1 Introduction
Just over half a century ago, Doi and Naganuma discovered a Hecke-equivariant lifting map from to weight k elliptic modular forms to weight
$(k, k)$
Hilbert modular forms for a real quadratic field F [Reference Doi and NaganumaDN70]. This is a special case of cyclic base change [Reference Jacquet and LanglandsJL70], which has now become a basic and useful tool in the theory of automorphic forms and automorphic representations. By the exceptional isogeny

the Doi-Naganuma lifting is also an instance of a theta lifting from
$\operatorname {SL}_2$
to
$\mathrm {O}(2, 2)$
[Reference KudlaKud78].
1.1 A problem posed by Gross and Zagier
In the seminal paper [Reference Gross and ZagierGZ86], Gross and Zagier proved their formula relating the central derivative of some Rankin-Selberg L-function attached to a weight 2 level N newform f and the Néron-Tate height pairing of f-isopytic components of Heegner points in the Jacobian of the modular curve
$X_0(N)$
. This was extended in [Reference Gross, Kohnen and ZagierGKZ87] to describe the positions of these Heegner points in the Jacobian using Fourier coefficients of modular forms. In the degenerate case
$N = 1$
, the Gross-Zagier formula yields a beautiful factorization formula of the norm of differences of singular moduli [Reference Gross and ZagierGZ85].
To calculate the archimedean contribution to the height pairings, one requires the automorphic Green function

on
$X_0(N) \times X_0(N)$
. It is an eigenfunction with respect to the Laplacians in
$z_1$
and
$z_2$
with eigenvalue
$s(1-s)$
. The function vanishes when one of the
$z_i$
approaches the cusps and has a logarithmic singularity along the diagonal. In fact, these properties characterize it uniquely. Using Hecke operators acting on either
$z_1$
or
$z_2$
, we can define

where
$R_N := \{\left (\begin {smallmatrix} a & b\\ Nc & d \end {smallmatrix}\right ): a, b, c, d \in \mathbb {Z}\}$
. Then
$G^{\Gamma _0(N), m}_s$
has a logarithmic singularity along the m-th Hecke correspondence
$T_m \subset X_0(N)^2$
(see (1.2) in Chapter II of [Reference Gross and ZagierGZ86]).
For integral parameters
$s = r + 1 \in \mathbb {N}_{\ge 2}$
, these functions are called higher Green functions. In Section V.1 of [Reference Gross, Kohnen and ZagierGKZ87], two problems about these functions were raised. The first one was to give an interpretation of their values at Heegner points as archimedean contributions of certain higher weight height pairings. This was answered by Zhang in [Reference ZhangZha97] (see also [Reference XueXue10]), where the Néron-Tate height pairing of Heegner points is replaced by the arithmetic intersection of Heegner cycles on Kuga-Sato varieties.
The second problem dealt with the algebraicity of higher Green functions at a single CM point. Let
$M^{!, \infty }_{-2r}(\Gamma _0(N))$
be the space of weakly holomorphic modular forms for
$\Gamma _0(N)$
of weight
$-2r < 0$
with poles only at the cusp infinity (see (2.27)). Given
$f = \sum _{m \gg -\infty } c(m) q^m \in M^{!, \infty }_{-2r}(\Gamma _0(N))$
, we call the following linear combination of higher Green functions

the principal higher Green function associated to f. Along the divisor

the function
$G_{r+1, f}^{\Gamma _0(N)}$
has a logarithmic singularity. Using Serre duality, this function is the same as the higher Green function defined via relations in Section V.4 of [Reference Gross and ZagierGZ86] (see Remark 2.5). We say that it is rational when f has rational Fourier coefficients at the cusp infinity. Even though the theory of complex multiplication does not directly apply as in the case of automorphic Green functions, the value of a rational, principal higher Green function
$G^{\Gamma _0(N)}_{r+1, f}$
at a single CM point on
$X_0(N)\times X_0(N)$
should be algebraic in nature, predicted by the following conjecture (see, for example, [Reference MellitMel08] and [Reference ViazovskaVia11]).
Conjecture 1.1. Suppose
$f \in M_{-2r}^{!, \infty }(\Gamma _0(N))$
has rational Fourier coefficients at the cusp infinity. Then for any CM point
$(z_1, z_2) \in X_0(N)^2 \backslash Z_f$
with
$z_j$
having discriminant
$d_j < 0$
, there exists
$\alpha = \alpha (z_1, z_2) \in \overline {\mathbb {Q}} \subset \mathbb {C}$
such that

Over the years, there have been a lot of partial results toward this conjecture. When
$d_1d_2$
is a perfect square, this conjecture was proved in [Reference ZhangZha97] conditional on the nondegeneracy of the height pairing of CM cycles. Using regularized theta liftings, an analytic proof was given in [Reference ViazovskaVia11] with restrictions on
$N, d_j$
and later in full generality in [Reference Bruinier, Ehlen and YangBEY21]. When
$d_1d_2$
is not a perfect square, less was known before. For
$N = 1, z_1 = i$
and
$r = 1$
, Mellit proved the conjecture in his thesis [Reference MellitMel08] using an algebraic approach. When one averages over the full Galois orbit of the CM point
$(z_1, z_2)$
, the conjecture follows from [Reference Gross, Kohnen and ZagierGKZ87] for r even. More partial results are available when one averages over different Galois orbits [Reference LiLi22, Reference Bruinier, Ehlen and YangBEY21] when
$N = 1$
.
Motivated by Conjecture 1.1, the first and third author, together with S. Ehlen, considered its generalization to the setting of orthogonal Shimura varieties in [Reference Bruinier, Ehlen and YangBEY21]. More precisely, let
$\mathrm {V}$
be a rational quadratic space of signature
$(n, 2)$
with
$n \ge 1$
, and
$X_K$
be the Shimura variety associated to
${\tilde {H}}_{\mathrm {V}} := \operatorname {GSpin}(\mathrm {V})$
and a compact open subgroup
$K \subset {\tilde {H}}_{\mathrm {V}}(\hat {\mathbb {Q}})$
. For a nonnegative integer r and a vector-valued harmonic Maass form f of weight
$1- n/2 -2r$
, denote by
$\Phi _{f}^r$
its regularized theta lift (see [Reference BruinierBru02] or equation (2.41)). This function is an eigenfunction of the Laplacian on
$X_K$
and has a logarithmic singularity along the special divisor
$Z_f$
associated to f (see (2.42)). We call it a higher Green function on
$X_K$
and say that it is principal, resp. rational, if f is weakly holomorphic, resp. has rational principal part Fourier coefficients. When
$\mathrm {V} = M_2(\mathbb {Q})$
and
$X_K = X_0(N)^2$
, the function
$\Phi _{f}^r$
becomes
$G^{\Gamma _0(N)}_{r + 1, f}$
(see Corollary 2.4).
For a totally real field F of degree d and an F-quadratic space
$W = E$
with
$E/F$
a quadratic CM extension, suppose there is an isometric embedding

which in particular implies that
$n+2 \ge 2d$
. Then we obtain a CM cycle
$Z(W)$
on
$X_K$
from a torus
$T_W$
in
${\tilde {H}}_{\mathrm {V}}$
(see section 2.4 for details). Note that
$Z(W)$
is defined over F and is the big CM cycle
$Z(W, z_0^\pm )$
in [Reference Bruinier, Kudla and YangBKY12]. We denote by
$Z(W)_{\mathbb {Q}}$
the union of the F-conjugates of
$Z(W)$
. If F is quadratic, we write
$Z(W)_{\mathbb {Q}} = Z(W) \cup Z(W)'$
.
In [Reference LiLi23], the second author studied the algebraicity of the difference of a rational, principal
$\Phi ^r_f$
at two CM points in
$Z(W)$
and was able to verify the analogue of Conjecture 1.1 in that setting. This opens up the possibility of proving Conjecture 1.1 when one proves an algebraicity result for the averaged value
$\Phi ^r_f(Z(W))$
. In this paper, we complete this step by proving the following result complementary to [Reference LiLi23].
Theorem 1.2 (Algebraicity and factorization).
Let
$\Phi _{f}^r$
be a rational, principal higher Green function on
$X_K$
. Suppose that
$E/\mathbb {Q}$
is a biquadratic CM number field with the real quadratic subfield
$F = \mathbb {Q}(\sqrt {D})$
, and
$Z(W) \cap Z_f = \emptyset $
. Then there exists a positive integer
$\kappa $
and
$a_1, a_2 \in F^\times $
such that

For any prime
$\mathfrak {p}$
of F, the value
$ \kappa ^{-1} \operatorname {ord}_{\mathfrak {p}}(a_j)$
is given in (5.6). When
$n = 2$
, we have
$a_j = 1$
for
$j \equiv r \bmod {2}$
.
Remark 1.3. The denominator
$\kappa $
appears as a consequence of our matching of sections (see Propositions 4.7 and 4.11) and only depends on
$Z(W)$
and r when f has integral Fourier coefficients.
Remark 1.4. Theorem 1.2 also applies to the case
$r = 0$
when f has zero constant term, in which case
$\Phi _f^0 = \Phi _f$
is the regularized Borcherds lift of f and we have
$a_2 = 1$
.
Combining Theorem 1.2 with the main result in [Reference LiLi23], we deduce the algebraicity of a rational, principal higher Green function at a single CM point when
$E/\mathbb {Q}$
is biquadratic – hence, Conjecture 1.1 in particular.
Theorem 1.5. In the setting of Theorem 1.2, there exists
$\kappa \in \mathbb {N}$
and Galois equivariant maps
$\alpha _1, \alpha _2: T_W(\hat {\mathbb {Q}}) \to E^{\mathrm {ab}}$
such that

for all
$[z_0, h] \in Z(W_{})$
. Furthermore, for
$n =2$
, we can choose
$\alpha _j(h) = 1$
for
$j \equiv r \bmod {2}$
; that is, there exists a Galois-equivariant map
$\alpha : T_W(\hat {\mathbb {Q}}) \to E^{\mathrm {ab}}$
such that

for all
$h \in T_W(\hat {\mathbb {Q}})$
. In particular, Conjecture 1.1 is true.
1.2 Comparison to previous works
There has been an extensive literature on the CM-value of regularized theta lifts. When
$r = 0, n = 2$
and f is weakly holomorphic, the CM-value
$\Phi _f(Z(W)_{\mathbb {Q}})$
was the subject of the classical work of Gross-Zagier on singular moduli [Reference Gross and ZagierGZ85] and generalizations by the first and third author [Reference Bruinier and YangBY06]. More generally for arbitrary n, totally real field F and harmonic Maass form f, the value
$\Phi _f(Z(W)_{\mathbb {Q}})$
is the archimedean contribution of the derivative of a Rankin-Selberg L-function involving the shadow
$\xi (f)$
at
$s = 0$
[Reference Bruinier and YangBY09, Reference Bruinier, Kudla and YangBKY12, Reference Andreatta, Goren, Howard and PeraAGHMP18].
A crucial ingredient in these works is a real-analytic Hilbert Eisenstein series
$E^{*}$
of parallel weight
$1$
over F. It is an incoherent Eisenstein series in the sense of the Kudla program [Reference KudlaKud97]. The arithmetic Siegel-Weil formula predicts that the Fourier coefficients of its derivative,
$E^{*, \prime }$
, are arithmetic degrees of special cycles [Reference Howard and YangHY11, Reference Howard and YangHY12], which are logarithms of rational numbers.
Using suitable weight 1 harmonic Maass forms in place of incoherent Eisenstein series, the first and third author, together with S. Ehlen, could prove the algebraicity result for higher Green function at a partially averaged CM cycle and deduce the Gross-Zagier conjecture for
$X_K = X_0(1)^2$
when the class group of one of the imaginary quadratic fields in E is an elementary 2 group [Reference Bruinier, Ehlen and YangBEY21, Theorem 1.2]. However, the factorization of the ideal generated by the algebraic numbers is not explicitly given.
Our main result in Theorem 1.2 goes far beyond these aforementioned works in an essential way by studying the regularized theta lifts at the partially averaged CM cycle
$Z(W)$
, which is in general only half of
$Z(W)_{\mathbb {Q}}$
and a priori defined over F. For
$r = 0$
and f weakly holomorphic, this means that
$\Phi _f(Z(W))$
is the logarithm of a number in the real quadratic field F and therefore cannot be related to the Fourier coefficients of incoherent Eisenstein series!
Furthermore, this partial average is quite different, yet more natural, than the one studied in [Reference Bruinier, Ehlen and YangBEY21]. Instead of using the weight 1 harmonic Maass form loc. cit., which is an elliptic modular form, we explicitly construct a Hilbert modular form
$\tilde {\mathcal {I}}$
, serving as a companion and substitute for the incoherent Eisenstein series, and obtain precise information concerning its Fourier coefficients. This is the main innovation of the paper and allows us to prove the exact factorization formula for the ideal generated by the algebraic numbers in the spirit of [Reference Gross and ZagierGZ85], which was not possible in [Reference Bruinier, Ehlen and YangBEY21]. Most importantly, we are able to achieve this for arbitrary open compact subgroup K, just as in [Reference LiLi23] for the difference of two CM-values, whereas the ingredients in [Reference Bruinier, Ehlen and YangBEY21] could only handle the level 1 case. This enables us to prove Theorem 1.5 for arbitrary level K, which encompasses the case in Conjecture 1.1. In that sense, this paper is the complement to [Reference LiLi23], both in results and methods, for biquadratic E.
Besides the analytic approach to Conjecture 1.1, which originated from the work of Viazovska for
$F = \mathbb {Q} \oplus \mathbb {Q}$
[Reference ViazovskaVia11], there is also an algebraic approach in [Reference ZhangZha97, Reference MellitMel08]. However, one must overcome serious obstacles to prove Theorem 1.5 via this approach. For
$F = \mathbb {Q} \oplus \mathbb {Q}$
, one needs to assume in an essential way the nondegeneracy of the restriction of the Gillet-Soulé height pairing, which is defined on Kuga-Sato varieties, to the subgroup of the Chow group spanned by CM cycles [Reference ZhangZha97, Theorem 5.2.2]. The nondegeneracy of this height pairing on a slightly larger subgroup is conjectured by Beilinson [Reference BeilinsonBei87] and Bloch [Reference BlochBlo84] (see Conjecture 1.3.1 in [Reference ZhangZha97]). For real quadratic F, one needs to find a substitute for the Kuga-Sato variety, construct canonical models, and define suitable cycles and arithmetic intersections such that the archimedean contribution is given by the CM-values of higher Green functions.Footnote
1
Assuming that the conjecture of Beilinson and Bloch holds in this case, one can then deduce the result in Theorem 1.5. For
$n = 2$
and concrete families of CM points, it is possible verify Conjecture 1.1 by explicit constructions of cycles and calculations (see [Reference MellitMel08]). In general, it is not clear at all how to construct suitable cycles, not to mention remove the nondegeneracy assumption. However, it would be very interesting to see if Theorem 1.5, which is proved via the analytic approach, can be used to prove the conjectural nondegeneracy when restricted to the above subgroup of the Chow group in [Reference ZhangZha97].
1.3 Proof strategy
For simplicity, we focus on the case
$n = 2$
, from which the general case is not hard to derive (see Section 5 for details). Applying the strategy in [Reference KudlaKud03] and the Rankin-Cohen operator, one can express
$\Phi ^r_f(Z(W)) + (-1)^r \Phi ^r_f(Z(W)')$
as an F-linear combination of Fourier coefficients of the holomorphic part of
$E^{*, \prime }$
, which are logarithms of rational numbers. This is a standard procedure involving the Siegel-Weil formula and Stokes’ Theorem (see, for example, the proof of Theorem 3.5 in [Reference LiLi21]). A crucial property of the incoherent Eisenstein series is the following differential equation: (see [Reference Bruinier, Kudla and YangBKY12, Lemma 4.3])

Here,
$L_j$
are lowering operators in the j-th variable, and
$\Phi ^{(\epsilon _1, \epsilon _2)} = \Phi _f \otimes \Phi ^{(\epsilon _1, \epsilon _2)}_\infty $
are Siegel-Weil sections in the degenerate principal series
$I(0, \chi )$
with
$\chi =\chi _{E/F}$
being the quadratic Hecke character of F associated to
$E/F$
(see Section 2.6 for details). In particular,
$E^*(g, 0, \Phi ^{(\epsilon , -\epsilon )})$
is a coherent Eisenstein series of weight
$(\epsilon , -\epsilon )$
for
$\epsilon = \pm 1$
. To prove Theorem 1.2, it suffices to understand
$\Phi ^r_f(Z(W)) - (-1)^r \Phi ^r_f(Z(W)')$
, which means we need a substitute of
$E^{*, \prime }$
on the left-hand side of (1.7) such that the right-hand side is
$E^*(g, 0, \Phi ^{(1, -1)}) - E^*(g, 0, \Phi ^{(-1, 1)})$
.
To obtain this minus sign, we apply the exceptional isogeny in (1.1) and view the coherent Eisenstein series as modular forms on the group
$H_0 := {\mathrm {SO}}(V_0)$
for the quadratic space
$V_0$
of signature
$(2, 2)$
defined in Section 3.1. Since
$E/\mathbb {Q}$
is biquadratic, there is an odd character
$\varrho = \varrho _f \cdot {\mathrm {sgn}}$
of
$[F^1]=F^1 \backslash \mathbb {A}_F^1$
such that
$\chi = \varrho \circ {\mathrm {Nm}}^-$
(see Remark 2.1). By viewing
$\varrho $
as an automorphic form on
$H_1 := {\mathrm {SO}}(V_1)$
for the quadratic space
$V_1 = (F, {\mathrm {Nm}})$
, we can consider its theta lift following the diagram

where
$G = \operatorname {SL}_2$
. The first map lifts
$\varrho $
to a weight one holomorphic cusp form
$\vartheta (g', \varphi ^-_1, \varrho )$
on G, which was first studied by Hecke. Here,
$\varphi _1^\pm $
is a Schwartz function on
$V_1(\mathbb {A})$
whose archimedean component
$\varphi ^\pm _{\infty }$
is the Schwartz function in
$V_1(\mathbb {R})$
defined in (2.65). Then the second map lifts it to a coherent Eisenstein series and is an instance of the Rallis tower property ([Reference RallisRal84]).Footnote
2
From this,
$\theta _0 \circ \theta _1$
gives us the equation

where
$\theta _0$
is a theta kernel for the quadratic space
$V_0$
,
$\varphi ^{(\pm 1, \mp 1)} = \varphi _0^{(\pm 1, \mp 1)} \otimes \varphi _1^-$
is a Schwartz function on
$V(\mathbb {A})$
with
$V := V_0 \oplus V_1$
and
$\Phi ^{(\pm 1, \mp 1)} = F_{\varphi , \varrho }$
is the section defined in (3.37). Our first main result is Theorem 3.3, which ensures that all coherent Eisenstein series can be realized as such lifts. This is reduced to the corresponding local problem and solved in Section 3.5.
To construct
$\tilde {\mathcal {I}}$
, we first modify the character
$\varrho $
to the function
$\tilde {\varrho }_C$
on
$H_1(\mathbb {A})$
defined in (2.76). It is a preimage of
$\varrho $
under the first order invariant differential operator
$t \frac {d}{dt}$
on
$H_1(\mathbb {R}) \cong \mathbb {R}^\times $
, and hence not a classical automorphic form on
$H_1$
. We call its lift
$\vartheta (g', \varphi ^+_1, \tilde {\varrho }_C)$
to G a deformed theta integral, since the archimedean component of
$\tilde {\varrho }_C$
is essentially
$\log t$
and comes from the first term in the Laurent expansion of
$t^s$
at
$s = 0$
. This deformed theta integral was first studied in [Reference Charollois and LiCL20]. It satisfies the following important property (see Theorem 2.7):

Here, L is the lowering operator on G, and
$\mathrm {error}$
is the special value of the theta kernel on
$V_1$
.
We now define
$\tilde {\mathcal {I}}(g) := \mathcal {I}(g, \varphi ^{(1, 1)}, \tilde {\varrho }_C)$
in (4.2) using the theta kernel
$\theta _0(g', g, \varphi _0^{(1, 1)})$
with the archimedean component of
$\varphi ^{(1, 1)}_0$
being the Schwartz function
$\varphi ^{(1, 1)}_{0, \infty }$
defined in (4.1) (the integral is similar to (1.9)). A key observation is that there is an identity between the actions of the universal enveloping algebras of
$H_0$
and G, which gives in this special case (see (4.7) and Lemma 4.1)

where
$L_1, L_2$
, resp. L, are differential operators for the variable
$g \in H_0$
, resp.
$g' \in G$
. Putting these together, we see that
$\tilde {\mathcal {I}}$
satisfies the following property (see the proof of Proposition 4.2 with
$r = 0$
for details):

Up to this the term
$\mathrm {error}'$
, which is a manifestation of the error term in (1.10), we have constructed the Hilbert modular form satisfying the desired analogue of the differential equation (1.7).
In addition to satisfying the differential equation, we still need to better understand the Fourier coefficients of
$\tilde {\mathcal {I}}$
and compare them to those of
$E^{*, \prime }$
. This is done in Section 4.2, where we show that they are logarithms of algebraic numbers and give precise factorization information. To achieve this, we introduce a new local section with an s-variable in (3.54) and match it with the standard section involving the s-variable up to an error of
$O(s^m)$
for any positive integer m. This builds upon the results in Section 3.5 and is accomplished in Theorem 3.14. These new local sections are of independent interest, as they do not come from pullback of the standard section on
$H \cong {\mathrm {SO}}(3, 3)$
. In fact, they do not even tensor together to form a global section with an s-variable.
Finally, we still need to handle the term arising from the error on the right-hand side in (1.10). This boils down to proving the rationality of a Millson theta lift, which is given in Proposition 4.9. For this, we need the Fourier expansion of such a lift computed in [Reference Alfes-Neumann and SchwagenscheidtANS18], and to choose the matching section with a suitable invariance property. Proceeding essentially as in [Reference Gross, Kohnen and ZagierGKZ87] or [Reference Bruinier, Ehlen and YangBEY21], with
$E^{*, \prime }$
replaced by its sum with
$\tilde {\mathcal {I}}$
, we complete the proof of Theorem 1.2.
1.4 Outlook and organization
The factorization of the algebraic numbers appearing in the Fourier coefficients of
$\tilde {\mathcal {I}}$
are very closely related to the Fourier coefficients of
$E^{*, \prime }$
, which suggests that they should reflect the non-archimedean part of the arithmetic intersection between integral versions of
$Z_f$
and
$Z(W)$
defined over the ring of integers of F. It would be very interesting to relate this arithmetic intersection to special values of derivatives of L-functions as in [Reference Bruinier, Kudla and YangBKY12] by applying and refining the results in [Reference Andreatta, Goren, Howard and PeraAGHMP18].
It would be interesting to investigate the analogues of Theorems 1.2 and 1.5 for other CM, étale
$\mathbb {Q}$
-algebras
$E/\mathbb {Q}$
. When
$E/\mathbb {Q}$
has degree 4, there are four cases
-
1.
$E/\mathbb {Q}$ is biquadratic,
-
2.
$E/\mathbb {Q}$ is a product of imaginary quadratic fields,
-
3.
$E/\mathbb {Q}$ is cyclic,
-
4.
$E/\mathbb {Q}$ is a non-Galois, quartic extension.
The CM points
$Z(W)$
have a moduli interpretation as abelian surfaces with CM by the reflex CM algebra
$E^\#$
. The present paper treats case (1). In cases (2) and (3), the reflex algebras
$E^\#$
are quartic, abelian field extensions of
$\mathbb {Q}$
, and the CM cycle
$Z(W)$
is already defined over
$\mathbb {Q}$
. In the last case,
$E^\#/\mathbb {Q}$
is a quartic, non-Galois field, and
$Z(W)$
is defined over a real quadratic field. We plan to extend the ideas and techniques in this paper to prove the analogue of Theorem 1.5 in cases (2)–(4). One difficulty that arises is that the quadratic space of signature
$(3, 3)$
will have Witt rank less than 3, making it impossible to apply the Siegel-Weil formula to identify the theta integral with an Eisenstein series. Instead, one could try to add a twist to the theta integral (see [Reference LiLi16]), compute its Fourier expansion and match it with that of an Eisenstein series. When
$E/\mathbb {Q}$
is a field of degree greater than 4, the Hilbert Eisenstein series are over totally real fields of degree greater than 2 and hence do not arise from theta integral of elliptic modular forms. For such cases, one would need some new ideas.
In addition, there are other applications of these expected results. For cases (2) and (3), by combining the analogue of Theorem 1.2 and the idea in [Reference LiLi21], we hope to obtain a non-existence result of genus 2 curves with CM Jacobian and having everywhere good reduction in certain families, generalizing the main result in [Reference Habegger and PazukiHP17]. In the last case, we expect a variation of our construction to lead to a proof of the factorization conjecture of CM-values of twisted Borcherds product in [Reference Bruinier and YangBY07].
The paper is organized in the following way. Section 2 contains preliminaries. Most of these are standard, except for Section 2.8, which contains the adelic version of the results in [Reference Charollois and LiCL20]. Section 3 matches the coherent Eisenstein series with the Doi-Naganuma lift of Hecke’s cusp form. Section 4 defines
$\tilde {\mathcal {I}}$
and studies its various properties. Finally, we give the proofs of Theorems 1.2 and 1.5 in the last section.
2 Preliminaries
In this section, we introduce some preliminary notions, most of which are standard from the literature. The only material not easily found in the literature are in Sections 2.7 and 2.8 concerning the weight one cusp forms of Hecke in the adelic language, which are translated from the results in [Reference Charollois and LiCL20] in the classical language.
Let
$\mathbb {N}$
denote the set of positive integers and
$\mathbb {N}_0 := \mathbb {N} \cup \{0\}$
. For a number field E, let
$\mathbb {A}_E$
be its ring of adeles,
$\hat E$
the finite adeles and
$\mathbb {A} =\mathbb {A}_{\mathbb {Q}}$
with
$\psi = \psi _f \psi _\infty $
its usual additive character. For an algebraic group
$\mathrm {G}$
over E, denote
$[\mathrm {G}] = \mathrm {G}(E)\backslash \mathrm {G}(\mathbb {A}_E)$
. As usual, let
$G = \operatorname {SL}_2$
with standard Borel
$B = MN \subset G$
. Denote also

and

Throughout the paper, F will be a real quadratic field (unless stated otherwise). Let
$\prime \in \operatorname {Gal}(F/\mathbb {Q})$
be the nontrivial element. It induces an automorphism of
$\mathbb {A}_F, \mathbb {A}_F^\times $
and
$F_p := F \otimes \mathbb {Q}_p$
for each prime
$p \le \infty $
, If p is a finite prime that splits in F (resp. is the infinite place), then F has two embeddings into
$\mathbb {Q}_p$
(resp.
$\mathbb {R}$
), and
$F_p$
is a 2-dimensional vector space over
$\mathbb {Q}_p$
(resp.
$\mathbb {R}$
). For
$\lambda \in F$
, let
$\lambda _1, \lambda _2$
denote the images under those embeddings. We will also sometimes use
$\lambda $
to represent the pair
$(\lambda _1, \lambda _2)$
in
$\mathbb {Q}_p^2$
(resp.
$\mathbb {R}^2$
), and
$\lambda '$
would represent
$(\lambda _2, \lambda _1)$
. We have the incomplete Gamma function

2.1 Differential operators
For a real-analytic function f on
$G(\mathbb {R})$
, the Lie algebra
${\mathfrak {sl}}_2(\mathbb {C})$
acts via

We define the raising and lowering operator

If f is right
$K_\infty $
-equivariant of weight k, then we have

where
$R_{\tau , k}$
and
$L_{\tau , k}$
are the usual raising and lowering operators given by

We say that f is holomorphic, resp. anti-holomorphic, if
$L(f) = 0$
, resp.
$R(f) = 0$
.
For
$ r\in \mathbb {N}_0$
and
$k_1, k_2 \in {\tfrac {1}{2}}\mathbb {Z}$
, define

in
$\mathbb {Q}[X, Y]$
. We omit
$(k_1, k_2)$
from the notation when it is
$(1, 1)$
, in which case

with
$P_r(x)$
the r-th Legendre polynomial given explicitly by

where
$r_0 := \lfloor r/2 \rfloor $
. The second identity comes from (3.133) on page 38 of [Reference GouldGou72] and direct calculation. We thank Zhiwei Sun for pointing us to this reference.
From the differential equation satisfied by
$P_r$
, we have

For
$A \in {\mathfrak {sl}}_2(\mathbb {C})$
, denote

in
${\mathfrak {sl}}_2(\mathbb {C})^2$
. Then we have two operators

on real-analytic functions on
$G(\mathbb {R})^2$
. If
$f: G(\mathbb {R})^2 \to \mathbb {C}$
is holomorphic and right
$K_\infty ^2$
-equivariant of weight
$(k_1, k_2)$
, then
$\mathrm {RC}_{r, (k_1, k_2)}(f)^\Delta : G(\mathbb {R}) \to \mathbb {C}$
is holomorphic and right
$K_\infty $
-equivariant of weight
$k_1 + k_2 + 2r$
, and the operator
$\mathrm {RC}$
is usually called the Rankin-Cohen operator. Here,
$f^\Delta (g) := f(g^\Delta ) = f(g, g)$
is the restriction of f to the diagonal
$G(\mathbb {R}) \subset G(\mathbb {R})^2$
. In fact, we have (see [Reference Bruinier, van der Geer, Harder and ZagierBvdGHZ08, Proposition 19])

For example, if
$f(g_{z_1}, g_{z_2}) = {\mathbf {e}}(m_1 z_1 + m_2 z_2)$
, then

From Lemma 2.2 in [Reference LiLi23], we know that there are unique constants
$c^{(r; k_1, k_2)}_{\ell } \in \mathbb {Q}$
such that

whenever
$k_1 + k_2 + 2r < 2$
.
2.2 Quadratic space associated to real quadratic field
Let
$F = \mathbb {Q}(\sqrt {D})$
be a real quadratic field, which becomes a
$\mathbb {Q}$
-quadratic space of signature
$(1, 1)$
with respect to the quadratic form

for any
$a \in \mathbb {Q}^\times $
. We denote this quadratic space by
$V_a $
and identify

This is an isometry, where the quadratic form on
$\mathbb {R}^2$
is given by
$Q((x_1, x_2)) = x_1x_2$
. The special orthogonal group
$H_a := \mathrm {SO}({V_a })$
satisfies

where
$ F^1 := \{\mu \in F: {\mathrm {Nm}}(\mu ) = 1\}$
acts on V via multiplication. Furthermore, we identify

where
$t \in \mathbb {R}^\times $
acts on
$V_a (\mathbb {R}) = F \otimes _{\mathbb {Q}} \mathbb {R} \cong \mathbb {R}^2$
via
$(x_1, x_2) \mapsto (t x_1, t^{-1}x_2)$
. Note that the invariant measure on
$H(\mathbb {R}) \cong \mathbb {R}^\times $
is
$\frac {dt}{|t|}$
. We have a character
${\mathrm {sgn}} : H_a (\mathbb {R}) \cong \mathbb {R}^\times \to \{\pm 1\}$
. Denote

its kernel, which is the connected component of
$H_a(\mathbb {R})$
containing the identity. We also denote

which is an index 2 subgroup of
$H_a (\mathbb {Q})$
.
Remark 2.1. Let
$\chi = \chi _{E/F}$
be a Hecke character associated to a quadratic extension
$E/F$
. Suppose
$E/\mathbb {Q}$
is biquadratic. Then
$\chi \mid _{\mathbb {A}^\times }$
is trivial and
$\chi $
factors through the map
${\mathrm {Nm}}^-: \mathbb {A}_F^\times \to H_a (\mathbb {A})$
; that is, there exists
$\varrho = \varrho _{E/F}: H_a (\mathbb {A}) \to \{\pm 1\}$
such that

Note that
$\varrho $
is odd if and only if E is totally imaginary. We also denote the compact subgroup

Note that
$H_a (\mathbb {Q})\backslash H_a (\hat {\mathbb {Q}})/K_\varrho $
is a finite set.
2.3 The Weil representation and theta functions
Let
$(V, Q)$
be a rational quadratic space of signature
$(p, q)$
, and
$H_V := {\mathrm {SO}}(V)$
. For a subfield
$E \subset \mathbb {C}$
, we denote
${\mathcal {S}}(\hat V; E)$
, resp.
${\mathcal {S}}(V_p; E)$
, to denote the space of Schwartz functions on
$\hat V := V(\hat {\mathbb {Q}})$
, resp.
$V_p := V(\mathbb {Q}_p)$
, with values in E, which is an E-vector space. We omit E from the notation if it is
$\mathbb {Q}$
. However, we write
${\mathcal {S}}(V(\mathbb {R}))$
and
${\mathcal {S}}(V(\mathbb {A}))$
for the space of Schwartz function on
$V(\mathbb {R})$
and
$V(\mathbb {A})$
valued in
$\mathbb {C}$
, respectively.
For a latticeFootnote
3
$L \subset V$
, we denote
$L^\vee \subset V$
its dual lattice,
$\hat L := L \otimes \hat {\mathbb {Z}}$
,
$\hat L^\vee := L^\vee \otimes \hat {\mathbb {Z}}$
and

for
$m \in \mathbb {Q}, \mu \in L^\vee /L$
. The finite dimensional E-subspace

is spanned by
$\{\phi _{L + \mu } : \mu \in \hat L^\vee /\hat L \cong L^\vee /L\}$
with

For full sublattices
$M \subset L \subset V$
, it is clear that

As above, we also denote
${\mathcal {S}}(L) := {\mathcal {S}}(L; \mathbb {Q})$
. Furthermore, since

we have
${\mathcal {S}}(\hat V; E) = {\mathcal {S}}(\hat V) \otimes _{\mathbb {Q}} E$
for any subfield
$E \subset \mathbb {C}$
.
Suppose
$V = V_1 \oplus V_2$
is a decomposition of rational quadratic spaces. For any lattice
$L_i \subset V_i$
, we have
${\mathcal {S}}(L_1 \oplus L_2; E) = {\mathcal {S}}(L_1; E) \otimes {\mathcal {S}}(L_2; E) \subset {\mathcal {S}}(\hat V_1; E) \otimes {\mathcal {S}}(\hat V_2; E)$
via the natural restriction map. This also gives us

Combining with equation (2.18), we see that for any given
$\phi \in {\mathcal {S}}(\hat V; E)$
, we can find a lattice
$L = L_1 \oplus L_2 \subset V$
such that
$L_i \subset V_i$
and

Let
$\tilde G(\mathbb {A}) := \mathrm {Mp}_2(\mathbb {A})$
be the metaplectic cover of
$G(\mathbb {A})$
. The group
$\tilde G(\mathbb {A}) \times H_V(\mathbb {A})$
acts on
${\mathcal {S}}(V(\mathbb {A}))$
via the Weil representation
$\omega = \omega _{V, \psi }$
(see [Reference KudlaKud94, section 5] for explicit formula). For each prime
$p \le \infty $
, we also have the local Weil representation
$\omega _p$
of
$G(\mathbb {Q}_p)$
acting on
${\mathcal {S}}(V(\mathbb {Q}_p); \mathbb {C})$
. Then
$\omega _f := \otimes _{p < \infty } \omega _p$
gives a representation of
$G(\hat {\mathbb {Q}})$
on
${\mathcal {S}}(\hat V; \mathbb {C})$
.
For any lattice
$L \subset V$
, the subspace
${\mathcal {S}}(L; \mathbb {C})$
defined in (2.16) is a
$K_f$
-invariant subspace with
$K_f := G(\hat {\mathbb {Z}})$
. It has a unitary pairing
$\langle , \rangle $
with the vector space
$\mathbb {C}[L^\vee /L] := \oplus _{\mu \in L^\vee /L} \mathbb {C} \mathfrak {e}_\mu $
given by

More generally, if
$L = L_1 \oplus L_2$
, we have
$L^\vee /L = L_1^\vee /L_1 \oplus L_2^\vee / L_2$
and
$\mathbb {C}[L^\vee /L] \cong \mathbb {C}[L_1^\vee /L_1] \otimes \mathbb {C}[L_2^\vee /L_2]$
. Therefore, we can extend the pairing above to

with
$\mathfrak {v}_i \in \mathbb {C}[L_i^\vee /L_i]$
and
$\phi \in {\mathcal {S}}(L_2; \mathbb {C})$
.
With respect to the perfect pairing in (2.21), the unitary dual of
$\omega _f$
is the representation
$\rho _L$
on
$\mathbb {C}[L^\vee /L]$
given by

where
$(p, q)$
is the signature of
$V(\mathbb {R})$
. This is the Weil representation on finite quadratic modules used by Borcherds in [Reference BorcherdsBor98]. If we identify
${\mathcal {S}}(L; \mathbb {C})$
and
$\mathbb {C}[L^\vee /L]$
with
$\mathbb {C}^{|L^\vee /L|}$
via the bases
$\{\phi _\mu : \mu \in \hat L^\vee /\hat L\}$
and
$\{\mathfrak {e}_\mu : \mu \in L^\vee / L\}$
, respectively, then
$\omega _f = \overline {\rho _L} = \rho _L^{-1}$
. For full sublattices
$M \subset L$
, the following trace map

intertwines the Weil representation and is compatible with the inclusion in (2.18) in the sense that

for any
$\mathfrak {v} \in \mathbb {C}[L^\vee /L], \phi \in {\mathcal {S}}(L; \mathbb {C})$
.
The following result will be very useful for us later.
Lemma 2.2. For any prime p, the local Weil representation
$\omega _p$
descends to
${\mathcal {S}}(V(\mathbb {Q}_p); \mathbb {Q}(\zeta _{p^\infty }))$
with

the maximal abelian extension of
$\mathbb {Q}$
ramified only at p.
Proof. Via
$L'/L = \oplus _p L^{\prime }_p/L_p$
with
$L_p := L \otimes \mathbb {Z}_p$
, we can write
$\rho _L = \otimes _p \rho _{p}$
with
$\rho _p$
the Weil representation associated to the finite quadratic module
$L^{\prime }_p/L_p$
and identify
$\omega _p = \rho _{p}^{-1}$
. It is well-known that any finite quadratic module can be written in the form
$M'/M$
for an even integral lattice M [Reference NikulinNik79]. Therefore, it suffices to prove the claim with
$\omega _p$
replaced by
$\rho _M$
for an even integral lattice M with quadratic form valued in
$\mathbb {Z}[1/p]$
. This follows then directly from formula (2.23) and Milgram’s formula [Reference BorcherdsBor98, Corollary 4.2].
As usual, we let
$H_{k, L}(\Gamma )$
denote the space of harmonic Maass forms valued in
$\mathbb {C}[L^\vee /L]$
of weight
$k \in \frac {1}{2}\mathbb {Z}$
and representation
$\rho _L$
on a congruence subgroup
$\Gamma \subset \operatorname {SL}_2(\mathbb {Z})$
(see [Reference Bruinier and FunkeBF04, section 3]). It contains the subspace
$M^!_{k, L}(\Gamma )$
of vector-valued weakly holomorphic modular forms. Post-composing with
$\operatorname {Tr}^L_M$
in (2.24) induces a map
$H_{k, L}(\Gamma ) \to H_{k, M}(\Gamma )$
, which preserves holomorphicity and rationality of holomorphic part Fourier coefficients. If
$L^\vee /L$
is trivial (resp.
$\Gamma = \operatorname {SL}_2(\mathbb {Z})$
), then we drop L (resp.
$\Gamma $
) from the above notations. Furthermore, we let

and denote for
$f(\tau ) = \sum _{m \in \mathbb {Q},~ \mu \in L^\vee /L} c(m, \mu ) q^m \mathfrak {e}_\mu \in M^!_{k, L}$

the principal part of f.
In [Reference McGrawMcG03, Theorem 4.3], McGraw extended the representation
$\rho _L$
to the metaplectic cover of
${\mathrm {GL}}_2$
. To simplify the notation, we recall it here for lattices with even rank, in which case this extension factors through
${\mathrm {GL}}_2$
. Using the short exact sequence

we can identify
${\mathrm {GL}}_2 \cong \operatorname {SL}_2 \rtimes T$
. Then
$\omega _f$
extends to a
$\mathbb {Q}$
-linear action of
${\mathrm {GL}}_2(\hat {\mathbb {Q}}) = {\mathrm {GL}}_2(\mathbb {Q}){\mathrm {GL}}_2(\hat {\mathbb {Z}})$
on
${\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})$
via

for
$ g \in \operatorname {SL}_2(\hat {\mathbb {Q}}), a \in \hat {\mathbb {Q}}^\times , \phi \in {\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})$
, where
$\sigma _a \in \operatorname {Gal}(\mathbb {Q}^{\mathrm {ab}}/\mathbb {Q})$
satisfies
$\sigma _a(\psi _f(a')) = \psi _f(aa')$
for all
$a , a' \in \hat {\mathbb {Q}}^\times $
and acts on
${\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})$
via its action on
$\mathbb {Q}^{\mathrm {ab}}$
. This gives us

For each
$p < \infty $
, this gives an extension of
$\omega _p$
to a
$\mathbb {Q}$
-linear action of
${\mathrm {GL}}_2(\mathbb {Q}_p)$
on
${\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))$
, which satisfies

For
$\varphi \in {\mathcal {S}}(V(\mathbb {A}))$
, we have the theta function

for
$(g, h) \in (G \times H_V)(\mathbb {A})$
as usual. For a lattice
$L \subset V$
, we also denote

the vector-valued theta function with
$\phi _\infty $
the Gaussian.
2.4 CM points and higher Green functions
We follow [Reference Bruinier, Kudla and YangBKY12] and [Reference Bruinier, Ehlen and YangBEY21] to recall CM points and higher Green functions. Let
$(\mathrm {V}, Q)$
be a rational quadratic space of signature
$(n, 2)$
, and
${\tilde {H}} = {\tilde {H}}_{\mathrm {V}} := \mathrm {GSpin}(\mathrm {V})$
. For an open compact subgroup
$K \subset {\tilde {H}}(\hat {\mathbb {Q}})$
, we have the associated Shimura variety
$X_K$
, whose
$\mathbb {C}$
-points are given by

Here,
$\mathbb {D} = \mathbb {D}^+ \sqcup \mathbb {D}^-$
is the symmetric space associated to
$\mathrm {V}(\mathbb {R})$
. For
$m \in \mathbb {Q}$
and
$\varphi \in {\mathcal {S}}(\hat {\mathrm {V}}; \mathbb {C})$
, one can define the special divisor
$Z(m, \varphi )$
on
$X_K$
(see, for example, [Reference Bruinier, Ehlen and YangBEY21, section 2]).
The CM points on
$X_K$
can be described as follows. For a totally real field F of degree d with real embeddings
$\sigma _j, 1 \le j \le d$
, denote
$\alpha _j := \sigma _j(\alpha )$
for
$\alpha \in F$
. Suppose
$\alpha _{j_0} < 0$
for some
$j_0$
and
$\alpha _j> 0$
when
$j \neq j_0$
. Then a CM quadratic extension
$E/F$
becomes an F-quadratic space
$W = W_\alpha = E$
with respect to the quadratic form
$Q_\alpha := \alpha {\mathrm {Nm}}_{E/F}$
. Suppose there is an isometric embedding as in (1.5). Then the subspace
$W \otimes _{F, \sigma _{j_0}} \mathbb {R} \subset \mathrm {V}(\mathbb {R})$
is a negative 2-plane and determines a point
$z_0^\pm \in \mathbb {D}^\pm $
with a choice of orientation. For convenience, we denote

The group
$\mathrm {Res}_{F/\mathbb {Q}}({\mathrm {SO}}(W))$
is contained in
${\mathrm {SO}}(\mathrm {V})$
, whose preimage in
${\tilde {H}}_{\mathrm {V}}$
is a torus denoted by
$T_W$
. Note that
$T_W(\mathbb {Q}) = E^\times /F^1$
. We denote the CM cycle on
$X_K$
associated to
$T_W$
by

with
$K_W := K \cap T_W(\hat {\mathbb {Q}})$
. It is defined over F, and its Galois conjugates are the CM cycles
$Z(W(j))$
with
$1 \le j \le d$
, where
$W(j)$
is the neighborhood F-quadratic spaces at
$\sigma _j$
of admissible incoherent
$\mathbb {A}_F$
-quadratic space
$\mathbb {W}$
associated to W (see [Reference Bruinier and YangBY11, Reference Bruinier, Kudla and YangBKY12] for details). Note that
$W(j) = W_{\alpha (j)}$
for some
$\alpha (j) \in F^\times $
and
$\alpha (j_0) = \alpha $
. For totally positive
$t \in F$
, we define the ‘Diff’ set

following [Reference KudlaKud97]. Note that this set is finite and odd (see [Reference Yang and YinYY19, Proposition 2.7]).
When F is real quadratic (i.e.
$d = 2$
), for
$\alpha \in F^\times $
with
${\mathrm {Nm}}(\alpha ) < 0$
, we set

Then
$({\alpha ^\vee })^\vee = \alpha $
. Note that
${\alpha ^\vee }$
is not necessarily the Galois conjugate of
$\alpha $
!
Denote

Let
$L \subset \mathrm {V}$
be an even integral lattice such that K stabilizes
$\hat L$
. For
$\mu \in L^\vee /L$
and
$m \in \mathbb {Z} + Q(\mu )$
, we write
$Z(m, \mu ) := Z(m, \phi _\mu )$
. The automorphic Green function on
$X_K \backslash Z(m, \mu )$
is defined by

for
$\mathrm {Re}(s)> \sigma _0 + 1$
, where
$F(a, b, c; z)$
is the Gauss hypergeometric function [Reference Abramowitz and StegunAS64, Chapter 15]. At
$Z(m, \mu )$
, the function
$\Phi _{m, \mu }$
has logarithmic singularity.
At
$s = \sigma _0 + 1 + r$
with
$r \in \mathbb {N}$
, the function
$\Phi _{m, \mu }(z, h, s)$
is called a higher Green function. For a harmonic Maass form
$f = \sum _{m, \mu } c(m, \mu ) q^{-m} \phi _\mu + O(1) \in H_{k - 2r, L}$
with
$k:= -2\sigma _0$
, define

to be the associated higher Green function. Following from the work of Borcherds [Reference BorcherdsBor98] and generalization by Bruinier [Reference BruinierBru02] (also see [Reference Bruinier, Ehlen and YangBEY21, Proposition 4.7]), the function
$\Phi _f^r$
has the following integral representation:

where
${\mathcal {F}}_T$
is the truncated fundamental domain of
$\operatorname {SL}_2(\mathbb {Z}) \backslash \mathbb {H}$
at height
$T> 1$
and
$d\mu $
is the invariant measure given in (3.20). It has logarithmic singularity along the special divisor

on
$X_K$
. Note that
$[z_0, h] \in Z(W) \cap Z_f$
if and only if

for some
$m, \mu $
with
$c(-m, \mu ) \neq 0$
.
2.5 Product of modular curves as a Shimura variety
We follow and slightly modify [Reference Yang and YinYY19, section 3] to express
$X_0(N) \times X_0(N)$
as
$\mathrm {O}(2, 2)$
orthogonal Shimura variety. Consider
$(\mathrm {V}, Q) = (M_2(\mathbb {Q}), \det )$
, and the lattice

for any
$N \in \mathbb {N}$
. Then the dual lattice
$L^\vee $
is given by

and
$L^\vee /L \cong (\mathbb {Z}/N\mathbb {Z})^2$
is isomorphic to that of a scaled hyperbolic plane.
For
$g_j \in \operatorname {SL}_2(\mathbb {Q})$
and
$\Lambda \in \mathrm {V}(\mathbb {Q})$
, the map

gives
$\operatorname {SL}_2 \times \operatorname {SL}_2 \cong \operatorname {Spin}(\mathrm {V})$
and identifies
$H_{\mathrm {V}}$
as a subgroup of
${\mathrm {GL}}_2 \times {\mathrm {GL}}_2$
[Reference Yang and YinYY19, section 3.1]. Let
$K(N) := K(\Gamma _0(N)) \subset {\mathrm {GL}}_2(\hat {\mathbb {Z}})$
be the open compact subgroup in [Reference Yang and YinYY19, section 3.1] and
$K := (K(N) \times K(N)) \cap H_{\mathrm {V}}(\hat {\mathbb {Q}})$
. Then the map

induces an isomorphism

with
$X_K$
the Shimura variety for
$H_{\mathrm {V}}$
.
Under the map (2.44), the inverse images of the discriminant kernel
$\Gamma _L \subset {\mathrm {SO}}(L)$
are

which contains
$ \Gamma _1(N) \times \Gamma _1(N)$
and is a normal subgroup of
$\Gamma _0(N)^2$
satisfying

The group
${\mathrm {SO}}(L^\vee /L) := {\mathrm {SO}}(L)/\Gamma _L \cong \Gamma _0(N)^2/\Gamma ^\Delta _0(N) \cong (\mathbb {Z}/N\mathbb {Z})^\times $
acts on
$L^\vee /L \cong (\mathbb {Z}/N\mathbb {Z})^2$
via

and the induced linear map on
$\mathbb {C}[L^\vee /L]$
intertwines the Weil representation
$\rho _L$
.
Now given
$f \in M^!_{k}(\Gamma _0(N))$
for
$k \in 2\mathbb {Z}$
, we can lift it to a vector-valued modular form in
$M^!_{k, \rho _L}$
via the following map:

This map and its generalizations are well-studied (see, for example, [Reference ScheithauerSch09]), whose properties are summarized in the following result.
Lemma 2.3. When
$k < 0$
, we have

for all
$f \in M^{!, \infty }_{k}(\Gamma _0(N))$
, on which space the map
$\mathrm {vv}$
is an isomorphism with the inverse given by

Furthermore, it preserves the rationality of the Fourier expansion at the cusp infinity.
Proof. See Proposition 4.2 in [Reference ScheithauerSch09] and Proposition 6.12, Corollary 6.14 in [BHK+20].
As a consequence, we can relate the higher Green function
$G^{\Gamma _0(N)}_{r+1, f}$
from the introduction to one on the Shimura variety
$X_K$
.
Corollary 2.4. Under the isomorphism (2.46), we have

with
$G^{\Gamma _0(N)}_{r+1, f}$
the higher Green function defined in (1.4) for
$f \in M^{!, \infty }_{-2r}(\Gamma _0(N))$
with
$r> 0$
.
Proof. Under (2.46), the divisor

on
$X_K$
is simply the m-th Hecke correspondence
$T_m$
on
$X_0(N) \times X_0(N)$
. Therefore, the two sides of (2.50) have logarithmic singularity along the same divisor. Using Corollary 4.2 and Theorem 4.4 of [Reference Bruinier, Ehlen and YangBEY21], we see that their difference is a smooth function in
$L^2(X_0(N)^2)$
and an eigenfunction of the Laplacians in
$z_1$
and
$z_2$
with eigenvalue
$r(1-r) < 0$
. By fixing
$z_2$
, this difference is an eigenfunction of the Laplacian on
$X_0(N)$
with negative eigenvalue, which vanishes identically. This holds for any
$z_2 \in X_0(N)$
, and we obtain (2.50).
Remark 2.5. Following Section V.4 of [Reference Gross and ZagierGZ86], we call a set of integers
$\{\lambda _m: m \in \mathbb {N}\}$
a relation for
$S_{2-k}(\Gamma _0(N))$
if only finitely many
$\lambda _m$
are nonzero and

for all
$\sum _{m \ge 1} a_m q^m \in S_{2-k}(\Gamma _0(N))$
. Since
$g_0 \in S_{2-k}(\Gamma _0(N))$
for all
$g \in S_{2-k, L^-}$
, we have

for
$f_P = \sum _{m \ge 1} \lambda _m q^{-m} \mathfrak {e}_0$
. By Serre duality [Reference BorcherdsBor99], there exists
$f \in M^!_{k, L}$
with
$\mathrm {prin}(f) = f_P$
.
Suppose
$E_1, E_2$
are imaginary quadratic fields such that
$E = E_1E_2$
is biquadratic containing a real quadratic field F (i.e.,
$E_1 \neq E_2$
). Then for any CM points
$z_j \in E_j$
, the point
$(z_1, z_2) \in \mathbb {H}^2$
is sent to
$Z(W_\alpha ) \cup Z(W_{\alpha ^\vee }) \subset X_K$
under the isomorphism in (2.46) (see Section 3.2 in [Reference Yang and YinYY19] for details).
2.6 Eisenstein series
We recall coherent and incoherent Eisenstein series for the group
$G = \operatorname {SL}_2$
following [Reference Bruinier, Kudla and YangBKY12]. Let F be a totally real field of degree d and discriminant
$D_F$
,
$E/F$
be a quadratic CM extension with absolute discriminant
$D_{E}$
and
$\chi = \chi _{E/F} = \otimes _{v \le \infty } \chi _v$
the associated Hecke character. For a standard section
$\Phi \in I(s, \chi )$
with

we can form the Eisenstein series

where
$\Lambda (s, \chi )$
is the completed L-function for
$\chi $
(see equation (4.6) in [Reference Bruinier, Kudla and YangBKY12]). When
$\Phi = \otimes _v \Phi _v$
, the Eisenstein series
$E(g, s, \Phi )$
has the Fourier expansion

and for
$t \in F^\times $
,

where
$W^*_{t, v}$
is the normalized local Whittaker function defined by

For simplicity, we denote

We will be interested in the case when
$\Phi $
is a Siegel-Weil section.
Given
$\alpha \in F^\times $
, we view
$W_\alpha = E$
as an F-quadratic space with quadratic form
$Q_\alpha (z) := \alpha z\bar z$
. Denote
$\omega _\alpha $
the associated Weil representation. We have an
$\operatorname {SL}_2$
-equivariant map
$\lambda _\alpha : {\mathcal {S}}(W_\alpha (\mathbb {A}_F)) \to I(0, \chi )$
given by

At each place v, there are local versions of
$\omega _\alpha $
and
$\lambda _\alpha $
as well, denoted by
$\omega _{\alpha , v}$
and
$\lambda _{\alpha , v}$
. When
$\Phi = \lambda _\alpha (\phi )$
, resp.
$\Phi _v = \lambda _{\alpha , v}(\phi _v)$
, we replace
$\Phi $
, resp.
$\Phi _v$
, from the notations above by
$\phi $
, resp.
$\phi _v$
.
Let
$Z(W_\alpha )$
be the CM points on
$X_K$
as in Section 2.4 and
$W_{\alpha , \mathbb {Q}} \subset \mathrm {V}$
the rational quadratic space as in (1.5). A special case of the Siegel-Weil formula (see [Reference Bruinier, Kudla and YangBKY12, Theorem 4.5]) gives us

for any
$\phi \in {\mathcal {S}}(W_{\alpha , \mathbb {Q}}(\mathbb {A})) = {\mathcal {S}}(W_\alpha (\mathbb {A}_F))$
. Here,
$C = \deg (Z(W_\alpha ))/2$
.
Suppose only the j-th real embedding of
$\alpha $
is negative. Denote
and
with
$-1$
at the j-th slot. The sections
and
are coherent and incoherent, respectively. For all
$\phi \in {\mathcal {S}}(W_\alpha (\hat F); \mathbb {C})$
, the Eisenstein series
$E^*(g, s, \Phi (j))$
is holomorphic of weight
at
$s = 0$
and

is called a coherent Eisenstein series. However, the Eisenstein series
$E^*(g, s, \Phi )$
vanishes at
$s = 0$
, and its derivative

is called an incoherent Eisenstein series, which is related to the coherent Eisenstein series via the differential equation [Reference Bruinier, Kudla and YangBKY12, Lemma 4.3]

for all
$1 \le j \le d$
. Furthermore, it has the Fourier expansion

where
$\mathcal {E}^*(\tau , \phi )$
has exponential decay near the cusp infinity and

Here,
$a_0(\phi )$
is an explicit constant (see (2.24) in [Reference Yang and YinYY19]) and

The coefficient
$\tilde {W}_t(\phi )$
is given by (see [Reference Yang and YinYY19, Proposition 2.7])

when
$\mathrm {Diff}(W_\alpha , t) = \{\mathfrak {p}\}$
2.7 Hecke’s cusp form
Denote
$\omega _a = \omega _{V_a , \psi }$
the Weil representation and
$\theta _a := \theta _{V_a}$
the theta function as in (2.32). For a bounded, integrable function
$\rho : H_a (\mathbb {Q}) \backslash H_a (\mathbb {A})/K \to \mathbb {C}$
, consider the following theta lift:

The measure
$dh$
is the product measure of the local measures
$dh_p$
, where
$dh_p$
is normalized such that the maximal compact subgroup in
$H_a(\mathbb {Q}_p)$
has volume 1. Such integral was first considered by Hecke in [Reference HeckeHec27] when
$\rho = \varrho $
is an odd, continuous character – that is,

with
$\varrho _f$
a continuous character on
$H_a(\hat {\mathbb {Q}})$
and
$\varphi = \varphi ^\pm _\infty \varphi _f$
with

Notice that
$\varphi ^\pm _\infty $
satisfies
$\varphi ^\pm _\infty (-x_1, -x_2) = - \varphi ^\pm _\infty (x_1, x_2)$
.
In this case, the m-th Fourier coefficient of
$\vartheta _a $
is given by

for
$m \in \mathbb {Q}$
. To evaluate it, we apply the usual unfolding trick

When
$m = 0$
, we have
$\lambda = 0$
since
$V_a $
is anisotropic and
$\varphi (0) = \varphi ^\pm _\infty (0)\varphi _f(0) = 0$
. When
$Q_a (\lambda ) = m \neq 0$
, the group
$H_{a , \lambda }$
is trivial. We can then write
$g = g_\tau g_f$
with
$g_\tau = n(u)m(\sqrt {v}) \in G(\mathbb {R})$
and
$g_f \in G(\hat {\mathbb {Q}})$
and obtain

The group
$H_a (\mathbb {Q})$
acts on
$V_{a , m}(\mathbb {Q})$
transitively. The archimedean integral can be evaluated as

This is 0 if
$\pm {\mathrm {sgn}}(m) < 0$
by the change of variable
$t \mapsto 1/t$
and can be otherwise evaluated using the lemma below.
Lemma 2.6. For any
$\beta> 0$
, we have

Therefore, we have

when
$\pm m> 0$
and
$\lambda \in V_{a , m}(\mathbb {Q}) $
. Otherwise, it is 0. The integral in (2.66) can be evaluated locally. Notice that it always converges as
$\varphi _f(h^{-1} \lambda )$
has compact support as a function of
$h \in H_a (\hat {\mathbb {Q}})$
.
2.8 The Deformed theta integral
We recall the real-analytic modular form of weight one constructed in [Reference Charollois and LiCL20] using the notations of Section 2.7. Let
$\varrho $
be an odd, continuous character as in (2.64), and
$K_\varrho \subset H_a (\hat {\mathbb {Z}})$
the open compact subgroup defined in (2.14). The intersection
$H_a(\mathbb {Q})^+ \cap K_\varrho $
, where
$H_a(\mathbb {Q})^+ := H_a(\mathbb {Q}) \cap H_a(\mathbb {R})^+$
, is a cyclic subgroup

of the totally positive units in
${\mathcal {O}}$
. Then we have

where
$C \subset H_a (\hat {\mathbb {Q}})$
is a finite subset of elements representing
$H_a (\mathbb {Q})^+ \backslash H_a (\hat {\mathbb {Q}}) / K_\varrho $
. So given
$h = (h_f, h_\infty )$
, we can find
$\alpha \in H_a (\mathbb {Q}), t \in H_a (\mathbb {R})^+, k_1 \in K_\varrho , \xi \in C$
all depending on h such that

though the choice is not unique. This gives us the identification

by sending
$h = (\alpha k_1 \xi , \alpha t) \in H_a (\mathbb {A})$
as in (2.69) to
$t \in H_a (\mathbb {R})^+$
in the
$\xi $
-component. Just like the decomposition (2.68), this isomorphism depends on the choice of the set of representatives C. Similarly, we have

Using the Fourier coefficient
$W_m$
in (2.66) and the decomposition in (2.71), we can write

for
$g \in G(\hat {\mathbb {Q}})\times B(\mathbb {R})$
, where
$\varphi ^\pm = \varphi _f \varphi _\infty ^\pm $
with
$\varphi _f \in {\mathcal {S}}(\hat {V}_a)$
being
$K_\varrho $
-invariant and

Although
$\varphi ^{0, \pm }_\infty $
is not a Schwartz function on
$\mathbb {R}^2$
, the sum above still converges absolutely. For
$g \in B(\mathbb {R})$
, the quantity
$\omega _a(g) \varphi ^{0, \pm }_\infty $
is defined with the usual formula of the Weil representation, and
$\vartheta _a(g, \varphi ^\pm , \varrho )$
is right
${\mathrm {SO}}_2(\mathbb {R})$
-equivariant with weight
$\pm 1$
. We also have a left
$G(\mathbb {Q})$
-invariant function

on
$G(\mathbb {A})$
for
$\varphi \in {\mathcal {S}}(V_a(\mathbb {A}))$
, which is independent of the choice of C.
We now define a function
$\lg _C : H_a (\mathbb {Q}) \backslash H_a (\mathbb {A}) / {K_\varrho } \to [0, 1)$
by

Note that
$\{0\} = \frac {1}{2}$
. Unlike the function considered by Hecke,
$\lg _C$
cannot be written as the product of functions on
$H_a (\hat {\mathbb {Q}})$
and
$H_a (\mathbb {R})$
. Denote

Given
$\varphi = \varphi _f \varphi _\infty \in {\mathcal {S}}(V_a(\mathbb {A}))$
for some
${K_\varrho }$
-invariant
$\varphi _f \in {\mathcal {S}}(\hat V_a; \mathbb {C})$
, the deformed theta integral
$\vartheta _a (g, \varphi , \tilde {\varrho }_C)$
, where
$\vartheta _a $
is defined in (2.63), was studied in [Reference Charollois and LiCL20], To describe its Fourier expansion, denote

for
$g = (g_f, g_\tau ) \in G(\mathbb {A})$
with
$\varphi ^{0, *} = \varphi _f \varphi ^{0, *}_\infty $
and

Note that
$\vartheta ^*_a$
is not necessarily left-
$G(\mathbb {Q})$
invariant. But it is modular after applying the lowering operator as

Similarly for
$\xi \in H_a(\hat {\mathbb {Q}})$
, define

Here, we have employed the rapidly decaying function

where
$B(\mathbb {R})\subset G(\mathbb {R})$
acts via
$\omega _a$
and
${\mathrm {SO}}_2(\mathbb {R})$
acts with weight 1. Also, we denote

which depends on the choice of C and satisfies

We recall some results.
Theorem 2.7. Let
$\tilde {\varrho }_C$
be as in (2.76) and
$\varphi _f \in V_a(\hat {\mathbb {Q}})$
a right-
${K_\varrho }$
invariant function. Then the integral
$\vartheta _a(g, \varphi ^+, \tilde {\varrho }_C)$
defines a
$G(\mathbb {Q})$
-invariant function in
$g \in G(\mathbb {A})$
of weight 1 with respect to
${\mathrm {SO}}_2(\mathbb {R})$
. Furthermore, it has the Fourier expansion

where
$\varphi ^{0, +} = \varphi _f \varphi ^{0, +}_\infty $
is defined in (2.73), and satisfies

Proof. This follows essentially from Proposition 5.5 in [Reference Charollois and LiCL20]. For completeness, we include a different (and slightly shorter) proof here. As in the evaluation of
$\vartheta _a(g, \varphi , \varrho )$
in (2.72), we have

To verify (2.83), we start with

Substituting this into the left-hand side of (2.83) proves it.
Now to calculate the Fourier expansion, it suffices prove the claim

For each
$\beta \in \Gamma _{\varrho } \backslash V_a(\mathbb {Q})$
, we choose the unique representative
$\beta _0 \in \Gamma _{\varrho } \beta $
such that
$|\beta _0/\beta _0'| \in [1, \varepsilon _{\varrho }^2)$
. We can then write
$J(\beta , v) = J_1(\beta _0, v) + J_2(\beta _0, v)$
, where

For
$J_1$
, we have

For
$J_2$
, the limit vanishes unless
$|\beta _0/\beta _0'| = 1$
, in which case

Putting these together proves claim (2.84).
Finally, we record a result as a direct consequence of Theorem 4.5 in [Reference Charollois and LiCL20] (see also Section 5 in [Reference Li and SchwagenscheidtLS22]).
Proposition 2.8. For any
$\varphi _f \in {\mathcal {S}}(\hat V_1; \mathbb {C})$
, there exists a real-analytic modular form
$\tilde \Theta _{a, C}(g, \varphi ^-, \varrho ) = \tilde \Theta ^+_{a, C}(g, \varphi ^-, \varrho ) + \tilde \Theta ^*_{a, C}(g, \varphi ^-, \varrho )$
such that
$L \tilde \Theta _{a, C} = \Theta _{a}$
and
$\sqrt {v}\tilde \Theta ^+_{a, C}(g_\tau , \varphi ^-, \varrho )$
is holomorphic in
$\tau $
with Fourier coefficients in
$\mathbb {Q}(\varphi _f)$
.
3 Doi-Naganuma lift of Hecke’s cusp form
In this section, we are interested in computing the
$\mathrm {O}(2, 2)$
theta lift of Hecke’s cusp form from Section 2.7 and realize it as coherent Hilbert Eisenstein series from 2.6 over real quadratic fields. The main result of this section is the global matching Theorem 3.3, where we show that any coherent Eisenstein series can be realized as such a theta lift. This global statement follows from its local counterpart in Theorem 3.10, which is improved further in Theorem 3.14 to allow matching deformed local sections. This last result will be crucial for us in proving the factorization result in Proposition 4.7 later.
3.1 Quadratic spaces
Let
$V_{\pm 1}$
be as in Section 2.2,
$\ell ^+, \ell ^-$
be isotropic lines such that
$\ell ^+ \oplus \ell ^-$
is a hyperbolic plane and denote

We can realize

with
$ \det $
as the quadratic form and furthermore write

where
$Q_D(x) = Dx^2$
. So V has Witt rank 3 and admits the isotropic decomposition

with
$V^\pm $
maximal totally isotropic subspaces. For a
$\mathbb {Q}$
-algebra R (e.g.,
$R \in \{\mathbb {Q}, \mathbb {Q}_p, \mathbb {R}, \hat {\mathbb {Q}} , \mathbb {A}\}$
), we will use

to represent elements in
$V(R)$
. Define elements
$f_j^\pm \in V$
by

Then
$\{f_j^\pm : j = 1, 2, 3\} \subset V^\pm $
is a
$\mathbb {Q}$
-basis of
$V^\pm $
. With respect to the ordered basis
$(f_1^+, f_2^+, f_3^+, f_1^-, f_2^-, f_3^-)$
, the Gram matrix of Q is
$\left (\begin {smallmatrix} 0 & {I_3} \\ I_3 & 0\end {smallmatrix}\right)$
. For
$i = 1, 2, 3$
, the following linear transformations

are easily checked to be in
$\mathrm {O}(V)$
. The unimodular lattice

provides V with an integral structure. Similarly for
$?\in \{00, 0, 1\}$
, the lattice
$V_{?, \mathbb {Z}} := V_{\mathbb {Z}} \cap V_? $
in
$V_?$
gives it with an integral structure.
For
$? \in \{00, 0, 1, -1, \emptyset \}$
, we write

which are subgroups of
${\tilde {H}}$
and H, respectively, by acting trivially on
$V^\perp _?$
and have the following exact sequence:

For any commutative ring R, we have explicitly

via the action of
$\gamma \in {\mathrm {GL}}_2({\mathcal {O}} \otimes _{\mathbb {Z}} R)$
on
$V_{\mathbb {Z}, 0}(R)$

For any
$\mathbb {Q}$
-algebra R, we also have
${\tilde {H}}_?(R) = \operatorname {GSpin}(V_{?, \mathbb {Z}})(R)$
for
$? \in \{00, 0, 1, \emptyset \}$
. Therefore, through
$\iota $
, we have

and will represent elements in
$H_0$
by their preimages in
$G_F$
. Denote
$T_0 := \iota ^{-1}(T) \subset {\tilde {H}}_0$
Then the relations among these groups can be visualized in the following diagram:

Here, the horizontal and vertical arrows are natural inclusions and surjections of algebraic groups, respectively, and the diagonal arrows are induced by
$\iota $
. Let
$B_F \subset G_F$
be the standard parabolic subgroup, and
$B_0 := \iota ^{-1}(B_F) \subset G_0$
. They can be visualized as

which gives us

via
$\iota $
. We also denote

the diagonal, which will play a crucial role in the local matching result in Section 3.5.
Now let
$P \subset H$
be the Siegel parabolic stabilizing
$V^+$
, whose Levi factor is isomorphic to
${\mathrm {GL}}(V^+)$
. Then
$P_0 := P \cap H_0 \subset H$
is the subgroup stabilizing the line
$\ell ^+$
and acting trivially on
$V_1$
. The preimage of
$P_0 H_{-1} \subset H_0$
in
${\tilde {H}}_0$
is given by
$B_0 T_0$
. Combining with (3.15), we obtain

For
$\alpha \in F^\times , \beta \in F$
, we then have
$m(\alpha ), n(\beta ) \in G_0(\mathbb {Q}) \subset {\tilde {H}}(\mathbb {Q})$
. It is easy to check that

for a Schwartz function
$\varphi \in {\mathcal {S}}(V(\mathbb {A}))$
.
3.2 Theta integral
Let
$\theta _0(g, g_1, \varphi _0)$
denote the theta function on
$[G \times H_0]$
associated to
$\varphi _0 \in {\mathcal {S}}(V_0(\mathbb {A}))$
. Suppose
$\varphi _{0, \infty }$
is in the polynomial Fock space
$\mathbb {S}(V_0(\mathbb {R}))$
(see Section 4.1). Using
$\mathbb {S}(V_0(\mathbb {R})) = \mathbb {S}(V_{00}(\mathbb {R})) \otimes \mathbb {S}(U_D(\mathbb {R}))$
, we can then restrict
$\theta _0$
to the subgroup
$[G \times H_{00}]$
, view it as a function on
$[G \times G_{00}]$
, and write

where
$\varphi _0= \sum _{j \in J} \varphi _{00, j} \varphi _{D, j}$
with
$\varphi _{00, j} \in {\mathcal {S}}(V_{00}(\mathbb {A}))$
and
$\varphi _{D, j} \in {\mathcal {S}}(U_D(\mathbb {A}))$
.
We now define

for f a bounded, integrable function on
$[G]$
. Note that the measure
$dg$
is normalized so that
$[G]$
has volume 1. In particular, for a right
$G(\hat {\mathbb {Z}})$
-invariant function
$\phi $
on
$[G]$
, we have

When
$f(g) = \vartheta _1(g, \varphi _1, \rho )$
for a bounded, integrable function
$\rho $
on
$H_1(\mathbb {Q})\backslash H_1(\mathbb {A})$
, the integral
$I_0$
above becomes

with
$\varphi = \varphi _0 \otimes \varphi _1$
.
For our purpose,
$\rho = \varrho $
will be an odd, continuous character as in (2.64), and
$\varphi = \varphi _f \varphi ^{ ( \epsilon , - \epsilon )}_\infty $
for
$\epsilon = \pm 1$
with

and
$\varphi ^\pm _\infty $
defined in (2.65). Here, we have identified
$V(\mathbb {R}) = \mathbb {R}^2 \oplus V_{1}(\mathbb {R}) \oplus V_{-1}(\mathbb {R}) \cong (\mathbb {R}^2)^{\oplus 3}$
via (2.11). For any
$\theta \in \mathbb {R}, \epsilon = \pm 1$
, we have

where
$\omega $
is the Weil representation of
$G(\mathbb {R})$
on
$V(\mathbb {R})$
. However, for
$h(\theta ) = (\kappa (\theta ), 1), h'(\theta ) = (1, \kappa (\theta )) \in H_0(\mathbb {R})$
with any
$\theta \in \mathbb {R}$
, it is easy to check that

So
$\varphi ^{(\epsilon , -\epsilon )}_\infty $
is equivariant of weight
$(\epsilon , -\epsilon )$
with respect to the connected component
${\mathrm {SO}}_2(\mathbb {R}) \times {\mathrm {SO}}_2(\mathbb {R})$
of the maximal compact of
$H_0(\mathbb {R})$
. Later, we will also consider the following integral

which is similar to
$\mathcal {I}(h_0, \varphi , \varrho )$
and well-defined as

for all
$g, h_0, h_1$
and
$\varphi _f \in {\mathcal {S}}(\hat V)$
. When
$\varphi = \varphi _0 \otimes \varphi _1$
, we have

where
$\Theta _a$
(with
$a = 1$
) is defined in (2.74).
3.3 Fourier transform and Siegel-Weil formula
We follow [Reference Gan, Qiu and TakedaGQT14] to recall the Siegel-Weil formula needed for our purpose, which goes from the split orthogonal group to the symplectic group. The range we need is in the 1st term range, and was originally proved in [Reference Kudla and RallisKR94]. Let
$\varphi = \varphi _\infty \varphi _f \in {\mathcal {S}}(V(\mathbb {A}))$
with
$\varphi _\infty $
as in (3.22) above. For
$(g, h) \in G(\mathbb {A}) \times H(\mathbb {A})$
, we have the theta function
$\theta (g, h, \varphi )$
and are interested in the value of the convergent integral
$I(h, \varphi )$
defined in (3.21).
For a rational quadratic space
$(V, (,)_V)$
, suppose
$V = U^+ + U^- + V_{\circ }$
with
$U^+, U^-$
complementary totally isotropic subspaces and
$V_{\circ } = (U^+ + U^-)^\perp $
. Let
$W = X + Y$
denote the symplectic space of rank 2 over
$\mathbb {Q}$
with the symplectic pairing
$\langle , \rangle _W$
. The rational vector space
$\mathbb {W} := V \otimes W$
is then a symplectic space with respect to the pairing

From this, we have the Fourier transform
${\mathscr {F}}_{U^+}: {\mathcal {S}}(V(\mathbb {A})) \to {\mathcal {S}}(((U^- \otimes W)+ V_{\circ })(\mathbb {A}))$
defined by

with
$\eta = (\eta _1, \eta _2) \in (U^-)^2(\mathbb {A}) \cong (U^- \otimes W)(\mathbb {A})$
and
$\eta _i \in U^-(\mathbb {A})$
. Here,
$du^+$
is the Usual Haar measure on
$U^+(\mathbb {A}) = \mathbb {A}$
. Note that we have
$(u^+, \eta _1, v_{\circ }) \in (U^+\otimes X + U^-\otimes X + V_{\circ })(\mathbb {A}) = V(\mathbb {A})$
. Note that on
${\mathcal {S}}((U^- \otimes W + V_{\circ })(\mathbb {A}))$
, the Weil representation
$\omega $
acts as

which makes
${\mathscr {F}}_{U^+}$
an intertwining map.
For V in (3.1), we can take
$U^\pm = V^\pm $
and
$V_{\circ }$
trivial with
$V^\pm $
defined in (3.3). Another possibility is to take
$U^\pm = \ell ^\pm $
and
$V_{\circ } = V_1 \oplus V_{-1}$
, which will be used in calculating the Fourier expansion of the theta integral
$I_0$
in (3.19). To simplify notations, we write

and use them to represent the Fourier transform at the finite and infinite places as well. For example,
${\mathscr {F}}_1$
is given by

for
$\varphi \in {\mathcal {S}}(V(\mathbb {A}))$
. As
${\mathscr {F}}_1$
acts as
${\mathscr {F}}_1^{\prime } \otimes \mathrm {id}$
on
${\mathcal {S}}(V(\mathbb {A})) = {\mathcal {S}}(V_0(\mathbb {A})) \otimes {\mathcal {S}}(V_1(\mathbb {A}))$
, we will abuse notation and write
${\mathscr {F}}_1 = {\mathscr {F}}_1^{\prime }$
, which acts on
${\mathcal {S}}(V_0(\mathbb {A}))$
.
For a place
$v \le \infty $
of
$\mathbb {Q}$
and corresponding local field
$k = \mathbb {Q}_v$
, recall we have the Siegel-Weil section

where
$I^H_v(s) = \operatorname {Ind}^{H(k)}_{P(k)}(|\cdot |^s)$
is the degenerate principal series. The image of
$\Phi _v$
is a submodule of
$I^H_v(0)$
denoted by
$R_v(W)$
. When
$v < \infty $
, it is known that (see [Reference Gan, Qiu and TakedaGQT14, Proposition 5.2(ii)])

It is clear that

for any
$g \in G(k)$
.
Given any
$\phi = \otimes _v \phi _v \in {\mathcal {S}}((V^-\otimes W)(\mathbb {A}))$
, we denote
$\Phi _s(\phi ) \in I^H(s)$
the standard section satisfying
$\Phi _0(\phi ) = \otimes _{v} \Phi _v(\phi _v)$
. We can then form the Eisenstein series

which has meromorphic continuation to
$s \in \mathbb {C}$
and is holomorphic at
$s = 0$
. The regularized Siegel-Weil formula by Kudla-Rallis gives then the following equality (see [Reference Gan, Qiu and TakedaGQT14, Theorem 7.3(ii)]):

As a special case of the proposition in Section 2 of [Reference MœglinMœg97], following an argument in [Reference Gelbart, Piatetski-Shapiro and RallisGPSR87], we have the following lemma.
Lemma 3.1. For any
$h \in H(\mathbb {A})$
, we have

Proof. We will show that
$ P(\mathbb {Q})\backslash H(\mathbb {Q})\cong (B(F)\backslash G(F) ) \times H_1(\mathbb {Q})$
with the map induced by (3.11). First, we have
$P(\mathbb {Q})\backslash H(\mathbb {Q}) = (P \cap (H_0 \times H_1))(\mathbb {Q})\backslash (H_0 \times H_1)(\mathbb {Q})$
. Let
$H_{-1} \subset H_0$
denote the image of
${\mathrm {SO}}(V_{-1})$
, which is isomorphic to
$H_1$
, and
$P_0 := P \cap H_0$
. Then
$P \cap (H_0 \times H_1) = P_0P_1^\Delta $
with
$P_1^\Delta \cong H_1$
the image of the diagonal embedding of
$H_1$
into
$H_{-1} \times H_1$
. From this, we obtain

Equation (3.17) then finishes the proof.
Suppose
$\varrho = \otimes _{p \le \infty } \varrho _p$
is an odd character of
$H_1(\mathbb {A})/H_1(\mathbb {Q})$
and

a totally odd character of
$\mathbb {A}_F^\times /F^\times $
, which can be viewed as a character on
$B_0(\mathbb {A})$
. Denote

From (3.14), we see that

with
$I(s, \chi )$
and
$I_v(s, \chi _v)$
defined in (2.52). Using the formula (3.31) and Lemma 3.1, we can rewrite the function
$\mathcal {I}(g_0, \varphi , \varrho )$
in (3.21) as

for
$g_0 \in G_0(\mathbb {A})$
, where
$E^{G_0}_{B_0}(s', F_{\varphi , \varrho })$
is the Eisenstein series for the standard section associated to

Note that
$F_{\varphi , \varrho , s}$
is not a standard section (i.e., it depends on s when restricted to any open compact subgroup of
$G_0(\hat {\mathbb {Q}})$
).
If
$\varrho = \otimes _{p \le \infty } \varrho _p$
and
$\varphi = \otimes _{p \le \infty } \varphi _p$
, then
$F_{\varphi , \varrho , s}$
is a product of local integrals.

Recall that
$dh_1$
is normalized so that the maximal compact subgroup of
$H_1(\mathbb {Q}_p)$
has volume 1. We have explicitly

with
$d\lambda $
the self-dual measure on
$F_p$
such that
$\int _{{\mathcal {O}}_{F_p}} d\lambda = |D|_p^{1/2}$
. From this, we see that

for all
$a \in \mathbb {Z}_p^\times $
and
$t_0 \in T_0(\mathbb {Z}_p)$
. At all but finitely many cases, the function
$F_{\varphi _p, \varrho _p, s}$
is given explicitly as follows.
Lemma 3.2. Suppose p is unramified in E and
$\varphi _p$
is the characteristic function of the maximal lattice
$V_{\mathbb {Z}} \otimes \mathbb {Z}_p \subset V_p$
. Then

for all
$g_p \in G_0(\mathbb {Z}_p)$
.
Proof. Since
$\varphi _p$
is
$G_0(\mathbb {Z}_p)$
-invariant, we can suppose
$g_p = 1$
.
If p is inert in F, then

for all
$h_1 \in H_1(\mathbb {Q}_p) = H_1(\mathbb {Z}_p) = {\mathcal {O}}_{F_p}^1 \subset {\mathcal {O}}_{F_p}^\times $
, and
$ F_{\varphi _p, \varrho _p, s}(g_p) = \int _{H_1(\mathbb {Z}_p)} dh_1 = 1$
.
If p is split in F, we have
$F_p \cong \mathbb {Q}_p^2$
,
$H_1(\mathbb {Q}_p) = \{(\alpha , \alpha ^{-1}) \in F_p: \alpha \in \mathbb {Q}_p^\times \} \cong \mathbb {Q}_p^\times $
and
$\chi _v = \chi _{v'}$
is a character of
$\mathbb {Q}_p^\times $
. Straightforward (though involved) calculations show that

For
$h_1 = (\alpha , \alpha ^{-1})$
with
$o(\alpha ) = m$
, we have

Since p is unramified in E, we have
$\varrho _p((\alpha , \alpha ^{-1})) = \epsilon ^{o(\alpha )}$
with
$\epsilon := \varrho _p((p, p^{-1})) = \chi _v(p) = \chi _{v'}(p)$
. Putting these together then gives us

This finishes the proof.
3.4 Matching global sections
The function
$\mathcal {I}(g_0, \varphi ^{(k, k')}, \varrho )$
is a Hilbert modular form of weight
$(k, k')$
. We want to suitably choose
$\varrho $
and
$\varphi _f$
and compare this function to a coherent Eisenstein series.
Let
$\chi = \chi _{E/F}$
be a Hecke character associated to a quadratic extension
$E/F$
with
$E/\mathbb {Q}$
biquadratic, and
$\varrho : \mathbb {A}_F^\times /F^\times \to \mathbb {C}^\times $
the character satisfying (2.13), whose kernel in
$H_1(\hat {\mathbb {Z}})$
is denoted by
$K_\varrho $
. Let
$\alpha \in F^\times , W_\alpha $
be the same as in Section 2.6. For our purpose, we will choose
$\phi ^{(k, k')}_\infty \in {\mathcal {S}}(W_\alpha (F \otimes \mathbb {R}))$
to be eigenfunctions of
$K_\infty = \operatorname {SL}_2(\mathbb {R})^2$
with weight
$(k, k')$
and normalized to have

The matching result we will prove is the following.
Theorem 3.3. For
$\alpha \in F^\times $
with
${\mathrm {Nm}}(\alpha ) < 0$
, given any
$\phi _f \in {\mathcal {S}}(\hat W_{\alpha })$
, there exists
$\varphi _f \in {\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})$
such that
$\omega _f(-1)\varphi _f = - \varphi _f$
for
$-1 \in H_1(\hat {\mathbb {Q}})$
, it is invariant with respect to the compact subgroup
$G(\hat {\mathbb {Z}}) T^\Delta (\hat {\mathbb {Z}})K_\varrho \subset G(\mathbb {A}) \times H_{}(\mathbb {A})$
and satisfies

Here,
$\varphi = \varphi _f \varphi ^{(\epsilon , -\epsilon )}_{\infty }$
with
$\epsilon := {\mathrm {sgn}}(\alpha _1) = -{\mathrm {sgn}}(\alpha _2)$
and
$\varphi ^{(\pm 1, \mp 1)}_{ \infty }$
defined in (3.22), and
$\phi = \phi _f \phi ^{(\epsilon , -\epsilon )}_\infty $
. In particular, we have the equality

Remark 3.4. The constants
$\Lambda (0, \chi ) = \Lambda (1, \chi ) = \frac {\sqrt {D_E/D}}{\pi ^2} L(1, \chi )$
and
$\sqrt {D_E}$
are in
$\mathbb {Q}^\times $
.
Remark 3.5. For
$L \subset W_{\alpha }(\hat {\mathbb {Q}})$
a lattice and
$\mu \in L^\vee /L$
, suppose
$\varphi _\mu \in {\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})$
satisfies (3.43) with
$\phi _f = \phi _{L + \mu }$
. Then it is easy to see that

is a (non-holomorphic) vector-valued modular form of weight 0 on
$\operatorname {SL}_2(\mathbb {Z})$
with representation
$\overline {\rho _L}$
.
Remark 3.6. If we decompose
$V_{0} = U \oplus U^\perp $
with
$U = \ell ^+ + \ell ^-$
the hyperbolic plane, then it is easy to see that
$T_0 \subset {\mathrm {SO}}(U) \subset H_0$
. Therefore, for any
$\varphi \in {\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})^{T^\Delta (\hat {\mathbb {Z}})}$
, we can write it as

such that
$\varphi _{U, j} \in {\mathcal {S}}(\hat U; \mathbb {Q}^{\mathrm {ab}})^{T^\Delta (\hat {\mathbb {Z}})}$
and
$\varphi _{U^\perp , j} \in {\mathcal {S}}(\hat U^\perp; \mathbb {Q}^{\mathrm {ab}})^{T^\Delta (\hat {\mathbb {Z}})}$
for all
$j \in J$
. This in particular implies that
$\varphi _{U^\perp , j}$
is
$T(\hat {\mathbb {Z}})$
-invariant (i.e., it is
$\mathbb {Q}$
-valued by (2.30)).
Proof of Theorem 3.3.
Suppose
$\phi = \otimes _{v \le \infty }\phi _v$
. By Theorem 3.10, there exists
$\varphi _p \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{}$
invariant with respect to
$G(\mathbb {Z}_p) T^\Delta (\mathbb {Z}_p)$
and satisfying (3.52). Furthermore,
$\varphi _p$
is the characteristic function of the maximal lattice in
$V_p$
for all but finitely many p. Therefore,
$\varphi _f := \bigotimes _{p < \infty } \varphi _p$
is in
${\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})^{(G\cdot T^\Delta )(\hat {\mathbb {Z}})}$
and satisfies

Since
$\varrho _f(-1) = {\mathrm {sgn}}(-1) = -1$
, the function
$\omega _f(-1)\varphi _f$
with
$-1 \in H_1(\hat {\mathbb {Q}})$
also satisfies these conditions, and we can replace
$\varphi _f$
by
$(\varphi _f - \omega _f(-1)\varphi _f)/2$
so that
$\omega _f(-1) \varphi _f = - \varphi _f$
. Furthermore, we have
$F_{\omega _f(h) \varphi _f, \varrho _f} = F_{\varphi _f, \varrho _f}$
for all
$h \in K_\varrho $
, and can therefore average over
$K_\varrho $
to ensure that
$\varphi _f$
is
$K_\varrho $
-invariant.
To prove (3.44), it suffices to check that
$F_{\varphi ^{(\epsilon , -\epsilon )}_\infty , \varrho _\infty }(g) = \pi ^{-1} \lambda _\alpha (\phi _\infty ^{(\epsilon , -\epsilon )})(g)$
for
$g = (g_{\tau _1}, g_{\tau _2})$
. Using

where we have used the change of variable
$x = v_1\lambda _1 - v_2\lambda _2, y = t^{-1} \lambda _1 + t \lambda _2$
, we obtain

However, we have

This finishes the proof.
The requirement that
$\varphi _f$
in Theorem 3.3 is invariant with respect to
$T^\Delta (\hat {\mathbb {Z}})$
will be important to deduce important rationality results in Section 4.3. We give a taste of such results in the following lemma.
Lemma 3.7. If
$\varphi _0 \in {\mathcal {S}}(V_0; \mathbb {Q}^{\mathrm {ab}})$
is invariant with respect to
$T^\Delta (\hat {\mathbb {Z}}) \subset ({\mathrm {GL}}_2 \times H_0)(\hat {\mathbb {Z}})$
, then
${\mathscr {F}}_1(\varphi _0) \in {\mathcal {S}}(((\ell ^- \otimes W) + V_{-1})(\hat {\mathbb {Q}}); \mathbb {Q}^{\mathrm {ab}})$
satisfies

for any
$\sigma _a \in \mathrm {Gal}(\mathbb {Q}^{\mathrm {ab}}/\mathbb {Q})$
associated to
$a \in \hat {\mathbb {Z}}^\times $
as in section 2.3. In particular, we have

for all
$r \in \hat {\mathbb {Q}}, \nu \in \hat F$
.
Proof. Using the expression for
${\mathscr {F}}_1$
in (3.29), we can write

For the second step, we moved
$\sigma _a$
inside the integral as
$\varphi _0$
is a Schwartz function and the integral is a finite sum. The third and fourth steps used (2.29) and the invariance of
$\varphi _0$
under
$(t, \iota (t)) \in T^\Delta (\hat {\mathbb {Z}})$
, respectively. Equation (3.46) now follows from (3.45) via (2.29).
3.5 Matching local sections I
The goal of this section is to prove Theorem 3.10, the non-archimedean local counterpart of the matching result 3.3. For this purpose, we fix a prime
$p < \infty $
throughout this section. The main input to Theorem 3.10 is the following surjectivity result.
Proposition 3.8. Let
$\varrho _p$
and
$\chi _p$
be as in (3.33). Then the following map

is surjective. Furthermore, if
$\Phi \in I_p^{G_0}(\chi _p)$
is valued in
$\mathbb {Q}(\zeta _{p^\infty })$
, then there exists
$\varphi \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{G(\mathbb {Z}_p)}$
satisfying
$\beta (\varphi ) = \Phi $
. Here,
$\mathbb {Q}(\zeta _{p^\infty }) \subset \mathbb {Q}^{\mathrm {ab}}$
is the subfield defined in (2.26).
Proof. Using (3.35), we can suppose
$\Phi = \otimes _{v \mid p} \Phi _v$
with
$\Phi _v \in I_v(0, \chi _v)$
. Since
$F_{\omega (g)\varphi , \varrho _p} = F_{\varphi , \varrho _p}$
for all
$g \in G(\mathbb {Z}_p)$
and
$\varphi \in {\mathcal {S}}(V_p; \mathbb {C})$
, it suffices to prove the surjectivity of
$\beta $
on
${\mathcal {S}}(V_p; \mathbb {C})$
. To do this, we will use the m-th Fourier coefficient of
$\Phi _v \in I_v(0, \chi _v)$
for
$m \in F_v$
, which is defined by

with
$\psi _v$
an additive character of
$F_v$
. For
$m = (m_v)_{v \mid p} \in F_p$
and
$\varphi \in {\mathcal {S}}(V_p; \mathbb {C})$
, we denote

with
$\psi _p := \prod _{v \mid p} \psi _v$
. Now, the
$G(F_v)$
-module
$I_v(0, \chi _v)$
can be written as

with
$R(W_\alpha )$
the image of
$\lambda _{\alpha , v}$
and irreducible. So
$I_v(0, \chi _v)$
is irreducible if and only if
$\chi _v$
is trivial. Otherwise, it is the direct sum of two irreducible submodules. Furthermore, for
$\Phi \in R(W_\alpha )$
, the coefficient
$W_m(\Phi )$
is zero unless
$m/\alpha \in \operatorname {N}_{E_{v}/F_{v}}E_{v}^\times $
. We then have two cases to consider, depending on whether
$\chi _v = \chi _{v'}$
is trivial or not.
When
$\chi _v = \chi _{v'}$
is trivial, Lemma 3.12 gives us
$\varphi $
such that
$W_m( \varphi ) \neq 0$
for some
$m \in F_p^\times $
. So for
$v \mid p$
, the restriction of
$\mathrm {im}(\beta ) \subset I_p^{G_0}(\chi _p)$
to
$G(F_v)$
gives a nonzero section in
$I_v(0, \chi _v)$
and generates a nontrivial, irreducible sub
$G(F_v)$
-module. As
$I_v(0, \chi _v)$
is irreducible, the map
$\beta $
is surjective. When
$\chi _v = \chi _{v'}$
is nontrivial, we again apply Lemma 3.12 to obtain a submodule
$R \subset I_v(0, \chi _v) = R(W_{\alpha _0}) \oplus R(W_{\alpha _1})$
from
$\mathrm {im}(\beta )$
such that
$\pi _i(R)$
is nontrivial with
$\pi _i: I_v(0, \chi _v) \to R(W_{\alpha _i})$
the projection. As
$R(W_{\alpha _i})$
is irreducible, we have
$\pi _i(R) = \pi _i(I_v(0, \chi _v))$
. Consider
$R_i := \ker \pi _i \cap R$
as a submodule of the irreducible module
$\ker \pi _i$
. As
$R(W_{\alpha _0})$
and
$R(W_{\alpha _1})$
are not isomorphic [Reference Kudla and RallisKR92, Proposition 3.4],
$R_i$
cannot be trivial for both
$i = 0, 1$
, otherwise,
$R \cong \pi _i(R) = R(W_{\alpha _i})$
. Thus,
$R_i = \ker \pi _i \subset R$
for an i, which implies
$R = I_v(0, \chi _v)$
and proves surjectivity.
When
$\Phi = \otimes _{v \mid p} \Phi _v$
has value in
$\mathbb {Q}(\zeta _{p^\infty })$
, we apply the surjectivity of
$\beta $
and the discussion in Section 2.3 to choose
$\varphi _j \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{G(\mathbb {Z}_p)}$
and
$c_j \in \mathbb {C}$
such that

satisfies
$\beta (\varphi ) = \Phi $
and J is minimal. Therefore,
$F_{\varphi , \varrho _p} = \sum _{j = 1}^J c_j F_{\varphi _j, \varrho _p}$
is valued in
$\mathbb {Q}(\zeta _{p^\infty })$
. By the minimality of J, the section
$F_{\varphi _j, \varrho _p}$
is not identically zero for all j. Therefore, the set
$\{1, c_1, \cdots , c_J\} \subset \mathbb {C}$
is linearly dependent over
$\mathbb {Q}(\zeta _{p^\infty })$
. The minimality of J then implies that
$J = 1$
and
$c_1 \in \mathbb {Q}(\zeta _{p^\infty })$
, and hence,
$\varphi \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{G(\mathbb {Z}_p)}$
.
Using this proposition, we can match any continuous function on
$G_0(\mathbb {Z}_p)$
via the map
$\beta $
. Furthermore, we can incorporate Galois action to obtain the following result.
Proposition 3.9. In the setting of Proposition 3.8, given any continuous function
$\Phi : G_0(\mathbb {Z}_p) \to \mathbb {C}$
satisfying

for all
$m(a), n(b) \in B_0(\mathbb {Z}_p), k \in G_0(\mathbb {Z}_p)$
, there exists
$\varphi \in {\mathcal {S}}(V_p; \mathbb {C})^{G(\mathbb {Z}_p)}$
such that
$F_{\varphi , \varrho _p}(g) = \Phi (g)$
for all
$g \in G_0(\mathbb {Z}_p)$
. Furthermore, if
$\Phi $
takes values in
$\mathbb {Q}(\zeta _{p^\infty })$
and satisfies

with
$ t_0 = \iota (t(a)) \in {\tilde {H}}_0(\mathbb {Z}_p), t(a) = \left (\begin {smallmatrix} a & \\ & 1 \end {smallmatrix}\right ) \in T \subset {\mathrm {GL}}_2(\mathbb {Z}_p)$
for all
$a \in \mathbb {Z}_p^\times $
and
$g \in G_0(\mathbb {Z}_p)$
, then
$\varphi \in {\mathcal {S}}(V; \mathbb {Q}(\zeta _{p^\infty }))^{G(\mathbb {Z}_p)}$
can be chosen to be
$T^\Delta (\mathbb {Z}_p)$
-invariant.
Proof. A continuous function
$\Phi $
on
$G_0(\mathbb {Z}_p)$
satisfying (3.50) can be uniquely extended to a section
$\tilde \Phi \in I^{G_0}_p(\chi _p)$
by setting

with
$g = m(a)n(b)k$
the Iwasawa decomposition of g. Therefore, the first claim is a direct consequence of Proposition 3.8.
For the second claim, we take any
$\varphi \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{G(\mathbb {Z}_p)}$
and observe that

for any
$(t, t_0) \in T^\Delta (\mathbb {Z}_p)$
with
$t = t(a), t_0 = \iota (t(a))$
and
$g \in G_0(\mathbb {Z}_p)$
. Here, we used (3.39) for the first line. By averaging
$\varphi $
over
$T^\Delta (\mathbb {Z}_p)$
, we can suppose that it is
$T^\Delta (\mathbb {Z}_p)$
-invariant. This finishes the proof.
We are now ready to state and prove the local matching result. This is just the Kudla matching principle [Reference KudlaKud03] in some sense.
Theorem 3.10. For any
$\phi _v \in {\mathcal {S}}(W_\alpha (F_v))$
with
$v \mid p$
, there exists
$\varphi _p \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{(G\cdot T^\Delta )(\mathbb {Z}_p)}$
such that

In addition, if p is unramified in E and co-prime to
$\alpha $
, and
$\phi _v$
is the characteristic function of the maximal lattice in
$W_\alpha (F_v)$
, then we can choose
$\varphi _p$
to be the characteristic function of the maximal lattice in
$V_p$
.
Proof. Suppose p and
$\alpha $
are co-prime,
$E_p/\mathbb {Q}_p$
is unramified and
$\phi _v = \operatorname {Char}({\mathcal {O}}_{E_v}), \varphi _p = \operatorname {Char}({\mathcal {O}}_{F_p} \times \mathbb {Z}_p^2 \times {\mathcal {O}}_{F_p})$
. Then it is easy to check that
$F_{\varphi _p, \varrho _p}$
and
$\prod _{v \mid p} \lambda _{\alpha , v}(\phi _v)$
are both right
$G({\mathcal {O}}_{F_p})$
-invariant. Since they are both in
$\prod _{v \mid p} I(0, \chi _v)$
, we only need to check that

by the Iwasawa decomposition of
$G(F_p)$
. This is given precisely by Lemma 3.2 and proves (3.52) for all but finitely many places.
When
$\phi _v$
is
$\mathbb {Q}$
-valued, we can use (2.29) to check that

for any
$(t, t_0) \in T^\Delta (\mathbb {Z}_p)$
with
$t = t(a), t_0 = \iota (t(a))$
and
$g \in G_0(\mathbb {Z}_p)$
. Proposition 3.9 combined with Remark 3.4 then completes the proof.
Finally, we record the two local calculation lemmas used in proving Proposition 3.8.
Lemma 3.11. Suppose
$F_p/\mathbb {Q}_p$
is non-split with valuation ring
${\mathcal {O}}_p$
, uniformizer
$\varpi $
, residue field size q, and a nontrivial additive character
$\psi $
. For a character
$\varrho $
of
$H_1(\mathbb {Q}_p) = F_p^1 \subset {\mathcal {O}}_p^\times $
, let

be the conductors of
$\psi $
and
$\varrho $
, respectively, where
$K_n:=F_p^1 \cap (1+ \varpi ^n {\mathcal {O}}_p)$
. Then

for some
$m \in {\mathcal {O}}_p^\times $
if and only if
$n(\varrho ) \le n(\psi )$
.
Proof. Let

Then
$f \in {\mathcal {S}}(F_p)$
and its Fourier transformation with respect to
$\psi $
is

First, assume that there is some
$h_0\in F_p^1$
such that
$h_0-m \in \varpi ^{n(\psi )}{\mathcal {O}}_p$
. Then

Next, we assume that there no
$h_0\in F_p^1$
such that
$h_0-m \in \varpi ^{n(\psi )}{\mathcal {O}}_p$
but some
$h_0 \in F_p^1$
with
$h_0-m \in \varpi ^{n(\psi )-1}{\mathcal {O}}_p$
. Then

Finally, if there is no
$h_0 \in F_p^1$
with
$h_0-m \in \varpi ^{n(\psi )-1}{\mathcal {O}}_p$
, then
$\hat f (m) =0$
. Now the lemma is clear.
Lemma 3.12. When
$\chi _v$
is trivial, there exists
$\phi \in {\mathcal {S}}(V_p)$
such that
$F_{\phi , \varrho _p}$
is nontrivial. When
$\chi _v$
is nontrivial, then for any
$\epsilon = (\epsilon _v)_{v \mid p}$
with
$\epsilon _v = \pm 1$
, there exists
$\phi ^\epsilon \in {\mathcal {S}}(V_p)$
and
$m^\epsilon \in F_p^\times $
such that
$W_{m^\epsilon }(\phi ^\epsilon ) \neq 0$
and
$m^\epsilon = (m^{\epsilon _v}_v)_{v \mid p}$
with
$\chi _v(m^{\epsilon _v}_v) = \epsilon _v$
.
Proof. When
$\chi _v$
is trivial, the character
$\varrho _p$
of
$H_1(\mathbb {Q}_p)$
is also trivial. Suppose
$\phi $
is the characteristic function of the maximal lattice in
$V_p$
; then the integral in (3.38) is positive at
$g = 1$
, which means
$F_{\phi , \varrho _p}$
is nontrivial.
Suppose now that
$\chi _v$
, hence
$\varrho _p$
, is nontrivial. We can suppose that
$n(\psi ) = 0$
. When
$\phi = \phi _0 \otimes \phi _1$
with
$\phi _i \in {\mathcal {S}}(V_{i, p})$
, we can apply (3.38) to write

We first assume that p is non-split and use the notation in Lemma 3.11. In this case, there is a unique place v of F above p, and
$\epsilon = \pm 1$
. For
$n \ge \max \{n(\varrho )+1, 1\}$
and
$\beta \in {\mathcal {O}}_p^\times $
, let
$ \phi _1 = \phi _{1, \beta } = \operatorname {Char}(\beta +p^n \mathcal O_p) $
and

Then

where

for some nonzero constant C. Here,
$f=1$
or
$2$
depending on whether
$F/\mathbb {Q}$
is ramified or inert at p. Using
$\varrho (h) = \varrho (h^{-1})$
, we have

If
$F_p/\mathbb {Q}_p$
is inert, then
$\chi _p$
is ramified and nontrivial when restricted to
${\mathcal {O}}_p^\times $
. By Lemma 3.11, we can find
$m_0 \in \varpi ^{ - n} {\mathcal {O}}_p^\times $
such that

as
$ n(\psi (\varpi ^{ - n} \cdot )) = n \ge n(\varrho )$
. We can choose
$\beta = \beta ^\pm $
such that
$\chi (m_0/(\beta ^\pm )') = \pm 1$
. Then taking
$\phi ^{\pm } = \phi _{0} \otimes \phi _{1, \beta ^\pm }$
proves the Lemma. If
$F_p/Q_p$
is ramified, then
$E_v/F_v$
is inert and
$\chi _v(\varpi ) = -1, \chi _v \mid _{{\mathcal {O}}_p^\times } = 1$
. Again by Lemma 3.11, we can find
$m_j \in \varpi ^{- n + j} {\mathcal {O}}_p^\times $
for
$j = 0, 1$
such that

as
$ n(\psi (m_j \cdot )) = n - j \ge n(\varrho )$
. Therefore,
$\phi ^\pm = \phi _0 \otimes \phi _{1, 1}$
satisfies the Lemma.
Finally, we come to the case when
$p = v_1v_2$
splits and
$\eta := \chi _{v_1} = \chi _{v_2}$
is nontrivial. In this case,
$F_p=F_{v_1} \times F_{v_2} =\mathbb {Q}_p^2$
and
$\eta = \varrho _p$
is a character of
$\mathbb {Q}_p^\times \cong H_1(\mathbb {Q}_p)$
. For
$m \in F_p$
, we write
$m=(m_1, m_2)$
with
$m_j \in \mathbb {Q}_p$
and

So we take
$\phi _1 = \phi _{1, 1} \otimes \phi _{1,2}$
with
$\phi _{1, j} \in {\mathcal {S}}(\mathbb {Q}_p)$
and
$\phi _{0} \in {\mathcal {S}}(M_2(\mathbb {Q}_p))$
. Simple calculation gives us

Taking
$\phi _{1, j} =\operatorname {Char}(1+p^n\mathbb {Z}_p)$
,
$n \ge \max \{ 1, n(\eta )\}$
, and

the same calculation as above gives

for some nonzero constant C.
When
$\eta $
is ramified, we take
$n=n(\rho )$
,
$m_1=p^l$
with
$l \ge n$
and
$m_2 =m_0 p^{-n}\in p^{-n}\mathbb {Z}_p^\times $
and obtain (write
$o(x) =\mbox {ord}_{p}x$
)

for some
$m_2$
.
When
$\eta $
is unramified, we have
$n(\eta ) = 0$
and take
$n = 1$
. For
$m_j \in p^{-1} \mathbb {Z}_p^\times $
,
$j=1, 2$
, we have

If we sum this over
$m_j \in p^{-1}\mathbb {Z}_p^\times $
, then the result is nonzero. So there exists
$m_j \in p^{-1}\mathbb {Z}_p^\times $
such that
$W_{(m_1, m_2)}(\phi _0 \otimes \phi _1) \neq 0$
. For
$m_j \in \mathbb {Z}_p^\times $
,
$j=1, 2$
, we have

as
$\eta (p) =-1$
. Replacing
$\phi _1$
by
$\phi _1^{\prime }=\operatorname {Char}(1+p^n \mathbb {Z}_p, p + p^n \mathbb {Z}_p)$
, the same calculation gives

when
$m_j \in \mathbb {Z}_p^\times $
and
$m_{3-j} \in p^{-1}\mathbb {Z}_p^\times $
. This completes the proof.
3.6 Matching local sections II
In order to give the factorization result, we also need a matching result involving the following local sections with the s parameter. When
$p = vv'$
splits in F, we define a slightly modified section

This function on
$G_0(\mathbb {Q}_p)$
depends on the choice of
$v \mid p$
and has the following property.
Lemma 3.13. When
$|s| < 1$
, the integral in (3.54) converges absolutely and defines a rational function in
$p^{s}$
defined over
$\mathbb {Q}(\zeta _{p^\infty })(\varphi _p)$
. Furthermore, when restricted to the first (resp. second) components of
$G_0(\mathbb {Q}_p) \cong G(F_p) \cong G(F_v) \times G(F_{v'})$
, it defines a section in
$I(s, \chi _v)$
(resp.
$I(-s, \chi _{v'})$
).
Proof. For the first claim, one can suppose
$\omega (g_p){\mathscr {F}}_1(\varphi _p)$
is the characteristic function of

with
$C_1 \subset \mathbb {Q}_p^2$
a compact subset and
$a_j, b_j \in \mathbb {Z}, r_j, t_j \in \mathbb {Q}_p$
. As in Lemma 3.2, the integral defining
$F_{\varphi _p, \varrho _p, v, s}(g_p)$
is given by

Suppose
$(0, 0) \in C_1$
; otherwise, the integral vanishes identically. When
$|\alpha |_p \ge p^N$
for N sufficiently large, we have
$|\alpha ^{-1} \lambda _2|_p$
very small for all
$\lambda _2 \in p^{a_2}\mathbb {Z}_p + r_2$
. Therefore, when
$t_2 \not \in p^{b_2}\mathbb {Z}_p$
or
$r_1 \not \in p^{a_1}\mathbb {Z}_p$
, the integral over those
$\alpha $
with
$|\alpha |_p \ge p^N$
is zero. When
$t_2 \in p^{b_2}\mathbb {Z}_p$
and
$r_1 \in p^{a_1}\mathbb {Z}_p$
, we have

which converges when
$|s| < 1$
and defines a rational function in
$p^s$
. The same argument takes care of the case when
$|\alpha |_p$
is sufficiently small. This proves the first claim.
For the second claim, it is clear from the definition that
$F_{\varphi _p, \varrho _p, v, s}$
is locally constant as
$\varphi _p$
is a Schwartz function. For the transformation property, we have

for
$\alpha = (\alpha _1, \alpha _2) \in F_p^\times = (\mathbb {Q}_p^\times )^2$
. A change of variable plus
$\varrho (\alpha /\alpha ') = \chi _v(\alpha ) \chi _{v'}(\alpha )$
and
$|\alpha |_v = |\alpha _1|_p, |\alpha '|_v = |\alpha _2|_p$
then finishes the proof.
Now, we will extend the matching result in Theorem 3.10 to standard sections.
Theorem 3.14. In the setting of Theorem 3.10, suppose
$p = vv'$
splits and let
$\lambda _{\alpha , v, s}(\phi _v) \in I_v(s, \chi _v)$
denote the standard section associated to
$\lambda _{\alpha , v}(\phi _v) \in I_v(0, \chi _v)$
for
$\phi _v \in {\mathcal {S}}(W_\alpha (F_v))$
. For any
$r \in \mathbb {N}$
, there exists
$\varphi ^{}_p \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{(G\cdot T^\Delta )(\mathbb {Z}_p)}$
such that

for all
$g = (g_v, g_{v'}) \in G_0(\mathbb {Q}_p)$
, where
$\mathcal {L}(s) := (1 - p^{-2}) |D_E|_p^{-1} L(1+s, \chi _v) L(1-s, \chi _{v'})$
Remark 3.15. If
$\varrho _p$
is unramified and
$\varphi _p$
is the characteristic function of the maximal lattice in
$V_p$
, then

by a similar calculation as in Lemma 3.2, and
$F_{\varphi _p, \varrho _p, v, s}/(L(1+s, \chi _v)L(1-s, \chi _{v'})(1-p^{-2}))$
is already the standard section
$\lambda _{\alpha , v, s}(\phi _v)(g_v) \lambda _{\alpha , v, -s}(\phi _{v'})(g_{v'})$
.
Proof. For
$ \varphi \in {\mathcal {S}}(V_p; \mathbb {C})$
and
$g \in G_0(\mathbb {Q}_p)$
, we write

It is easy to check from definition that
$F_{\varphi , \varrho _p, v, s}: G_0(\mathbb {Z}_p) \to \mathbb {C}$
satisfies (3.50), and hence, so does the function
$a_n(\varphi , g)$
for all
$n \ge 0$
. Now we define

with
$(p, p^{-1}) \in H_1(\mathbb {Q}_p)$
. An easy induction shows that
$ F_{\varphi ^{(n)}, \varrho _p, v, s} = \tfrac {1}{n!}(\tfrac {1 - p^{2s}}{2})^n F_{\varphi ^{}, \varrho _p, v, s}, $
which implies

When
$\varphi \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))$
is
$(G\cdot T^\Delta )(\mathbb {Z}_p)$
-invariant, so is the function
$\varphi ^{(n)} \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))$
. Furthermore, the function
$a_n(\varphi , \cdot ): G_0(\mathbb {Z}_p) \to \mathbb {Q}(\zeta _{p^\infty })$
satisfies conditions (3.50) and (3.51). By Proposition 3.9, there exists
$\varphi _{n} \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{(G\cdot T^\Delta )(\mathbb {Z}_p)}$
such that

for all
$k \in G_0(\mathbb {Z}_p)$
.
Now we prove the theorem by induction on r. The case
$r = 1$
is just the content of Theorem 3.10. Note that
$|D|_p = 1$
when p splits in F. Now suppose we have
$\varphi $
satisfying (3.56) for some
$r \ge 1$
. As

is a standard section, it satisfies
$ \Phi _s(k) = \Phi _0(k) $
when
$k \in G_0(\mathbb {Z}_p)$
. So in that case, we have

with
$c_r := (-\log p)^{-r} \partial ^r_s \mathcal {L}(s)\mid _{s = 0}$
rational. If we set

then equations (3.58) and (3.59) give us

So
$\tilde \varphi $
satisfies the claim for
$r +1$
. This completes the proof.
Now, we can state a consequence of the matching result in Theorem 3.14.
Proposition 3.16. For matching sections
$\varphi _p \in {\mathcal {S}}(V_p; \mathbb {Q}(\zeta _{p^\infty }))^{G(\mathbb {Z}_p)}$
and
$\{\phi _v \in {\mathcal {S}}(W_\alpha (F_v)): v \mid p\}$
as in Theorem 3.14 with
$r = 2$
, we have

for all
$t \in F_p^\times $
. Furthermore, when
$p = vv'$
is a split prime and
$W_{t, v}^*(\phi _v) = 0$
, then we have

Remark 3.17. Since
$o_{v'}(h) = - o_v(h)$
for all
$h \in H_1(\mathbb {Q}_p)$
, the left-hand side of (3.61) gets a minus sign if
$o_v(h)$
is replaced by
$o_{v'}(h)$
on the right-hand side.
Proof. To prove (3.60). we apply the definition of
$W^*_{t, v}$
in (2.54) and Theorem 3.10 to obtain

Now the right-hand side of (3.60) can be rewritten as

For the second line, we have used
$\omega (m(a))\varphi _p = \varphi _p$
for all
$a \in \mathbb {Z}_p^\times $
since
$\varphi _p$
is
$G(\mathbb {Z}_p)$
-invariant. Equation (3.60) now follows from applying Proposition 3.18 to
$\varphi = \omega _p(h_1)\varphi _p$
.
We now prove (3.61). Let
$ \Phi _s \in I(s, \chi _v)I(-s, \chi _{v'})$
be the standard sections extending
$ \lambda _{\alpha , v}(\phi _v)\lambda _{\alpha , v'}(\phi _{v'})$
. By Theorem 3.14 with
$r = 2$
, we have

Therefore, using
$W^{*}_{t, v}(\phi _v) = 0$
, we have

Applying Proposition 3.18 and continuing as in the second half of the proof of (3.60) then proves (3.61).
We end this section with the following technical result used in the proof of the previous proposition.
Proposition 3.18. For any
$\varphi \in {\mathcal {S}}(V_p; \mathbb {C})^{G(\mathbb {Z}_p)}$
and
$t \in F_p^\times $
, we have

Proof. Since the left-hand side is essentially the Fourier transform of
$(\omega (wn(\beta )){\mathscr {F}}(\varphi _p))(0)$
as a function of
$\beta \in F_p$
, it suffices to calculate the inverse Fourier transform of the right-hand side, though we need to be careful about the singularity of right-hand side when
$t \not \in F_p^\times $
. To take care of this, we define

for
$\epsilon> 0$
and
$t \in F_p$
. Note that
$|t| := \min \{|t_1|, |t_2|\}$
when
$t = (t_1, t_2) \in F_p = \mathbb {Q}_p^2$
, Given any fixed
$t \in F_p^\times $
, the limit
$\lim _{\epsilon \to 0} G_\epsilon (t, \varphi )$
exists and is the right-hand side of (3.62). Also for any fixed
$\epsilon> 0$
, the function
$G_\epsilon (t, \varphi )$
is a Schwartz function on
$F_p$
. Its inverse Fourier transform is given by

where
$D_\epsilon \subset F_p$
is the
$\epsilon $
-neighborhood of
$0$
. Note that

where
$\varphi _\beta := \omega (w_2 n(\beta ))\varphi $
and
$w_i \in H(\mathbb {Q})$
is defined in (3.6). Therefore,
$\hat G_\epsilon (\beta , \varphi )$
is given by

with
$s = -\tilde t$
and
$ds = d \tilde t$
. Using the
$G(\mathbb {Z}_p)$
-invariance of
$\varphi $
, we have

for all
$u \in \mathbb {Z}_p^\times , n \in \mathbb {Z}$
and
$a \in (p^n\mathbb {Z}_p)^2 - (p^{n+1}\mathbb {Z}_p)^2 $
. Applying this, we can evaluate the first part as

Then for any fixed
$t \in F_p^\times $
, we have

Since
${\mathscr {F}}_1(\varphi )$
is a Schwartz function, we can replace the domain
$\mathbb {Q}_p^\times \times D_{\epsilon /|u|}$
with a compact open subset independent of
$\beta $
and interchange the order of integration to compute the integral over
$\beta $
first, which yields

When
$\epsilon $
is sufficiently small, we have
$t \not \in D_\epsilon $
and
$t + su \neq 0$
, in which case
$E_\epsilon = 0$
. This finishes the proof.
4 Doi-Naganuma lift of the deformed theta integral
In this section, we will define and study the properties of the function
$\tilde {\mathcal {I}}$
discussed in the introduction. In particular, we will calculate its Fourier coefficients and images under lowering differential operators. The actions of differential operators follow from those on the theta kernel, which are given in Section 4.1. The Fourier expansion computations are carried out in Section 4.2, with the main result being Proposition 4.7. Section 4.3 contains rationality results about theta lifts that will be needed to handle the error term mentioned in Section 1.3.
Choose
$\varphi ^{(1, 1)} := \varphi _f\varphi ^{(1, 1)}_{\infty } \in {\mathcal {S}}(V(\mathbb {A}))^{K_\varrho }$
with
$\varphi ^{(1, 1)}_\infty := \varphi ^{(1, 1)}_{0, \infty } \otimes \varphi ^+_\infty \in {\mathcal {S}}(V_0(\mathbb {R})) \otimes {\mathcal {S}}(V_1(\mathbb {R}))$
and

For
$\tilde {\varrho }_C$
as in (2.76), we can define

We will now analyze various properties of this integral.
4.1 Lowering operator action
To calculate the action of differential operators on
$\tilde {\mathcal {I}}$
, it suffices to understand the effect on
$\varphi _0$
via the Weil representation, which can be done in the Fock model. For this, we follow the appendices in [Reference Funke and MillsonFM06] and [Reference LiLi22] (see also [Reference Kudla and MillsonKM90]).
We identify
$(V_0(\mathbb {R}), Q) = (M_2(\mathbb {R}), \det ) \cong \mathbb {R}^{2, 2}$
with the basis

which identifies
${\mathcal {S}}(\mathbb {R}^{2, 2}) \cong {\mathcal {S}}(V_0(\mathbb {R}))$
and gives us

The polynomial Fock space is the subspace
$\mathbb {S}(\mathbb {R}^{2, 2}) \subset {\mathcal {S}}(\mathbb {R}^{2, 2})$
spanned by functions of the form
$\prod _{1 \le j \le 4} D_j^{r_j} \varphi ^\circ $
for
$r_j \in \mathbb {N}_0$
, where
$\varphi ^\circ \in {\mathcal {S}}(\mathbb {R}^{2, 2})$
is the Gaussian

and
$D_r$
are operators on
${\mathcal {S}}(\mathbb {R}^{2, 2})$
defined by

There is an isomorphism
$\iota : \mathbb {S}(\mathbb {R}^{2, 2}) \to \mathscr {P}(\mathbb {C}^4) = \mathbb {C}[\mathfrak {z}_1, \mathfrak {z}_2, \mathfrak {z}_3, \mathfrak {z}_4]$
such that
$ \iota ( \varphi ^\circ ) = 1$
,
$D_r$
acts as
$(-1)^{\lfloor (r -1)/2\rfloor } i \mathfrak {z}_r $
. We now set

Then using (4.4), the Schwartz functions
$\varphi ^{(\epsilon , \epsilon ')}_{0, \infty } \in {\mathcal {S}}(V_0(\mathbb {R}))$
in (3.22) and (4.1) become

Let
$(W, \langle , \rangle )$
be the
$\mathbb {R}$
-symplectic space of dimension 2, and
$\mathbb {W} := V_0(\mathbb {R}) \otimes W$
the symplectic space with the skew-symmetric form
$(,)\otimes \langle , \rangle $
. The Lie algebra
${\mathfrak {sp}}(\mathbb {W} \otimes \mathbb {C})$
acts on
${\mathcal {S}}(V_0)$
through the infinitesimal action induced by
$\omega $
, which we also denote by
$\omega $
. In
${\mathfrak {sp}}(\mathbb {W} \otimes \mathbb {C})$
, we have the subalgebra
${\mathfrak {sp}}(W \otimes \mathbb {C}) \times \mathfrak {o}(V_0 \otimes \mathbb {C})$
. Through
$\iota $
, the elements
$L, R \in {\mathfrak {sl}}_2(\mathbb {C}) \cong {\mathfrak {sp}}(W \otimes \mathbb {C})$
defined in (4.11) act on
$\mathbb {C}[\mathfrak {z}_1, \mathfrak {z}_2, \mathfrak {z}_3, \mathfrak {z}_4]$
as (see [Reference Funke and MillsonFM06, Lemma A.2])

Using the isomorphism

we see that the elements
$L_1 = (L, 0), L_2 = (0, L_2), R_1 = (R, 0), R_2 = (0, R)$
in
$ {\mathfrak {sl}}_2(\mathbb {C})^2$
act on
$\mathbb {C}[\mathfrak {z}_1, \mathfrak {z}_2, \mathfrak {z}_3, \mathfrak {z}_4]$
through
$\iota $
as (see [Reference Funke and MillsonFM06, Lemma A.1])

For convenience, we slightly abuse notation and write
$L, R, L_j, R_j$
for their corresponding actions on
$\mathscr {P}(\mathbb {C}^4)$
.
When we consider the decomposition
$V_0 = V_{00} \oplus U_{D}$
in (3.2), the map
$\iota $
induces
$\mathbb {S}(V_{00}(\mathbb {R})) \cong \mathscr {P}(\mathbb {C}^3) = \mathbb {C}[\mathfrak {z}_1, \mathfrak {z}_3, \mathfrak {z}_4]$
and
$\mathbb {S}(U_{D}(\mathbb {R})) \cong \mathscr {P}(\mathbb {C}) = \mathbb {C}[\mathfrak {z}_2]$
. For
$a, b, c \in \mathbb {N}_0$
, we also define
$\varphi ^{(a, b)}_{00, \infty } \in \mathbb {S}(V_{00}(\mathbb {R}))$
and
$\varphi ^c_{D, \infty } \in \mathbb {S}(\mathbb {R})$
by

For
$r \in \mathbb {N}_0$
, we have the operators
$\mathrm {RC}_{r}, \widetilde {\mathrm {RC}}_{r}$
defined in (2.8) that also act on
$ \mathscr {P}(\mathbb {C}^4)$
. They are related by the following lemma.
Lemma 4.1. In the notations above, we have

where
$p_r(X, Y) := - (\tilde {Q}_r(X, Y) - (X + Y)^r)/Y \in \mathbb {Q}[X, Y]$
for all
$r \in \mathbb {N}_0$
.
Proof. It is easy to check that

This leads directly to the second equation in (4.11) from the definition. To prove the first equation, it is enough to verify

which follows from (2.6).
Proposition 4.2. Let
$\phi _f \in {\mathcal {S}}(W_\alpha (\hat F))$
and
$\varphi _f \in {\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})$
be matching sections as in Theorem 3.3 and denote
$\epsilon := {\mathrm {sgn}}(\alpha _1) = - {\mathrm {sgn}}(\alpha _2)$
. Then for
$\tilde {\mathcal {I}}$
defined in (4.2), we have

where
$\mathcal {I}_f$
is defined in (3.23) and
$\varphi ^{(\pm 1, \mp 1)} = \varphi _f \otimes \varphi ^{(\pm 1, \mp 1)}_{\infty } \in {\mathcal {S}}(V(\mathbb {A}))$
is defined in (3.22).
Proof. Suppose
$\varphi _f = \varphi _{0, f} \otimes \varphi _{1, f}$
and denote
$\varphi ^\pm _1 = \varphi _{1, f} \varphi ^\pm _\infty , \varphi _{0}^{(k, k')} = \varphi _{0, f}\varphi ^{(k, k')}_{0, \infty }$
. Then

by Lemma 4.1 and (4.7). Now applying Stokes’ theorem and Theorem 2.7 gives us

with
$\epsilon = \pm 1$
. Since
$R \vartheta _1(g', \varphi _1^-, \varrho ) = 0$
, we can apply Stokes’ theorem, Lemma 4.1 and Theorem 3.3 to obtain

Putting these together finishes the proof.
To understand the first term on the right-hand side of (4.13), recall the decomposition for
$\theta _0(g', g^\Delta , \varphi _0)$
in (3.18) when
$\varphi _{0, \infty } \in \mathbb {S}(V_0)$
. This allows us to define

for
$\varphi _0 = \varphi _{00} \otimes \varphi _{D}$
with
$\varphi _{00} \in {\mathcal {S}}(V_{00}(\mathbb {A})), \varphi _D \in {\mathcal {S}}(U_D(\mathbb {A}))$
, and
$R^{\prime }_j$
the raising operator on
$g^{\prime }_j$
. In the Fock model,
$R^{\prime }_1$
, resp.
$R^{\prime }_2$
, acts on
$\mathbb {C}[\mathfrak {z}_1, \mathfrak {z}_3, \mathfrak {z}_4]$
, resp.
$\mathbb {C}[\mathfrak {z}_2]$
, as

This definition also extends by linearity to all
$\varphi _0 \in {\mathcal {S}}(V_0(\mathbb {A}))$
satisfying
$\varphi _{0, \infty } \in \mathbb {S}(V_0(\mathbb {R}))$
. We now record the following lemma.
Lemma 4.3. For
$r \in \mathbb {N}_0$
, denote
$r_0 := \lfloor r/2\rfloor $
. Then

Proof. Suppose
$\varphi _{0, f} = \varphi _{00, f} \otimes \varphi _{D, f}$
. Then equations in (4.7) imply

From (4.12) and the definition of
$P_r$
in (2.5), we have

Substituting the definition (4.10) finishes the proof.
The following technical lemma concerns a change of regularized integrals and follows from the proof of Lemma 5.4.3 in [Reference LiLi22].
Lemma 4.4. Given
$\varphi _{i, f} \in {\mathcal {S}}(\hat V_i)$
with
$i = 0, 1$
, let
$\Gamma \subset \operatorname {PSL}_2(\mathbb {Z}) \subset G_{00}(\mathbb {Q})$
be a congruence subgroup that acts trivially on
$\varphi _{0, f}$
. For any
$a \ge 1, b \ge 0$
and
$f \in M^!_{-2b}(\Gamma )$
, we have

4.2 Fourier expansion of
$\tilde {\mathcal {I}}$
To evaluate the Fourier expansion of
$\tilde {\mathcal {I}}$
in (4.2), we change to a mixed model of the Weil representation using the partial Fourier transform
${\mathscr {F}}_1$
defined in (3.28).
Throughout the section, we write

for
$(z_1, z_2) \in \mathbb {H}^2$
with
$z_i = x_i + iy_i$
and
$g_{\boldsymbol \tau } \in G(\mathbb {R})$
with
${\boldsymbol \tau } = {\boldsymbol u} + i{\boldsymbol v} \in \mathbb {H}$
. Then (3.27) implies

Also when
$\varphi _{0, \infty } = \varphi ^{(k, k')}_{0, \infty }$
with
$k, k' = \pm 1$
as given in (3.22) and (4.1), we have

with
$r \in \mathbb {R}, \nu = (\nu _1, \nu _2) \in \mathbb {R}^2$
. After applying Poisson summation and unfolding, we can rewrite the theta kernel
$\theta _0(g, g_0, \varphi _0)$
as

For a bounded, integrable function f on
$[G]$
such that
$\theta _0(g, g_0, \varphi _0) f(g)$
is right
$\operatorname {SL}_2(\hat {\mathbb {Z}}){\mathrm {SO}}_2(\mathbb {R})$
-invariant, we have

which can be written as
$ I_0(g_0, \varphi _0, f) = I^0_0(g_0, \varphi _0, f) + I^+_0(g_0, \varphi _0, f)$
with (see, for example, equation (4.2) in [Reference KudlaKud16])

Using the
$\operatorname {SL}_2(\hat {\mathbb {Z}})$
-invariance of
$\theta _0(g, g_0, \varphi _0)f(g)$
, we can rewrite for any
$N \in \mathbb {N}$

with
$\Gamma _\infty ^N := \{n(Nb): b \in \mathbb {Z}\} \subset \Gamma _\infty $
.
For our purpose, we are interested in the case when
$f(g) = \vartheta _1(g, \varphi _1, \rho )$
with
$\rho \in \{ \varrho , \tilde {\varrho }_C\}, \varphi _1 = \varphi ^\pm _1$
and
$\varphi _0 = \varphi _0^{(k, k')}$
for
$k, k' = \pm 1$
. In that case, we have
$I_0(g_0, \varphi _0, f) = \mathcal {I}(g_0, \varphi , \rho )$
with
$\varphi = \varphi _0 \otimes \varphi _1$
, and denote

This can be extended by
$\mathbb {Q}$
-linearity to all
$\varphi = \varphi _f \varphi _\infty \in {\mathcal {S}}(V(\mathbb {A}))$
with
$\varphi _f \in {\mathcal {S}}(\hat V)$
and
$\varphi _\infty \in \mathbb {S}(V(\mathbb {R}))$
.
The constant term
$I^0_0$
in the Fourier expansion of
$I_0(g_0, \varphi _0, f)$
is independent of
$x_1, x_2$
and can be evaluated by the change of model in section 3.3. For our purpose, we will state a decay result needed to prove Theorem 5.1
Lemma 4.5. Suppose there is
$s \in \mathbb {R}$
such that
$|f(g_\tau )| \ll v^s$
for all
$\tau $
in the usual fundamental domain of
$\operatorname {SL}_2(\mathbb {Z})\backslash \mathbb {H}$
. Then

for any
$a, b, c \in \mathbb {N}_0$
satisfying
$a + b + c \ge 1$
. When
$a = b= c = 0$
, the limit exists.
Remark 4.6. It is easy to check that
$f(g) = \vartheta (g, \varphi _1, \tilde {\varrho }_C)$
fulfills the condition in the lemma.
Proof. Let
${\mathcal {F}}$
denote the fundamental domain. Then we can use (4.18) to obtain

Since
${\mathscr {F}}_1(\varphi _{0, f})$
is a Schwartz function, we can suppose the sum over
$\lambda \in F$
is replaced by a sum over the translate of a lattice. For the integral on the second line, we can trivially estimate it by

From this, we see that
$| y^{-1}I^0_0((g_z, g_z), \varphi _0, f)|$
is bounded independent of y, and the second claim holds. This also gives the first claim for
$a = b = 0$
and
$c \ge 1$
. The other cases follow from first applying
$\partial _{y_1}^a\partial _{y_2}^b$
to
$ \frac { I^0_0(g_0, \varphi _0, f)}{\sqrt {y_1y_2}}$
and then conducting the same estimate.
We will now evaluate the non-constant term
$\mathcal {I}^+$
. Let
$\varrho $
and
$\tilde {\varrho }_C$
be as in (2.76), and
$K_\varrho \subset H_1(\hat {\mathbb {Z}}), \Gamma _\varrho \subset H_1^+(\mathbb {Q})$
be as in (2.14) and (2.67), respectively. For
$f(g) = \vartheta _1(g, \varphi ^-_1, \varrho )$
and
$\varphi _{0, \infty } = \varphi ^{(\pm 1, \mp 1)}_{0, \infty }$
defined in (3.22), we can apply (4.18), (4.19) and (4.21) to obtain

if
$\mp r \nu> 0$
, and zero otherwise. After the change of variable
$t = r \nu '$
, we have

In the case of
$\mathcal {I}^+(g_0, \varphi , \tilde {\varrho }_C)$
, we have the following result.
Proposition 4.7. Given
$\phi _f \in {\mathcal {S}}(\hat W_\alpha )$
, let
$\varphi _f \in {\mathcal {S}}(\hat V)^{G(\hat {\mathbb {Z}})T^\Delta (\hat {\mathbb {Z}})K_\varrho }$
be a matching section as in Theorem 3.3. For
$\tilde {\varrho }_C$
as in (2.76), we have

where
$\tilde {c}_t(\varphi _f, \varrho ) \in \mathbb {C}$
and
$e_t(\varphi _f; \cdot ): \mathbb {R}_{> 0}^2 \to \mathbb {C}$
are given in (4.27) and (4.28) below. There is a constant
$M \in \mathbb {N}$
such that
$\tilde {c}_t(\varphi _f, \varrho ) = 0$
when
$ t \not \in M^{-1}{\mathcal {O}}$
.
Let

be a finite set of primes, and then there exists
$\kappa \in \mathbb {N}$
and
$\beta (t, \phi _f) \in F^\times $
such that
$\tilde {c}_t(\varphi _f, \varrho ) = -\frac {2}{\kappa }\log |\beta (t, \phi _f)/\beta (t, \phi _f)'|$
and

when
$\mathrm {Diff}(W_\alpha , t) \subset S_C^c$
with
$\tilde {W}_t$
defined in (2.62). Furthermore, the function
$e_t$
satisfies

for all
$a, b, c \in \mathbb {N}_0$
.
Remark 4.8. Note that we have

Proof. Suppose
$\varphi _f = \varphi _{0, f} \otimes \varphi _{1, f}$
, as the general case follows by linearity. We first prove (4.23). Using (4.20), it is enough to evaluate
$\mathfrak {F}_{r, \nu }(\varphi ^{(1, 1)}_{0, \infty })(z_1, z_2, \vartheta _1(\cdot , \varphi ^+_1, \tilde {\varrho }_C))$
. If we set
$t := r\nu '$
, then we have by (4.18), (4.19), and Theorem 2.7

with
$\tilde {c}_{r, \nu '}(\varphi , \varrho )$
and
$e_{r, \nu }(\varphi; y, y')$
given by

Here, we have set

So if we set


then equation (4.23) holds by (4.20). Since
$\varphi _f$
is a Schwartz function, the sum defining
$\tilde {c}_t$
is finite and equals to 0 when
$t \not \in M^{-1} {\mathcal {O}}$
for some
$M \in \mathbb {N}$
depending only on
$\varphi _f$
.
To prove (4.25), notice that

By Theorem 3.10 and Lemma 3.7, there exists
$c \in \mathbb {N}$
such that
$2c {\mathscr {F}}_1(\varphi _f)((0, r), \nu , \lambda ) \in \mathbb {Z}$
for all
$r \in \hat {\mathbb {Q}}$
and
$\nu , \lambda \in \hat F$
. Then we can write

For any split rational prime
$p = \mathfrak {p} \mathfrak {p}'$
with any
$\mathfrak {p} \not \in S_C$
, we have

since
$t \gg 0, \operatorname {ord}_{\mathfrak {p}}(\beta ) = \operatorname {ord}_{\mathfrak {p}}(\xi ^{-1} \beta )$
and
$\varrho _f(h) = {\mathrm {sgn}}(h)$
, where
$h_1 = -\xi h \in H_1(\hat {\mathbb {Q}})$
and
$\varrho (h_1) = - \varrho (\xi )$
. The last step follows from (2.71).
Notice that the quantity above factors as the following product of sums of local integrals

Applying (3.60) turns the first line on the right-hand side into
$2 \prod _{v < \infty ,~ v \nmid p} W^*_{t, v}(\phi _v)$
. If this is nonzero, then
$\mathrm {Diff}(W_\alpha , t)$
is either
$\{\mathfrak {p}\}$
or
$\{\mathfrak {p}'\}$
as it has odd size. If
$\mathrm {Diff}(W_\alpha , t) = \{\mathfrak {p}\}$
, then
$W^*_{t, \mathfrak {p}}(\phi _{\mathfrak {p}}) = 0$
and Proposition 3.16 tell us that the second line on the right-hand side becomes

as
$\operatorname {ord}_{\mathfrak {p}} (h_{1, p}^{-1} t') = \operatorname {ord}_{\mathfrak {p}} (h_{1, p}^{-1}) + \operatorname {ord}_{\mathfrak {p}}( t')$
. This gives us

Repeating the above argument together with Remark 3.17, we obtain
$\kappa ^{-1} \operatorname {ord}_{\mathfrak {p}'} \beta (t, \varphi _f)= -\tilde {W}_t(\phi _f)/2$
. Putting this together gives us (4.25), where the case with
$\mathrm {Diff}(W_\alpha , t) = \{\mathfrak {p}'\}$
is obtained similarly.
Now we will prove (4.26). Since
$\varphi $
has compact support, the summation over
$\xi $
and
$\beta $
in
$e_{r, \nu }$
and the summation over r in (4.28) are finite sums, it suffices to establish (4.26) with
$e_t$
replaced by
$e^*_{r, \nu }$
and
$e^\dagger _{r, \nu , \beta }$
with
$r> 0$
. For any fixed
$C, \epsilon> 0$
,
$s \in \mathbb {R}$
and
$a, b, c \in \mathbb {N}_0$
, we have

when B is in a compact subset of
$\mathbb {R}$
and
$A> 0$
is sufficiently large. Furthermore, it is easy to see that there exists
$C> 0$
such that

for all
${\boldsymbol v}> 0$
. Combining these then proves (4.26).
4.3 Rationality of theta lifts
Recall that the rational quadratic space
$V_\alpha $
is the restriction of scalars of the F-quadratic space
$W_\alpha $
. The following result shows that the Millson theta lift preserves rationality.
Proposition 4.9. Let
$f = \sum _{\mu \in L_{}^\vee /L_{}} f_\mu \mathfrak {e}_\mu \in M^!_{-2r, \rho _{L_{}}}$
be weakly holomorphic for some
$r \in \mathbb {N}$
and lattice
$L_{} \subset V_{\alpha }$
. For
$\varphi ^{(1, r)}_{} = \varphi ^{(1, r)}_{00, \infty } \varphi _f$
with
$\varphi _{ f} \in {\mathcal {S}}(\hat V_{00}; \mathbb {Q}^{\mathrm {ab}})^{T^\Delta (\hat {\mathbb {Z}})}$
, let
$\Gamma (N) \subset \operatorname {SL}_2(\mathbb {Z})$
be a congruence subgroup contained in
$\ker (\rho _L)$
that fixes
$\varphi _f$
. The following regularized integralFootnote
4

defines a weakly holomorphic modular form of weight
$-r + 1/2 < 0$
. Suppose f has rational Fourier coefficients at the cusp
$\infty $
, so does
$I^M(\tau , f_\mu , \varphi _f)$
for all
$\mu \in L_{}^\vee /L_{}$
.
Remark 4.10. When
$r = 0$
and f has vanishing constant term, the same proof shows that the weakly holomorphic modular form
$I^M(\tau , f_\mu , \varphi _f)$
has rational Fourier coefficients up to algebraic multiples of weight
$1/2$
unary theta series.
Proof. We will use the Fourier expansion of Millson theta lift calculated in Theorem 5.1 of [Reference Alfes-Neumann and SchwagenscheidtANS18], which we now recall. Fix an orientation on
$V_{00}(\mathbb {R})$
. For an isotropic line
$\ell \subset V_{00}$
, let
$G_{00, \ell } \subset G_{00}$
be its stabilizer and
$\gamma _\ell \in \operatorname {SL}_2(\mathbb {Z})$
such that
$\gamma _\ell ^{-1} G_{00, \ell } \gamma _\ell = G_{00, \ell _\infty }$
with
$\ell _\infty = \mathbb {Q} v_\infty $
and
$v_\infty := \left (\begin {smallmatrix} 1 & 0\\ 0 & 0 \end {smallmatrix}\right )$
. Denote
$c_\ell (m, \mu )$
the m-th Fourier coefficient of
$f_\mu \mid _{-2r} \gamma _\ell $
. If
$x \in V_{00}(\mathbb {Q})$
satisfies
$\sqrt {-Q(x)} = d \in \mathbb {Q}_{> 0}$
, then
$x^\perp $
is a hyperbolic plane spanned by two isotropic lines
$\ell _x$
and
$\ell _{-x}$
such that
$x, \gamma _{\ell _x}v_\infty ,\gamma _{\ell _{-x}}v_\infty $
is positively oriented. We can then define
$r_{x} \in \mathbb {Q}$
by

Suppose
$r \ge 1$
. From [Reference Alfes-Neumann and SchwagenscheidtANS18, Theorem 5.1], we know that
$[\operatorname {SL}_2(\mathbb {Z}): \Gamma (N)] \cdot I^M_{}$
is weakly holomorphic of weight
$-r + 1/2 < 0$
with principal part given byFootnote
5

Note that the inner sum above vanishes for d sufficiently large by Proposition 4.7 in [Reference Bruinier and FunkeBF04], and
$I^M$
is uniquely determined by its principal part because its weight is negative. Now we can enlarge N such that
$ N r_{\pm x} w \in \mathbb {Z} $
whenever
$ c_{\ell _{\pm x}}(w, \mu ) \neq 0$
. Then given a prime
$p \nmid N$
, for an element
$x \in \Gamma _{L}\backslash V_{00, -d^2}(\mathbb {Q})$
to have a representative
$\tilde {x} \in V_{00}$
such that both
$t(p) \cdot \tilde {x}$
and
$\tilde {x}$
are both p-integral is equivalent to finding a p-integral representative
$\left (\begin {smallmatrix} A & B\\ B & C \end {smallmatrix}\right )$
with
$p \nmid A$
. Note that the set

is a finite set for any
$\varphi \in {\mathcal {S}}(\hat V_{00})$
.
For any
$\sigma _a \in \operatorname {Gal}(\mathbb {Q}^{\mathrm {ab}}/\mathbb {Q})$
with
$a \in \hat {\mathbb {Z}}^\times $
, we have
$t(a) \in T(\hat {\mathbb {Z}}) \subset {\mathrm {GL}}_2(\hat {\mathbb {Z}})$
. Choose an odd prime
$p \nmid N$
such that
$a \equiv p \bmod {N}$
and every
$ x\in S_d(\varphi _{ f})$
has a p-integral representative
$\tilde {x} \in V_{00}(\mathbb {Q})$
such that
$t(p) \cdot \tilde {x}$
is p-integral. Let
$\tilde {S}_d(\varphi _{ f})$
be such a set of representatives.
Denote
$\tilde {x}' := t(p)\cdot \tilde {x}$
for
$\tilde {x} \in \tilde {S}_d(\varphi _{ f})$
, which is p-integral and satisfies
$\ell _{\pm \tilde {x}' } = t(p) \cdot \ell _{\pm \tilde {x}}$
and

when
$c_{\ell _{\tilde {x}}}(w)$
or
$c_{\ell _{-\tilde {x}}}(w)$
is nonzero. By equation (2.29),
$\Gamma (N) \subset \ker (\rho _L)$
and the fact that f has rational Fourier coefficients, we then have

These imply that

for all
$d> 0$
,
$\tilde {x} \in \tilde S_d(\varphi _f)$
and
$w \in \mathbb {Q}_{< 0}$
. Finally, we have
$\varphi _{ f}(\tilde {x}) = \varphi _{ f}(t(p)^{-1}\tilde {x}')$
, which implies

since
$\varphi _{ f} \in {\mathcal {S}}(L_{})$
and p is co-prime to the level of
$L_{}$
. Here,
$\iota $
is the map defined in (3.10). The map
$\tilde {x} \mapsto \tilde {x}'$
then gives a bijection between
$S_d(\varphi _{ f})$
and
$S_d(\omega (\iota (t(a)))\varphi _{ f})$
. From this, we then obtain

As
$(t(a), \iota (t(a))) \in T^\Delta (\hat {\mathbb {Z}})$
, and
$\varphi _{f}$
is
$T^\Delta (\hat {\mathbb {Z}})$
-invariant, the modular form
$I^M(\tau , f_\mu , \varphi _f)$
has rational Fourier coefficients.
From Propositions 2.8 and 4.9, we can deduce the following result.
Proposition 4.11. Let
$r \in \mathbb {N}_0$
and
$f \in M^!_{-2r, \rho _L}$
as in Proposition 4.9. For all
$\mu \in L^\vee /L$
and congruence subgroup
$\Gamma (N) \subset \ker (\rho _L)$
fixing
$\varphi _\mu \in {\mathcal {S}}(\hat V; \mathbb {Q}^{\mathrm {ab}})^{(G \cdot T^\Delta )(\hat {\mathbb {Z}})}$
, which is a matching section of
$\phi _{\mu }$
as in Theorem 3.3, the regularized integral

is a rational number.
Proof. Since
$\varphi _\mu $
is
$G(\hat {\mathbb {Z}})$
-invariant, we can rewrite the constant
$c_\mu := c_\mu (f)$
as

with
$\varphi ^r_\mu := \widetilde {\mathrm {RC}}_r(\varphi _\mu ^{(1, -1)} - (-1)^r \varphi _\mu ^{(-1, 1)}) \in {\mathcal {S}}(V(\mathbb {A}))$
. Using Lemma 4.4, we can switch the regularized integral in g with the limit in
$T'$
. Then by the rational decomposition
$V = V_{00} \oplus U_{D} \oplus V_1$
, we can write

with
$ \varphi _{00, \mu , j} \in {\mathcal {S}}(\hat V_{00}; \mathbb {Q}^{\mathrm {ab}})^{T^\Delta (\hat {\mathbb {Z}})}, \varphi _{1, \mu , j} \in {\mathcal {S}}(\hat V_1)$
and
$\varphi _{D, \mu , j} \in {\mathcal {S}}(\hat U_D)$
. The constant
$c_\mu $
can then be rewritten as

where
$c_{\mu , j}$
is defined by

Now with Lemma 4.3, we obtain

where
$G_{\mu , j}$
is a weakly holomorphic modular form of weight
$-1$
defined by

and has rational Fourier coefficient at the cusp
$\infty $
by Proposition 4.11. As
$\varphi _\mu $
is
$\operatorname {SL}_2(\hat {\mathbb {Z}})$
-invariant, the function
$(v')^{-1/2} \sum _{j \in J} \Theta _{1}(g_{\tau '}, \varphi _{1, \mu , j}^-, \varrho ) G_{\mu , j}(\tau ')$
is
$\operatorname {SL}_2(\mathbb {Z})$
-invariant in
$\tau '$
. Applying Proposition 2.8 and Stokes’ Theorem, we then have

This finishes the proof.
5 Proofs of theorems
In this section, we will prove Theorem 1.2. First, we state and prove the case for
$\mathrm {O}(2, 2)$
.
Theorem 5.1. Let E be a biquadratic CM number field with real quadratic subfield F. Let
$W = W_\alpha $
be an F-quadratic space and
$W_{{\alpha ^\vee }}$
its neighborhood quadratic space as in Section 2.4. Suppose
$\alpha _1 < 0 < \alpha _2$
. For
$r \in \mathbb {N}_0$
and a lattice
$L \subset W_{\mathbb {Q}}$
, suppose

is a weakly holomorphic modular form with rational Fourier coefficients. Furthermore, suppose it has vanishing constant term when
$r = 0$
. Then there exists
$\kappa , M \in \mathbb {N}$
depending on f such that

where
$\beta (t, \phi _f) \in F^\times $
is nonzero which has the property

Proof. By the Siegel-Weil formula in (2.56), we have

with
$C' := \deg (Z(W ))/(2\Lambda (0, \chi )) \in \mathbb {Q}^\times $
. For each
$\mu \in L^\vee /L$
, let
$\varphi _{\mu } \in {\mathcal {S}}(\hat V)^{(G\cdot T^\Delta )(\hat {\mathbb {Z}})}$
be a matching section of
$\phi _{\mu }$
as in Theorem 3.3. Then we can apply Proposition 4.2 to obtain

By Proposition 4.11, we know that
$c_\mu (f) \in \mathbb {Q}$
for all
$\mu \in L^\vee /L$
. For the other term, we can apply Stokes’ theorem to obtain

For the last step, we have applied Remark 4.8 to replace
$\tilde {\mathcal {I}}$
with
$\mathcal {I}^0 + \mathcal {I}^+$
, used Lemma 4.5 to see that
$\mathcal {I}^0$
contribute nothing, and substitute in the Fourier coefficients of
$\mathrm {RC}_r\mathcal {I}^+$
in terms of
$\tilde {c}_\lambda ( \varphi _{\mu }, \varrho )$
, the
$\lambda $
-th Fourier coefficient of
$\tilde {\mathcal {I}}(g_\tau ^\Delta , \varphi ^{(1, 1)}_\mu , \tilde {\varrho }_C)$
. Note that
$m^r P_r((\lambda - \lambda ')/m)$
appears by (2.9) and is a rational multiple of
$\sqrt {D}^r$
. As the sum above is finite, we can choose C such that
$S_C^c$
contains
$\mathrm {Diff}(W , t)$
for all the t that appears in this sum. Finally, the knowledge about the factorization of these coefficients in Proposition 4.7 finishes the proof.
Corollary 5.2. In the setting of Theorem 5.1, suppose that
$Z_f$
does not intersect with
$Z(W )$
when
$r = 0$
. Then there exists
$\kappa \in \mathbb {N}$
and
$\gamma (\lambda , \phi _\mu ) \in F^\times $
such that

and

Remark 5.3. The constant
$\frac {\deg (Z(W ))}{\Lambda (0, \chi )}$
can be explicitly given when
$X_K = X_0(1)^2$
(see [Reference LiLi21, Remark 3.6]).
Proof. By Theorem 5.10 in [Reference Bruinier, Ehlen and YangBEY21], we have

where
$a_t(\phi _\mu )$
is t-th the Fourier coefficient of the holomorphic part of the incoherent Eisenstein series, and given in (2.61). Adding this to (5.1) and applying (5.2) finishes the proof.
Now we can prove Theorem 1.2.
Proof of Theorem 1.2.
Write
$\mathrm {V} = \mathrm {V}_\circ \oplus W_{\mathbb {Q}}$
and
$L_W := L \cap W_{\mathbb {Q}},~ L_\circ := L \cap \mathrm {V}_\circ $
. Then
$L_\circ \oplus L_W \subset L$
is a full sublattice, and we can write

with
$\tilde {f}(\tau _1, \tau _2) := \langle \operatorname {Tr}^L_{L_\circ \oplus L_W}(f(\tau _1)), \overline {\Theta _{L_\circ }(\tau _2)}\rangle _{L_\circ }$
. Using (2.10), we have

where
$k_1 = -2r + 1 - \frac {n}{2}, k_2 = \frac {n}{2}-1$
and

has rational Fourier coefficients. Therefore, we have

If f has the Fourier expansion

then the
$(m, \mu )$
-th Fourier coefficient of
$\tilde {f}_\ell $
, denoted by
$c_\ell (m, \mu )$
, can be expressed as

with
$Q_{\ell , (k_1, k_2)}(X, Y) \in \mathbb {Q}[X, Y]$
defined in (2.4). In particular, when
$Z(W) \cap Z_f = \emptyset $
, we have
$c_r(0, \mu ) = 0$
for all
$0 \le \ell \le r$
and
$\mu \in L^\vee /L$
as
$c(-Q(\lambda _\circ ), (\lambda _\circ , \mu )) = 0$
for all
$\lambda _\circ \in L_\circ ^\vee $
and
$\mu \in L_W^\vee /L_W$
by (2.43).
By Corollary 5.2, we can write

with
$\gamma _\ell (\lambda , \phi _\mu ) \in F^\times $
having factorization as in (5.4) independent of
$\ell $
, and express
$\Phi ^r_f(Z(W))$
as in (1.6) such that

for all prime
$\mathfrak {p}$
of F.
Finally, we prove Theorem 1.5.
Proof of Theorem 1.5.
By the main result in [Reference LiLi23], we have
$\tilde {\kappa } \in \mathbb {N}$
and Galois equivariant maps
$\tilde \alpha _j: T_W(\hat {\mathbb {Q}}) \to E^{\mathrm {ab}}$
satisfying

for all
$h, h' \in T_W(\hat {\mathbb {Q}})$
. Furthermore, when
$n = 2$
, we have
$\tilde \alpha _j = 1$
for
$j \equiv r \bmod {2}$
. Setting
$\alpha _j(h) := a_j \prod _{[z_0, h'] \in Z(W) \backslash [z_0, h]} \frac {\tilde \alpha _j(h)}{\tilde \alpha _j(h')}$
and
$\kappa := \tilde {\kappa } |Z(W)|$
and applying equations (5.7) and (1.6) proves the first two claims. Combining with Corollary 2.4, we see that Conjecture 1.1 holds.
Acknowledgements
We thank Claudia Alfes, Chao Li and Shaul Zemel for helpful comments on an earlier version. We thank Zhiwei Sun for his help on formula (2.5). We also thank Mingkuan Zhang and the anonymous referees for carefully reading through the draft and giving useful remarks to help clear confusions and improve the exposition. J.B. and Y.L. were supported by the LOEWE research unit USAG and by the Deutsche Forschungsgemeinschaft (DFG) through the Collaborative Research Centre TRR 326 ‘Geometry and Arithmetic of Uniformized Structures’, project number 444845124. Y.L. was also supported by DFG through the Heisenberg Program ‘Arithmetic of real-analytic automorphic forms’, project number 539345613. T.Y. was partially supported by the UW-Madison Kelley Mid-Career Award, the Dorothy Gollmar Chair’s Fund and Van Vleck research fund of the department of mathematics, UW-Madison.
Competing Interest
The authors have no competing interest to declare.