Introduction
Let
$(V,q)$
be a nondegenerate quadratic space over
$\mathbb {Q}$
of dimension
$n\geq 3$
with associated bilinear form
$\langle \cdot ,\cdot \rangle $
, satisfying
$\langle v,v\rangle = 2 q(v)$
, and let
$G={\mathrm {SO}}_{V}$
denote the special orthogonal group of V, viewed as a reductive algebraic group over
$\mathbb {Q}$
. The real Lie group
$G(\mathbb {R})$
is determined up to isomorphism by the signature
$(r,s)$
of the real quadratic space
. In the special case of signature
$(r,2)$
, arithmetic quotients attached to G give rise to a collection of Shimura varieties endowed with a systematic supply of algebraic cycles of all possible codimensions which have been the object of extensive study in recent decades. These cycles conjecturally arise, notably, as Fourier coefficients of modular generating series with coefficients in the arithmetic Chow groups of orthogonal Shimura varieties, following a general program initiated by Kudla [Reference Kudla, Darmon and Zhang32].
The present work proposes a conjectural framework for extending some aspects of the arithmetic theory to the setting of arbitrary real signature, in which case the Archimedean symmetric spaces associated with G is not endowed with any complex structure unless
$r = 2$
or
$s = 2.$
This framework rests on the notion of rigid meromorphic cocycles for p-arithmetic subgroups of G, whose study is initiated in this paper.
Fix an odd prime p for which the p-adic quadratic space admits a self-dual
$\mathbb {Z}_p$
-lattice
$\Lambda $
(i.e., the bilinear pairing
$\langle \cdot ,\cdot \rangle $
identifies
$\Lambda $
with its
$\mathbb {Z}_p$
-dual, or, equivalently, the induced
-valued pairing on
$\Lambda /p\Lambda $
is nondegenerate). Furthermore, fix a
$\mathbb {Z}[1/p]$
-lattice
$L\subseteq V$
on which the quadratic form q is
$\mathbb {Z}[1/p]$
-valued with associated dual lattice

and discriminant module

Let
$\mathrm {Spin}_V$
denote the spin group of V (see, for example, [Reference Bruinier9, §2.3]). It is a non-split double cover of G which fits into an exact sequence

of algebraic groups. For any extension
$K/\mathbb {Q}$
, the spinor norm

is the connecting homomorphism in the long exact sequence of -cohomology

obtained by taking
$G_K$
-invariants in (1). Denote by
$G(K)^{+}$
the kernel of this spinor norm, and let

be a p-arithmetic congruence subgroup that lies in the kernel of the real spinor norm. It is a discrete subgroup of
$G(\mathbb {R})^{+}\times G(\mathbb {Q}_p)$
.
Chapter 1 attaches to V certain Archimedean and p-adic symmetric spaces, denoted
$X_\infty $
and
$X_p$
respectively. The former is a real analytic Riemannian manifold of dimension
$rs$
equipped with an action of
$G(\mathbb {R})$
, and the latter has the structure of a rigid analytic space over
$\mathbb {Q}_p$
of dimension
$n-2$
on which
$G(\mathbb {Q}_p)$
acts naturally. The p-arithmetic group
$\Gamma $
acts on both
$X_\infty $
and
$X_p$
, and acts discretely on the product
$X_\infty \times X_p$
.
Chapter 2 describes special cycles on
$X_\infty ,$
special divisors on
$X_p$
, and the attendant structures that arise from them. More precisely, the symmetric space
$X_\infty $
of dimension
$rs$
is equipped with a systematic collection of special cycles of real codimension s, which were studied by Kudla and Millson (see [Reference Kudla and Millson33]). They are indexed by vectors
$v\in V$
of positive length, with
$\Delta _{v,\infty }$
denoting the codimension s cycle attached to v. Let
$L_{\circ }$
be the intersection of L with a self dual
$\mathbb {Z}_p$
-lattice in
$V_{\mathbb {Q}_p}$
and let
${\Gamma _{\!\circ }}\subseteq \Gamma $
be an arithmetic subgroup of
$\Gamma $
preserving
$L_{\circ }$
. Given
$m\in \mathbb {Q}_{>0}$
and
$\beta \in {\mathbb D}_{L}$
that is fixed by
${\Gamma _{\!\circ }}$
, the formal sum

descends to a cycle of real codimension s on the quotient
${\Gamma _{\!\circ }}\backslash X_\infty $
. In signature
$(r,2)$
, it corresponds to a Heegner divisor on the associated orthogonal Shimura variety. Regardless of the signature, the cycles
$\Delta _{v,\infty }$
admit p-adic avatars having codimension one in
$X_p$
, which are referred to as rational quadratic divisors. Combining the real and p-adic objects leads to the study of so-called Kudla–Millson divisors – namely, special classes in the s-th cohomology of
$\Gamma $
with values in a module
$\mathrm {Div^{\dagger }_{rq}}(X_p)$
of rational quadratic divisors satisfying a suitable ‘local finiteness’ condition which is spelled out in Chapter 2. Under the assumption that
$\Gamma $
acts trivially on the discriminant module
$\mathbb {D}_{L}$
there is – in analogy with (3) – for each
$m\in \mathbb {Q}_{>0}$
and
$\beta \in {\mathbb D}_{L}$
a Kudla–Millson divisor

admitting a representative s-cocycle whose values are supported on the rational quadratic divisors attached to vectors
$v\in \beta $
satisfying
$q(v)=m$
. In particular, if V is definite,
${\mathscr {D}}_{m,\beta }$
defines a divisor on the quotient
$\Gamma \backslash X_p$
.
Chapter 3 introduces the rigid meromorphic cocycles themselves. To this end, let
$\mathcal {M}_{\mathrm {rq}}^{\times }$
be the multiplicative group of rigid meromorphic functions on
$X_p$
with divisor in
$\mathrm {Div^{\dagger }_{rq}}(X_p)$
, equipped with the natural
$\Gamma $
-module structure via left translation on the arguments:

A class in
$\mathrm {H}^s(\Gamma , \mathcal {M}_{\mathrm {rq}}^{\times })$
is called a rigid meromorphic cocycle
Footnote
1
for
$\Gamma $
if its divisor is a Kudla–Millson divisor. Theorem 3.19 and Theorem 3.25 describe the construction of a p-adic variant of Borcherds’ singular theta lift in certain special signatures - namely, in signature
$(n,0)$
for
$n\geq 4$
, in signature
$(3,1)$
, and in signature
$(4,1)$
- having rigid meromorphic cocycles whose divisors are prescribed Kudla-Millson divisors. This implies that rigid meromorphic cocycles exist in abundance, at least in the latter special signatures.
Chapter 4 defines a notion of special points on the symmetric space
$X_p$
(relative to the action of
$G(\mathbb {Q})$
). These special points are meant to generalise CM points on orthogonal Shimura varieties, but the tori that arise from their stabilisers are not compact at infinity unless
$r =0$
or
$s = 0.$
If
$r\geq s$
, one can make sense of the value of a rigid meromorphic cocycle J at a special point x, leading to an analytically defined, and a priori transcendental, numerical quantity
$J[x]\in \mathbb {C}_p$
.
The main conjecture of this work, formulated in Conjecture 4.38 of Section 4.4, asserts that the quantities
$J[x]$
are in fact algebraic. More precisely,
$J[x]$
is predicted to lie in a class field of an appropriate reflex field attached to the special point x. This conjecture suggests a program to extend the scope of the Shimura–Taniyama–Weil theory of complex multiplication to the setting of arbitrary, not necessarily CM, number fields. More precisely, it suggests that the algebraicity properties of CM points on orthogonal Shimura varieties attached to quadratic spaces of signature
$(n,2)$
can be meaningfully extended to special points on p-adic symmetric spaces for orthogonal groups attached to quadratic spaces of arbitrary real signature.
Chapter 5 concludes with a discussion of rigid meromorphic cocycles in a few of the simplest concrete settings. When V is positive-definite (i.e., when
$s=0$
), the group
$\Gamma $
acts discretely on
$X_p$
, and, when
$r=3$
or
$4$
, the quotient
$\Gamma \backslash X_p$
can be identified with the
$\mathbb {C}_p$
-points of a Shimura curve or a quaternionic Hilbert–Blumenthal surface, respectively. Rigid meromorphic cocycles, which belong to
$\mathrm {H}^0(\Gamma , \mathcal {M}_{\mathrm {rq}}^{\times })$
, define rational functions on these varieties, and special points
$\Gamma \backslash X_p$
correspond to CM points. These considerations lead to a proof of Conjecture 4.38 when
$s=0$
and
$r\leq 5$
. Upcoming work of the second and third authors will supply more details in the definite setting. For this reason, Chapter 5 focuses on the more subtle indefinite setting.
Section 5.1 discusses the first case going truly beyond the setting of complex multiplication, where V is a quadratic space of signature
$(2,1)$
and special points correspond to real quadratic elements of the Drinfeld p-adic upper half-plane. The study of this setting, which serves as the basic prototype for the general conjecture of this paper, was initiated in [Reference Darmon and Vonk18] and a great deal of evidence, both experimental and theoretical, has been amassed in its support (cf., for instance, [Reference Darmon, Pozzi and Vonk16], [Reference Gehrmann22], [Reference Guitart, Masdeu and Xarles26], [Reference Darmon and Vonk17]).
Section 5.2 ends with a discussion of the first nontrivial setting not covered by these prior works, that of a quadratic space of signature
$(3,1)$
where the associated p-arithmetic group is a congruence subgroup of the Bianchi group
, with K an imaginary quadratic field. The resulting ‘Bianchi cocycles’ play the role of meromorphic Hilbert modular forms arising as Borcherds lifts of weakly holomorphic modular functions. Their ‘values’ at special points of
$\Gamma \backslash (\mathcal {H}_p \times \mathcal {H}_p)$
belong conjecturally to abelian extensions of quadratic extensions of real quadratic fields with a single complex place and can be envisaged as the counterpart of CM values of Hilbert modular functions.
1 Symmetric spaces
1.1 Quadratic lattices
Some notations and basic facts regarding lattices in quadratic spaces are collected in this section.
Fix a principal ideal domain R with
$\mathrm {char}(R)\neq 2$
and let K be its field of fractions. Let
$(W,q_W)$
be a finite-dimensional nondegenerate quadratic space over K with associated bilinear form
$b_W(\cdot ,\cdot )$
; that is,
$b_W(w,w)=2q(w)$
for all
$w\in W$
. An R-lattice in W is a finitely generated R-submodule
$\Lambda \subseteq W$
of maximal rank. The dual of an R-lattice
$\Lambda \subseteq W$
is

The dual of an R-lattice is also an R-lattice. Moreover, the equality

holds. An R-lattice
$\Lambda $
is contained in its dual if and only if
$b_W(\Lambda ,\Lambda )\subseteq R$
. In that case, the discriminant module of
$\Lambda $
is defined to be the quotient

It is a finitely generated R-torsion module. A lattice
$\Lambda \subseteq W$
is called self-dual if
$\Lambda =\Lambda ^{\#}$
. Let
$\Lambda \subseteq W$
be a self-dual lattice and let
$\mathfrak {q}\subseteq R$
be a maximal ideal with
$\mathrm {char}(R/\mathfrak {q})\neq 2$
. Then the reduction of the quadratic form
$q_W$
defines a nondegenerate quadratic form on the quotient
$\Lambda /\mathfrak {q}\Lambda $
.
The following basic lemma describes the behaviour of self-dual lattices under orthogonal projection.
Lemma 1.1. Let
$(W,q_W)$
be a finite-dimensional nondegenerate quadratic space over K with an orthogonal direct sum decomposition:
$W=W_1 \bigoplus W_2$
with
$W_1 \perp W_2.$
Denote the orthogonal projections by
$\pi _i\colon W \rightarrow W_i$
,
$i=1,2$
. Let
$\Lambda \subseteq W$
be a self-dual lattice and put
$\Lambda _i=\Lambda \cap W_i$
,
$i=1,2$
. Then the following holds:
-
(a)
$\Lambda _i=\pi _i(\Lambda )^{\#}$ for
$i=1,2$
-
(b)
$\pi _1(\Lambda )/\Lambda _1\cong \pi _2(\Lambda )/\Lambda _2$ as R-modules
Proof. Fix
$w\in W_1$
. Since
$\Lambda $
is self-dual, it follows that
$w\in \Lambda _1$
if and only if
$b_W(w,\lambda )\subseteq R$
for all
$\lambda \in \Lambda $
. Write
$\lambda =\pi _1(\lambda )+\pi _2(\lambda )$
. Then, by definition of the orthogonal projection, we have

Thus, it follows that
$w\in \Lambda _1$
if and only if
$w\in \pi _1(\Lambda )^{\#}$
. This proves the first claim.
For the second claim, let
$w_1$
be an element of
$\pi _1(\Lambda )$
. By definition, there exists
$w_1^{\prime }\in \pi _2(\Lambda )$
such that
$w_1+w_1^{\prime }\in \Lambda $
. Moreover, the coset
$w_1^{\prime }+\Lambda _2$
is independent of the choice of
$w_1^{\prime }$
. In case
$w_1\in \Lambda _1$
, one may take
$w_1^{\prime }=0$
. Hence, the map

is well-defined. It is easy to check that this defines an isomorphism of R-modules.
1.2 Archimedean symmetric spaces
In this section, and indeed throughout most of the paper, a quadratic space V of signature
$(r,s)$
will be fixed, following the running hypotheses and notations described in the introduction.
1.2.1 Definition
The real symmetric space attached to
$V_{\mathbb {R}}$
is the manifold, denoted
$X_\infty $
, parametrizing maximal negative-definite subspaces of
$V_{\mathbb {R}}$
. Note that
$X_\infty $
is also in natural bijection with the set of maximal positive-definite subspaces of
$V_{\mathbb {R}}$
, under the map
$Z\mapsto Z^\perp $
.
The real orthogonal group
$\mathrm {O}_V(\mathbb {R})$
, the special orthogonal group
$G(\mathbb {R})={\mathrm {SO}}_V(\mathbb {R})$
, and its connected component
$G(\mathbb {R})^+$
, all act transitively on
$X_\infty $
. The stabilisers of a point
$Z\in X_\infty $
in these three groups are

where the group in the middle consists of elements of determinant
$1$
. It follows that

Since the orthogonal group of a nondegenerate quadratic form of rank n has dimension
$n(n-1)/2$
, the dimension of the real manifold
$X_\infty $
is given by

As
$\mathrm {O}_{Z^\perp }(\mathbb {R})\times \mathrm {O}_Z(\mathbb {R}) \subseteq O_V(\mathbb {R})$
is a maximal compact subgroup (see, for example, [Reference Hilgert and Neeb28][Proposition 17.2.5]), the polar decomposition [Reference Hilgert and Neeb28][Proposition 16.1.9] implies that the quotient
$X_\infty $
is contractible. Since
$X_\infty $
is an open subset of the Grassmannian of s-dimensional subspaces in
$V_{\mathbb {R}}$
, the tangent space to
$X_\infty $
at Z is canonically identified with

The penultimate isomorphism is obtained by restricting an endomorphism of
$V_{\mathbb {R}}$
to Z and composing it with the natural surjection
$V_{\mathbb {R}}\rightarrow V_{\mathbb {R}}/Z$
.
The spaces Z and
$Z^\perp $
are endowed with natural positive inner products obtained from the restrictions of q and
$-q$
, respectively. The same is true as well for their duals, and for
$\operatorname {\mathrm {Hom}}_{\mathbb {R}}(Z,Z^\perp ) = Z^\vee \otimes Z^{\perp }$
. The resulting euclidean structure on the tangent spaces
$T_Z(X_\infty )$
endows
$X_\infty $
with a Riemannian structure and a well-defined metric and volume form, which are
$\mathrm {O}_{V}(\mathbb {R})$
-invariant. By [Reference Helgason27][Chapter V, Theorem 3.1],
$X_\infty $
has non-positive sectional curvature everywhere. In particular, the Cartan–Hadamard theorem implies that, for any point
$Z\in X_\infty $
, the exponential map

is a diffeomorphism and any two points in
$X_\infty $
are connected by a unique geodesic.
Every arithmetic subgroup
${\Gamma _{\!\circ }}\subseteq G(\mathbb {Q})$
acts discretely on
$X_\infty $
. The arithmetic quotient
${\Gamma _{\!\circ }}\backslash X_\infty $
has finite volume and is compact when V is anisotropic (i.e., contains no
$v\neq 0$
with
$q(v)=0$
) (see, for example, [Reference Bruinier9, (2.25)]).
1.2.2 Functoriality of
$X_{\infty }$
More generally, given a finite-dimensional real nondegenerate quadratic space
$(W,q_W)$
, one defines
$X_\infty (W,q_W)$
to be the space of maximal negative-definite subspaces of W. Sending a maximal negative-definite subspace of W to its orthogonal complement yields an isomorphism between the two symmetric spaces
$X_\infty (W,q_W)$
and
$X_\infty (W,-q_W)$
. Let
$(W',q_W^{\prime })$
be another real nondegenerate quadratic space such that the dimension of a maximal negative subspace of
$(W,q_W)$
and
$(W',q_W^{\prime })$
agree. For any isometric embedding
$f\colon (W,q_W) \hookrightarrow (W',q_W^{\prime })$
, the induced map

is an isometric embedding of Riemannian manifolds. In particular, if f is an isomorphism of quadratic spaces, then it induces an isometry
$f\colon X_{\infty }(W,q_W) \simeq X_{\infty } (W',q_W^{\prime })$
of symmetric spaces.Footnote
2
The latter observation implies that the symmetric space
$X_{\infty }$
is isometric to
$X_{\infty }(V_0,q_0)$
for any particular quadratic space
$(V_0,q_0)$
over
$\mathbb {R}$
having signature
$(r,s)$
. It is sometimes convenient to work with a particular model quadratic space
$(V_0,q_0)$
. In the below examples for small values of s, specializing to some particular models allows us to recover classical descriptions of certain well-known symmetric spaces.
1.2.3 Examples of
$X_{\infty }$
for small s
Signature
$(r,0)$
. In case
$V_{\mathbb {R}}$
is positive or negative-definite, the group
$G(\mathbb {R})$
is compact and the space
$X_\infty $
consists of a single point.
Signature
$(r,1)$
. When
$s=1$
, the space
$X_\infty $
is isometric to the set of lines in the negative cone of
$(r+1)$
-dimensional Minkowski space; that is,
$X_\infty $
consists of all lines
$[x_1:\cdots : x_r: x_{r+1}]\in \mathbb {P}^r(\mathbb {R})$
satisfying the inequality

Thus,
$X_\infty $
can be identified with the open unit ball in r-dimensional euclidean space, endowed with its natural hyperbolic structure.
Signature
$(r,2)$
. The case
$s=2$
plays a special role in the theory. First, it represents essentially the only setting where
$X_\infty $
has the structure of a Hermitian symmetric domain, possessing a complex structure. Second, the corresponding symmetric space
$X_\infty $
admits a convenient description which motivates the definition of the p-adic symmetric space
$X_p$
taken up in Section 1.3.
A point in
$X_\infty $
corresponds to a negative-definite two-dimensional space
$W\subseteq V_{\mathbb {R}}$
. The choice of an orthogonal basis
$(v_1,v_2)$
of
$ W$
satisfying
$q(v_1) = q(v_2)$
determines an isotropic vector
in the complexification
$V_{\mathbb {C}}$
of V satisfying
$\langle \xi _W, \bar \xi _W\rangle <0$
. The complex line spanned by
$\xi _W$
depends only on the orientation induced by
$(v_1,v_2)$
on W. This suggests identifying
$X_\infty $
with an open subset of the quadric of isotropic lines in
$V_{\mathbb {C}}$
. More precisely, let Q be the smooth quadric over
$\mathbb {Q}$
attached to V by setting

Letting
$[v]$
denote the class in
$Q(K)$
of the nonzero isotropic vector
$v\in V_K$
, we define

where
$v\mapsto \overline v$
denotes the complex conjugation on
$V_{\mathbb {C}}$
. The Hermitian symmetric space
$\widetilde {X}_\infty $
has two connected components which are interchanged by
$[v]\mapsto [\overline {v}]$
. The map sending the line
$[v]\in Q(\mathbb {C})$
spanned by
$v = v_1 + i v_2$
, with
$v_1, v_2 \in V_{\mathbb {R}}$
, to the negative-definite two-dimensional subspace spanned by
$v_1$
and
$v_2$
gives a two-to-one map
$\widetilde {X}_\infty \rightarrow X_\infty $
and identifies each of the connected components with
$X_\infty $
. Because
$s = 2$
, the orthogonal complement of
$\{v_1,v_2\}$
in
$V_{\mathbb {R}}$
is positive-definite. In particular, it does not contain any isotropic vectors. It follows that

This is, in fact, an equality except if
$r=2$
. In signature
$(2,2)$
, the space
$\widetilde {X}_\infty '$
has four connected components: two of them are given by
$\widetilde {X}_\infty $
, while the other two correspond to positive-definite planes in
$V_{\mathbb {R}}$
. For further discussion of the symmetric space
$X_\infty $
in signature
$(r,2)$
and its description in terms of what is called the projective model, see [Reference Bruinier9, §2.4].
We now specialise further to discuss indefinite quadratic spaces of dimensions
$3$
and
$4$
, in which
$V_{\mathbb {R}}$
and
$X_\infty $
admit useful concrete descriptions:
Signature
$(1,2)$
. When
$V_{\mathbb {R}}$
is of signature
$(1,2)$
(resp. signature
$(2,1)$
), it can be identified with the subspace of trace zero elements
$\mathrm {M}_2^0(\mathbb {R})$
of the space of real
$2\times 2$
matrices, with quadratic form given by
$q=\det $
(resp.
$q=-\det $
). The special orthogonal group
$G(\mathbb {R})$
is identified with
${\mathrm {PGL}}_2(\mathbb {R})$
, and
$g\in G(\mathbb {R})$
acts on
$V_{\mathbb {R}}$
by conjugation:

Isotropic vectors in
$V_{\mathbb {C}}$
correspond to nonzero linear endomorphisms of
$\mathbb {C}^2$
of trace 0 and determinant 0 – that is, nonzero nilpotent endomorphisms. Such an endomorphism of
$\mathbb {C}^2$
is uniquely determined by its kernel, which equals its image, leading to an identification

of the base change of the quadric Q to
$\mathbb {R}$
with the projective line, on which
$G(\mathbb {R})={\mathrm {PGL}}_2(\mathbb {R})$
acts via Möbius transformations. Under this identification, the hermitian symmetric space
$\widetilde {X}_\infty \subseteq Q(\mathbb {C})$
is mapped to the union of the upper and lower complex half space. In coordinates, the identification can be written as follows: the nilpotent endomorphism of
$\mathbb {C}^2$
with kernel (and image) spanned by
equals

up to scaling. The line spanned by
$M_\tau $
belongs to
$\widetilde {X}_\infty $
if and only if this vector is not defined over
$\mathbb {R}$
. The map
$\tau \mapsto M_\tau $
therefore identifies the complex upper half space
$\mathcal {H}$
with a connected component of
$\widetilde {X}_\infty $
:

Signature
$(2,2)$
. When
$V_{\mathbb {R}}$
is of signature
$(2,2)$
, it can be identified with the space
$\mathrm {M}_2(\mathbb {R})$
of real
$2\times 2$
matrices endowed with the determinant quadratic form. The spin group
$\mathrm {Spin}_V(\mathbb {R})$
is identified with the group
${\mathrm {SL}}_2(\mathbb {R})\times {\mathrm {SL}}_2(\mathbb {R})$
. The map from
${\mathrm {SL}}_2(\mathbb {R})\times {\mathrm {SL}}_2(\mathbb {R})$
to
$G(\mathbb {R})^+$
sends
$(\gamma _1,\gamma _2)$
to the orthogonal transformation of
$V_{\mathbb {R}}$
given by left and right multiplication:

This induces an identification of
$({\mathrm {SL}}_2(\mathbb {R})\times {\mathrm {SL}}_2(\mathbb {R}))/\langle \pm 1\rangle $
with
$G(\mathbb {R})^+$
(cf. the discussion in [Reference Bruinier9, §2.3]). Isotropic vectors in
$V_{\mathbb {C}}$
correspond to linear endomorphisms of
$\mathbb {C}^2$
of rank one. Such an endomorphism is uniquely determined, up to scaling, by its kernel and image, leading to an identification

of the quadric
$Q_{\mathbb {R}}$
with the product of two copies of the projective line. Given
$\tau _1, \tau _2 \in \mathbb {C}$
, the endomorphism of
$\mathbb {C}^2$
with image and kernel spanned by the vectors
respectively
equals

up to scaling. Furthermore, the line spanned by
$M_{\tau _1,\tau _2}$
belongs to the space
$\widetilde {X}_\infty '$
defined in (7) if and only if
$\tau _1,\tau _2\notin \mathbb {R}$
(i.e., if both
$\tau _1$
and
$\tau _2$
belong to the union of the complex upper and lower half planes). The map
$(\tau _1,\tau _2) \mapsto [M_{\tau _1,\tau _2}]$
thus identifies
$\mathcal {H}\times \mathcal {H}$
with one connected component of
$\widetilde {X}^{\prime }_\infty $
, and therefore, one may identify

Under this identification, the action of
$G(\mathbb {R})^+$
on
$X_\infty $
corresponds to the natural action of
$({\mathrm {SL}}_2(\mathbb {R})\times {\mathrm {SL}}_2(\mathbb {R}))/\langle \pm 1\rangle $
on
$\mathcal {H}\times \mathcal {H}$
by Möbius transformations on each coordinate.
Signature
$(3,1)$
. When
$V_{\mathbb {R}}$
is of signature
$(3,1)$
, it can be identified with the space

where
$M'$
denotes the principal involution on
$M_2(\mathbb {C})$
satisfying
$MM' = \det (M)$
, and
$\overline {M}$
is the complex conjugate of M. The action

of
${\mathrm {PSL}}_2(\mathbb {C})$
on
$V_{\mathbb {R}}$
by twisted conjugation identifies this group with the connected component
$G(\mathbb {R})^{+}$
of the special orthogonal group of
$V_{\mathbb {R}}$
. A point in
$X_\infty $
, represented by a negative line in
$V_{\mathbb {R}}$
, is spanned by a unique vector of the form
with
$z\bar z+t<0$
. The map

is a bijection.
The case of signature
$(3,1)$
is noteworthy for representing the first scenario in ranks
$3$
or
$4$
where
$X_\infty $
does not admit a complex structure. It will be discussed at greater length in Section 5.2.
1.3 p-adic symmetric spaces
We introduce a p-adic analogue of the symmetric space
$X_\infty $
. To that end, let
$\mathbb {C}_p$
be the completion of a fixed algebraic closure of
$\mathbb {Q}_p$
. Moreover, write
$\mathcal {O}_{\mathbb {C}_p}$
for its ring of integers with maximal ideal
$\mathfrak {m}_{\mathbb {C}_p}$
.
1.3.1 Definition
Recall from the introduction that p is assumed to be an odd prime for which
$V_{\mathbb {Q}_p}$
admits a self-dual lattice
$\Lambda $
. When
$n\geq 3$
, the nondegenerate quadratic
$\mathbb {F}_p$
-space
$\Lambda /p\Lambda $
is isotropic (i.e., contains a nonzero isotropic vector by the Chevalley–Warning theorem). The same is thus true for
$\Lambda $
by Hensel’s Lemma. It follows that
$V_{\mathbb {Q}_p}$
is isotropic as well: in fact, the Witt indices of
$V_{\mathbb {Q}_p}$
and
$\Lambda /p\Lambda $
(the dimensions of a maximal isotropic subspace) are essentially maximal: equal to
$(n-1)/2$
if n is odd, and greater or equal to
$(n-2)/2$
if n is even. This follows from a standard induction argument reducing the statement for
$V_{\mathbb {Q}_p}$
to the same statement for the orthogonal complement of a hyperbolic plane in
$V_{\mathbb {Q}_p}$
, and likewise for
$\Lambda /p\Lambda $
.
The description of
$X_\infty $
(or rather, of its double cover
$\widetilde {X}_{\infty }$
) given in (7) in the special case of signature
$(r,2)$
motivates the following definition of the p-adic symmetric space
$X_p$
, which shall be adopted for all signatures:

The running assumption on
$V_{\mathbb {Q}_p}$
implies that the boundary
$Q(\mathbb {Q}_p)$
of
$X_p$
is nonempty.
1.3.2 Rigid analytic structure
Like the real symmetric space
$X_\infty $
of (6) in signature
$(r,2)$
, the space
$X_p$
admits a natural analytic structure: in this instance, a rigid analytic structure over
$\mathbb {Q}_p$
obtained by expressing
$X_p$
as an increasing union of affinoid subsets, as will now be described.
The choice of a self-dual
$\mathbb {Z}_p$
-lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
extends the projective space
$\mathbb {P}_{V_{\mathbb {Q}_p}}$
over
$\mathbb {Q}_p$
to a smooth proper model
$\mathbb {P}_\Lambda $
over
$\mathrm {Spec}(\mathbb {Z}_p)$
. Since
$\Lambda $
is self-dual, the quadric of isotropic lines in
$\Lambda /p\Lambda $
is nonsingular over
$\mathbb {F}_p$
. It follows that the zero locus of q in
$\mathbb {P}_\Lambda $
extends the quadric
$Q_{\mathbb {Q}_p}$
to a smooth integral model
$Q_{\Lambda }$
over
$\mathrm {Spec}(\mathbb {Z}_p)$
. By the valuative criterion for properness,

In more concrete terms, put
$\Lambda _{\mathcal {O}_{\mathbb {C}_p}}=\Lambda \otimes _{\mathbb {Z}_p} \mathcal {O}_{\mathbb {C}_p}$
and let

denote the sets of primitive vectors in
$\Lambda $
and
$\Lambda _{\mathcal {O}_{\mathbb {C}_p}}$
, respectively. Each
$\xi \in Q(\mathbb {Q}_p)$
(resp.
$Q(\mathbb {C}_p)$
) can be represented by an isotropic vector
$v_\xi $
in
$\Lambda '$
(resp. in
$\Lambda _{\mathcal {O}_{\mathbb {C}_p}}'$
) which is well-defined up to multiplication by
$\mathbb {Z}_p^{\times }$
(resp.
$\mathcal {O}_{\mathbb {C}_p}^{\times }$
) and extends
$\xi $
to a point of
$Q_{\Lambda }$
over
$\mathbb {Z}_p$
or
$\mathcal {O}_{\mathbb {C}_p}$
, respectively.
For each integer
$k\geq 0$
and every self-dual lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
, the associated basic affinoid subset
$X_{p,\Lambda }^{\leq k}\subseteq X_p$
is defined to be

where the subscript
$0$
denotes the set of isotropic vectors. The sets
$X_{p,\Lambda }^{\leq k}$
are invariant under the compact group
${\mathrm {SO}}(\Lambda )$
.
Lemma 1.2. The space
$X_p$
is the increasing union of the
$X_{p,\Lambda }^{\leq k}$
.
Proof. It suffices to prove the inclusion

where
$A^c$
denotes the complement of a subset A inside the projective space
$\mathbb {P}_V(\mathbb {C}_p).$
If
$\xi $
belongs to
$ \bigcap (X_{p,\Lambda }^{\leq k})^c$
, then for each
$k \geq 1$
, there is a vector
$w_k \in (\Lambda ')_0$
satisfying

By the compactness of
$(\Lambda ')_0$
, the sequence
$(w_k)_{k\geq 1}$
contains a convergent subsequence, whose limit is an element
$w\in (\Lambda ')_0$
satisfying

It follows that
$\xi $
belongs to
$X_p^c$
, as was to be shown.
Remark 1.3. It is easily checked that the covering
$X_{p,\Lambda }^{\leq 0}\subseteq X_{p,\Lambda }^{\leq 1} \subseteq X_{p,\Lambda }^{\leq 2}\subseteq \ldots $
fulfils the conditions of [Reference Kiel29, Definition 2.3]. Hence, the p-adic symmetric space
$X_p$
is a rigid analytic Stein space.
1.3.3 Low-dimensional examples of
$X_p$
There is a related p-adic symmetric space that features prominently in the literature (cf. [Reference Shen42, Appendix A]). Namely, let
$Q^{\mathrm {wa}}$
be the connected rigid analytic space over
$\mathbb {Q}_p$
for which

By definition, the p-adic symmetric space
$X_p$
is an admissible open subspace of
$Q^{\mathrm {wa}}$
. In small dimensions, the two spaces agree and admit familiar descriptions, as will be explained below.
Example 1.4 (Three-dimensional quadratic spaces).
The dimension of a totally isotropic subspace of a three-dimensional quadratic space is at most one, and hence,

Let us consider the special case that
$V_{\mathbb {Q}_p}$
is given by the space of
$2 \times 2$
matrices with trace zero. Just as in the discussion of real quadratic spaces of signature
$(2,1)$
, one obtains an identification of
$G(\mathbb {Q}_p)$
with
${\mathrm {PGL}}_2(\mathbb {Q}_p)$
and of the quadric
$Q_{\mathbb {Q}_p}$
with the projective line over
$\mathbb {Q}_p$
. The same reasoning that led to (8) leads to the analogous identification

of
$X_p$
with the Drinfeld p-adic upper half plane
$\mathcal {H}_p$
, on which
$G(\mathbb {Q}_p) = {\mathrm {PGL}}_2(\mathbb {Q}_p)$
acts via Möbius transformations.
Example 1.5 (Four-dimensional quadratic space of Witt index 2).
An explicit model of the four-dimensional quadratic space over
$\mathbb {Q}_p$
of Witt index
$2$
is given by

The group
${\mathrm {SL}}_2(\mathbb {Q}_p) \times {\mathrm {SL}}_2(\mathbb {Q}_p)$
acts on
$V_{\mathbb {Q}_p}$
via (9), just as in the real case, inducing an isomorphism

as well as an identification

of rigid analytic varieties. The action of
$G(\mathbb {Q}_p)^{+}$
on
$X_p$
is by componentwise Möbius transformations under these identifications.
Let
$\xi =[v_\xi ]$
be an element of
$Q(\mathbb {C}_p)$
that is not contained in
$X_p$
(i.e., for which there is an element
$[w]\in Q(\mathbb {Q}_p)$
with
$\langle w,v_\xi \rangle =0).$
The two-dimensional regular quadratic space
$V'=w^\perp /\operatorname {\mathrm {span}}(w)$
contains an isotropic vector and hence is a hyperbolic plane over
$\mathbb {Q}_p$
. Hence, the image of
$v_\xi $
in
$V^{\prime }_{\mathbb {C}_p}$
is a multiple of a
$\mathbb {Q}_p$
-rational isotropic vector. Let
$w'\in V_{\mathbb {Q}_p}$
be a preimage of this isotropic vector. By construction, w and
$w'$
span a totally isotropic subspace
$D\subseteq V_{\mathbb {Q}_p}$
and
$v_\xi $
is an element of
$D_{\mathbb {C}_p}.$
Thus, v is not an element of
$Q^{\mathrm {wa}}$
, and it follows that

The two examples above are the only cases in which
$X_p=Q^{\mathrm {wa}}$
holds. Indeed, assume we are not in one of these situations. Decompose
$V_{\mathbb {Q}_p}$
into an orthogonal direct sum

with H being an hyperbolic plane. If V if four-dimensional, U is anisotropic by assumption. The line generated by any isotropic vector
$w_U \in U_{\mathbb {C}_p}$
is clearly contained in
$Q^{\mathrm {wa}}$
but not in
$X_p$
. If
$\dim V \geq 5$
, the quadratic space U fulfils the assumptions we imposed on
$V_{\mathbb {Q}_p}$
. Thus, replacing
$V_{\mathbb {Q}_p}$
by U, one can construct the p-adic symmetric space
$X_p(U)$
as an admissible open subset of the quadric of isotropic lines in U. Take any point of
$X_p(U)$
. Then, its image in
$Q(\mathbb {C}_p)$
is contained in
$Q^{\mathrm {wa}}$
but not in
$X_p$
. This also shows that the formation of
$X_p$
is – in contrast to the formation of
$Q^{\mathrm {wa}}$
– not functorial with respect to isometric embeddings of quadratic spaces.
1.3.4 GIT characterization of
$X_p$
We now turn to a description of the points in
$X_p$
and
$Q^{\mathrm {wa}}$
in the spirit of ‘geometric invariant theory’, characterizing them in terms of their stabilisers and their orbits under the action of split subtori in
$G(\mathbb {Q}_p)$
. This in turn relates these spaces to p-adic period spaces appearing in the theory of local Shimura varieties (see Remark 1.9 below). Split subtori of orthogonal groups are closely related to totally isotropic subspaces. More precisely, consider a nondegenerate quadratic space
$(W,q_W)$
over field K with
$\mathrm {char}(K)\neq 2$
with associated bilinear form
$b_W(\cdot ,\cdot )$
, and a nontrivial morphism

of a split torus T over K to the special orthogonal group
${\mathrm {SO}}_W$
. Remember that a split torus over K is just a finite product of copies of
$\mathbb {G}_{m,K}$
. Let
$X^\ast (T)$
be the character lattice of T – that is, the group of all (K-rational) homomorphisms from T to
$\mathbb {G}_{m,K}$
. The action of T on W via
$\mu $
induces the decomposition

into eigenspaces

For
$w_1\in W_{\chi _1}$
and
$w_2\in W_{\chi _2}$
, one computes

It follows that
$W_{\chi _1}$
and
$W_{\chi _2}$
are orthogonal for all characters
$\chi _1,\chi _2\in X^\ast (T)$
with
$\chi _1\neq \chi _2^{-1}$
. In particular,
$W_{\chi }$
is isotropic if
$\chi $
is nontrivial. One easily deduces that in that case,
$W_{\chi }\oplus W_{\chi ^{-1}}$
is a sum of hyperbolic planes and that the eigenspace
$W_{0}$
of the trivial character is a nondegenerate subspace.
In the two examples of Section 1.3.3, the symmetric space
$X_p$
consists of exactly the points in
$Q(\mathbb {C}_p)$
that are not stabilized by a
$\mathbb {Q}_p$
-split torus in
$G_{\mathbb {Q}_p}$
. Indeed, a split subtorus of
${\mathrm {SL}}_2(\mathbb {Q}_p)$
has exactly two fixed points in
$\mathbb {P}^1(\mathbb {C}_p)$
, both of which already lie in
$\mathbb {P}^1(\mathbb {Q}_p)$
. Vice versa, every element in
$\mathbb {P}^1(\mathbb {Q}_p)$
is a fixed point of a split subtorus in
${\mathrm {SL}}_2(\mathbb {Q}_p)$
. The following illustrative example shows: in dimension
$> 4$
, yet more points of
$Q(\mathbb {C}_p)$
beyond those stabilized by a
$\mathbb {Q}_p$
-split torus in
$G_{\mathbb {Q}_p}$
must be removed in order to obtain
$X_p$
from
$Q(\mathbb {C}_p)$
.
Example 1.6. Assume that the dimension of
$V_{\mathbb {Q}_p}$
is of dimension
$5$
. Write
$V_{\mathbb {Q}_p}$
as the orthogonal direct sum

with H being a hyperbolic plane. Let
$0\neq w_H\in H$
be an isotropic vector,
$w_U\in U_{\mathbb {C}_p}$
an isotropic vector such that

and put . It is easy to check that
$v^\perp \cap V_{\mathbb {Q}_p}=\mathbb {Q}_p w.$
In particular, the image
$\xi =[v]$
of v in
$ Q(\mathbb {C}_p)$
does not belong to
$X_p$
. We argue next that
$\xi $
is not stabilized by any
$\mathbb {Q}_p$
-split torus.
Suppose that
$T\subseteq G_{\mathbb {Q}_p}$
is a
$\mathbb {Q}_p$
-split torus that stabilizes the line
$\xi =[v]$
with corresponding eigenspace decomposition

The dimension of the subspace
$\dim V_{\mathbb {Q}_p,0}$
must be odd, as it is the orthogonal complement of a direct sum of hyperbolic planes. In particular, it is nontrivial. However, since
$G_{\mathbb {Q}_p}$
, and hence T, acts faithfully on
$V_{\mathbb {Q}_p}$
, we must have that
$V_{\mathbb {Q}_p, \chi ^{\pm 1}} \neq \{0\}$
for some nontrivial character
$\chi \in X^{\ast }(T)$
. It follows that
$\dim v^\perp \cap V_{\mathbb {Q}_p} \geq 2$
. Indeed,
-
○ Since T stabilizes
$\xi =[v]$ , the vector v is an eigenvector for the action of
$T.$ If
$v \in V_{\mathbb {C}_p,0}$ , then it is perpendicular to
$V_{\mathbb {Q}_p, \chi } \oplus V_{\mathbb {Q}_p,\chi ^{-1}}$ , which has dimension at least 2. If
$v \in V_{\mathbb {C}_p,\chi ^{\pm 1}}$ , however, then v is perpendicular to
$V_{\mathbb {Q}_p,0} \oplus V_{\mathbb {Q}_p, \chi ^{\mp 1}}$ , which has dimension at least 2
Since
$v^{\perp } \cap V_{\mathbb {Q}_p} = \mathbb {Q}_p w$
is only one-dimensional, the above gives a contradiction. Thus, the stabilizer of
$\xi $
contains no
$\mathbb {Q}_p$
-split torus.
The point
$\xi $
in the example above is not stabilized by a
$\mathbb {Q}_p$
-split subtorus of
$G_{\mathbb {Q}_p}$
. Nevertheless, it is still unstable with respect to such a subtorus in the sense described next. Recall the definition of (semi)stability with respect to one-parameter subgroups: fix a one-parameter
$\mathbb {Q}_p$
-subgroup

(i.e., a nontrivial homomorphism of algebraic groups over
$\mathbb {Q}_p$
). A point
$\xi \in Q(\mathbb {C}_p)$
with lift
$v_{\xi }\in V_{\mathbb {C}_p}$
is called stable with respect to
$\mu $
if
-
○ the stabilizer of
$v_{\xi }$ in
$\mathbb {C}_p^{\times }$ is finite and
-
○ the
$\mathbb {C}_p^{\times }$ -orbit of
$v_{\xi }$ is Zariski-closed in
$V_{\mathbb {C}_p}$ .
It is called semistable with respect to
$\mu $
if
$0$
does not lie in the closure of
$\mathbb {C}_p^{\times }$
-orbit of
$v_\xi $
. Every stable point is clearly also semistable. As the character lattice of
$\mathbb {G}_m$
is canonically isomorphic to
$\mathbb {Z}$
, we can write the eigenspace decomposition as

It is easy to see that
$\xi $
is stable with respect to
$\mu $
if and only if
$v_\xi $
is neither contained in

In particular, if x is fixed by
$\mathbb {G}_{m,\mathbb {C}_p}$
, then x is not stable. Similarly, a point
$\xi $
is semistable if
$v_\xi $
is neither contained in

Proposition 1.7. Let
$\xi $
be an element of
$Q(\mathbb {C}_p)$
.
-
(a) The following conditions are equivalent:
-
○
$\xi \in X_p$ and
-
○
$\xi $ is stable with respect to every one-parameter
$\mathbb {Q}_p$ -subgroup of
$G_{\mathbb {Q}_p}$ .
-
-
(b) The following conditions are equivalent:
-
○
$\xi \in Q^{\mathrm {wa}}$ and
-
○
$\xi $ is semistable with respect to every one-parameter
$\mathbb {Q}_p$ -subgroup of
$G_{\mathbb {Q}_p}$ .
-
Proof. (a): Suppose there exists a one-parameter
$\mathbb {Q}_p$
-subgroup
$\mu \colon \mathbb {G}_{m,\mathbb {Q}_p}\hookrightarrow G_{\mathbb {Q}_p}$
such that
$\xi $
is not stable with respect to
$\mu $
. After possibly replacing
$\mu $
by
$-\mu $
, we may assume that
$v_{\xi }$
is an element of the nonnegative part
$V_{\mathbb {C}_p, \geq 0}$
of the eigenspace decomposition induced by
$\mu $
. Let
$n_0> 0$
be an integer with
$V_{\mathbb {C}_p,n_0}\neq 0$
, which exists as
$\mu $
is nontrivial. Then
$\langle w,v_{\xi }\rangle =0$
for all
$w\in V_{\mathbb {Q}_p,n_0}$
, and hence,
$\xi $
is not an element of
$X_p$
.
Conversely, suppose that there exists an isotropic vector
$w\in V_{\mathbb {Q}_p}$
such that
$\langle w,v_{\xi }\rangle =0.$
There exists an isotropic vector
$w'\in V_{\mathbb {Q}_p}$
with

Let U be the orthogonal complement of
$\operatorname {\mathrm {span}}\{w,w'\}$
. This yields a one-parameter
$\mathbb {Q}_p$
-subgroup
$\mu \colon \mathbb {G}_{m,\mathbb {Q}_p}\rightarrow G_{\mathbb {Q}_p}$
as follows:

In other words, there is an eigenspace decomposition of the form

Because w and
$v_{\xi }$
are perpendicular, we see that
$v_{\xi }\in V_{\mathbb {C}_p,1}\oplus V_{\mathbb {C}_p,0}$
, and therefore,
$\xi $
is not stable with respect to
$\mu $
.
The proof of (b) follows from the observation that every isotropic subspace
$D\subseteq V_{\mathbb {Q}_p}$
can be realized as
$V_{\mathbb {Q}_p,>0}$
for some one-parameter subgroup.
Remark 1.8. Let
$\mathcal {F}$
be a generalized flag variety for some split reductive group J over
$\mathbb {Q}_p$
. Voskuil and van der Put analysed when the subspace of
$\mathcal {F}$
given by those points that are stable with respect to all one-parameter
$\mathbb {Q}_p$
-subgroups of J agrees with the space of points which are semistable with respect to all one-parameter
$\mathbb {Q}_p$
-subgroups of J (cf. [Reference Voskuil and van der Put48]), which explains the phenomenon observed in Section 1.3.3.
Remark 1.9. The rigid analytic space
$Q^{\mathrm {wa}}$
naturally appears in the theory of local Shimura varieties as a p-adic period domain. To any triple
$(H,\mu ,b)$
consisting of
-
○ a connected reductive group H over
$\mathbb {Q}_p$ ,
-
○ a cocharacter
$\mu \colon \mathbb {G}_{m,\overline {\mathbb {Q}}_p}\to \mathrm {H}_{\overline {\mathbb {Q}}_p}$ and
-
○ an element
$b\in H(\breve {\mathbb {Q}}_p)$ , where
$\breve {\mathbb {Q}}_p$ denotes the completion of the maximal unramified extension of
$\mathbb {Q}_p$ ,
one can attach the weakly admissible period domain
$\mathcal {F}^{\mathrm {wa}}(H,\mu ,b)$
([Reference Chen, Fargues and Shen13, §2]). It is an open rigid analytic subvariety of the Flag variety
$\mathcal {F}(H,\mu )$
attached to the pair
$(H,\mu )$
defined via p-adic Hodge theoretic means. Moreover, they are stable under the action of the Frobenius stabilizer
$J_b$
of b in H (see [Reference Chen, Fargues and Shen13, p. 229]). In fact, Totaro has shown in [Reference Totaro46] that the weakly admissible period domain is given by the subspace of the Flag variety
$\mathcal {F}(H,\mu )$
consisting of those points that are semistable with respect to all one-parameter
$\mathbb {Q}_p$
-subgroups of
$J_b$
. In case H is a special orthogonal group associated to a quadratic space W over
$\mathbb {Q}_p$
and the cocharacter is given by acting on a two-dimensional hyperbolic subspace of W via scaling, the associated flag variety is just the quadric of isotropic lines in W. In view of these results, the local reductive group
$G_{\mathbb {Q}_p}$
should be viewed as the Frobenius stabilizer
$J_b$
for some ‘nearby’ orthogonal group.
2 Kudla–Millson divisors
This chapter introduces the key notion of Kudla–Millson divisors. In the theory of rigid meromorphic cocycles, they play the role of Heegner divisors on orthogonal Shimura varieties.
2.1 Homological algebra notation
Some notations and conventions regarding homological algebra are collected in this section.
Let R be a (not necessarily commutative) ring. A chain complex of R-modules is a chain complex of left R-modules concentrated in nonnegative degree – that is, a complex of the form

The category of such chain complexes is denoted by
$\mathrm {Ch}_{\geq 0}(R)$
. For a chain complex
and an integer
$t\geq 0$
, write
for the chain complex given by

The canonical t-truncation of
is the subcomplex given by

By construction, the homology of the canonical t-truncation is given by

A quasi-isomorphism of chain complexes is a morphism of chain complexes that induces the identity on their cohomology. A left R-module M will often be viewed as a chain complex concentrated in degree
$0$
. A resolution of M is a chain complex
together with a quasi-isomorphism
.
Given a group
$\mathcal {G}$
and a commutative ring R, write
$R[\mathcal {G}]$
for the group ring of
$\mathcal {G}$
over R. Let us remind ourselves that homomorphisms from resolutions of the trivial
$\mathcal {G}$
-module
$\mathbb {Z}$
to shifts of
$\mathcal {G}$
-modules give rise to classes in group cohomology. More precisely, given a resolution
in
$\mathrm {Ch}_{\geq 0}(\mathbb {Z}[\mathcal {G}])$
, there are canonical homomorphisms

for all
$\mathcal {G}$
-modules M and
$t\geq 0$
. First note that a homomorphism

of chain complexes is just a
$\Gamma $
-equivariant homomorphism
$f_t\colon A_t \rightarrow M$
such that the composition
$f_t\circ d_{t+1}\colon A_{t+1}\rightarrow M$
is equal to zero. In particular, if one considers the bar resolution
of
$\mathbb {Z}$
given by

with differentials

and augmentation map

one gets back the classical notion of an M-valued t-cocycle on
$\mathcal {G}$
. Thus, by definition, every element of
defines a class in
$\mathrm {H}^{t}(\mathcal {G},M)$
. For a general resolution
, there exists a homomorphism
of chain complexes such that
$\varepsilon _0=\varepsilon \circ g_0$
by [Reference Weibel49][Theorem 2.2.6]. Pulling back
along
gives a class in
$\mathrm {H}^{t}(\mathcal {G},M)$
. As
is unique up to chain homotopy by [Reference Weibel49][Theorem 2.2.6], the class is independent of the chosen chain homomorphism. Note that in the discussion above the bar resolution could have been replaced by any projective resolution of the trivial
$\mathcal {G}$
-module
$\mathbb {Z}$
.
Given a
$\mathcal {G}$
-set I, let
$\underline {I}$
denote the category whose objects are the elements of I, with morphisms given by

Definition 2.1. An R-linear representation of I is a functor from
$\underline {I}$
to the category of R-modules.
More concretely, an R-linear representation of I consists of
-
○ a collection of R-modules
$M_{i}$ indexed by
$i\in I$ ,
-
○ for each
$g\in \mathcal {G}$ and
$i\in I$ , an R-linear homomorphism
$$ \begin{align*}g \colon M_{i} \longrightarrow M_{g i}\end{align*} $$
$\mathcal {G}$ represents the identity transformation from
$M_i$ to
$M_i$ for each i, and the following diagram commutes, for all
$i\in I$ and
$g, g' \in \mathcal {G}$ :
(17)
Given an R-linear representation
$M = ( M_i)_{i\in I}$
of I, both the direct sum
$\oplus _{i\in I} M_{i}$
and the product
$\prod _{i\in I} M_{i}$
are equipped with a
$\mathcal {G}$
-module structure via

Likewise, a chain complex of R-linear representations of I is a functor from
${\underline I}$
to the category of complexes of R-modules, that is, a collection of chain complexes
of R-modules indexed by
$i\in I$
, together with homomorphisms of complexes

for all
$g \in \mathcal {G}$
and
$i\in I$
, for which the analogue of (17) commutes.
2.2 Special cycles on
$X_\infty $
2.2.1 Definitions
A vector
$v\in V_{\mathbb {R}}$
is said to be of positive length if
$q(v)> 0$
. Denote by
$V_{\mathbb {R},+}$
(resp.
$V_{+}$
) the subset of
$V_{\mathbb {R}}$
(resp. V) of vectors of positive length. In [Reference Kudla and Millson33], Kudla and Millson attach to each
$v\in V_{\mathbb {R},+}$
a real topological cycle
$\Delta _{v,\infty }\subseteq X_\infty $
, consisting of the maximal negative subspaces that are orthogonal to v or, equivalently, the set of maximal positive subspaces that contain v:

The cycle
$\Delta _{v,\infty }$
is identified with the Archimedean symmetric space attached to
$(\mathbb {R} v)^\perp $
, a quadratic space of signature
$(r-1,s)$
. Hence,
$\Delta _{v,\infty }$
is of dimension
$(r-1)s$
and defines a cycle of real codimension s in
$X_\infty $
. Alternatively, let

be the tautological rank s vector bundle over
$X_\infty $
whose fibre over Z (relative to the second coordinate projection) is the vector space Z itself. A positive vector v defines a section
$s_v\colon X_\infty \rightarrow \tau $
sending Z to
$(v_Z,Z)$
, where
$v_Z$
is the orthogonal projection of v onto Z. The cycle
$\Delta _{v,\infty }$
is the zero locus of this section.
Examples for small s
When
$s=1$
, the cycle
$\Delta _{v,\infty }$
is identified with the set of real lines in the negative cone that intersect the hyperplane
$v^\perp \subseteq V_{\mathbb {R}}$
. It is isomorphic to an open unit ball in
$(r-1)$
-dimensional euclidean space, embedded as a codimension one submanifold of
$X_\infty $
. The complement of
$\Delta _{v,\infty }$
is the union of two disjoint open subsets of
$X_\infty $
.
When V is of signature
$(2,1)$
, this gives a qualitative description of the cycle
$\Delta _{v,\infty }$
as a hyperbolic line segment joining two points of the boundary of
$X_\infty $
relative to the ‘open unit disc’ model of
$X_\infty $
. Alternatively, under the identification of
$\mathcal {H}$
with
$X_\infty $
sending
$\tau $
to
given in (8), the cycle
$\Delta _{v,\infty }$
attached to the positive vector
corresponds to the set of
$\tau =x+iy\in \mathcal {H}$
for which v belongs to the real vector space spanned by

that is, for which
$\tau $
satisfies the equation

This leads the familiar geodesic cycle in
$\mathcal {H}$
joining the real roots of the quadratic equation
$ax^2+bx+c$
of positive discriminant
$b^2-4ac$
, which plays a key role in defining the rigid meromorphic cocycles of [Reference Darmon and Vonk18].
2.2.2 Orientations
We recall some of the generalities concerning orientations on special cycles
$\Delta _{v,\infty }$
as described in [Reference Kudla and Millson33, pp. 130-131].
An orientation on a real vector space W of dimension d is a choice of a connected component
$\mathfrak {o}\subseteq \wedge ^d W -\{0\}\simeq \mathbb {R}-\{0\}$
. A basis
$(w_1,\ldots ,w_d)$
of W is then said to be positively oriented, or simply positive, if
$w_1 \wedge \cdots \wedge w_d$
belongs to
$\mathfrak {o}$
. An orientation on W induces one on its dual
$W^\ast $
, by declaring that the dual basis of any positive basis for W is positive. Given oriented real vector spaces
$W_1$
and
$W_2$
with positive bases
$\{w_{1,1},\ldots , w_{1,k}\}$
and
$\{w_{2,1},\ldots , w_{1,\ell }\}$
, the tensor product
$W_1\otimes W_2$
is oriented by the convention that the basis
$\{w_{1,i}\otimes w_{2,j}\ \vert \ 1\leq i \leq k, 1\leq j \leq \ell \}$
ordered lexicographically from right to left is positive. The space
$\operatorname {\mathrm {Hom}}_{\mathbb {R}}(W_1,W_2)\cong W_1^\ast \otimes _{\mathbb {R}} W_2$
is oriented accordingly. Note that if
$\dim _{\mathbb {R}} W_2=1$
, changing the orientation on
$W_2$
changes the orientation on
$W_1\otimes _{\mathbb {R}} W_2$
if and only if
$\dim _{\mathbb {R}} W_1$
is odd.
Fix an orientation on
$V_{\mathbb {R}}$
once and for all. Because
$X_{\infty }$
is contractible, the real vector bundle
$\tau \rightarrow X_{\infty }$
is orientable. An orientation on one of the fibres
$\tau _Z$
of
$\tau $
thus equips each maximal negative-definite subspace of
$V_{\mathbb {R}}$
with a consistent choice of orientation. This choice can be described concretely by writing any maximal negative subspace
$Z'\in X_\infty $
as
$gZ$
for some
$g\in G(\mathbb {R})^+$
and declaring a basis of
$Z'$
to be positive if it is the image under g of a positive basis of Z. Similar remarks apply to
$Z^\perp $
; alternately, a basis for a maximal positive-definite subspace
$W\subseteq V_{\mathbb {R}}$
is said to be positively oriented if completing it by a positive basis of
$W^\perp $
yields a positive basis of
$V_{\mathbb {R}}$
.
Let
$\Delta _{v,\infty }$
be the special cycle attached to
$v\in V$
, and let
$Z\in \Delta _{v,\infty }$
. The subspace
$U\subseteq V_{\mathbb {R}}$
generated by v is oriented by saying that
$\{v\}$
is a positive basis. Following the reasoning in (5), there are canonical isomorphisms

where
$\nu _Z(X_\infty )$
denotes the normal space of
$\Delta _{v,\infty }$
at Z. These isomorphisms lead to orientations on both the tangent and normal spaces of
$X_\infty $
at Z. In other words, the tangent space
$T_Z(\Delta _{v,\infty })$
is oriented by the convention that a positive basis of
$T_Z(\Delta _{v,\infty })$
completed by a positive basis of
$\nu _Z(\Delta _{v,\infty })$
is a positive basis of
$T_Z(X_\infty )$
. The cycle
$\Delta _{v,\infty }$
is equipped with the induced orientation. The spaces
$\Delta _{v,\infty }$
and
$\Delta _{-v,\infty }$
carry the same orientation if s is even, while they carry opposite orientations if s is odd. One immediately deduces that the action of
$g \in G(\mathbb {R})^{+}$
preserves the chosen orientation on the cycles
$\Delta _{v,\infty }$
; that is, there is an equality of oriented cycles

2.2.3 Geodesic properties of special cycles on
$X_{\infty }$
Let
$v \in V_{\mathbb {R}}$
be a positive vector. The reflection

is an element of the orthogonal group
$O_{V}(\mathbb {R})$
and thus acts isometrically on
$X_\infty $
.
Lemma 2.2. The locus in
$X_\infty $
fixed by
$s_v$
equals
$\Delta _{v,\infty }$
for every
$v\in V_{\mathbb {R},+}$
.
Proof. Clearly, if
$\langle z,v \rangle = 0$
, then
$s_v(z) = z$
. Thus, if
$Z\in X_\infty $
is orthogonal to v, then Z is fixed by
$s_v$
pointwise and hence setwise. Thus,
$\Delta _{v,\infty }$
is contained in the
$s_v$
-fixed locus of
$X_\infty .$
Conversely, suppose Z is a (maximal) negative definite subspace of
$V_{\mathbb {R}}$
for which
$s_v(Z) = Z$
. Suppose
$\langle z,v \rangle \neq 0$
for some
$z \in Z$
. Then both z and
$s_v(z) = z - 2 \frac {\langle z,v \rangle }{\langle v,v \rangle } \cdot v$
, lie in Z. Since
$2 \frac {\langle z,v \rangle }{\langle v,v \rangle } \neq 0$
, it follows that v is an element of Z as well. This contradicts Z being a negative-definite subspace. It follows that the
$s_v$
-fixed locus of
$X_\infty $
is contained in
$\Delta _{v,\infty }$
.
Corollary 2.3. For every
$v\in V_{\mathbb {R},+}$
, the cycle
$\Delta _{v,\infty } \subset X_{\infty }$
is a totally geodesic submanifold; that is,
-
(a) every geodesic in
$X_{\infty }$ tangent to
$\Delta _{v,\infty }$ at some point is entirely contained in
$\Delta _{v,\infty }$ , and
-
(b) every geodesic in
$X_{\infty }$ connecting two distinct points of
$\Delta _{v,\infty }$ is entirely contained in
$\Delta _{v,\infty }$ .
Proof. Since
$X_\infty $
is non-positively curved, there is a unique geodesic connecting any two distinct points. Now let
$Z,W$
be two distinct points in
$\Delta _{v,\infty }$
. By Lemma 2.2, Z and W are both fixed by
$s_v$
. Since
$s_v$
acts isometrically on
$X_\infty $
, it maps geodesics to geodesics. Thus, the unique geodesic connecting Z to W must be fixed setwise. Since Z and W are fixed by
$s_v$
, it follows that the entire geodesic connecting Z and W is fixed pointwise by
$s_v$
. Thus, the geodesic connecting Z to W is contained in the fixed locus of
$s_v$
, which equals
$\Delta _{v,\infty }$
by Lemma 2.2. This proves (b).
The tangent space at Z to
$\Delta _{v,\infty }$
, the fixed locus of
$s_v$
in
$X_\infty $
by Lemma 2.2, can be identified with the fixed locus of
$(ds_v)_Z$
acting on
$T_Z(X_\infty ).$
Thus, if a geodesic
$\gamma $
in
$X_\infty $
is tangent to
$T_Z(\Delta _{v,\infty })$
at
$Z = \gamma (0)$
, say, then
$(ds_v)_Z$
fixes
$\gamma '(0).$
Since
$s_v$
is an isometry fixing
$\gamma (0)$
and
$\gamma '(0)$
, it fixes the entire geodesic
$\gamma $
pointwise. Thus,
$\gamma $
is entirely contained in the fixed locus of
$s_v$
, which equals
$\Delta _{v,\infty }$
by Lemma 2.2. This proves (a).
2.2.4 Homology of complements of special cycles
Given a topological space X, denote by the singular chain complex of X with integer coefficients; that is,
$C_q(X)$
is the free abelian group generated by all continuous maps from the standard q-simplex to X. If A is a subspace of X, write

for the relative singular chain complex with associated relative homology groups
$\mathrm {H}_q(X,A)$
for
$q\geq 0$
.
Proposition 2.4. Let v be an element of
$V_{\mathbb {R}}$
of positive length. Then

Proof. We may assume that
$s\geq 1$
, the case
$s=0$
being trivial.
The assertion is a consequence of the long exact sequence for relative homology and the following claim:
$X_\infty \!-\! \Delta _{v,\infty }$
is homotopy equivalent to an
$(s-1)$
-sphere. Indeed, since
$\Delta _{v,\infty }$
is a totally geodesic submanifold of the non-positively curved symmetric space
$X_\infty $
, the exponential map induces a diffeomorphism

for every
$Z\in \Delta _{v,\infty }$
. But for any finite dimensional
$\mathbb {R}$
-vector space V, the complement of a proper subspace
$W\subsetneq V$
of codimension s is homotopy equivalent to
$(V/W) \!-\! \{0\}$
, which in turn is homotopy equivalent to an
$(s-1)$
-sphere.
Let Z be an element of
$\Delta _{v,\infty }$
. The chosen orientations on
$T_Z(\Delta _{v,\infty })$
and
$T_Z(X_\infty )$
induce an orientation on the quotient
. By the proof of Lemma 2.4, one can identify the relative homology
$\mathrm {H}_s(X_\infty ,X_\infty \!-\! \Delta _{v,\infty })$
with the reduced homology
$\widetilde {\mathrm {H}}_{s-1}(\mathbb {S}_W)$
of any sphere
$\mathbb {S}_W$
in W. By continuously choosing normal vectors on
$S_W$
that point outwards, this determines an orientation on
$S_W$
and hence an identification

Lemma 2.5. If
$\gamma \in G(\mathbb {R})^{+}$
fixes v, then the induced map

is the identity.
Proof. Let
$G_v$
denote the stabiliser of v in
$G(\mathbb {R})^+$
and let W be the orthogonal complement of v. The group
$G_v$
is the connected component
${\mathrm {SO}}(W)^+$
of
${\mathrm {SO}}(W)$
, consisting of the elements with trivial spinor norm. The map
$\gamma \mapsto \gamma _\ast $
induces a homomorphism from
$G_v$
to
$\mathbb {Z}^{\times } = \pm 1$
which is therefore trivial since
$G_v$
is connected. The claim follows.
2.2.5 Kudla-Millson cycles intersecting a compact region in
$X_{\infty }$
Lemma 2.6. Let
$C\subseteq {X}_\infty $
be a compact subset and
$m>0$
a fixed real number. Then the set

is compact.
Proof. Consider the set

and let
$p_1\colon \Omega \rightarrow C$
and
$p_2\colon \Omega \rightarrow V_{\mathbb {R}}$
be the two coordinate projections. The map
$p_1$
, whose fibres are spheres of radius m in r-dimensional Euclidean space, is proper. Hence,
$\Omega $
is compact, and the same follows for
$K = p_2(\Omega )$
.
2.3 Quadratic divisors on
$X_p$
Recall that a vector
$v\in V_{\mathbb {Q}_p}$
is said to be anisotropic if
$q(v)\neq 0$
.
2.3.1 Definition
The quadratic divisor on
$X_p$
attached to an anisotropic vector
$v\in V_{\mathbb {Q}_p}$
is the subset
$\Delta _{v,p}\subseteq X_p$
given by

In case
$v\in V_{+}$
, the divisor
$\Delta _{v,p}$
is called rational quadratic. By definition,

for all
$g \in G(\mathbb {Q}_p)$
and all anisotropic vectors
$v\in V_{\mathbb {Q}_p}$
.
Example 2.7 (Three-dimensional quadratic spaces).
Following the notations and identifications in Example 1.4, the cycle
$\Delta _{v,p}\subseteq \mathcal {H}_p$
associated to the vector
is given by the equation

This set is nonempty precisely when the discriminant
$q(v) = b^2-4ac$
is not a square in
$\mathbb {Q}_p$
or, equivalently, when the orthogonal complement of v in
$V_{\mathbb {Q}_p}$
is anisotropic.
Example 2.8 (Four-dimensional quadratic spaces).
With the notations of Example 1.5, the rational quadratic divisor associated to the vector is given by the equation

This subset of
$\mathcal {H}_p\times \mathcal {H}_p$
is always nonempty, and is the p-adic analogue of a Hirzebruch–Zagier divisor on a Hilbert modular surface. In particular, the rational quadratic divisor attached to the identity matrix is simply the diagonal in
$\mathcal {H}_p \times \mathcal {H}_p$
.
Remark 2.9. More generally, for
$n\geq 4$
, the rational quadratic divisor
$\Delta _{v,p}\subseteq X_p$
is always nonempty. (See Lemma 2.11 below.)
2.3.2 Quadratic divisors intersecting affinoids
Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be a self-dual
$\mathbb {Z}_p$
-lattice. Recall that
$\Lambda '$
denotes the set of primitive vectors of
$\Lambda $
in the sense of (11). Given a nonzero vector
$v\in V_{\mathbb {Q}_p}$
there exists an integer k and a primitive vector
$v_0\in \Lambda '$
such that
$v=p^k v_0$
. The order of v with respect to
$\Lambda $
is

and the isotropy level of v with respect to
$\Lambda $
is given by

Since v determines
$v_0$
up to multiplication by
$\mathbb {Z}_p^{\times }$
, both
${\mathrm {ord}}_{\Lambda }(v)$
and
$\mathrm {iso}_{\Lambda }(v)$
are well-defined.
As explained in Section 1.3.2, the
$\mathbb {Z}_p$
-lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
allows us to express
$X_p$
as an increasing union of affinoids:

The following p-adic counterpart of Lemma 2.6 relates this filtration to the isotropy level.
Lemma 2.10. Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be a self-dual
$\mathbb {Z}_p$
-lattice and
$v\in V_{\mathbb {Q}_p}$
an anisotropic vector. Then

In particular, for every
$m\in \mathbb {Q}_p^{\times }$
and every integer
$k\in \mathbb {Z}$
, the set

is contained in
$p^{-\ell }\Lambda $
for
$\ell \leq \frac {k-{\mathrm {ord}}_p(m)}{2}$
.
Proof. We may assume that
$v\in \Lambda '$
and thus that
. We note that
$q(v)= 0$
modulo
$p^{d}$
, and hence, the image of the vector v in
$\mathbb {P}({\Lambda }/p^{d}\Lambda )$
lies on the quadric
$Q_\Lambda (\mathbb {Z}/p^{d}\mathbb {Z})$
. Since this quadric is smooth, Hensel’s Lemma implies that there is an isotropic vector
$\widetilde v\in \Lambda '$
satisfying

This implies that if
$z\in \Lambda ^{\prime }_{\mathcal {O}_{\mathbb {C}_p}}$
is orthogonal to v, then

and therefore that
$[z]\notin X_{p,\Lambda }^{\leq k}$
.
The following is a partial converse to the previous lemma:
Lemma 2.11. Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be a self-dual
$\mathbb {Z}_p$
-lattice and
$v \in V_{\mathbb {Q}_p}$
with
$q(v)\neq 0$
such that the orthogonal complement of v in
$V_{\mathbb {Q}_p}$
is not a hyperbolic plane. Then

Proof. As before, we assume that
$v\in \Lambda '$
and hence,
. Let
$V_1$
be the orthogonal complement of v. Write
$\pi \colon V_{\mathbb {Q}_p} \rightarrow V_1$
for the orthogonal projection onto
$V_1$
and put
$\Lambda _1=\Lambda \cap V_1$
. The task at hand is to find an isotropic and primitive vector
$z\in \Lambda _1$
such that

for all
$w\in (\Lambda ')_0$
. Let us first give a bound on the set
$\pi ((\Lambda ')_0)$
. By Lemma 1.1, the projection
$\pi (\Lambda )$
is the dual lattice of
$\Lambda _1$
, and the quotient
$\pi (\Lambda )/\Lambda _1$
is isomorphic to the discriminant module of the
$\mathbb {Z}_p$
-lattice
$\mathbb {Z}_p v \subseteq \mathbb {Q}_p v$
. In particular, it is cyclic of order
$p^d$
. By assumption, v is not isotropic modulo
$p^{d+1}$
, and therefore,
$v\not \equiv w \bmod p^{d+1}$
for every isotropic vector
$w\in (\Lambda ')_0$
. Since
$\langle \cdot ,\cdot \rangle $
is a perfect bilinear form on
$\Lambda $
, its reduction modulo
$p^{d+1}$
of defines a perfect bilinear form on
$\Lambda /p^{d+1}\Lambda $
. It follows that for every
$w\in (\Lambda ')_0$
, there exists an element
$\bar {u}\in \Lambda /p^{d+1}\Lambda $
with

By Hensel’s Lemma, one can lift
$\bar {u}$
to an element
$u\in \Lambda _1$
. In particular, we see that

Let
$\varpi $
be a square-root of
$p^{d}$
in
$\mathcal {O}_{\mathbb {C}_p}$
and write
$\widetilde {\Lambda }_1\subseteq \pi (\Lambda )_{\mathcal {O}_{\mathbb {C}_p}}$
for the preimage of
$\varpi (\pi (\Lambda )/\Lambda _1)_{\mathcal {O}_{\mathbb {C}_p}}\subseteq (\pi (\Lambda )/\Lambda _1)_{\mathcal {O}_{\mathbb {C}_p}}$
under the reduction map. It is the unique self-dual
$\mathcal {O}_{\mathbb {C}_p}$
-lattice
$\widetilde {\Lambda }_1\subseteq (V_1)_{\mathbb {C}_p}$
such that

Moreover, each quotient of subsequent lattices in the chain above is annihilated by
$p^{\lceil d/2 \rceil }.$
By (20), we may write

with
$\widetilde {w}$
a primitive vector of
$\widetilde {\Lambda }_1$
and
$m_w\leq d$
. Since
$\Lambda $
is compact, it follows that the set
$\{\widetilde {w} \bmod {\mathfrak {m}_{\mathcal {O}_{\mathbb {C}_p}} \widetilde {\Lambda }_1}\ \vert \ w\in (\Lambda ')_0\} \subseteq \widetilde {\Lambda }_1 /\mathfrak {m}_{\mathcal {O}_{\mathbb {C}_p}} \widetilde {\Lambda }_1$
is finite. Assume for the moment that
$n\geq 4$
and let
$Q_{\widetilde {\Lambda }_1}\subseteq \mathbb {P}(\widetilde {\Lambda }_1)$
be the smooth quadric cut out by the equation
$q=0$
. The
$\overline {\mathbb {F}}_p$
-valued points of
$Q_{\widetilde {\Lambda }_1}$
are not contained in any finite union of hyperplanes. Thus, by Hensel’s Lemma, there exists a primitive vector
$\widetilde {z}\in \widetilde {\Lambda }_1$
such that

for every
$w\in (\Lambda ')_0.$
In case
$n=3$
, the statement above follows by a direct calculation using that
$\Lambda _1$
is an anisotropic lattice by assumption. There exists a rational number
$m_z\leq d/2$
such that
$z=p^{m_z}\widetilde {z}$
is primitive in
$\Lambda _{\mathcal {O}_{\mathbb {C}_p}}$
. Thus,

for all
$w\in (\Lambda ')_0$
; that is, the class
$[z]$
lies in the intersection
$\Delta _{v,p}\cap X_{p,\Lambda }^{\leq \lceil 3d/2\rceil }.$
Remark 2.12. Let
$v_1,v_2\in V_{\mathbb {Q}_p}$
with
$q(v_i)\neq 0$
and
$\Delta _{v_i,p}\neq \emptyset $
for
$i=1,2$
. Replacing
$(\Lambda ')_0$
by its union with a suitably normalized multiple of
$v_2$
, similar arguments as in the proof of Lemma 2.11 show that
$\Delta _{v_1,p}=\Delta _{v_2,p}$
if and only if there exists
$a\in \mathbb {Q}_p^{\times }$
with
$v_{1}=a v_{2}$
. Moreover, given a self-dual
$\mathbb {Z}_p$
-lattice,
$\Lambda \subseteq V_{\mathbb {Q}_p}$
. Then,
$\Delta _{v_1,p}=\Delta _{v_2,p}$
if and only
$\Delta _{v_1,p}\cap X_{p,\Lambda }^{\leq k }=\Delta _{v_2,p}\cap X_{p,\Lambda }^{\leq k}$
for some k with
$\Delta _{v_1,p}\cap X_{p,\Lambda }^{\leq k }\neq \emptyset $
.
2.3.3 Locally finite, rational quadratic divisors
Definition 2.13. A formal integer linear combination

of rational quadratic divisors on
$X_p$
is said to be locally finite if it satisfies the following equivalent conditions:
-
1. For each affinoid subset
$\mathcal {A}\subseteq X_p$ , the formal sum
-
2. For one (and thus for each) self-dual lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$ and each
$k\geq 0$ , the formal sum
$\Delta \cap X_{p,\Lambda }^{\leq k}$ is a divisor.
The equivalence of (1) and (2) follows from Lemma 1.2. Definition 2.13 generalises the locally finite divisors that are studied in [Reference Darmon and Vonk18] and [Reference Gehrmann22], which are infinite sums of real quadratic points in
$\mathcal {H}_p$
whose support has finite intersection with each affinoid subset of
$\mathcal {H}_p$
. Write
$\mathrm {Div^{\dagger }_{rq}}(X_p)$
for the module of locally finite, rational quadratic divisors. The group
$G(\mathbb {Q})$
naturally acts on
$\mathrm {Div^{\dagger }_{rq}}(X_p)$
.
Let
$L\subseteq V$
be the
$\mathbb {Z}[1/p]$
-lattice fixed in the introduction. The most common construction of locally finite divisors rests on the following lemma:
Lemma 2.14. Let C be a compact subset of
$X_\infty $
, m a positive rational number, and
$\mathcal {A}$
an affinoid subset of
$X_p$
. Then the set

is finite.
Proof. Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
a self-dual
$\mathbb {Z}_p$
-lattice. Since any affinoid subset is contained in one of the form
$X_{p,\Lambda }^{\le k}$
, we may assume that
$\mathcal {A} = X_{p,\Lambda }^{\le k}$
without loss of generality. By Lemma 2.6, X is contained in a compact subset of
$V_{\mathbb {R}}$
, and by Lemma 2.10, it is contained in
$p^{-l} \Lambda \cap L$
for some
$l\in \mathbb {Z}$
, a discrete subset of
$V_{\mathbb {R}}$
. The finiteness of X follows.
2.3.4 Compactly supported products
Let
$\mathcal {O}\subseteq V_{+}$
be a finite union of
$\Gamma $
-orbits, such as the set of all
$v\in L$
satisfying
$q(v)=m$
for a given m. Lemma 2.14 motivates the following definition:
Definition 2.15. Let
$M = (M_v)_{v\in \mathcal {O}}$
be a representation of
$\mathcal {O}$
in the sense of Definition 2.1 with values in abelian groups. The compactly supported product of the
$M_v$
is the
$\Gamma $
-submodule of
$ \prod _{v\in \mathcal {O}} M_{v}$
given by

Let
$(\mathbb {Z})_{v\in \mathcal {O}}$
be the representation of
$\mathcal {O}$
, whose objects are all equal to
$\mathbb {Z}$
and whose arrows attached to
$\gamma \in \Gamma $
are the identity maps. After multiplying
$\mathcal {O}$
with a positive rational number, we may assume that
$\mathcal {O}$
is contained in L. Lemma 2.14 implies that the natural
$\Gamma $
-equivariant homomorphism

is well-defined.
For later purposes, let us discuss an important feature of compactly supported products. The compactly supported product of a chain complex of
$\mathcal {O}$
-representations is the subcomplex

defined by taking the compactly supported product in each degree. One easily deduces that taking compactly supported products commutes with taking homology:
Lemma 2.16. For all
$q\geq 0$
,

2.4 Kudla–Millson divisors
Suppose for the moment that
$s=0$
and that
$\Gamma $
is torsion-free. Then
$\Gamma $
acts discretely on the p-adic locally symmetric space
$X_p$
and the quotient
$\Gamma \backslash X_p$
is a smooth rigid analytic variety. The group

can be identified with the space of divisors on
$\Gamma \backslash X_p$
whose pullback to
$X_p$
is a (locally finite) rational quadratic divisor.
For general
$s\ge 0$
the ‘Archimedean components’ of rational quadratic divisors define real codimension s cycles on the Archimedean locally symmetric space
$X_\infty $
. This suggests studying classes in
$\mathrm {H}^s(\Gamma , \mathrm {Div^{\dagger }_{rq}}(X_p))$
. Indeed, we will attach to every subset
$\mathcal {O}\subseteq V_+$
that is a finite union of
$\Gamma $
-orbits a class
${\mathscr {D}}_{\mathcal {O}}\in \mathrm {H}^s(\Gamma , \mathrm {Div^{\dagger }_{rq}}(X_p))$
. Informally speaking,
${\mathscr {D}}_{\mathcal {O}}$
should be the cohomology class associated via (15) to the assignment

where c is an s-chain on
$X_\infty $
and
$\Delta _{v,\infty }\cap c$
is the signed count of intersection points of c and
$\Delta _{v,\infty }$
. The main obstacle to making this rigorous arises from the possibility that, in general, c and
$\Delta _{v,\infty }$
do not intersect transversally. In what follows, this difficulty is overcome by constructing a
$G(\mathbb {Q})$
-invariant subcomplex
for which
-
○ the inclusion
is a quasi-isomorphism and
-
○ every
$c\in \mathfrak {C}_s$ intersects every Archimedean cycle
$\Delta _{v,\infty }$ nicely.
2.4.1 A subcomplex of the singular chain complex
Every
$\gamma \in G(\mathbb {Q})$
induces homomorphisms

as well as homomorphisms

for each
$v\in V$
of positive length, which fulfil relation (17). Hence,
is a representation of the
$\Gamma $
-module
$V_+$
. The product of the quotient maps

defines a
$G(\mathbb {Q})$
-equivariant homomorphism

of chain complexes.
By Proposition 2.4, the homology of the chain complex

is concentrated in degree s. Therefore, the inclusion

of the canonical s-truncation is a quasi-isomorphism. Since taking homology commutes with products, the inclusion

is a quasi-isomorphism as well.
Define to be the pullback of
along
. More concretely,

and

Proposition 2.17. The inclusion

is a quasi-isomorphism. In particular, is a resolution of the constant
$\mathbb {Z}[\Gamma ]$
-module
$\mathbb {Z}$
.
Proof. Since
$X_\infty $
is contractible, the homology of
is concentrated in degree
$0$
. Moreover, its zeroth homology group is generated by the class of any point
$Z\in X_\infty $
. Since countable unions of proper closed submanifolds have measure zero, it follows that there exists a point

and thus, the inclusion induces a surjective map on homology groups. Injectivity on homology follows from the next general lemma on chain complexes.
Lemma 2.18. Let be a quasi-isomorphism of chain complexes and
a homomorphism of chain complexes. Then the pullback

of f along g induces injective maps on homology.
Proof. Let
$(x,y)\in A_n \times _{B_n} C_n$
be a cycle; that is, assume that

Suppose that the image of
$(x,y)$
in the homology of
$C_n$
is equal to zero. That means that there exists
${z\in C_{n+1}}$
such that
$\widetilde {f}(x,y)=y=d_{n+1}(z)$
for some
$z\in C_{n+1}$
. Since
$f(x)=g(y)=g(d_{n+1}(z))=d_{n+1}(g(z))$
is a boundary and f is a quasi-isomorphism, we see that x is a boundary as well. Thus, there exists an element
$u\in A_{n+1}$
such that
$x=d_{n+1}(u)$
. It follows that
$d_{n+1}(u,z)=(x,y)$
is a boundary, which proves the claim.
2.4.2 Divisor valued cohomology classes attached to
$\Gamma $
-orbits
Let
$\mathcal {O}\subseteq V_{+}$
be a finite union of
$\Gamma $
-orbits. By definition, the image of the
$\Gamma $
-equivariant map

is contained in the compactly supported product. Thus, restricting this map to gives a homomorphism

The natural quotient map to homology induces a
$\Gamma $
-equivariant quasi-isomorphism

Moreover, the identification (18) induces a
$\Gamma $
-equivariant isomorphism

The composition

defines a morphism of
$\mathbb {Z}[\Gamma ]$
-chain complexes and, therefore, by (15) a class

Let us unravel this construction. It is convenient to introduce the following signed intersection number for
$c\in \mathfrak {C}_s$
and
$\Delta _v$
,
$v\in V_{\mathbb {R},+}$
: by definition,

The intersection number
$\Delta _{v,\infty }\cap c$
is the image of
$d_s(c)$
in the reduced homology group
$\tilde {H}_{s-1}(X_\infty \!-\! \Delta _{v,\infty })$
that we identify with
$\mathbb {Z}$
by (18). If c is smooth and does intersect
$\Delta _{v,\infty }$
transversely, this is simply the signed count of intersection points of c with
$\Delta _{v,\infty }$
, where the latter is oriented according to the definitions spelled out in Section 2.2.2. Now the class
${\mathscr {D}}_{\mathcal {O}}^{\Gamma }$
is in fact given by the s-cocycle

2.4.3 Restriction to a subgroup
Let
$\Gamma '\subseteq \Gamma $
be a finite index subgroup and

the restriction map on cohomology. The group
$\Gamma '$
fulfils all the conditions that were imposed on
$\Gamma $
. Moreover, the set
$\mathcal {O}$
is also a finite union of
$\Gamma '$
-orbits. Thus, we can define the class

The following proposition follows directly from the construction.
Proposition 2.19. Let
$\Gamma '\subseteq \Gamma $
be a finite index subgroup. Then

Because of Proposition 2.19, the class
${\mathscr {D}}_{\mathcal {O}}^{\Gamma }$
does not depend on the group
$\Gamma $
preserving
$\mathcal {O}$
in an essential way, and we adopt the abbreviation

when the underlying group
$\Gamma $
is clear from the context.
Remark 2.20. If
$\mathcal {O}$
decomposes into the union of
$\Gamma $
-orbits
$\mathcal {O}_1,\ldots ,\mathcal {O}_h$
, then

Moreover,
${\mathscr {D}}_{a\mathcal {O}}=\mathrm {sign}(a)^{s}\cdot {\mathscr {D}}_{\mathcal {O}}$
for every
$a\in \mathbb {Q}^{\times }$
. If V is three-dimensional and
$\mathcal {O}$
consists of a single
$\Gamma $
-orbit, the class
${\mathscr {D}}_{\mathcal {O}}$
is obviously zero if
$\Delta _{v,p}=\emptyset $
for one (and thus any)
$v\in \mathcal {O}$
. It is natural to wonder whether the latter relations generate all relations between the divisors attached to
$\Gamma $
-orbits in
$V_{+}$
. This is known in signature
$(3,0)$
and
$(2,1)$
. The case that V is the three-dimensional quadratic space of rational
$2\times 2$
matrices of trace
$0$
and
$\Gamma ={\mathrm {SL}}_2(\mathbb {Z}[1/p])$
is implicitly treated in [Reference Darmon and Vonk18, Proof of Thm. 1.24], where it is shown that the parabolic part of
is generated by the classes
${\mathscr {D}}_{\mathcal {O}_i}$
as the
$\mathcal {O}_i$
range over all
$\Gamma $
-orbits of vectors. A more general case is treated in [Reference Gehrmann22]. It is unclear whether such a strong result can reasonably be expected to hold in higher rank.
2.4.4 Kudla–Millson divisors attached to Schwartz functions
The class
${\mathscr {D}}_{\mathcal {O}}$
attached to a
$\Gamma $
-orbit
$\mathcal {O}\subseteq V_+$
should be viewed as an analogue of the connected cycles on orthogonal Shimura varieties introduced by Kudla in [Reference Kudla31, Section 3]. In the following, we introduce the analogue of weighted cycles in the sense of [Reference Kudla31, Section 5]. Let
$\mathbb {A}^{p}_{f}$
denote the ring of finite adeles away from p – that is, the restricted product over all completions
$\mathbb {Q}_\ell $
for all rational primes
$\ell \neq p$
. Put
and write

for the space of
$\mathbb {Z}$
-valued Schwartz functions on
$V_{\mathbb {A}^{p}_{f}}$
with its natural
$G(\mathbb {A}^{p}_{f})$
-action. The group
$G(\mathbb {Q})$
acts on
$S(V_{\mathbb {A}^{p}_{f}})$
via the diagonal embedding
$G(\mathbb {Q})\hookrightarrow G(\mathbb {A}^{p}_{f})$
. Let
$\Phi \in S(V_{\mathbb {A}^{p}_{f}})$
be a
$\Gamma $
-invariant Schwartz function and m a positive rational number. For every nonzero integer r, the set
is a finite union of
$\Gamma $
-orbits. Moreover, it is empty for all but finitely many r. Thus, the cohomology class

is well-defined. Its definition is inspired by the formula for the restriction of a weighted cycle to a connected component of an orthogonal Shimura variety (cf. [Reference Kudla31, Proposition 5.4]).
Definition 2.21. The space of Kudla–Millson divisors of level
$\Gamma $
is the subspace

generated by the classes
${\mathscr {D}}_{m,\Phi }$
with
$m\in \mathbb {Q}_{>0}$
and
$\Phi \in S(V_{\mathbb {A}^{p}_{f}})^\Gamma $
.
2.4.5 Kudla–Millson divisors attached to cosets
The following class of Kudla–Millson divisors features prominently in the construction of p-adic Borcherds products in Section 3.3. Assume for the moment that
$\Gamma $
acts trivially on the discriminant module
$\mathbb {D}_{L}$
. Let
$\widehat {L}$
be the completion of the lattice L inside
$V_{\mathbb {A}^{p}_{f}}$
. Then for any
$\beta \in \mathbb {D}_{L}$
, the coset
$\beta +\widehat {L}$
is a
$\Gamma $
-invariant compact open subset of
$V_{\mathbb {A}^{p}_{f}}$
, and thus, the characteristic function
$\mathbf {1}_{\beta +\widehat {L}}$
is a
$\Gamma $
-invariant Schwartz function.
Definition 2.22. The Kudla–Millson divisor
${\mathscr {D}}_{m,\beta }\in \mathcal {KM}(\Gamma )$
is defined as

Denote by
$\mathcal {O}_L(m,\beta )$
the set of vectors
$v\in \beta +L$
such that
$q(v)=m$
. The equality

follows directly from the definition. If
$v\in \beta $
, then clearly,
$-v\in -\beta $
. Since
$q(v)=q(-v)$
, the equality

follows from the discussion of orientations in Section 2.2.2. Moreover, the fact that
$q(pv) = p^2 q(v)$
implies that Kudla–Millson divisors are ‘p-ordinary’ in the following sense:
Proposition 2.23. For all
$\beta \in {\mathbb D}_{L}$
and all positive rational numbers m,

2.5 Explicit cocycles for small values of s
When
$s=0$
, the space
$X_\infty $
consists of a single point and the class
${\mathscr {D}}_{\mathcal {O}}$
is simply the
$\Gamma $
-invariant locally finite divisor

In particular, it is nonzero whenever
$\Delta _{v,p}\neq \emptyset $
for one (and thus for all)
$v\in \mathcal {O}$
.
2.5.1 Transversal base points
When
$s\geq 1$
, the complex
is rather big. For both computational and theoretical purposes, it is desirable to replace it by something more manageable.
Consider the bar resolution defined in (16). Every point
$Z\in X_\infty $
gives rise to a map of chain complexes of
$\Gamma $
-modules

sending
$(\gamma _0, \ldots , \gamma _q)\in B(\Gamma )_q$
to the barycentric simplex
$[\gamma _0 Z,\ldots , \gamma _q Z]$
with corners
$\gamma _0 Z, \ldots , \gamma _q Z$
. Let us briefly recall the construction of barycentric simplices: let

denote the standard q-dimensional simplex in
$\mathbb {R}^{q+1}$
. The choice of a
$(q+1)$
-tuple of points
$(Z_0, Z_1,\ldots , Z_q)$
in
$X_\infty $
determines a map

sending
$(t_0,t_1,\ldots , t_q)$
to the unique minimumFootnote
3
of the real-valued function
$g\colon X_\infty \rightarrow \mathbb {R}$
defined by

The map
$\Phi _{Z_0,\ldots ,Z_s}$
maps the vertices of
$\Delta _q$
to the points
$Z_0,\ldots , Z_q$
, and we set

The boundary of this q-dimensional simplex is given by

where
$[Z_0,\ldots , \hat {Z_i},\ldots Z_q]$
denotes the
$(q-1)$
-dimensional simplex in
$X_\infty $
obtained by removing the i-th vertex from
$[Z_0,\ldots , Z_q]$
.
Definition 2.24. A point
$Z\in X_\infty $
is a transversal base point if
factors over the subcomplex
.
Let
$Z\in X_\infty $
be a transversal base point. For
$c\in B(\Gamma )_s$
and
$v\in V_{\mathbb {R},+}$
, define the intersection product

Then a cocycle representing the class
$\mathcal {D}_{\mathcal {O}}$
is given by the map

2.5.2 Transversal base points in signature
$(r,1)$
It is natural to ask whether transversal base points exist. This question can be answered affirmatively if
$s=1$
. Indeed, in that case, the only condition required of Z is that it not be an element of the union of all
$\Delta _{v,\infty }$
,
$v\in \mathcal {O}$
, and it was already observed in the proof of Proposition 2.17 that such points exist. So in case
$s=1$
, we may chose Z as above and get the following description of the class attached to the orbit of v: the space
$X_\infty \!-\! \Delta _{v,\infty }$
decomposes into two connected components
$\mathrm {H}_v^+$
and
$\mathrm {H}_v^-$
for all
$v\in \mathcal {O}$
, where the choice of
$\mathrm {H}_v^+$
is determined by the chosen orientations. Then the intersection product is given by

2.6 Modular symbols
Assume henceforth that the signature of V is
$(r,1)$
, and that
$Q(\mathbb {Q})\neq \emptyset $
(i.e., that V has an isotropic vector). Meyer’s theorem implies that this second condition is automatic when
$r \geq 4$
. It is also satisfied for the quadratic space of rational
$2\times 2$
matrices of trace
$0$
treated at length in [Reference Darmon and Vonk18], where Q is a conic with a rational point and hence
$Q(\mathbb {Q})={\mathbb P}_1(\mathbb {Q})$
. The existence of a rational isotropic vector in V implies that the locally symmetric space
${\Gamma _{\!\circ }}\backslash X_\infty $
is noncompact for any arithmetic subgroup
${\Gamma _{\!\circ }}\subseteq G(\mathbb {Q})$
. This complication is compensated by the availability of a theory of modular symbols which provides a valuable tool for computing the (parabolic) cohomology of
${\Gamma _{\!\circ }}$
concretely, and which we now proceed to describe.
2.6.1 Abstract modular symbols
The maximal isotropic subspaces of V are rational lines on the boundary of the negative cone in
$V_{\mathbb {R}}$
. In particular, they are one-dimensional (i.e., for any two distinct elements
$\ell _1=[w_1],\ \ell _2=[w_2]\in Q(\mathbb {Q})$
the inequality

holds).
The p-arithmetic group
$\Gamma $
acts on the set
$Q(\mathbb {Q})$
via the inclusion
$\Gamma \subseteq G(\mathbb {Q})$
. Therefore, it acts naturally on the free abelian group
$\mathbb {Z}[Q(\mathbb {Q})]$
generated by
$Q(\mathbb {Q})$
and its subgroup

Definition 2.25. Let M be any
$\Gamma $
-module. The space of M-valued modular symbols is the
$\Gamma $
-module

Given a modular symbol
$m\in \mathrm {MS}(M)$
, it is customary to set
for any
$r,s\in Q(\mathbb {Q})$
. The resulting function

then satisfies the familiar two and three term relations

The short exact sequence of
$\Gamma $
-modules

is split in the category of abelian groups, and thus, applying
$\operatorname {\mathrm {Hom}}_{\mathbb {Z}}(-, M)$
leads to the short exact sequence

Let

be the boundary map of the long exact cohomology sequence obtained by taking
$\Gamma $
-invariants in (23). The stabilizer
$P_\ell \subseteq G$
of a point
$\ell \in Q(\mathbb {Q})$
is a maximal parabolic subgroup of G. Write
$\Gamma _\ell =\Gamma \cap P_\ell (\mathbb {Q})$
for the stabilizer of
$\ell $
in
$\Gamma $
.
Lemma 2.26. Let M be a
$\Gamma $
-module such that
$M^{\Gamma _\ell }=0$
for all
$\ell \in Q(\mathbb {Q})$
. The homomorphism
$\delta _M$
is injective.
Proof. Analyzing the long exact sequence induced by the short exact sequence (23), we see that it is enough to show that

By [Reference Borel8, Theorem 7.3], the group
$\Gamma $
acts on
$Q(\mathbb {Q})$
with finitely many orbits
$\Gamma \ell _1,\ldots ,\Gamma \ell _h$
. Thus, it is enough to show that

for every
$i=1,\ldots , h$
. For each
$\ell \in Q(\mathbb {Q})$
, the
$\mathbb {Z}[\Gamma ]$
-module
$\operatorname {\mathrm {Hom}}_{\mathbb {Z}}(\mathbb {Z}[\Gamma \ell ],M)$
can be identified with the coinduction of M from
$\Gamma _\ell $
to
$\Gamma $
. Frobenius reciprocity implies that

which is equal to zero by assumption.
2.6.2 Eichler transformations
We now discuss how certain Kudla–Millson divisors can be lifted to unique
$\Gamma $
-invariant modular symbols. This bridges the gap to the formulation of the theory in terms of modular symbols as formulated in [Reference Darmon and Vonk18]. It also gives a criterion for showing that Kudla–Millson divisors are nonzero. To simplify notation, we put
$\delta =\delta _M$
in case
$M=\mathrm {Div^{\dagger }_{rq}}(X_p)$
.
As a first step, we prove that the homomorphism
$\delta $
is injective. In view of Lemma 2.26, the task at hand is to produce enough elements in
$\Gamma _\ell $
. To that end, let
$w\in V$
be an isotropic vector and
$u\in V$
a vector orthogonal to w. The Eichler transformation associated to the pair
$(w,u)$
is defined by

A simple calculation shows that
$E(w,u)$
is an orthogonal transformation of V that fixes w. In particular,
$E(w,u)\in P_{\ell }(\mathbb {Q})$
, where
$\ell $
denotes the line spanned by w. Moreover, the equality

holds for all
$u_1,u_2\in \ell ^{\perp }$
. Let
$L_\circ \subseteq V$
be any
$\mathbb {Z}$
-lattice with
$q(L_\circ )\subseteq \mathbb {Z}$
and
$w\in L_\circ $
. Then
$E(w,u)\in {\mathrm {SO}}(L_\circ )$
for every
$u\in L_\circ $
that is orthogonal to w. It follows that for every
$u\in \ell ^{\perp }$
and every arithmetic subgroup
${\Gamma _{\!\circ }}\subseteq G(\mathbb {Q})$
, there exists an integer
$t\neq 0$
such that
$E(w,xu)\in {\Gamma _{\!\circ }}$
for all
$x\in t\mathbb {Z}$
.
Lemma 2.27. The map

is injective.
Proof. Let
$\ell \subseteq V$
be an isotropic line. Fix a nondegenerate
$\mathbb {Z}_p$
lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
and a vector
$v\in V_+$
such that
$\Delta _{v,p}\cap X_{p,\Lambda }^{\leq k}\neq \emptyset $
for some
$k\geq 0$
. Put
, which is an arithmetic subgroup of the parabolic
$P_\ell (\mathbb {Q})$
. Then
$\Delta _{\gamma v,p}\cap X_{p,\Lambda }^{\leq k}\neq \emptyset $
for all
$\gamma \in \Gamma _{\ell ,\Lambda }$
. We claim that that the orbit of v under
$\Gamma _{\ell ,\Lambda }$
contains infinitely many pairwise linearly independent elements. The claim together with Remark 2.12 implies that there are no
$\Gamma _\ell $
-invariant locally finite rational quadratic divisors on
$X_p$
. The assertion of the lemma then follows from Lemma 2.26.
To prove the claim, chose a generator w of
$\ell $
that is primitive in the
$\mathbb {Z}$
-lattice
$L_\circ =L\cap \Lambda $
. First assume that v and w are orthogonal. Let
$t\neq 0$
be an integer such that
$E(w,xv)\in \Gamma _{\ell ,\Lambda }$
for all
$x\in t\mathbb {Z}$
. The vectors
$E(w,xv)(v)=v+2xq(v)w$
for
$x\in t \mathbb {Z}$
are pairwise linearly independent. Now assume that v and w are not orthogonal. The
$\mathbb {Q}$
-span H of v and w is a nondegenerate subspace of V. Take any nonzero vector
$u\in H^{\perp }$
. As before, there exists an integer
$t\neq 0$
such that
$E(w,xu)\in \Gamma _{\ell ,\Lambda }$
for all
$x\in t\mathbb {Z}$
and the the vectors
$E(w,xu)(v)=v - \langle v, w\rangle xu - q(xu)\langle v, w \rangle w$
are pairwise linearly independent because
$\langle v, w\rangle \neq 0$
.
2.6.3 Modular symbols and intersections with geodesics
Given distinct isotropic lines
$\ell _{-},\ell _{+}\in Q(\mathbb {Q})$
, we let
$\Pi (\ell _{-},\ell _{+})_{\mathbb {R}}\subseteq V_{\mathbb {R}}$
denote the
$\mathbb {R}$
-plane spanned by
$\ell _{-}$
and
$\ell _{+}$
. The set

defines a one-dimensional flat subspace of
$X_\infty $
. Recall that
$X_\infty $
consists of the negative lines of
$V_{\mathbb {R}}$
. In particular, there is a natural embedding of
$X_\infty $
into
$\mathbb {P}(V_{\mathbb {R}})$
. We may compactify
$X_\infty $
by adding
$Q(\mathbb {R})$
as the boundary. Let
$\overline {X}_\infty $
be the resulting closed subspace of
$\mathbb {P}(V_{\mathbb {R}})$
. Then the closure of
$[\ell _{-},\ell _{+}]$
in
$\overline {X}_\infty $
is just the geodesic from
$\ell _{-}$
to
$\ell _{+}$
, and we equip it with the induced orientation.
Let
$v\in V$
be a vector of positive length such that its orthogonal complement in V is anisotropic. This implies that the cycles
$\Delta _{v,\infty }$
and
$[\ell _{-},\ell _{+}]$
intersect transversally, and we write
$\left ( \Delta _{v,\infty } \cap [\ell _{-},\ell _{+}] \right ) $
for its signed intersection number. By the assumption that
$v^\perp $
is anisotropic over
$\mathbb {Q}$
, the closure
$\overline {\Delta }_{v,\infty }$
of
$\Delta _{v,\infty }$
in
$\overline {X}_\infty $
does not intersect
$Q(\mathbb {Q})$
. The complement of
$\overline {\Delta }_{v,\infty }$
in
$\overline {X}_\infty $
decomposes into two connected components
$\overline {H}_v^{+}$
and
$\overline {H}_v^{-}$
. The intersection number is zero if and only if
$\ell _{-}$
and
$\ell _{+}$
are on the same connected component. Choosing generators
$w_{-}$
and
$w_{+}$
of
$\ell _{-}$
and
$\ell _{+}$
with
$\langle w_{-}, w_{+}\rangle < 0$
, this happens if and only if
$\langle v, w_{-} \rangle $
and
$\langle v, w_{+} \rangle $
have the same sign. Concretely, the intersection number can be computed by

2.6.4 Finiteness of intersections
Lemma 2.28. Fix a pair of distinct elements
$[w_{-}],[w_{+}]\in Q(\mathbb {Q})$
. We may assume that
$\langle w_{-}, w_{+}\rangle = -1.$
Fix a rational number
$d> 0$
and a
$\mathbb {Z}$
-lattice
$L_\circ $
in V. There are only finitely many vectors
$v \in L_\circ $
satisfying
-
○
$q(v)= d$ ,
-
○
$\langle v, w_{-} \rangle , \langle v, w_{+} \rangle $ are both nonzero and have opposite sign.
Proof. Let
$\Pi $
be the subspace generated by
$w_{-}$
and
$w_{+}$
. The plane
$\Pi $
is hyperbolic and the restriction of
$\langle \cdot ,\cdot \rangle $
to
$\Pi $
is nondegenerate. So there is an orthogonal decomposition
$V = \Pi \oplus \Pi ^{\perp }$
, and
$\Pi ^{\perp }$
is positive-definite. We may write
$v = v^{||} + v^{\perp }$
with respect this decomposition. The projection
$v^{||}$
equals

We calculate:

where the last equality follows because
$\langle v,w_{-} \rangle $
and
$\langle v,w_{+} \rangle $
have opposite signs. There is some integer N such that
$N w_{-}, N w_{+} \in L.$
Then

So

since both summands are positive. There are only finitely many possibilities for the vector
$v^\perp $
since
$rN^2 \cdot v^{\perp }$
is a vector in the lattice
$L_\circ \cap \Pi ^\perp $
equipped with the quadratic form Q which is positive-definite on
$\Pi ^\perp .$
Also,
$ \left | \langle v,Nw_{+} \rangle \right |$
and
$ \left | \langle v, Nw_{-} \rangle \right |$
are positive integers (they are nonzero by assumption), and their product has bounded size. Thus, there are only finitely many possibilities for both of
$\langle v,w_{-} \rangle $
and
$\langle v,w_{+}\rangle $
and hence only finitely many possibilities for
$v^{||}.$
Hence, there are only finitely many possibilities for
$v = v^{||} + v^{\perp }$
, proving the Lemma.
2.6.5 Lifting Kudla–Millson divisors to modular symbols
Proposition 2.29. Let
$\mathcal {O}$
be a
$\Gamma $
-orbit of vectors of positive length in V such that the orthogonal complement in V of one (and thus every)
$v\in \mathcal {O}$
is anisotropic. For every pair of rational isotropic lines
$\ell _{-}, \ell _{+} \in Q(\mathbb {Q})$
, the formal sum

is locally finite. Furthermore, the assignment

defines a
$\Gamma $
-invariant
$\mathrm {Div^{\dagger }_{rq}}(X_p)$
-valued modular symbol that lifts the divisor
${\mathscr {D}}_{\mathcal {O}}$
– that is,

If, in addition,
$\mathcal {O}\neq -\mathcal {O} $
and the orthogonal complement in V of one (and thus every)
$v\in \mathcal {O}$
is not a hyperbolic plane, then the divisor valued cohomology class
${\mathscr {D}}_{\mathcal {O}}$
is nonzero.
Proof. Replacing Lemma 2.6 by Lemma 2.28, local finiteness of the divisor follows as before. That
$\widetilde {{\mathscr {D}}}_{\mathcal {O}}$
defines a
$\Gamma $
-invariant modular symbol is a formal calculation reducing to equivariance of intersection numbers and of Kudla-Millson cycles. Comparing the two different intersection products (22) and (24) yields the equality
$\delta (\widetilde {{\mathscr {D}}}_{\mathcal {O}})={\mathscr {D}}_{\mathcal {O}}$
. In order to show that
${\mathscr {D}}_{\mathcal {O}}$
is nonzero, it is enough to show that
$\widetilde {{\mathscr {D}}}_{\mathcal {O}}$
is nonzero by Corollary 2.27. Remark 2.12 implies that
$\Delta _{v_1,p}\neq \Delta _{v_2,p}$
for all vectors
$v_1,v_2\in \mathcal {O}$
with
$v_1\neq v_2$
. Thus, it is enough to prove the existence of
$v\in \mathcal {O}$
and
$\ell _{-},\ell _{+}\in Q(\mathbb {Q})$
such that
$\Delta _{v,\infty }\cap [\ell _{-},\ell _{+}]\neq 0$
. But this is obvious from the description (24) of the intersection product.
Whether the orthogonal complement of v is anisotropic or not only depends on
$q(v)$
by Witt’s cancellation theorem. This leads to the following definition.
Definition 2.30. A positive rational number
$m\in \mathbb {Q}_{>0}$
is called compact with respect to
$(V,q)$
if the orthogonal complement of one (and thus all)
$v\in V$
with
$q(v)=m$
is anisotropic.
Proposition 2.29 immediately implies that Kudla–Millson divisors attached to compact
$m\in \mathbb {Q}_{>0}$
can be lifted to modular symbols:
Corollary 2.31. Assume that
$\Gamma $
acts trivially on
$\mathbb {D}_L$
. Let
$m\in \mathbb {Q}_{>0}$
be compact with respect to
$(V,q)$
and
$\beta \in \mathbb {D}_L$
. Then there exists a
$\Gamma $
-invariant modular symbol
$\widetilde {{\mathscr {D}}}_{m,\beta }\in \mathrm {MS}(\mathrm {Div^{\dagger }_{rq}}(X_p))^{\Gamma }$
such that

Remark 2.32. For rational quadratic spaces, the discussion above is only applicable in small dimensions. Indeed, by Meyer’s theorem, there are no vectors
$v\in V$
of positive length whose orthogonal complement in V is anisotropic if
$r\geq 5$
. However, in the setup of [Reference Darmon and Vonk18] where V is the quadratic space of rational
$2\times 2$
matrices of trace
$0$
, every Kudla–Millson divisor
${\mathscr {D}}_{\mathcal {O}}$
can be lifted to a modular symbol since, if the orthogonal complement of v in V contains an isotropic vector, then
$\Delta _{v,p}$
is empty anyway.
3 Rigid meromorphic cocycles and p-adic Borcherds products
This chapter introduces the notion of rigid meromorphic cocycles and constructs an analogue of Borcherds’ singular theta lift under certain restrictive assumptions on the signature of V.
3.1 Rigid meromorphic cocycles
3.1.1 Definitions
Recall from the introduction the multiplicative group
$\mathcal {M}_{\mathrm {rq}}^{\times }$
of rigid meromorphic functions on the rigid analytic
$\mathbb {Q}_p$
-variety
$X_p$
whose divisor belongs to
$\mathrm {Div^{\dagger }_{rq}}(X_p)$
. The
$G(\mathbb {Q}_p)$
-equivariant homomorphism

induces the divisor map

Definition 3.1. A rigid meromorphic cocycle of level
$\Gamma $
is a class in
$J\in \mathrm {H}^s(\Gamma , \mathcal {M}_{\mathrm {rq}}^{\times })$
whose image under the divisor map is a Kudla–Millson divisor. Let

denote the space of rigid meromorphic cocycles.
3.1.2 Lifting obstructions
By definition, there is an exact sequence of
$G(\mathbb {Q}_p)$
-modules

with
$\mathcal {A}$
being the ring of rigid analytic functions on
$X_p$
.
Proposition 3.2. The map
$\mathrm {div}$
is surjective (i.e., the sequence

is exact).
Proof. Let

be a locally finite, rational quadratic divisor. After fixing a self-dual
$\mathbb {Z}_p$
-lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
, the locally finite divisor
${\mathscr {D}}$
can be written as a sum

where the
${\mathscr {D}}_d$
are finite divisors by Lemma 2.11. Let v be a vector that contributes to
${\mathscr {D}}_{d}$
. One may suppose that v is primitive, and choose an isotropic vector
$\tilde v\in \Lambda '$
satisfying

Fix a primitive representative
$v_\xi \in \Lambda _{\mathcal {O}_{\mathbb {C}_p}}'$
in the sense of (11) for every
$\xi \in X_p$
. By definition of
$X_p$
, the expression
$\langle \tilde v, v_\xi \rangle $
is nonzero for all
$\xi \in X_p$
. The function

is independent of the choice of
$v_\xi $
and defines a rigid meromorphic function of
$\xi \in X_p$
with divisor
$\Delta _{v,p}$
. If
$k< d$
, the restriction of
$f_v$
to
$X_{p,L}^{\leq k}$
is analytic by Lemma 2.10 and, furthermore, (26) implies that

It follows that the rational functions

with divisor
${\mathscr {D}}_d$
converge uniformly to
$1$
on any affinoid. The infinite product

therefore converges to a rigid meromorphic function on
$X_p$
with divisor
${\mathscr {D}}$
. The lemma follows.
Remark 3.3. Note that the rigid meromorphic function
$f_{\mathscr {D}}$
constructed in the proof is far from unique since it depends on the choice of a system of isotropic vectors
$\{\tilde v\}_{v\in V}$
. A different choice would multiply
$f_{\mathscr {D}}$
by an element of
$\mathcal {A}^{\times }$
.
The short exact sequence of Proposition 3.2 induces a long exact sequence in cohomology. Let

be the induced boundary homomorphism. It measures the obstruction of lifting a divisor-valued cohomology class to a cohomology class with values in meromorphic functions. Thus, its image should be viewed as an analogue of the divisor class group.
The Gross–Kohnen–Zagier theorem of Borcherds (cf. [Reference Borcherds5]) states that certain generating series of Heegner divisors in divisor class groups of orthogonal Shimura varieties are modular forms. This leads naturally to the following question: Assume that
$\Gamma $
acts trivially on
$\mathbb {D}_L$
. Let
$\mathbf {e}_\beta $
denote the standard basis element of
$\mathbb {Z}[\mathbb {D}_L]$
corresponding to an element
$\beta \in \mathbb {D}_L$
. What modularity properties does the formal Fourier series

have? Note that Proposition 2.23 suggests that it is enough to consider only those
$m\in \mathbb {Q}^{>0}$
that lie in
$\mathbb {Z}_{(p)}$
– that is, those that satisfy
${\mathrm {ord}}_p(m)\geq 0$
. We will partially answer this question in Section 3.3.
3.1.3 The cycle class map
For higher-dimensional orthogonal Shimura varieties, the kernel of the cycle class map from the divisor class group to the second singular cohomology group is frequently torsion. This is used in the proof of [Reference Yuan, Zhang and Zhang50, Theorem 1.3] to deduce the modularity of the generating series of Heegner cycles in the Chow group from Kudla and Millson’s modularity theorem for topological cycles. The analogue of the cycle class map in the current setup is the homomorphism

Remark 3.4. If
$n=3$
and
$s=0$
, one can identify
$\mathrm {H}^1(\Gamma ,\mathcal {A}^{\times })$
with the Picard group of the Mumford curve
$\Gamma \backslash X_p$
, while
$\mathrm {H}^{1}(\Gamma ,\mathcal {A}^{\times }/\mathbb {Q}_p^{\times })$
is canonically isomorphic to
$\mathbb {Z}$
, and the homomorphism
$\mathrm {cyc}$
agrees with the usual degree map. See [Reference Darmon and Vonk19, Appendix A] for a discussion of this case and its parallel with rigid meromorphic cocycles in signature
$(2,1)$
.
Similarly as in the setting of orthogonal Shimura varieties, the kernel of the homomophism
$\mathrm {cyc}$
tends to be torsion frequently, as the next proposition demonstrates. The proof uses a description of the cohomology with constant coefficients of p-arithmetic groups due to Blasius, Franke and Grunewald. Essential in their theorem is the assumption that the ambient semi-simple group is almost
$\mathbb {Q}$
-simple. Remember that a semi-simple algebraic group over a field K is called almost K-simple if it contains no nontrivial smooth connected normal subgroup. The group G is almost
$\mathbb {Q}$
-simple if
$n\neq 4$
. In dimension four, this is not necessarily the case, as the example of the quadratic space of rational
$2\times 2$
matrices demonstrates.
Proposition 3.5. Let
$n\geq 4$
and
$r\geq s$
. If
$n=4$
, assume that
$V_{\mathbb {Q}_p}$
is a sum of two hyperbolic planes and that G is almost
$\mathbb {Q}$
-simple. Then the kernel of
$\mathrm {cyc}$
is finite in the following cases:
-
○ s is even
-
○
$s=1$
-
○
$s\equiv 1\bmod 4$ and r is odd
Furthermore, in the case
$s=0$
, the exponent of
$\ker (\mathrm {cyc})$
divides
$p-1$
.
Proof. It is enough to show that the group
$\mathrm {H}^{s+1}(\Gamma ,\mathbb {Q}_p^{\times })$
is finite. Since
$\Gamma $
is a p-arithmetic group, it is of type (VFL) by [Reference Serre41, no 2.4, Theorem 4]. The remark on page 101 of loc.cit. implies that for every finitely generated abelian group A, the cohomology groups
$\mathrm {H}^{i}(\Gamma ,A)$
,
$i\geq 0$
, are finitely generated and that

for every flat
$\mathbb {Z}$
-module A and every
$i\geq 0$
. Thus, it is enough to prove that
$\mathrm {H}^{s+1}(\Gamma ,\mathbb {C})=0$
. The main theorem of [Reference Blasius, Franke and Grunewald3] describes
$\mathrm {H}^{i}(\Gamma ,\mathbb {C})$
in terms of automorphic forms. To be more precise, let
$r_p$
denote the
$\mathbb {Q}_p$
-rank of
$G_{\mathbb {Q}_p}$
, which is equal to the Witt index of
$V_{\mathbb {Q}_p}$
. Then, [Reference Blasius, Franke and Grunewald3][Theorem 1] yields a decomposition

The first summand is generated by certain cuspidal automorphic representations and which contribute to the cohomology in degree
$i-r_p$
of some locally symmetric space attached to G. (To be more precise, the local component of those automorphic representations has to be Steinberg.) By a theorem of Vogan–Zuckerman (see [Reference Vogan and Zuckerman47, Theorem 8.1]), there is no cuspidal cohomology in degrees below s. The existence of a self-dual
$\mathbb {Z}_p$
-lattice in
$V_{\mathbb {Q}_p}$
forces the Witt index of
$V_{\mathbb {Q}_p}$
to be greater or equal to
$2$
if
$n\geq 5$
. Hence, we have
$r_p\geq 2$
in all cases, which implies that

The second summand in (29) can be described via
$G(\mathbb {R})$
-invariant differential i-forms on
$X_\infty $
. These are in one-to-one correspondence with classes in
$\mathrm {H}^{i}(X_\infty ^{\vee },\mathbb {C})$
, where
$X_\infty ^{\vee }$
denotes the compact dual of
$X_\infty $
. More precisely,
$X_\infty ^{\vee }$
is the Grassmannian of oriented s-planes in
$\mathbb {R}^{r+s}$
. If s is even, the rational cohomology of
$X_\infty ^{\vee }$
is concentrated in even degrees (see [Reference Borel7]). In case
$s=1$
, the oriented Grassmannian
$X_\infty ^{\vee }$
is just the r-sphere, and thus, its cohomology is concentrated in degree
$0$
and r. Finally, if r and s are both odd, then the cohomology of
$X_\infty ^{\vee }$
is generated by Pontryagin classes of the tautological bundles over
$X_\infty ^{\vee }$
, whose degrees are by definition divisible by
$4$
, and a class in degree
$r+s-1$
(cf. [Reference Takeuchi45]). Thus, we deduce that
$\mathrm {H}^{s+1}_{\mathrm {const}}(\Gamma ,\mathbb {C})=0$
in all cases.
Remark 3.6. In the definite case (that is,
$s=0$
), the vanishing of the first cohomology group of
$\Gamma $
in the above cases also follows from Margulis’ normal subgroup theorem (see [Reference Margulis35, Chapter VIII, Theorem 2.6]).
Remark 3.7. In case of signature
$(3,1)$
the group
$G_{\mathbb {R}}$
is almost
$\mathbb {R}$
-simple, and therefore, G is almost
$\mathbb {Q}$
-simple.
3.2 Vector-valued modular forms
We give a reminder on the Weil representation attached to finite quadratic modules and vector-valued modular forms. Using a
$U_{p^2}$
-operator on vector-valued modular forms, we prove a crucial upper bound on the level of certain linear combinations of components of vector-valued modular forms (see Lemma 3.12 below). Finally, we state the modularity of theta series of definite quadratic forms as well as Funke and Millson’s theorem on the modularity of intersection numbers of special cycles with modular symbols.
3.2.1 The Weil representation attached to a finite quadratic module
Following the treatment in [Reference Strömberg44], we introduce the Weil representation attached to a finite quadratic module. Let
$\mathcal {H}\subseteq \mathbb {C}$
denote the complex upper half plane and
$\mathrm {GL}_2(\mathbb {R})^+\subseteq \mathrm {GL}_2(\mathbb {R})$
the subgroup of matrices with positive determinant. The group
$\mathrm {GL}_2(\mathbb {R})^+$
acts on
$\mathcal {H}$
via Möbius transformations; that is,

For , we let

be the usual automorphy factor.
Definition 3.8. The metaplectic two-fold cover
$\widetilde {\mathrm {GL}}_2(\mathbb {R})^+$
of
$\mathrm {GL}_2(\mathbb {R})^+$
is the group of pairs
$(g,\phi )$
with
$g\in \mathrm {GL}_2(\mathbb {R})^+$
and
$\phi \colon \mathcal {H}\rightarrow \mathbb {C}$
a square root of the automorphy factor attached to g (i.e, a holomorphic function on
$\mathcal {H}$
satisfying
$\phi (\tau )^2=j(g,\tau )$
). The multiplication is given by the law

Let
${\mathrm {Mp}}_2(\mathbb {Z})$
be the preimage of
${\mathrm {SL}}_2(\mathbb {Z})$
in
$\widetilde {\mathrm {GL}}_2(\mathbb {R})^+$
. It is generated by the two elements

Note that

A finite quadratic module
$(D,q_D)$
is a finite group D together with a function

such that
-
○
$q_D(n x)= n^2 q_D(x)$ for all
$x\in D$ ,
$n\in \mathbb {Z}$ and
-
○ the function
$B_D(x,y)=q_D(x+y)-q_D(x)-q_D(y)$ is a perfect symmetric pairing on D.
The level
$N_D$
of D is the smallest positive integer such that
$N_D\cdot q_D(x)=0$
for all
$x\in D$
. It is clearly bounded by the exponent of D. Its signature
$\mathrm {sign}(D)\in \mathbb {Z}/8\mathbb {Z}$
and Witt invariant
$\sigma _w(D)\in \mu _8$
are defined by the formula

Denote by
$\mathbf {e}_x$
, with
$x\in D$
, the standard basis of
$\mathbb {C}[D]$
. The Weil representation

attached to D is determined by the action of the two generators

See [Reference Strömberg44][Section 5] for the proof that this indeed defines a representation of
${\mathrm {Mp}}_2(\mathbb {Z})$
. Note that one has

3.2.2 Explicit actions
In case
$\mathrm {sign}(D)\equiv 0 \bmod 2$
, the action of
${\mathrm {Mp}}_2(\mathbb {Z})$
on
$\rho _D$
factors through
${\mathrm {SL}}_2(\mathbb {Z})$
. Moreover, it is trivial on the principal congruence subgroup
$\Gamma (N_D)$
. If
$\mathrm {sign}(D)\equiv 1 \bmod 2$
, then
$4$
divides
$N_D$
. In particular,
$\mathrm {sign}(D)$
is even if
$2\nmid |D|$
. The explicit formula for the action of T together with [Reference Borcherds6, Theorem 5.4], implies the following well-known description of the action of
$\Gamma _0(N_D)$
on
$\rho _D$
.
Lemma 3.9. Suppose that
$2\nmid |D|$
and let x be an element of D. Then

3.2.3 Decomposition of Weil representation
The finite quadratic module D decomposes canonically as the direct sum

over all primes
$\ell $
, where
denotes the submodule of
$\ell ^\infty $
-torsion elements. Each
$D_\ell $
is also a finite quadratic module. The Weil representation of D decomposes accordingly:
Proposition 3.10 ([Reference Zemel51], Proposition 3.2).
Let
$\{\ell _{1},\ldots ,\ell _k\}$
be a finite set of distinct primes containing all prime divisors of
$|D|$
. The canonical map

is an isomorphism of
${\mathrm {Mp}}_2(\mathbb {Z})$
-representations.
In particular, one may decompose
$D=D_p\oplus D^{(p)}$
, where
$D^{(p)}=\oplus _{\ell \neq p} D_\ell $
, and get the following decomposition of finite Weil representations:

3.2.4 Finite quadratic modules from lattices
Let
$L_\circ \subseteq V$
be an even
$\mathbb {Z}$
-lattice – that is,
$q(w)\in \mathbb {Z}$
for all
$w\in L_\circ $
. (Typically, it will be assumed that
$L_\circ \otimes \mathbb {Z}[1/p]= L$
but this is not essential.) The function

makes the discriminant module
$\mathbb {D}_{L_\circ }$
into a finite quadratic module. Moreover, the equality

holds by Milgram’s formula (see [Reference Milnor and Husemoller37][Appendix 4]). We are mostly interested in lattices of the form
$L_\circ =L\cap p^t\Lambda $
with
$\Lambda \subseteq V_{\mathbb {Q}_p}$
a self-dual
$\mathbb {Z}_p$
-lattice and
$t\geq 0$
. The projection from
$\mathbb {Q}/\mathbb {Z}$
to
$\mathbb {Q}/\mathbb {Z}[1/p]$
has a unique section, which identifies
$\mathbb {Q}/\mathbb {Z}[1/p]$
with the prime to p-part of
$\mathbb {Q}/\mathbb {Z}$
. The quadratic form q induces a map

which makes
$\mathbb {D}_L$
into a finite quadratic module. Moreover, there is a canonical isomorphism

of finite quadratic modules. Similarly, the canonical homomorphism
$\mathbb {Q}/\mathbb {Z}\rightarrow \mathbb {Q}_p/\mathbb {Z}_p$
induces an isomorphism between the
$p^\infty $
-torsion submodule of
$\mathbb {Q}/\mathbb {Z}$
and
$\mathbb {Q}_p/\mathbb {Z}_p$
. Thus, the quadratic form q makes
$\mathbb {D}_{p^t\Lambda }=p^{-t}\Lambda /p^t\Lambda $
into a finite quadratic module, and there is a canonical isomorphism

of finite quadratic modules. Thus, (31) gives the decomposition

Moreover, self-duality of
$\Lambda $
implies that

and

3.2.5 Extending the Weil representation
Fix a finite quadratic module D. Following Bruinier and Stein (cf. [Reference Bruinier and Stein11]), we extend the action of
${\mathrm {Mp}}_2(\mathbb {Z})$
on
$\rho _D$
to a larger group in order to define Hecke operators on
$\rho _D$
-valued modular forms. To that end, let
$\mathbb {Z}_{(N_D)}$
be the localization of
$\mathbb {Z}$
away from
$N_D$
; that is, we invert all primes not dividing
$N_D$
. We define

and

In [Reference Bruinier and Stein11, §4,5], Bruinier and Stein construct a projective, unitary representation of
$\mathcal {Q}(N_D)$
on
$\mathbb {C}[D]$
. In case that
$\mathrm {sign}(D)$
is even, this is an honest representation. In general, there exists a central extension
$\mathcal {Q}_1(N_D)$
of
$\mathcal {Q}(N_D)$
by
$\{\pm 1\}$
and a unitary representation of
$\mathcal {Q}_1(N_D)$
on
$\mathbb {C}[D]$
that will also denote by
$\rho _D$
. The following explicit description of the action will be an important ingredient in the study of the
$U_{p^2}$
-operator.
Lemma 3.11 ([Reference Bruinier and Stein11], Lemma 3.6).
For every integer m with
$(m,N_D)=1$
, the formula

holds.
Let
$\widetilde {\mathcal {H}}(N_D)$
be the preimage of
$\mathcal {H}(N_D)$
in
$\widetilde {\mathrm {GL}}_2(\mathbb {R})^+$
. The group
$\mathcal {Q}_2(N_D)$
is defined to be the group of all tuples
$(g,\phi ,r,t)$
such that
$(g,\phi )\in \widetilde {\mathcal {H}}(N_D)$
and
$(g,r,t)\in \mathcal {Q}_1(N_D)$
with multiplication induced from the canonical embedding

It acts on
$\mathbb {C}[D]$
via its projection onto
$\mathcal {Q}_1(N_D)$
.
Let
$\sqrt {\cdot }$
denote the principal branch of the holomorphic square root. The injection

defines a group homomorphism that satisfies

We often view
${\mathrm {Mp}}_2(\mathbb {Z})$
as a subgroup of
$\mathcal {Q}_2(N_D)$
via the embedding
$L_D$
.
3.2.6 Slash operator and Fourier expansions
Let D be a finite quadratic module and
$k\in \frac {1}{2}\mathbb {Z}$
a half-integer. The metaplectic group
$\widetilde {\mathrm {GL}}_2(\mathbb {R})^+$
acts on functions
$f\colon \mathcal {H}\rightarrow \mathbb {C}$
from the right via

This induces an action of the group
$\mathcal {Q}_2(N_D)$
via the projection
$\mathcal {Q}_2(N_D) \rightarrow \widetilde {\mathcal {H}}(N_D)$
. Any function
$f\colon \mathcal {H}\rightarrow \mathbb {C}[D]$
can be uniquely written as a sum
$f=\sum _{x\in D} f_x\cdot \mathbf {e}_x$
for some functions
$f_x\colon \mathcal {H} \rightarrow \mathbb {C}$
. Given such a function and
$\gamma \in \mathcal {Q}_2(N_D)$
, we define

Let
$w\geq 1$
be an integer and suppose that f is satisfies
$f\vert _{k,T^{w}}=f$
or, in other words,
$f_x(\tau +w) \cdot e^{-2 \pi i\cdot w q(x)}=f_x(\tau )$
for all
$x\in D$
. Then the function
$\tau \mapsto f_x(\tau ) e^{-2 \pi i \cdot q(x)\tau }$
is periodic with period w. In case f is holomorphic, it thus has a Fourier expansion of the form

Moreover, in case
$w=1$
, the Fourier coefficients
$a_f(m,x)$
vanish unless
$m\in q(x)+\mathbb {Z}$
.
3.2.7 Vector-valued modular forms
(Cf. [Reference Borcherds4] for a more detailed exposition of this material.) Let
$\mathscr {G}\subseteq \mathcal {Q}_2(N_D)$
be a subgroup that is commensurable with
$L({\mathrm {Mp}}_2(\mathbb {Z}))$
. A vector-valued modular form of weight k, type
$\rho _D$
and level
$\mathscr {G}$
is a function

such that
-
(i)
$f\vert _{k,\gamma }= f$ for all
$\gamma \in \mathscr {G}$ ,
$\tau \in \mathcal {H}$ ,
-
(ii) f is holomorphic and
-
(iii) f is holomorphic at the cusps; that is, for every
$\gamma \in {\mathrm {Mp}}_2(\mathbb {Z})$ , the Fourier coefficients of
$f\vert _{k,\gamma }$ satisfy
$$\begin{align*}a_{f\vert_{k,\gamma}}(m,x)=0\quad \forall m< 0. \end{align*}$$
A modular form f is uniquely determined by its Fourier expansion. We denote by
$M_{k,D}(\mathscr {G})$
the vector space of modular forms of weight k, type
$\rho _D$
and level
$\mathscr {G}$
. Let N be any positive integer. We denote by
$\mathscr {G}_0(N)\subseteq {\mathrm {Mp}}_2(\mathbb {Z})$
the preimage of the congruence subgroup
$\Gamma _0(N)$
under the quotient map
${\mathrm {Mp}}_2(\mathbb {Z})\rightarrow {\mathrm {SL}}_2(\mathbb {Z})$
. Note that (30) implies that
$M_{k,D}(\mathscr {G}_0(N))=\{0\}$
if
$\mathrm {sign}(D)\neq 2k \bmod 2$
.
More generally, let A be an abelian group. A formal power series

is a modular form of weight k, type
$\rho _D$
and level
$\mathscr {G}$
if for every homomorphism
$\psi \colon A\to \mathbb {C}$
, the formal power series

is the Fourier expansion of a modular form of weight k, type
$\rho _D$
and level
$\mathscr {G}$
.
3.2.8 A lemma on vector valued modular forms
Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be a self-dual
$\mathbb {Z}_p$
-lattice and
$t\geq 1$
an integer. An element of
$Q_\Lambda (\mathbb {Z}/p^t\mathbb {Z})$
is simply a free
$\mathbb {Z}/p^t\mathbb {Z}$
-submodule
$\ell \subseteq \Lambda /p^t\Lambda $
such that
$q(v)\equiv 0 \bmod p^t$
for all
$v\in \ell $
. Given such an isotropic line, we define the set

As explained in Section 3.2.4, we may decompose the discriminant module
$\mathbb {D}_{L\cap p^t\Lambda }$
as a direct sum
$\mathbb {D}_{L\cap p^t\Lambda }=\mathbb {D}_{L}\oplus \mathbb {D}_{p^t\Lambda }$
and
$\mathbb {D}_{p^t\Lambda }=p^{-t}\Lambda /p^t\Lambda $
. The following lemma is a technical but crucial input in the proof of our main theorem.
Lemma 3.12. Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be a self-dual
$\mathbb {Z}_p$
-lattice,
$\ell \in Q(\mathbb {Z}/p^t\mathbb {Z})$
an isotropic line, and
$f\in M_{k,\mathbb {D}_{L\cap p^t\Lambda }}$
. There exists a modular form
$g\in M_{k,\mathbb {D}_L}(\mathscr {G}_0(p))$
such that

Note that, although the level of the discriminant module at p gets arbitrary large, the level of the modular form g stays bounded. This is a consequence of the level lowering properties of the
$U_{p^2}$
-operator. The proof of Lemma 3.12 will be given at the end of Section 3.2.10.
3.2.9 Evaluation maps
Let
$f\colon \mathcal {H}\rightarrow \mathbb {C}[D]$
be a function and
$\mu =\sum _{x_p\in D_p} a_{x_p} \cdot \mathbf {e}_{x_p}$
an element of
$\mathbb {C}[D_p]$
. The function
$\mathscr {E}_\mu (f)\colon \mathcal {H}\rightarrow \mathbb {C}[D^{(p)}]$
is defined via

Lemma 3.13. Let
$\mathscr {G}\subseteq {\mathrm {Mp}}_2(\mathbb {Z})$
be a finite index subgroup,
$f\in M_{k,D}(\mathscr {G})$
and
$\mu \in \mathbb {C}[D_p]$
. Then
$\mathscr {E}_\mu (f)\in M_{k,D^{(p)}}(\mathscr {G}_\mu )$
, where
$\mathscr {G}_\mu $
denotes the stabilizer of
$\mu $
in
$\mathscr {G}$
.
Proof. This is an immediate consequence of (31).
Given a self-dual
$\mathbb {Z}_p$
-lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
, an integer
$t\geq 1$
, and an isotropic line
$\ell \in Q_\Lambda (\mathbb {Z}/p^t\mathbb {Z})$
, we put

Corollary 3.14. For every
$f \in M_{k,\mathbb {D}_{L\cap p^{t}\Lambda }}$
, one has

Proof. The action of
${\mathrm {Mp}}_2(\mathbb {Z})$
on
$\mathbb {D}_{p^t\Lambda }$
factors through
${\mathrm {SL}}_2(\mathbb {Z})$
. Let
${x_p}\in \mathbb {D}_\Lambda $
be any class that is represented by a vector in
$\Lambda $
. Then, Lemma 3.9 together with the discussion in Section 3.2.4 implies that

Thus, the operator
$\rho _{\mathbb {D}_\Lambda }(\gamma )$
permutes the elements in
$\{\mathbf {e}_{x_p}\ \vert \ x_p\in P(\Lambda ,\ell )\}$
, which implies that it stabilizes the element
$\mu (\Lambda ,\ell )$
. The assertion follows from Lemma 3.13.
3.2.10 Hecke operators
For any subgroup
$\mathscr {G}\subseteq \mathcal {Q}_2(N_D)$
as above and any
$\alpha =(g,\phi ,r,t) \in \mathcal {Q}_2(N_D)$
, the double coset
$\mathscr {G} \alpha \mathscr {G}$
decomposes into a finite union of left cosets:

The Hecke operator

does not depend on the choice of coset representatives
$\xi _i$
. We are only interested in the case
$p\nmid N_D$
and the special element

Let A be an abelian group. Given a formal Fourier series

we define

Proposition 2.23 implies that the formal Fourier series defined in (28) satisfies

This
$U_{p^2}$
-operator satisfies the following analogue of the classical level lowering property of the
$U_p$
-operator on scalar-valued modular forms of integral weight (cf. [Reference Atkin and Lehner1, Lemmas 6,7], and [Reference Li34, Lemma 1]):
Proposition 3.15. Let D be a finite quadratic module with
$p\nmid N_D$
and
$t\geq 1$
an integer. Then

In particular, one has
$U_{p^2}(f)\in M_{k,D}(\mathscr {G}_0(p^t))$
for all
$f\in M_{k,D}(\mathscr {G}_0(p^t))$
. Moreover, if
$t\geq 3$
, then

Proof. Consider the element

Lemma 4.7 and Lemma 4.8 of [Reference Bruinier and Stein11] imply that if

with
$\gamma _i \in \mathscr {G}_0(p^t)$
is a decomposition into disjoint left cosets, then

is also a disjoint decomposition into left cosets. In particular, we may choose

The desired equality of operators then directly follows from the last computation in the proof of [Reference Bruinier and Stein11, Theorem 4.10].
Now let
$t\geq 3$
. Given integers
$N,M\geq 1$
, we let
$\mathscr {G}_0(N,M)$
be the preimage in
${\mathrm {Mp}}_2(\mathbb {Z})$
of the congruence subgroup

Lemma 4.8 of loc.cit implies that

Averaging a modular form in
$M_{k,D}(\Gamma _0(p^{t-2},p^2))$
over the elements
$\gamma _i$
gives a form of level
$\Gamma _0(p^{t-2})$
.
Remark 3.16. Proposition 3.15 combined with equation (32) shows that, in case the formal power series
$Z_L$
is a modular form, it is of level
$\mathscr {G}_0(p)$
.
We are finally able to prove Lemma 3.12: Corollary 3.14 together with Proposition 3.15 implies that the modular form
$g=(U_{p^2}^{t}\circ \mathscr {E}_{\mu (\Lambda ,\ell )})(f)$
has level
$\mathscr {G}_0(p)$
. By construction, it has the desired Fourier coefficients.
3.2.11 Modularity of theta series
Let us suppose for the moment that
$s=0$
; that is, the quadratic form q is positive-definite. Let
$L_\circ \subseteq V$
be any
$\mathbb {Z}$
-lattice such that
$q(L_\circ )\subseteq \mathbb {Z}$
. Since q is positive-definite, the set
is finite for any
$x\in \mathbb {D}_{L_\circ }$
and any
$m\in \mathbb {Q}_{\geq 0}$
. We put
. It is well-known (cf. [Reference Borcherds4, Thm. 4.1] in the special case where
$b^-=0$
) that the theta series

is an element of
$M_{n/2,\mathbb {D}_{L_\circ }}$
. The following hyperbolic analogue when
$s=1$
is due to Funke and Millson (see [Reference Funke and Millson21, Theorem 1.7])
Theorem 3.17 (Funke–Millson).
Assume that
$s=1$
and that
$Q(\mathbb {Q})\neq \emptyset $
. For every
$\mathbb {Z}$
-lattice
$L_\circ \subseteq V$
such that
$q(L_\circ )\subseteq \mathbb {Z}$
and every pair of rational isotropic lines
$\ell _{-},\ell _{+}\in Q(\mathbb {Q})$
, there exists a modular form

such that

holds for all
$(m,\beta )$
with m compact with respect to
$(V,q)$
(see Definition 2.30).
3.3 Modularity theorems
For the remainder of this chapter, we assume that
$\Gamma $
acts trivially on the discriminant module
$\mathbb {D}_L$
. We state the main theorems on the existence of lifts of linear combinations of Kudla–Millson divisors to rigid meromorphic cocycles. To this end, for any ring R, consider the R-module

We attach to an element
$\underline {c}=(c_{m,\beta })$
of
$\mathcal {C}_{\mathbb {D}_L}\hspace {-0.2em}(\mathbb {Z})$
the Kudla–Millson divisor

Moreover, we define
$\mathcal {F}_{\mathbb {D}_L}\hspace {-0.2em}(R)$
to be the module of formal q-series of the form

For every ring R, the canonical pairing

is nondegenerate. The homomorphism

that maps a modular form to the positive part of its Fourier expansion is injective. We consider
$M_{k,\mathbb {D}_L}(\mathscr {G}_0(N))$
as a submodule of
$\mathcal {F}_{\mathbb {D}_L}\hspace {-0.2em}(\mathbb {C})$
via this embedding. Since
$M_{k,\mathbb {D}_L}(\mathscr {G}_0(N))$
is finite-dimensional, it is equal to its double orthogonal complement with respect to the pairing above. By the main theorem of [Reference McGraw36],
$M_{k,\mathbb {D}_L}$
has a basis of modular forms whose Fourier expansions have integer coefficients. The same arguments show that
$M_{k,\mathbb {D}_L}(\mathscr {G}_0(N))$
has such a basis for every integer
$N\geq 1$
, which implies that

3.3.1 Main theorems in the definite case
Theorem 3.18. Assume that
$s=0$
. Let
$\underline {c}$
be an element of
$\mathcal {C}_{\mathbb {D}_L}\hspace {-0.2em}(\mathbb {Z})$
satisfying

Then there exists
$\hat {J}\in \mathrm {H}^{0}(\Gamma ,\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times })$
with

Proof. The proof of this theorem will be given in Section 3.4.
Theorem 3.19. Let
$n\geq 4$
and
$s=0$
. If
$n=4$
, assume that
$V_{\mathbb {Q}_p}$
is a sum of two hyperbolic planes and that G is almost
$\mathbb {Q}$
-simple. Let
$\underline {c}$
be an element of
$\mathcal {C}_{\mathbb {D}_L}\hspace {-0.2em}(\mathbb {Z})$
satisfying

There exists a rigid meromorphic cocycle
$J\in \mathcal {RMC}(\Gamma )$
with

Proof. Theorem 3.18 implies that
$\mathrm {cyc}([{\mathscr {D}}_{\underline {c}}])=0$
. As the kernel of the cycle class map is finite of exponent dividing
$p-1$
by Proposition 3.5, it follows that
$[(p-1)\cdot {\mathscr {D}}_{\underline {c}}]=0$
. The claim follows.
Theorem 3.20. Let
$n\geq 4$
and
$s=0$
. If
$n=4$
, assume that
$V_{\mathbb {Q}_p}$
is a sum of two hyperbolic planes and that G is almost
$\mathbb {Q}$
-simple. There exist
$a_{0,\beta }\in \mathrm {H}^{1}(\Gamma ,\mathcal {A}^{\times })\otimes \mathbb {Q}$
,
$\beta \in \mathbb {D}_{L}$
, such that the formal power series

is a modular form of weight
$n/2$
, type
$\rho _{\mathbb {D}_L}$
and level
$\mathscr {G}_0(p)$
with coefficients in
$\mathrm {H}^{1}(\Gamma ,\mathcal {A}^{\times })\otimes \mathbb {Q}$
.
Proof. Let
$\psi \colon \mathrm {H}^{1}(\Gamma ,\mathcal {A}^{\times }) \to \mathbb {C}$
be any a homomorphism of abelian groups. Theorem 3.19 implies that

It follows from (33) that
$\psi (Z_L)(\tau )$
can be completed to an element of
$M_{k,\mathbb {D}_L}(\mathscr {G}_0(N))$
for all
$\psi $
. The theorem follows.
Remark 3.21. Crucial in this argument is the finiteness of the kernel of the cycle class map. This fails in dimension three. Indeed, the kernel of the cycle class map is the Jacobian of the Mumford curve
$\Gamma \backslash X_p$
. In [Reference Beneish, Darmon, Gehrmann and Roset2], Theorem 3.20 is proven in signature
$(3,0)$
via a study of p-adic deformations of ternary theta series attached to the definite quadratic space V.
Remark 3.22. If the signature is
$(4,0)$
and
$V_{\mathbb {Q}_p}$
is a sum of hyperbolic planes, one can identify the quotient
$\Gamma \backslash X_p$
with a quaternionic Shimura surface. Under this identification, Kudla–Millson divisors and Heegner divisors match up. Thus, Theorem 3.20 gives a p-adic analytic proof of the Gross–Kohnen–Zagier theorem for these Shimura surfaces. (The second and third authors plan to flesh out the details for this argument in a forthcoming work.)
3.3.2 Main theorems when
$s=1$
There are analogous results in some hyperbolic cases where
$s=1$
. Assume that
$Q(\mathbb {Q})\neq \emptyset $
. We call an element
$\underline {c}=(c_{m,\beta })\in \mathcal {C}_{\mathbb {D}_L}\hspace {-0.2em}(R)$
compact with respect to
$(V,q)$
if
$c_{m,\beta }=0$
for all m that are not compact with respect to
$(V,q)$
. Let
$\underline {c}=(c_{m,\beta })\in \mathcal {C}_{\mathbb {D}_L}\hspace {-0.2em}(\mathbb {Z})$
be compact with respect to
$(V,q)$
. By Corollary 2.31, we may lift the Kudla–Millson divisor
${\mathscr {D}}_{\underline {c}}$
to a
$\Gamma $
-invariant
$\mathrm {Div^{\dagger }_{rq}}(X_p)$
-valued modular symbol
$\widetilde {{\mathscr {D}}}_{\underline {c}}$
.
Remark 3.23. As mentioned before, the existence of positive rational numbers that are compact with respect to
$(V,q)$
is a strong hypothesis which forces the dimension of V to be less or equal than
$5$
. In particular, Theorems 3.24 and 3.25 below are only applicable in signature
$(3,1)$
and
$(4,1)$
. Theorem 3.25 for the three-dimensional quadratic space of rational
$2\times 2$
matrices of trace
$0$
is proven in [Reference Darmon and Vonk19].
Theorem 3.24. Let
$s=1$
,
$Q(\mathbb {Q})\neq \emptyset $
and
$\underline {c}\in \mathcal {C}_{\mathbb {D}_L}\hspace {-0.2em}(\mathbb {Z})$
compact with respect to
$(V,q)$
. Under the condition that

for every
$f\in M_{n/2,\mathbb {D}_L}(\mathscr {G}_0(p))$
, there exists
$\hat {J}\in \mathrm {MS}(\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times })^\Gamma $
such that

Proof. The proof of this theorem is the content of Section 3.5.
Theorem 3.25. Let
$n\geq 4$
,
$s=1$
,
$Q(\mathbb {Q})\neq \emptyset $
and and
$\underline {c}\in \mathcal {C}_{\mathbb {D}_L}\hspace {-0.2em}(Z)$
compact with respect to
$(V,q)$
. If
$n=4$
, assume that
$V_{\mathbb {Q}_p}$
is a sum of two hyperbolic planes. Suppose that

for every
$f\in M_{n/2,\mathbb {D}_L}(\mathscr {G}_0(p))$
. There exists a nonzero integer
$c\in \mathbb {Z}$
and a rigid meromorphic cocycle
$J\in \mathcal {RMC}(\Gamma )$
such that

Proof. Since
$\delta (\hat {J})\in \mathrm {H}^{1}(\Gamma ,\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times })$
is a lift of
${\mathscr {D}}_{\underline {c}}$
, it follows that
$\mathrm {cyc}([{\mathscr {D}}_{\underline {c}}])=0$
. As in the proof of Theorem 3.19, we conclude by using Proposition 3.5.
3.4 Proof of modularity in the definite case
The aim of this section is to prove Theorem 3.18 and Theorem 3.24 by constructing explicit infinite p-adic products that are reminiscent of Borcherds products.
3.4.1 Convergence of infinite p-adic products
Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be a self-dual lattice. The p-adic Borcherds products will be given by products of functions of the form

where
$v_\xi \in V_{\mathbb {C}_p}$
is any representative of
$\xi $
. The next lemma gives a criterion for the convergence of such products. Its proof is a slight modification of the proof of Proposition 3.2.
Lemma 3.26. Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be a self-dual
$\mathbb {Z}_p$
-lattice. Suppose
$v_1,v_2 \in \Lambda ' $
satisfy
-
○
$v_1 \equiv \alpha \cdot v_2 \bmod {p^k \Lambda }$ for some
$\alpha \in \mathbb {\mathbb {Z}}_p^{\times }$ and
-
○
$q(v_i)=0 \bmod {p^k \mathbb {Z}_p}$ .
Then for every
$l\leq k$
and every
$\xi $
lying in the standard affinoid
$X_{p,\Lambda }^{\leq l}$
,

In particular, if a sequence of pairs
$v_1^{i}, v_2^{i}\in \Lambda '$
satisfies
-
○
$v_1^{i} \equiv \alpha _i \cdot v_2^{i} \bmod {p^{k_i} \Lambda }$ for some
$\alpha _i \in \mathbb {Z}_p^{\times }$ ,
-
○
$q(v_1^i)=0 \bmod {p^{k_i} \mathbb {Z}_p}$ .
-
○
$k_i \to +\infty $ ,
then the product

converges to an element of
$\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times }$
.
Proof. Suppose
$\xi $
is represented by a primitive vector
$v_\xi \in \Lambda ^{\prime }_{O_{\mathbb {C}_p}}.$
The inequality

holds for all
$\xi \in X_{p,\Lambda }^{\leq n}$
since
$k\geq l$
. Thus, we can compute

for all such
$\xi $
, which proves the first claim. The second claim is a direct consequence of this inequality.
3.4.2 Convergence
Assume for the remainder of this section that
$s=0$
. Fix elements
$\beta _1,\ldots \beta _l \in \mathbb {D}_{L}$
, positive rational numbers
$m_1,\ldots ,m_l\in \mathbb {Z}_{(p)}^{>0}$
, and integers
$c_{m_1,\beta _1},\ldots ,c_{m_l,\beta _l}\in \mathbb {Z}$
such that

for all modular forms
$f \in M_{n/2,\mathbb {D}_{L}}(\mathscr {G}_0(p))$
. Consider the Kudla–Millson divisor

Note that since the quadratic form q is definite, the set
$\mathcal {O}_L(m_i,\beta _i)\cap \Lambda $
is finite for every
$\mathbb {Z}_p$
-lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
. The following lemma states the convergence of p-adic Borcherds products.
Proposition 3.27. For every self-dual
$\mathbb {Z}_p$
-lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
the infinite product

converges in
$\mathcal {M}_{\mathrm {rq}}^{\times } / \mathbb {Z}_p^{\times }$
. Moreover, the following equality holds:

Proof. After scaling by appropriate powers of p, we get the equality

in
$\mathcal {M}_{\mathrm {rq}}^{\times } / \mathbb {Z}_p^{\times }$
, where the factors are given by

and

for
$t\geq 1$
. The conditions
$q(v) = m_i p^{2t}$
and v is primitive in
$\Lambda $
imply in particular that v generates an isotropic line in
$\Lambda / p^t \Lambda .$
It is then natural to decompose
$J_{\Lambda ,t}$
further as a product over isotropic lines:

where
$\hat {J}_{t,\ell }$
is the sub-product of
$\hat {J}_t$
determined by insisting that the image of
$v \bmod p^t\Lambda $
generates
$\ell $
.
We claim that the total sum of exponents in the product expression for
$\hat {J}_{\Lambda ,t,\ell }$
equals 0. The desired sum of exponents equals

Lemma 3.12 applied to the theta series of the definite quadratic lattice
$L\cap p^t \Lambda $
implies that
$a_{m_i,\beta _i}$
is the
$(m_i,\beta _i)$
-Fourier coefficient of a modular form of weight
$n/2$
, type
$\rho _{\mathbb {D}_L}$
and level
$\mathscr {G}_0(p)$
. Therefore, the claim follows by the hypothesis made on the integers
$c_{m_i,\beta _i}$
.
By Lemma 3.26, it follows that the product
$\xi \mapsto \prod _{t = 0}^\infty \hat {J}_{\Lambda ,t}(\xi )$
converges. The last assertion is an immediate consequence of the construction of
$\hat {J}_{\underline {c},\Lambda }$
.
3.4.3 Action of the p-arithmetic group
One easily deduces the following transformation law of
$\hat {J}_{\underline {c},\Lambda }\in \mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times }$
under the action of the p-arithmetic group
$\Gamma $
from the construction.
Lemma 3.28. Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be a self-dual
$\mathbb {Z}_p$
-lattice and
$\gamma \in \Gamma $
. The equality

holds.
3.4.4 Independence of lattice at p
We show that the function
$\hat {J}_{\underline {c},\Lambda }$
is independent of the choice of lattice at p. The proof relies on the properties of p-neighbouring lattices, which we are going to recall first. Remember that two self-dual
$\mathbb {Z}_p$
-lattices in
$V_{\mathbb {Q}_p}$
are called p-neighbours if their intersection is of index p in each of them. The following alternative description of p-neighbours of self-dual lattices is well-known (see, for example, the discussion after [Reference Chenevier14, Lemma 4.3]):
Lemma 3.29. Let
$\Lambda _1,\Lambda _2\subseteq V_{\mathbb {Q}_p}$
be self-dual
$\mathbb {Z}_p$
-lattices, which are p-neighbours. Then there exists a primitive isotropic vector
$w\in \Lambda ^{\prime }_1$
such that

The following important property of p-neighbours goes back to the seminal work of Kneser (cf. [Reference Kneser30]):
Lemma 3.30. Any two self-dual
$\mathbb {Z}_p$
-lattices in
$V_{\mathbb {Q}_p}$
can be connected by a finite chain of self-dual
$\mathbb {Z}_p$
-lattices in which any two consecutive lattices are p-neighbours.
Lemma 3.31. Let
$\Lambda _1,\Lambda _2 \subseteq V_{\mathbb {Q}_p}$
be self-dual
$\mathbb {Z}_p$
-lattices. The equality

holds in
$\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times }$
.
Proof. By Lemma 3.30, it is enough to consider the case that
$\Lambda _1$
and
$\Lambda _2$
are p-neighbours. Thus, by Lemma 3.29, there exists a primitive isotropic vector
$w\in \Lambda ^{\prime }_1$
such that

From the definition, one immediately gets that

Suppose that
${\mathrm {ord}}_{\Lambda _2}(v)\geq -t$
for some
$t \geq 1$
; that is, v is of the form

with
$a\in \mathbb {Z}_p$
and
$u \in \Lambda _1$
satisfying
$\langle u,w \rangle \equiv 0 \bmod p.$
Then
${\mathrm {ord}}_{\Lambda _1}(v) < -t$
if and only if
${\mathrm {ord}}_{\Lambda _1}(v)=-t-1$
. This happens exactly when
-
(i)
${\mathrm {ord}}_{\Lambda _2}(v) = -t$ and
-
(ii) p does not divide a.
Now let
$v_1,v_2 \in V_{\mathbb {Q}_p}$
with
${\mathrm {ord}}_{\Lambda _2}(v_i) \geq -t$
for
$i=1,2$
such that
$p^{t}v_1$
and
$p^{t}v_2$
generate the same line in
$\Lambda _2/p^t\Lambda _2$
. Using the above criterion, one readily checks that the orders of
$v_1$
and
$v_2$
with respect to
$\Lambda _1$
are either both less than
$-t$
or both greater or equal to
$-t$
. In particular, the second factor in (35) is equal to a product of factors
$\hat {J}_{\Lambda _2,k,\ell }^{-1}$
for certain isotropic lines
$\ell \subseteq \Lambda _2/p^k\Lambda _2$
. As in the proof of Proposition 3.27, we deduce that this product converges to
$1$
in
$\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times }$
. By symmetry of the p-neighbour relation, the first factor in (35) also converges to
$1$
.
3.4.5 Proof of Theorem 3.18
Let
$\Lambda \subseteq V_{\mathbb {Q}_p}$
be any self-dual
$\mathbb {Z}_p$
-lattice. By Lemma 3.31, the function

is independent of the choice of
$\Lambda \subseteq V_{\mathbb {Q}_p}$
. It is therefore
$\Gamma $
-invariant by Lemma 3.28, and by construction, its divisor is equal to
${\mathscr {D}}_{\underline {c}}$
. Hence, Theorem 3.18 follows.
3.5 Proof of modularity in the hyperbolic case
The proof of the main theorem in the hyperbolic setting is very close to the one in the definite setting. We will only indicate the main differences. Throughout this section, we assume that
$s=1$
and
$Q(\mathbb {Q})\neq \emptyset $
. We fix elements
$\beta _1,\ldots \beta _l \in \mathbb {D}_{L}$
, positive rational numbers
$m_1,\ldots ,m_l\in \mathbb {Z}_{(p)}^{>0}$
, and integers
$c_{m_1,\beta _1},\ldots ,c_{m_l,\beta _l}\in \mathbb {Z}$
. Moreover, we assume that each
$m_i$
is compact with respect to
$(V,q)$
.
3.5.1 Convergence in the hyperbolic case
Let us fix rational isotropic lines
$\ell _{-},\ell _{+}\in Q(\mathbb {Q})$
. The set
$\{v\in \mathcal {O}_L(m_i,\beta _i)\cap \Lambda \ \vert \ \Delta _{v,\infty }\cap [\ell _{-},\ell _{+}]\neq 0\}$
is finite by Lemma 2.28. Replacing the modularity of theta series of definite quadratic lattices by Theorem 3.17 in the proof of Proposition 3.27 yields the following:
Proposition 3.32. For every self-dual
$\mathbb {Z}_p$
-lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
, the infinite product

converges in
$\mathcal {M}_{\mathrm {rq}}^{\times } / \mathbb {Z}_p^{\times }$
. Moreover, the following equality holds:

3.5.2 Properties
From the construction, one immediately deduces the following:
Lemma 3.33. For every self-dual
$\mathbb {Z}_p$
-lattice, the assignment

defines an
$\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times }$
-valued modular symbol. Moreover,

for every
$\gamma \in \Gamma $
.
Again, by replacing the modularity of theta series of definite lattices by Theorem 3.17, the same arguments as in the proof of Lemma 3.31 give the following:
Lemma 3.34. Let
$\ell _{-},\ell _{+}\in Q(\mathbb {Q})$
be rational isotropic vectors and
$\Lambda _1,\Lambda _2 \subseteq V_{\mathbb {Q}_p}$
self-dual
$\mathbb {Z}_p$
-lattices. The equality

holds in
$\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times }$
.
To summarize, for every every self-dual
$\mathbb {Z}_p$
-lattice
$\Lambda \subseteq V_{\mathbb {Q}_p}$
, the assignment

defines a
$\Gamma $
-invariant
$\mathcal {M}_{\mathrm {rq}}^{\times }/\mathbb {Z}_p^{\times }$
-valued modular symbol, whose divisor is given by
$\widetilde {{\mathscr {D}}_{\underline {c}}}$
, which proves Theorem 3.24.
4 Special points and singular moduli
This chapter introduces the notion of special points on
$X_p$
and defines the value of a rigid meromorphic cocycle at such a point. Since a special point is an element of
$X_p$
whose stabiliser in G contains a suitable kind of maximal torus, a general discussion in Section 4.1 of maximal tori in orthogonal groups precedes the definition, which is given in Section 4.2. Section 4.3 describes a Galois action on special points akin to the Galois action on CM points of Shimura varieties, which is the basis of a Shimura reciprocity law for special values of rigid meromorphic cocycles.
As in the previous chapters, V is a quadratic space over
$\mathbb {Q}$
of signature
$(r,s)$
and dimension
$n = r+s$
. It is assumed throughout this chapter that
$r\geq s$
, and the algebraic closure
$\overline {\mathbb {Q}}$
is considered as a subfield of
$\mathbb {C}_p$
via a fixed chosen embedding. Given a field K with a fixed algebraic closure
$\overline {K}$
, the absolute Galois group of K is denoted by
$\mathcal {G}_{K}=\operatorname {\mathrm {Gal}}(\overline {K}/K)$
.
4.1 Maximal tori in orthogonal groups
A subtorus T of G is an algebraic subgroup
$T \subset G$
defined over
$\mathbb {Q}$
that becomes isomorphic over
$\overline {\mathbb {Q}}$
to a product of multiplicative groups. Given such a torus, we write
$X^{\ast }(T)$
for the group of
$\overline {\mathbb {Q}}$
-rational characters of
$T_{\overline {\mathbb {Q}}}$
. As in Section 1.3.4, we attach to each
$\chi \in X^*(T)$
the associated eigenspace

for the action of
$T_{\overline {\mathbb {Q}}}$
on
$V_{\overline {\mathbb {Q}}}$
. The set
$\mathcal {W}_T^\circ \subseteq X^\ast (T)$
of characters
$\chi $
for which
$V_\chi \neq \{0\}$
is a finite
$\mathcal {G}_{\mathbb {Q}}$
-set of cardinality
$\le n$
, and
$V_{\overline {\mathbb {Q}}}$
is equipped with the weight space decomposition

As explained in Section 1.3.4, the weight spaces
$V_{\chi }$
and
$V_{\chi '}$
are orthogonal to each other unless
$\chi \chi '=1$
. The nondegeneracy of the quadratic form therefore implies that the dimensions of
$V_{\chi }$
and
$V_{\chi ^{-1}}$
agree, for all
$\chi \in X^\ast (T)$
. In particular, the involution
$\sigma $
sending
$\chi $
to
$\chi ^{-1}$
operates on
$\mathcal {W}_T^\circ $
.
The torus T is maximal (with respect to inclusion of tori) if and only if the eigenspaces in (36) are all one-dimensional (i.e., if
$\mathcal {W}_T^\circ $
has cardinality n). In that case, the
$\mathbb {Q}$
-algebra

of linear transformations that commute with
$T(\mathbb {Q})$
becomes isomorphic to
${\overline {\mathbb {Q}}}^n$
when tensored with
${\overline {\mathbb {Q}}}$
(i.e., is an étale
$\mathbb {Q}$
-algebra of dimension n). It is also endowed with a canonical involution
$\sigma $
sending a linear transformation to its adjoint. When
$n=\dim (V)$
is odd, T acts trivially on a unique nondegenerate one-dimensional subspace of V and acts faithfully on its orthogonal complement, denoted
$V_T$
. When n is even, T acts faithfully on
. In both cases, write
$E_T$
for the image of
$E_T^{\circ }$
in the endomorphism ring of
$V_T$
. The action of
$\sigma $
naturally descends to one on
$E_T$
. The group
$T(\mathbb {Q})$
can then be recovered from the pair
$(E_T, \sigma )$
as

The dimension of the torus T is therefore given by

where
$F_T$
is the étale subalgebra of
$\sigma $
-fixed elements of
$E_T$
. One easily checks that
$F_T$
is an étale
$\mathbb {Q}$
-algebra whose dimension is half the dimension of
$E_T$
(see, for example, [Reference Brusamarello, Chuard-Koulmann and Morales12, Proposition 3.3]).
4.1.1 Description of V in terms of T
The following Lemma, which also plays a crucial role in [Reference Darmon and Vonk17], was pointed out to the authors by Jan Vonk (Cf. [Reference Brusamarello, Chuard-Koulmann and Morales12, Proposition 3.9]).
Lemma 4.1. There are unique hermitian bilinear pairings (relative to
$\sigma $
)

satisfying

and likewise with
$E_T$
replaced by
$F_T$
.
Proof. The étale algebra
$E_T$
is canonically identified with its
$\mathbb {Q}$
-linear dual via the trace form. For any
$v_1, v_2\in V_T$
, the
$\mathbb {Q}$
-valued functional
$\lambda \mapsto \langle \lambda v_1, v_2\rangle $
on
$E_T$
can therefore be represented as
$\lambda \mapsto \mathrm {Tr}^{E_T}_{\mathbb {Q}}(\lambda x)$
for a unique
$x\in E_T$
, depending on
$v_1$
and
$v_2$
. Setting
, it is readily checked that the resulting
$E_T$
-valued function on
$V_T \times V_T$
has the asserted bilinearity properties. The same argument works with
$E_T$
replaced by
$F_T$
.
The
$F_T$
and
$E_T$
-valued pairings are trace compatible; that is,

The quantity
$\langle v,v\rangle _{E_T}$
belongs to
$F_T$
for any nonzero vector v in the one-dimensional
$E_T$
-vector space
$V_T$
, and it follows that the quantity

is well-defined up to norms from
$E_T^{\times }$
. Its image in
$F_T^{\times }/\mathrm { N}^{E_T}_{F_T}(E_T^{\times })$
is a well-defined invariant of the torus T.
Lemma 4.2. The quadratic space
$V_T$
is isomorphic to
$E_T$
equipped with the quadratic form

Proof. The choice of an
$E_T$
-module generator
$v\in V_T$
determines a
$\mathbb {Q}$
-vector space isomorphism

Under this identification, the original pairing on
$V_T$
is transported to the pairing
$\langle \!\langle \ , \ \rangle \!\rangle $
on
$E_T$
given by

so that, in particular,

The lemma follows.
Lemma 4.2 leads to a useful description of the quadratic space V. If n is odd, the orthogonal complement of
$V_T$
in V is one-dimensional and nondegenerate, with quadratic form given by
${q_\alpha (x)=\alpha x^2}$
for some
$\alpha \in \mathbb {Q}^{\times }$
which is uniquely determined up to squares. As quadratic spaces over
$\mathbb {Q}$
, one has

Elements of the torus
$T\subseteq G$
act
$E_T$
-linearly on
$V_T$
viewed as a one-dimensional
$E_T$
-vector space, and preserve the
$\sigma $
-hermitian form
$\langle \ , \ \rangle _{E_T}$
. Hence, T is identified with the unitary group

acting on
$E_T$
via multiplication under the identification (37). (Cf. [Reference Brusamarello, Chuard-Koulmann and Morales12, Proposition 3.3], for instance.)
4.1.2 Aside: spinor norms
This section recalls a well-known formula for the spinor norm of elements of
$\mathrm {U}(E_T,\sigma )$
, viewed as a subgroup of the group of orthogonal group of the quadratic form
$\langle \ , \ \rangle _{F_T}$
. Recall that a norm one element of
$E_T^{\times }/F_T^{\times }$
is of the form
$a/a^{\sigma }$
, for some
$a\in E_T^{\times }$
, by Hilbert’s Theorem 90.
For extra generality, fix a field k of characteristic different from
$2$
.
Lemma 4.3. Let
$l/k$
be a quadratic étale algebra and denote by
$\sigma $
the nontrivial k-automorphism of l. Fix an element
$\nu \in k^{\times }$
and consider the k-vector space l with the quadratic form over k given by
. Then the spinor norm of multiplication with
$a/a^{\sigma }$
is equal to
$\mathrm {N}^{l}_k(a) \bmod (k^{\times })^2$
, for every
$a\in l^{\times }$
.
Proof. Given any
$a\in l$
, the reflection about a, denoted
$\tau _a$
, is the transformation sending a to
$-a$
and fixing the orthogonal complement of a relative to the trace form. The spinor norm is defined by the values it takes on these reflections, by the rule

Assume without loss of generality that
$\nu =1$
, since the spinor norm of elements in
${\mathrm {SO}}(q_k)$
is independent of scaling of the quadratic form. One can check that
$\tau _a$
is given by the explicit formula

In particular, the equality
$\tau _a \circ \tau _1 = a/a^\sigma $
holds, and the claim follows by taking the spinor norm on both sides of this equality.
The following lemma describes the behaviour of spinor norms under restriction of scalars.
Lemma 4.4. Let
$l/k$
be a finite étale algebra. Let
$(W,q_\ell )$
be a nondegenerate quadratic space over l. Then W equipped with
defines a nondegenerate quadratic space over k. There is a canonical embedding
${\mathrm {SO}}(q_l)\subseteq {\mathrm {SO}}(q_k)$
, and the following diagram involving the spinor norms
$\mathrm {sn}_{q_l}$
and
$\mathrm {sn}_{q_k}$
associated to
$q_l$
resp.
$q_k$
is commutative:

Proof. In case l is a field, this is [Reference Scharlau40, Chapter 2, Lemma 5.5] and [Reference Scharlau40, Chapter 9, Example 3.7]. The proof carries over verbatim to the more general case.
Combining the two lemmas above one deduces the following:
Corollary 4.5. Let
$k_1/k$
be a finite étale algebra,
$\nu $
an element of
$k_1^{\times }$
and
$l/k_1$
a quadratic étale algebra with involution
$\sigma $
. Equip the k-vector space l with the nondegenerate quadratic form
$q_k(\lambda )=\mathrm {Tr}^{l}_{k_1}(\nu \cdot \mathrm {N}^{k_1}_{k}(\lambda ))$
. For every
$a\in l^{\times }$
, the spinor norm of multiplication with
$a/a^\sigma $
is equal to
$\mathrm {N}^{l}_{k}(a) \bmod (k^{\times })^2$
.
4.1.3 Maximally
$\mathbb {R}$
-split tori
Under the identification
$T\cong U(E_T,\sigma )$
, every arithmetic subgroup of
$T(\mathbb {Q})$
is a finite index subgroup of the group
${\mathcal {O}}_{E_T/F_T}^{\times }$
of relative units of
$E_T/F_T$
, which fits into the exact sequence

in which the last map has finite cokernel, since the image of the norm map contains the squares of units in
$\mathcal {O}_{F_T}^{\times }$
, and therefore, the cokernel is of exponent
$2$
and finitely generated.
The connected component of the commutative real Lie group
$T(\mathbb {R})$
is isomorphic to

where
${\mathbb S}_1 \subseteq \mathbb {C}_1^{\times }$
is the unit circle. The integer t is called the real rank of T. The
$\mathbb {Z}$
-rank of every arithmetic subgroup of
$T(\mathbb {Q})$
is less or equal than t by Dirichlet’s unit theorem for algebraic tori. (See, for example, [Reference Shyr43].) It is equal to t if and only if T does not have a
$\mathbb {Q}$
-split subtorus.
Proposition 4.6. For every maximal torus
$T\subseteq G$
of real rank t, one has
$t\le s$
. Moreover, equality holds if and only if
$F_T$
is totally real and admits exactly s real places that split into pairs of real places of
$E_T$
.
Proof. Let a be the number of real places of
$F_T$
that lie below a complex place of
$E_T$
, let b be the number of real places of
$F_T$
that lie below two real places of
$E_T$
, and let c be the number of complex places of
$F_T$
, so that

Let
$\nu \in F_T^{\times }$
be as in Lemma 4.2. The signature of the quadratic form
$q_\nu $
can be read off directly from the invariants
$a,b$
and c above. More precisely, write
$a = a_+ + a_-$
, where
$a_+$
is the number of real places v of
$F_T$
that lie below a complex place of
$E_T$
and for which
$v(\nu )>0$
, and
$a_-$
is the number of those places for which
$v(\nu )<0$
. A direct calculation using the expression of
$q_T$
in terms of the norm form to
$F_T$
shows that

while the real rank of
$T(\mathbb {R})\simeq (\mathbb {C}_1^{\times })^a \times (\mathbb {R}^{\times })^b \times (\mathbb {C}^{\times })^c$
is given by

Comparing the equations for s and t shows that
$t\le s$
, with equality if and only if
$a_- = c = 0$
. The proposition follows.
Proposition 4.6 motivates the following:
Definition 4.7. A maximal torus
$T\subseteq G$
for which the equality
$t=s$
is satisfied is called a maximally
$\mathbb {R}$
-split torus in G.
When
$\Gamma $
is an arithmetic subgroup of G and T is a maximally
$\mathbb {R}$
split torus in G, the group
$\Gamma \cap T(\mathbb {Q})$
is abelian of maximal rank s among abelian subgroups of
$\Gamma $
. Points of the p-adic symmetric space
$X_p$
with such maximal stabiliser, and certain special values of rigid meromorphic cocycles at such special points, are of great interest in connection with explicit class field theory for non-totally real fields.
4.1.4 Weight spaces and fixed points
In this section, it is assumed throughout that T is a maximal torus. Because the
$\chi $
-isotypic line
$V_{\chi }$
is isotropic for all
, it defines a point
$x_\chi \in Q(\overline {\mathbb {Q}})$
. The map
$\chi \mapsto x_\chi $
sets up an inclusion of
$\mathcal {W}_T$
into
$Q(\overline {\mathbb {Q}})$
which is compatible with the action of
${\mathcal {G}_{\mathbb {Q}}}$
on both sides. The involution
$\sigma $
on
$\mathcal {W}_T$
sending
$\chi $
to
$\chi ^{-1}$
commutes with the action of
${\mathcal {G}}_{\mathbb {Q}}$
and induces an involution
$\sigma $
on
$\mathcal {W}_T\subseteq Q(\overline {\mathbb {Q}})$
satisfying

In particular, it also commutes with the action of
$G_{\mathbb {Q}}$
, just like its counterpart acting on
$\mathcal {W}_T$
.
Henceforth, we will identify
$\mathcal {W}_T$
with its image in
$Q(\overline {\mathbb {Q}})$
and with the corresponding collection of T-stable isotropic lines in
$V_{\overline {Q}}$
.
Thus, we have shown the following:
Lemma 4.8. Let
$T\subseteq G$
be a maximal torus. Then for all
$\tau \in \mathcal {G}_{\mathbb {Q}}$
and for all
$\chi \in \mathcal {W}_T$
with associated
$x_\chi \in Q(\overline {\mathbb {Q}})$
,

Note that the direct sum is a hyperbolic plane over
$\overline {\mathbb {Q}}$
and that

expresses
$V_{T,\overline {\mathbb {Q}}}$
as the orthogonal direct sum of these hyperbolic planes.
The weight space decomposition of a maximal torus
$T\subseteq G$
can be further described in terms of the étale algebra
$E_T\subseteq \mathrm {End}(V)$
with involution associated to T. Given
$\chi \in \mathcal {W}_T$
, the action of
$E_T$
on the isotropic line
$x_\chi $
determines a
$\mathbb {Q}$
-algebra homomorphism

whose restriction to
$\mathrm {U}(E_T,\sigma )\subseteq E_T^{\times }$
agrees with
$\chi $
. These
$2d$
algebra homomorphisms are all distinct, and thus, they exhaust the set of all
$\mathbb {Q}_p$
-algebra homomorphisms from
$E_T$
to
$\overline {\mathbb {Q}}$
. The bijection

is
$\mathcal {G}_{\overline {\mathbb {Q}}}$
-equivariant. In particular, two points
$\chi ,\chi '\in \mathcal {W}_T$
lie in the same
$\mathcal {G}_{\mathbb {Q}}$
-orbit if and only if
$\ker (\varphi _\chi )=\ker (\varphi _{\chi '})$
. Moreover, the relation

holds.
4.1.5 Fixed points in
$X_p$
The ring homomorphism
$\varphi _x$
induces a
$\mathbb {Q}_p$
-algebra homomorphism

Two points
$\chi ,\chi '\in \mathcal {W}_T$
lie in the same
$\mathcal {G}_{\mathbb {Q}_p}$
-orbit if and only if
$\ker (\varphi _{x,p})=\ker (\varphi _{x',p})$
. In case n is odd, one views
$\varphi _{x,p}$
a homomorphism
$\varphi _{x,p}\colon (E_T\oplus \mathbb {Q})\otimes _{\mathbb {Q}} \mathbb {Q}_p \rightarrow \mathbb {C}_p$
via extension by zero. The following lemma clarifies when a fixed point of
$T(\mathbb {C}_p)$
belongs to
$X_p$
.
Lemma 4.9. Let
$T\subseteq G$
a maximal torus and
$\chi \in \mathcal {W}_T$
a fixed point of
$T(\mathbb {C}_p)$
.
-
(a) Assume that n is even. Then x belongs to
$X_p$ if and only if the subspace
$\ker (\varphi _{\sigma (x),p})\subseteq E_T\otimes \mathbb {Q}_p$ contains no nonzero isotropic vectors (relative to the quadratic form
$q_\nu $ ).
-
(b) Assume that n is odd. Then x belongs to
$X_p$ if and only if the subspace
$\ker (\varphi _{\sigma (x),p})\subseteq (E_T\oplus \mathbb {Q})\otimes \mathbb {Q}_p$ contains no nonzero isotropic vectors.
Proof. Under the identification
$V\cong E_T$
respectively
$V\cong E_T\oplus \mathbb {Q}$
, the orthogonal complement of the isotropic line x is given by the kernel of
$\varphi _{\sigma (x)}$
. Thus, the lemma follows from the definition of
$X_p$
.
Corollary 4.10. Let
$T\subseteq G$
be a maximal torus. Assume that
$E_T\otimes \mathbb {Q}_p$
is a field. Then every fixed point
$x\in \mathcal {W}_T$
belongs to
$X_p$
.
Proof. The claim follows from the preceding lemma since every
$\mathbb {Q}_p$
-algebra homomorphism
$\varphi \colon E_T\otimes \mathbb {Q}_p \rightarrow \mathbb {C}_p$
is injective.
Definition 4.11. A point
$x \in X_p$
is called a toric point if it is stabilized by
$T(\mathbb {C}_p)$
where
$T\subseteq G$
is a maximal torus. In that case,
$(x,T)$
is called a toric pair.
Remark 4.12. Given a toric point
$x\in X_p$
there are in general several maximal tori
$T\subseteq G$
that stabilise x. But under the assumption that the étale algebra
$E_T$
is a field, T is uniquely determined by x. Indeed, the group
$\mathcal {G}_{\mathbb {Q}}$
acts transitively on
${\mathcal N}_T\simeq \mathrm {Hom}(E_T, \overline {\mathbb {Q}})$
by Galois theory, and the torus T can be recovered as the stabiliser in G of x, or equivalently, of all the points of
${\mathcal N}_T$
.
In general, the algebra
$E_T$
associated to a maximal torus
$T\subseteq G$
does not have to be a field. But the existence of a fixed point in
$X_p$
has strong implications on the splitting behaviour of p-adic places in the extension
$E_T/F_T$
:
Lemma 4.13. Let
$(x,T)$
be a toric pair. Then every p-adic place of the étale
$\mathbb {Q}$
-algebra
$F_T$
is non-split in
$E_T$
. Moreover, y and
$\sigma (y)$
lie in the same
$\mathcal {G}_{\mathbb {Q}_p}$
-orbit for all
$y\in \mathcal {W}_T$
.
Proof. Let
$S_p$
be the set of p-adic places of
$F_T$
. The étale
$\mathbb {Q}_p$
-algebra
$F_T \otimes \mathbb {Q}_p$
decomposes into the product over the completions
$F_{\mathfrak {p}}$
,
$\mathfrak {p}\in S_p$
. For
$\mathfrak {p} \in S_p$
, we put
$E_{\mathfrak {p}} = E_T \otimes _{F_T} F_{\mathfrak {p}}$
. The involution
$\sigma $
induces an
$F_{\mathfrak {p}}$
-linear involution on
$E_{\mathfrak {p}}$
. We may decompose
$T_{\mathbb {Q}_p}$
as follows:

Assume that there exists a place
$\mathfrak {p}\in S_p$
that is split in
$E_T$
. Then
$E_{\mathfrak {p}}\cong F_{\mathfrak {p}} \times F_{\mathfrak {p}}$
, and thus, the unitary group
$\mathrm {U}(E_{\mathfrak {p}} ,\sigma )$
is isomorphic to the Weil restriction of
$\mathbb {G}_{m,F_{\mathfrak {p}}}$
from
$\mathbb {F}_p$
to
$\mathbb {Q}_p$
. In particular, it contains a
$\mathbb {Q}_p$
-split subtorus. Proposition 1.7 now implies that no fixed point of
$T(\mathbb {C}_p)$
lies in
$X_p$
. This proves the first claim. The second claim is a direct consequence of the first one.
A priori the involution
$\sigma $
does depend not only on the toric point x but the toric pair
$(x,T)$
. Lemma 4.13 implies that for every toric pair
$(x,T)$
, the subspace of
$V_{\overline {\mathbb {Q}}}$
generated by the
$\mathcal {G}_{\mathbb {Q}}$
-conjugates of x is a direct sum of hyperbolic planes, and in particular, it is nondegenerate. By construction, it is the base change of a (nondegenerate) rational subspace
$V_x\subseteq V$
that is stable under the action of T. Let
$T\rightarrow {\mathrm {SO}}(V_x)$
be the induced homomorphism, whose image is a maximal torus in
${\mathrm {SO}}(V_x)$
that we denote by
$T_x$
. The inclusion
${\mathrm {SO}}(V_x)\hookrightarrow {\mathrm {SO}}(V)$
restricts to a homomorphism
$T_x \hookrightarrow T$
which is a right inverse to the homomorphism
$T\longrightarrow T_x$
. It follows that
$T_x$
is a direct factor of T. By construction, the étale
$\mathbb {Q}$
-algebra
$E_{T_x}$
(with involution
$\sigma $
) associated to the torus
$T_x$
is a field. Applying Remark 4.12 to the quadratic space
$V_x$
and the torus
$T_x$
we see that
$T_x$
does not depend on the choice of T. In particular, the assignment
$x\mapsto \sigma (x)$
is independent of the choice of the torus T stabilizing x. There is an equality
$\varphi _x(E_T)=\varphi _x(E_{T_x})$
of subfields of
$\mathbb {C}_p$
.
4.2 Special points and values of rigid meromorphic cocycles
After defining oriented special points and the value of a rigid meromorphic cocycle at such a point, a crude form of the algebraicity conjecture on these special values will be formulated in this section.
4.2.1 Special points
Definition 4.14. A point
$x \in X_p$
is said to be special if there exists a maximal
$\mathbb {R}$
-split torus
$T\subseteq G$
such that
$T(\mathbb {C}_p)$
stabilizes x. In that case,
$(x,T)$
is called a special pair.
Let
$(x,T)$
be a special pair. When V is three-dimensional, the étale algebra
$E_T$
with involution is a quadratic field in which p is non-split. It is imaginary when
$s=0$
, and real when
$s=1$
.
When V is four-dimensional, the algebra
$E_T$
can be of one of the following types:
-
1. A quadratic extension of a real quadratic field with exactly
$2s$ real embeddings. The étale
$\mathbb {Q}_p$ -algebra
$E_T\otimes \mathbb {Q}_p$ is either a field or a direct sum of two quadratic extensions of
$\mathbb {Q}_p$ .
-
2. A direct sum
$E_1 \oplus E_2$ of quadratic fields in which the prime p does not split, and where exactly s of the fields are real quadratic.
In all of these cases, the torus T is equal to the stabilizer of x in G.
Lemma 4.15. The étale algebra
$E_T\otimes \mathbb {Q}_p$
over
$\mathbb {Q}_p$
associated to a special point
$x\in X_p$
admits a decomposition

where
$E_{1}$
is a field and
$E_2$
an étale algebra over
$\mathbb {Q}_p$
of rank
$\le 4$
when n is even, and rank
$\le 2$
when n is odd.
Proof. Let
$E_1$
denote the quotient of
$E_T \otimes \mathbb {Q}_p$
through which its action of the isotropic line attached to x factors. This action realises
$E_1$
as a subfield of
$\mathbb {C}_p$
. The orthogonal direct summand of
$V\otimes \mathbb {Q}_p$
identified with
$E_2$
(resp.
$E_2\oplus \mathbb {Q}_p$
) when n is even (resp. odd) is orthogonal to the
$\mathbb {C}_p$
-line attached to x, and hence anisotropic since x belongs to
$X_p$
. This shows that the rank of
$E_2$
is
$\le 4$
when n is even and is
$\le 2$
when n is odd.
In particular, the previous lemma implies that

Only when the first equality is strict can the stabiliser of x contain more than one maximal torus.
4.2.2 Oriented special points
The following proposition controls the rank of p-arithmetic subgroups of the stabilizer of special points of
$X_p$
.
Proposition 4.16. Let
$(x,T)$
be a special pair. The rank of every p-arithmetic subgroup of
$T(\mathbb {Q})$
is equal to s.
Proof. By Proposition 1.7, the base change of T to
$\mathbb {Q}_p$
does not contain any
$\mathbb {Q}_p$
-split subtorus. In particular, T does not contain any
$\mathbb {Q}$
-split subtorus. Hence, the
$\mathbb {Z}$
-rank of every p-arithmetic subgroup of
$T(\mathbb {Q})$
is equal to the real rank of T by Dirichlet’s S-unit theorem for tori (see [Reference Shyr43]). The claim now follows from the definition of special points.
In particular, the subgroup
$\Gamma _T=\Gamma \cap T(\mathbb {Q})$
is a finitely generated abelian group of rank s. The inflation- restriction sequence in group homology shows that
$H_s(\Gamma _T,\mathbb {Z})$
maps to
$H_s(\mathbb {Z}^s,\mathbb {Z})\simeq \mathbb {Z}$
with finite kernel and cokernel, and hence that

Since
$\mathbb {Z}^s$
acts freely and discretely on the contractible space
$\mathbb {R}^s$
, the homology of
$\mathbb {Z}^s$
is identified with that of a real s-torus
$(\mathbb {R}/\mathbb {Z})^s$
, and the choice of a generator of
$H_s(\mathbb {Z}^s,\mathbb {Z})$
amounts to choosing an orientation on this torus. This motivates the following:
Definition 4.17. An oriented special point (of level
$\Gamma $
) is a triple
$\vec {x}=(x,T,o)$
consisting of a special pair
$(x,T)$
together with a generator o of
$\mathrm {H}_s(\Gamma _{T}, \mathbb {Z})$
modulo torsion.
Remark 4.18. Since the order of torsion elements in
$G(\mathbb {Q})$
is bounded from above, it follows that the torsion in
$\mathrm {H}_s(\Gamma _T,\mathbb {Z})$
is bounded from above, where
$(x,T)$
ranges over all special pairs.
In the present context, oriented special points play the role of special points or CM points in the classical theory of Shimura varieties as described in the seminal work of Shimura, Deligne and Mumford.
4.2.3 Regularity
A meromorphic function
$f\in \mathcal {M}_{\mathrm {rq}}^{\times }$
is said to be regular at a point
$\xi \in X_p$
if
$\xi $
does not lie in the support of the divisor of f. Let
$\vec {x}$
be an oriented special point and
$J\in \mathrm {H}^{s}(\Gamma ,\mathcal {M}_{\mathrm {rq}}^{\times })$
. Restriction to
$\Gamma _{T}$
followed by taking the cap product with the orientation o yields a homomorphism

and we write
$J_{\vec {x}}$
for the image of J under this map.
Definition 4.19. A class
$J\in \mathrm {H}^{s}(\Gamma ,\mathcal {M}_{\mathrm {rq}}^{\times })$
is said to be regular at the oriented special point
$\vec {x}$
if the homology class
$J_{\vec {x}}\in (\mathcal {M}_{\mathrm {rq}}^{\times })_{\Gamma _{T}}$
admits a representative
$\widetilde {J}_{\vec {x}}\in \mathcal {M}_{\mathrm {rq}}^{\times }$
that is regular at x.
Remark 4.20. In the case where
$E_T$
is a field and n is even, any function in
$\mathcal {M}_{\mathrm {rq}}^{\times }$
is automatically regular at every special point x, since no nonzero vector in V is orthogonal to x.
For a
$\Gamma $
-invariant subset
$S \subseteq X_p$
, let
$(\mathcal {M}_{\mathrm {rq}}^{\times })_S$
denote those rigid meromorphic functions on
$X_p$
which are regular at all
$\xi \in S$
. If a class J lifts to
$\mathrm {H}^s(\Gamma , (\mathcal {M}_{\mathrm {rq}}^{\times })_{\Gamma x})$
, then it is clearly regular at
$(\vec {x},T)$
. It is worth noting that the p-adic Borcherds products constructed in Theorems 3.19 and 3.25 define natural representative cocycles in
$\mathrm {Z}^s(\Gamma , (\mathcal {M}_{\mathrm {rq}}^{\times })_{S})$
where S are certain unions of rational quadratic divisors.
4.2.4 Values of cohomology classes at oriented special points
Lemma 4.21. Let J be a class in
$\mathrm {H}^{s}(\Gamma ,\mathcal {M}_{\mathrm {rq}}^{\times })$
that is regular at
$\vec {x}$
and let
$\widetilde {J}_{\vec {x}}$
and
$\widetilde {J}_{\vec {x}}^{\prime }\in \mathcal {M}_{\mathrm {rq}}^{\times }$
be two representatives of
$J_{\vec {x}}$
that are regular at x. Then

Proof. Let
$f\in \mathcal {M}_{\mathrm {rq}}^{\times }$
be a function and
$\gamma \in \Gamma _T$
such that
$f/\gamma f$
is regular at x. It is enough to show that
$(f/ \gamma f)(x)=\pm 1$
. We may write f as a product

with
$h\in \mathcal {M}_{\mathrm {rq}}^{\times }$
regular at x,
$v_0\in V_{\mathbb {Q}_p}$
isotropic,
$v_i\in V_+$
,
$i=1,\ldots , k$
, pairwise linearly independent such that
$x \in \Delta _{v_i,p}$
, and
$a_i\in \mathbb {Z}$
. As before,
$v_\xi \in V_{\mathbb {C}_p}$
denotes a generator of the isotropic line
$\xi \in X_p$
. Let g be an element of
$G(\mathbb {Q})$
and
$v\in V$
be an anisotropic eigenvector of g with eigenvalue
$\lambda $
. Then
$\lambda $
is either equal to
$1$
or
$-1$
since

It follows that we can normalize the vectors
$v_i$
,
$i=1,\ldots ,k$
, in such a way that, if
$\gamma v_i$
and
$v_j$
are collinear for some
$\gamma \in \Gamma _T$
, then their ratio is
$\pm 1$
. Moreover, since
$(h/\gamma h)(x)=h(x)/h(\gamma ^{-1}x)=h(x)/h(x)=1$
, we may assume that
$h=1$
. That the function

is regular at x implies that each vector
$v_i$
(modulo sign) appears equally often in the denominator and numerator of the first product. Thus, we deduce that

The claim now follows since x is a fixed point of
$\gamma $
.
Definition 4.22. Let J be a class in
$\mathrm {H}^{s}(\Gamma ,\mathcal {M}_{\mathrm {rq}}^{\times })$
that is regular at the oriented special point
$\vec {x}=(x,T,o)$
. The value of J at
$\vec {x} =(x,T,o)$
is defined by

where
$\widetilde {J}_{\vec {x}}\in \mathcal {M}_{\mathrm {rq}}^{\times }$
is any representative of
$J_{\vec {x}}$
that is regular at x.
Note that the value
$J[\vec {x}]$
lies in the field of definition of x over
$\mathbb {Q}_p$
. In case
$s=0$
, one has
$\mathrm {H}^0(\Gamma _x,\mathbb {Z})=\mathbb {Z}$
. If the orientation o of
$\vec {x}=(x,T,o)$
corresponds to the integer n under this identification, then
$J[\vec {x}]$
is simply the n-th power of the value of the
$\Gamma $
-invariant function J at x. For general s scaling, the orientation defines an action of
$\mathbb {Z}$
on the set of oriented special points. One clearly has

In particular, if the orientation o of
$\vec {x} $
is torsion,
$J[\vec {x}]$
is a root of unity whose order is bounded from above by Remark 4.18.
4.2.5
$\Gamma $
-invariance of special values
Let
$(x,T)$
be a special pair and g an element of
$G(\mathbb {Q})$
. Then
$(g x, g T g^{-1})$
is also a special pair and
$\Gamma _{g T g^{-1}}=g\Gamma _T g^{-1}$
. Write

for the homomorphism induced by conjugation with g. If
$\vec {x}=(x,T,o)$
is an oriented special point, then so is
. Since conjugation by
$\gamma \in \Gamma $
induces the identity on
$\mathrm {H}^{s}(\Gamma , \mathcal {M}_{\mathrm {rq}}^{\times })$
, we immediately get the following:
Proposition 4.23. Suppose that a class
$J\in \mathrm {H}^{s}(\Gamma ,\mathcal {M}_{\mathrm {rq}}^{\times })$
is regular at the oriented special point
$\vec {x}$
. Then J is regular at
$\gamma .\vec {x}$
for all
$\gamma \in \Gamma $
. Moreover, the equality

holds.
4.2.6 Algebraicity of special values
In general, we do not expect the values of rigid meromorphic cocycles at oriented special points to be algebraic. Indeed, in signature
$(n,0)$
, every constant function with values in
$\mathbb {Q}_p^{\times }$
is a rigid meromorphic cocycle. More generally, since p-arithmetic groups are of type (VFL), the homomorphism

has finite kernel and cokernel and the group
$\mathrm {H}^s(\Gamma ,\mathbb {Z})$
is finitely generated. It follows that for every
$J\in \mathrm {H}^s(\Gamma ,\mathbb {Q}_p^{\times })$
, there exists a finitely generated submodule
$\Pi _J\subseteq \mathbb {Q}_p^{\times }$
such that

for all oriented special points
$\vec {x}$
. Moreover, the rank of
$\Pi _J$
is bounded by the rank of
$\mathrm {H}^s(\Gamma ,\mathbb {Z})$
. We now state a crude form of the algebraicity conjecture:
Conjecture 4.24. Let
$J\in \mathcal {RMC}(\Gamma )$
be a rigid meromorphic cocycle. There exists a finitely generated submodule
$\Pi _J\subseteq \mathbb {Q}_p^{\times }$
such that

for every oriented special point
$\vec {x}$
at which J is regular. Moreover, the rank of
$\Pi _J$
is less or equal to the rank of
$\operatorname {\mathrm {rk}}_{\mathbb {Z}} \mathrm {H}^s(\Gamma ,\mathbb {Z})$
.
By the discussion following Definition 4.22, the conjecture is vacuous for all points whose orientation is torsion.
Remark 4.25. With similar methods as in the proof of Proposition 3.5, one can determine the rank of
$\mathrm {H}^s(\Gamma ,\mathbb {Z})$
. For example, in signature
$(r,1)$
, one deduces that
$\mathrm {H}^1(\Gamma ,\mathbb {Z})$
is always torsion, and hence, Conjecture 4.24 predicts that the special values of rigid meromorphic cocycles are in fact algebraic. This favourable scenario occurs in the special case of signature
$(2,1)$
explored in [Reference Darmon and Vonk18].
4.3 Class group action on special points
Following Deligne’s treatise of Shimura varieties in [Reference Deligne20], we associate to a toric point x a cocharacter
$\mu _x$
, a reflex field
$E_x$
, and a reflex norm
$r_x$
. This leads to an action of the idele class group of
$E_x$
on the set of oriented special points modulo
$\Gamma $
imitating the Galois action on CM points of Shimura varieties.
4.3.1 The cocharacter
$\mu _x$
and its reflex field
Let
$T\subseteq G$
be a maximal torus and
$x\in Q(\mathbb {C}_p)$
a point stabilized by
$T(\mathbb {C}_p)$
. Remember that x corresponds to a nontrivial weight
$\chi $
of T and, in particular, that x is algebraic. We consider x as an isotropic line in
$V_{\overline {\mathbb {Q}}}$
. As before, write
$\sigma (x)$
for the T-weight space corresponding to the weight
$\chi ^{-1}$
and consider the hyperbolic summand
$\mathrm {H}_{x}=x\oplus \sigma (x)$
of
$V_{\overline {\mathbb {Q}}}$
attached to x by (39). The assignment

defines a cocharacter
$\mu _{x,T}\colon \mathbb {G}_{m,\overline {\mathbb {Q}}} \rightarrow T_{\overline {\mathbb {Q}}}$
.
Assume for the moment that x is an element of
$X_p$
. Then, by the discussion following Lemma 4.13, the point
$\sigma (x)$
does only depend on x and not T. Hence, the cocharacter
$\mu _{x,T}$
is independent of T and defines a homomorphism

where
$T_x$
is the factor of T associated to x in Section 4.1.5.
Definition 4.26. Let x be a toric point. The reflex field
$E_x \subseteq \mathbb {C}_p$
of x is the field of definition of
$\mu _x$
.
The fields of definition of the points x and
$\sigma (x)$
agree, and thus, the reflex field
$E_x$
is just the field of definition of the point
$x\in X_p$
. It is a finite extension of
$\mathbb {Q}$
. More precisely, one has the following equality of fields:

The isomorphism
$\varphi _x\colon E_{T_x}\rightarrow E_x$
induces an isomorphism

of algebraic groups over
$\mathbb {Q}$
.
4.3.2 Abstract reflex norm associated to a cocharacter
Let T be a torus over
$\mathbb {Q}$
with cocharacter lattice
$X_\ast (T)$
. Let
$\mu \colon \mathbb {G}_{m,E} \rightarrow T_E$
be a cocharacter defined over a finite extension E of
$\mathbb {Q}$
. Write
for the Weil restriction of the multiplicative group
$\mathbb {G}_{m,E}$
from E to
$\mathbb {Q}$
. As explained in the proof of [Reference Oesterlé39, Chapter II, Theorem 2.4], there exists a canonical isomorphism

By Frobenius reciprocity, the
$\mathcal {G}_{E}$
-equivariant map of cocharacter lattices

induces the
$\mathcal {G}_{\mathbb {Q}}$
-equivariant homomorphism

We are led to the following definition:
Definition 4.27. Let
$T / \mathbb {Q}$
be a torus and
$\mu \colon \mathbb {G}_{m,E} \rightarrow T_E \in X(T)$
be a cocharacter defined over a finite extension E of
$\mathbb {Q}$
. The reflex norm attached to the pair
$(\mu ,E)$
is the unique
$\mathbb {Q}$
-rational homomorphism

that induces the homomorphism
$\overline {r_\mu }$
on cocharacter lattices.
4.3.3 Reflex norm
$r_x$
attached to fixed points in
$X_p$
The above construction can be applied to the cocharacter
$\mu _x$
of Section 4.3.1.
Definition 4.28. Let
$x\in X_p$
be a toric point with corresponding cocharacter
$\mu _x$
and reflex field
$E_x$
. The associated reflex norm

is the abstract reflex norm associated to the pair
$(\mu _x, E_x)$
.
Lemma 4.29. Let
$x \in X_p$
be a toric point and let
$g \in G(\mathbb {Q})$
. Then g stabilizes x if and only if g commutes with
$\mu _x$
. In particular, if g stabilizes x, then g commutes with the image of the reflex norm
$r_x$
.
Proof. The cocharacter
$\mu _x$
defines an action of
$\mathbb {G}_{m,E}$
on
$V_E$
with distinct one-dimensional weight spaces spanned by x and
$\sigma (x)$
. So if g commutes with
$\mu _x$
, then g must preserve these weight spaces and so must stabilize x and
$\sigma (x)$
.
Conversely, by Galois conjugacy of x and
$\sigma (x)$
(see Lemma 4.13),
$g \in G(\mathbb {Q})$
stabilizes x if and only if it stabilizes
$\sigma (x)$
. Since g scales the lines spanned by x and
$\sigma (x)$
and preserves the orthogonal decomposition
$H \oplus \mathrm {H}^{\perp }$
, where H is the hyperbolic plane spanned by x and
$\sigma (x)$
, it follows that g commutes with
$\mu _x$
. If g stabilizes x, and hence commutes with
$\mu _x$
by the above, it follows directly that g commutes with
$r_x$
.
4.3.4 Cocharacter lattices of maximal tori in orthogonal groups
In the following, we will give an explicit description of the reflex norm. We first give a characterization of the cocharacter lattice
$X_\ast (T)$
of a maximal
$\mathbb {Q}$
-rational torus
$T\subseteq G$
in terms of the set
$\mathcal {W}_T$
of isotropic eigenlines of T. To this end, put

The absolute Galois group
$\mathcal {G}_{\mathbb {Q}}$
of
$\mathbb {Q}$
acts naturally on
$\mathcal {W}_T$
, and so it also does on
$X_\ast ^{\mathrm {geom}}(T)$
. Given
$f\in X_\ast ^{\mathrm {geom}}(T)$
and
$t\in \mathbb {C}_p^{\times }$
, we define the orthogonal operator

via
$\mu _f(t)v=t^{f(x)}v$
if
$v\in x$
,
$x\in \mathcal {W}_T$
and
$\mu _f(t)v=v$
if v is orthogonal to every
$x\in \mathcal {W}_T$
. One easily checks that
$\mu _f(t)\in T(\mathbb {C}_p)$
, and thus,
$\mu _{f}$
defines a cocharacter of T.
Lemma 4.30. Let
$T\subseteq G$
be a
$\mathbb {Q}$
-rational maximal torus. The mapping

is a
$\mathcal {G}_{\mathbb {Q}}$
-equivariant isomorphism.
Proof. Via this homomorphism, the function
$f_x= \mathbf {1}_x - \mathbf {1}_{\sigma (x)}$
is mapped to
$\mu _{x,T}$
. The set of cocharacters
$\{\mu _{x,T}\ \vert \ x\in \mathcal {W}_T \}$
generates
$X_\ast (T)$
. This proves surjectivity. Bijectivity follows immediately since both modules are finitely generated free
$\mathbb {Z}$
-modules of the same rank. Lemma 4.8 implies that
$\tau (f_x)=f_{\tau (x)}$
. Thus, for proving Galois equivariance, it suffices to show that

for all
$\tau \in \mathcal {G}_{\mathbb {Q}}$
, which follows directly from the definition of
$\mu _{x,T}$
.
Given a toric pair
$(x,T)$
, define
$\mathcal {W}_x\subseteq \mathcal {W}_T$
to be the
$\mathcal {G}_{\mathbb {Q}}$
-orbit of
$\chi _x$
and write
$X_{\ast }^{\mathrm {geom}}(T_x)\subseteq X_\ast ^{\mathrm {geom}}(T)$
for the subset of those functions, which are supported on
$\mathcal {W}_x$
. By Lemma 4.30, there is a canonical identification

Since
$E_x$
is the field of definition of x, the map

is a well-defined isomorphism of
$\mathcal {G}_{\mathbb {Q}}$
-sets, which induces the isomorphism of
$\mathcal {G}_{\mathbb {Q}}$
-modules

By (40), the right-hand side of this homomorphism can be identified with the cocharacter lattice of
$\underline {E}_x^{\times }$
. Thus, restricting (42) to
$X_\ast ^{\mathrm {geom}}(T_x)$
yields the map

which is induced by the homomorphism of tori

4.3.5 Concrete identification of the reflex norm
The concrete reflex norm attached to a toric point
$x\in X_p$
is the homomorphism

of algebraic groups over
$\mathbb {Q}$
.
Proposition 4.31. For every toric point
$x\in X_p$
, the equality

holds. In particular, the kernel of
$r_x$
is equal to the subgroup
$\underline {F}_x^{\times }\subseteq \underline {E}_x^{\times }$
.
Proof. For the first claim, it is enough to prove that both homomorphisms induce the same map on cocharacter groups. Under the identification (41), the reflex norm induces the following map on cocharacter groups:

Thus, the composition
$\varphi _x \circ r_x$
induces the map

which is readily identified as the one induced by
$r_\sigma $
.
The second claim follows since
$\ker (r_\sigma )=\underline {F}_x^{\times }$
and
$\varphi _x\colon T_x \to \mathrm {U}(E_x,\sigma )$
is an isomorphism.
Proposition 4.31 together with Corollary 4.5 immediately implies the following:
Corollary 4.32. Let
$x \in X_p$
be a toric point. The formula

holds for all
$e\in E_x^{\times }$
.
4.3.6 Class group action on toric points
Denote by
$\mathbb {A}$
(resp.
$\mathbb {A}^{p}_{f}$
) the ring of adeles (resp. the ring of adeles away from p and
$\infty $
). If E is a finite extension of E, put
and
. Recall that
is the special orthogonal group of V, viewed as a group over
$\mathbb {Z}$
after fixing a lattice
$V_{\mathbb {Z}}\subseteq V$
. Fix an open compact subgroup
$U^{p}\subseteq G(\mathbb {A}^{p}_{f})$
and an open subgroup
$U_p\subseteq G(\mathbb {Q}_p)$
containing
$G(\mathbb {Q}_p)^+$
and put
. For a coset
$h\in G(\mathbb {A})/U$
, consider the stabilizer
of h in
$G(\mathbb {Q})$
. This is a p-arithmetic subgroup that lies in the kernel of the real spinor norm. The set

of adelic toric points (of level U) is equipped with the diagonal
$G(\mathbb {Q})$
-action. For a toric point x, let
$x_U\subseteq \Sigma _U$
be the set of pairs with first coordinate equal to x and write
for its image in
$G(\mathbb {Q})\backslash \Sigma _U$
. Given a point
$(x,h)\in \Sigma _U$
, the reflex norm defines an action of the idele group
$\mathbb {I}_{E_x}=\underline {E}_x^{\times }(\mathbb {A})$
on
$x_U$
:

Lemma 4.33. Let
$x\in X_p$
be a toric point. The action of
$\mathbb {I}_{E_x}$
on
$x_{U}$
descends to a well-defined action on
$\overline {x_U}$
.
Proof. Suppose
$ (x,h)$
and
$ (x,h')$
are
$G(\mathbb {Q})$
-equivalent points of
$x_U$
(i.e.,
$ g.(x,h) = (x,h')$
for some
$g \in G(\mathbb {Q})$
). In particular,
$gx = x$
holds, which by Lemma 4.29 implies that g commutes with the image of the reflex norm. It follows that

which proves the claim.
To an adelic toric point
$(x,h)\in \Sigma _U$
, we attach the relative class group

Corollary 4.34. Let
$(x,h)$
be an adelic toric point. The action of
$\mathbb {I}_{E_x}$
on the orbit of
$(x,h)$
in
$\overline {x}_U$
factors through
$C_{x,h}$
.
Proof. By Proposition 4.31, the subgroup
$\mathbb {I}_{F_x}$
is identified with
$\ker (r_x)(\mathbb {A})$
, which acts trivially on
$x_U$
and hence also on
$\overline {x}_U$
. Since
$r_x(E_x^{\times })$
is a subgroup of
$G(\mathbb {Q})$
, the subgroup
$E_x^{\times }$
acts trivially on
$\overline {x}_U$
. Finally,
$r_x^{-1}(h U \mathrm {H}^{-1})$
acts trivially on h by definition. The result follows.
4.3.7 Class group action on oriented special points
Definition 4.35. An adelic oriented special point
$(\vec {x},h)$
of level U consists of a coset
$h\in G(\mathbb {A})/U$
and an oriented special point
$\vec {x}$
of level
$\Gamma _h$
. The set of oriented adelic special points of level U will be denoted by
$\Sigma _U^{\mathrm {or}}$
.
The set
$\Sigma _U^{\mathrm {or}}$
carries a natural
$G(\mathbb {Q})$
-action: let
$(\vec {x},h)$
be an adelic oriented special point given by the tuple
$(x,T,h,o)$
and
$g\in G(\mathbb {Q})$
. The adelic oriented special pair
$g(\vec {x},h))$
is given by the tuple
$(gx,g T g^{-1}, gh, c_g(o))$
, where as before

denotes the homomorphism induced by conjugation by g. Consider the subset
$x_U^{\mathrm {or}}\subseteq \Sigma _U^{\mathrm {or}}$
of all adelic oriented special points with first coordinate equal to x. As in Section 4.3.6, the reflex norm induces an action of
$\mathbb {I}_{E_x}$
on
$x_U^{\mathrm {or}}$
: given
$t\in \mathbb {I}_{E_x}$
and
$(x,T,h,o)\in x_U^{\mathrm {or}}$
, then
$t\star (\vec {x},h)$
is associated to the special pair
$(x,T)$
, the coset
$r_x(t)h$
and the orientation
$c_t(o)$
where

is the homomorphism induced by conjugation with
$r_x(t)$
. With similar arguments as in Lemma 4.33 and Corollary 4.34, one proves the following:
Proposition 4.36. Let
$(\vec {x},h)$
be an adelic oriented special point. The action of
$\mathbb {I}_{E_x}$
on
$x_U$
descends to an action on
. Moreover, the action of
$\mathbb {I}_{E_x}$
on the orbit of the image of
$(\vec {x},h)$
in
$\overline {x}_U^{\mathrm {or}}$
factors through the relative class group
$C_{x,h}$
.
4.3.8 Connected components
Definition 4.37. Two cosets
$h,h'\in G(\mathbb {A})/U$
lie in the same component if their images in
$G(\mathbb {Q})\backslash G(\mathbb {A})/U$
agree.
Assume that
$h,h'\in G(\mathbb {A})/U$
lie in the same component and choose
$g\in G(\mathbb {Q})$
with
$h =g h'$
. The element g is unique up to left multiplication by elements of
$\Gamma _h$
. Let J be an element of
$\mathrm {H}^{s}(\Gamma _h,\mathcal {M}_{\mathrm {rq}}^{\times })$
and
$(\vec {x},h')$
an adelic oriented special point. Then
$g.\vec {x}$
is a special point of level
$\Gamma _h$
. We say that J is regular at
$(\vec {x},h')$
if it is regular at
$g.\vec {x}$
. This regularity condition is independent of the choice of g. Moreover, Proposition 4.23 implies that the special value

is independent of the choice of g as well.
The product of the local spinor norm maps defined a group homomorphism

Since the spin group of a nondegenerate quadratic space is semi-simple and
$G(\mathbb {Q}_p)$
is not compact, strong approximation implies that (44) induces a bijection

Let
$E_U/\mathbb {Q}$
be the abelian extension associated to
$C_{x,h}$
by class field theory. The description of the Galois action on connected components of orthogonal Shimura varieties in [Reference Kudla31, Section 1] suggests that the ‘connected components’ of the space
$G(\mathbb {Q})\backslash (X_\infty \times X_p \times G(\mathbb {A})/U)$
are ‘defined over
$E_U$
’.
Let
$x\in X_p$
be a toric point. The diagram

is commutative by Corollary 4.5 and Proposition 4.31. Thus, for every
$h\in G(\mathbb {A})/U$
, the norm map
$\mathrm {N}^{E_x}_{\mathbb {Q}}\colon \mathbb {I}_E \rightarrow \mathbb {I}_{\mathbb {Q}}$
descends to a homomorphism
$C_{x,h}\rightarrow C_{U}$
. We denote its kernel by
$C_{x,h}^{U}\subseteq C_{x,h}$
. The commutativity of the diagram above implies that h and
$r_x(t)h$
lie in the same component for all
$t\in C_{x,h}^{U}$
. In particular, if
$(\vec {x},h)$
is an adelic oriented special point and
$t\in C_{x,h}^{U}$
, the special value
$J[t\star (\vec {x},h)]$
is defined for all
$J\in \mathrm {H}^{s}(\Gamma _h,\mathcal {M}_{\mathrm {rq}}^{\times })$
.
4.4 The conjectural reciprocity law
We formulate a reciprocity law regarding the algebraicity and field of definition of the values of rigid meromorphic cocycles at oriented special pairs, akin to the classical Shimura reciprocity law for singular moduli on Shimura varieties. Throughout this section, we fix an open subgroup
$U^p \subseteq G(\mathbb {A}^{p,\infty })$
and an open subgroup
$U_p\subseteq G(\mathbb {Q}_p)$
containing
$G(\mathbb {Q}_p)^+$
and put
$U=U^p U_p G(\mathbb {R})^+$
. As before, for a coset
$h\in G(\mathbb {A})/U$
, put
. Let
$(\vec {x},h)$
be an oriented adelic special point and
$C_{x,h}$
the relative class group that is associated to it by (43). Denote by
$\mathrm {H}_{x,h} / E_x$
the abelian extension associated to
$C_{x,h}$
by class field theory; that is, the Artin reciprocity map furnishes a canonical isomorphism

By functoriality of the Artin reciprocity map, the subgroup
$C_{x,h}^{U}$
corresponds to the group of those field automorphisms that act trivially on
$\mathrm {H}_{x,h}\cap E_U$
. Regarding rationality properties of adelic rigid meromorphic cocycles and their values at adelic special pairs, we are guided by the following two principles:
-
○ The set
$G(\mathbb {Q})\backslash \Sigma _U^{\mathrm {or}}$ behaves like the collection of special points on a Shimura variety with its attendant canonical model. The Shimura reciprocity law for classical Shimura varieties then makes the following plausible: The image of the adelic special pair
$(\vec {x},h)$ in
$G(\mathbb {Q})\backslash \Sigma _U^{\mathrm {or}}$ is defined over the abelian extension
$\mathrm {H}_{x,h} / E_x$ and
$\tau \in \operatorname {\mathrm {Gal}}(\mathrm {H}_{x,h} / E_x)$ acts via
$\mathrm {rec}^{-1}(\tau )$ on it.
-
○ Rigid meromorphic cocycles J behave as though they are meromorphic functions with divisors defined over
$E_U$ on a ‘Shimura variety’
$Y_U$ whose collection of special points coincides with
$G(\mathbb {Q})\backslash \Sigma _U^{\mathrm {or}}$ .
These principles, in the settings of [Reference Darmon and Vonk18], [Reference Gehrmann22] and [Reference Guitart, Masdeu and Xarles26] involving quadratic spaces of signatures
$(2,1)$
and
$(2,2)$
, lead to predictions for which a reasonable amount of experimental and theoretical evidence has been gathered. Extrapolating it to orthogonal groups of general ranks and signatures, one is led to the following analogue of the Shimura reciprocity law for special points on orthogonal Shimura varieties:
Conjecture 4.38. Let
$h\in G(\mathbb {A})/U$
be a coset. For every rigid meromorphic cocycle
$J\in \mathcal {RMC}(\Gamma _h)$
of level
$\Gamma _h$
, there exists a finitely generated submodule
$\Pi _J\subseteq \mathbb {Q}_p^{\times }$
of rank less or equal to the rank of
$\mathrm {H}^s(\Gamma _h,\mathbb {Z})$
such that

for every oriented special pair
$(\vec {x},h)$
at which J is regular. Moreover, for every
$\alpha \in \Pi _J$
such that
$J[(\vec {x},h)]/\alpha \in \mathrm {H}_{x,h} E_U$
, there exists
$\alpha ^{\prime }\in \Pi _J$
such that

5 Examples and numerical experiments
This closing chapter reports on a few numerical experiments illustrating Conjecture 4.38. These experiments only scratch the surface, and it would be worthwhile to extend their scope by developing systematic, practical approaches for computing rigid meromorphic cocycles and their special values.
Note that the case
$s=0$
of definite quadratic spaces represents the most tractable setting for Conjecture 4.38. Because
$G(\mathbb {R})$
is compact, the p-arithmetic group
$\Gamma $
acts discretely on
$X_p$
, and a rigid meromorphic cocycle is just a
$\Gamma $
-invariant rigid meromorphic function on
$X_p$
whose divisor is supported on a finite union of
$\Gamma $
-orbits of rational quadratic divisors. In ranks
$3$
and
$4$
, the theory of p-adic uniformization of Shimura curves and quaternionic Shimura surfaces places Conjecture 4.38 within the purview of the classical theory of complex multiplication by reducing it to the study of CM points on certain orthogonal Shimura varieties. (The articles [Reference Giampietro and Darmon23] and [Reference Daas15] exploit the p-adic uniformization of Shimura curves to obtain nontrivial results on singular moduli in the case of quadratic spaces of signature
$(3,0)$
and
$(4,0)$
. Forthcoming work by the second and third authors will give a detailed discussion of the more general case.) This setting has already been explored from a practical, computational angle in the literature, notably in [Reference Greenberg25], [Reference Negrini38] and [Reference Giampietro and Darmon23] for the special case of signature
$(3,0)$
, and upcoming work of Negrini and Hassan describes similar calculations in signature
$(4,0)$
. The remainder of this chapter will therefore focus on the hyperbolic case where
$s=1$
and rigid meromorphic cocycles are one-cocycles on p-arithmetic groups.
5.1 Signature
$(2,1)$
and class fields of real quadratic fields
A prototypical example of a three-dimensional quadratic space can be obtained by letting B be a quaternion algebra over
$\mathbb {Q}$
, and setting

where
$\mathrm {Norm}$
denotes the reduced norm. For example, when
$V=\mathrm {M}_2(\mathbb {Q})$
, equipped with the negative of the trace form, this quadratic space is the direct sum of a hyperbolic plane and the form
$ x^2$
, and is of signature
$(2,1)$
. More generally, V is of signature
$(2,1)$
whenever
$\mathrm {M}_2(\mathbb {Q})$
is replaced by an indefinite quaternion algebra over
$\mathbb {Q}$
in this construction.
The exact sequence (2) with
$k=\mathbb {Q}$
and
$G=B^{\times }/\mathbb {G}_m$
then becomes

where the penultimate map is induced from the natural inclusion of
$B_1^{\times }$
into
$B^{\times }$
, and the rightmost map is induced from the norm modulo squares. In particular, one then has

Rigid meromorphic cocycles attached to quadratic spaces of signature
$(2,1)$
have already been considered in the literature, notably in [Reference Darmon and Vonk18] where
$V $
is the space of trace zero elements in
$\mathrm {M}_2(\mathbb {Z}[1/p])$
and
$\Gamma = {\mathrm {SL}}_2(\mathbb {Z}[1/p])$
, and in [Reference Guitart, Masdeu and Xarles26] and [Reference Gehrmann22], where more general quadratic spaces arising from the trace zero elements in certain indefinite quaternion algebras are considered. The constructions of loc.cit. are in one sense finer than what will be described in this section because they exploit the circumstance, peculiar to rank
$3$
, that the rational quadratic divisor
$\Delta _{v,p}$
on
$\mathcal {H}_p$
attached to a vector
$v\in B$
is of the form

where
$\tau _v$
and
$\tau _v^{\prime }$
are defined over a quadratic extension of
$\mathbb {Q}_p$
and interchanged by the Galois automorphism. This is slightly cruder than the construction of [Reference Darmon and Vonk18], which constructs divisor-valued cohomology classes whose divisor involves
$\tau _v$
but not
$\tau _v^{\prime }$
. Furthermore, the collection of
$\Gamma $
-orbits of vectors of length d is more rich and subtle in the rank
$3$
settings studied in loc. cit., since it is a principal homogeneous space for the narrow class group of (an order in)
$\mathbb {Q}(\sqrt {d})$
. A rigid meromorphic cocycle with divisor concentrated on a single orbit is in some sense ‘defined over’ the associated narrow ring class field and does not arise in general as a Borcherds lift. A similar phenomenon occurs in classical Borcherds theory, where not every meromorphic function on a modular or Shimura curve with CM divisor can be realised as a Borcherds lift, although such converse results have been proved in certain cases for quadratic spaces of higher rank (cf. [Reference Bruinier10], for example).
Let be the space of trace zero matrices endowed with the negative of the determinant quadratic form, for which
$q(v) = -\det (v)$
and

The matrix

corresponds to the quadratic form
$[a,2b,c]$
of discriminant
$4(b^2-ac)$
, whose roots,

have opposite signs precisely when
$ac<0$
. Let
$\Phi $
be the Schwartz function on the finite adelic space
$V_{{\mathbb A}_f}$
whose component
$\Phi _\ell $
at odd primes
$\ell $
is the characteristic function of the space
$M_2(\mathbb {Z}_\ell )_0$
of trace zero matrices with coefficients in
$\mathbb {Z}_\ell $
, and whose component at
$2$
is defined in terms of the odd Dirichlet character
$\chi _4$
of conductor
$4$
by setting

Let
$\Phi ^{(p)}$
be the Schwartz function on the adelic space
$V_{{\mathbb A}_f^{(p)}}$
with the component at p ignored. The stabiliser of this Schwartz function is the p-arithmetic analogue of the Hecke congruence group
$\Gamma _0(2)$
:

For each
$d>0$
, let

be the locally finite divisor on
$\mathcal {H}_p$
attached to
$\Phi ^{(p)}$
and d, where the sum is taken over the matrices v as in (47) of length
$d = b^2-ac$
satisfying
$ac<0$
, and
$\delta _v = \mathrm {sign}(a)$
. The degree of
$\Delta _d(0,\infty ) \cap \mathcal {H}_p^{\le n}$
is equal to the
$d p^{2n}$
-th Fourier coefficient
$b(dp^{2n})$
of the modular form of weight
$3/2$
and level
$4$
given by

where
$\theta (q) = \sum _{n\in \mathbb {Z}} q^{n^2}$
is the usual unary theta series and
$E_1(1,\chi _4)$
is the weight one Eisenstein series attached to
$\chi _4$
. Some of the coefficients of this generating series are listed in the table below:

When
$p=3$
, the relevant space of modular forms of weight
$3/2$
and level
$12$
is spanned by
$g(q)$
and
$g(q^3)$
. The Borcherds theory described in [Reference Darmon and Vonk19] implies that the locally finite divisors

which are of the form
$\sum c(d) \Delta _d(0,\infty )$
with

are the divisors of a rigid meromorphic period function in the sense of [Reference Darmon and Vonk18] – but on
$\Gamma $
rather than on
${\mathrm {SL}}_2(\mathbb {Z}[1/p])$
. Let

denote the rigid meromorphic cocycles with these divisors. They were calculated with
$200$
digits of
$3$
-adic accuracy and evaluated at the RM point
$\tau _{11} = \frac {1+\sqrt {11}}{2}$
of discriminant
$44$
. Since the narrow Hilbert class field of
$\mathbb {Q}(\sqrt {11})$
is the biquadratic field
$\mathbb {Q}(i, \sqrt {11})$
, Conjecture 4.38 asserts that
$J_t[\tau _{11}]$
is defined over this field, for
$t=1,2,3,4$
. Indeed, one finds that, at least to within this accuracy,

These Gaussian integers are all of norm one, and the primes that divide their denominators are all inert in
$\mathbb {Q}(\sqrt {11})$
. This experiment extends the framework of [Reference Darmon and Vonk18] ever so slightly by working with a lattice that differs from the space of all integral binary quadratic forms. (In the setting of signature
$(2,1)$
, this generalisation is pursued more systematically in Antoine Giard’s PhD thesis [Reference Giard24].) It also illustrates a feature of Conjecture 4.38, namely that the field of definition of
$J[\tau ]$
is a class field that depends only on
$\tau $
and not on J, insofar as the rigid meromorphic cocycles constructed in the orthogonal group framework are all ‘defined over
$\mathbb {Q}$
’.
5.2 Signature
$(3,1)$
and Bianchi cocycles
Let be the standard rational Minkowski space, endowed with the quadratic form

It admits a convenient description as a subspace of
$\mathrm {M}_2(K)$
, with
. The matrix ring
$\mathrm {M}_2(K)$
is endowed with the standard anti-involution
$M\mapsto M'$
as well as an involution
$M \mapsto \overline {M}$
coming from the Galois automorphism of K, defined by

and we define

equipped with the quadratic form

We will adopt the convenient notational shorthand

The group
${\mathrm {SL}}_2(K)$
operates as isometries of V by twisted conjugation,

leading to a homomorphism

The real symmetric space
$X_\infty $
is a real
$3$
-dimensional manifold, and is identified with
$\mathbb {C} \times \mathbb {R}^{>0}$
by sending the line spanned by the vector
$[\beta ;u,v]$
of negative norm to

Suppose that p is split in K – that is,
$K\otimes _{\mathbb {Q}} \mathbb {Q}_p \cong \mathbb {Q}_p^2$
. The p-adic symmetric space
$\mathcal {H}_p \times \mathcal {H}_p$
is identified with
$X_p$
via the association

For any
$M=[\alpha; b,c]\in V$
of positive norm, the divisor
$\Delta _{M,p}$
can be identified with the set of
$(\tau _1,\tau _2)$
for which

Hence, a function
$F_M\in \mathcal {M}_{\mathrm {rq}}^{\times }$
having
$\Delta _{M,p}$
as divisor can be defined by setting

Consider the standard lattice

on which the group

acts naturally by isometries, as above. Concrete instances of rigid meromorphic cocycles for the Bianchi group
$\Gamma $
can be described using modular symbols, as we now explain. The group
$\Gamma $
acts naturally on the boundary
$\mathbb {C} \cup \{\infty \} = \mathbb {P}_1(\mathbb {C})$
of
$X_\infty $
by Möbius transformations and preserves the subset
$\mathbb {P}_1(K)$
of K-rational points, whose stabilisers in
$\Gamma $
are parabolic subgroups. Given
$r,s\in \mathbb {P}_1(K)$
, denote by
$(r,s)$
the open hyperbolic geodesic path joining r and s in
$X_\infty $
.
Given an integer
$d>0$
, one can formally define a modular symbol with values in
$\mathcal {M}_{\mathrm {rq}}^{\times }$
by setting, for all
$r,s\in \mathbb {P}_1(K)$
,

where
$F_M\in \mathcal {M}_{\mathrm {rq}}^{\times }$
is the rational function on
$X_p$
having
$\Delta _{M,p}$
as divisor, and
$\Delta _{M,\infty }$
is the real two-dimensional cycle in the three-manifold
$X_\infty $
attached to the vector M. If these expressions converge for all
$r,s$
, then they clearly define a modular symbol with values in
$\mathcal {M}_{\mathrm {rq}}^{\times }$
. One also hopes that they satisfy some type of
$\Gamma $
-equivariance property, since

where

is a multiplicative factor of automorphy that depends on
$\gamma $
and
$\tau $
but not on M.
Assuming that the assignment
$(r,s) \mapsto J_d(r,s)$
is
$\Gamma $
-equivariant, the theory of modular symbols allows us to reduce the calculation of
$J_d$
to that of
$J_d(0,\infty )$
. Following a terminology of Zagier that is taken up in [Reference Darmon and Vonk18], the function

is called the rigid meromorphic period function on
$X_p$
associated to
$J_d$
(or to the integer d). The formula defining this rigid meromorphic period function is given by

where

To make it more concrete, we invoke the following lemma:
Lemma 5.1. Let
$M = [\alpha; b,c]\in V$
be a vector of positive norm with associated real two-dimensional cycle
$\Delta _{M,\infty } \subseteq X_\infty $
. Then

Proof. The two-dimensional region attached to
$M = [\alpha ,b,c]$
is the set of
$(z,t) = (\beta /v, (uv-\beta \overline {\beta })/v^2) \in \mathbb {C}\times \mathbb {R}^{>0}$
satisfying the equality

that is,

This equation can be rewritten in terms of the coordinates
$(z,t)$
, as

or equivalently, after dividing by
$-c$
,

This equation can be further rewritten as

This two-dimensional region intersects the boundary of
$ \mathbb {C} \times \mathbb {R}^{>0}$
(with equation
$t=0$
in the circle on the complex z-plane centred at
$\overline {\alpha }/c$
and of square-radius
$(\alpha \overline {\alpha }-bc)/c^2$
). Since this circle contains
$z=0$
in its interior precisely when
$bc<0$
, the lemma follows.
Thanks to this lemma, equation (48) can be written as

There are two difficulties that arise in this definition:
-
1. The term
$\mathrm {sgn}(c)$ is problematic to make sense of when
$c=0$ , which occurs when the two-cycle in
$X_\infty $ attached to
$[\alpha ;b,c]$ intersects
$(0,\infty )$ improperly.
-
2. Even if such improper intersections do not occur, one has
$\Phi _{d,j}^\circ (\tau _1,\tau _2) = \pm 1$ for all j, because the symmetry
$[\alpha; b,c]\mapsto -[\alpha ;b,c]$ forces trivial cancellation in the product.
To remedy the first problem, it can simply be assumed that d is not a norm from
$\mathbb {Z}[i]$
, that is, is not a sum of two squares, to ensure that there are no length d vectors
$[\alpha ;b,c]$
in the lattice with
$bc=0$
. This holds, for instance, if d is exactly divisible by a prime which is congruent to
$3$
modulo
$4$
.
To address the second problem, the sum over a full lattice needs to be replaced by a sum over a coset of one lattice in another (or a linear combination of such sums) that is not preserved by multiplication by
$-1$
. For instance, let
$\chi _4$
be the odd Dirichlet character of conductor
$4$
and weight the factor attached to
$[\alpha ;b,c]$
in the product defining
$\Phi _{d,j}^\circ $
by a further exponent of
$\chi _4(c)$
, by setting

This further weighting by an odd character causes the factors attached to
$[\alpha; b,c]$
and
$[-\alpha; -b, -c]$
to reinforce each other rather than cancelling out. The following suggestive formula relates the degree of the function
$\Phi _{d,j}$
to to the Fourier coefficients of modular forms.
Lemma 5.2. For all
$j\geq 1$
and all d which are not a sum of two squares, the rational function
$\Phi _{d,j}$
has degree equal to

Proof. By (49),

The inner sum is readily recognised as the
$dp^{2j}-(r^2+s^2)$
-th coefficient of the weight one Eisenstein series attached to
$\chi _4$
:

and hence, we can write

where
$r_2(m)$
is the number of ways of representing m as a sum of two squares, which is also the m-th Fourier coefficient of the binary theta series attached to the quadratic form
$x^2+y^2$
. This theta series is also equal to a multiple of
$E_1(1,\chi _4)$
, a very simple instance of the Siegel-Weil formula, and thus, one finds

where
$b(n)$
is the n-th Fourier coefficient of
$ 2 E_1(1,\chi _4)^2$
. This modular form of weight two, level
$4$
and trivial nebentypus character is the quaternary theta series whose Fourier coefficients give the number of ways of representing an integer as a sum of
$4$
squares. It is also the Eisenstein series of weight two and level
$4$
with q-expansion given by

Hence, the equation

holds, as was to be shown.
Since
$\sigma (3) =4 $
and
$\sigma (7) = 8$
, we may let
and conclude that the rational functions

are of weight
$0$
for all j. This weight zero condition is not sufficient in and of itself to ensure the convergence of the infinite product
$\prod _j \Phi _{{\mathscr {D}},j}$
as a rigid meromorphic function. Numerical calculations indeed reveal that this product does not converge, even for
$p=5$
, the smallest prime that is split in
$\mathbb {Q}(i)$
. The Borcherds theory developed in the body of this paper suggests that the obstruction to this product converging as a rigid meromorphic function on the
$5$
-adic analytic space
$X_5$
lies in the space
$M_2(20)$
of modular forms of weight
$2$
and level
$20$
. In addition to the weight two Eisenstein series, this space contains a unique normalised cusp form

The fact that
$2a_3(g) - a_7(g) =-6 \ne 0$
explains why the product of the
$\Phi _{{\mathscr {D}},j}$
fails to converge for this choice of
${\mathscr {D}}$
. Inspection reveals that the functional

vanishes identically on the cusp form g, and on the weight two Eisenstein series, since the coefficients
$b_n$
of (50) satisfy

This motivates setting

and studying the rational functions

as well as the infinite product

The p-adic Borcherds theory of Theorem 3.25 predicts that this product should converge uniformly to a rigid meromorphic function on all affinoid subsets of
$\mathcal {H}_p\times \mathcal {H}_p$
.
The function
$ [ \alpha; b,c] \mapsto \chi _4(c)$
is invariant under the action of
$\Gamma $
, and hence, there is a unique
$\mathcal {M}_{\mathrm {rq}}^{\times }$
-valued modular symbol
$J_{\mathscr {D}}$
satisfying

This rigid meromorphic cocycle (or rather, its associated rigid meromorphic period function
$\Phi _{{\mathscr {D}}}$
) was calculated on the computer to an accuracy of
$160$
significant
$5$
-adic digits, using the same kind of iterative algorithms that are described in [Reference Darmon and Vonk18] in signature
$(2,1)$
. The calculation took around a week on a standard machine. The greater time complexity of the calculation compared to the
$O(2,1)$
setting lies in the fact that the
$5$
-adic rigid meromorphic period function
$\Phi _{{\mathscr {D}}}$
is eventually stored as a
$6\times 6$
array (rather than a vector of length
$6$
) of power series in two variables (rather than one) of degree
$\leq 160$
with coefficients in
$\mathbb {Z}/5^{160}\mathbb {Z}$
. Once it has been computed and recorded in a file, evaluating the associated rigid meromorphic cocycle
$J_{\mathscr {D}}$
at a special point typically requires only a few seconds.
Special points on
$X_p= \mathcal {H}_p\times \mathcal {H}_p$
can be divided into three types:
-
(a) Small RM points whose reflex field is a real quadratic field. Examples of small RM points are given by tuples
$(\tau ,\tau )$ , where
$\tau $ is an RM point in
$\mathcal {H}_p$ . The reflex field of such a points is the real quadratic field
$\mathbb {Q}(\tau )$ .
-
(b) Small CM points whose reflex field is an imaginary quadratic field. Examples of small CM points are given by tuples
$(\tau ,\tau ')$ , where
$\tau $ is an RM point in
$\mathcal {H}_p$ and
$\tau '$ is its Galois conjugate. The reflex field of such a point is the imaginary quadratic subfield of the biquadratic field
$\mathbb {Q}(\tau ,i)$ which differs from
$\mathbb {Q}(i)$ .
-
(c) Big special points whose reflex fields are ATR extensions of real quadratic fields – that is, quartic extensions of
$\mathbb {Q}$ with exactly one complex place.
We discuss the three classes in turn.
Evaluations at small
$RM$
points. Adopt the shorthand

The following identities give a good illustration of Conjecture 4.38:



with

The algebraic numbers on the right-hand side of these putative evaluations all belong to an unramified abelian extension of the reflex field
$\mathbb {Q}(\tau )$
and have norm
$1$
to this reflex field. More precisely, the Shimura reciprocity law in this case predicts that they are in the minus part (multiplicatively) for the action of complex conjugation which lies in the center of the absolute Galois group of
$\mathbb {Q}(\tau )$
. In particular, they lie in the unit circle relative to all complex embeddings of
$\overline {\mathbb {Q}}$
. It is also worth noting that the primes that arise in the factorisation of
$J_{{\mathscr {D}}}[\tau ]$
lie above rational primes that are either inert or ramified in the reflex field. These observations evoke patterns that were already observed in [Reference Darmon and Vonk18], and this is no coincidence. Indeed, the points
$(\tau ,\tau )$
are the images of the RM point
$\tau \in \mathcal {H}_5$
under the diagonal embedding
$\mathcal {H}_5\subseteq \mathcal {H}_5\times \mathcal {H}_5$
, and the restriction of
$J_{{\mathscr {D}}}$
to the diagonal gives rise to a rigid meromorphic cocycle for a subgroup of
${\mathrm {SL}}_2(\mathbb {Z}[1/p])$
, namely, the intersection of
$\Gamma $
with
${\mathrm {SL}}_2(\mathbb {Q})$
– that is, the
$\Gamma _0(2)$
-type congruence subgroup of
${\mathrm {SL}}_2(\mathbb {Z}[1/p])$
considered in Section 5.1. Hence, the evaluations (51), (52) and (53) do not represent a truly new verification of Conjecture 4.38 for a Bianchi cocycle, since they can be reduced to the conjectures of [Reference Darmon and Vonk18].
Evaluations at small
$CM$
points. Values at small CM points of the form
$(\tau ,\tau ')$
where
$\tau $
is an RM point in
$ \mathcal {H}_5$
– for which the p-adic ‘CM type’ has been flipped, so that the associated reflex field is now an imaginary quadratic field – have no counterpart in the setting of rigid meromorphic cocycles for
$O(2,1)$
, and are thus significantly more interesting a priori. The first numerical evaluations reveal that



Comparing with (51), (52) and (53), it seems that these more exotic values tend to be of significantly smaller height. This is confirmed by further experiments. Letting
$\tau _D\in \mathcal {H}_5$
be an RM point of discriminant
$D>0$
, it was observed that

Another novel and still poorly understood phenomenon is the relatively frequent occurrence of instances where
$J_{{\mathscr {D}}}[\tau _D,\tau _D^{\prime }]$
cannot be evaluated because the point
$(\tau _D,\tau _D^{\prime })$
occurs among the poles and zeroes of the rational functions in the infinite product defining
$J_{{\mathscr {D}}}\{r,s\}$
with
$r,s\in {\mathbb P}_1(\mathbb {Q}(i))$
. This happens for

It is unclear whether these zeroes and poles are unavoidable or are an artefact of the modular symbol algorithm and might have been avoided by a more careful choice of a unimodular path in this algorithm.
By contrast, the diagonal restriction of
$J_{{\mathscr {D}}}\{r,s\}$
has zeroes and poles on a collection of RM points with finitely many discriminants, and it is expected that
$J_{{\mathscr {D}}}[\tau ,\tau ]=1$
occurs only in extremely rare instances as well. Despite their relative paucity, interesting algebraic values at small CM points do arise. For instance, numerical evaluations of

to
$160$
digits of
$5$
-adic precision suggest that


Unlike what happens for the small
$RM$
values, these algebraic invariants have nontrivial norms to
$\mathbb {Q}$
:


Observe that the primes that appear in the right-hand sides of (59) and (60) are all ramified or inert in the respective reflex fields
$\mathbb {Q}(\sqrt {-17})$
and
$\mathbb {Q}(\sqrt {-37})$
, and are relatively small compared to the height of the special value. This suggests that these special values admit factorisations analogous to those established by Gross and Zagier for differences of singular moduli. The numerical evidence, however limited, strongly supports this conclusion, but we have not been able to formulate a very precise statement along these lines.
The numerical evaluations in (56), (57) and (58) provide the first substantial piece of experimental evidence for Conjecture 4.38 in the case of Bianchi cocycles, beyond what directly follows from this conjecture in the signature
$(2,1)$
case.
Evaluations at big special points
Nontrivial values of
$J_{{\mathscr {D}}}$
at big special points are expected to be defined over fields of fairly large degree (abelian extensions of quadratic extensions of real quadratic fields with a single complex place) and to be of sizeable height. A straightforward recognition of these values seems to lie somewhat beyond the
$5$
-adic precision for
$J_{{\mathscr {D}}}$
that the authors were able to calculate. Recently, Xevi Guitart, Marc Masdeu and the second author have implemented a more systematic and sophisticated approach to the computer calculation of rigid meromorphic Bianchi cocycles and succeeded in convincingly identifying certain values at big special points as algebraic numbers. They will report on their successful experiments in future work.
Acknowledgements
The authors are grateful to the Centre de Recherches Mathématiques in Montreal, which continued to host in-person activities during the 2020 thematic semester on number theory, at the height of the COVID-19 pandemic. Notable among these was a ‘p-adic Kudla seminar’ whose participants provided a congenial and stimulating atmosphere for the elaboration of this article. We would like to thank Tonghai Yang, Sören Sprehe and the anonymous referee for their comments and corrections, which greatly improved the exposition of the paper.
Competing interest
The authors have no competing interests to declare.
Financial support
The work on this article began while Lennart Gehrmann was visiting McGill University, supported by Deutsche Forschungsgemeinschaft, and he thanks both of these institutions. More recently, Lennart Gehrmann has received funding from the Maria Zambrano Grant for the attraction of international talent in Spain and from Deutsche Forschungsgemeinschaft via the grant SFB-TRR 358/1 2023 – 491392403. Henri Darmon was supported by NSERC Discovery grant RGPIN-2018-04062. Mike Lipnowski was supported by NSERC Discovery grant RGPIN-2018-04784 and NSF CAREER grant 2338933.