Skip to main content
Log in

Poincaré square series for the Weil representation

  • Published:
The Ramanujan Journal Aims and scope Submit manuscript

Abstract

We calculate the Jacobi Eisenstein series of weight \(k \ge 3\) for a certain representation of the Jacobi group, and evaluate these at \(z = 0\) to give coefficient formulas for a family of modular forms \(Q_{k,m,\beta }\) of weight \(k \ge 5/2\) for the (dual) Weil representation on an even lattice. The forms we construct have rational coefficients and contain all cusp forms within their span. We explain how to compute the representation numbers in the coefficient formulas for \(Q_{k,m,\beta }\) and the Eisenstein series of Bruinier and Kuss p-adically to get an efficient algorithm. The main application is in constructing automorphic products.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Borcherds, R.: Automorphic forms with singularities on Grassmannians. Invent. Math. 132(3), 491–562 (1998). https://doi.org/10.1007/s002220050232. ISSN 0020-9910

    Article  MathSciNet  MATH  Google Scholar 

  2. Borcherds, R.: The Gross–Kohnen–Zagier theorem in higher dimensions. Duke Math. J. 97(2), 219–233 (1999). https://doi.org/10.1215/S0012-7094-99-09710-7. ISSN 0012-7094

    Article  MathSciNet  MATH  Google Scholar 

  3. Bruinier, J.: Borcherds products on O(2, \(l\)) and Chern classes of Heegner divisors, volume 1780 of Lecture Notes in Mathematics. Springer, Berlin (2002a). ISBN 3-540-43320-1. https://doi.org/10.1007/b83278

    Book  Google Scholar 

  4. Bruinier, J.: On the rank of Picard groups of modular varieties attached to orthogonal groups. Compos. Math. 133(1), 49–63 (2002b). https://doi.org/10.1023/A:1016357029843. ISSN 0010-437X

    Article  MathSciNet  MATH  Google Scholar 

  5. Bruinier, J., Kuss, M.: Eisenstein series attached to lattices and modular forms on orthogonal groups. Manuscr. Math. 106(4), 443–459 (2001). https://doi.org/10.1007/s229-001-8027-1. ISSN 0025-2611

    Article  MathSciNet  MATH  Google Scholar 

  6. Bykovskiĭ, V.: A trace formula for the scalar product of Hecke series and its applications. Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI), 226 (Anal. Teor. Chisel i Teor. Funktsiĭ. 13): 14–36, 235–236, 1996. ISSN 0373-2703. https://doi.org/10.1007/BF02358528

    Article  MathSciNet  Google Scholar 

  7. Cowan, R., Katz, D., White, L.: A new generating function for calculating the Igusa local zeta function. Adv. Math. 304, 355–420 (2017). https://doi.org/10.1016/j.aim.2016.09.003. ISSN 0001-8708

    Article  MathSciNet  MATH  Google Scholar 

  8. Dittmann, M., Hagemeier, H., Schwagenscheidt, M.: Automorphic products of singular weight for simple lattices. Mathematische Zeitschrift 279(1), 585–603 (2015). https://doi.org/10.1007/s00209-014-1383-6. ISSN 1432-1823

    Article  MathSciNet  MATH  Google Scholar 

  9. Eichler, M., Zagier, D.: The theory of Jacobi forms, volume 55 of Progress in Mathematics. Birkhäuser Boston Inc., Boston (1985). ISBN 0-8176-3180-1. https://doi.org/10.1007/978-1-4684-9162-3

    Book  Google Scholar 

  10. Igusa, J.: Complex powers and asymptotic expansions. I. Functions of certain types. J. Reine Angew. Math., 268/269:110–130 (1974). ISSN 0075-4102. https://doi.org/10.1515/crll.1974.268-269.110. Collection of articles dedicated to Helmut Hasse on his seventy-fifth birthday, II

  11. Raum, M.: Computing genus 1 Jacobi forms. Math. Comp. 85(298), 931–960 (2016). https://doi.org/10.1090/mcom/2992. ISSN 0025-5718

    Article  MathSciNet  MATH  Google Scholar 

  12. Scheithauer, N.: The Weil representation of \({{\rm SL}}_2(\mathbb{Z})\) and some applications. Int. Math. Res. Not. 8, 1488–1545 (2009). ISSN 1073-7928

    Article  Google Scholar 

  13. Shintani, T.: On construction of holomorphic cusp forms of half integral weight. Nagoya Math. J., 58, 83–126 (1975). ISSN 0027-7630. http://projecteuclid.org/euclid.nmj/1118795445

    Article  MathSciNet  Google Scholar 

  14. Strömberg, F.: Weil representations associated with finite quadratic modules. Mathematische Zeitschrift 275(1), 509–527 (2013). https://doi.org/10.1007/s00209-013-1145-x. ISSN 1432-1823

    Article  MathSciNet  MATH  Google Scholar 

  15. Zagier, D.: Modular forms whose Fourier coefficients involve zeta-functions of quadratic fields, Vol. 627, pp. 105–169. Lecture Notes in Mathematics (1977)

    Google Scholar 

  16. Ziegler, C.: Jacobi forms of higher degree. Abh. Math. Sem. Univ. Hamburg 59, 191–224 (1989). https://doi.org/10.1007/BF02942329. ISSN 0025-5858

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

I am grateful to Richard Borcherds, Jan Hendrik Bruinier, Sebastian Opitz, and Martin Raum for helpful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Brandon Williams.

Appendices

Appendix A: Averaging operators

For applications to automorphic products, we do not need the Eisenstein series \(E_{k,\beta }\) for any nonzero \(\beta \in {\varLambda }'/{\varLambda }\) with \(q(\beta ) \in {\mathbb {Z}}\). This is essentially because the constant terms \({\mathfrak {e}}_{\beta }\), \(\beta \ne 0\) are not counted towards the principal part of the input function F in Borcherds’ lift. However, the \(E_{k,\beta }\) are still necessary in order to span the full space of modular forms.

It seems difficult to apply the formula for \(E_{k,\beta }\) in [5] directly since the Kloosterman sums there do not reduce to Ramanujan sums. A brute-force way to find \(E_{k,\beta }\) is to search for \({\varDelta } \cdot E_{k,\beta }\) as a linear combination of Poincaré square series, but this is usually messy. Instead, we mention here that averaging over Schrödinger representations allows one in many (but not all) cases to read off the coefficients of all \(E_{k,\beta }\) from those of \(E_{k,0}\).

Definition 33

Let \(\beta \in {\varLambda }'/{\varLambda }\) have denominator \(d_{\beta }\). The averaging operator attached to \(\beta \) is

$$\begin{aligned} A_{\beta } : M_k(\rho ^*)&\rightarrow M_k(\rho ^*) \\ A_{\beta } F(\tau )&= \frac{1}{d_{\beta ^2}} \sum _{\lambda ,\mu \in {\mathbb {Z}}/d_{\beta }^2 {\mathbb {Z}}} \sigma _{\beta }^*(\lambda ,\mu ,0) F(\tau ). \end{aligned}$$

This is well defined because \(\sigma _{\beta }^*(\lambda ,\mu ,0)\) depends only on the remainder of \(\lambda \) and \(\mu \) mod \(d_{\beta }^2\), and it defines a modular form because

$$\begin{aligned} \Big ( \sigma _{\beta }^*(\zeta ) F \Big ) |_{k,\rho ^*} M (\tau ) = \sigma _{\beta }^*(\zeta \cdot M) F(\tau ) \end{aligned}$$

for all \(\zeta \in {\mathcal {H}}\) and \(M \in \tilde{\varGamma }.\)

Explicitly, if the components of F are written out as

$$\begin{aligned} F(\tau ) = \sum _{\gamma \in {\varLambda }'/{\varLambda }} f_{\gamma }(\tau ) {\mathfrak {e}}_{\gamma }, \end{aligned}$$

then

$$\begin{aligned} A_{\beta } F(\tau ) = \frac{1}{d_{\beta }^2} \sum _{\gamma \in {\varLambda }'/{\varLambda }} \sum _{\lambda \in {\mathbb {Z}}/d_{\beta }^2 {\mathbb {Z}}} \sum _{\mu \in {\mathbb {Z}}/d_{\beta }^2 {\mathbb {Z}}} {\mathbf {e}}\Big ( -\mu \langle \beta , \gamma \rangle + \lambda \mu \cdot q(\beta ) \Big ) f_{\gamma }(\tau ) {\mathfrak {e}}_{\gamma - \lambda \beta }. \end{aligned}$$

The sum over \(\mu \) is nonzero exactly when \(\langle \beta , \gamma \rangle - \lambda q(\beta ) \in {\mathbb {Z}}\), in which case it becomes \(d_{\beta }^2\); therefore,

$$\begin{aligned} A_{\beta } F(\tau ) = \sum _{\lambda \in {\mathbb {Z}}/d_{\beta }^2 {\mathbb {Z}}} \sum _{\begin{array}{c} \gamma \in {\varLambda }'/ {\varLambda } \\ \langle \beta , \gamma \rangle - \lambda q(\beta ) \in {\mathbb {Z}} \end{array}} f_{\gamma }(\tau ) {\mathfrak {e}}_{\gamma - \lambda \beta }. \end{aligned}$$

In the special case that \(q(\beta ) \in {\mathbb {Z}}\), this is a constant multiple of the modified averaging operator

$$\begin{aligned}A'_{\beta } F(\tau ) = \sum _{\begin{array}{c} \gamma \in {\varLambda }'/{\varLambda } \\ \langle \beta , \gamma \rangle \in {\mathbb {Z}} \end{array}} \left( \sum _{\lambda \in {\mathbb {Z}}/d_{\beta } {\mathbb {Z}}} f_{\gamma + \lambda \beta }(\tau ) \right) {\mathfrak {e}}_{\gamma }, \end{aligned}$$

making this easier to compute.

When \(F = E_{k,0}\) is the Eisenstein series

$$\begin{aligned} E_{k,0}(\tau ) = \frac{1}{2} \sum _{c,d} (c \tau + d)^{-k} \rho ^*(M)^{-1} {\mathfrak {e}}_0, \end{aligned}$$

then we get

$$\begin{aligned} A_{\beta } E_{k,0}(\tau )&= \frac{1}{2 d_{\beta }^2} \sum _{c,d} (c \tau + d)^{-k} \sum _{\lambda ,\mu \in {\mathbb {Z}}/d_{\beta }^2 {\mathbb {Z}}} \sigma _{\beta }^* (\lambda ,\mu ,0) \rho ^*(M)^{-1} {\mathfrak {e}}_0 \\&= \frac{1}{2 d_{\beta }^2} \sum _{c,d} (c \tau + d)^{-k} \sum _{\lambda , \mu \in {\mathbb {Z}}/d_{\beta }^2 {\mathbb {Z}}} {\mathbf {e}}\Big ( \lambda \mu q(\beta ) \Big ) \rho ^*(M)^{-1} {\mathfrak {e}}_{-\lambda \beta } \\&= \sum _{\begin{array}{c} \lambda \in {\mathbb {Z}}/d_{\beta }^2 {\mathbb {Z}} \\ \lambda q(\beta ) \in {\mathbb {Z}} \end{array}} E_{k,\lambda \beta }. \end{aligned}$$

In many cases this makes it possible to find all the Eisenstein series \(E_{k,\beta }.\)

Example 34

Let S be the Gram matrix \(S = \begin{pmatrix} -8 \end{pmatrix}.\) The Eisenstein series of weight 5 / 2 is

$$\begin{aligned} E_{5/2,0}(\tau )&= \Big ( 1 - 24q - 72q^2 - 96q^3 - \ldots \Big ) {\mathfrak {e}}_{0} \\&\quad + \Big ( -\frac{1}{2}q^{1/16} - 24q^{17/16} - 72 q^{33/16} - \ldots \Big ) ({\mathfrak {e}}_{1/8} + {\mathfrak {e}}_{7/8}) \\&\quad + \Big ( -5 q^{1/4} - 24q^{5/4} - 125q^{9/4} - \ldots \Big ) ({\mathfrak {e}}_{1/4} + {\mathfrak {e}}_{3/4} ) \\&\quad + \Big ( -\frac{25}{2} q^{9/16} - \frac{121}{2} q^{25/16} - 96 q^{41/16} - \ldots \Big ) ({\mathfrak {e}}_{3/8} + {\mathfrak {e}}_{5/8}) \\&\quad + \Big ( -46 q - 48q^2 - 144q^3 - \ldots \Big ) {\mathfrak {e}}_{1/2}. \end{aligned}$$

Averaging over the Schrödinger representation attached to \(\beta = (1/2)\) gives

$$\begin{aligned} E_{5/2,0}(\tau ) + E_{5/2,1/2}(\tau )&= \Big ( 1 - 70q - 120q^2 - 240q^3 - \ldots \Big ) ({\mathfrak {e}}_0 + {\mathfrak {e}}_{1/2}) \\&\quad + \Big ( -10q^{1/4} - 48q^{5/4} - 250q^{9/4} - \ldots \Big ) ({\mathfrak {e}}_{1/4} + {\mathfrak {e}}_{3/4}), \end{aligned}$$

from which we can read off the Fourier coefficients of \(E_{5/2,1/2}(\tau ).\)

Appendix B: Calculating the Euler factors at \(p = 2\)

We will summarize the calculations of Appendix B in [7] as they apply to our situation.

Proposition 35

Let \(f(X) = \bigoplus _{i \in {\mathbb {N}}_0} 2^i Q_i(X) \oplus L + c\) be a \({\mathbb {Z}}_2\)-integral quadratic polynomial in normal form, and assume that all \(Q_i\) are given by \(Q_i(v) = v^T S_i v\) for a symmetric (not necessarily even) \({\mathbb {Z}}_2\)-integral matrix \(S_i\). For any \(j \in {\mathbb {N}}_0\), define

$$\begin{aligned} \mathbf {Q}_{(j)} := \bigoplus _{\begin{array}{c} 0 \le i \le j \\ i \equiv j \, (2) \end{array}} Q_i, \; \; {\mathbf {r}}_{(j)} = {\mathrm {rank}}(\mathbf {Q}_{(j)}), \; \; {\mathbf {p}}_{(j)} = 2^{\sum _{0 \le i < j} {\mathbf {r}}_{(i)}}. \end{aligned}$$

Let \(\omega \in {\mathbb {N}}_0\) be such that \(Q_i = 0\) for all \(i > \omega .\) Then:

  1. (i)

    If \(L = 0\) and \(c = 0\), let \(r = \sum _i {\mathrm {rank}}(Q_i)\); then the Igusa zeta function for f at 2 is

    $$\begin{aligned} \zeta _{Ig}(f;2;s)&= \sum _{0 \le \nu < \omega - 1} \frac{2^{-\nu s}}{{\mathbf {p}}_{(\nu )}} I_0(\mathbf {Q}_{(\nu )}, \mathbf {Q}_{(\nu +1)}, Q_{\nu +2}) \; \\&\quad \quad \quad + \left[ \frac{2^{-s(\omega - 1)}}{{\mathbf {p}}_{(\omega - 1)}} I_0(\mathbf {Q}_{(\omega -1)}, \mathbf {Q}_{(\omega )}, 0) + \frac{2^{-\omega s}}{{\mathbf {p}}_{(\omega )}} I_0(\mathbf {Q}_{(\omega )}, \mathbf {Q}_{(\omega - 1)}, 0) \right] \cdot \\&\qquad \qquad \times (1 - 2^{-2s - r})^{-1}. \end{aligned}$$
  2. (ii)

    If \(L(x) = bx\) for some \(b \ne 0\) with \(v_2(b) = \lambda \) and if \(v_2(b) \le v_2(c)\), then

    $$\begin{aligned} \zeta _{Ig}(f;2;s)&= \sum _{0 \le \nu< \lambda - 2} \frac{2^{-\nu s}}{{\mathbf {p}}_{(\nu )}} I_0({\mathbf {Q}}_{(\nu )}, {\mathbf {Q}}_{(\nu +1)}, Q_{\nu +2}) \\&\quad \quad \quad + \sum _{\max \{0,\lambda - 2\} \le \nu < \lambda } \frac{2^{-\nu s}}{{\mathbf {p}}_{(\nu )}} I_0^{\lambda - \nu }({\mathbf {Q}}_{(\nu )}, {\mathbf {Q}}_{(\nu +1)}, Q_{i+2})\\&\quad \quad \quad + \frac{2^{-\lambda s}}{{\mathbf {p}}_{(\lambda )}} \cdot \frac{1}{2 - 2^{-s}}. \end{aligned}$$
  3. (iii)

    If \(L(x) = bx\) with \(b \ne 0\) and \(v_2(c) < v_2(b) \le v_2(c) + 2\), let \(\kappa = v_2(c)\); then

    $$\begin{aligned} \zeta _{Ig}(f;2;s)&= \sum _{0 \le \nu < \lambda - 2} \frac{2^{-\nu s}}{{\mathbf {p}}_{(\nu )}} I_{c / 2^{\nu }}({\mathbf {Q}}_{(\nu )}, {\mathbf {Q}}_{(\nu + 1)}, Q_{\nu + 2}) \\&\quad \quad \quad + \sum _{\max \{0,\lambda - 2\} \le \nu \le \kappa } \frac{2^{-\nu s}}{{\mathbf {p}}_{(\nu )}} I^{\lambda - \nu }_{c / 2^{\nu }}( {\mathbf {Q}}_{(\nu )}, {\mathbf {Q}}_{(\nu + 1)}, Q_{\nu + 2})\\&\quad \quad \quad + \frac{1}{{\mathbf {p}}_{(\kappa + 1)}} 2^{-\kappa s}. \end{aligned}$$
  4. (iv)

    If \(L = 0\) or \(L(x) = bx\) with \(v_2(b) > v_2(c) + 2\), let \(\kappa = v_2(c)\); then

    $$\begin{aligned} \zeta _{Ig}(f;2;s)&= \sum _{0 \le \nu \le \kappa } \frac{2^{-\nu s}}{{\mathbf {p}}_{(\nu )}} I_{c / 2^{\nu }}({\mathbf {Q}}_{(\nu )}, {\mathbf {Q}}_{(\nu + 1)}, Q_{\nu + 2}) + \frac{1}{{\mathbf {p}}_{(\kappa + 1)}} 2^{-\kappa s}. \end{aligned}$$

Here, \(I_a^b(Q_0,Q_1,Q_2)(s)\) are helper functions that we describe below, and we set \(I_a(Q_0,Q_1,Q_2) = I^{\infty }_a(Q_0,Q_1,Q_2).\) Note that not every unimodular quadratic form \(Q_i\) over \({\mathbb {Z}}_2\) can be written in the form \(Q_i(v) = v^T S_i v\); but \(2 \cdot Q_i\) can always be written in this form, and replacing f by \(2 \cdot f\) only multiplies \(\zeta _{Ig}(f;2;s)\) by \(2^{-s}\), so this does not lose generality.

Every unimodular quadratic form over \({\mathbb {Z}}_2\) that has the form \(Q_i(v) = v^T S_i v\) is equivalent to a direct sum of at most two one-dimensional forms \(a \cdot \mathrm {Sq}(x) = ax^2\); at most one elliptic plane \(\mathrm {Ell}(x,y) = 2x^2 + 2xy + 2y^2\); and any number of hyperbolic planes \(\mathrm {Hyp}(x,y) = 2xy.\) This decomposition is not necessarily unique. It will be enough to fix one such decomposition.

The following proposition explains how to compute \(I_a^b(Q_0,Q_1,Q_2)(s).\)

Proposition 36

Define the function

$$\begin{aligned} \mathrm {Ig}(a,b,\nu ) = {\left\{ \begin{array}{ll} \frac{2^{-\nu s}}{2 - 2^{-s}}: &{} v_2(a) \ge \mathrm {min}(b,\nu ); \\ 2^{-v_2(a) s}: &{} v_2(a) < \mathrm {min}(b,\nu ). \end{array}\right. } \end{aligned}$$

(Here, \(v_2(0) = \infty \).) For a unimodular quadratic form Q of rank r, fix a decomposition into hyperbolic planes, at most one elliptic plane and at most two square forms as above. Let \(\varepsilon =1\) if Q contains no elliptic plane and \(\varepsilon = -1\) otherwise. Define functions \(H_1(a,b,Q)\), \(H_2(a,b,Q)\), and \(H_3(a,b,Q)\) as follows:

  1. (i)

    If Q contains no square forms, then

    $$\begin{aligned} H_1(a,b,Q)&= (1 - 2^{-r}) \mathrm {Ig}(a,b,1); \\ H_2(a,b,Q)&= \Big ( 1 - 2^{-r/2} \varepsilon \Big ) \cdot \Big ( \mathrm {Ig}(a,b,1) + 2^{-r/2}\varepsilon \mathrm {Ig}(a,b,2) \Big ); \\ H_3(a,b,Q)&= 0. \end{aligned}$$
  2. (ii)

    If Q contains one square form \(cx^2\), then

    $$\begin{aligned} H_1(a,b,Q)&= \mathrm {Ig}(a,b,0) - 2^{-r} \mathrm {Ig}(a,b,1); \\ H_2(a,b,Q)&= (1 - 2^{-(r-1)/2} \varepsilon ) \mathrm {Ig}(a,b,0) - 2^{-r} \mathrm {Ig}(a,b,2) \\&\quad \quad \quad +2^{-(r+1)/2}\varepsilon (\mathrm {Ig}(a,b,2) + \mathrm {Ig}(a+c,b,2)); \\ H_3(a,b,Q)&= 2^{-r} (\mathrm {Ig}(a+c,b,3) - \mathrm {Ig}(a+c,b,2)). \end{aligned}$$
  3. (iii)

    If Q contains two square forms \(cx^2\), \(dx^2\), and \(c + d \equiv 0 \, (4)\), then

    $$\begin{aligned} H_1(a,b,Q)&= \mathrm {Ig}(a,b,0) - 2^{-r} \mathrm {Ig}(a,b,1); \\ H_2(a,b,Q)&= \mathrm {Ig}(a,b,0) - 2^{-r/2}\varepsilon \mathrm {Ig}(a,b,1) + (2^{-r/2}\varepsilon - 2^{-r}) \mathrm {Ig}(a,b,2); \\ H_3(a,b,Q)&= (-1)^{(c+d)/4} 2^{-r} (\mathrm {Ig}(a,b,3) - \mathrm {Ig}(a,b,2)). \end{aligned}$$
  4. (iv)

    If Q contains two square forms \(cx^2\), \(dx^2\), and \(c + d \not \equiv 0 \, (4)\), then

    $$\begin{aligned} H_1(a,b,Q)&= \mathrm {Ig}(a,b,0) - 2^{-r} \mathrm {Ig}(a,b,1); \\ H_2(a,b,Q)&= (1 - 2^{-(r-2)/2} \varepsilon ) \mathrm {Ig}(a,b,0) + 2^{-r/2} \varepsilon (\mathrm {Ig}(a,b,1)\\&\quad \quad \quad + \mathrm {Ig}(a+c,b,2)) - 2^{-r} \mathrm {Ig}(a,b,2); \\ H_3(a,b,Q)&= -2^{1-r} \mathrm {Ig}(a,b,1) + 2^{-r} (\mathrm {Ig}(a,b,2) + \mathrm {Ig}(a+c+d,b,3) ). \end{aligned}$$

Let \(\varepsilon _1 = 1\) if \(Q_1\) contains no elliptic plane and \(\varepsilon _1 = -1\) otherwise, and let \(r_1\) denote the rank of \(Q_1\). Then \(I_a^b(Q_0,Q_1,Q_2)\) is given as follows:

  1. (1)

    If both \(Q_1\) and \(Q_2\) contain at least one square form, then

    $$\begin{aligned} I_a^b(Q_0,Q_1,Q_2) = H_1(a,b,Q_0). \end{aligned}$$
  2. (2)

    If \(Q_1\) contains no square forms but \(Q_2\) contains at least one square form, then

    $$\begin{aligned} I_a^b(Q_0,Q_1,Q_2) = H_2(a,b,Q_0). \end{aligned}$$
  3. (3)

    If both \(Q_1\) and \(Q_2\) contain no square forms, then

    $$\begin{aligned} I_a^b(Q_0,Q_1,Q_2) = H_2(a,b,Q_0) + 2^{-r_1 / 2} \varepsilon _1 H_3(a,b,Q_0). \end{aligned}$$
  4. (4)

    If \(Q_1\) contains one square form \(cx^2\), and \(Q_2\) contains no square forms, then

    $$\begin{aligned} I_a^b(Q_0,Q_1,Q_2)= & {} H_1(a,b,Q_0) \\&\quad + 2^{-(r_1 + 1)/2} \varepsilon _1 (H_3(a,b,Q_0) + H_3(a+2c,b,Q_0)). \end{aligned}$$
  5. (5)

    If \(Q_1\) contains two square forms \(cx^2\) and \(dx^2\) such that \(c+d \equiv 0\, (4)\), and \(Q_2\) contains no square forms, then

    $$\begin{aligned} I_a^b(Q_0,Q_1,Q_2) = H_1(a,b,Q_0) + 2^{-r_1 / 2} \varepsilon _1 H_3(a,b,Q_0). \end{aligned}$$
  6. (6)

    If \(Q_1\) contains two square forms \(cx^2\) and \(dx^2\) such that \(c +d \not \equiv 0 \, (4)\), and \(Q_2\) contains no square forms, then

    $$\begin{aligned} I_a^b(Q_0,Q_1,Q_2) = H_1(a,b,Q_0) + 2^{-r_1/2} \varepsilon _1 H_3(a+c,b,Q_0). \end{aligned}$$

Proof

In the notation of [7],

$$\begin{aligned} \mathrm {Ig}(a,b,\nu ) = \mathrm {Ig}(z^{a + 2^b {\mathbb {Z}}_2 + 2^{\nu } {\mathbb {Z}}_2}) \end{aligned}$$

and

$$\begin{aligned} H_1(a,b,Q) = \mathrm {Ig}\Big (z^{a + 2^b {\mathbb {Z}}_2} \hat{H}_Q(z)\Big ) \end{aligned}$$

and

$$\begin{aligned} H_2(a,b,Q) = \mathrm {Ig}\Big (z^{a + 2^b {\mathbb {Z}}_2} \tilde{H}_Q(z)\Big ) \end{aligned}$$

and

$$\begin{aligned} H_3(a,b,Q) = \mathrm {Ig}\Big (z^{a + 2^b {\mathbb {Z}}_2} (H_Q(z) - \tilde{H}_Q(z))\Big ). \end{aligned}$$

This calculation of \(I_a^b(Q_0,Q_1,Q_2)\) is available in Appendix B of [7]. Finally, the calculation of \(\zeta _{Ig}(f;2;s)\) is given in Theorem 4.5 loc. cit. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Williams, B. Poincaré square series for the Weil representation. Ramanujan J 47, 605–650 (2018). https://doi.org/10.1007/s11139-017-9986-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11139-017-9986-2

Keywords

Mathematics Subject Classification

Navigation