version of) differential -oriented maps in [HS05] differs from the differential stable relative -structure in [YY21, Definition 5.12]. Fix a fundamental cocycle . Given a proper smooth map , a topological tangential -orientation consists of a choice of embedding for some , a tubular neighborhood of in with a vector bundle structure , and a classifying map . A differential tangential -orientation consists of its lift to a differential function
(3.5) |
Then the map (3.3) is given by (3.2) and the Pontryagin-Thom construction. The resulting pushforward maps depend on the various choices.
Remark 3.6.
However, using the assumption that is rationally even, in the case where is a submersion the situation is simple. First of all, the relative tangent bundle makes sense, and we restrict our attention to the case where we are given a differential stable -structure on (as opposed to the more general notion of differential stable relative tangential -structure on in [YY21, Definition 5.12]). Then, associated to such there is a canonical set of choices of which gives the same map (3.3). We explain this point in details in Appendix A. We call such a lift of (Definition A.56). The map (3.3) defined by any choice of a lift of is the unique map denoted by
(3.7) |
We simply call it the differential pushforward map (Definition A.52 and Proposition A.57).
In the case where is a submersion and equipped with a differential stable -structure on , there is also the corresponding pushforward map on the level of differential forms. The Chern-Dold character (1.8) of the multiplicative genus is the element
(3.8) |
For example, for , the Chern-Dold chacacter is trivial, . For and , the Chern-Dold characters are the Todd polynomial and the polynomial, respectively. Applying the Chern-Weil construction in (2.8), we get the Chern-Dold character form for the relative tangent bundle,
(3.9) |
Using this, the pushforward map on is given by
(3.10) |
Restricted to the closed forms, the induced homomorphism on the cohomology, , is compatible with the Chern-Dold character (1.8) for and the topological pushforward (3.1). The differential pushforward map in (3.7) is compatible with the map (3.10) (tangential version of (A.27)).
In particular, if , for every -dimensional differential stable tangential -cycle over , the differential pushforward map (3.7) is
(3.11) |
As we explain in the last part of Appendix A.1, an important property of the pushforward is the following Bordism formula, relating the pushforward of differential forms (3.10) on the bulk and the differential pushforward (3.11) on the boundary.
Fact 3.12 (Bordism formula, [Bun12, Problem 4.235]).
For any morphism in , the following diagram commutes.
Example 3.13.
In the case , the nontrivial degree of pushforwards are called the higher holonomy function which appears in the definition of Chern-Simons invariants. In terms of the Cheeger-Simons model of in terms of differential characters [CS85], it is given by the evaluation on the fundamental cycle. In particular for the case it is the usual holonomy, and the Bordism formula is satisfied because of the relation between the curvature and the holonomy for -connections.
Example 3.14.
In the case , Freed and Lott [FL10] constructed a model of in terms of hermitian vector bundles with hermitian connections, and the refinement of pushforwards when is odd, , is given by the reduced eta invariants. The Bordism formula is a consequence of the Atiyah-Patodi-Singer index theorem.
3.2. Differential Pushforwards in terms of functors
As a preparation to the main Subsection 3.3, in this subsection we translate the data of differential pushforwards into functors from .
Definition 3.15.
In the above settings, let be a manifold, be an integer and . Let be an integer with . Then define the functor of Picard groupoids,
(3.16) |
by the following.
-
•
On objects, we set
(3.17) - •
The well-definedness of the functor follows by the Bordism formula in Fact 3.12.
As is easily shown by the Bordism formula, the formula (3.17) defines the homomorphism
(3.18) |
As expected, the transformation (3.16) is induced by the first arrow in (3.27). To show this, first remark that for any spectrum and its any fundamental cycle , the forgetful functor gives the equivalence of Picard groupoids,
(3.19) |
where the left hand side means the simplicial fundamental groupoid, whose objects are differential functions , and morphisms are bordism classes of differential functions . The right hand side is the fundamental groupoid for the space , which is equipped with the structure of a Picard groupoid by [HS05, Example B.7].
We have a functor of Picard groupoids555The right hand side is the Picard groupoid associated to the homomorphism of abelian groups. In general, a homomorphism between abelian groups associates a Picard groupoid , whose objects are elements of , and a morphism from to is given by an element such that . ,
(3.20) |
given by assigning the element for an object and the integration of the curvature for a morphism.
Proposition 3.21.
Proof.
Choose a differential function representing . For each object in , choose a Hopkins-Singer’s differential -structure lifting . By the discussion in Appendix A.2 and its tangential variant in Appendix A.3, we get a functor
(3.23) |
Indeed, for objects, given with the chosen lift , denote the underlying embedding and tubular neighborhood by . We have differential functions and . Applying the (-version of the) left vertical arrow of (A.40) to them and using the open embedding (the Pontryagin-Thom collapse), we get the differential function .
For morphisms , choose any representative and smooth map (not necessarily a submersion) so that it coincides with a collar coordinates of each objects near the endpoints, respectively. The structure induces , in particular the topological structure , on . Take any Hopkins-Singer’s differential tangential -oorientation (Appendix A.3) for which coincides with the chosen lifts at the boundary, and whose underlying map classifies . Then applying the same procedure as that we did for objects above, we get a differential function which restricts at the boundary to the ones assigned to objects above. Since any of the choices we have made is unique up to bordisms, the resulting morphism in the right hand side of (3.23) is uniquely determined. This gives the desired functor.
3.3. The construction and the proof
In this main subsection, we state and prove the main result of this article. Let be a multiplicative tangential structure groups, be a rationally even ring spectrum and be a homomorphism of ring spectra. Fix an integer such that . We construct a natural transformation
(3.25) |
on (Definition 3.36).
The main result of this paper is the following topological characterization of the transformation (3.25).
Theorem 3.26 (=Theorem 1.2).
In the above settings, let be a manifold and be an integer. For and , the element coincides with the following composition,
(3.27) |
Here we denoted .
As a preparation, we show that there exists a canonical homomorphism666 The existsnce of a canonical pairing is used in [FMS07, Proposition 6], in particular in the last arrow of the second line of the proof of that proposition. They do not state any condition on , but they use the assumption implicitely.
(3.28) |
where denotes the group of chain maps of complexes of abelian groups. Indeed, by [HS05, (4.57)], we have a canonical isomorphism777 This isomorphism does not follow from the axiom of differential cohomology theory in [BS10]. For more on this point, see [BS10, Section 5].
(3.29) |
Here, for any abelian group , the cohomology theory is represented by the spectrum , where is the Moore spectrum. As explained there, this is because the differential function complexes can be fits into the homotopy Cartesian square [HS05, (4.12)]. Applied to and , we get the identification
(3.30) |
An element induces the element for any , and using and , we get the induced homomorphisms on , which we also denote as
(3.31) | ||||
(3.32) |
The homomorphism (3.31) coincides with the one obtained by the map in (2.1). The homomorphism (3.28) is given by mapping to the pair . The well-definedness follows by the construction.
On the other hand, by [HS05, Corollary B.17]888 Note that there is an obvious typo of the degree in the statement of [HS05, Corollary B.17]. (also see [YY21, Fact 2.6] and the explanations there), we have an isomorphism for any spectra ,
(3.33) |
where means the group of natural isomorphism classes of functors of Picard groupoids. By (3.19), (3.20) and (3.33), we get a homomorphism
(3.34) |
It directly follows from the definition of the identification (3.29) that we have
(3.35) |
Definition 3.36 ().
In the settings explained in the beginning of this subsection, for each manifold we define a homomorphism of abelian groups
(3.37) |
by the following. For and , set , where
-
•
The element is obtained by applying (3.31) on the coefficient of .
- •
The fact that the pair satisfies the compatibility condition follows from the well-definedness of (3.28) and the fact that in Definition 3.15 is a functor.
Now we prove Theorem 3.26.
Proof of Theorem 3.26(=Theorem 1.2).
We use the argument in [YY21, Subsection 4.2]. Recall that, for an element we associated a functor in [YY21, (4.46)]. In the proof of [YY21, Theorem 4.51], the natural isomorphism
(3.38) |
where for the former we use the model of by [HS05, Corollary B.17], was given as follows. Using the equivalence (2.5), we have . The map (3.38) is given by mapping the isomorphism class of the functor to .
3.4. Examples
3.4.1. The holonomy theory (1) : [YY21, Example 4.54]
Here we explain the easiest example of the “Holonomy theory (1)” which appeared in [YY21, Example 4.54]. This corresponds to the case , is the usual orientation, and .
Recall that, given a manifold and a hermitian line bundle with unitary connection over , we get the element
(3.40) |
On the other hand, in the case we have the canonical choice of an element in , namely the Anderson self-duality element . Thus we have the homomorphism
Using the model of in terms of hermitian vector bundles with -connections (for example see [HS05, Example 2.7]), the pair defines a class . We have the following.
Proposition 3.41.
We have the following equality in ,
(3.42) |
Moreover, the element coincides with the following composition,
Proof.
The last statement follows from (3.42) and Theorem 3.26. The equality (3.42) follows from the fact that the self-duality element induces the canonical isomorphism and , together with the following well-known facts about (for example see [HS05, Section 2.4]). The element satisfies
and, given a map from an oriented -dimensional closed manifold , the pushforward
∎
3.4.2. The classical Chern-Simons theory : [YY21, Example 4.56]
Here we explain the classical Chern-Simons theory which appeared in [YY21, Example 4.56]. This is essentially a generalization of Subsection 3.4.1, corresponding to the case , is the usual orientation, and .
Recall that, given a compact Lie group and an element , the corresponding classical Chern-Simons theory of level is defined by choosing an -classifying object in the category of manifolds with principal -bundles with connections, and fixing an element lifting . Then we have the element
(3.43) |
whose equivalence class in does not depend on the lift .
Proposition 3.44.
The element coincides with the following composition,
Proof.
The classifying map induces an equivalence . Moreover, by the pullback of the universal connection it is refined to a functor of Picard groupoids,
(3.45) |
which is naturally isomorphic to the above one under the equivalences .
We have the element
(3.46) |
Recall that an element associates a functor by [YY21, (4.46)]. We claim that the functors associated to the elements (3.46) and (3.43) are related by
Indeed, this follows from the fact that the Chern-Simons invariants are given by the pushforward in differential ordinary cohomology [YY21, (4.58)]. Applying Theorem 3.26 to the element (3.46), we get the result. ∎
3.4.3. The theory of massive free complex fermions : [YY21, Example 4.62]
Here we explain the example of the theory on massive free complex fermions which appeared in [YY21, Example 4.62]. This example corresponds to the case , and .
Recall that, given a hermitian vector bundle with unitary connection over a manifold , we get an element
On the other hand, in the case we have the canonical choice of an element in , namely the self-duality element . Thus we have the homomorphism
(3.47) |
Using the model of in terms of hermitian vector bundles with unitary connections by Freed-Lott ([FL10]), we have the class .
Proposition 3.48.
We have the following equality in ,
(3.49) |
Moreover, the element coincides with the following composition,
Proof.
The last statement follows from (3.49) and Theorem 3.26. Denote the Bott element by . The equality (3.49) follows from the fact that the self-duality element induces the canonical isomorphism and , together with the following facts about in [FL10]. The element satisfies
and, given a map from an oriented -dimensional closed manifold with a physical Spinc-structure , the pushforward is given by
∎
3.4.4. An interpretation of Subsection 3.4.3 - Taking anomaly theories of free spinor field theories
Here we explain an interpretation of the result in Subsection 3.4.3 in terms of anomalies of free spinor field theories.
We briefly recall the explanation in [Fre19, Lecture 11] and [FH21, Section 9] about free spinor field theory and its anomalies. A real spinor representation of and a (contractible) choice of nonnegative symmetric invariant bilinear pairing determine an -dimensional possibly anomalous theory called the free real spinor field theory . The spinor representation gives an element .
An anomalous -dimensional field theory is formulated as a boundary theory of -dimensional invertible field theory called the associated anomaly theory, which is classified by . In this case the relevant structure group is . As explained in the references, the anomaly theory associated to has partition function given by suitable fraction of exponentiated reduced eta invariants, depending on mod .
Freed and Hopkins suggested the following conjecture.
Conjecture 3.50 ([Fre19, Conjecture 11.23], [FH21, Conjecture 9.70]).
The -dimensional anomaly theory associated to the -dimensional free real spinor field theory correponds to the element in given by the following composotion.
(3.51) |
Here is the Atiyah-Bott-Shapiro map and is the Anderson self-duality element for the -theory.
The composition 3.51 is the special case of the composition (3.27) for . But actually at this point we do not have a proof for Conjecture 3.50, since we do not have the complete understanding of the pushforward in differential -theory. Before explaining the details, let us explain the complexified version where we can actually show the corresponding statement using the result in Subsection 3.4.3.
In the complexified settings, we have the corresponding story. A complex spinor representation gives a class . In this case nontrivial classes appears only when is even, so we focus on this case.
Proposition 3.52 (Complex version of Conjecture 3.50).
The -dimensional anomaly theory associated to the -dimensional free complex spinor field theory correponds to the element in given by the following composotion.
(3.53) |
Proof.
We apply the result in Subsection 3.4.3 for . In this case, we simply have and . We know from Proposition 3.48 that, in the case is the generator, the composition (3.53) equals to the element
(3.54) |
This is indeed the anomaly theory for , whose partition function is given by the exponentiated eta invariants. ∎
As we see from the proof, our result can be useful even when . The general case of nontrivial can be regarded as giving the parametrized version. Also we see that the proof uses the knowledge of pushforward in differential -theory as reduced eta invariants.
Let us go back to the real case. We can apply Theorem 1.2 to deduce that the composition (3.51) gives the element999 In the case , we have . In other cases, since is or , we have canonically.
(3.55) |
The remaining problem is to understand this element, which is equivalent to understanding the pushforward in differential -theory. As far as the author is aware of, we do not have the enough understanding of this pushforward to verify Conjecture 3.50.
Remark 3.56.
In the examples in this subsection, we used the Anderson self-duality elements in for . However, the results in this subsection do not use the self-duality, and indeed there are many other interesting examples given by non-self-duality elements in . For example, in the analysis of anomalies of the heterotic string theories in [TY21], we encounter such examples when and with the Witten genus and .
Appendix A Differential pushforwards for proper submersions
As mentioned in Subsection 3.1, there are certain subtleties regarding the formulations of differential pushforwards. In this appendix, we explain that there is a nice theory on differential pushforwards for proper submersions under the assumption that is rationally even. The author believe that the results in this Appendix well-known among experts. It is convenient to start with multiplicative differential extensions which are not necessarily the one given by the Hopkins-Singer. The minimal requirements for the differential extension are,
-
•
For real vector bundles over manifolds, the properly supported differential cohomology groups
(A.1) are defined with a module structure over , so that they refine properly supported cohomologies and forms.
-
•
If we have a vector bundle and we have an open embedding in the total space of another vector bundle , we have the corresponding map
refining the topological and form counterparts.
-
•
We have the desuspension map,
refining the topological and form counterparts.
Since we are assuming is rationally even, the Hopkins-Singer’s differential extension admits a canonical multiplicative structure by [Upm15], and the above properties are also satisfied.
A.1. The normal case
In this subsection we explain the normal case. The content of this subsection basically follows the unpublished survey by Bunke [Bun12, Section 4.8–4.10]. Let and be multiplicative with rationally even, and assume we are given a homomorphism of ring spectra,
(A.2) |
where is the Thom spectrum. Then for each real vector bundle of rank over a topological space equipped with a stable -structure , we get the Thom class , where we denote . Its multiplication gives the Thom isomorphism . Its Chern-Dold character is an element . We set
Definition A.3 (Differential Thom classes, , homotopy).
Let be a smooth real vector bundle over a manifold of rank equipped with a stable -structure .
-
(1)
A differential Thom class is an element such that is the Thom class for .
-
(2)
For such a , we define
(A.4) -
(3)
A homotopy between two differential Thom classes and is a differential Thom class for with for such that
(A.5) The homotopy class of is denoted by .
In particular, if and are homotopic, we have . Thus we use the notation .
Lemma A.6.
Let and be as before, and be the Thom class for . Assume we are given an element such that .
-
(1)
There exists a differential Thom class with .
-
(2)
The set of homotopy classes of differential Thom classes with is a torsor over
(A.7)
Proof.
If is equipped with a stable differential -structure , applying the Chern-Weil construction (2.8) to , we have
(A.8) |
This satisfies .
For of rank represented by with , we associate a differential stable -structure on which is represented by . For a topological stable -structure , we define in the same way.
If we have a homotopy class of diffential Thom classes for , the integration
defines a well-defined homotopy class of differential Thom classes for . Moreover, by Lemma A.6, the above integration gives a bijection between the sets of homotopy classes of diffential Thom classes for and for .
Proposition A.9.
101010 In the proof we use the assumption that is rationally even. However, by a small modification of the proof, this assumption can be weakened to . As a result, the results in this subsection hold under this weaker condition. The same remark applies to Proposition A.47.There exists a unique way to assign a homotopy class of differential Thom classes to every real vector bundle with differential stable -structure such that the following three conditions hold.
-
(1)
It is compatible with pullbacks.
-
(2)
We have .
-
(3)
We have .
Moreover, the resulting homotopy class only depends on the homotopy class (Definition 2.3 (4)) of differential stable -structure .
Proof.
By the condition (2), it is enough to consider only such that is represented by a representative of the form , i.e., without stabilization.
The proof basically follows that for [Bun12, Problem 4.197]. Suppose we have of rank over with with a representative . Take a manifold with an -connected map . We can factor the classifying map for as with smooth. Take a -connection on the pullback of the universal bundle, and denote by the resulting differential -structure on by . We have maps and covering . We may assume that pulls back to by .
The difference of any two choices of the homotopy classes of differential Thom classes on is measured by an element in by Proposition A.6. The pullback map is zero because is -connected and we have since is rationally even. Thus, taking any homotopy class of differential Thom classes for , the pullback
(A.10) |
defines a homotopy class of differential Thom classes for which does not depend on the choice of . By the condition (1) and (2), we are forced to define the required homotopy class as
(A.11) |
by taking any on .
We need to check that the element (A.11) does not depend on the other choices made above. But this easily follows from the cofinality of such choices. Namely, given two choices with the underlying manifolds for , we may take another with maps so that , and other data on which pulls back to those given on . From this, we conclude that the elements (A.11) defined using and coincide with the one defined using , so the element (A.11) is well-defined. By the arguments so far, they satisfy the required conditions and the uniqueness.
For the last statement, changing a differential stable -structure on to a homotopic one amounts to changing the vector bundle map by a homotopy while fixing and in the above procedure. Pulling back the homotopy class by such a homotopy, we get a homotopy of differential Thom classes between the differential Thom classes pulled back at the endpoints. This completes the proof. ∎
Now we turn to differential pushforwards for proper submersions. Let be a proper submersion between manifolds of relative dimension , and assume it is equipped with a differential stable normal -structure (Definition 2.4) on the relative tangent bundle . Take a representative of . It induces a differntial stable -structure on which we denote , represented by . By Proposition A.9 we have a differential Thom class whose homotopy class is canonically determined,
(A.12) |
If we stabilize to , the homotopy classes of (A.11) are related as Proposition A.9 (2).
Now, choose an embedding over (i.e., ) for large enough, a tubular neighborhood of in with a vector bundle structure so that it is a map over (this is possible because is a submersion). Replacing larger if necessary, choose an isomorphism of vector bundles over so that the isomorphism is homotopic to . The isomorphism induces a differential stable -structure on , and the element (A.12) induces a differential Thom class on denoted by
(A.13) |
We consider the composition,
(A.14) |
where the first map uses the module structure of the properly supported , and the middle arrow is induced by the open embedding .
Proposition A.15.
The composition (A.14) only depends on the differential stable normal -structure on .
Proof.
The above procedure includes the following ambigiuities : the choice of representing and the choice of the data of embedding with a tubular neighborhood and an isomorphism . The independence on directly follows from the last statement of Proposition A.9.
First we show the independence on the choice of , with the other data fixed. Since its homotopy class is fixed by Proposition A.9, any two choices , , are connected by a homotopy . Its pullback by gives a homotopy between the corresponding differential Thom classes on . Denote the inclusion by for . Consider the following commutative diagram,
(A.20) |
The commutativity of the middle square is because the vector bundle structure is a map over . The commutativity of the right square is by the homotopy formula ([BS10, Lemma 1]).
Take any element . Then the image of under the composition of the bottom arrows in (A.20) is equal to the difference of the elements in obtained by applying to the composition (A.14) using and . By the commutativity of (A.20), it is enough to check that the element maps to zero under the composition of the top arrows in (A.20). Indeed, since is a map over , we can factor the upper middle horizontal integration in (A.20) on , and the result is equal to
(A.21) |
and by (recall (A.5))
so (A.21) is equal to
as desired. Thus we conclude that, fixing the data of an embedding with a tubular neighborhood, the composition (A.14) only depends on the homotopy class .
Now consider the stabilization of the embeddings, increasing to and to . By the condition (2) in Proposition A.9 and the result so far, we also conclude that the composition (A.14) is invariant under this stabilization.
The desired independence of (A.14) on the remaining choices is also proved in a parallel way, by choosing corresponding objects on the cylinder so that they restrict to stabilizations of the given ones on the endpoints. This completes the proof of Proposition A.15.
∎
Thus we define the following.
Definition A.22.
By the construction, the following diagram commutes.
(A.27) |
In this sense, Definition A.22 refines the pushforwards on and .
An important property of differential pushforwards is the Bordism formula [Bun12, Problem 4.235], which says that if we have a bordism , the differential pushforwards at the boundary can be computed by the integration of the characteristic form on the bordism. Its normal variant is stated in the form we use in this paper as Fact 3.12. To prove it, we need to consider differential pushforwards for proper maps which is not submersions, namely boundary defining functions . The result easily follows by the homotopy formula ([BS10, Lemma 1]). For the details of the proof we refer [Bun12, Problem 4.235].
A.2. Differential pushforwards in Hopkins-Singer’s differential extensions
Now we turn to the Hopkins-Singer’s differential extensions. As we explain, the definition of differential pushforwards in [HS05] differs from the one in Subsection A.3. In this subsubsection, we clarify their relation in the settings of our interest (Proposition A.45).
Fix fundamental cocycles and for and , respectively. Since is rationally even, the Hopkins-Singer’s model admits a canonical multiplicative structure by [Upm15]. We briefly explain it here. We only explain the even-degrees. The remaining cases are induced by requiring the compatibility with the -integration. Let and be even integers, and denote by a multiplication map. We need to choose a reduced cochain so that
(A.28) |
Since is rationally even, we have by the proof of [BS10, Lemma 3.8]. Thus any two choices of such cochains differ by a coboundary. Using we get the map of differential function spaces ([HS05, Remark 4.17]),
(A.29) |
for any manifold . Also choose a natural cochain homotopy cobounding the difference between on forms and on singular cochains as in [HS05, (3.8)], [Upm15, Section 6]. Any two such choices are naturally cochain homotopic. It induces the map
(A.30) |
for any . Combining (A.29) and (A.30), we get the multiplication map,
(A.31) |
This does not depend on any of the choices above. For a real vector bundle , in the same way we get a map using the properly supported differential functions ([HS05, Section 4.3])
(A.32) |
which gives the module structure,
(A.33) |
As we mentioned in Subsection 3.1, Hopkins-Singer’s normal differential -orientations are defined in terms of differential functions to . A differential pushforward is defined by fixing a map of differential function spaces
(A.34) |
whose underlying map factors as . Here is induced by , so that represents . We can take a so that , and it is determined up to coboundary because . We may take (A.34) to be the composition
(A.35) |
where the first map uses . Let be a real vector bundle, and consider the following diagram.
(A.40) |
Here the top horizontal arrow uses , the left vertical arrow uses the cochain homotopy , and the remaining arrows are as before. The two triangles commute. The square does not commute on the level of differential function spaces, but we can easily check that the difference is a coboundary so the induced maps on the differential cohomology level,
(A.41) |
are the same. Using the top factorization of (A.40), we see that (A.41) factors as
(A.42) |
To put them into the picture in Subsection A.1, apply the discussions there in the case and . Assume that we have a proper submersion equipped with a differential stable normal -structure on the relative tangent bundle , represented by . A Hopkins-Singer’s normal differential -orientation ([HS05, Section 4.9.2]) consists of choices of an embedding over , a tubular neighborhood with a vector bundle structure , an isomorphism as in Subsection A.1 (in general is not required to be a map over ), and a lift of a classifying map for the induced -structure on to a differential function . Then, the differential function represents a differential Thom class for ,
(A.43) |
where we denoted by the differential cohomology class represented by a differential function . Now we define the following.
Definition A.44.
Let be a proper submersion between manifolds of relative dimension , and assume it is equipped with a differential stable normal -structure on . A Hopkins-Singer’s normal differential -orientation is said to be a lift of if, in the notations above,
-
•
The vector bundle structure is a map over , and
-
•
The homotopy class of the differential Thom class is the one associated to by Proposition A.9 (applied to ).
In particular, this means that,
where .
Now assume we are given an element . Then, in [HS05, Section 4.10] the differential pushforward of is formulated as follows. Take a differential function representing . Apply the left bottom composition of (A.40) for the vector bundle to the pair to get a differential function in . By the open embedding we get a differential function in , and this represents the desired element . By the discussion so far, the result is the same if we use the top right composition in (A.40), and it is given by the composition (A.42). Then we see that the above definition of exactly translates into the definition of differential pushforwards (A.14) in the Subsection A.1. Thus we conclude that
A.3. The tangential case
Now we explain the tangential variants of the last Subsections A.1 and A.2. The constructions and verifications are parallel to the normal case, so we go briefly.
In this case, we are given a homomorphism of ring spectra,
(A.46) |
where is the Madsen-Tillmann spectrum. is constructed as a direct limit of Thom spaces of stable normal bundles to the universal bundles over approximations of ’s, so classifies vector bundles with stable normal -structures. Then for each real vector bundle of rank over a topological space equipped with a topological stable normal -structure , we get the Thom class , whose multiplication gives the Thom isomorphism .
We formulate the notion of differential Thom classes as a differential refinements of the Thom classes, as well as differential forms and homotopies in the same way as Definition A.3. By the exactly the same proof, the classification result of differential Thom classes corresponding to Lemma A.6 also holds in the case here. The Chern-Dold character for (A.46) is an element . If is equipped with a stable normal -structure , the characteristic form (A.8) is replaced by the form , where we use the Chern-Weil construction in (2.9). Then, the same proof as that of Proposition A.9 shows the following.
Proposition A.47.
There exists a unique way to assign a homotopy class of differential Thom classes to every real vector bundle with differential stable normal -structure such that the following three conditions hold.
-
(1)
It is compatible with pullbacks.
-
(2)
We have .
-
(3)
We have .
Moreover, the resulting homotopy class only depends on the homotopy class (Definition 2.4 (4)) of differential stable normal -structure .
Let be a proper submersion between manifolds of relative dimension , equipped with a differential stable -structure on the relative tangent bundle represented by . Choose an embedding over for large enough, a tubular neighborhood of in with a vector bundle structure so that it is a map over (this is possible because is a submersion). Then we get an isomorphism
(A.48) |
of vector bundles over . As a result, we get a differential stable normal -structure on the vector bundle , represented by . For , Proposition A.47 assigns a differential Thom class whose homotopy class is canonically determined,
(A.49) |
We consider the composition,
(A.50) |
The following proposition can be shown in the same way as Proposition A.15.
Proposition A.51.
The composition (A.50) only depends on the differential stable -structure on .
Proposition A.51 allows us to define the following.
Definition A.52.
Let be a proper submersion between manifolds of relative dimension , equipped with a differential stable -structure on the relative tangent bundle . We define the differential pushforward map,
to be the composition (A.50).
Now we turn to the Hopkins-Singer’s models as in Subsection A.2. Take fundamental cocycles and for and , respectively. As explained there, admits a canonical multiplicative structure. In the normal case, a differential pushforward is defined by a map of differential function spaces
(A.53) |
whose underlying map factors as . By the same argument to the normal case, the map (A.53) induces the map of differential cohomologies for any real vector bundle ,
(A.54) |
Let be a proper submersion between manifolds of relative dimension , equipped with a differential stable -structure on the relative tangent bundle represented by . A Hopkins-Singer’s tangential differential -orientation consists of choices of an embedding , a tubular neighborhood , a vector bundle structure as in the first part of this subsection (in general is not required to be a map over ), and a lift of a classifying map for of the induced normal structure to a differential function . Then, the differential function represents a differential Thom class for ,
(A.55) |
Now we define the following.
Definition A.56.
In the above settings, Hopkins-Singer’s tangential differential -orientation is said to be a lift of if, in the notations above,
-
•
The vector bundle structure is a map over , and
-
•
The homotopy class of the differential Thom class is the one associated to by Proposition A.47 (applied to ).
In particular, this means that,
where .
Let us take . By the same procedure as in the last paragraph of Subsection A.2, the tangential variant of [HS05, Section 4.10] using the map (A.53) and the open embedding produces the element . We get
Proposition A.57.
Acknowledgment
The author is grateful to Kiyonori Gomi and Kazuya Yonekura for helpful discussion and comments. She is supported by Grant-in-Aid for JSPS KAKENHI Grant Number 20K14307 and JST CREST program JPMJCR18T6.
References
- [BS10] U. Bunke and T. Schick, Uniqueness of smooth extensions of generalized cohomology theories, J. Topol. 3 (2010) 110–156.
- [BS12] by same author, Differential K-theory: a survey, Global differential geometry, Springer Proc. Math., vol. 17, Springer, Heidelberg, 2012, pp. 303–357. https://doi.org/10.1007/978-3-642-22842-1_11.
- [BSSW09] U. Bunke, T. Schick, I. Schröder, and M. Wiethaup, Landweber exact formal group laws and smooth cohomology theories, Algebr. Geom. Topol. 9 (2009) 1751–1790.
- [Bun12] U. Bunke, Differential cohomology, 2013. arXiv:1208.3961 [math.AT].
- [CS85] J. Cheeger and J. Simons, Differential characters and geometric invariants, Geometry and topology (College Park, Md., 1983/84), Lecture Notes in Math., vol. 1167, Springer, Berlin, 1985, pp. 50–80. https://doi.org/10.1007/BFb0075216.
- [FH21] D. S. Freed and M. J. Hopkins, Reflection positivity and invertible topological phases, Geom. Topol. 25 (2021) 1165–1330.
- [FL10] D. S. Freed and J. Lott, An index theorem in differential -theory, Geom. Topol. 14 (2010) 903–966.
- [FMS07] D. S. Freed, G. W. Moore, and G. Segal, The uncertainty of fluxes, Comm. Math. Phys. 271 (2007) 247–274.
- [Fre19] D. S. Freed, Lectures on field theory and topology, CBMS Regional Conference Series in Mathematics, vol. 133, American Mathematical Society, Providence, RI, 2019. https://doi.org/10.1090/cbms/133. Published for the Conference Board of the Mathematical Sciences.
- [GS21] D. Grady and H. Sati, Differential KO-theory: constructions, computations, and applications, Adv. Math. 384 (2021) Paper No. 107671, 117.
- [HS05] M. J. Hopkins and I. M. Singer, Quadratic functions in geometry, topology, and M-theory, J. Differential Geom. 70 (2005) 329–452.
- [Klo08] K. R. Klonoff, An index theorem in differential K-theory, ProQuest LLC, Ann Arbor, MI, 2008. http://gateway.proquest.com/openurl?url_ver=Z39.88-2004&rft_val_fmt=info:ofi/fmt:kev:mtx:dissertation&res_dat=xri:pqdiss&rft_dat=xri:pqdiss:3321109. Thesis (Ph.D.)–The University of Texas at Austin.
- [Rud98] Y. B. Rudyak, On Thom spectra, orientability, and cobordism, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 1998. With a foreword by Haynes Miller.
- [TY21] Y. Tachikawa and M. Yamashita, Topological modular forms and the absence of all heterotic global anomalies, 2021. arXiv:2108.13542 [hep-th].
- [Upm15] M. Upmeier, Algebraic structure and integration maps in cocycle models for differential cohomology, Algebr. Geom. Topol. 15 (2015) 65–83.
- [YY21] M. Yamashita and K. Yonekura, Differential models for the Anderson dual to bordism theories and invertible QFT’s, I, arXiv:2106.09270 [math.AT].