1 Introduction
In the present paper, we study the following two topics. First, we give a formulation of a deformation of Dirac operator along orbits on a possibly noncompact manifold equipped with a group action to get an equivariant index and a K-homology cycle representing the index. Second, we apply this framework to Hamiltonian torus manifolds to define geometric quantization from the viewpoint of index theory. In particular, we give proofs of a [Q,R]=0 type theorem and a Danilov-type formula for the toric case in the possibly noncompact setting. The proofs are based on the same perspective, taken in [Reference Fujita8, Reference Fujita, Furuta and Yoshida11] by the author and joint works with Furuta and Yoshida, namely, the localization of index to lattice points. These results give a simplification and a generalization of [Reference Fujita8, Reference Fujita, Furuta and Yoshida11]. They also make more clear the relation with a similar construction in [Reference Braverman6].
Geometric quantization of symplectic manifolds originates from ideas in physics. However, nowadays, it is related to several topics in various branches of mathematics. One of them is the index theory of Dirac operator. In fact, in some cases, the quantization can be regarded as an index of the spin
$^c$
Dirac operator associated with a compatible almost complex structure. This approach is called spin
$^c$
quantization. Studying quantization from the viewpoint of index theory, K-theory, K-homology, and KK-theory is an active area of research.
Geometric quantization in the compact setting has been extensively studied. The noncompact case has also been studied to some extent. For example, such a generalization is important for quantization of Hamiltonian loop group space in [Reference Loizides and Song18]. In addition, the noncompact setting plays an essential role to obtain localization phenomena in geometric quantization as below. On the other hand, unlike the compact manifold case, the index of Dirac operator on a noncompact or open manifold is not well defined in a straightforward way. To get the index in a possibly generalized sense, it is necessary to take an appropriate boundary condition or to consider additional structure such as a fiber bundle structure or a nice group action.
In [Reference Braverman6], Braverman gave a formulation to define an equivariant index in a noncompact setting. This framework originates in a proof of [Q,R]=0 in [Reference Tian and Zhang26] and was applied to a solution of the Vergne conjecture in [Reference Ma and Zhang20]. He used a deformation of the Dirac operator by the Clifford action of the vector field generated by the moment map.Footnote 1 On the other hand, in a series of papers [Reference Fujita, Furuta and Yoshida9–Reference Fujita, Furuta and Yoshida11] with Furuta and Yoshida, the author developed an index theory on open manifolds using a family of partly defined fiber bundle structures and a deformation of Dirac operator. The deformation in [Reference Fujita, Furuta and Yoshida9–Reference Fujita, Furuta and Yoshida11] is given by first-order differential operators, a family of Dirac operators along fibers, which need not use a group action essentially. We call it FFY’s deformation for short. Both Braverman’s and FFY’s deformations are motivated by Witten’s pioneering work [Reference Witten27], and in the equivariant case, these deformations have the same nature, that is, a deformation along the orbits. Both of the resulting indices satisfy the excision formula, which leads us to the localization of index. Here, we summarize the differences between Braverman’s and FFY’s deformations.
-
• Braverman’s deformation:
-
(1) can be applied to compact group actions (not necessarily abelianFootnote 2 ), and
-
(2) realizes a localization of index to the zero level set of the moment map and fixed points (or critical points of the norm square of the moment map).
-
-
• FFY’s deformation:
-
(1) can be applied to torus fibrations (e.g., Lagrangian torus fibrations), and
-
(2) realizes a localization of index to the inverse images of the lattice points (or Bohr–Sommerfeld fibers).
-
As an application of the FFY’s second point above, a geometric proof of [Q,R]=0 for the torus action case based on the localization of index is obtained in [Reference Fujita, Furuta and Yoshida11]. There is another application in [Reference Fujita8], which gives a proof of Danilov’s formula. Danilov’s formula can be regarded as a localization of the geometric quantization of toric manifolds to lattice points in the momentum polytope. The proof in [Reference Fujita8] realizes such a picture of localization faithfully.
In the present paper, we give a framework of a deformation of Dirac operator in a similar manner as in the torus-equivariant setting for FFY’s deformation. We use a single differential operator along orbits for the deformation, which satisfies some acyclicity and boundedness condition. We call it an acyclic orbital Dirac-type operator (Definitions 2.1 and A.1). Though it is similar to the acyclic compatible system in [Reference Fujita, Furuta and Yoshida9] or [Reference Fujita, Furuta and Yoshida10], the definition of the acyclic orbital Dirac-type operator is much simpler due to the presence of the global torus action and the isotypic component decomposition of the space of sections. Another difference is that the deformation by an acyclic orbital Dirac-type operator gives an transversally elliptic operator in the sense of Atiyah [Reference Atiyah3]. We summarize our first main results:
Theorem 1.1 (Theorem 2.4 and Corollary 2.5) Under suitable technical assumptions, we can construct an acyclic orbital Dirac-type operator, which gives an equivariant index valued in the formal completion of the representation ring and a natural K-homology cycle representing the index.
The above acyclic orbital Dirac-type operator (Definition 2.1) is a combination of Kasparov’s orbital Dirac operator [Reference Kasparov14] and Braverman’s deformation term, which, in fact, becomes the Braverman-type Clifford action shifted by a weight when it is restricted to each isotypic component. The second main result is the following.
Theorem 1.2 (Theorem 2.10) Under suitable technical assumptions, the equivariant index defined by the acyclic orbital Dirac-type operator coincides with the equivariant index defined by Braverman’s deformation.
As a corollary of Braverman’s index theorem in [Reference Braverman6], our equivariant index is also equal to Atiyah’s transverse index in [Reference Atiyah3] under the same assumptions.
Finally, we apply the above construction to the setting of noncompact Hamiltonian torus manifolds with possibly noncompact fixed-point sets, allowing us to define the spin
$^c$
quantization of it as an equivariant index (Definition 5.1). Our quantization has a localization property to integral lattice points due to its origin. The third main result is the following.
Theorem 1.3 (Theorems 5.3 and 5.5) For the quantization of Hamiltonian torus manifolds defined by an acyclic orbital Dirac-type operator, we have proofs of the following
$:$
-
(1)
$[Q,R]$ =0 theorem for integral regular values of the circle action case, and
-
(2) a Danilov-type formula for toric case.
The proofs of the above theorems apply also to the compact case, giving simple alternative proofs for [Reference Fujita8, Reference Fujita, Furuta and Yoshida11].Footnote
3
Because our equivariant index can be identified with Atiyah’s transverse index, the proof of the first statement in Theorem 1.3 gives an alternative proof of the Vergne conjecture in [Reference Ma and Zhang20]. In the toric case, the lattice points in the momentum polytope are closely related to the geometric quantization obtained by a real polarization. There are several results concerning the coincidence between the spin
$^c$
(or Kähler) quantization and the quantization based on the real polarization from the viewpoint of the index theory. For example, see [Reference Andersen1, Reference Fujita, Furuta and Yoshida9, Reference Kubota15, Reference Yoshida28]. Theorem 5.5 can be regarded as such a coincidence in the noncompact setting.
This paper is organized as follows. In Section 2, we construct a K-acyclic orbital Dirac-type operator (Definition 2.1 and Theorem 2.4) for a complete manifold equipped with an action of a compact torus K. This operator arises naturally in the situation of Hamiltonian actions on symplectic manifold. In Section 2.3, we show that our equivariant index is equal to the equivariant index obtained by Braverman’s deformation (Theorem 2.10). In Section 3, we summarize the product formula in useful two ways (Propositions 3.3 and 3.10). Because the product formula itself can be obtained in the abstract framework of index theory of Fredholm operators, we just confirm our setup and statements. We also present two practical formulas, which have key roles in Section 5. In Section 4, we show a vanishing formula of index for fixed-point subsets (Theorem 4.2), which is also important in the construction in Section 5. In Section 5, by using the constructions and discussions in the previous sections, we define quantization of Hamiltonian torus manifolds as an equivariant index (Definition 5.1). For our quantization, we show [Q,R]=0 theorem (Theorem 5.3) and a Danilov-type formula for toric case (Theorem 5.5). The proofs are straightforward from the localization property of our index to lattice points and product formulas. In Section 6, we explain some future problems concerning quantization of Hamiltonian loop group spaces and a relation between the deformation and KK-product. In Appendix A, we give a general machinery to have an equivariant index and a K-homology cycle by using K-acyclic orbital Dirac-type operator. We show that a deformation by a K-acyclic orbital Dirac-type operator has a compact resolvent on each isotypic component of the space of
$L^2$
-sections (Corollary A.3), and hence, it gives an equivariant (K-Fredholm) index and a K-homology cycle in a natural way (Definition A.2). We also show that the resulting Fredholm index is equal to that obtained from a deformation using a large parameter instead of the proper function (Theorem A.5). This deformation is closer to the deformation studied in [Reference Fujita, Furuta and Yoshida9, Reference Fujita, Furuta and Yoshida10].
1.1 Notations
We fix some notations.
For a compact Lie group K, let
$\textrm {Irr}(K)$
be the set of all isomorphism classes of finite-dimensional irreducible unitary representations of K. We frequently do not distinguish an element
$\rho \in \textrm {Irr}(K)$
and its corresponding representation space. Each unitary representation
$\mathcal {H}$
of K has the K-isotypic component decomposition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu1.png?pub-status=live)
where each isotypic component
$\mathcal {H}^{(\rho )}$
is defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu2.png?pub-status=live)
We also use the similar notation
$A^{(\rho )}$
for the restriction of a K-equivariant linear map A to the isotypic component. The representation ring of K is denoted by
$R(K)$
, which is generated by
$\textrm {Irr}(K)$
. We denote its formal completion by
$R^{-\infty }(K)$
, namely
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu3.png?pub-status=live)
Note that
$R(K)$
can be identified with the subgroup consisting of finite support elements in
$R^{-\infty }(K)$
by taking the coefficients in each irreducible representation.
Let
$\mathcal {H}$
be a Hilbert space with inner product
$(\cdot , \cdot )$
, A and B self-adjoint operators on
$\mathcal {H}$
, which have common domain. We write
$A\geq B$
if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu4.png?pub-status=live)
for all
$u\in \mathcal {H}$
in the domain of A. If
$\mathcal {H}$
has a
$\mathbb {Z}/2$
-grading and A is an odd Fredholm operator with the decomposition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu5.png?pub-status=live)
according to the grading, then its
$\mathbb {Z}/2$
-graded Fredholm index is defined as the super-dimension of
$\ker (A)$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu6.png?pub-status=live)
Let M be a Riemannian manifold and
$W\to M$
a vector bundle over M equipped with a Hermitian metric
$\langle \cdot ,\cdot \rangle _W=\langle \cdot ,\cdot \rangle $
. This metric gives rise to an
$L^2$
-inner product on the space of compactly supported sections
$\Gamma _c(W)$
of W, which is denoted by
$(\cdot , \cdot )_W=(\cdot , \cdot )$
. The associated
$L^2$
-norm and
$L^2$
-completion are denoted by
$\|\cdot \|_W=\|\cdot \|$
and
$L^2(W)$
, respectively.
In this paper, we mean a generalized Dirac operator by a Dirac(-type) operator. Namely, for a vector bundle W over a Riemannian manifold M equipped with a structure of a Clifford module bundle over
$TM$
, a first-order differential operator D acting on
$\Gamma _c(W)$
is called a Dirac(-type) operator if D is a formally self-adjoint operator whose principal symbol is equal to the Clifford action on W. When W has a
$\mathbb {Z}/2$
-grading, we impose that a Dirac operator is an odd operator.
2 Acyclic orbital Dirac-type operator for torus action
2.1 Construction of
$\boldsymbol{D}_{\boldsymbol{K}}$
Let K be a compact torus with Lie algebra
${\mathfrak {k}}$
. We fix an inner product on
$\mathfrak {k}$
and identify
${\mathfrak {k}}^*={\mathfrak {k}}$
. We often identify
$\textrm {Irr}(K)$
with
$\Lambda ^*$
, where we put
$\Lambda :=\textrm {ker}(\textrm {exp}:\mathfrak {k}\to K)$
. Let M be a complete Riemannian manifold and W a
$\mathbb {Z}/2$
-graded Clifford module bundle over M. Suppose that K acts on M in an isometric way and the action lifts to W as a unitary action. Take a K-invariant Hermitian connection
$\nabla $
of W.
For
$\xi \in \mathfrak {k}$
, we denote the induced infinitesimal action of
$\xi $
on M by
$\underline {\xi }^M$
. Let
$\mathcal {L}_{\xi }:\Gamma (W)\to \Gamma (W)$
be the induced derivative defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu7.png?pub-status=live)
for
$s\in \Gamma (W)$
. Let
$\mu :M\to \textrm {End}(W)\otimes {\mathfrak {k}}^*$
be the map defined by Kostant’s formula:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqn1.png?pub-status=live)
Fix an orthonormal basis
$\{\xi _1, \ldots , \xi _n\}$
of
$\mathfrak {k}$
.
Definition 2.1 We define the orbital Dirac-type operator
$D_K:\Gamma _c(W)\to \Gamma _c(W)$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu8.png?pub-status=live)
Lemma 2.1 The orbital Dirac-type operator
$D_K$
satisfies the following conditions.
-
(1)
$D_K$ is a first-order self-adjoint differential operator, which contains only differentials along K-orbits.
-
(2)
$D_K$ anticommutes with the Clifford multiplication of the transverse direction to orbits. Namely, for any K-invariant function h on M, one has
$$ \begin{align*} D_Kc(dh)+c(dh)D_K=0. \end{align*} $$
-
(3) For any Dirac-type operator D acting on
$\Gamma (W)$ , the anticommutator
$$ \begin{align*} DD_K+D_KD \end{align*} $$
Proof (1) follows from the definition of
$D_K$
. (2) follows from the anticommutativity between
$c(dh)$
and
$c(\underline {\xi _i}^M)$
for any K-invariant function h. By using (2), one can show (3) by the computation of the anticommutator:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu11.png?pub-status=live)
▪
Remark 2.2 The differential term
$\displaystyle \sum _{i=1}^nc(\underline {\xi _i^M})\mathcal {L}_{\xi _i}$
in
$D_K$
is the orbital Dirac operator in the sense of Kasparov [Reference Kasparov14]. On the other hand, the multiplication term
$\displaystyle \sum _{i=1}^nc(\underline {\xi _i^M})\mu _{\xi _i}$
is equal to
$c(\underline {\mu })$
for
$\underline {\mu }:=\displaystyle \sum _{i=1}^n\underline {\xi _i^M}\mu _{\xi _i}$
, which gives the deformation studied by Braverman [Reference Braverman6]. For each
$\rho \in \textrm {Irr}(K)$
, one has
$\mathcal {L}_{\xi _i}=\sqrt {-1}\rho (\xi _i)$
on each isotypic component
$L^2(W_L)^{(\rho )}$
, and hence,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu12.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu13.png?pub-status=live)
where
$\underline {\rho }$
is the infinitesimal action induced by
$\rho \in \mathfrak {k}^*=\mathfrak {k}$
. In other words,
$D_K$
gives a kind of shift of Braverman’s deformation. We investigate the relation between our deformation and Braverman’s deformation in the next section.
For
$\rho \in \textrm {Irr}(K)$
, let
$Z_{\rho }:=\textrm {Zero}(\underline {\rho }-\underline {\mu })$
be the set of points in M at which the vector field
$\underline {\rho }-\underline {\mu }$
vanishes. Note that
$Z_\rho $
coincides with the set of critical points of
$|{\rho }-{\mu }|^2$
in M, and it contains
$M^K\cup \mu ^{-1}(\rho )$
. The above description of
$D_K$
implies the following.
Proposition 2.3 For
$x\in M$
and
$\rho \in \textrm {Irr}(K)$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu14.png?pub-status=live)
Let D be the Dirac operator acting on
$\Gamma (W)$
, which is defined by the connection
$\nabla $
. For each
$\rho \in \textrm {Irr}(K)$
, we put
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu15.png?pub-status=live)
Then, because
$(D_K^{(\rho )}|_{K\cdot x})^2$
is a strictly positive operator on
$\Gamma (W|_{K\cdot x})^{(\rho )}$
for any
$x\in V_\rho $
, there exists a constant
$C_{\rho , x}$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu16.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu17.png?pub-status=live)
hold for any
$s\in \Gamma (W_L|_{K\cdot x})^{(\rho )}$
.
Theorem 2.4 If the following conditions are satisfied, then
$(D_K,\{V_\rho \}_{\rho \in \textrm {Irr}(K)})$
is a K-acyclic orbital Dirac-type operator on
$(M,W)$
in the sense of Definition A.1.
-
(1) For each
$\rho \in \textrm {Irr}(K)$ , the critical point set
$Z_{\rho }$ is compact.
-
(2) There exists
$C>0$ such that
$$ \begin{align*}C^{-1} < \displaystyle\sum_{i=1}^n|\underline{\xi_i^M}| < C \end{align*} $$
-
(3) For each
$\rho \in \textrm {Irr}(K)$ , we have
$$ \begin{align*}\mathop{\mathrm{sup}}\limits\{C_{\rho,x} \ | \ x\in V_\rho\}<\infty. \end{align*} $$
-
(4) For each
$\rho \in \textrm {Irr}(K)$ , we have
$$ \begin{align*} \inf_{x\in V_\rho}\{\kappa \ | \ \kappa \ \textrm{is the minimum eigenvalue of} \ (D_K|_{K\cdot x})^2 \ \textrm{on } \ L^2(W|_{K\cdot x})^{(\rho)} \}>0. \end{align*} $$
In particular, if M has a cylindrical (resp. periodic) end and all the data have translationally invariance (resp. periodicity), then the conditions (2)–(4) are satisfied. Moreover, if there are two such data, then the product of them satisfies these conditions.
By using the above K-acyclic orbital Dirac-type operator, we have a family of deformations of the Dirac operator D,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu21.png?pub-status=live)
or
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu22.png?pub-status=live)
as in Corollary A.3, Definition A.2, and Corollary A.6. As a consequence of Theorem 2.4, Corollary A.3, Definition A.2, and Corollary A.6, we have the following.
Corollary 2.5 Under the condition in Theorem 2.4, the family of deformations
$\hat D_\rho ($
or
$D_{\rho , t})$
gives a K-equivariant index
$[\hat D]=[M,W, D_K]\in R^{-\infty }(K)$
and a K-homology cycle, which represents it.
For latter convenience, we investigate the case of Hermitian manifold in detail. We assume that the metric on M is induced from a K-invariant Hermitian structure
$(g, J)$
and the Clifford module bundle W is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu23.png?pub-status=live)
for a K-equivariant Hermitian line bundle with Hermitian connection
$(L,\nabla ^L)$
over M, where
$T_{\mathbb {C}}M=TM$
is the vector bundle regarded as a complex vector bundle by J. This W carries a structure of
$\mathbb {Z}/2$
-graded
$\textrm {Cl}(TM)$
-module bundle with the Clifford multiplication
$c:TM\to \textrm {End}(W)$
defined by the exterior product and its adjoint. In this case,
$\mu $
is a map to
$\mathfrak {k}^*$
determined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu24.png?pub-status=live)
and we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu25.png?pub-status=live)
For
$x\in M$
, let
$H^0(K\cdot x; L|_{K\cdot x})$
be the space of global parallel sections on
$(L, \nabla ^L)|_{K\cdot x}$
, which is a vector space of dimension at most one. Suppose that
$H^0(K\cdot x; L|_{K\cdot x})\neq 0$
and s is its nontrivial element, then we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu26.png?pub-status=live)
for all
$\xi \in \mathfrak {k}$
. This equation implies that
$\mu _\xi (x)$
is an integer for all
$\xi $
, and hence, we have the following.
Proposition 2.6 If
$H^0(K\cdot x; L|_{K\cdot x})\neq 0$
for
$x\in M$
, then we have
$\rho :=\mu (x)\in \Lambda ^*$
and
$H^0(K\cdot x; L|_{K\cdot x})=\mathbb {C}_{(\rho )}$
, where
$\mathbb {C}_{(\rho )}$
is the one-dimensional representation of K whose weight is given by
$\rho $
. In particular, if
$H^0(K\cdot x; L|_{K\cdot x})\neq 0$
, then we have
$K\cdot x\subset Z_\rho $
for
$\rho =\mu (x)\in \Lambda ^*$
.
Remark 2.7 If
$M=(M,\omega )$
is a symplectic manifold whose dimension is twice of the dimension of K, the K-action is an effective Hamiltonian torus action, and
$(L,\nabla ^L)$
is a prequantizing line bundle, i.e., the curvature form of
$\nabla ^L$
is equal to
$-\sqrt {-1}\omega $
, then the condition
$H^0(K\cdot x; L|_{K\cdot x}) \neq 0$
is equivalent to the Bohr–Sommerfeld condition for the orbit
$K\cdot x$
, which is essential in the geometric quantization by the real polarization.
2.2 Noncomplete case and localization formula
As we will mention in the end of Appendix A.4, the index associated with the K-acyclic orbital Dirac-type operator can be defined for noncomplete situation. For instance, suppose that the first conditionFootnote
4
in Theorem 2.4 is satisfied. We take a K-invariant compact submanifold
$X_\rho $
with boundary as a neighborhood of
$Z_{\rho }$
and attach a cylinder
$\partial X_\rho \times [0,\infty )$
to
$\partial X_\rho $
, so that we have a K-invariant complete Riemannian manifold
$\tilde X_\rho $
with K-invariant cylindrical end. Let
$\tilde \mu $
,
$\tilde W$
, and
$\tilde D_K$
be the extensions of
$\mu $
, W, and
$D_K$
on
$\tilde X_\rho $
such that they have translational invariance and
$\ker (\tilde D_K^{(\rho )}|_{K\cdot x})=\ker ((\tilde D_K^{(\rho )}|_{K\cdot x})^2)=\ker (|\underline {\rho }-\underline {\mu }|^2)=0$
for any
$x\in \partial X_\rho \times (0,\infty )$
. These data define a Fredholm operator on
$L^2(\tilde W)^{(\rho )}$
as in Corollary A.3. Though we agree that it is a little bit strange notation,Footnote
5
we denote this index by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqn2.png?pub-status=live)
We decompose
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu27.png?pub-status=live)
into the disjoint union of the connected components, where
$Z_{\rho ,\alpha }$
is a connected component other than
$\mu ^{-1}(\rho )$
. This description enable us to get more refined decomposition of (2.2) into the summation of local contributions from each component, which we denote by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu28.png?pub-status=live)
The excision formula implies the following localization formula.
Theorem 2.8 If the conditions in Theorem 2.4 are satisfied, then the index
$[\hat D]=[M, W, D_K]\in R^{-\infty }(K)$
defined by the K-acyclic orbital Dirac-type operator
$D_K$
satisfies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu29.png?pub-status=live)
for each
$\rho \in \textrm {Irr}(K)$
.
In Section 5, we discuss the case of Hamiltonian circle action and symplectic toric case. In these cases, one has the vanishing
$[Z_{\rho ,\alpha }]=0$
, and hence, we realize the localization of the index
$[M,W,D_K]$
into the lattice points
$\textrm {Irr}(K)=\Lambda ^*$
.
2.3 Relation with Braverman’s deformation
In [Reference Braverman6], Braverman studied a Witten-type deformation of the Dirac operator and its equivariant index on noncompact K-manifold. In a symplectic geometric setting, Braverman’s deformation is given by the Clifford multiplication of the Hamiltonian vector field of the norm square of the moment map. In particular, in the setting in Section 2.1 (not necessarily K is a torus), we can consider the Braverman’s deformation as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu30.png?pub-status=live)
where
$h:M\to \mathbb {R}$
is a K-invariant function called an
${\textit {admissible function}}$
, which satisfies a suitable growth condition. Braverman showed several fundamental properties of
$D_{\mu }$
. In particular, he showed that
$D_{\mu }$
is a K-Fredholm operator and the resulting index in
$R^{-\infty }(K)$
is independent of a choice of the admissible function. Moreover, the index is equal to Atiyah’s transverse index. After that, his equivariant index has been applied in several directions, for instance, a solution to Vergne’s conjecture by Ma and Zhang [Reference Ma and Zhang20].
In this section, we consider the same setup in Section 2.1 and assume the followings to make the situation simple.
Assumption 2.9 We assume that the conditions in Theorem 2.4 are satisfied together with the cylindrical end condition,Footnote 6 and:
-
• The moment map
$\mu :M\to \textrm {End}(W)\otimes \mathfrak {k}^*$ defined by Kostant’s formula
$($ 2.1
$)$ is proper in the sense that each inverse image of a compact subset of
$\mathfrak {k}$ by
$\mu $ is compact.
-
• The differential of the function
$|\mu |:M\to \mathbb {R}$ is
$L^{\infty }$ -bounded on the outside of the compact subset
$\mu ^{-1}(0)$ .
Note that the second condition is satisfied for the symplectic setting and the genuine moment map
$\mu $
by taking J as an
$\omega $
-compatible almost complex structure.
We show the following.
Theorem 2.10 Under Assumption 2.9, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu31.png?pub-status=live)
Remark 2.11 As it is noted in [Reference Fujita7, Example 5.2] the above equality does not hold in general without properness of
$\mu $
or completeness of M.
As a corollary of Braverman’s index theorem [Reference Braverman6, Theorem 5.5] we also have the following.
Corollary 2.12 Under Assumption 2.9,
$[\hat D]=[M, W,D_K]\in R^{-\infty }(K)$
is equal to the transverse index in the sense of Atiyah [Reference Atiyah3].
We first note that under Assumption 2.9, we can take f as in Appendix A, so that
$f=|\mu |$
on the outside of a compact neighborhood of the compact subset
$\mu ^{-1}(0)$
. Moreover, we can take an admissible function h to be
$f_{\rho }^4=\varphi _{\rho }^4f^4$
for each
$\rho \in \textrm {Irr}(K)$
, where
$\varphi _\rho $
is the cutoff function for
$V_{\rho }=M\setminus Z_{\rho }$
as in (A.1).
To show Theorem 2.10, we fix
$\rho \in \textrm {Irr}(K)$
and consider the following one-parameter family in the setting in Section 2.1:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu32.png?pub-status=live)
for
$\epsilon \in [0,1]$
. We show that for each
$\rho $
, an unbounded operator
$\mathbb {D}_\epsilon ^{(\rho )}$
on
$L^2(W)^{(\rho )}$
gives a norm-continuous family of the bounded transformations, such as
$\displaystyle \frac {\mathbb {D}_\epsilon }{\sqrt {1+\mathbb {D}_\epsilon ^2}}$
, and hence, the equality
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu33.png?pub-status=live)
holds. We use the following criteria.
Lemma 2.13 ([Reference Nicolaescu21, Proposition 1.6])
Let
$A_0$
and A be unbounded self-adjoint operators on a Hilbert space such that
$\textrm {dom}(A_0) \cap \textrm {dom}(A)$
is dense. Suppose that the family of operators
$A_\epsilon = A_0 + \epsilon A \ (\epsilon \geq 0)$
is essentially self-adjoint and, for each
$\epsilon \geq 0$
, the following conditions hold:
-
(1)
$A_\epsilon $ has a gap in its spectrum.
-
(2)
$\textrm {dom}(A_\epsilon )\subset \textrm {dom}(A).$
-
(3) There exist constants
$C, C'>0$ such that
$C' A^2\leq A_\epsilon ^2+C$ .
Then, the family of bounded transforms
$\epsilon \mapsto \frac {A_\epsilon }{\sqrt {1+A_\epsilon ^2}}$
is norm-continuous.
As in [Reference Loizides and Song19, Remark 4.10] it suffices to show the third condition in Lemma 2.13 in our situation.
Hereafter, we mainly consider the isotypic component of operators. Even if so, we often omit the superscript
$(\cdot )^{(\rho )}$
of the isotypic component for simplicity and use the notation as
$D:L^2(W)^{(\rho )}\to L^2(W)^{(\rho )}$
and so on. As we noted in Remark 2.2, one can write as
$\displaystyle D_K=\sqrt {-1}c(\underline {\rho }-\underline {\mu })$
on
$L^2(W)^{(\rho )}$
, and hence, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu34.png?pub-status=live)
Then, the third condition in Lemma 2.13 is equivalent to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu35.png?pub-status=live)
for some constants
$C, C'>0$
. Because
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu36.png?pub-status=live)
by using an orthonormal basis
$\{\xi _1,\ldots , \xi _n\}$
of
$\mathfrak {k}$
, and our boundedness condition on
$|\underline {\xi _i^M|}$
, it suffices to show the following.
Lemma 2.14 There exist constants
$C, C'>0$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu37.png?pub-status=live)
holds for all
$\epsilon \in [0,1]$
.
Proof On
$L^2(W)^{(\rho )}$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu38.png?pub-status=live)
On the other hand, there exist constants
$C_1>0$
and
$C_2>0$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu39.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu40.png?pub-status=live)
where we get the inequality in a similar way as in the proof of Proposition A.2 and we use the assumption on cylindrical end, so that we can take
$C_2$
uniformly. So we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu41.png?pub-status=live)
On the other hand, because
$\mu $
is proper and
$M^K$
is compact,
$|\epsilon \underline {\rho }-\underline {\mu }|$
is uniformly positive on the outside of a compact subset, and hence, there exists
$C'>0$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu42.png?pub-status=live)
Because
$f_\rho =|\mu |$
on the outside of a compact subset, there exists
$C>0$
independent from
$\epsilon \in [0,1]$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu43.png?pub-status=live)
Finally, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu44.png?pub-status=live)
and hence,
$(\mathbb {D}_\epsilon )^2+C>C'f_\rho ^8$
. ▪
3 Product formula
For latter convenience, we summarize the product formula for our index and some useful formulas derived from it. Instead of giving full general setting, we explain typical two situations, which will be used in the subsequent sections. We follow the basic formulation of the product formula of indices as in [Reference Atiyah and Singer4], and we give a formulation to adapt that in [Reference Fujita, Furuta and Yoshida10, Section 3.3]. Though we use terminologies in Appendix A, the main applications are the acyclic orbital Dirac-type operators constructed in Section 2 and Theorem 2.4.
3.1 Direct product
For
$i=0,1$
, let
$K_i$
be a torus. Let
$M_i$
be a complete Riemannian manifold and
$W_i\to M_i$
a
$\mathbb {Z}/2$
-graded Clifford module bundle on which
$K_i$
acts in an isometric way. Suppose that there exists a
$K_i$
-acyclic orbital Dirac-type operator
$(D_{K_i}, \{V_{i,\rho _i}\}_{\rho _i\in \textrm {Irr}(K_i)})$
on
$(M_i, W_i)$
. Put
$M:=M_0\times M_1$
and define a Clifford module bundle W over M by the outer tensor product
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu45.png?pub-status=live)
for the projections onto the first and second factors of M. For
$\rho =(\rho _0, \rho _1)\in \textrm {Irr}(K_0)\times \textrm {Irr}(K_1)$
, we define
$V_{\rho }$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu46.png?pub-status=live)
whose complement in M is compact. Let
$D_K:\Gamma (W)\to \Gamma (W)$
be an operator defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu47.png?pub-status=live)
where
$\varepsilon _{W_0}: W_0\to W_0$
is the grading operator on
$W_0$
. Because
$D_{K_0}(\varepsilon _{W_0}D_{K_1})+(\varepsilon _{W_0}D_{K_1})D_{K_0}=0$
, one has the following.
Lemma 3.1
$(D_K, \{V_{\rho }\}_{\rho \in \textrm {Irr}(K)})$
is a K-acyclic orbital Dirac-type operator on
$(M,W)$
.
Dirac operators
$D_i$
on
$W_i$
give rise to the Dirac operator D on W:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu48.png?pub-status=live)
For each
$\rho _i\in \textrm {Irr}(K_i)$
, we take a
$K_i$
-invariant cutoff function
$\varphi _{i, \rho _i}$
on
$M_i$
with
$\varphi _{i,\rho _i}|_{M_i\setminus V_{i,\rho _i}}\equiv 0$
as in (A.1). For
$\rho =(\rho _1, \rho _2)\in \textrm {Irr}(K)$
, define a function
$\varphi _\rho :M\to [0,1]$
by
$\varphi _\rho :=\varphi _{0,\rho _0} \varphi _{1,\rho _1}$
, which gives a cut-off function with
$\varphi _{\rho }|_{\widetilde M\setminus \widetilde V_{\rho }}\equiv 0$
. Then, we have a Fredholm operator on
$L^2(W)^{(\rho )}$
as the deformation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu49.png?pub-status=live)
In particular, we have the index
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu50.png?pub-status=live)
On the other hand, we have the sum of the deformations
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu51.png?pub-status=live)
which is also Fredholm on
$L^2(W)^{(\rho )}$
. In fact, by using the similar estimate in the proof of Proposition A.7, one can see that
$\hat D_\rho '$
is coercive (see [Reference Anghel2] or Proposition A.7) on the outside of a compact subset containing
$\varphi _{0,\rho _0}^{-1}(0)\cup \varphi _{1,\rho _1}^{-1}(0)=\varphi _\rho ^{-1}(0)$
.
Lemma 3.2
$\textrm {index}(\hat D_{\rho }^{\prime })=\textrm {index}(\hat D_\rho )=[M](\rho )$
.
Proof This follows from the fact that the deformation of D by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu52.png?pub-status=live)
gives a family of coercive operators by using the similar argument in the proof of Proposition A.7. ▪
Now, consider the Fredholm operator
$D_1+t\varphi _{1,\rho _1}^4D_{K_1}$
on
$L^2(W_1)^{(\rho _1)}$
and we put
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu53.png?pub-status=live)
as the
$\mathbb {Z}/2$
-graded finite-dimensional vector space. Then, there is a natural embedding
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu54.png?pub-status=live)
whose image is preserved by
$(D_0+t\varphi _{\rho _0}^4D_{K_0})\otimes \textrm {id}$
. Let
$D_{\rho _0,E_{\rho _1}}$
be the restriction of
$(D_0+t\varphi _{\rho _0}^4D_{K_0})\otimes \textrm {id}$
on this image, which gives a Fredholm operator on
$L^2(W_0\otimes E_{\rho _1})^{(\rho _0)}$
.
Proposition 3.3 We have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu55.png?pub-status=live)
If we write
$\textrm {index}(D_0+t\varphi _{\rho _0}^4D_{K_0})=E_{\rho _0}^+-E_{\rho _0}^-$
as an element in the K-group K(pt)
$\cong \mathbb {Z}$
, then we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu56.png?pub-status=live)
Proof This follows from Lemma 3.2 and the fact that the above construction satisfies [Reference Fujita, Furuta and Yoshida10, Assumption 3.14]. ▪
Hereafter, we exhibit examples and useful formulas. These examples give local models in the computation in Section 5.
Example 3.4 (Cylinder) Let
$M_1$
be the cotangent bundle of the circle
$T^*S^1\cong \mathbb {R}\times S^1$
equipped with the standard symplectic structure, almost complex structure, and the natural
$S^1$
-action on the
$S^1$
-factor. Let
$(r,\theta )$
be the coordinate on
$M_1$
. Fix
$\rho \in \textrm {Irr}(S^1)\cong \mathbb {Z}$
and put
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu57.png?pub-status=live)
where
$\mathbb {C}_{(\rho )}$
is the one-dimensional Hermitian vector space with
$S^1$
-action of weight
$\rho $
. We take a connection
$\nabla $
on
$L_\rho $
defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu58.png?pub-status=live)
where
$\mu :\mathbb {R}\to \mathbb {R}$
is a smooth nondecreasing
$S^1$
-invariant function such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu59.png?pub-status=live)
We take a Clifford module bundle
$W_{1,\rho }$
as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu60.png?pub-status=live)
with the Clifford action
$c:T^*M_1\to \textrm {End}(W_{1,\rho })$
given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu61.png?pub-status=live)
These structures give rise to a Dolbeault–Dirac operator D and an
$S^1$
-acyclic orbital Dirac-type operator
$(D_{1,\rho }, \{V_{1,\rho , \tau }\}_{\tau })$
with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu62.png?pub-status=live)
and all the data satisfy the condition in Theorem 2.4. In particular, we have the resulting index as an element in
$R^{-\infty }(S^1)$
. We denote it by
$[M_{1,\rho }]$
. By the direct computation, one has the following.
Proposition 3.5
$[M_{1,\rho }]$
is the delta function supported at
$\rho \in \textrm {Irr}(S^1)$
. Namely, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu63.png?pub-status=live)
Example 3.6 (Vector space) Consider
$M_2=\mathbb {C}$
with the standard
$S^1$
-action. Let
$B_{\delta }(0)$
be the open disc centered at the origin with radius
$\delta>0$
. Here, we take an
$S^1$
-invariant metric on
$M_2$
, so that it is standard on
$B_{\frac {1}{4}}(0)$
and isometric on the outside of
$B_{\frac {3}{4}}(0)$
to that on the subset
$\left \{r\geq \frac {3}{4}\right \}\times S^1$
of
$M_1$
. Put
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu64.png?pub-status=live)
We take a connection
$\nabla $
on
$L_\rho $
and a Clifford module bundle
$W_{2,\rho }$
, so that they are standard on
$B_{\frac {1}{4}}(0)$
and isomorphic to those on
$\left \{r>\frac {3}{4}\right \}\times S^1\subset M_1$
in Example 3.4 under the identification between
$M_2\setminus B_{\frac {3}{4}}(0)$
. These structures give rise to a Dirac operator D and an
$S^1$
-acyclic orbital Dirac-type operator
$(D_{2,\rho }, \{V_{2,\rho ,\tau }\}_{\tau })$
with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu65.png?pub-status=live)
and all the data satisfy the condition in Theorem 2.4. We denote the resulting index by
$[M_{2,\rho }]$
. By the direct computation, one has the following.
Proposition 3.7
$[M_{2,\rho }]$
is the delta function supported at
$\rho \in \textrm {Irr}(S^1)$
. Namely, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu66.png?pub-status=live)
Example 3.8 (Product of cylinders and discs)
Let
$l,m$
be nonnegative integers and M the product of l copies of the cylinder
$M_1$
and m copies of the disc
$M_2$
in the previous examples:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu67.png?pub-status=live)
There is the natural induced action of
$K:=(S^1)^{l+m}$
on M. We use the natural identifications
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu68.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu69.png?pub-status=live)
Take
$\rho =(\rho _1, \ldots , \rho _{l},\rho _1^{\prime } ,\ldots , \rho _{k}^{\prime })\in \textrm {Irr}(K)$
and consider the corresponding structures
$(M_1, W_{1,\rho _i}, D_{1,\rho _i}, \{V_{1,\rho _i,\tau }\}_{\tau \in \textrm {Irr}(S^1)})$
and
$(M_2, W_{2,\rho _j^{\prime }}, D_{2,\rho _j^{\prime }}, \{V_{2,\rho _j^{\prime }, \tau }\}_{\tau \in \textrm {Irr}(S^1)})$
. Using the outer tensor product, we can define the product of the Clifford module bundle
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu70.png?pub-status=live)
which is a Clifford module bundle over M. The products
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu71.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu72.png?pub-status=live)
induce a K-acyclic orbital Dirac-type operator on
$(M,W)$
, where for operators
$A:\mathcal {H}_0\to \mathcal {H}_0$
and
$B:\mathcal {H}_1\to \mathcal {H}_1$
on
$\mathbb {Z}/2$
-graded Hilbert spaces, their product
$A\boxtimes B :\mathcal {H}_0\otimes \mathcal {H}_1\to \mathcal {H}_0\otimes \mathcal {H}_1$
is defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu73.png?pub-status=live)
with the grading operator
$\varepsilon _0$
of
$\mathcal {H}_0$
. In fact, the data
$(D_K, \{V_\tau \}_\tau )$
satisfy the conditions in Theorem 2.4, in particular, we have the resulting index
$[M_\rho ]\in R^{-\infty }(K)$
. The product formula (Proposition 3.3) implies the following equality.
Proposition 3.9 We have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu74.png?pub-status=live)
Namely,
$[M_{\rho }]$
is the delta function supported at
$\rho \in \textrm {Irr}(K)$
.
This structure serves as a local model of a neighborhood of the fiber of the moment map of symplectic toric manifold in Section 5.3.
3.2 Fiber bundle over a closed manifold
Let X be a closed Riemannian manifold,
$E\to X$
a
$\mathbb {Z}/2$
-graded Clifford module bundle over X, and
$P\to X$
a principal G-bundle for a compact Lie group G. Consider a K-acyclic orbital Dirac-type operator
$(D_K, \{V_{\rho }\}_{\rho \in \textrm {Irr}(K)})$
on
$(M,W)$
as in Theorem 2.4. Suppose that
$G\times K$
acts on
$W\to M$
in an isometric way and
$(D_K, \{V_{\rho }\}_{\rho \in \textrm {Irr}(K)})$
is G-invariant. Consider the diagonal action of G on
$P\times M$
and the quotient manifold
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu75.png?pub-status=live)
which has a structure of M-bundle
$\pi :\widetilde M\to X$
. Let
$\widetilde W\to \widetilde M$
be the vector bundle defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu76.png?pub-status=live)
which has a structure of a Clifford module bundle over
$\widetilde M$
by using an appropriate connection of P. One can define operators
$\widetilde D_W$
and
$\widetilde D_E$
on
$\widetilde W$
as lifts (by using a trivialization of P and a partition of unity if necessary) of Dirac operators
$D_W$
on W and
$D_E$
on E. Then,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu77.png?pub-status=live)
is a Dirac operator on
$\widetilde W$
.
For
$\rho \in \textrm {Irr}(K)$
, let
$\widetilde V_{\rho }$
be the open subset defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu78.png?pub-status=live)
whose complement in
$\widetilde M$
is compact.
$D_K$
induces an operator
$\widetilde D_K$
on
$\widetilde W$
. One can see that
$(\widetilde D_K, \{\widetilde V_{\rho }\}_{\rho \in \textrm {Irr}(K)})$
is a K-acyclic orbital Dirac-type operator on
$(\widetilde M, \widetilde W)$
. In particular, we have a Fredholm operator
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu79.png?pub-status=live)
on
$L^2(\widetilde W)^{(\rho )}$
, where
$\widetilde \varphi _\rho :\widetilde M\to [0,1]$
is the cut-off function induced from the cut-off function
$\varphi _\rho $
on M as in (A.1). In this way, we have an element
$[\widetilde M]\in R^{-\infty }(K)$
defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu80.png?pub-status=live)
Now, consider the Fredholm operator
$D_W+t\varphi _{\rho }^4D_{K}$
on
$L^2(W)^{(\rho )}$
and we put
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu81.png?pub-status=live)
as the
$\mathbb {Z}/2$
-graded finite-dimensional vector space. Then, there is a natural embedding
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu82.png?pub-status=live)
whose image is preserved by
$\widetilde D_E$
. Let
$D_{E, {\rho }}$
be the restriction of
$\widetilde D_E$
on this image, which gives a Fredholm operator on
$L^2(E\otimes E_{\rho })$
, because the symbol of
$D_{E, {\rho }}$
is equal to the tensor product of
$\textrm {id}_{E_{\rho }}$
and the symbol of
$D_E$
, in particular, it is an elliptic operator on the closed manifold X.
Proposition 3.10 For each
$\rho \in \textrm {Irr}(K)$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu83.png?pub-status=live)
If we write
$\textrm {index}(D_E)=E_{0}^+-E_{0}^-$
as an element in the K-group K(pt)
$\cong \mathbb {Z}$
, then we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu84.png?pub-status=live)
Proof This follows from the fact that the above construction satisfies [Reference Fujita, Furuta and Yoshida10, Assumption 3.14]. ▪
Example 3.11 Let K be a torus. Consider
$M=T^*K$
with the K-acyclic orbital Dirac-type operator
$(D_K, \{V_{\rho }\}_{\rho \in \textrm {Irr}(K)})$
defined as the product of Example 3.4. Suppose that we take a Clifford module bundle by using
$\mathbb {C}_{(\rho )}$
for a fixed
$\rho \in \textrm {Irr}(K)$
. Then, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu85.png?pub-status=live)
Let X be a closed Riemannian manifold,
$E\to X$
a Clifford module bundle, and
$P\to X$
a principal K-bundle. Let
$\widetilde M$
be the M-bundle over X defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu86.png?pub-status=live)
Proposition 3.10 ensures us that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu87.png?pub-status=live)
where
$\textrm {index}(E)$
is the index of a Dirac operator on E. This example serves as a local model of a neighborhood of the inverse image of the moment map of Hamiltonian torus action in Section 5.2.
4 Vanishing theorem for fixed points
In this section, we show the following vanishing theorem for our index, which is a modification of [Reference Fujita, Furuta and Yoshida11, Theorem 6.1] and plays an important role in the subsequent section. Though we only use the circle action case in this paper, we give a slight general version below.
For a torus K, we consider a K-acyclic orbital Dirac-type operator on a Hermitian manifold M with a K-equivariant line bundle
$L\to M$
as in the end of Section 2.1. We fix and use the Clifford module bundle
$W_{\rho }=\wedge ^{\bullet } T_{\mathbb {C}}M\otimes L\otimes \mathbb {C}_{(\rho )}$
, where
$\mathbb {C}_{(\rho )}$
is the one-dimensional irreducible representation of K with weight
$\rho $
. We put the following assumptions.
Assumption 4.1 Together with the conditions in Theorem 2.4, we assume the followings.
-
• A compact Lie group H acts on M, which commutes with K-action, and all the additional data are
$H\times K$ -equivariant.
-
•
$Z_\rho $ is equal to the fixed-point set
$M^K$ , and it is a closed connected submanifold of M.
-
• The fixed-point set
$L^K$ is equal to the image of
$M^K$ in
$L|_{M^K}$ by the zero section.
Theorem 4.2 Under Assumption 4.1, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu88.png?pub-status=live)
To show it, we show a reducing rank lemma. Suppose that there exists a subtorus
$K'$
of K and
$\rho '\in \textrm {Irr}(K')$
such that the following conditions are satisfied.
-
• The restriction of
$\rho $ to
$K'$ -action is
$\rho '$ , i.e.,
$\iota _{K'}^*(\rho )=\rho '$ .
-
•
$Z_{\rho '}=\textrm {Zero}(\underline {\rho '}-\underline {\mu '})$ is compact for
$\mu ':=\iota _{K'}^*\circ \mu $ .
-
• The differential operator
$$ \begin{align*} D_{K'}=\sum_{i=1}^{{\dim}K'}c(\underline{\xi_i^M})(\mathcal{L}_{\xi_i}-\sqrt{-1}\mu_i) \end{align*} $$
$V_{\rho '}:=M\setminus Z_{\rho '}$ give a
$\rho '$ -acyclic orbital Dirac-type operator on
$(M,W_\rho )$ .
The deformation
$\hat D_{\rho '}=D+t\varphi _{\rho '}^4D_{K'}$
gives a Fredholm operator on the isotypic component
$L^2(W_\rho )^{(\rho ')}$
for
$t\gg 0$
, where
$\varphi _{\rho '}$
is a cut-off function for
$V_{\rho '}$
as in (A.1). On the other hand, the condition
$\iota _{K'}^*(\rho )=\rho '$
implies that
$L^2(W_\rho )^{(\rho )}$
is a subspace of
$L^2(W\rho )^{(\rho ')}$
and
$(\hat D_{\rho '})^{(\rho ')}$
preserves it. We define
$\textrm {index}(\hat D_{\rho ',\rho })$
as its Fredholm index:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu90.png?pub-status=live)
We can incorporate H-action and regard them as H-equivariant indices
$\textrm {index}_{H}(\cdot )$
.
Lemma 4.3
$[Z_{\rho }]=\textrm {index}_H(\hat D_{\rho })=\textrm {index}_H(\hat D_{\rho ',\rho })\in R(H)$
.
Proof By taking a basis of
$\mathfrak {k}$
, which is an extension of a basis of
$\mathfrak {k}'$
, we may assume that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu91.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu92.png?pub-status=live)
We also define
$D_{K,K'}$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu93.png?pub-status=live)
Take and fix cut-off functions
$\varphi _{\rho }$
for
$V_\rho $
and
$\varphi _{\rho '}$
for
$V_{\rho '}$
as in (A.1). We put
$\varphi _{\rho ,\rho '}:=\varphi _\rho \varphi _{\rho '}$
. There exists
$t>0$
such that the deformation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqn3.png?pub-status=live)
gives a Fredholm operator on the isotypic component
$L^2(W_\rho )^{(\rho )}$
. The almost same argument in the proof of Theorem 2.10 implies that for any
$t'\geq t$
, the deformation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu94.png?pub-status=live)
is Fredholm on
$L^2(W_\rho )^{(\rho )}$
and its Fredholm index is the same as that of (4.1). On the other hand, for fixed such t, the family
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu95.png?pub-status=live)
satisfies the coercivity on the interior of
$\varphi _{\rho ,\rho '}^{-1}(1)$
for
$t'\geq t$
large enough. It implies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu96.png?pub-status=live)
The excision property implies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu97.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu98.png?pub-status=live)
which complete the proof. ▪
Remark 4.4 To show Lemma 4.3, we do not use the assumption
$Z_{\rho }=M^K$
.
Proposition 4.5 Theorem 4.2 is true when M is a small open disc around the origin of a Hermitian vector space on which the K-action is linear and
$M^K$
consists of the origin.
Proof By considering the tensor product, it suffices to prove in the case that
$\rho $
is the trivial representation
${\textbf {0}}$
. We can choose an appropriate generic circle subgroup
$K_1$
of K, so that
$K_1$
acts on M with
$M^{K_1}=\{0\}$
and the
$K_1$
-action on
$L|_0$
is nontrivial. In fact, let
$\rho _1,\ldots , \rho _{\dim M}\in \textrm {Irr}(K)$
be the weights appeared in the linear action on M, all of which are nonzero by the assumption
$M^{K}=\{0\}$
, then we can take a splitting of the differential of the representation
$K\to U(1)$
on
$L|_0$
such that the image of the splitting in
${\mathfrak {k}}$
is rational and is not perpendicular to any
$\rho _i$
. The subgroup of the image gives the desired circle subgroup. By Lemma 4.3, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu99.png?pub-status=live)
On the other hand, [Reference Fujita, Furuta and Yoshida11, Proposition 6.8] and Theorem A.5 imply
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu100.png?pub-status=live)
and we complete the proof. ▪
Proof of Theorem 4.2 The claim follows from Proposition 4.5 and the product formula (Proposition 3.10) with the same argument in [Reference Fujita, Furuta and Yoshida11, Section 6.4]. ▪
5 Quantization of noncompact Hamiltonian torus manifolds
In this section, by using the ingredients established in the previous sections, we define quantization of noncompact symplectic manifolds equipped with Hamiltonian group action and show [Q,R]=0 for circle action case and a Danilov-type formula for toric action case.
5.1 Definition: general case
Let K be a compact torus and M a symplectic manifold equipped with Hamiltonian K-action. Suppose that there exists a K-equivariant prequantizing line bundle
$(L,\nabla )$
, and let
$\mu :M\to \mathfrak {k}^*$
be the associated moment map. We use the Clifford module bundle
$W=\wedge ^\bullet T_{\mathbb {C}} M\otimes L$
for a K-invariant compatible almost complex structure. We assume the following for the moment:
Assumption 5.1 For each
$\rho \in \textrm {Irr}(K)$
, the zero set
$Z_{\rho }=\textrm {Zero}(\underline \rho -\underline \mu )$
is compact.
Definition 5.1 We define its quantization
$\mathcal {Q}_K(M)\in R^{-\infty }(K)$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqn4.png?pub-status=live)
where
$\widetilde X_{\rho }$
is a complete manifold containing
$Z_{\rho }$
as its neighborhood on which the Dirac operator along orbits defined as in Definition 2.1 gives a
$\rho $
-acyclic orbital Dirac-type operator for
$\widetilde X_\rho \setminus Z_{\rho }$
.
The excision property guarantees that the number
$\mathcal {Q}_K(M)(\rho )$
is independent from the choice of such
$\widetilde X_{\rho }$
. Theorem 2.8 enables us to describe
$\mathcal {Q}_K(M)(\rho )$
into the sum of local contributions
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu101.png?pub-status=live)
It would be natural to expect the vanishing of
$[Z_{\rho , \alpha }]$
. One possible way to show this vanishing is using a combination of the coincidence of
$[Z_{\rho ,\alpha }]$
with the transverse index and vanishing results for it, e.g., by Paradan [Reference Paradan22]. In the subsequent subsections, instead of using them, we have the vanishing of
$[Z_{\rho , \alpha }]$
for the circle action case and the toric case based on Theorem 4.2, and we define the quantization
$\mathcal {Q}_K(M)$
under a weaker assumption than Assumption 5.1.
The quantization
$\mathcal {Q}_K(M)$
is a generalization of K-equivariant spin
$^c$
quantization using the index of Dolbeault–Dirac operator in the compact case, which is often denoted by
$RR_K(M)$
and called the equivariant Riemann–Roch number or Riemann–Roch character.
5.2 [Q,R]=0 for noncompact Hamiltonian torus manifolds
In this subsection, we consider the case
$K=S^1$
. Because, in this case, one has
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu102.png?pub-status=live)
for each
$\rho \in \textrm {Irr}(K)=\Lambda ^*$
and
$\mu (M^{K})\subset \Lambda ^*$
, the quantization
$\mathcal {Q}_K(M)$
has a localization property to
$\Lambda ^*$
. Moreover, one has a decomposition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu103.png?pub-status=live)
which gives us a decomposition of the index
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu104.png?pub-status=live)
where we use the notation as in Section 2.2. Proposition 2.6 and Theorem 4.2 imply that we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu105.png?pub-status=live)
This observation enables us to define
$\mathcal {Q}_K(M)$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu106.png?pub-status=live)
without Assumption 5.1. We only need the following assumption.
Assumption 5.2 The preimage of each lattice point in
$\Lambda ^*$
is compact.
This definition leads us to a proof of [Q,R]=0, the principal of “quantization commutes with reduction,”as in [Reference Fujita, Furuta and Yoshida11] in the noncompact case.
For a regular value
$\xi \in {\mathfrak {k}}^*$
of
$\mu :M\to {\mathfrak {k}}^*$
, let
$M_\xi $
be the symplectic quotient at
$\xi $
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu107.png?pub-status=live)
which is a closed symplectic manifold (orbifold) under Assumption 5.2. Moreover, if a regular value
$\rho $
is an element of
$\textrm {Irr}(K)$
, then there exists a natural prequantizing line bundle over
$M_\rho $
, and hence, one can define the Riemann–Roch number
$RR(M_\rho )$
as the index of the Dolbeault–Dirac operator associated with a K-invariant compatible almost complex structure.
Theorem 5.3 Suppose that
$\rho \in \textrm {Irr}(K)$
is a regular value of the moment map
$\mu :M\to {\mathfrak {k}}^*$
. Then, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu108.png?pub-status=live)
Proof A neighborhood of
$\mu ^{-1}(\rho )$
in M can be identified with the product
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu109.png?pub-status=live)
by the Darboux-type theorem (see [Reference Fujita, Furuta and Yoshida11, Lemma 7.1] for example), which has a structure of
$T^*K$
-bundle over
$M_\rho $
. By applying the product formula in Example 3.11, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu110.png?pub-status=live)
▪
Remark 5.4
-
(1) Even for a higher rank torus case, by choosing a circle subgroup generic enough, one can give a proof of Theorem 5.3 by induction.
-
(2) Due to Corollary 2.12, the quantization
$\mathcal {Q}_K(M)$ can be identified with Atiyah’s transverse index. Theorem 5.3 gives an alternative proof of Vergne’s conjecture for torus case to Ma–Zhang’s proof in [Reference Ma and Zhang20], which uses Braverman’s deformation.
-
(3) The above construction and a proof of Theorem 5.3 are essentially the same as those in [Reference Fujita7].
5.3 A Danilov-type formula for noncompact toric manifolds
Now, we focus on the symplectic toric case. Namely, we assume that K is a torus with
$2\dim (K)=\dim (M)$
. In this case, Assumption 5.2 is automatically satisfied, because the preimage of each point is a single orbit. We can define the quantization
$\mathcal {Q}_K(M),$
as it is noted in the previous section. In fact, for each
$\rho \in \textrm {Irr}(K)$
, the image
$\mu (Z_{\rho ,\alpha })$
of the component of
$Z_{\rho }$
is contained in the boundary of the momentum polytope
$\mu (M)$
, and one can see
$[\mu ^{-1}(\rho )]=1$
and
$[Z_{\rho ,\alpha }]=0$
by the same argument in [Reference Fujita8, Section 6.1] together with Proposition 2.6 and Theorem 4.2. These observations enable us to define
$\mathcal {Q}_K(M)\in R^{-\infty }(K)$
and give the following description, which is a noncompact generalization of Danilov’s formula.
Theorem 5.5 Under the above setup, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu111.png?pub-status=live)
where the right-hand side is an element in
$R^{-\infty }(K)$
, which is characterized by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu112.png?pub-status=live)
Remark 5.6 In a general framework of geometric quantization, one uses an additional structure called a polarization, which is an integrable Lagrangian distribution of the complexification of the tangent bundle. One typical example is a Kähler polarization, which is defined as a compatible complex structure. Our quantization is the spin
$^c$
quantization, which is a quantization based on a polarization relaxed the integrality condition in the Kähler polarization. The quantization is given by the Fredholm index of the Dolbeault–Dirac operator. The other example is a real polarization, which is defined by the tangent bundle along fibers of the Lagrangian fibration. In the real polarization case, it is known that the quantization can be described by Bohr–Sommerfeld fibers, which are characterized by the existence of nontrivial global parallel sections of the prequantizing line bundle on the orbits. The moment map of toric manifolds can be regarded as a real polarization with singular fibers. In the toric case, the Bohr–Sommerfeld fibers are nothing other than the inverse images of the integral lattice points in the momentum polytope.
One important topic in geometric quantization is the problem of independence from the polarizations. There are several results supporting the coincidence between the quantizations obtained by the spin
$^c$
polarization and the real polarization from the viewpoint of index theory, such as [Reference Andersen1, Reference Fujita, Furuta and Yoshida9, Reference Kubota15, Reference Yoshida28]. Theorem 5.5 can be considered as a noncompact version of the above results.
Remark 5.7 In [Reference Fujita8], we gave a proof of Danilov’s formula for compact symplectic toric manifolds (or more generally for toric origami manifolds) using a localization formula based on the theory of the acyclic compatible fibration/system developed in [Reference Fujita, Furuta and Yoshida10]. Because one can see that the acyclic compatible fibration constructed on a given toric manifold does not have a product structure in general, we cannot apply the product formula and have to compare the resulting index with the index of the product. One remarkable difference in the computation of the local contribution is that our deformation by
$D_K$
fits into the local product structure of a neighborhood of
$\mu ^{-1}(\rho )$
. In particular, we can apply the product formula directly.
6 Comments and further discussions
6.1 Application to quantization of Hamiltonian loop group spaces
Quantization of Hamiltonian loop group spaces is studied in various directions. In particular, Loizides and Song [Reference Loizides and Song18] studied it from the viewpoint of index theory and KK-theory. Their construction is based on their previous work [Reference Loizides, Meinrenken and Song16] with Meinrenken in which they constructed a spinor bundle over a proper Hamiltonian loop group space and a nice finite-dimensional noncompact submanifold in it, which is transverse to the orbits of the loop group action. One key ingredient in [Reference Loizides and Song18] is to associate a K-homology cycle to such a noncompact manifold. They established an index theory using the
$\textrm {C}^*$
-algebraic condition, which they call the
$(\Gamma , K)$
-admissibility, where K is a compact Lie group and
$\Gamma $
is a countable discrete group with proper length function. They showed that in the proper Hamiltonian loop group space case, the
$(\Lambda , T)$
-admissibility is satisfied for a maximal torus T of K, and the resulting K-homology class has an antisymmetric property with respect to some Weyl group action of K, which gives rise to quantization as an element in the fusion ring of K.
In this paper, we constructed a similar K-homology cycle without using
$(\Gamma , K)$
-admissibility. In the subsequent research, we will investigate an approach of quantization of Hamiltonian loop group spaces by incorporating the action of the integral lattice
$\Lambda $
in our construction appropriately. In such an approach, it would be interesting to understand how the localization phenomenon of our index is reflected in the quantization of loop group spaces.
There is another related work by Takata. In [Reference Takata24], an
$LS^1$
-equivariant index is constructed as an element in the fusion ring from the viewpoint of KK-theory and noncommutative geometry. He also developed an index theorem in infinite-dimensional setting in [Reference Takata23, Reference Takata25]. It would be also interesting to investigate how our construction is positioned in Takata’s theory.
6.2 Deformation as KK-products
Motivated by the pioneering work by Kasparov [Reference Kasparov14], Loizides et al. showed in [Reference Loizides, Rodsphon and Song17] that the K-homology class which is obtained by Braverman’s deformation factors as a KK-product of the Dirac class and a KK-class arising from the deformation. It is desirable to understand our deformation using the acyclic orbital Dirac-type operator as a KK-product.
A Appendix: K-acyclic orbital Dirac-type operator
In this appendix, we give a general machinery to have an equivariant index and a K-homology cycle using a deformation by differential operator along orbits. Though the machinery itself works for nonabelian case, we do not know a good example except
$D_K$
as in Definition 2.1 for the moment.
A.1 Setup and definition
Let M be a complete Riemannian manifold and
$W\to M $
a
$\mathbb {Z}/2$
-graded
$\textrm {Cl}(TM)$
-module bundle with the Clifford multiplication
$c:TM\cong T^*M\to \textrm {End}(W)$
.
Let K be a compact Lie group acting on M in an isometric way. We assume that the K-action lifts to a unitary action of W. Take and fix a K-invariant Dirac-type operator
$D:\Gamma _c(W)\to \Gamma _c(W)$
.
Definition A.1 (
$\rho $
-acyclic and K-acyclic orbital Dirac-type operator)
Let
$\rho $
be an element of
$\textrm {Irr}(K)$
. A pair
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu113.png?pub-status=live)
is called a
$\rho $
-acyclic orbital Dirac-type operator on
$(M,W)$
if the following conditions are satisfied.
-
(1)
$D_K:\Gamma _c(W)\to \Gamma _c(W)$ is a K-invariant first-order self-adjoint differential operator such that:
-
(a)
$D_K$ contains only differentials along K-orbits, and the restriction to each K-orbit is an elliptic operator on the orbit.
-
(b)
$D_K$ is finite propagation speed, i.e., the principal symbol
$\sigma (D_K):T^*M\to \textrm {End}(W)$ satisfies
$$ \begin{align*} \mathop{\mathrm{sup}}\limits \{\|\sigma(D_K)(v)\| \ | \ v\in T^*M, \ |v|=1 \} < \infty. \end{align*} $$
-
(c)
$D_K$ anticommutes with the Clifford multiplication of the transverse direction to orbits. Namely, for any K-invariant function h on M, one has
$$ \begin{align*} D_Kc(dh)+c(dh)D_K=0. \end{align*} $$
-
(d) The isotypic component
$D_K^{(\rho )}$ gives a bounded operator on
$L^2(W)^{(\rho )}$ .
-
-
(2)
$V_{\rho }$ is an open subset of M such that
$M\setminus V_\rho $ is compact.
-
(3) We have
$$ \begin{align*} \ker(D_{K}|_{V_\rho})^{(\rho)}=0. \end{align*} $$
-
(4) There exists a constantFootnote 7
$C_{\rho }>0$ such that
$$ \begin{align*} |( (DD_{K}+D_{K}D)s, s )_W|\leq C_{\rho}( D_{K}^2 s,s)_W \end{align*} $$
$$ \begin{align*} |( D_{K}s, s )_W|\leq C_{\rho}( D_{K}^2 s,s)_W \end{align*} $$
$s\in \Gamma _c(W|_{V_\rho })^{(\rho )}$ .
-
(5) There exists a constant
$\kappa _\rho>0$ such that
$$ \begin{align*} \kappa_\rho( s, s)_W \leq ( D_{K}^2 s,s)_W \end{align*} $$
$s\in \Gamma _c(W|_{V_\rho })^{(\rho )}$ .
If a family of open subsets
$\{V_{\rho }\}_{\rho \in \textrm {Irr}(K)}$
gives a
$\rho $
-acyclic orbital Dirac-type operator
$(D_K,V_\rho )$
for each
$\rho \in \textrm {Irr}(K)$
, then we call
$(D_K, \{V_{\rho }\}_{\rho \in \textrm {Irr}(K)})$
the K-acyclic orbital Dirac-type operator.
The completeness of M implies that there exists a K-invariant smooth proper function
$f:M\to [1,\infty )$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu120.png?pub-status=live)
We take and fix such f. For each
$\rho \in \textrm {Irr}(K)$
, we take and fix a K-invariant cut-off function
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqn5.png?pub-status=live)
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu121.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu122.png?pub-status=live)
We put
$f_\rho :=\varphi _\rho f$
. We consider the deformation of D defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu123.png?pub-status=live)
Because D and
$D_K$
have finite propagation speed,
$\hat D_\rho $
is an essentially self-adjoint operator on
$L^2(W)$
. Moreover, one can see that
$\hat D_\rho $
is transversally elliptic in the sense of Atiyah [Reference Atiyah3]. In fact, for any K-invariant function
$h:M\to \mathbb {R}$
, because
$D_K$
commutes with the multiplication by h, one has
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu124.png?pub-status=live)
which is invertible unless
$dh=0$
.
Hereafter, we mainly consider the isotypic component
$\hat D_{\rho }^{(\rho )}$
. Even if so, we often omit the superscript
$(\cdot )^{(\rho )}$
of the isotypic component for simplicity and use the notation as
$\hat D_{\rho }:L^2(W)^{(\rho )}\to L^2(W)^{(\rho )}$
and so on.
Remark A.1 The Clifford module structure and Dirac-type condition are not so essential. In fact, we can establish almost all propositions, definitions, etc., below for more general vector bundles and elliptic operators with finite propagation speed. However, because we do not have applications of such generalizations, we only handle with Clifford module bundles and Dirac-type operators in the present paper.
A.2 Compactness and K-Fredholm property
Let
$(D_K, \{V_{\rho }\}_{\rho \in \textrm {Irr}(K)})$
be a K-acyclic orbital Dirac-type operator on
$(M,W)$
. We take and fix a family of functions
$\{f,\{\varphi _\rho \}_{\rho \in \textrm {Irr}(K)}\}$
as above.
Proposition A.2 For each
$\rho \in \textrm {Irr}(K)$
, there exists a smooth K-invariant proper function
$\Phi _{\rho }:M\to \mathbb {R}$
such that
$\Phi _{\rho }$
is bounded below and we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu125.png?pub-status=live)
as self-adjoint operators on
$L^2(W)^{(\rho )}$
.
Proof Because
$f_\rho $
is K-invariant, we have an equality on
$\Gamma _c(W)^{(\rho )}$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu126.png?pub-status=live)
Now, for any
$s\in \Gamma _c(W)^{(\rho )}$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu127.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu128.png?pub-status=live)
Summarizing the above inequalities, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu129.png?pub-status=live)
Now, put
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu130.png?pub-status=live)
Because
$f_\rho $
is proper and bounded below, the function
$g_\rho $
is also proper and bounded below. Note that
$M_-:=g_\rho ^{-1}((-\infty , 0])$
is a compact subset of M, and hence, by the boundedness of
$D_K$
(1(d) in Definition A.1), there exists a constant
$C_{\rho , M_-}>0$
such that we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu131.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu132.png?pub-status=live)
As a consequence, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu133.png?pub-status=live)
for
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu134.png?pub-status=live)
which is K-invariant, proper, and bounded below. ▪
As a corollary, we have the following compactness by [Reference Loizides and Song18, Proposition B.1].
Corollary A.3 For any
$\rho \in \textrm {Irr}(K)$
, a bounded operator
$((\hat D_\rho ^2)^{(\rho )}+1)^{-1}$
on
$L^2(W)^{(\rho )}$
is a compact operator. In particular,
$(\hat D_\rho )^{(\rho )}$
is a Fredholm operator on
$L^2(W)^{(\rho )}$
.
Definition A.2 Define an element
$[\hat D]\in R^{-\infty }(K)$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu135.png?pub-status=live)
for each
$\rho \in \textrm {Irr}(K)$
. We also use the notations
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu136.png?pub-status=live)
Hereafter, we often write
$[\hat D](\rho )=\textrm {index}(\hat D_{\rho })\in \mathbb {Z} $
instead of
$\textrm {index}((\hat D_{\rho })^{(\rho )})$
.
In general, a K-equivariant operator A on a
$\mathbb {Z}/2$
-graded Hilbert space
${\mathcal {H}}$
with isometric K-action is called K-Fredholm if each isotypic component
$A^{(\rho )}:{\mathcal {H}}^{(\rho )}\to {\mathcal {H}}^{(\rho )}$
is Fredholm. Such a K-Fredholm operator A defines an element in
$R^{-\infty }(K)$
denoted by a formal expression:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu137.png?pub-status=live)
Corollary A.3 and Definition A.2 imply that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu138.png?pub-status=live)
is a K-Fredholm operator and
$[\hat D]$
is its index in
$R^{-\infty }(K)$
.
A.3 K-homology cycle representing the class
$[\hat D]$
We consider the same setup as in the previous sections. For each
$\rho \in \textrm {Irr}(K)$
, we put
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu139.png?pub-status=live)
which is a bounded operator acting on
$L^2(W)^{(\rho )}$
with
$\|F_{\rho }\|=1$
. We can see that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqn6.png?pub-status=live)
gives a bounded operator on
$\displaystyle L^2(W)=\bigoplus _{\rho \in \textrm {Irr}(K)}L^2(W)^{(\rho )}$
.
It is known that the formal completion
$R^{-\infty }(K)$
can be identified with the K-homology group of the group C
$^*$
-algebra
$\textrm {K}^0(C^*(K))$
, which is also identified with the KK-group
$\textrm{K K} (C^*(K),\mathbb {C})$
. These groups are generated by triples consisting of a Hilbert space, a C
$^*$
-representation of
$C^*(K)$
, and a bounded operator on the Hilbert space satisfying certain boundedness and compactness. See [Reference Blackadar5, Reference Higson and Roe12, Reference Kasparov14] for basic definitions on K-homology or KK-theory. Corollary A.3 implies the following.
Proposition A.4 The bounded operator F as in (
A.2
) together with the natural representation of
$C^*(K)$
on
$L^2(W)$
gives a K-homology cycle which represents
$[\hat D]$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu140.png?pub-status=live)
A.4 Relation with Fujita–Furuta–Yoshida’s deformation
In this section, we consider another deformation of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu141.png?pub-status=live)
for
$\rho \in \textrm {Irr}(K)$
using a K-acyclic orbital Dirac-type operator
$(D_K, \{V_{\rho }\}_{\rho \in \textrm {Irr}(K)})$
, where
$\varphi _{\rho }$
is the cut-off function as in (A.1). This type of deformation was studied for an acyclic compatible system in a series of papers [Reference Fujita, Furuta and Yoshida9–Reference Fujita, Furuta and Yoshida11]. One main differenceFootnote
8
between the above deformation and
$\hat D_{\rho }$
is the presence of a proper function f. To compare them, we introduce a one-parameter family
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu142.png?pub-status=live)
which acts on
$L^2(W)$
. We show the following.
Theorem A.5 For each
$\rho \in \textrm {Irr}(K)$
, there exists
$t_{\rho }>0$
such that
$\{\mathbb {D}_{\epsilon }\}_{\epsilon \in [0,1]}$
gives a family of Fredholm operator on
$L^2(W)^{(\rho )}$
for any
$t>t_\rho $
and its Fredholm index does not depend on
$\epsilon $
and t. In particular, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu143.png?pub-status=live)
Corollary A.6 Define
$[D_t]\in R^{-\infty }(K)$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu144.png?pub-status=live)
for each
$\rho \in \textrm {Irr}(K)$
. Then, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu145.png?pub-status=live)
Note that because both
$D_K$
and D are essentially self-adjoint,
$\mathbb {D}_\epsilon $
is also an essentially self-adjoint operator on
$L^2(W)^{(\rho )}$
. Theorem A.5 follows from the following estimate, which is also known as the coercivity in [Reference Anghel2]. In fact, as in [Reference Fujita, Furuta and Yoshida10], the
$\mathbb {Z}/2$
-graded Fredholm index of a coercive family of essentially self-adjoint operators does not depend on a parameter of the family.
Proposition A.7 There exist an open subset
$U_\rho $
and a constant
$t_\rho>0$
such that
$M\setminus U_{\rho }$
is compact and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu146.png?pub-status=live)
holds for any
$s\in \Gamma _c(W)^{(\rho )}$
with
$\textrm {supp}(s)\subset U_\rho $
,
$\epsilon \in [0,1]$
, and
$t>t_\rho $
, where
$\kappa _\rho>0$
is the constant as in
$($
A.1
$)$
of Definition A.1.
Proof We take
$U_\rho '$
to be the interior of
$\varphi _\rho ^{-1}(1)$
and put
$h:=(1-\epsilon )f^4 +\epsilon t$
. On
$U_\rho '$
, consider the square
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu147.png?pub-status=live)
For any
$s\in \Gamma _c(W)^{(\rho )}$
with
$\textrm {supp}(s)\subset U_\rho '$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu148.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu149.png?pub-status=live)
It implies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu150.png?pub-status=live)
Now, put
$t_\rho :=4\|df\|_\infty C_\rho +C_\rho +1$
and define
$U_\rho $
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu151.png?pub-status=live)
Then, on
$U_\rho $
when
$t>t_\rho $
, we have
$ (-4\|df\|_\infty C_\rho -C_\rho +h)h> t_\rho . $
Finally, we haveFootnote
9
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu152.png?pub-status=live)
▪
Theorem A.5 implies that one can adopt the deformation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqnu153.png?pub-status=live)
without the proper function f to discuss the equivariant index
$[M](\rho )=[\hat D](\rho )=\textrm {index}(\hat D_{\rho })$
. It also impliesFootnote
10
that
$[M](\rho )$
satisfies the excision formula, sum formula, invariance under continuous deformations, and product formula as stated in [Reference Fujita, Furuta and Yoshida10, Section 3]. In particular, if there are two data
$(M, W, D, D_K, V_{\rho })$
and
$(M^{\prime },W^{\prime }, D^{\prime },D^{\prime }_K,V^{\prime }_\rho )$
for the same K and
$\rho \in \textrm {Irr}(K)$
, which are isomorphic on neighborhoods of compact subsets
$M\setminus V_\rho $
and
$M^{\prime }\setminus V_{\rho }^{\prime }$
, then the excision formula implies that the resulting indices coincide:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220809030118010-0201:S0008414X2100016X:S0008414X2100016X_eqn7.png?pub-status=live)
It ensures us to define the index starting from a noncomplete manifold by taking an appropriate completion, for instance, a cylindrical end as in [Reference Fujita, Furuta and Yoshida10, Section 7.1] or [Reference Loizides and Song18, Section 4.7]. We gave an explanation of such a construction in Section 2.2 and used in Section 5.
Acknowledgment
This work had been done while the author stayed at the Department of Mathematics, University of Toronto, and the Department of Mathematics and Statistics, McMaster University. The author would like to thank their hospitality, especially for L. Jeffrey and M. Harada. He also would like to thank Y. Loizides for explaining his work and having fruitful discussion about abelian case. Finally, the author is grateful to the referee for pointing out several mistakes in the preliminary version.