This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\@setabstract

Explicit formula for the (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2}) theta lift via Bruhat decomposition Formule explicite pour la correspondance thêta (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2}) via la décomposition de Bruhat

Peter Xu
Abstract.

Using combinations of weight-11 and weight-22 of Kronecker-Eisenstein series to construct currents in the distributional de Rham complex of a squared elliptic curve, we find a simple explicit formula for the type II (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2}) theta lift without smoothing, analogous to the classical formula of Siegel for periods of Eisenstein series. For KK a CM field, the same technique applies without change to obtain an analogous formula for the (GL2(K),K×)(\mathrm{GL}_{2}(K),K^{\times}) theta correspondence.

Utilisant des combinaisons des séries et courants Kronecker-Eisenstein de poids 11 et 22 dans des complexes de Rham d’une courbe elliptique carrée, nous trouvons une formule simple explicite pour la correspondance thêta de type II pour (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2}) sans stabilisation; cela est un analogue de la formule classique de Siegel pour les périodes des séries d’Eisenstein. Pour KK un corps avec CM, la même technique s’applique pareil pour obtenir une telle formule pour la correspondance (GL2(K),K×)(\mathrm{GL}_{2}(K),K^{\times}).

1. Introduction

In the article [BCG1] and its sequel [BCG3], the authors construct a theta lift from the (n1)(n-1)-homology of GLn\mathrm{GL}_{n} to modular forms for GL2\mathrm{GL}_{2}. This construction uses an equivariant polylogarithm class; in particular, when n=2n=2, the class in question corresponds to the algebraic GL2\mathrm{GL}_{2}-action on the squred universal elliptic curve over the upper half-plane (or its modular curve quotients).

This class can be computed by various means: in [BCG1], the lifts of certain modular geodesics are computed via the seesaw principle, while in [BCG3], certain stabilizations of the lift are computed by relating them to values “at the boundary,” i.e. by partial modular symbols. Related cocycles valued in KK-theory were constructed in [SV] and [BPPS] (for general nn; the regulators of these cocycles these correspond to parabolic stabilizations of the theta lift. We also construct both parabolic and non-parabolic stabilized versions of the KK-theory cocycle for general nn [X3].

In the previous paper [X4], we constructed explicit elements in the GL2\mathrm{GL}_{2}-equivariant distributional de Rham complex of a torus to re-derive classical “unstabilized” formulas for periods of weight-22 Eisenstein series. The principle used was that in cohomology, these Eisenstein series arise as pullbacks of an equivariant polylogarithm class on (S1)2(S^{1})^{2}, allowing us to compute them via equivariant-geometric means. In the present article, by applying the same technique to the squared universal elliptic curve, we obtain novel, similarly explicit formulas for the (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2})-theta lift. These formulas especially shed light on the Eisenstein component of the lift, exhibiting the particular combiniation of weight-22 Eisenstein series which arise from specializations at parabolic matrices.

Due to the nature of the theta lift’s construction, instead of the equivariant distributional de Rham complex of [X4], we will use the equivariant weight-22 distributional Dolbeault complex of our squared universal elliptic curve. The explicit elements in loc. cit. were comprised of weight-11 and weight-22 Bernoulli polynomials B1(z)B_{1}(z) and B2(z)B_{2}(z); these will be replaced with theta series E1(τ,z)E_{1}(\tau,z) and E2(τ,z)E_{2}(\tau,z) interpolating weight-11 and 22 Eisenstein series. To find the explicit forms of these elements, as in loc. cit., we will need to use an explicit version of the Bruhat decomposition of

GL2()=BwB\mathrm{GL}_{2}(\mathbb{Q})=BwB

where

w=(0110)w=\begin{pmatrix}0&1\\ 1&0\end{pmatrix}

represents a Weyl element, and BGL2()B\subset\mathrm{GL}_{2}(\mathbb{Q}) is the upper triangular Borel subgroup.

1.1. Related work and future directions

A similar computation using complexes to compoute equivariant cohomology was completed in [KS, §3], in a more general setting using a generalized Kronecker-Eisenstein series. Our calculation can be considered as a more explicated form of the calculation there, for the self-product of the universal elliptic curve with trivial coefficients, except that:

  1. (1)

    we treat the entire action of GL2\mathrm{GL}_{2} at once, instead of just working with the equivariance under one subtorus at a time;

  2. (2)

    we construct also a cocycle for a larger group with no geometric action on the fibers (GL2()\mathrm{GL}_{2}(\mathbb{Q}) instead of GL2()\mathrm{GL}_{2}(\mathbb{Z}), or GL2(K)\mathrm{GL}_{2}(K) in the CM case);

  3. (3)

    we do not rely on any stabilization or smoothing coming from a degree-zero torsion cycle.

We restrict to this setting both because it is particularly arithmetically rich, and allows for a considerably simpler and less technical exposition. Futhermore, by eschewing smoothing, the resulting formulas are more transparently related to classical objects such as classical holomorphic Eisenstein series or the period formulas of Siegel (as in [X4]), and corresponds directly to the theta kernel of integration of [BCG1]. Furthermore, our formulas can be specialized (with cohomological meaning) at arbitrary non-zero torsion points, thanks to Theorem 3.5.

As we noted in our previous article [X4] in a simpler setting, our same methods should produce analogous formulas with twisted coefficients corresponding to higher-weight automorphic sheaves; the resulting formulas will involve progressively more complicated combinations of higher-weight specializations of Kronecker-Eisenstein series. Another natural generalization would be to use the more complicated decomposition of parabolic subgroups for GLn\mathrm{GL}_{n} and nn-fold products of elliptic curves; in this setting, our unsmoothed formulas grow in complexity very quickly with nn, and it is not clear to us if there exists a symbol formalism that can help one understand this complexity. Furthermore, the direct analogue of Theorem 3.5 does not hold for n>2n>2, so the ability to specialize at torsion sections is more limited. Nevertheless, in the smoothed setting, things are much simpler and can be captured by symbols coming from elementary linear algebra; we follow this approach for elliptic schemes in [X3].

1.2. Acknowledgements

I would like to thank first and foremost Nicolas Bergeron, to whose ideas I owe all of my work in this subject area; and who very kindly accommodated me in Paris in February 2023, listening to and helping to develop my ideas even with the limited energy he had at the time. This article is dedicated in his memory. I would also like to thank Romyar Sharifi, Marti Roset Julia, and Pierre Charollois for helpful consultations during the writing process.

2. Eisenstein theta lift for (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2})

The theta lift we are considering was originally constructed purely analytically in [BCG1], but we will find the algebraic construction of [BCG3] more useful; the comparison between the two is also proven in loc. cit. (See also the author’s thesis [X].)

To begin, we consider the complex upper half-plane \mathcal{H}, and the universal elliptic curve over it given by the complex uniformization

E:=(×)/{(τ,+τ)}.E:=(\mathcal{H}\times\mathbb{C})/\{(\tau,\mathbb{Z}+\tau\mathbb{Z})\}.

Here, we will write τ=x+yi\tau=x+yi for the coordinate on the base, and zz for the fiberwise coordinate on \mathbb{C}. This fiber bundle has the complex analytic action (as a fiber bundle) of GL2()\mathrm{GL}_{2}(\mathbb{Z}) by

(abcd)(τ,z)=(aτ+bcτ+d,zcτ+d).\begin{pmatrix}a&b\\ c&d\end{pmatrix}(\tau,z)=\left(\frac{a\tau+b}{c\tau+d},\frac{z}{c\tau+d}\right).

We will also write z=uτvz=u-\tau v for u,vu,v\in\mathbb{R}; this then allows GL2()\mathrm{GL}_{2}(\mathbb{Z}) to act on the column vector (u,v)(u,v) by the standard left action.

The principal space we will work on is the self-product T:=E×ET:=E\times_{\mathcal{H}}E, which we will coordinatize by τ\tau on the base, and z1z_{1} and z2z_{2} on the two elliptic curves. For us, the important property of TT is that it has a fiberwise action by endomorphisms of M2()M_{2}(\mathbb{Z}), coming from the endomorphism action of \mathbb{Z} on any elliptic scheme by isogenies, i.e. the integer aa\in\mathbb{Z} acts by the finite multiplication-by-aa isogeny [a][a]_{*} of degree a2a^{2}. In particular, the group GL2()\mathrm{GL}_{2}(\mathbb{Z}) acts by automorphisms on TT.

For any torsion-free subgroup HGL2()H\subset\text{GL}_{2}(\mathbb{Z}), the Y(H):=H\Y(H):=H\backslash\mathcal{H} can be equipped with an algebraic structure making it the fine moduli space of elliptic curves over \mathbb{C}-schemes with HH-level structure. There is also the corresponding universal elliptic curve

E(H):=H\(×(/2)).E(H):=H\backslash(\mathcal{H}\times(\mathbb{C}/\mathbb{Z}^{2})).

and its square

T(H):=E(H)×Y(H)E(H),T(H):=E(H)\times_{Y(H)}E(H),

whose fiber over a point of the moduli space is the square of the corresponding elliptic curve. In this article, for technical reasons, we will primarily work over \mathcal{H} rather than after quotient by HH; however, the output of our construction will visibly descend through these quotients.

Let ΓGL2()\Gamma\subset\mathrm{GL}_{2}(\mathbb{Z}) be any subgroup; we will work with the Γ\Gamma-equivariant cohomology of Γ\Gamma-subspaces of TT. In this article, we will follow the approach pioneered by [KS] to define the theta lift algebraically in coherent cohomology of relative sheaves of differentials over \mathcal{H}; see the author’s thesis [X] for details about how this relates to the algebraic construction of [BCG3].

Following [KS], if we fix an auxiliary integer c>1c>1, there is a unique (“polylogarithm”) cohomology class

zΓ(c)HΓ1(TT[c],ΩTT[c]/2)(0)z_{\Gamma}^{(c)}\in H^{1}_{\Gamma}(T-T[c],\Omega^{2}_{T-T[c]/\mathcal{H}})^{(0)}

characterized (as indicated by the superscript (0)(0)) by being invariant by [a][a]_{*} for all integers aa relatively prime to cc, and having residue

[T[c]c2{0}]HΓ,T[c]2(T,ΩT/2)[T[c]-c^{2}\{0\}]\in H^{2}_{\Gamma,T[c]}(T,\Omega^{2}_{T/\mathcal{H}})

where HΓ,T[c]H^{\bullet}_{\Gamma,T[c]} denotes the equivariant cohomology with support. Let STS\subset T be a nonempty Γ\Gamma-fixed arrangement of T[c]T[c]-translates of elliptic subgroups of TT; we restrict the class zΓ(c)z_{\Gamma}^{(c)} to

HΓ1(TS,ΩTS/2)(0):=limSfSHΓ1(TSf,ΩTSf/2)(0)H^{1}_{\Gamma}(T-S,\Omega^{2}_{T-S/\mathcal{H}})^{(0)}:=\varinjlim_{S_{f}\subset S}H^{1}_{\Gamma}(T-S_{f},\Omega^{2}_{T-S_{f}/\mathcal{H}})^{(0)}

where the limit is over finite subarrangements. The various TSfT-S_{f} are cofinally fiberwise Stein manifolds over the contractible base \mathcal{H}, so have vanishing higher coherent cohomology; thus, in the Hochschild-Serre spectral sequence we have a Hochschild-Serre edge map

e:HΓ1(TS,ΩTS/2)(0)H1(Γ,H0(TS,ΩTS/2)):=H1(Γ,limH0(TSf,ΩTSf/2))(0)e:H^{1}_{\Gamma}(T-S,\Omega^{2}_{T-S/\mathcal{H}})^{(0)}\to H^{1}(\Gamma,H^{0}(T-S,\Omega^{2}_{T-S/\mathcal{H}})):=H^{1}(\Gamma,\varinjlim H^{0}(T-S_{f},\Omega^{2}_{T-S_{f}/\mathcal{H}}))^{(0)}

where here we use the exactness of filtered colimits.

Thanks to a similar argument as the one to uniquely define zΓ(c)z_{\Gamma}^{(c)}, using projectors built from isogenies, e(zΓ(c))e(z_{\Gamma}^{(c)}) can be refined to a class which we denote

ZΓ(c)H1(Γ,H0(TS,ΩTS/2)(0))Z^{(c)}_{\Gamma}\in H^{1}(\Gamma,H^{0}(T-S,\Omega^{2}_{T-S/\mathcal{H}})^{(0)})

i.e. a cohomology class valued in prime-to-cc isogeny-fixed forms (instead of being isogeny-fixed only up to coboundary).

This group cohomology class valued in 22-forms on (an open subspace of) the bundle TT parameterizes a family of theta lifts: if one contracts with the GL2()\mathrm{GL}_{2}(\mathbb{Z})-fixed vector field z1z2\partial z_{1}\otimes\partial z_{2} and pulls back by any Γ\Gamma-fixed torsion section xx disjoint from SS, then

ΘΓ,x(c):=x(ZΓ(c)z1z2)H1(Γ,H0(,ω2))\Theta^{(c)}_{\Gamma,x}:=x^{*}(Z^{(c)}_{\Gamma}\cap\partial z_{1}\otimes\partial z_{2})\in H^{1}(\Gamma,H^{0}(\mathcal{H},\omega^{\otimes 2}))

can be viewed via integration as a map from H1(Γ)H_{1}(\Gamma) to a weight-22 modular form inside H0(,ω2)H^{0}(\mathcal{H},\omega^{\otimes 2}), as the values of the integral will be fixed by any level structure HH stabilizing the section xx. Note that for any given nonzero xx, one can always find a disjoint STS\subset T invariant by its stabilizer, so all torsion sections have associated pullbacks. Furthermore, since one can always restrict away from the union of two different SS, it is clear the pulled back class is independent of the choice of removed sub-elliptic curves. If we fix any point of the base τ\tau\in\mathcal{H}, we can similarly define the fiberwise version of these cycles by restriction,

Zτ(c)H1(Γ,H0(TτSτ,ΩTτSτ2)(0)),Θτ,x(c):=x(Zτ(c)z1z2)H1(Γ,)Z^{(c)}_{\tau}\in H^{1}(\Gamma,H^{0}(T_{\tau}-S_{\tau},\Omega^{2}_{T_{\tau}-S_{\tau}})^{(0)}),\Theta^{(c)}_{\tau,x}:=x^{*}(Z^{(c)}_{\tau}\cap\partial z_{1}\otimes\partial z_{2})\in H^{1}(\Gamma,\mathbb{C})

where the subscript τ\tau refers to the fiber over that point. (We here omit the Γ\Gamma from the notation for brevity, since these classes are compatible under restriction of Γ\Gamma in any case.)

Out of technical convenience, we will obtain our explicit formulas for the fiberwise classes Θτ,x(c)\Theta^{(c)}_{\tau,x}. Since the action of Γ\Gamma is trivial for these classes, however, this will also imply formulas for the classes ΘΓ,x(c)\Theta^{(c)}_{\Gamma,x}.111With a little extra work, one can also prove similar formulas for the “spread out” classes ZΓ(c)Z^{(c)}_{\Gamma}, but this is not the focus of this article.

2.1. Computing with complexes

We now define the main tool we need, in this setting: the weight-22 distributional Dolbeault complex. Let XX be a complex manifold of complex dimension dd, and write 𝒜Xp,q\mathcal{A}^{p,q}_{X} for the sheaves of smooth complex differentials on XX of holomorphic degree pp and antiholomorphic degree qq; these sheaves are flabby on XX, so we (a bit abusively) use this same notation for their global sections. These forms support the pair of anti-commuting Dolbeault differentials

:𝒜Xp,q𝒜Xp+1,q,¯:𝒜Xp,q𝒜Xp,q+1\partial:\mathcal{A}^{p,q}_{X}\to\mathcal{A}^{p+1,q}_{X},\overline{\partial}:\mathcal{A}^{p,q}_{X}\to\mathcal{A}^{p,q+1}_{X}

which are square zero, satisfy the Leibniz rule, and sum to the usual exterior derivative on differential forms. For an integer w0w\geq 0, we recall that the holomorphic (w,0)(w,0)-differentials are precisely the kernel of ¯\overline{\partial} inside AXw,0A^{w,0}_{X}.

Ordinarily, the Dolbeault resolution of ΩX2\Omega^{2}_{X} is constructed from these sheaves, but instead, we dualize: first, write 𝒜X,cp,q\mathcal{A}^{p,q}_{X,c} for the space of compactly supported relative (p,q)(p,q)-differentials on XX, carrying the same differentials as the usual ones (though they do not form a sheaf). Then we define

𝒟Xp,q=hom(𝒜X,cdp,dq,)\mathcal{D}_{X}^{p,q}=\hom(\mathcal{A}^{d-p,d-q}_{X,c},\mathbb{C})

which we call the space of (p,q)(p,q)-currents on XX. Unlike the compactly supported forms, these do form a sheaf, since they can be restricted along inclusions of opens in adjunction to the pushforward of compactly supported forms. Since these pushforwards are injective, the sheaf restrictions maps are surjective and thus each 𝒟Xp,q\mathcal{D}_{X}^{p,q} is flabby. We therefore again abusively use the same notation for the sheaf and its global sections. For these properties of currents, the original reference is [dR].

The Dolbeault operators

:𝒟Xp,q𝒟Xp+1,q,¯:𝒟Xp,q𝒟Xp,q+1\partial:\mathcal{D}^{p,q}_{X}\to\mathcal{D}^{p+1,q}_{X},\overline{\partial}:\mathcal{D}^{p,q}_{X}\to\mathcal{D}^{p,q+1}_{X}

are defined as the graded adjoints of the exterior derivative on forms, i.e.

(c)(η):=(1)degcc(η),(¯c)(η):=(1)degcc(¯η)(\partial c)(\eta):=(-1)^{\deg c}c(\partial\eta),(\overline{\partial}c)(\eta):=(-1)^{\deg c}c(\overline{\partial}\eta)

where deg\deg refers to the total degree (i.e. p+qp+q for a (p,q)(p,q)-current).

There is a natural inclusion

νX:𝒜Xp,q𝒟X/p,q\nu_{X}:\mathcal{A}^{p,q}_{X}\to\mathcal{D}^{p,q}_{X/}

given by

ω(ηXωη)\omega\mapsto\left(\eta\mapsto\int_{X}\omega\wedge\eta\right)

where η\eta is a compactly supported smooth (dp,dq)(d-p,d-q)-form; the integral makes sense and is finite since the wedge product is also compactly supported, and of degree (d,d)(d,d). It is clear that νX\nu_{X} commutes with the Dolbeault differentials \partial and ¯\overline{\partial}. Furthermore, since (p,q)(p,q)-compactly supported forms have covariant functoriality by flat holomorphic maps and contravariant by proper holomorphic maps, by adjunction, (p,q)(p,q)-relative currents can be pulled back by flat maps and pushed forward by proper ones. One can check that the push-pull formula for differential implies that νX\nu_{X} is functorial for flat pullback and proper pushforward of complex manifolds.222For orientation-reversing maps, one has to be careful about this comparison map, since they introduce an extra sign; this was an issue we had to work with in [X4]. However, all of our maps will be morphisms of complex manifolds, and hence orientation-preserving.

We now come to the main cohomological result about these currents:

Lemma 2.1.

The complex

(2.1) ΩXw𝒟Xw,0¯𝒟Xw,1¯\Omega^{w}_{X}\hookrightarrow\mathcal{D}^{w,0}_{X}\xrightarrow{\overline{\partial}}\mathcal{D}^{w,1}_{X}\xrightarrow{\overline{\partial}}\ldots

is an acyclic resolution of sheaves. Further, the map of ¯\overline{\partial}-complexes induced by νX\nu_{X}

(2.2) 𝒜Xw,𝒟Xw,\mathcal{A}^{w,\bullet}_{X}\to\mathcal{D}^{w,\bullet}_{X}

is a quasi-isomorphism for each degree ww.

Proof.

This is a consequence of the ¯\overline{\partial}-Poincaré (or Dolbeault-Grothendieck) lemma due to [Skoda] which implies the exactness of the for sheaves of holomorphic forms/currents on any complex manifold (not relative to a base). The version for forms can be found in [Serre], which implies the quasi-isomorphism (2.2). ∎

In particular, for 22-dimensional XX and w=2w=2, the complex

𝒟X2,0¯𝒟X2,1¯𝒟X2,2\mathcal{D}^{2,0}_{X}\xrightarrow{\overline{\partial}}\mathcal{D}^{2,1}_{X}\xrightarrow{\overline{\partial}}\mathcal{D}^{2,2}_{X}

is functorial in XX (in the sense described previously) and computes the coherent cohomology of ΩX2\Omega^{2}_{X}. (Here, we recall that XX is assumed to be relative dimension 22, so there are no further terms.)

Having proved this key cohomological property, we now introduce an important class of currents: associated to closed complex submanifolds ZXZ\subset X of codimension rr, we have a closed current of integration

δZ𝒟Xr,r\delta_{Z}\in\mathcal{D}^{r,r}_{X}

defined by

δZ(ω):=Zω.\delta_{Z}(\omega):=\int_{Z}\omega.

Now consider a fixed squared elliptic fiber Tτ=Eτ×EτT_{\tau}=E_{\tau}\times E_{\tau}. Then the class of a closed current ω𝒟Tτ2,1\omega\in\mathcal{D}^{2,1}_{T_{\tau}} having residue 𝒞H0(Tτ[c])\mathcal{C}\in H^{0}(T_{\tau}[c]) along the residue map

H1(TτTτ[c],ΩTτTτ[c]2)H0(Tτ[c])H^{1}(T_{\tau}-T_{\tau}[c],\Omega^{2}_{T_{\tau}-T_{\tau}[c]})\to H^{0}(T_{\tau}[c])

is equivalent to dω=δ𝒞d\omega=\delta_{\mathcal{C}} (where this is interpreted as a suitable linear combination of the currents of integration along points in the support of 𝒞\mathcal{C}); see for example [X, (3.3)] from the author’s thesis.

Since the distributional Dolbeault complex is a functorial complex computing the coherent cohomology of holomorphic differentials, the coresponding equivariant coherent cohomology of ΩX2\Omega^{2}_{X} can thus be computed by the double complex C(Γ,𝒟X2,)C^{\bullet}(\Gamma,\mathcal{D}^{2,\bullet}_{X}), where ΓGL2()\Gamma\subset\mathrm{GL}_{2}(\mathbb{Z}) acts on the currents by pushforward.

Analogously to the non-equivariant case, then, an element ω\omega of this double complex for TT restricts to a representative of a class in HΓ1(TτTτ[c],ΩTτTτ[c]2)H^{1}_{\Gamma}(T_{\tau}-T_{\tau}[c],\Omega^{2}_{T_{\tau}-T_{\tau}[c]}) with residue

[Tτ[c]c4{0}]H0(Tτ[c])Γ[T_{\tau}[c]-c^{4}\{0\}]\in H^{0}(T_{\tau}[c])^{\Gamma}

if and only if the total differential of ω\omega is δTτ[c]c4δ0C0(Γ,𝒟Tτ2,2)\delta_{T_{\tau}[c]}-c^{4}\delta_{0}\in C^{0}(\Gamma,\mathcal{D}^{2,2}_{T_{\tau}}). In particular, if we can find such a class which is in the trace-fixed part (𝒟Tτ2,)(0)(\mathcal{D}^{2,\bullet}_{T_{\tau}})^{(0)}, this notation meaning the part which is invariant by [a][a]_{*} for all integers aa relatively prime to cc, then it will represent the class

Zτ(c)HΓ1(TτTτ[c],ΩTτTτ[c]2).Z_{\tau}^{(c)}\in H^{1}_{\Gamma}(T_{\tau}-T_{\tau}[c],\Omega^{2}_{T_{\tau}-T_{\tau}[c]}).

3. Constructing the lift

3.1. Kronecker-Eisenstein series as theta functions

We now know that in principle, one can compute Eisenstein classes by finding suitable lifts inside the distributional de Rham complex. It remains to find currents realizing these lifts.

Recall the Kronecker-Eisenstein series [Weil]

(3.1) Ka(s,τ,z,u)=ω+τ(ω+z¯)a|ω+z|2sexp(2πiωu¯ω¯uττ¯)K_{a}(s,\tau,z,u)=\sum_{\omega\in\mathbb{Z}+\mathbb{Z}\tau}^{\prime}\frac{(\overline{\omega+z})^{a}}{|\omega+z|^{2s}}\exp\left(2\pi i\frac{\omega\overline{u}-\overline{\omega}u}{\tau-\overline{\tau}}\right)

for τ\tau\in\mathcal{H}, z,uz,u\in\mathbb{C}, and aa\in\mathbb{Z}; the superscript apostrophe denotes that the sum omits any term where ω+z=0\omega+z=0. This series is convergent for Res>1+a/2\text{Re}\,s>1+a/2, but has meromorphic continuation to all ss\in\mathbb{C}, with simple poles possibly only at s=0s=0 (when z+τz\in\mathbb{Z}+\mathbb{Z}\tau and a=0a=0) s=1s=1 (when u+τu\in\mathbb{Z}+\mathbb{Z}\tau and a=0a=0). We define special notations, following [BCG1, §9], for the specializations we need:

(3.2) E1(τ,z)\displaystyle E_{1}(\tau,z) :=i2πK1(1,τ,z,0),\displaystyle:=\frac{i}{2\pi}K_{1}(1,\tau,z,0),
(3.3) E2(τ,z)\displaystyle E_{2}(\tau,z) :=14πyK2(1,τ,z,0).\displaystyle:=-\frac{1}{4\pi y}K_{2}(1,\tau,z,0).

Since these functions are manifestly (+τ)(\mathbb{Z}+\mathbb{Z}\tau)-periodic in zz for fixed τ\tau, we can consider E1(τ,z)E_{1}(\tau,z) and E2(τ,z)E_{2}(\tau,z) as functions on EτE_{\tau}. Further, E1(τ,z)E_{1}(\tau,z) is odd and E2(τ,z)E_{2}(\tau,z) is even in zz, for any τ\tau; in particular, we see that E1(τ,0)=0E_{1}(\tau,0)=0. As noted in [Weil], these two functions specialize at torsion points

E1(τ,αβτ),E2(τ,αβτ);(α,β)(/)2{(0,0)}E_{1}(\tau,\alpha-\beta\tau),E_{2}(\tau,\alpha-\beta\tau);(\alpha,\beta)\in(\mathbb{Q}/\mathbb{Z})^{2}\setminus\{(0,0)\}

to holomorphic modular forms in τ\tau of weight 11, respectively 22, which are the classical Eisenstein series of those weights, of level corresponding to the stabilizer of (α,β)(\alpha,\beta) in SL2()\mathrm{SL}_{2}(\mathbb{Z}).

From [BCG1, (9.6)] together with [BCG1, Proposition 20], we observe that we have the distribution relations as functions

(3.4) [a](E1(τ,z)dz)\displaystyle[a]_{*}(E_{1}(\tau,z)\,dz) =E1(τ,z)dz,\displaystyle=E_{1}(\tau,z)\,dz,
(3.5) [a]E2(τ,z)\displaystyle[a]_{*}E_{2}(\tau,z) =E2(τ,z)\displaystyle=E_{2}(\tau,z)

for any integer a{0}a\in\mathbb{Z}\setminus\{0\}. Further, [BCG1, Theorem 19] and [BCG1, §9.4] together imply that, considered as a current on EτE_{\tau},

¯E1(τ,z)dz=δ0volEτ=δ02idzdz¯y\overline{\partial}E_{1}(\tau,z)\,dz=\delta_{0}-\text{vol}_{E_{\tau}}=\delta_{0}-\frac{2idz\wedge d\overline{z}}{y}

where here we give the formula for the normalized volume form on EτE_{\tau}. Away from the zero section (ignoring the current of integration δ0\delta_{0}), this derivative holds on the level of functions. Furthermore, [Weil, (35)] says that

¯E2(τ,z)=i2yE1(τ,z)dz¯\overline{\partial}E_{2}(\tau,z)=\frac{i}{2y}\cdot E_{1}(\tau,z)\,d\overline{z}

as functions (and hence currents) on EτE_{\tau}. Thus, both E1E_{1} and E2E_{2} are smooth away from the zero section, where E2E_{2} is once-differentiable and E1E_{1} has a singularity. Note that the formulas in [BCG1, §9.4] imply that on a punctured neighborhood of the zero section, E1E_{1} looks like it has a simple log pole (even though its actual value at zero is 0). It therefore makes sense to consider forms like E1(τ,z)dzE_{1}(\tau,z)\,dz and E2(τ,z)dzE_{2}(\tau,z)\,dz as (1,0)(1,0)-currents on EτE_{\tau} by considering them as kernels of integration as in the map νX\nu_{X} we defined for smooth forms, since the corresponding integrals will converge for compactly supported smooth forms.

3.2. Explicit isogeny-fixed currents

We now construct a certain element of total degree 11 (or 33, if one adds in the Hodge weight 22)

ζτC(GL2(),(𝒟Tτ2,)(0))\zeta_{\tau}\in C^{\bullet}(\mathrm{GL}_{2}(\mathbb{Q}),(\mathcal{D}_{T_{\tau}}^{2,\bullet})^{(0)})

whose total differential is the cocycle

δ0volTτZ0(Γ,(𝒟Tτ2,2)(0))=[(𝒟Tτ2,)(0)]Γ\delta_{0}-\text{vol}_{T_{\tau}}\in Z^{0}(\Gamma,(\mathcal{D}_{T_{\tau}}^{2,2})^{(0)})=[(\mathcal{D}_{T_{\tau}}^{2,\bullet})^{(0)}]^{\Gamma}
Lemma 3.1.

If we construct such a ζτ\zeta_{\tau}, then for any integer c>1c>1, the restriction of ([c]c4)ζτ([c]^{*}-c^{4})\zeta_{\tau} to any ΓGL2()\Gamma\subset\mathrm{GL}_{2}(\mathbb{Z}) and TτTτ[c]TτT_{\tau}-T_{\tau}[c]\subset T_{\tau} represents the class Zτ(c)Z_{\tau}^{(c)}.

Proof.

If the total differential of ζτ\zeta_{\tau} is δ0volE\delta_{0}-\text{vol}_{E}, then the total differential of ([c]c4)ζτ([c]^{*}-c^{4})\zeta_{\tau} is δTτ[c]c4δ0\delta_{T_{\tau}[c]}-c^{4}\delta_{0}. Furthermore, it remains invariant by all isogenies prime to cc; hence by the discussion at the end of the previous section, the result follows. ∎

To construct this lift, we must specify ζτ0,1C0(GL2(),(𝒟Tτ2,1)(0))\zeta_{\tau}^{0,1}\in C^{0}(\mathrm{GL}_{2}(\mathbb{Q}),(\mathcal{D}_{T_{\tau}}^{2,1})^{(0)}) and ζτ1,0C1(GL2(),(𝒟Tτ2,0)(0))\zeta_{\tau}^{1,0}\in C^{1}(\mathrm{GL}_{2}(\mathbb{Q}),(\mathcal{D}_{T_{\tau}}^{2,0})^{(0)}) such that ¯ζτ1,0=GL2()ζτ0,1\overline{\partial}\zeta_{\tau}^{1,0}=\partial_{\mathrm{GL}_{2}(\mathbb{Q})}\zeta_{\tau^{0,1}} where the latter differential refers to the group coboundary map.

We therefore fix the choice

ζτ0,1:=E1(τ,z1)dz1δz2=0+E1(τ,z2)dz2volz1=0\zeta_{\tau}^{0,1}:=E_{1}(\tau,z_{1})\,dz_{1}\,\delta_{z_{2}=0}+E_{1}(\tau,z_{2})\,dz_{2}\wedge\text{vol}_{z_{1}=0}

whose image under ¯\overline{\partial} is can be computed as

(3.6) ¯ζτ0,1\displaystyle\overline{\partial}\zeta_{\tau}^{0,1} =¯[E1(τ,z2)dz2δz2=0+E1(τ,z1)dz1volz1=0]\displaystyle=\overline{\partial}[E_{1}(\tau,z_{2})\,dz_{2}\,\delta_{z_{2}=0}+E_{1}(\tau,z_{1})\,dz_{1}\wedge\text{vol}_{z_{1}=0}]
(3.7) =(δ0volz1=0δz2=0)+(volz1=0δz2=0volz1=0volz2=0)\displaystyle=(\delta_{0}-\text{vol}_{z_{1}=0}\,\delta_{z_{2}=0})+(\text{vol}_{z_{1}=0}\,\delta_{z_{2}=0}-\text{vol}_{z_{1}=0}\wedge\text{vol}_{z_{2}=0})
(3.8) =δ0volTτ\displaystyle=\delta_{0}-\text{vol}_{T_{\tau}}

It remains therefore to find, for each γGL2()\gamma\in\mathrm{GL}_{2}(\mathbb{Q}), a lift to (𝒟Tτ2,0)(0)(\mathcal{D}_{T_{\tau}}^{2,0})^{(0)} of

(γ1)ζτ0,1=(γ1)[E1(τ,z2)dz2δz2=0+E1(τ,z1)dz1volz1=0].(\gamma_{*}-1)\zeta_{\tau}^{0,1}=(\gamma_{*}-1)[E_{1}(\tau,z_{2})\,dz_{2}\,\delta_{z_{2}=0}+E_{1}(\tau,z_{1})\,dz_{1}\wedge\text{vol}_{z_{1}=0}].

We observe that there are no more choices to be made; these lifts are unique:

Lemma 3.2.

The trace-fixed complex

0(𝒟Tτ2,0)(0)(𝒟Tτ2,1)(0)(𝒟Tτ2,2)(0)0\to(\mathcal{D}^{2,0}_{T_{\tau}})^{(0)}\to(\mathcal{D}^{2,1}_{T_{\tau}})^{(0)}\to(\mathcal{D}^{2,2}_{T_{\tau}})^{(0)}

is left-exact.

Proof.

This can be proven identically to [SV, Lemma 6.2.1], as the only trace-fixed cohomology of ΩTτ2\Omega^{2}_{T_{\tau}} is in top degree, coming from the form yielding the fundamental class of the 44-torus TτT_{\tau} in the Hodge-de Rham spectral sequence.333In fact, one can show using Fourier series that this complex is almost right-exact as well, except for one dimension of cohomology on the right. This is unnecessary for us, so we omit it.

To find formulas for these lifts for a fully general matrix γ\gamma is complicated if approached directly: the pushforward action on currents can yield expressions with arbitrarily many terms. Instead, we decompose our matrices to simplify the calculation.

3.3. Telescoping with the Bruhat decomposition

The main observation we need is that we have the “telescoping” relation

(3.9) (γ2γ11)ζTτ0,1=γ2(γ11)ζTτ0,1+(γ21)ζTτ0,1(\gamma_{2}\gamma_{1}-1)\zeta^{0,1}_{T_{\tau}}=\gamma_{2}(\gamma_{1}-1)\zeta^{0,1}_{T_{\tau}}+(\gamma_{2}-1)\zeta^{0,1}_{T_{\tau}}

reducing the problem of finding a lift for the product of matrices to the problem for the individual matrices. Thus, the problem of finding lifts can be reduced to a set of generators of GL2()\mathrm{GL}_{2}(\mathbb{Q}).

We recall that the Bruhat decomposition says that

GL2()=BwB=(0)(0110)(0)\mathrm{GL}_{2}(\mathbb{Q})=BwB=\begin{pmatrix}*&*\\ 0&*\end{pmatrix}\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\begin{pmatrix}*&*\\ 0&*\end{pmatrix}

where BB denotes the upper triangular Borel subgroup, and ww the antidiagonal Weyl element. We can make this more explicit by writing

(3.10) γ:=(abcd)\displaystyle\gamma:=\begin{pmatrix}a&b\\ c&d\end{pmatrix} =(1a0c)(0110)(1d/c0detγc)\displaystyle=\begin{pmatrix}1&a\\ 0&c\end{pmatrix}\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\begin{pmatrix}1&d/c\\ 0&-\frac{\det\gamma}{c}\end{pmatrix}
(3.11) =(1a/c01)(100c)(0110)(1ddetγ01)(100detγc)\displaystyle=\begin{pmatrix}1&a/c\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ 0&c\end{pmatrix}\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\begin{pmatrix}1&-\frac{d}{\det\gamma}\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ 0&-\frac{\det\gamma}{c}\end{pmatrix}

whenever c0c\neq 0, for an arbitrary matrix γGL2()\gamma\in\mathrm{GL}_{2}(\mathbb{Q}). Here, we have further decomposed the appearing upper triangular matrices by factoring them into a diagonal times a unipotent matrix. When c=0c=0, our matrix is already in the Borel, so we can directly write

(ab0d)=(1b/d01)(a00d)\begin{pmatrix}a&b\\ 0&d\end{pmatrix}=\begin{pmatrix}1&b/d\\ 0&1\end{pmatrix}\begin{pmatrix}a&0\\ 0&d\end{pmatrix}

The point of this is that ww, diagonal, and unipotent matrices all act in very computable ways on ζTτ0,1\zeta^{0,1}_{T_{\tau}}, enabling us to find lifts:

  1. (1)

    All diagonal matrices act trivially on ζτ0,1\zeta_{\tau}^{0,1}, by the isogeny properties of E1E_{1} we proved earlier together with the fact that the volume form of a torus is isogeny-invariant. Hence the corresponding lifts are zero.

  2. (2)

    For the element ww, we can write (w1)ζTτ(w-1)\zeta_{T_{\tau}} as

    E1(τ,z2)dz2δz1=0+E1(τ,z1)dz1volz1=0E1(τ,z1)dz1δz2=0E1(τ,z2)dz2volz2=0E_{1}(\tau,z_{2})\,dz_{2}\,\delta_{z_{1}=0}+E_{1}(\tau,z_{1})\,dz_{1}\wedge\text{vol}_{z_{1}=0}-E_{1}(\tau,z_{1})\,dz_{1}\,\delta_{z_{2}=0}-E_{1}(\tau,z_{2})\,dz_{2}\wedge\text{vol}_{z_{2}=0}
    =¯[E1(τ,z1)E1(τ,z2)dz1dz2]=\overline{\partial}[E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})\,dz_{1}\wedge dz_{2}]

    which is a very simple lift.

  3. (3)

    For a unipotent matrix γu\gamma_{u} with upper-right entry uu, we find that the term with the current δz2=0\delta_{z_{2}=0} cancels in (γu1)ζTτ(\gamma_{u}-1)\zeta_{T_{\tau}} (since the coordinate z1z_{1} is fixed), and what remains is

    (3.12) 2iuyE1(τ,z2)dz1dz2dz2¯=¯[4uE2(τ,z2)dz1dz2]\frac{2iu}{y}E_{1}(\tau,z_{2})\,dz_{1}\wedge dz_{2}\wedge d\overline{z_{2}}=\overline{\partial}\left[4uE_{2}(\tau,z_{2})\,dz_{1}\wedge dz_{2}\right]

Combining these with our explicit Bruhat decomposition, we find the following formulas: in the case when c=0c=0, we obtain

(3.13) ζTτ1,0(γ)=4bdE2(τ,z2)dz1dz2.\zeta^{1,0}_{T_{\tau}}(\gamma)=\frac{4b}{d}E_{2}(\tau,z_{2})\,dz_{1}\wedge dz_{2}.

In the case when c0c\neq 0, we use the full Bruhat decomposition to obtain the following:

Proposition 3.3.

The unique lift ζTτ1,0\zeta^{1,0}_{T_{\tau}} on a matrix γGL2()\gamma\in\mathrm{GL}_{2}(\mathbb{Q}) is given by the expression

(3.14) [(a1c0)4dcdetγE2(τ,z2)+1c(1a0c)E1(τ,z1)E1(τ,z2)+4acE2(τ,z2)]dz1dz2.\left[\begin{pmatrix}a&1\\ c&0\end{pmatrix}_{*}\frac{4d}{c\det\gamma}E_{2}(\tau,z_{2})+\frac{1}{c}\begin{pmatrix}1&a\\ 0&c\end{pmatrix}_{*}E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})+\frac{4a}{c}E_{2}(\tau,z_{2})\right]\,dz_{1}\wedge dz_{2}.

Notice that here, we have to introduce factors coming from the determinants of the matrices to go from pushforwards of forms to pushforwards of functions.

We therefore have obtained full formulas for ζTτ\zeta_{T_{\tau}}, whose cc-stabilizations yield the kernel classes Zτ(c)Z_{\tau}^{(c)} for the Eisenstein theta lift after restriction to GL2()\mathrm{GL}_{2}(\mathbb{Z}).

For a general matrix γ\gamma, there does not appear to be a substantial further simplification of the preceding formulas, besides writing out the pushforwards as sums.

3.4. Specializations at torsion points

The actual “theta lift” for (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2}) is generally taken to be valued after contraction pullback by torsion sections. Fix an arithmetic subgroup ΓGL2()\Gamma\subset\mathrm{GL}_{2}(\mathbb{Z}); in order to compare our results with the cohomological theta lift, we will restrict ζτ\zeta_{\tau} to Γ\Gamma.

Remark 3.4.

This arithmeticity hypothesis is only necessary to compare with the cohomological construction of the theta lift: one can still obtain pulled-back cocycles on, say, SS-arithmetic subgroups for some set of inverted places SS, so long as they fix some torsion sections (of order necessarily prime to the places in SS). We do not write down this extension here, but the formulas can be obtained by our same methods.

In this context, we wish to consider the image of Zτ(c)Z_{\tau}^{(c)} under the composite

(3.15) HΓ1(TτTτ[c],ΩTτTτ[c]2)z1z2HΓ1(TτTτ[c],𝒪TτTτ[c])DHΓ1({τ},)=H1(Γ,)H^{1}_{\Gamma}(T_{\tau}-T_{\tau}[c],\Omega^{2}_{T_{\tau}-T_{\tau}[c]})\xrightarrow{\cap\frac{\partial}{\partial z_{1}}\otimes\frac{\partial}{\partial z_{2}}}H^{1}_{\Gamma}(T_{\tau}-T_{\tau}[c],\mathcal{O}_{T_{\tau}-T_{\tau}[c]})\xrightarrow{D^{*}}H^{1}_{\Gamma}(\{\tau\},\mathbb{C})=H^{1}(\Gamma,\mathbb{C})

where DD is any Γ\Gamma-fixed torsion cycle DD disjoint from Tτ[c]T_{\tau}[c]. (Here, we are slightly abusive in writing the pullback DD^{*}; this is actually a sum of pullbacks over the various torsion sections in the support of DD.) Analogous to the case of a single torsion section, we will write Θτ,D(c)\Theta_{\tau,D}^{(c)} for this image.

We wish to interpret this composite in terms of the explicit double complex representative ζτ\zeta_{\tau}; the main issue is that currents cannot in general be pulled back by closed immersions.

The technical tool we need to remedy this is the introduction of a variant of the distributional Dolbeault complex: For any ΓGL2()\Gamma\subset\mathrm{GL}_{2}(\mathbb{Q}), let HΓTτH_{\Gamma}\subset T_{\tau} be the Γ\Gamma-orbit of the lines {z1=0}\{z_{1}=0\} and {z2=0}\{z_{2}=0\} and their translates by cc-torsion sections inside TτT_{\tau}; this is a union of infinitely many elliptic subschemes (and their cc-torsion translates). For any finite subarrangement HHΓH\subset H_{\Gamma}, we define a complex 𝒟Tτ,H2,\mathcal{D}_{T_{\tau},H}^{2,\bullet} via the pullback square

(3.16) 𝒟Tτ,H{\mathcal{D}_{T_{\tau},H}^{\bullet}}𝒟Tτ2,{\mathcal{D}_{T_{\tau}}^{2,\bullet}}𝒜TτH2,{\mathcal{A}_{T_{\tau}-H}^{2,\bullet}}𝒟TτH2,{\mathcal{D}_{T_{\tau}-H}^{2,\bullet}}

which results in an identification of 𝒟Tτ,H2,i\mathcal{D}_{T_{\tau},H}^{2,i} with the (2,i)(2,i)-currents such that their restriction to TτHT_{\tau}-H are given by smooth (2,i)(2,i)-forms. Here, the bottom horizontal map is the earlier-defined inclusion, and the right vertical map is the restriction dual to the pushforward of compactly-supported differential forms. We define then

𝒟Tτ,HΓ2,:=limH𝒟Tτ,H2,\mathcal{D}_{T_{\tau},H_{\Gamma}}^{2,\bullet}:=\varinjlim_{H}\mathcal{D}_{T_{\tau},H}^{2,\bullet}

where the limit runs over HH finite subarrangements of HΓH_{\Gamma}, along the natural inclusion maps. The group Γ\Gamma permutes the pullback diagrams for each HH (sending it to that of γH\gamma H), and these assemble to give a pushforward action on 𝒟Tτ,HΓ2,\mathcal{D}_{T_{\tau},H_{\Gamma}}^{2,\bullet}. Further, because the bottom row in each pullback diagram is a quasi-isomorphism, we see that 𝒟Tτ,HΓ2,\mathcal{D}_{T_{\tau},H_{\Gamma}}^{2,\bullet} computes the cohomology of ΩTτ2\Omega^{2}_{T_{\tau}}, just as 𝒟Tτ2,\mathcal{D}_{T_{\tau}}^{2,\bullet} does. Furthermore, analogously to the full distributional de Rham complex, we have a left exact sequence

0(𝒟Tτ,HΓ2,0)(0)(𝒟Tτ,HΓ2,1)(0)(𝒟Tτ,HΓ2,2)(0)0\to(\mathcal{D}_{T_{\tau},H_{\Gamma}}^{2,0})^{(0)}\to(\mathcal{D}_{T_{\tau},H_{\Gamma}}^{2,1})^{(0)}\to(\mathcal{D}_{T_{\tau},H_{\Gamma}}^{2,2})^{(0)}

meaning that ζτ\zeta_{\tau}, considered in this more refined complex, is still uniquely determined. The important new phenomenon for us is that if Γ\Gamma fixes the torsion cycle DTτD\subset T_{\tau} disjoint from HΓH_{\Gamma}, then there is a composite pullback map

𝒟Tτ,HΓ2,0limHΩTτH2z1z2𝒪TτHD\mathcal{D}_{T_{\tau},H_{\Gamma}}^{2,0}\to\varinjlim_{H}\Omega^{2}_{T_{\tau}-H}\xrightarrow{\cap\frac{\partial}{\partial z_{1}}\otimes\frac{\partial}{\partial z_{2}}}\mathcal{O}_{T_{\tau}-H}\xrightarrow{D^{*}}\mathbb{C}

which induces the composite (3.15).

We now concern ourselves with the case that DD is supported on NN-torsion for an integer N>1N>1. From the preceding discussion, we can conclude that

[D([c]c4)ζτ]=Θτ,D(c)[D^{*}([c]^{*}-c^{4})\zeta_{\tau}]=\Theta_{\tau,D}^{(c)}

so long as DD is disjoint from HΓH_{\Gamma}. Note that if c1(modN)c\equiv 1\pmod{N}, we can write the left-hand side as

(1c4)[Dζτ].(1-c^{4})[D^{*}\zeta_{\tau}].

The restriction on DD is rather irritating, as it depends on the arbitrary choice of coordinates z1,z2z_{1},z_{2} we used to choose our lift ζτ0,1\zeta_{\tau}^{0,1}: for any given torsion section, we could simply start with a different lift to obtain formulas for the pullback. However, this would result in a somewhat unsatisfying lack of unity in our formulas. Luckily, we can use a trick to bypass this issue entirely, and make the formulas valid even for “bad” torsion sections:

Theorem 3.5.

Suppose c1(modN)c\equiv 1\pmod{N}. Then for any NN-torsion cycle DD disjoint from the identity, we have that

[Dζτ]=11c4Θτ,D(c)[D^{*}\zeta_{\tau}]=\frac{1}{1-c^{4}}\Theta_{\tau,D}^{(c)}

The proof consists of “bootstrapping” from torsion points disjoint from HΓH_{\Gamma} to all of them. In order to do this, we will need the following lemma allowing us to “improve” our current-valued cocycles to be form-valued:

Lemma 3.6.

There is an injection

Mτ(𝒟Tτ,HΓ2,0)(0)M_{\tau}\hookrightarrow(\mathcal{D}^{2,0}_{T_{\tau},H_{\Gamma}})^{(0)}

where MτM_{\tau} is defined to be the module of (2,0)(2,0)-forms on Tτ{0}T_{\tau}-\{0\} spanned by the GL2()\mathrm{GL}_{2}(\mathbb{Q})-orbit of E2(z2)dz2E_{2}(z_{2})\,dz_{2} and E1(τ,z1)E1(τ,z2)dz1dz2E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})\,dz_{1}\wedge dz_{2}, given by sending

ω(ηTωη).\omega\mapsto\left(\eta\mapsto\int_{T}\omega\wedge\eta\right).
Proof of lemma.

The only non-formal assertion here is that this map is injective. Notice that the map considering smooth forms (i.e. the map νTτ\nu_{T_{\tau}} defined before) as currents via kernels of integration, or even continuous forms, is clearly injective: by integrating against dz¯1dz¯2d\overline{z}_{1}\wedge d\overline{z}_{2} times a bump function on any small open set, we see that the zero current can only come from a form which vanishes almost everywhere, which hence must be zero by continuity. Thus, the depth of this lemma’s assertion comes precisely from the discontinuities of the forms in the orbit of E1(τ,z1)E1(τ,z2)dz1dz2E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})\,dz_{1}\wedge dz_{2} along codimension-11 sub-elliptic curves.444Indeed, to appreciate the delicacy, observe if we change Tτ{0}T_{\tau}-\{0\} to TT in its statement, the statement becomes false: see [BG, Proposition 3.7] for an example of a relation between the weight-11 and weight-22 series everywhere except the zero section.

We note the following property of E1(τ,z1)E1(τ,z2)dz1dz2E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})\,dz_{1}\wedge dz_{2}: suppose x=(x1,x2)Tτx=(x_{1},x_{2})\in T_{\tau} is a point lying on one of the subcurves of discontinuity STτS\subset T_{\tau} (so x1=0x_{1}=0 or x2=0x_{2}=0) but not equal to zero. Then take any small v=(v1,v2)2v=(v_{1},v_{2})\in\mathbb{C}^{2} not parallel to the curve of discontinuity of xx, so that x±vx\pm v does not lie in SS. By the oddness of E1E_{1}, we find that the average of the translates by ±v\pm v vanishes as we shrink vv:

limϵ0+12(E1(τ,z1+ϵv1)E1(τ,z2+ϵv2)+E1(τ,z1ϵv1)E1(τ,z2ϵv2))=0\lim_{\epsilon\to 0^{+}}\frac{1}{2}(E_{1}(\tau,z_{1}+\epsilon v_{1})E_{1}(\tau,z_{2}+\epsilon v_{2})+E_{1}(\tau,z_{1}-\epsilon v_{1})E_{1}(\tau,z_{2}-\epsilon v_{2}))=0

as an equality of coefficients of dz1dz2dz_{1}\wedge dz_{2}.

By moving this argument around by the general linear action, this applies to any nonzero point on a codimension 11 discontinuity stratum of a function in the orbit of E1(τ,z1)E1(τ,z2)dz1dz2E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})\,dz_{1}\wedge dz_{2}.

Now consider an arbitrary ωMτ\omega\in M_{\tau}, and suppose that ω\omega gives the trivial (2,0)(2,0)-current when considered as a kernel of integration. Then, ω\omega must be identically zero outside a finite union of sub-elliptic curves through the identity. Consider an arbitrary nonzero point xx on one of these subcurves STτS\subset T_{\tau}, and pick some decomposition

ω=ω1+ω2\omega=\omega_{1}+\omega_{2}

where ω1\omega_{1} consists of a sum of terms in the orbit of E2(τ,z2)dz1dz2E_{2}(\tau,z_{2})\,dz_{1}\wedge dz_{2} or E1(τ,z1)E1(τ,z2)dz1dz2E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})\,dz_{1}\wedge dz_{2} which do not have a discontinuity along SS, and ω2\omega_{2} consists of a sum of terms from the latter orbit which do have a discontinuity along SS. Pick now some vector vCv\in C such that the the line segment between the points x±vx\pm v intersects no other discontinuity locus of any term in ω1\omega_{1} or ω2\omega_{2}; this is always possible since vv is nonzero and there are only finitely many terms to consider (and hence subcurves to avoid).

Then we find that

(3.17) 12(limϵ0+ω2|x+ϵv+ω2|xϵv)\displaystyle\frac{1}{2}\left(\lim_{\epsilon\to 0^{+}}\omega_{2}|_{x+\epsilon v}+\omega_{2}|_{x-\epsilon v}\right) =12(limϵ0+ω|x+ϵv[ω1]|x+ϵv+ω|xϵv[ω1]|xϵv)\displaystyle=\frac{1}{2}\left(\lim_{\epsilon\to 0^{+}}\omega|_{x+\epsilon v}-[\omega_{1}]|_{x+\epsilon v}+\omega|_{x-\epsilon v}-[\omega_{1}]|_{x-\epsilon v}\right)
(3.18) =12(limϵ0+[ω1]|x+ϵv[ω1]|xϵv)\displaystyle=\frac{1}{2}\left(\lim_{\epsilon\to 0^{+}}-[\omega_{1}]|_{x+\epsilon v}-[\omega_{1}]|_{x-\epsilon v}\right)

since ω\omega is identically zero on a neighborhood of xx in TτST_{\tau}-S. But ω1\omega_{1} is continuous in a neighborhood of xx by assumption, so this expression is just ω1|z=x-\omega_{1}|_{z=x}. On the other hand, the average of the two limits we started with is zero from the preceding discussion, so we conclude that ω1|z=x=0\omega_{1}|_{z=x}=0. We also have ω2|z=x=0\omega_{2}|_{z=x}=0 because forms in the orbit of E1(τ,z1)E1(τ,z2)dz1dz2E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})\,dz_{1}\wedge dz_{2} are zero along their discontinuity loci by construction. We hence conclude that ω|z=x=0\omega|_{z=x}=0; since this applies to any nonzero point xx, we conclude that ω\omega is the zero form on Tτ{0}T_{\tau}-\{0\}. This concludes the proof of injectivity. ∎

Proof of theorem.

Thanks to the lemma, we can consider ζτ\zeta_{\tau} to be a cocycle valued in MτM_{\tau}. Let Tτ[N]T_{\tau}[N]^{\prime} denote the primitive NN-torsion, and let STτS\subset T_{\tau} be a sub-elliptic curve which is not the vanishing locus of either z1z_{1} or z2z_{2}. Let x0S[N]x_{0}\in S[N]^{\prime} be any point, and write Γ1(x0)GL2()\Gamma_{1}(x_{0})\subset\mathrm{GL}_{2}(\mathbb{Z}) for its stabilizer. By construction, xHΓ1(x0)x\not\in H_{\Gamma_{1}(x_{0})}, and so by the previous discussion,

(3.19) [x0ζτ]=11c4Θτ,x0(c)[x_{0}^{*}\zeta_{\tau}]=\frac{1}{1-c^{4}}\Theta^{(c)}_{\tau,x_{0}}

for any c1(modN)c\equiv 1\pmod{N}. In fact, noticing that Γ1(x0)\Gamma_{1}(x_{0}) must stabilize the entire elliptic sub-curve SS, this formula holds for any point xS[N]x\in S[N]^{\prime}.

We now observe that zτ(c)z1z2z_{\tau}^{(c)}\cap\frac{\partial}{\partial z_{1}}\otimes\frac{\partial}{\partial z_{2}} induces, by restriction, a class

(3.20) Θτ,N(c)HGL2()1(Tτ[N],)H1(GL2(),hom(Tτ[N],))\Theta_{\tau,N}^{(c)}\in H^{1}_{\mathrm{GL}_{2}(\mathbb{Z})}(T_{\tau}[N]^{\prime},\mathbb{C})\cong H^{1}(\mathrm{GL}_{2}(\mathbb{Z}),\hom(T_{\tau}[N]^{\prime},\mathbb{C}))

where the isomorphism comes from the fact that Tτ[N]T_{\tau}[N]^{\prime} is a union of contractible spaces, causing the collapse of the Hochschild-Serre spectral sequence. By functoriality of this spectral sequence, for any cycle DTτ[N]D\subset T_{\tau}[N]^{\prime} stabilized by Γ(D)\Gamma(D), the image of Θτ,N(c)\Theta_{\tau,N}^{(c)} under the composite of restriction and evaluation

H1(GL2(),hom(Tτ[N],))resH1(Γ(D),hom(Tτ[N],))𝐷H1(Γ(D),)H^{1}(\mathrm{GL}_{2}(\mathbb{Z}),\hom(T_{\tau}[N]^{\prime},\mathbb{C}))\xrightarrow{\text{res}}H^{1}(\Gamma(D),\hom(T_{\tau}[N]^{\prime},\mathbb{C}))\xrightarrow{D}H^{1}(\Gamma(D),\mathbb{C})

yields Θτ,D(c)\Theta^{(c)}_{\tau,D}.

Observe that there is a GL2()\mathrm{GL}_{2}(\mathbb{Z})-equivariant map

Mτhom(Tτ[N],),fdz1dz2(xf(x)).M_{\tau}\to\hom(T_{\tau}[N]^{\prime},\mathbb{C}),f\,dz_{1}\wedge dz_{2}\mapsto\left(x\mapsto f(x)\right).

We claim that the pushforward of ([c]c4)ζτ([c]^{*}-c^{4})\zeta_{\tau} under this map can be identified with Θτ,N(c)\Theta_{\tau,N}^{(c)}, which would then imply the desired result for arbitrary primitive NN-torsion cycles.

Indeed, there is an isomorphism of GL2()\mathrm{GL}_{2}(\mathbb{Z})-modules

hom(Tτ[N],)IndΓ1(x0)GL2()hom(S[N],),f(γfγ1)\hom(T_{\tau}[N]^{\prime},\mathbb{C})\to\text{Ind}_{\Gamma_{1}(x_{0})}^{\mathrm{GL}_{2}(\mathbb{Z})}\,\hom(S[N]^{\prime},\mathbb{C}),f\mapsto\left(\gamma\mapsto f\circ\gamma^{-1}\right)

where hom(S[N],)\hom(S[N]^{\prime},\mathbb{C}) has, naturally, a trivial action of GL2()\mathrm{GL}_{2}(\mathbb{Z}). Hence, by Shapiro’s lemma it suffices to show that ([c]c4)ζτ([c]^{*}-c^{4})\zeta_{\tau} and Θτ,N(c)\Theta_{\tau,N}^{(c)} agree upon restriction to Γ1(x0)\Gamma_{1}(x_{0}) under the quotient

hom(Tτ[N],)hom(S[N],)\hom(T_{\tau}[N]^{\prime},\mathbb{C})\twoheadrightarrow\hom(S[N]^{\prime},\mathbb{C})

dual to the obvious inclusion. But this is precisely (3.19), which we have already established. Assembling these identifications together for all N>1N>1 yields the full theorem. ∎

Thus, from (3.14), when specialized at any nonzero torsion sections (or combination thereof), yields the following formula for the Eisenstein theta lift of [BCG1]:

Theorem 3.7.

Let

(3.21) θτ[γ](z1,z2):={4bdE2(τ,z2)if c=0, else(a1c0)4dcdetγE2(τ,z2)+1c(1a0c)E1(τ,z1)E1(τ,z2)+4acE2(τ,z2)\theta_{\tau}[\gamma](z_{1},z_{2}):=\begin{cases}\frac{4b}{d}E_{2}(\tau,z_{2})&\text{if $c=0$, else}\\ \begin{pmatrix}a&1\\ c&0\end{pmatrix}_{*}\frac{4d}{c\det\gamma}E_{2}(\tau,z_{2})+\frac{1}{c}\begin{pmatrix}1&a\\ 0&c\end{pmatrix}_{*}E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})+\frac{4a}{c}E_{2}(\tau,z_{2})\end{cases}

Then given any Γ\Gamma-fixed combination of nonzero torsion sections

D=iti[(ui,vi)],D=\sum_{i}t_{i}[(u_{i},v_{i})],

we have

Θτ,D(c)(γ)=iciθτ[γ](a1,b2).\Theta^{(c)}_{\tau,D}(\gamma)=\sum_{i}c_{i}\theta_{\tau}[\gamma](a_{1},b_{2})\in\mathbb{C}.

Here, tit_{i} are integer coefficients, and uiu_{i} and viv_{i} are elements of (/N)2{(0,0)}(\mathbb{Z}/N)^{2}-\{(0,0)\}, thought of as NN-torsion sections on EτE_{\tau}.

Noticing that θτ\theta_{\tau} transforms like a weight-22 modular form in τ\tau, we can consider its specialization at torsion sections as a section of the weight-22 automorphic line bundle ω2\omega^{2} on any open modular curve over which the torsion sections are defined. It immediately follows:

Corollary 3.8.

If HH is any level structure fixing the torsion cycle DD, then with the same notation as above, we have

ΘΓ,D(c)(γ)=iciθτ[γ](ui,vi)H0(Y(H),ω2).\Theta^{(c)}_{\Gamma,D}(\gamma)=\sum_{i}c_{i}\theta_{\tau}[\gamma](u_{i},v_{i})\in H^{0}(Y(H),\omega^{2}).

where Y(H)Y(H) is the open modular curve of level HH.555It is also true that θτ\theta_{\tau} represents cocycles over distributions of torsion sections, where Γ\Gamma now acts nontrivially by permuting the sections. Because of the way we set up our machinery in this article, this is not immediate; however, it can be proven with only a little extra work.

These formulas are workable, but the presence of the pushforward matrices (which can be evaluated as finite sums over preimage torsion sections) make them slightly unwieldy. Analogously to the classical setting of periods of Eisenstein series [X4], the first term

iciθτ[γ](ui,vi)=4dcdetγici[(a1c0)E2(τ,z2)](ui,vi)\sum_{i}c_{i}\theta_{\tau}[\gamma](u_{i},v_{i})=\frac{4d}{c\det\gamma}\sum_{i}c_{i}\left[\begin{pmatrix}a&1\\ c&0\end{pmatrix}_{*}E_{2}(\tau,z_{2})\right]_{(u_{i},v_{i})}

can be simplified if we assume that γSL2()\gamma\in SL_{2}(\mathbb{Z}):

(3.22) 4dcdetγici[(a1c0)E2(τ,z2)](ui,vi)\displaystyle\frac{4d}{c\det\gamma}\sum_{i}c_{i}\left[\begin{pmatrix}a&1\\ c&0\end{pmatrix}_{*}E_{2}(\tau,z_{2})\right]_{(u_{i},v_{i})} =4dcicij,k(/c)2E2(τ,uiac(vi+(j,k)))\displaystyle=4\frac{d}{c}\sum_{i}c_{i}\sum_{j,k\in(\mathbb{Z}/c)^{2}}E_{2}\left(\tau,u_{i}-\frac{a}{c}(v_{i}+(j,k))\right)
(3.23) =4dcicij,k(/c)2E2(τ,1c(vi+(j,k)))\displaystyle=4\frac{d}{c}\sum_{i}c_{i}\sum_{j,k\in(\mathbb{Z}/c)^{2}}E_{2}\left(\tau,\frac{1}{c}(v_{i}+(j,k))\right)
(3.24) =4dciciE2(τ,vi)\displaystyle=\frac{4d}{c}\sum_{i}c_{i}E_{2}(\tau,v_{i})

Here, we use the distribution property of E2E_{2}, along with the fact that γ1\gamma^{-1} stabilizes DD. This latter fact implies that for all iIi\in I,

avicui=vσ(i)av_{i}-cu_{i}=v_{\sigma(i)}

for some permutation σ\sigma of the index set II such that ci=cσ(i)c_{i}=c_{\sigma(i)} for all iIi\in I.

Unfortunately, we do not see a natural way to simplify the E1E1E_{1}E_{1} term in any generality, analogously to the classical formulas for Eisenstein periods we discussed in [X4].

Remark 3.9.

Note that the value at τ=\tau=\infty theta lift θτ\theta_{\tau} yields precisely the classical formula for the weight-22 Eisenstein cocycle reproven in loc. cit; this follows immediately from the fact that at τ=\tau=\infty, the series Ei(τ,z)E_{i}(\tau,z) degenerates to the periodic Bernoulli polynomial Bi(z)2i\frac{B_{i}(z)}{2i}: this is immediate from the description of both functions by Hecke regularized (analytic continuation in ss), since it holds for ss with large enough real part that the Fourier series are absolutely convergent.

Hence as expected, the (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2})-Eisenstein theta lift’s degeneration at a cusp yields the (GL2,GL1)(\mathrm{GL}_{2},\mathrm{GL}_{1}) theta lift (in the sense described in [BCG1, §13]).

Remark 3.10.

Instead of a formula, one can obtain a more efficient ”continued fraction” algorithm for computing the lifts for matrices in SL2()GL2()\mathrm{SL}_{2}(\mathbb{Z})\subset\mathrm{GL}_{2}(\mathbb{Q}) by using its famous generators SS and TT and the recursion principle (3.9): this is presented for Eisenstein cocycles presented in [Scz1, §2.4], but works identically here by replacing the Bernoulli polynomials with our Eisenstein-Kronecker series. The outputs of this algorithm will coincide with the preceding formulas by uniqueness of ζTτ1,0\zeta^{1,0}_{T_{\tau}}, though this is not visibly obvious.

4. Some properties and applications

4.1. Hecke equivariance

As one expects for a theta lift, the cocycles θτ\theta_{\tau}, considered for the group SL2()\mathrm{SL}_{2}(\mathbb{Z}), satisfy a compatbility property between two kinds of Hecke operators: geometric Hecke operators coming from the variable τ\tau in the upper half-plane with its SL2()\mathrm{SL}_{2}(\mathbb{Z})-action, and a cohomological Hecke action coming from the matrix action fiberwiise.

Using our algebraic approach, one could prove this compatibility analogously to the approach in [SV, §6]. However, since a form of Hecke compatibility was already proven in [BCG3, Théorème 2.8] for a closely related cocycle, it is much easier for us to simply to import this result using our already-proven comparison.

To fix ideas, in this section we will consider the restriction of θτ\theta_{\tau} to Γ:=Γ1(N)SL2()\Gamma:=\Gamma_{1}(N)\subset\mathrm{SL}_{2}(\mathbb{Z}) for some integer N>1N>1, and a torsion section x=(0,x2):E2x=(0,x_{2}):\mathcal{H}\to E^{2} which descends to level Y1(N)Y_{1}(N). The below approach can be applied to broader contexts, but in this article we will remain in this setting.

We recall the definition of two kinds of Hecke operators for GL2\mathrm{GL}_{2} acting on [θτ][\theta_{\tau}]: a fiberwise action coming from group cohomology, and a geometric action coming from the Möbius action on τ\tau.

We write Δ\Delta to be the monoid of rank-22 integral matrices which stabilize (1,0)(/N)2(1,0)\in(\mathbb{Z}/N)^{2} for the standard left representation, so that ΓΔ\Gamma\subset\Delta. Given any double Γ\Gamma-coset in Δ\Delta, we can decompose it finitely as

ΓαΓ=iαiΓ.\Gamma\alpha\Gamma=\bigcup_{i}\alpha_{i}\Gamma.

As always, there are two different Δ\Delta-actions we need to consider: first, the “fiberwise” action, where γ\gamma acts by

(|detγ|γ1),(|\det\gamma|\gamma^{-1})^{*},

this choice made so that for γSL2()\gamma\in\mathrm{SL}_{2}(\mathbb{Z}) it coincides with the pushforward we have heretofore been considering, and the “modular” action, which sends

γ(τ,z1,z2):=(aτ+bcτ+d,z1cτ+d,z2cτ+d).\gamma\cdot(\tau,z_{1},z_{2}):=\left(\frac{a\tau+b}{c\tau+d},\frac{z_{1}}{c\tau+d},\frac{z_{2}}{c\tau+d}\right).

Then for any double coset ΓαΓ\Gamma\alpha\Gamma, the action of ΓαΓ\Gamma\alpha\Gamma on 11-cocyles can be defined as in [RW] (or [BCG3, §2.2.1]) by sending a 11-cocycle c:ΓMc:\Gamma\to M valued in a Δ\Delta-module MM to

γiαic(γi)\gamma\mapsto\sum_{i}\alpha_{i}c(\gamma_{i})

where γi\gamma_{i} is defined by the relation αiγ=γiασ(i)\alpha_{i}\gamma=\gamma_{i}\alpha_{\sigma(i)} for some permutation σ\sigma of the representatives αi\alpha_{i}. (Note that our conventions differ slightly from loc. cit, both here and for the pullback action of Δ\Delta; these two changes result in the same Hecke action.)

On invariants (with the modular or fiberwise action), the action of ΓαΓ\Gamma\alpha\Gamma is simpler to define, sending an element xMΓx\in M^{\Gamma} to

iαix.\sum_{i}\alpha_{i}x.

If we denote the fiberwise Hecke operator by T(α)T(\alpha) and the modular one by 𝐓(α)\mathbf{T}(\alpha), then (4), (5) of [BCG3, Théorème 2.8] tell us that

T(α)θτ,D=T(α)θτ,T(α)D,𝐓(α)θτ,D=T(α)θτ,𝐓(α)DT(\alpha)\circ\theta_{\tau,D}=T(\alpha)\circ\theta_{\tau,T(\alpha)D},\mathbf{T}(\alpha)\circ\theta_{\tau,D}=T(\alpha)\circ\theta_{\tau,\mathbf{T}(\alpha)D}

for any αΔ\alpha\in\Delta. In particular, let TpT_{p} and 𝐓p\mathbf{T}_{p} be the double coset operators associated to a prime pp, consisting of all matrices in Δ\Delta with determinant pp. When pp is relatively prime to NN, these form just a single double coset, else they may be a sum of multiple such operators.

Write δp\delta_{p} for the torsion cycle comprised of all pp-torsion points ((α1,β1),(α2,β2))Tτ[p]((\alpha_{1},\beta_{1}),(\alpha_{2},\beta_{2}))\in T_{\tau}[p] such that (α1,β1)(\alpha_{1},\beta_{1}) and (α2,β2)(\alpha_{2},\beta_{2}) are linearly dependent over /p\mathbb{Z}/p. Then we can compute that for any auxiliary integer c>1c>1, we have

Tp(Tτ[c]c4{0})=𝐓p(Tτ[c]c4{0})=([c]c4)(δp+p{0}).T_{p}(T_{\tau}[c]-c^{4}\{0\})=\mathbf{T}_{p}(T_{\tau}[c]-c^{4}\{0\})=([c]^{*}-c^{4})(\delta_{p}+p\{0\}).

Pulling back this equality by xx (since this commutes with the pullback action of Δ\Delta), we obtain:

Proposition 4.1.

We have the equality Tpθτ,x=𝐓pθτ,xT_{p}\theta_{\tau,x}=\mathbf{T}_{p}\theta_{\tau,x} for all primes pp, i.e.

θτ,x:H1(Γ1(N),)H0(Y1(N),ω2)\theta_{\tau,x}:H_{1}(\Gamma_{1}(N),\mathbb{Z})\to H^{0}(Y_{1}(N),\omega^{\otimes 2})

is equivariant for the Hecke subalgebra generated by {Tp}p\{T_{p}\}_{p} for the fiberwise, respectively modular Hecke actions on source and target.

Note in particular that this includes the entire anemic Hecke algebra (all operators of level prime to NN), but does not necessarily include UpU_{p} for pp dividing NN.

Using this Hecke equivariance, one can obtain explicit spectral decompositions of the (GL2,GL2)(\mathrm{GL}_{2},\mathrm{GL}_{2})-theta lift, by using the Rankin-Selberg formula for the inner product of an eigenform with the E1E1E_{1}E_{1} terms in (3.21). However, we will pursue this via a more systematic approach in future work, so do not go into it here.

4.2. CM elliptic curves

Suppose now that τ\tau satisfies a quadratic equation with rational coefficients; then the corresponding elliptic curve EτE_{\tau} has complex multiplication by an order 𝒪\mathcal{O} in K:=(τ)K:=\mathbb{Q}(\tau). Then we can extend the action of GL2()\mathrm{GL}_{2}(\mathbb{Z}) on TτT_{\tau} to an action of GL2(𝒪)\mathrm{GL}_{2}(\mathcal{O}), and therefore the action of GL2()\mathrm{GL}_{2}(\mathbb{Q}) on the trace-fixed distributional de Rham complex to an action of GL2(K)\mathrm{GL}_{2}(K). Note that in this case, we can take “trace-fixed” to include all isogenies built out of the “scalar” endomorphisms in 𝒪\mathcal{O}, because (3.4) generalizes to these isogenies [BK, Proposition 1.1.6].

Since this latter group has a Bruhat decomposition

GL2(K)=BK(11)BK\mathrm{GL}_{2}(K)=B_{K}\begin{pmatrix}&1\\ 1&\end{pmatrix}B_{K}

(where here BKB_{K} denotes the upper triangular Borel of this group) exactly as over \mathbb{Q}, the arguments of section 3 go through exactly as before, with the small detail that in equation 3.12, the lift must be

4u¯E2(τ,z2)dz1dz2.4\overline{u}E_{2}(\tau,z_{2})\,dz_{1}\wedge dz_{2}.

We hence conclude that the corresponding map given by

γθτ[γ](z1,z2):={4bdE2(τ,z2)if c=0, else(a1c0)4d¯cdetγ¯E2(τ,z2)+1c(1a0c)E1(τ,z1)E1(τ,z2)+4a¯c¯E2(τ,z2)\gamma\mapsto\theta_{\tau}[\gamma](z_{1},z_{2}):=\begin{cases}\frac{4b}{d}E_{2}(\tau,z_{2})&\text{if $c=0$, else}\\ \begin{pmatrix}a&1\\ c&0\end{pmatrix}_{*}\frac{4\overline{d}}{\overline{c\det\gamma}}E_{2}(\tau,z_{2})+\frac{1}{c}\begin{pmatrix}1&a\\ 0&c\end{pmatrix}_{*}E_{1}(\tau,z_{1})E_{1}(\tau,z_{2})+\frac{4\overline{a}}{\overline{c}}E_{2}(\tau,z_{2})\end{cases}

is a cocycle for GL2(K)\mathrm{GL}_{2}(K) valued in functions on Tτ{0}T_{\tau}-\{0\}, whose restriction to GL2(𝒪)\mathrm{GL}_{2}(\mathcal{O}) comes from equivariant polylogarithm class. Just as with the GL2()\mathrm{GL}_{2}(\mathbb{Q}) cocycle, this can be specialized at various torsion points. In this case, these specializations are just numbers instead of varying over an underlying symmetric space, so the “big cocycle” valued in forms on TτT_{\tau} may be of primary interest.

This imaginary quadratic cocycle can be viewed as being “for the dual pair (GL2(K),GL1(K))(\mathrm{GL}_{2}(K),\mathrm{GL}_{1}(K)),” and is approached analytically in the work [BCG2]. In particular, our formula above gives a simple expression in terms of Eisenstein-Kronecker numbers of the weight-(0,0)(0,0) cocycle denoted Φ0,0\Phi^{0,0} in loc. cit., and therefore also the values of Hecke LL-functions associated to weight-0 characters for the field KK, as in [BCG2, Theorem 1.2]. As in the GL2()\mathrm{GL}_{2}(\mathbb{Q}) case, by employing twisted versions of our complexes and taking different weight Eisenstein-Kronecker series, it is possible to obtain analogous formulas for the more general cocycles Φp,q\Phi^{p,q}, which could be an interesting direction of future work.

References

  • [BCG1] Nicolas Bergeron, Pierre Charollois and Luis Garcia “Transgressions of the Euler class and Eisenstein cohomology of GLN()\text{GL}_{N}(\mathbb{Z}).” In Japanese Journal of Mathematics 15, 2020, pp. 311–379
  • [BCG2] Nicolas Bergeron, Pierre Charollois and Luis Garcia “Eisenstein cohomology classes for GLN\text{GL}_{N} over imaginary quadratic fields.”, 2021 URL: https://drive.google.com/file/d/18_36dIOfsmoCmd3Q-6GT4OgPrTYCBYlL/view?usp=sharing
  • [BCG3] Nicolas Bergeron, Pierre Charollois and Luis Garcia “Cocycles de Sczech et arrangements d’hyperplans dans les cas additif, multiplicatif et elliptique.” In preparation.
  • [BG] Lev A. Borisov and Paul E. Gunnells “Toric modular forms and nonvanishing of LL-functions.” In Journal für die reine und angewandte Mathematik 539, 2001, pp. 149–165
  • [BK] Shinichi Kobayashi Kenichi Bannai “Algebraic theta functions and pp-adic interpolation of Eisenstein-Kronecker numbers.” In Duke Math Journal. 153.2, 2010, pp. 229–295
  • [BPPS] Cecilia Busuioc, Jeehoon Park, Owen Patashnick and Glenn Stevens “Modular symbols with values in Beilinson-Kato distributions.”, 2023 URL: https://arxiv.org/abs/2311.14620
  • [dR] Georges Rham “Variétés différentiables: formes, courants, formes harmoniques.” Hermann, 1955
  • [KS] Guido Kings and Johannes Sprang “Eisenstein-Kronecker classes, integrality of critical values of Hecke LL-functions and pp-adic interpolation.” Preprint. URL: https://arxiv.org/pdf/1912.03657.pdf
  • [RW] Young Rhie and George Whaples “Hecke operators in the cohomology of groups.” In Journal of the Mathematical Society of Japan. 22, 1970, pp. 431–442
  • [Scz1] Robert Sczech “Eisenstein cocycles for GL2()\mathrm{GL}_{2}(\mathbb{Q}) and values of LL-functions in real quadratic fields.” In Commentarii Mathematici Helvetici. 67, 1992, pp. 363–382
  • [Serre] Jean-Pierre Serre “Faisceaux analytiques sur l’espace projectif.” In Séminaire Henri Cartan. 6.18, 1953, pp. 1–10
  • [Skoda] Henri Skoda “A Dolbeault lemma for temperate currents.” In Annales Henri Lebesgue. 4, 2021, pp. 879–896
  • [SV] Romyar Sharifi and Akshay Venkatesh “Eisenstein cocycles in motivic cohomology.” Preprint., 2020 URL: https://arxiv.org/pdf/2011.07241
  • [Weil] André Weil “Elliptic Functions According to Eisenstein and Kronecker.”, Classics in Mathematics Springer-Verlag, 1999
  • [X] Peter Xu “Arithmetic Eisenstein theta lifts.”, 2023
  • [X3] Peter Xu “Explicit arithmetic Eisenstein cocycles: the elliptic case.” In preparation., 2024
  • [X4] Peter Xu “Classical periods of Eisenstein series and Bernoulli polynomials in the equivariant cohomology of a torus.” Preprint., 2024