1 Introduction
Let
$(V,Q)$
be a quadratic space over
$\mathbb {Q}$
of signature
$(p,q)$
, and let G be its orthogonal group. Let
$\mathbb {D}$
be the space of oriented negative q-planes in
$V(\mathbb {R})$
and
$\mathbb {D}^+$
one of its connected components. It is a Riemannian manifold of dimension
$pq$
and an open subset of the Grassmannian. The Lie group
$G(\mathbb {R})^+$
is the connected component of the identity and acts transitively on
$\mathbb {D}^+$
. Hence, we can identify
$\mathbb {D}^+$
with
$G(\mathbb {R})^+ /K $
, where K is a compact subgroup of
$G(\mathbb {R})^+$
and is isomorphic to
$\operatorname {SO}(p)\times \operatorname {SO}(q)$
. Moreover, let L be a lattice in
$ V(\mathbb {Q}),$
and let
$\Gamma $
be a torsion-free subgroup of
$G(\mathbb {R})^+$
preserving L.
For every vector v in
$V(\mathbb {R})$
such that
$Q(v,v)>0$
, there is a totally geodesic submanifold
$\mathbb {D}^+_v$
of codimension q consisting of all the negative q-planes that are orthogonal to v. Let
$\Gamma _v$
denote the stabilizer of v in
$\Gamma $
. We can view
$ \Gamma _v \backslash \mathbb {D}^+$
as a rank q vector bundle over
$\Gamma _v \backslash \mathbb {D}_v^+$
, so that the natural embedding
$\Gamma _v \backslash \mathbb {D}_v^+ $
in
$ \Gamma _v \backslash \mathbb {D}^+$
is the zero section. In [Reference Kudla and Millson6], Kudla and Millson constructed a closed
$G(\mathbb {R})^+$
-invariant differential form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn1.png?pub-status=live)
where
$G(\mathbb {R})^+$
acts on the Schwartz space
$\mathscr {S}(V(\mathbb {R}))$
from the left by
and on
$\Omega ^q(\mathbb {D}^+) \otimes \mathscr {S}(V(\mathbb {R}))$
from the right by
. In particular,
$\varphi _{KM}(v)$
is a
$\Gamma _v$
-invariant form on
$\mathbb {D}^+$
. The main property of the Kudla–Millson form is its Thom form property: if
$\omega $
in
$\Omega _c^{pq-q}(\Gamma _v \backslash \mathbb {D}^+)$
is a compactly supported form, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn2.png?pub-status=live)
Another way to state it is to say that in cohomology, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn3.png?pub-status=live)
where
$\operatorname {PD}(\Gamma _v \backslash \mathbb {D}^+_v)$
denotes the Poincaré dual class to
$\Gamma _v \backslash \mathbb {D}^+_v$
.
1.1 Kudla–Millson theta lift
In order to motivate the interest in the Kudla–Millson form, let us briefly recall how it is used to construct a theta correspondence between certain cohomology classes and modular forms. For simplicity,Footnote
1
assume that
$p+q$
is even, and let
$\omega $
be the Weil representation of the dual pair
$\operatorname {SL}_2(\mathbb {R}) \times G(\mathbb {R})$
in
$\mathscr {S}(V(\mathbb {R}))$
. We extend it to a representation in
$\Omega ^q(\mathbb {D}^+) \otimes \mathscr {S}(V(\mathbb {R}))$
by acting in the second factor of the tensor product. Building on the work of [Reference Weil11], Kudla and Millson [Reference Kudla and Millson7, Reference Kudla and Millson9] used their differential form to construct the theta series
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn4.png?pub-status=live)
where
$\tau =x+iy$
is in
$\mathbb {H}$
and
$g_\tau $
is the matrix
$\begin {pmatrix} \sqrt {y} & x\sqrt {y}^{-1} \\ 0 & \sqrt {y}^{-1} \end {pmatrix}$
in
$\operatorname {SL}_2(\mathbb {R})$
that sends i to
$\tau $
by Möbius transformation. This form is
$\Gamma $
-invariant, closed and holomorphic in cohomology in the sense that
$\frac {\partial }{\partial \overline {\tau }}\Theta _{KM}(\tau )$
is an exact form. Kudla and Millson showed that if we integrate this closed form on a compact q-cycle C in
$\mathcal {Z}_q(\Gamma \backslash \mathbb {D}^+)$
, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn5.png?pub-status=live)
is a modular form of weight
$\frac {p+q}{2}$
, where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn6.png?pub-status=live)
and the special cycles
$C_v$
are the images of the composition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn7.png?pub-status=live)
Thus, the Kudla–Millson theta series realizes a lift between the (co)-homology of
$\Gamma \backslash \mathbb {D}^+$
and the space of weight
$\frac {p+q}{2}$
modular forms.
1.2 The result
Let E be a
$G(\mathbb {R})^+$
-equivariant vector bundle of rank q over
$\mathbb {D}^+$
, and let
$E_0$
be the image of the zero section. By the equivariance, we also have a vector bundle
$\Gamma _v \backslash E$
over
$\Gamma _v \backslash \mathbb {D}^+$
. The Thom class of the vector bundle is a characteristic class
$\operatorname {Th}(\Gamma _v \backslash E)$
in
$ H^{q}(\Gamma _v \backslash E,\Gamma _v \backslash (E-E_0))$
defined by the Thom isomorphism (see Section 3.6). A Thom form is a form representing the Thom class. It can be shown that the Thom class is also the Poincaré dual class to
$\Gamma _v \backslash E_0$
. Let
$s_v \colon \Gamma _v \backslash \mathbb {D}^+ \longrightarrow \Gamma _v \backslash E$
be a section whose zero locus is
$\Gamma _v \backslash \mathbb {D}_v^+$
, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn8.png?pub-status=live)
Viewing it as a class in
$H^q(\Gamma _v \backslash \mathbb {D}^+)$
it is the Poincaré dual class of
$\Gamma _v \backslash \mathbb {D}_v^+$
. Since the Poincaré dual class is unique, property (1.3) implies that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn9.png?pub-status=live)
on the level of cohomology.
For arbitrary real oriented metric vector bundles, Mathai and Quillen used the Chern–Weil theory to construct in [Reference Mathai and Quillen10] a canonical Thom form on E. We denote by
$U_{MQ}$
the canonical Thom form in
$\Omega ^{q}(E)$
of Mathai and Quillen. Since
$U_{MQ}$
is
$\Gamma $
-invariant, it is also a Thom form for the bundle
$\Gamma _v \backslash E$
for every vector v. The main result is the following.
Theorem (Theorem 4.5)
For a natural choice of a bundle E and of a section
$s_v$
, we have
$\varphi _{KM}(v)= 2^{-\frac {q}{2}}e^{-\pi Q(v,v)} s_v^\ast U_{MQ}$
in
$\Omega ^q(\Gamma _v \backslash \mathbb {D}^+).$
The bundle E is the tautological bundle of the Grassmannian
$\mathbb {D}^+$
(see Section 3.6), and the section
$s_v$
is defined in Section 4.1.
For signature
$(2,q)$
, the spaces are Hermitian and the result was obtained by a similar method in [Reference Garcia3] using the work of Bismut–Gillet–Soulé.
1.3 Generalizations
More generally, for a positive nondegenerate r-subspace
$U \subset V$
spanned by vectors
$v_1, \dots , v_r$
, Kudla and Millson also construct an
$rq$
form
$\varphi _{KM}(v_1,\dots ,v_r)$
. This form can also be recovered by the Mathai–Quillen formalism (see (3) of Section 5). Furthermore, in [Reference Kudla and Millson7, Reference Kudla and Millson9], they not only construct forms for the symmetric space associated with
$\operatorname {SO}(p,q)$
, but also for the Hermitian space associated with
$U(p,q)$
. In this case, one should be able to recover their forms using the formalism of superconnections as in [Reference Mathai and Quillen10, Theorem 8.5]. We expect the computations to be closer to the computations done in [Reference Garcia3].
2 The Kudla–Millson form
2.1 The symmetric space
$\mathbb {D}$
Let
$(V,Q)$
be a rational quadratic space, and let
$(p,q)$
be the signature of
$V(\mathbb {R})$
. Let
$e_1, \dots , e_{p+q}$
be an orthogonal basis of
$V(\mathbb {R})$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn10.png?pub-status=live)
Note that we will always use letters
$\alpha $
and
$\beta $
for indices between
$1$
and p, and letters
$\mu $
and
$\nu $
for indices between
$p+1$
and
$p+q$
. A plane z in
$V(\mathbb {R})$
is a negative plane if
$Q\big|{}_{z} $
is negative definite. Let
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn11.png?pub-status=live)
be the set of negative-oriented q-planes in
$V(\mathbb {R})$
. For each negative plane, there are two possible orientations, yielding two connected components
$\mathbb {D}^+ $
and
$ \mathbb {D}^-$
of
$\mathbb {D}$
. Let
$z_0$
in
$\mathbb {D}^+$
be the negative plane spanned by the vectors
$e_{p+1}, \dots , e_{p+q}$
together with a fixed orientation. The group
$G(\mathbb {R})^+$
acts transitively on
$\mathbb {D}^+$
by sending
$z_0$
to
$gz_0$
. Let K be the stabilizer of
$z_0$
, which is isomorphic to
$\operatorname {SO}(p)\times \operatorname {SO}(q)$
. Thus, we have an identification
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn12.png?pub-status=live)
For z in
$\mathbb {D}^+$
, we denote by
$g_z$
any element of
$G(\mathbb {R})^+$
sending
$z_0$
to z.
For a positive vector v in
$V(\mathbb {R}),$
we define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn13.png?pub-status=live)
It is a totally geodesic submanifold of
$\mathbb {D}$
of codimension q. Let
$\mathbb {D}_v^+$
be the intersection of
$\mathbb {D}_v$
with
$\mathbb {D}^+$
.
Let z in
$\mathbb {D}^+$
be a negative plane. With respect to the orthogonal splitting of
$V(\mathbb {R})$
as
$z^\perp \oplus z$
, the quadratic form splits as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn14.png?pub-status=live)
We define the Siegel majorant at z to be the positive-definite quadratic form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn15.png?pub-status=live)
2.2 The Lie algebras
$\mathfrak {g}$
and
$\mathfrak {k}$
Let
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn16.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn17.png?pub-status=live)
be the Lie algebras of
$G(\mathbb {R})^+$
and
$K,$
where
$\mathfrak {so}(z_0)$
is equal to
$\mathfrak {so}(q)$
. The latter is the space of skew-symmetric q by q matrices. Similarly, we have
$\mathfrak {so}(z_0^\perp )$
equals
$\mathfrak {so}(p)$
. Hence, we have a decomposition of
$\mathfrak {k}$
as
$\mathfrak {so}(z_0^\perp ) \oplus \mathfrak {so}(z_0)$
that is orthogonal with respect to the Killing form. Let
$\epsilon $
be the Lie algebra involution of
$\mathfrak {g}$
mapping X to
$-X$
. The
$+1$
-eigenspace of
$\epsilon $
is
$\mathfrak {k}$
and the
$-1$
-eigenspace is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn18.png?pub-status=live)
We have a decomposition of
$\mathfrak {g} $
as
$\mathfrak {k} \oplus \mathfrak {p}$
and it is orthogonal with respect to the Killing form. We can identify
$\mathfrak {p}$
with
$\mathfrak {g}/\mathfrak {k}$
. Since
$\epsilon $
is a Lie algebra automorphism, we have that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn19.png?pub-status=live)
We identify the tangent space of
$\mathbb {D}^+$
at
$eK$
with
$\mathfrak {p}$
and the tangent bundle
$T\mathbb {D}^+$
with
$G(\mathbb {R})^+ \times _K \mathfrak {p}$
, where K acts on
$\mathfrak {p}$
by the
$\operatorname {Ad}$
-representation. We have an isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn20.png?pub-status=live)
A basis of
$\mathfrak {g}$
is given by the set of matrices
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn21.png?pub-status=live)
and we denote by
$\omega _{ij}$
, its dual basis in the dual space
$\mathfrak {g}^\ast $
. Let
$E_{ij}$
be the elementary matrix sending
$e_i$
to
$e_j$
and the other
$e_k$
’s to
$0$
. Then
$\mathfrak {p}$
is spanned by the matrices
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn22.png?pub-status=live)
and
$\mathfrak {k}$
is spanned by the matrices
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn23.png?pub-status=live)
2.3 Poincaré duals
Let M be an arbitrary m-dimensional real orientable manifold without boundary. The integration map yields a nondegenerate pairing [Reference Bott and Tu2, Theorem 5.11]
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn24.png?pub-status=live)
where
$H_c(M)$
denotes the cohomology of compactly supported forms on M. This yields an isomorphism between
$H^{q}(M)$
and the dual
$H_c^{m-q}(M)^\ast =\operatorname {Hom}(H_c^{m-q}(M),\mathbb {R})$
. If C is an immersed submanifold of codimension q in M, then C defines a linear functional on
$H_c^{m-q}(M)$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn25.png?pub-status=live)
Since we have an isomorphism between
$H_c^{m-q}(M)^\ast $
and
$H^{q}(M)$
, there is a unique cohomology class
$\operatorname {PD}(C)$
in
$H^q(M)$
representing this functional, i.e.,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn26.png?pub-status=live)
for every class
$[\omega ]$
in
$H_c^{m-q}(M)$
. We call
$\operatorname {PD}(C)$
the Poincaré dual class to C, and any differential form representing the cohomology class
$\operatorname {PD}(C)$
a Poincaré dual form to C.
2.4 The Kudla–Millson form
The tangent plane at the identity
$T_{eK} \mathbb {D}^+ $
can be identified with
$\mathfrak {p}$
and the cotangent bundle
$(T\mathbb {D}^+)^\ast $
with
$G(\mathbb {R})^+ \times _K \mathfrak {p}^\ast $
, where K acts on
$\mathfrak {p}^\ast $
by the dual of the
$\operatorname {Ad}$
-representation. The basis
$e_1, \dots , e_{p+q}$
identifies
$V(\mathbb {R})$
with
$\mathbb {R}^{p+q}$
. With respect to this basis, the Siegel majorant at
$z_0$
is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn27.png?pub-status=live)
Recall that
$G(\mathbb {R})^+$
acts on
$\mathscr {S}(\mathbb {R}^{p+q})$
from the left by
$(g \cdot f)(v)= f(g^{-1}v)$
and on
$\Omega ^q(\mathbb {D}^+) \otimes \mathscr {S}(\mathbb {R}^{p+q})$
from the right by
. We have an isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn28.png?pub-status=live)
by evaluating
$\varphi $
at the basepoint
$eK$
in
$G(\mathbb {R})^+/K$
, corresponding to the point
$z_0$
in
$\mathbb {D}^+$
. We define the Howe operator
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn29.png?pub-status=live)
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn30.png?pub-status=live)
where
$A_{\alpha \mu }$
denotes left multiplication by
$\omega _{\alpha \mu }$
. The Kudla–Millson form is defined by applying D to the Gaussian:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn31.png?pub-status=live)
Kudla and Millson showed that this form is K-invariant. Hence, by the isomorphism (2.19), we get a form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn32.png?pub-status=live)
In particular, since
$g^\ast \varphi _{KM}(v)=\varphi _{KM}(g^{-1}v)$
for any
$g \in G(\mathbb {R})^+$
, the form is
$\Gamma _v$
-invariant and defines a form on
$\Gamma _v \backslash \mathbb {D}^+$
. It is also closed and Kudla–Millson prove in [Reference Kudla and Millson8, Proposition 5.2] that it satisfies the Thom form property: for every compactly supported form
$\omega $
in
$\Omega ^{pq-q}_c(\Gamma _v \backslash \mathbb {D}^+)$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn33.png?pub-status=live)
3 The Mathai–Quillen formalism
We begin by recalling a few facts about principal bundles, connections, and associated vector bundles. For more details, we refer to [Reference Berline, Getzler and Vergne1, Reference Kobayashi and Nomizu5]. The Mathai–Quillen form is defined in Section 3.7 following [Reference Berline, Getzler and Vergne1] (see also [Reference Getzler, Cruzeiro and Zambrini4]).
3.1 K-principal bundles and principal connections
Let K be
$\operatorname {SO}(p)\times \operatorname {SO}(q)$
as before, and let P be a smooth principal K-bundle. Let
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn34.png?pub-status=live)
be the smooth right action of K on P and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn35.png?pub-status=live)
the projection map. For a fixed
$p $
in P, consider the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn36.png?pub-status=live)
Let
$V_pP$
be the image of the derivative at the identity
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn37.png?pub-status=live)
which is injective. It coincides with the kernel of the differential
$d_p\pi $
. A vector in
$V_pP$
is called a vertical vector. Using this map, we can view a vector X in
$\mathfrak {k}$
as a vertical vector field on P. The space P can a priori be arbitrary, but in our case, we will consider either:
-
(1) P is
$G(\mathbb {R})^+$ and
$R_k$ the natural right action sending g to
$gk$ . Then
$P/K$ can be identified with
$\mathbb {D}^+$ .
-
(2) P is
$G(\mathbb {R})^+ \times z_0$ and the action
$R_k$ maps
$(g,w)$ to
$(gk,k^{-1}w)$ . In this case,
$P/K$ can be identified with
$G(\mathbb {R})^+ \times _K z_0$ . It is the vector bundle associated with the principal bundle
$G(\mathbb {R})^+$ as defined below.
A principal K-connection on P is a
$1$
-form
$\theta _P $
in
$\Omega ^1(P, \mathfrak {k})$
such that:
-
•
$\iota _X \theta _P = X$ for any
$X $ in
$\mathfrak {k}$ ,
-
•
$R_k^\ast \theta _P=Ad(k^{-1}) \theta _P \quad $ for any k in K,
where
$\iota _X$
is the interior product
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn38.png?pub-status=live)
and we view X as a vector field on P. Geometrically, these conditions imply that the kernel of
$\theta _P$
defines a horizontal subspace of
$TP$
that we denote by
$HP$
. It is a complement to the vertical subspace, i.e., we get a splitting of
$T_pP$
as
$V_pP \oplus H_pP$
.
Let
$\mathfrak {g}$
be the Lie algebra of
$G(\mathbb {R})^+$
, and let
$\mathcal {P}$
be the orthogonal projection from
$\mathfrak {g}$
on
$\mathfrak {k}$
. After identifying
$\mathfrak {g}^\ast $
with the space
$\Omega ^1(G(\mathbb {R})^+)^{G(\mathbb {R})^+}$
of
$G(\mathbb {R})^+$
-invariant forms, we define a natural
$1$
-form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn39.png?pub-status=live)
called the Maurer–Cartan form, where
$X_{ij}$
is the basis of
$\mathfrak {g}$
defined earlier and
$\omega _{ij}$
its dual in
$\mathfrak {g}^\ast $
. After projection onto
$\mathfrak {k}$
, we get a form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn40.png?pub-status=live)
where we identify
$\Omega ^1(G(\mathbb {R})^+, \mathfrak {k})$
with
$\Omega ^1(G(\mathbb {R})^+) \otimes \mathfrak {k}$
. A direct computation shows that it is a principal K-connection on P, when P is
$G(\mathbb {R})^+$
.
If P is
$G(\mathbb {R})^+ \times z_0$
, then the projection
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn41.png?pub-status=live)
induces a pullback map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn42.png?pub-status=live)
The form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn43.png?pub-status=live)
is a principal connection on
$G(\mathbb {R})^+ \times z_0$
.
3.2 The associated vector bundles
Since
$z_0$
is preserved by K, we have an orthogonal K-representation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn44.png?pub-status=live)
where we will usually simply write
$kw$
instead of
${k\big|{}_{z_0} }w$
. We can consider the associated vector bundle
$P \times _K z_0$
which is the quotient of
$P \times z_0$
by K, where K acts by sending
$(p,w)$
to
$ (R_k(p), \rho (k)^{-1}w)$
. Hence, an element
$[p,w]$
of
$P \times _K z_0$
is an equivalence class where the equivalence relation identifies
$(p,w)$
with
$(R_k(p), \rho (k)^{-1}w)$
. This is a vector bundle over
$P/K$
with projection map sending
$[p,w]$
to
$\pi (p)$
. Let
$\Omega ^i(P/K,P \times _Kz_0)$
be the space of i-forms valued in
$P \times _Kz_0$
, when i is zero it is the space of smooth sections of the associated bundle.
In the two cases of interest to us, we define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn45.png?pub-status=live)
Note that in both cases, P admits a left action of
$G(\mathbb {R})^+$
and that the associated vector bundles are
$G(\mathbb {R})^+$
-equivariant. Moreover, it is a Euclidean bundle, equipped with the inner product
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn46.png?pub-status=live)
on the fiber. Let
$ \Omega ^i(P,z_0)$
be the space of
$z_0$
-valued differential i-forms on P. A differential form
$\alpha $
in
$\Omega ^i(P,z_0)$
is said to be horizontal if
$\iota _X\alpha $
vanishes for all vertical vector fields X. There is a left action of K on a differential form
$\alpha $
in
$\Omega ^i(P,z_0)$
defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn47.png?pub-status=live)
and
$\alpha $
is K-invariant if it satisfies
$k\cdot \alpha = \alpha $
for any k in
$K,$
i.e., we have
$R_k^\ast \alpha = \rho (k^{-1}) \alpha $
. We write
$\Omega ^i(P, z_0)^K$
for the space of K-invariant
$z_0$
-valued forms on P. Finally, a form that is horizontal and K-invariant is called a basic form and the space of such forms is denoted by
$\Omega ^i(P,z_0)_{\mathrm{bas}}$
.
Let
$X_1, \dots , X_N$
be tangent vectors of
$P/K$
at
$\pi (p),$
and let
$\widetilde {X}_i$
be tangent vectors of P at p that satisfy
$d_p\pi (\widetilde {X}_i)=X_i$
. There is a map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn48.png?pub-status=live)
defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn49.png?pub-status=live)
Proposition 3.1 The map is well-defined and yields an isomorphism between
$\Omega ^i(P/K,P \times _K z_0)$
and
$\Omega ^i(P,z_0)_{\mathrm{bas}}$
. In particular, if
$z_0$
is one-dimensional, then
$\Omega ^i(P/K)$
is isomorphic to
$\Omega ^i(P)_{\mathrm{bas}}$
.
Proof In the case where i is zero, the horizontally condition is vacuous and the isomorphism simply identifies
$\Omega ^0(P/K,P \times _K z_0)$
with
$\Omega ^0(P,z_0)^K$
. We have a map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn50.png?pub-status=live)
which is well defined since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn51.png?pub-status=live)
Conversely, every smooth section s in
$\Omega ^0(P/K,P \times _K z_0)$
is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn52.png?pub-status=live)
for some smooth function
$f_s$
in
$\Omega ^0(P,z_0)^K$
. The map sending s to
$f_s$
is inverse to the previous one. The proof is similar for positive i.
3.3 Covariant derivatives
A covariant derivative on the vector bundle
$P \times _K z_0$
is a differential operator
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn53.png?pub-status=live)
such that for every smooth function f in
$C^\infty (P/K),$
we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn54.png?pub-status=live)
The inner product on
$P \times _Kz_0$
defines a pairing
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn55.png?pub-status=live)
and we say that the derivative is compatible with the metric if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn56.png?pub-status=live)
for any two sections
$s_1$
and
$s_2$
in
$\Omega ^0(P/K,P\times _Kz_0)$
. There is a covariant derivative that is induced by a principal connection
$\theta _P$
in
$\Omega ^1(P) \otimes \mathfrak {k}$
as follows. The derivative of the representation gives a map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn57.png?pub-status=live)
which we also denote by
$\rho $
by abuse of notation. Note that for the representation (3.11), this is simply the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn58.png?pub-status=live)
since
$\mathfrak {k}$
splits as
$\mathfrak {so}(z_0^\perp ) \oplus \mathfrak {so}(z_0)$
. Composing the principal connection with
$\rho $
defines an element
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn59.png?pub-status=live)
In particular, if s is a section of
$P \times _K z_0$
, then we can identify it with a K-invariant smooth map
$f_s$
in
$\Omega ^0(P, z_0)^K$
. Since
$\rho (\theta _P)$
is a
$\mathfrak {so}(z_0)$
-valued form and
$\mathfrak {so}(z_0)$
is a subspace of
$\operatorname {End}(z_0),$
we can define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn60.png?pub-status=live)
Lemma 3.2 The form
$df_s+\rho (\theta _P) \cdot f_s$
is basic, hence gives a
$P \times _K z_0$
-valued form on
$P/K$
. Thus,
$d+\rho (\theta _P)$
defines a covariant derivative on
$P \times _K z_0$
. Moreover, it is compatible with the metric.
Proof See [Reference Berline, Getzler and Vergne1, p. 24]. For the compatibility with the metric, it follows from the fact that the connection
$\rho (\theta _P)$
is valued in
$\mathfrak {so}(z_0)$
that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn61.png?pub-status=live)
Hence, if we denote by
$\nabla _P$
is the covariant derivative defined by
$d+\rho (\theta _P),$
then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn62.png?pub-status=live)
Let us denote by
$\nabla _P$
the covariant derivative
$d+\rho (\theta _P)$
. It can be extended to a map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn63.png?pub-status=live)
by setting
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn64.png?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn65.png?pub-status=live)
We define the curvature
$R_P$
in
$\Omega ^2(P,\mathfrak {k})$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn66.png?pub-status=live)
for two vector fields X and Y on P. It is basic by [Reference Berline, Getzler and Vergne1, Proposition 1.13] and composing with
$\rho $
gives an element
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn67.png?pub-status=live)
so that we can view it as an element in
$\Omega ^2(P/K,P \times _K \mathfrak {so}(z_0)),$
where K acts on
$\mathfrak {so}(z_0)$
by the
$\operatorname {Ad}$
-representation. For a section s in
$\Omega ^0(P/K,P \times _K z_0)$
, we have [Reference Berline, Getzler and Vergne1, Proposition 1.15]
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn68.png?pub-status=live)
From now on, we denote by
$\nabla $
and
$\widetilde {\nabla }$
the covariant derivatives on E and
$\widetilde {E}$
associated with
$\theta $
and
$\widetilde {\theta }$
defined in (3.7) and (3.10). Let R and
$\widetilde {R}$
be their respective curvatures.
3.4 Pullback of bundles
The pullback of E by the projection map gives a canonical bundle
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn69.png?pub-status=live)
over E. We have the following diagram:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn70.png?pub-status=live)
The projection induces a pullback of the sections
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn71.png?pub-status=live)
We can also pullback the covariant derivative
$\nabla $
to a covariant derivative
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn72.png?pub-status=live)
on
$\pi ^\ast E$
. It is characterized by the property
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn73.png?pub-status=live)
Proposition 3.3 The bundles
$\widetilde {E}$
and
$\pi ^\ast E$
are isomorphic, and this isomorphism identifies
$\widetilde {\nabla }$
and
$\pi ^\ast \nabla $
.
Proof By definition,
$([g_1,w_1],[g_2,w_2])$
are elements of
$ \pi ^\ast E$
if and only if
$g^{-1}_1g_2$
is in K. We have a
$G(\mathbb {R})^+$
-equivariant morphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn74.png?pub-status=live)
This map is well defined and has as inverse
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn75.png?pub-status=live)
The second statement follows from the fact that
$\widetilde {\theta }$
is
$\pi ^\ast \theta $
.
3.5 A few operations on the vector bundles
We extend the K-representation
$z_0$
to
$\bigwedge ^j z_0$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn76.png?pub-status=live)
We consider the bundles
$P \times _K \wedge ^j z_0$
and
$ P \times _K \wedge z_0$
over
$P/K$
, where
$\bigwedge z_0$
is defined as
$ \bigoplus _i \bigwedge ^iz_0$
. Denote the space of differential forms valued in
$P \times _K \wedge ^j z_0$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn77.png?pub-status=live)
The total space of differential forms
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn78.png?pub-status=live)
is an (associative) bigraded
$C^\infty (P/K)$
-algebra, where the product is defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn79.png?pub-status=live)
This algebra structure allows us to define an exponential map by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn80.png?pub-status=live)
where
$\omega ^k$
is the k-fold wedge product
$\omega \wedge \cdots \wedge \omega $
.
Remark 3.1 Suppose that
$\omega $
and
$\eta $
commute. Then the binomial formula
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn81.png?pub-status=live)
holds and one can show that
$\exp (\omega +\eta )=\exp (\omega )+\exp (\eta )$
in the same way as for the real exponential map. In particular, the diagonal subalgebra
$\bigoplus \Omega _P^{i,i}$
is a commutative, since for two forms
$\omega $
and
$\eta $
in
$\Omega _P$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn82.png?pub-status=live)
and similarly for two sections s and t in
$\Omega ^0(P/K,P \times _K z_0)$
.
The inner product
$\langle - , - \rangle $
on
$z_0$
can be extended to an inner product on
$\bigwedge z_0$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn83.png?pub-status=live)
If
$e_1, \dots , e_q$
is an orthonormal basis of
$z_0$
, then the set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn84.png?pub-status=live)
is an orthonormal basis of
$\bigwedge z_0$
. We define the Berezin integral
$\int ^B$
to be the orthogonal projection onto the top dimensional component, that is the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn85.png?pub-status=live)
The Berezin integral can then be extended to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn86.png?pub-status=live)
where
$\int ^B s$
in
$C^\infty (P/K)$
is the composition of the section with the Berezinian in every fiber. Let
$s_1, \dots , s_q$
be a local orthonormal frame of
$P \times _K z_0$
. Then
$s_1 \wedge \cdots \wedge s_q$
is in
$\Omega ^0(P/K,\wedge ^q P \times _K z_0)$
and defines a global section. Hence, for
$\alpha $
in
$\Omega (P/K,P\times _K \wedge z_0),$
we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn87.png?pub-status=live)
Finally, for every section s in
$\Omega ^{0,1}$
, we can define the contraction
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn88.png?pub-status=live)
and extended by linearity, where the symbol
$ \, \widehat {\cdot } \,$
means that we remove it from the product. Note that when j is zero, then
$i(s)$
is defined to be zero. The contraction
$i(s)$
defines a derivation on
$\oplus \widetilde {\Omega }^{i,j}$
that satisfies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn89.png?pub-status=live)
for
$\alpha $
in
$\widetilde {\Omega }^{i,j}$
and
$\alpha '$
in
$\widetilde {\Omega }^{k,l}$
.
3.6 Thom forms
We denote by E the bundle
$G(\mathbb {R})^+\times _K z_0$
. On the fibers of the bundle, we have the inner product given by
. Let v be arbitrary vector in L and
$\Gamma _v$
its stabilizer. Since the bundle is
$G(\mathbb {R})^+$
-equivariant, we have a bundle
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn90.png?pub-status=live)
and let
$\operatorname {D}(\Gamma _v \backslash E)$
be the closed disk bundle. If we have a closed
$(q+i)$
-form on
$\Gamma _v \backslash E$
whose support is contained in
$\operatorname {D}(\Gamma _v \backslash E)$
, then it has compact support in the fiber and represents a class in
$H^{q+i}(\Gamma _v \backslash E,\Gamma _v \backslash E-\operatorname {D}(\Gamma _v \backslash E))$
. The cohomology group
$H^{\bullet }(\Gamma _v \backslash E,\Gamma _v \backslash E-\operatorname {D}(\Gamma _v \backslash E))$
is equal to the cohomology group
$H^{\bullet }(\Gamma _v \backslash E,\Gamma _v \backslash (E-E_0))$
that we used in the introduction, where
$E_0$
is the zero section. Fiber integration induces an isomorphism on the level of cohomology
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn91.png?pub-status=live)
known as the Thom isomorphism [Reference Bott and Tu2, Theorem 6.17]. When i is zero, then
$H^i(\Gamma _v \backslash \mathbb {D}^+)$
is
$\mathbb {R}$
and we call the preimage of
$1$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn92.png?pub-status=live)
the Thom class. Any differential form representating this class is called a Thom form, in particular, every closed q-form on
$\Gamma _v \backslash E$
that has compact support in every fiber and whose integral along every fiber is
$1$
is a Thom form. One can also view the Thom class as the Poincaré dual class of the zero section
$E_0$
in E, in the same sense as for (2.24).
Let
$\omega $
in
$\Omega ^j(E)$
be a form on the bundle, and let
$\omega _z$
be its restriction to a fiber
$E_z=\pi ^{-1}(z)$
for some z in
$\mathbb {D}^+$
. After identifying
$z_0$
with
$\mathbb {R}^q$
, we see
$\omega _z$
as an element of
$C^\infty (\mathbb {R}^q) \otimes \wedge ^j(\mathbb {R}^q)^\ast $
. We say that
$\omega $
is rapidly decreasing in the fiber, if
$\omega _z$
lies in
$\mathscr {S}(\mathbb {R}^q) \otimes \wedge ^j(\mathbb {R}^q)^\ast $
for every z in
$\mathbb {D}^+$
. We write
$\Omega ^j_{\textrm {rd}}(E)$
for the space of such forms.
Let
$\Omega ^\bullet _{\textrm {rd}}(\Gamma _v \backslash E)$
be the complex of rapidly decreasing forms in the fiber. It is isomorphic to the complex
$\Omega ^\bullet _{\textrm {rd}}(E)^{\Gamma _v}$
of rapidly decreasing
$\Gamma _v$
-invariant forms on E. Let
$H_{\textrm {rd}}(\Gamma _v \backslash E)$
the cohomology of this complex. The map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn93.png?pub-status=live)
is a diffeomorphism from the open disk bundle
$\operatorname {D}(\Gamma _v \backslash E)^\circ $
onto
$\Gamma _v \backslash E$
. It induces an isomorphism by pullback
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn94.png?pub-status=live)
which commutes with the fiber integration. Hence, we have the following version of the Thom isomorphism:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn95.png?pub-status=live)
The construction of Mathai and Quillen produces a Thom form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn96.png?pub-status=live)
which is
$G(\mathbb {R})^+$
-invariant (hence,
$\Gamma _v$
-invariant) and closed. We will recall their construction in the next section.
3.7 The Mathai–Quillen construction
As earlier, let
$\widetilde {E}$
be the bundle
$(G(\mathbb {R})^+ \times z_0) \times _K z_0$
. Let
$ \wedge ^j \tilde {E} $
be the bundle
$(G(\mathbb {R})^+ \times z_0) \times _K \wedge ^j z_0$
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn97.png?pub-status=live)
First, consider the tautological section
$\mathbf{s}$
of
$\widetilde {E}$
defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn98.png?pub-status=live)
This gives a canonical element
$\mathbf{s}$
of
$\widetilde {\Omega }^{0,1}$
. Composing with the norm induced from the inner product, we get an element
$\lVert \mathbf{s} \rVert ^2$
in
$\widetilde {\Omega }^{0,0}$
.
The representation
$\rho $
on
$z_0$
induces a representation on
$\wedge ^iz_0$
that we also denote by
$\rho $
. The derivative at the identity gives a map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn99.png?pub-status=live)
The connection form
$\rho (\widetilde {\theta })$
in
$\Omega ^1(G(\mathbb {R})^+ \times z_0,\wedge ^j z_0)$
defines a covariant derivative
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn100.png?pub-status=live)
on
$\wedge ^j \widetilde {E}$
. We can extend it to a map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn101.png?pub-status=live)
by setting
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn102.png?pub-status=live)
as in (3.30). The connection on
$\widetilde {\Omega }^{i,j}$
is compatible with the metric. Finally, the covariant derivative
$\widetilde {\nabla }$
defines a derivation on
$\oplus \widetilde {\Omega }^{i,j}$
that satisfies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn103.png?pub-status=live)
for any
$\alpha $
in
$\widetilde {\Omega }^{i,j}$
and
$\alpha '$
in
$\widetilde {\Omega }^{k,l}$
.
Taking the derivative of the tautological section gives an element
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn104.png?pub-status=live)
Let
$\mathfrak {so}(\widetilde {E})$
denote the bundle
$(G(\mathbb {R})^+ \times z_0) \times _K \mathfrak {so}(z_0)$
and consider the curvature
$\rho (\widetilde {R})$
in
$\Omega ^2(\widetilde {E}, \mathfrak {so}(\widetilde {E}))$
. We have an isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn105.png?pub-status=live)
The inverse sends
$v \wedge w$
to the endomorphism
$ u \mapsto \langle v,u \rangle w-\langle w,u \rangle v$
, and is the isomorphism from (2.11) restricted to
$z_0$
. Note that we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn106.png?pub-status=live)
Using this isomorphism, we can also identify
$\mathfrak {so}(\widetilde {E})$
and
$\wedge ^2 \widetilde {E}$
so that we can view the curvature as an element
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn107.png?pub-status=live)
Lemma 3.4 The form
lying in
$\widetilde {\Omega }^{0,0} \oplus \widetilde {\Omega }^{1,1} \oplus \widetilde {\Omega }^{2,2}$
is annihilated by
$\widetilde {\nabla }+ 2 \sqrt {\pi } i(\mathbf{s})$
. Moreover
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn108.png?pub-status=live)
for every form
$\alpha $
in
$\widetilde {\Omega }^{i,j}$
. Hence,
$\int ^B exp(-\omega )$
is a closed form.
Proof We have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn109.png?pub-status=live)
It vanishes, because we have the following:
-
∙
$i(\mathbf{s}) \lVert \mathbf{s} \rVert ^2=0$ since
$\lVert \mathbf{s} \rVert $ is in
$\widetilde {\Omega }^{0,0}$ ,
-
∙
$\widetilde {\nabla } \rho (\widetilde {R})=0$ by Bianchi’s identity,
-
∙
$\widetilde {\nabla }\lVert \mathbf{s} \rVert ^2= 2\langle \widetilde {\nabla }\mathbf{s},\mathbf{s} \rangle =-2 i(\mathbf{s}) \widetilde {\nabla }\mathbf{s} $ ,
-
∙
$\widetilde {\nabla }^2\mathbf{s}=\rho (\widetilde {R})\mathbf{s} =i(\mathbf{s})\rho (\widetilde {R})$ .
For the last point, we used (3.73), where we view
$\rho (\widetilde {R})$
as an element of
$\Omega ^2(E,\mathfrak {so}(\widetilde {E}))$
, respectively of
$\Omega ^2(E,\wedge ^2\widetilde {E})$
.
Let
$s_1 \wedge \cdots \wedge s_q$
in
$\Omega ^0(E,\wedge ^q \widetilde {E})$
be a global section, where
$s_1, \dots , s_q$
is a local orthonormal frame for
$\widetilde {E}$
. Then, for any
$\alpha $
in
$\widetilde {\Omega }^{i,j}$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn110.png?pub-status=live)
This vanishes if j is different from q, hence we can assume
$\alpha $
is in
$\widetilde {\Omega }^{i,q}$
. If we write
$\alpha $
as
$\beta s_1 \wedge \cdots \wedge s_q$
for some
$\beta $
in
$\Omega ^i(E)$
, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn111.png?pub-status=live)
On the other hand, since the connection on
$\widetilde {\Omega }^{i,q}$
is compatible with the metric, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn112.png?pub-status=live)
Then we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn113.png?pub-status=live)
Since
$\widetilde {\nabla } + 2 \sqrt {\pi }i(\mathbf{s})$
is a derivation that annihilates
$\omega $
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn114.png?pub-status=live)
for positive k. Hence, it follows that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn115.png?pub-status=live)
In [Reference Mathai and Quillen10], Mathai and Quillen define the following form:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn116.png?pub-status=live)
We call it the Mathai–Quillen form.
Proposition 3.5 The Mathai–Quillen form is a Thom form.
Proof From the previous lemma, it follows that the form is closed. It remains to show that its integral along the fibers is
$1$
. The restriction of the form
$U_{MQ}$
along the fiber
$\pi ^{-1}(eK)$
is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn117.png?pub-status=live)
and its integral over the fiber
$\pi ^{-1}(eK)$
is equal to
$1$
.
4 Computation of the Mathai–Quillen form
4.1 The section
$s_v$
Let
$\textrm {pr}$
denote the orthogonal projection of
$V(\mathbb {R})$
on the plane
$z_0$
. Consider the section
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn118.png?pub-status=live)
where
$g_z$
is any element of
$G(\mathbb {R})^+$
sending
$z_0$
to z. Let us denote by
$L_g$
the left action of an element g in
$G(\mathbb {R})^+$
on
$\mathbb {D}^+$
. We also denote by
$L_g$
the action on E given by
$L_{g}[g_z,v]=[gg_z,v]$
. The bundle is
$G(\mathbb {R})^+$
-equivariant with respect to these actions.
Proposition 4.1 The section
$s_v$
is well-defined and
$\Gamma _v$
-equivariant. Moreover, its zero locus is precisely
$\mathbb {D}^+_v$
.
Proof The section is well-defined, since replacing
$g_z$
by
$g_zk$
gives
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn119.png?pub-status=live)
Suppose that z is in the zero locus of
$s_v$
, that is to say
$\textrm {pr}(g_z^{-1}v)$
vanishes. Then
$g_z^{-1}v$
is in
$z_0^\perp $
. It is equivalent to the fact that
$z=g_zz_0$
is a subspace of
$v^\perp $
, which means that z is in
$\mathbb {D}_v^+$
. Hence, the zero locus of
$s_v$
is exactly
$\mathbb {D}^+_v$
. For the equivariance, note that we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn120.png?pub-status=live)
Hence, if
$\gamma $
is an element of
$\Gamma _v$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn121.png?pub-status=live)
We define the pullback
of the Mathai–Quillen form by
$s_v$
. It defines a form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn122.png?pub-status=live)
It is only rapidly decreasing on
$\mathbb {R}^q$
, and in order to make it rapidly decreasing everywhere we set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn123.png?pub-status=live)
It defines a form
$ \varphi \in \mathscr {S}(\mathbb {R}^{p+q}) \otimes \Omega ^q(\mathbb {D})^+$
.
Proposition 4.2
-
(1) For fixed v in
$V(\mathbb {R}),$ the form
$\varphi ^0(v)$ in
$\Omega ^q(\mathbb {D}^+)$ is given by
(4.7)$$ \begin{align} \varphi^0(v) & = (- 1)^{\frac{q(q+1)}{2}} (2\pi)^{-\frac{q}{2}} \exp \left (2 \pi {Q\big|{}_{z_0} }(v,v) \right ) \int^B \exp \left (-2 \sqrt{\pi} \nabla s_v+\rho(R)\right ). \end{align} $$
-
(2) It satisfies
$L_g^\ast \varphi ^0(v)=\varphi ^0(g^{-1}v)$ , hence
(4.8)$$ \begin{align} \varphi^0 \in \left [\Omega^q(\mathbb{D}^+) \otimes C^\infty(\mathbb{R}^{p+q}) \right ]^{G(\mathbb{R})^+}.\end{align} $$
-
(3) It is a Poincaré dual of
$\Gamma _v \backslash \mathbb {D}_v^+$ in
$\Gamma _v \backslash \mathbb {D}^+$ .
Proof
-
(1) Recall that
$\widetilde {\nabla }=\pi ^\ast \nabla $ and
$\widetilde {R}=\pi ^\ast R$ . We pullback by
$s_v$
Since
$\pi \circ s_v$ is the identity, we have
(4.9)$$ \begin{align} s_v^\ast \widetilde{\nabla} = s_v^\ast \pi^\ast \nabla = \nabla. \end{align} $$
Hence, the pullback connection
$s_v^\ast \widetilde {\nabla }$ satisfies
(4.10)since$$ \begin{align} s_v^\ast (\widetilde{\nabla} \mathbf{s})= (s_v^\ast \widetilde{\nabla}) ( s_v^\ast \mathbf{s}) = \nabla s_v, \end{align} $$
$s_v^\ast \mathbf{s}=s_v$ . We also have
$s_v^\ast \widetilde {R}=R$ and
(4.11)$$ \begin{align}s_v^\ast \lVert \mathbf{s} \rVert^2= \lVert s_v \rVert^2= \langle s_v , s_v \rangle=-{Q\big|{}_{z_0} }(v,v). \end{align} $$
The expression for
$\varphi ^0$ then follows from the fact that
$\exp $ and
$s_v^\ast $ commute.
-
(2) The bundle E is
$G(\mathbb {R})^+$ equivariant. By construction, the Mathai–Quillen form is
$G(\mathbb {R})^+$ -invariant, so
$L_g^\ast U_{MQ}=U_{MQ}$ . On the other hand, we also have
(4.12)$$ \begin{align}s_v \circ L_g(z)=L_g \circ s_{g^{-1}v}(z),\end{align} $$
and thus,
(4.13)$$ \begin{align} L_g^\ast \varphi^0(v)=L_g^\ast s_v^\ast U_{MQ}=\varphi^0(g^{-1}v). \end{align} $$
-
(3) Since
$s_v$ is
$\Gamma _v$ -equivariant, we view it as a section
(4.14)whose zero locus is precisely$$ \begin{align} s_v \colon \Gamma_v \backslash \mathbb{D}^+ \longrightarrow \Gamma_v \backslash E, \end{align} $$
$\Gamma _v \backslash \mathbb {D}_v^+$ . Let
$S_0$ (resp.
$S_v$ ) be the image in
$\Gamma _v \backslash E$ of the section
$s_v$ (resp. the zero section). By [Reference Bott and Tu2, Proposition 6.24(b)], the Thom form
$U_{MQ}$ is a Poincaré dual of the zero section
$S_0$ of E. For a form
$\omega $ in
$\Omega _c^{m-q}(\Gamma _v \backslash \mathbb {D}^+),$ we have
(4.15)$$ \begin{align} \int_{\Gamma_v \backslash \mathbb{D}^+} \varphi^0(v) \wedge \omega & = \int_{\Gamma_v \backslash \mathbb{D}^+} s_v^\ast \left ( U_{MQ} \wedge \pi ^\ast \omega \right ) \nonumber \\ & = \int_{S_v} U_{MQ} \wedge \pi ^\ast \omega \nonumber \\ & = \int_{S_v \cap S_0} \pi ^\ast \omega \nonumber \\ & = \int_{\Gamma_v \backslash \mathbb{D}_v^+} \omega. \end{align} $$
The last step follows from the fact that
$\pi ^{-1}(S_v \cap S_0)$ equals
$\Gamma _v \backslash \mathbb {D}_v^+$ .
As in (2.19), we have an isomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn133.png?pub-status=live)
by evaluating at the basepoint
$eK$
of
$G(\mathbb {R})^+/K$
that corresponds to
$z_0$
in
$\mathbb {D}^+$
. We will now compute
${ {\varphi ^0}\big|{}_{eK} }$
.
4.2 The Mathai–Quillen form at the identity
From now on, we identify
$\mathbb {R}^{p+q}$
with
$V(\mathbb {R})$
by the orthonormal basis of (2.1), and let
$z_0$
be the negative spanned by the vectors
$e_{p+1}, \ldots , e_{p+q}$
. Hence, we identify
$z_0$
with
$\mathbb {R}^q$
and the quadratic form is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn134.png?pub-status=live)
where
$x_{p+1}, \dots , x_{p+q}$
are the coordinates of the vector v.
Let
$f_v$
in
$\Omega ^0(G(\mathbb {R})^+,z_0)^K$
be the map associated with the section
$s_v$
, as in Proposition 3.1. It is defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn135.png?pub-status=live)
Then
$df_v+\rho (\theta ) f_v$
is the horizontal lift of
$\nabla s_v$
, as discussed in Section 3.1. Let X be a vector in
$\mathfrak {g}$
, and let
$X_{\mathfrak {p}}$
and
$X_{\mathfrak {k}}$
be its components with respect to the splitting of
$\mathfrak {g}$
as
$\mathfrak {p} \oplus \mathfrak {k}$
. We have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn136.png?pub-status=live)
In particular, we can evaluate on the basis
$X_{\alpha \mu }$
and get:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn137.png?pub-status=live)
So as an element of
$\mathfrak {p}^\ast \otimes z_0$
, we can write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn138.png?pub-status=live)
with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn139.png?pub-status=live)
Proposition 4.3 Let
$\rho (R_e)$
in
$\wedge ^2\mathfrak {p}^\ast \otimes \mathfrak {so}(z_0)$
be the curvature at the identity. Then after identifying
$\mathfrak {so}(z_0)$
with
$\wedge ^2 z_0$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn140.png?pub-status=live)
where
$\eta _\alpha ^2=\eta _\alpha \wedge \eta _\alpha $
.
Proof Using the relation
$E_{ij}E_{kl}=\delta _{il}E_{kj}$
, one can show that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn141.png?pub-status=live)
for two vectors
$X_{\alpha \nu }$
and
$ X_{\beta \mu }$
in
$\mathfrak {p}$
. Hence, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn142.png?pub-status=live)
On the other hand, since
$\eta _i(X_{jr})=\delta _{ij}e_r$
, we also have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn143.png?pub-status=live)
The lemma follows since
$\rho (X_{\nu \mu })=T(e_\nu \wedge e_\mu )$
in
$\mathfrak {so}(z_0)$
, because
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn144.png?pub-status=live)
Using the fact that the exponential satisfies
$\exp (\omega +\eta )=\exp (\omega )\exp (\eta )$
on the subalgebra
$\bigoplus \Omega ^{i,i}$
—see Remark 3.1—we can write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn145.png?pub-status=live)
We define the nth Hermite polynomial by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn146.png?pub-status=live)
The first three Hermite polynomials are
$H_0(x)=1$
,
$H_1(x)=2x$
, and
$H_2(x)=4x^2-2$
.
Lemma 4.4 Let
$\eta $
be a form in
$\bigoplus \Omega ^{i,i}$
. Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn147.png?pub-status=live)
where
$H_n$
is the nth Hermite polynomial.
Proof Since
$\eta $
and
$\eta ^2$
are in
$\bigoplus \Omega ^{i,i}$
, they commute and we can use the binomial formula:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn148.png?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn149.png?pub-status=live)
The conditions on k and l imply that n is less than or equal to
$2k$
. First, suppose that n is even. Then we have that k is between
$\frac {n}{2}$
and n, so that the sum above can be written
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn150.png?pub-status=live)
where in the second step, we let m be
$k-\frac {n}{2}$
. If n is odd, then k is between
$\frac {n+1}{2}$
and n, so that the sum can be written
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn151.png?pub-status=live)
Applying the lemma to (4.28), we get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn152.png?pub-status=live)
If
$n_1+ \cdots + n_p$
is different from q, then the Berezinian of
$\eta _1^{n_1} \wedge \cdots \wedge \eta _p^{n_p}$
vanishes and we get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn153.png?pub-status=live)
Note that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn154.png?pub-status=live)
where the sums are over all
$\mu _i$
’s between
$p+1$
and
$p+q$
. If
$n_1+\cdots +n_p$
is equal to
$q,$
we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn155.png?pub-status=live)
where the sums in the last two lines go over all tuples
$\underline {\alpha }=(\alpha _1, \dots , \alpha _q)$
with
$\alpha $
between
$1$
and p, and the value
$\alpha $
appears exactly
$n_{\alpha }$
-times in
$\underline {\alpha }$
. Hence
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn156.png?pub-status=live)
After multiplying by
$\exp \left (-\pi Q(v,v) \right )$
, we get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn157.png?pub-status=live)
The form is now rapidly decreasing in v, since the Siegel majorant is positive definite. We have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn158.png?pub-status=live)
Theorem 4.5 We have
$2^{-\frac {q}{2}}\varphi (v)=\varphi _{KM}(v)$
.
Proof It is a straightforward computation to show that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn159.png?pub-status=live)
Hence, applying this, we find that the Kudla–Millson form, defined by the Howe operators in (2.22), is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20241025122952586-0837:S0008414X23000573:S0008414X23000573_eqn160.png?pub-status=live)
5 Examples and remarks
-
(1) Let us compute the Kudla–Millson as above in the simplest setting of signature
$(1,1)$ . Let
$V(\mathbb {R})$ be the quadratic space
$\mathbb {R}^2$ with the quadratic form
$Q(v,w)=x'y+xy'$ , where x and
$x'$ (resp. y and
$y'$ ) are the components of v (respectively of w). Let
$e_1=\frac {1}{\sqrt {2}}(1,1)$ and
$e_2=\frac {1}{\sqrt {2}}(1,-1)$ . The one-dimensional negative plane
$z_0$ is
$\mathbb {R} e_2$ . If r denotes the variable on
$z_0$ , then the quadratic form is
${{Q}\big|{}_{z_0} }(r)=-r^2$ . The projection map is given by
(5.1)$$ \begin{align} \operatorname{pr} \colon V(\mathbb{R}) & \longrightarrow z_0 \nonumber \\ v=(x,x') & \longmapsto \frac{x-x'}{\sqrt{2}}. \end{align} $$
The orthogonal group of
$V(\mathbb {R})$ is
(5.2)and$$ \begin{align} G(\mathbb{R})^+=\left \{ \begin{pmatrix} t & 0 \\ 0 & t^{-1} \end{pmatrix},t>0 \right \}, \end{align} $$
$\mathbb {D}^+$ can be identified with
$\mathbb {R}_{>0}$ . The associated bundle E is
$\mathbb {R}_{>0} \times \mathbb {R}$ and the connection
$\nabla $ is simply d since the bundle is trivial. Hence, the Mathai–Quillen form is
(5.3)as in the proof of Proposition 3.5. The section$$ \begin{align} U_{MQ}=\sqrt{2}e^{-2\pi r^2}dr \in \Omega^1(E), \end{align} $$
$s_v \colon \mathbb {R}_{>0} \rightarrow E$ is given by
(5.4)where x and$$ \begin{align} s_v(t)=\left (t, \frac{t^{-1}x-tx'}{\sqrt{2}} \right ), \end{align} $$
$x'$ are the components of v. We obtain
(5.5)$$ \begin{align} s_v^\ast U_{MQ}=e^{-\pi \left ( \frac{x}{t}-tx' \right )^2}\left ( \frac{x}{t}+tx' \right ) \frac{dt}{t}. \end{align} $$
Hence, after multiplication by
$2^{-\frac {1}{2}}e^{-\pi Q(v,v)}$ , we get
(5.6)$$ \begin{align} \varphi_{KM}(x,x')= 2^{-\frac{1}{2}}e^{-\pi \left [ \left (\frac{x}{t} \right )^2+(tx')^2 \right ]}\left ( \frac{x}{t}+tx' \right ) \frac{dt}{t}. \end{align} $$
-
(2) The second example illustrates the functorial properties of the Mathai–Quillen form. Suppose that we have an orthogonal splitting of
$V(\mathbb {R})$ as
$\bigoplus _i^r V_i(\mathbb {R})$ . Let
$(p_i,q_i)$ be the signature of
$V_i(\mathbb {R})$ . We have
(5.7)$$ \begin{align} \mathbb{D}_{1} \times \cdots \times \mathbb{D}_{r} \simeq \left \{ z \in \mathbb{D} \; \vert \; z = \bigoplus_{i=1}^r z \cap V_i(\mathbb{R}) \right \}. \end{align} $$
Suppose, we fix
$z_0= z_0^1 \oplus \cdots \oplus z_0^r$ in
$\mathbb {D}^+_{1} \times \cdots \times \mathbb {D}^+_{r} \subset \mathbb {D}$ , where
$z_0^i$ is a negative
$q_i$ -plane in
$V_i(\mathbb {R})$ . Let
$G_i(\mathbb {R})$ be the subgroup preserving
$V_i(\mathbb {R})$ , let
$K_i$ be the stabilizer of
$z_0^i$ , and
$\mathbb {D}_i$ be the symmetric space associated with
$V_i(\mathbb {R})$ .
Over
$\mathbb {D}^+_{1} \times \cdots \times \mathbb {D}^+_{r}$ the bundle E splits as an orthogonal sum
$E_1 \oplus \cdots \oplus E_r$ , where
$E_i$ is the bundle
$G_i(\mathbb {R})^+ \times _{K_i} z_0^i$ . Moreover, the restriction of the Mathai–Quillen form to this subbundle is
(5.8)where$$ \begin{align} {{U_{MQ}}\big|{}_{E_1 \times \cdots \times E_r} }=U_{MQ}^1 \wedge \cdots \wedge U_{MQ}^r, \end{align} $$
$U_{MQ}^i$ is the Mathai–Quillen form on
$E_i$ . The section
$s_v$ also splits as a direct sum
$\oplus s_{v_i}$ , where
$v_i$ is the projection of v onto
$v_i$ . In summary, the following diagram commutes
(5.9)and we can conclude that(5.10)where$$ \begin{align} {{\varphi_{KM}(v)}\big|{}_{\mathbb{D}_1^+ \times \cdots \times \mathbb{D}_r^+} }=\varphi_{KM}^1(v_1)\wedge \cdots \wedge \varphi_{KM}^r(v_r), \end{align} $$
$\varphi _{KM}^i$ is the Kudla–Millson form on
$\mathbb {D}_i^+$ .
-
(2) Let
$U \subset V$ be a nondegenerate r-subspace spanned by vectors
$v_1, \dots , v_r$ . Let
$(p',q')$ be the signature of U. Let
$\mathbb {D}_U$ be the subspace
(5.11)When U is positive, i.e., when
$q'=0$ , then
$\mathbb {D}_U$ is in fact
(5.12)In particular, when U is spanned by a single positive vector v, then
$\mathbb {D}_U=\mathbb {D}_v$ , where
$\mathbb {D}_v$ is as in (2.4). Kudla and Millson construct an
$rq$ -form
$\varphi _{KM}(v_1,\dots ,v_r)$ that is a Poincaré dual to
$\Gamma _U \backslash \mathbb {D}_U$ in
$\Gamma _U \backslash \mathbb {D}$ , where
$\Gamma _U$ is the stabilizer of U in
$\Gamma $ . One of its properties [Reference Kudla and Millson8][Lemma. 4.1] is that
(5.13)$$ \begin{align} \varphi_{KM}(v_1,\dots,v_r)=\varphi_{KM}(v_1) \wedge \cdots \wedge \varphi_{KM}(v_r). \end{align} $$
Let us explain how this form can also be recovered by the Mathai–Quillen formalism. Consider the bundle
$E^r=E \oplus \cdots \oplus E$ of rank
$rq$ over
$\mathbb {D}$ . One can check that all the “ingredients” of the Mathai–Quillen form
$U_{MQ}(E^r)$ are compatible with respect to the splitting as a direct sum, so that we have
(5.14)$$ \begin{align} U_{MQ}(E^r)=U_{MQ}(E) \wedge \dots \wedge U_{MQ}(E). \end{align} $$
On the other hand, the zero locus of the section
of
$E^r$ is precisely
$\mathbb {D}_U$ . Hence, the pullback
(5.15)is a Poincaré dual of$\mathbb {D}_U$ . Moreover, by (5.14), we have
(5.16)$$ \begin{align} \varphi^0(v_1,\dots,v_r)=\varphi^0(v_1) \wedge \cdots \wedge \varphi^0(v_r). \end{align} $$
Finally, after setting
(5.17)we get(5.18)$$ \begin{align} 2^{-\frac{rq}{2}}\varphi(v_1,\dots, v_r) & = 2^{-\frac{rq}{2}} e^{-\pi \sum_{i=1}^r Q(v_i,v_i) }\varphi^0(v_1) \wedge \cdots \wedge \varphi^0(v_r) \nonumber\\ & = 2^{-\frac{rq}{2}} \varphi(v_1) \wedge \cdots \wedge \varphi(v_r) \nonumber\\ & = \varphi_{KM}(v_1) \wedge \cdots \wedge \varphi_{KM}(v_r) \nonumber \\ & = \varphi_{KM}(v_1,\dots,v_r). \end{align} $$
Acknowledgment
This project is part of my thesis and I thank my advisors Nicolas Bergeron and Luis Garcia for suggesting me this topic and for their support. I thank the anonymous referee for helpful comments and suggestions.