Skip to main content

Maass Waveforms and Low-Lying Zeros

  • Chapter
Analytic Number Theory

Abstract

The Katz–Sarnak Density Conjecture states that the behavior of zeros of a family of L-functions near the central point (as the conductors tend to zero) agrees with the behavior of eigenvalues near 1 of a classical compact group (as the matrix size tends to infinity). Using the Petersson formula, Iwaniec, Luo, and Sarnak proved that the behavior of zeros near the central point of holomorphic cusp forms agrees with the behavior of eigenvalues of orthogonal matrices for suitably restricted test functions ϕ. We prove similar results for families of cuspidal Maass forms, the other natural family of \(\mathrm{GL}_{2}/\mathbb{Q}\) L-functions. For suitable weight functions on the space of Maass forms, the limiting behavior agrees with the expected orthogonal group. We prove this for \(\mathop{\mathrm{supp}}(\hat{\phi }) \subseteq (-3/2,3/2)\) when the level N tends to infinity through the square-free numbers; if the level is fixed the support decreases to being contained in (−1, 1), though we still uniquely specify the symmetry type by computing the 2-level density.

To Professor Helmut Maier on his 60th birthday

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 54.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 9th printing (Dover, New York, 1972)

    Google Scholar 

  2. L. Alpoge, S.J. Miller, Low Lying Zeros of Maass Form L-Functions. Int Math Res Notices (2014), 24 pages, doi:10.1093/imrn/rnu012

    Google Scholar 

  3. B. Birch, H.P.F. Swinnerton-Dyer, Notes on elliptic curves. I. J. Reine Angew. Math. 212, 7–25 (1963)

    MathSciNet  MATH  Google Scholar 

  4. B. Birch, H.P.F. Swinnerton-Dyer, Notes on elliptic curves. II. J. Reine Angew. Math. 218, 79–108 (1965)

    MathSciNet  MATH  Google Scholar 

  5. J.B. Conrey, L-Functions and random matrices, in Mathematics Unlimited: 2001 and Beyond (Springer, Berlin, 2001), pp. 331–352

    Google Scholar 

  6. J.B. Conrey, H. Iwaniec, Spacing of zeros of Hecke L-functions and the class number problem. Acta Arith. 103(3), 259–312 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  7. H. Davenport, Multiplicative Number Theory (Graduate Texts in Mathematics), vol. 74, 2nd edn. (Springer, New York, 1980). Revised by H. Montgomery

    Google Scholar 

  8. E. Dueñez, S.J. Miller, The low lying zeros of a GL(4) and a GL(6) family of L-functions. Compos. Math. 142(6), 1403–1425 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  9. E. Dueñez, S.J. Miller, The effect of convolving families of L-functions on the underlying group symmetries. Proc. Lond. Math. Soc. (2009). doi:10.1112/plms/pdp018

    Google Scholar 

  10. H.M. Edwards, Riemann’s Zeta Function (Academic, New York, 1974)

    MATH  Google Scholar 

  11. A. Entin, E. Roditty-Gershon, Z. Rudnick, Low-lying zeros of quadratic Dirichlet L-functions, hyper-elliptic curves and random matrix theory. Geom. Funct. Anal. 23(4), 1230–1261 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  12. P. Erdős, H. Maier, A. Sárközy, On the distribution of the number of prime factors of sums a + b. Trans. Am. Math. Soc. 302(1), 269–280 (1987)

    Google Scholar 

  13. D. Fiorilli, S.J. Miller, Surpassing the ratios conjecture in the 1-level density of Dirichlet L-functions. Algebra & Number Theory 9(1), 13–52 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  14. F.W.K. Firk, S.J. Miller, Nuclei, primes and the random matrix connection. Symmetry 1, 64–105 (2009). doi:10.3390/sym1010064

    Article  MathSciNet  Google Scholar 

  15. P. Forrester, Log-Gases and Random Matrices. London Mathematical Society Monographs, vol. 34 (Princeton University Press, Princeton, 2010)

    Google Scholar 

  16. E. Fouvry, H. Iwaniec, Low-lying zeros of dihedral L-functions. Duke Math. J. 116(2), 189–217 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  17. P. Gao, N-level density of the low-lying zeros of quadratic Dirichlet L-functions. Ph.D Thesis, University of Michigan, 2005

    Google Scholar 

  18. D. Goldfeld, The class number of quadratic fields and the conjectures of Birch and Swinnerton-Dyer. Ann. Scuola Norm. Sup. Pisa 3(4), 623–663 (1976)

    MathSciNet  MATH  Google Scholar 

  19. B. Gross, D. Zagier, Heegner points and derivatives of L-series. Invent. Math. 84, 225–320 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  20. A. Güloğlu, Low-lying zeros of symmetric power L-functions. Int. Math. Res. Not. 9, 517–550 (2005)

    Article  Google Scholar 

  21. B. Hayes, The spectrum of Riemannium. Am. Sci. 91(4), 296–300 (2003)

    Article  Google Scholar 

  22. D. Hejhal, On the triple correlation of zeros of the zeta function. Int. Math. Res. Not. 7, 294–302 (1994)

    Google Scholar 

  23. C. Hughes, S.J. Miller, Low-lying zeros of L-functions with orthogonal symmtry. Duke Math. J. 136(1), 115–172 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  24. C. Hughes, Z. Rudnick, Linear statistics of low-lying zeros of L-functions. Q. J. Math. Oxf. 54, 309–333 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  25. H. Iwaniec, Introduction to the Spectral Theory of Automorphic Forms (Biblioteca de la Revista Matemática Iberoamericana, Madrid, 1995)

    MATH  Google Scholar 

  26. H. Iwaniec, E. Kowalski, Analytic Number Theory, vol. 53 (AMS Colloquium Publications, Providence, 2004)

    MATH  Google Scholar 

  27. H. Iwaniec, W. Luo, P. Sarnak, Low lying zeros of families of L-functions. Inst. Hautes Études Sci. Publ. Math. 91, 55–131 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  28. N. Katz, P. Sarnak, Random Matrices, Frobenius Eigenvalues and Monodromy, vol. 45 (AMS Colloquium Publications, Providence, 1999)

    MATH  Google Scholar 

  29. N. Katz, P. Sarnak, Zeros of zeta functions and symmetries. Bull. Am. Math. Soc. 36, 1–26 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  30. H. Kim, Functoriality for the exterior square of GL 2 and the symmetric fourth of GL 2. J. Am. Math. Soc. 16(1), 139–183 (2003)

    Article  MATH  Google Scholar 

  31. H. Kim, P. Sarnak, Appendix: Refined estimates towards the Ramanujan and Selberg conjectures, Appendix to “H. Kim, Functoriality for the exterior square of GL 2 and the symmetric fourth of GL 2. J. Am. Math. Soc. 16(1), 139–183 (2003)”

    Google Scholar 

  32. J. Levinson, S.J. Miller, The n-level density of zeros of quadratic Dirichlet L-functions. Acta Arith. 161, 145–182 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  33. C. Li, A. Knightly, Kuznetsov’s trace formula and the Hecke eigenvalues of Maass forms. Mem. Am. Math. Soc. 24(1055), 132 (fourth of 4 numbers) (2013)

    Google Scholar 

  34. J. Liu, Lectures on Maass Forms (Postech, 2007), http://www.prime.sdu.edu.cn/lectures/LiuMaassforms.pdf. Accessed 25–27 March 2007

  35. J. Liu, Y. Ye, Petersson and Kuznetsov trace formulas, in Lie Groups and Automorphic Forms, ed. by L. Ji, J.-S. Li, H.W. Xu, S.-T. Yau, AMS/IP Studies in Advanced Mathematics, vol. 37 (American Mathematical Society, Providence, 2006), pp. 147–168

    Google Scholar 

  36. H. Maier, On Exponential Sums with Multiplicative Coefficients, in Analytic Number Theory (Cambridge University Press, Cambridge 2009), pp. 315–323

    Google Scholar 

  37. H. Maier, C. Pomerance, Unusually large gaps between consecutive primes. Trans. Am. Math. Soc. 322(1), 201–237 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  38. H. Maier, A. Sankaranarayanan, On an exponential sum involving the Mbius function. Hardy-Ramanujan J. 28, 10–29 (2005)

    MathSciNet  MATH  Google Scholar 

  39. H. Maier, A. Sankaranarayanan, Exponential sums over primes in residue classes. Int. J. Number Theory 6(4), 905–918 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  40. H. Maier, G. Tenenbaum, On the set of divisors of an integer. Invent. Math. 76(1), 121–128 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  41. M. Mehta, Random Matrices, 2nd edn. (Academic, Boston, 1991)

    MATH  Google Scholar 

  42. S.J. Miller, 1- and 2-level densities for families of elliptic curves: evidence for the underlying group symmetries. Ph.D. Thesis, Princeton University, 2002

    Google Scholar 

  43. S.J. Miller, 1- and 2-level densities for families of elliptic curves: evidence for the underlying group symmetries. Compos. Math. 140, 952–992 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  44. S.J. Miller, D. Montague, An orthogonal test of the L-functions ratios conjecture, II. Acta Arith. 146, 53–90 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  45. S.J. Miller, R. Peckner, Low-lying zeros of number field L-functions. J. Number Theory 132, 2866–2891 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  46. S.J. Miller, R. Takloo-Bighash, An Invitation to Modern Number Theory (Princeton University Press, Princeton, 2006)

    MATH  Google Scholar 

  47. H. Montgomery, The pair correlation of zeros of the zeta function, in Analytic Number Theory. Proceedings of Symposia in Pure Mathematics, vol. 24 (American Mathematical Society, Providence, 1973), pp. 181–193

    Google Scholar 

  48. A. Odlyzko, On the distribution of spacings between zeros of the zeta function. Math. Comput. 48(177), 273–308 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  49. A. Odlyzko, The 1022-nd zero of the Riemann zeta function, in Proceedings of the Conference on Dynamical, Spectral and Arithmetic Zeta-Functions. American Mathematical Society, Contemporary Mathematics series, 2001, ed. by M. van Frankenhuysen, M.L. Lapidus. http://www.research.att.com/$\sim$amo/doc/zeta.html

    Google Scholar 

  50. A.E. Özlük, C. Snyder, Small zeros of quadratic L-functions. Bull. Aust. Math. Soc. 47(2), 307–319 (1993)

    Article  MATH  Google Scholar 

  51. G. Ricotta, E. Royer, Statistics for low-lying zeros of symmetric power L-functions in the level aspect. Forum Math. 23, 969–1028 (2011, preprint)

    Google Scholar 

  52. G.F.B. Riemann, Über die Anzahl der Primzahlen unter einer gegebenen Grösse, Monatsber. Königl. Preuss. Akad. Wiss. Berlin, Nov. 671–680 (1859) (see “H.M. Edwards, Riemann’s Zeta Function (Academic, New York, 1974)” for an English translation)

    Google Scholar 

  53. E. Royer, Petits zéros de fonctions L de formes modulaires. Acta Arith. 99(2), 147–172 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  54. M. Rubinstein, Low-lying zeros of L-functions and random matrix theory. Duke Math. J. 109, 147–181 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  55. M. Rubinstein, P. Sarnak, Chebyshev’s bias. Exp. Math. 3(3), 173–197 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  56. Z. Rudnick, P. Sarnak, Zeros of principal L-functions and random matrix theory. Duke Math. J. 81, 269–322 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  57. S.W. Shin, N. Templier, Sato-Tate theorem for families and low-lying zeros of automorphic L-functions. Invent. Math. preprint. http://arxiv.org/pdf/1208.1945v2

  58. R.A. Smith, The L 2-norm of Maass wave functions. Proc. Am. Math. Soc.  82(2), 179–182 (1981)

    MATH  Google Scholar 

  59. A. Yang, Low-Lying Zeros of Dedekind Zeta Functions Attached to Cubic Number Fields, preprint

    Google Scholar 

  60. M. Young, Low-lying zeros of families of elliptic curves. J. Am. Math. Soc. 19(1), 205–250 (2006)

    Article  MATH  Google Scholar 

Download references

Acknowledgements

We thank Eduardo Dueñez, Gergely Harcos, Andrew Knightly, Stephen D. Miller, and Peter Sarnak for helpful conversations in an earlier version. This work was done at the SMALL REU at Williams College, funded by NSF GRANT DMS0850577 and Williams College; it is a pleasure to thank them for their support. The second named author was also partially supported by the Mathematics Department of University College London, and the fifth named author was partially supported by NSF grants DMS0970067 and DMS1265673.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Steven J. Miller .

Editor information

Editors and Affiliations

Appendices

Appendix 1: Contour Integration

We prove Proposition 3.3 below. We restate it for the reader’s convenience.

Proposition 5.1.

Let T be an odd integer and X ≤ T. Let w T equal h T or H T, where these are the weight functions from Theorems  2.2 to 2.5. Then

$$ \displaystyle\begin{array}{rcl} \int _{\mathbb{R}}J_{2ir}(X)\frac{rw_{T}(r)} {\cosh (\pi r)} dr& =& c_{1}\sum _{k\geq 0}(-1)^{k}J_{ 2k+1}(X)(2k + 1)w_{T}\left (\left (k + \frac{1} {2}\right )i\right ) \\ & & \quad \left [+\ c_{2}T^{2}\sum _{ k\geq 1}(-1)^{k}J_{ 2kT}(X)k^{2}h(k)\right ] {}\end{array}$$
(86)
$$\displaystyle\begin{array}{rcl} & =& \ c_{1}\sum _{k\geq 0}(-1)^{k}J_{ 2k+1}(X)(2k + 1)w_{T}\left (\left (k + \frac{1} {2}\right )i\right ) \\ & & \quad \left [+O\left (Xe^{-c_{3}T}\right )\right ], {}\end{array}$$
(87)

where c 1 , c 2 , and c 3 are constants independent of X and T, and the terms in brackets are included if and only if w T = h T.

Proof.

In the proof below bracketed terms are present if and only if w T  = h T .

Recall that

$$\displaystyle\begin{array}{rcl} J_{\alpha }(2x)\ =\ \sum _{m\geq 0} \frac{(-1)^{m}x^{2m+\alpha }} {m!\varGamma (m +\alpha +1)}.& &{}\end{array}$$
(88)

By Stirling’s formula, \(\varGamma (m + 2ir + 1)\cosh (\pi r) \gg \vert m + 2ir + 1\vert ^{m+\frac{1} {2} }e^{-m}\). Hence by the Lebesgue Dominated Convergence Theorem (remember that w T (z) is of rapid decay as \(\vert \mathfrak{R}z\vert \rightarrow \infty \)) we may switch sum and integral to get

$$\displaystyle\begin{array}{rcl} & & \int _{\mathbb{R}}J_{2ir}(X)\frac{rw_{T}(r)} {\cosh (\pi r)} dr \\ & & \quad \quad \ =\ \sum _{m\geq 0}\frac{(-1)^{m}x^{2m}} {m!} \int _{\mathbb{R}} \frac{x^{2ir}rw_{T}(r)} {\varGamma (m + 1 + 2ir)\cosh (\pi r)}dr,{}\end{array}$$
(89)

where X = : 2x.

Now we move the line of integration down to \(\mathbb{R} - iR\), \(R\not\in \mathbb{Z} + \frac{1} {2}\) (and \(\not\in T\mathbb{Z}\) if \(w_{T} = h_{T}\)). To do this, we note the estimate (for A ≫ 1)

$$\displaystyle\begin{array}{rcl} \int _{\pm A\rightarrow \pm A-iR} \frac{x^{2ir}rw_{T}(r)} {\varGamma (m + 1 + 2ir)\cosh (\pi r)}dr& \ll & \int _{\pm A\rightarrow \pm A-iR}e^{m}x^{2R}\vert r\vert \vert m + 2ir \\ & & +1\vert ^{-m-\frac{1} {2} }\vert w_{T}(r)\vert dr \\ & \ll & _{T,R}A^{-2}, {}\end{array}$$
(90)

where again we have used the rapid decay of w T along horizontal lines, and B → BiR denotes the vertical line from \(B \in \mathbb{C}\) to \(B - iR \in \mathbb{C}\). (Rapid decay also ensures the integral along \(\mathbb{R} - iR\) converges absolutely.)

Note that the integrand

$$\displaystyle\begin{array}{rcl} \frac{x^{2ir}rw_{T}(r)} {\varGamma (m + 1 + 2ir)\cosh (\pi r)}& &{}\end{array}$$
(91)

has poles below the real axis precisely at \(r \in -i(\mathbb{N} + \frac{1} {2}) =\{ -\frac{1} {2}i,-\frac{3} {2}i,\ldots \}\), and, if w T  = h T , poles also at \(r \in -iT\mathbb{Z}^{+}\). The residue of the pole at \(r = -\frac{2k+1} {2} i\) (k ≥ 0) is, up to an overall constant independent of k,

$$\displaystyle\begin{array}{rcl} \frac{x^{2k+1}} {\varGamma (m + 1 + (2k + 1))}(-1)^{k}(2k + 1)w_{ T}\left (\frac{2k + 1} {2} i\right ).& &{}\end{array}$$
(92)

If w T  = h T , the residue of the pole at r = −ikT (k ≥ 1) is, up to another overall constant independent of k,

$$\displaystyle\begin{array}{rcl} \frac{x^{2kT}} {\varGamma (m + 1 + (2kT))}(-1)^{k}k^{2}T^{2} \frac{h(k)} {\cos (\pi kT)} = \frac{x^{2kT}} {\varGamma (m + 1 + (2kT))}k^{2}T^{2}h(k).& &{}\end{array}$$
(93)

Hence the sum of (89) becomes

$$\displaystyle\begin{array}{rcl} & & \sum _{m\geq 0}\frac{(-1)^{m}x^{2m}} {m!} \int _{\mathbb{R}} \frac{x^{2ir}rw_{T}(r)} {\varGamma (m + 1 + 2ir)\cosh (\pi r)}dr \\ & & \quad \quad =\ c_{6}\sum _{m\geq 0}\frac{(-1)^{m}x^{2m}} {m!} \Bigg(\sum _{0\leq k\ll R} \frac{x^{2k+1}} {\varGamma (m + 1 + (2k + 1))}(-1)^{k}(2k + 1)w_{ T} \\ & & \quad \quad \quad \quad \quad \left.\left (\frac{2k + 1} {2} i\right ) +\int _{\mathbb{R}-iR} \frac{x^{2ir}rw_{T}(r)} {\varGamma (m + 1 + 2ir)\cosh (\pi r)}dr\right. \\ & & \quad \quad \quad \quad \quad \Bigg[+c_{7}\sum _{m\geq 0}\frac{(-1)^{m}x^{2m}} {m!} \sum _{0\leq k\ll R} \frac{x^{2kT}} {\varGamma (m + 1 + (2kT))}(-1)^{k}k^{2}T^{2}h(k)\Bigg]\Bigg). {}\end{array}$$
(94)

Now we take R → . Note that

$$\displaystyle\begin{array}{rcl} \int _{\mathbb{R}-iR} \frac{x^{2ir}rw_{T}(r)} {\varGamma (m + 1 + 2ir)\cosh (\pi r)}dr& \,\ll & \,\int _{\mathbb{R}-iR}x^{2R}\vert r\vert \vert w_{ T}(r)\vert \vert m + 2ir + 1\vert ^{-m-\frac{1} {2} -2R}dr \\ & & \ll _{m}\ x^{2R}e^{ \frac{\pi R} {2T} }R^{-m-\frac{1} {2} -2R}, {}\end{array}$$
(95)

again by Stirling and rapid decay of w T on horizontal lines (that is, \(w_{T}(x + iy) \ll (1 + x)^{-4}e^{ \frac{\pi y} {2T} }\) since both h and H have all their derivatives supported in \(\left (-\frac{1} {4}, \frac{1} {4}\right )\)). This of course vanishes as R → .

Hence we see that

$$\displaystyle\begin{array}{rcl} & & \int _{\mathbb{R}}J_{2ir}(X)\frac{rw_{T}(r)} {\cosh (\pi r)} dr \\ & & \ \ \ \ \ \ =\ c_{8}\sum _{m\geq 0}\frac{(-1)^{m}x^{2m}} {m!} \sum _{k\geq 0} \frac{x^{2k+1}} {\varGamma (m + 1 + (2k + 1))}(-1)^{k}(2k + 1)w_{ T}\left (\frac{2k + 1} {2} i\right ) \\ & & \ \ \ \ \ \ \ \ \ \ \ \left [+c_{9}\sum _{m\geq 0}\frac{(-1)^{m}x^{2m}} {m!} \sum _{k\geq 0} \frac{x^{2kT}} {\varGamma (m + 1 + (2kT))}k^{2}T^{2}h(k)\right ]. {}\end{array}$$
(96)

Switching sums (via the exponential bounds on w T along the imaginary axis) and applying \(J_{n}(X) =\sum _{m\geq 0}\frac{(-1)^{m}x^{2m+n}} {m!(m+n)!}\) gives us the claimed calculation. For the bound on the bracketed term, use

$$\displaystyle\begin{array}{rcl} J_{n}(ny)\ \ll \ \frac{y^{n}e^{n\sqrt{1-y^{2}} }} {(1 + \sqrt{1 - y^{2}})^{n}}& &{}\end{array}$$
(97)

(see [1], p. 362) and bound trivially.

Appendix 2: An Exponential Sum Identity

The following proposition and proof are also used in [2].

Proposition 6.1.

Suppose X ≤ T. Then

$$\displaystyle\begin{array}{rcl} S_{J}(X)\ &:=& \ T\sum _{k\geq 0}(-1)^{k}J_{ 2k+1}(X)\frac{\tilde{\tilde{h}}\left (\frac{2k+1} {2T} \right )} {\sin \left (\frac{2k+1} {2T} \pi \right )} \\ & =& \ c_{10}T\sum _{\vert \alpha \vert <\frac{T} {2} }e\left (Y \sin \left ( \frac{\pi \alpha } {T}\right )\right )\tilde{\tilde{g}}\left (\frac{\pi Y } {T}\cos \left ( \frac{\pi \alpha } {T}\right )\right ) + O(Y ),{}\end{array}$$
(98)

where c 10 is some constant, g(x):= sgn (x)h(x), X =: 2πY, and for any f set \(\tilde{f}(x):= xf(x)\) .

Proof.

Observe that \(k\mapsto \sin \left (\pi k/2\right )\) is supported only on the odd integers, and maps 2k + 1 to (−1)k. Hence, rewriting gives

$$\displaystyle\begin{array}{rcl} S_{J}(X)\ =\ T\sum _{{ k\geq 0 \atop k\not\in 2T \mathbb{Z}} }J_{k}(X)\tilde{\tilde{h}}\left ( \frac{k} {2T}\right ) \frac{\sin \left (\frac{\pi k} {2} \right )} {\sin \left ( \frac{\pi k} {2T}\right )}.& &{}\end{array}$$
(99)

Since

$$\displaystyle\begin{array}{rcl} \frac{\sin \left (\frac{\pi k} {2} \right )} {\sin \left ( \frac{\pi k} {2T}\right )}\ =\ \frac{e^{\frac{\pi ik} {2} } - e^{-\frac{\pi ik} {2} }} {e^{ \frac{\pi ik} {2T} } - e^{-\frac{\pi ik} {2T} }}\ =\ \sum _{ \alpha =-\left (\frac{T-1} {2} \right )}^{\frac{T-1} {2} }e^{\frac{\pi ik\alpha } {T} }& &{}\end{array}$$
(100)

when k is not a multiple of 2T, we find that

$$\displaystyle\begin{array}{rcl} S_{J}(X)\ =\ T\sum _{\vert \alpha \vert <\frac{T} {2} }\sum _{{ k\geq 0 \atop k\not\in 2T \mathbb{Z}} }e\left ( \frac{k\alpha } {2T}\right )J_{k}(X)\tilde{\tilde{h}}\left ( \frac{k} {2T}\right ).& &{}\end{array}$$
(101)

Observe that, since the sum over α is invariant under α ↦ −α (and it is non-zero only for k odd!), we may extend the sum over k to the entirety of \(\mathbb{Z}\) at the cost of a factor of 2 and of replacing h by

$$\displaystyle{ g(x)\:=\ \mathrm{ sgn}(x)h(x). }$$
(102)

(This is because of the identity J k (x) = (−1)k J k (x) and the fact that we need only consider odd k.) Note that g is as differentiable as h has zeros at 0, less one. That is to say, \(\hat{g}\) decays like the reciprocal of a degree \(\mathop{\mathrm{ord}}_{z=0}h(z) - 1\) polynomial at . This will be crucial in what follows.

Next, we add back on the \(2T\mathbb{Z}\) terms and obtain

$$\displaystyle\begin{array}{rcl} \frac{1} {2}S_{J}(X)& =& T\sum _{\vert \alpha \vert <\frac{T} {2} }\sum _{k\in \mathbb{Z}}e\left ( \frac{k\alpha } {2T}\right )J_{k}(X)\tilde{\tilde{g}}\left ( \frac{k} {2T}\right ) - T^{2}\sum _{ k\in \mathbb{Z}}J_{2kT}(X)k^{2}h(k) \\ & =& T\sum _{\vert \alpha \vert <\frac{T} {2} }\sum _{k\in \mathbb{Z}}e\left ( \frac{k\alpha } {2T}\right )J_{k}(X)\tilde{\tilde{g}}\left ( \frac{k} {2T}\right ) + O\left (Xe^{-c_{11}T}\right ) \\ & =:& V _{J}(X) + O\left (Xe^{-c_{11}T}\right ), {}\end{array}$$
(103)

where we have bounded the term \(T^{2}\sum _{k\in \mathbb{Z}}J_{2kT}(X)k^{2}h(k)\) trivially via \(J_{n}(2x) \ll x^{n}/n!\).

Now we move to apply Poisson summation. Write X = : 2π Y. We apply the integral formula (for \(k \in \mathbb{Z}\))

$$\displaystyle{ J_{k}(2\pi x)\ =\ \int _{ -\frac{1} {2} }^{ \frac{1} {2} }e\left (kt - x\sin (2\pi t)\right )dt }$$
(104)

and interchange the sum and integral (via rapid decay of g) to get that

$$\displaystyle\begin{array}{rcl} V _{J}(X)\ =\ T\sum _{\vert \alpha \vert <\frac{T} {2} }\int _{-\frac{1} {2} }^{ \frac{1} {2} }\left (\sum _{k\in \mathbb{Z}}e\left ( \frac{k\alpha } {2T} + kt\right )\tilde{\tilde{g}}\left ( \frac{k} {2T}\right )\right )e\left (-Y \sin (2\pi t)\right )dt.& &{}\end{array}$$
(105)

By Poisson summation, (105) is just (interchanging the sum and integral once more)

$$\displaystyle\begin{array}{rcl} V _{J}(X)& =& T^{2}\sum _{ \vert \alpha \vert <\frac{T} {2} }\sum _{k\in \mathbb{Z}}\int _{-\frac{1} {2} }^{ \frac{1} {2} }\hat{g}''\left (2T(t - k)+\alpha \right )e\left (-Y \sin (2\pi t)\right )dt \\ & =& c_{12}T\sum _{\vert \alpha \vert <\frac{T} {2} }\int _{-\infty }^{\infty }\hat{g}''(t)e\left (Y \sin \left ( \frac{\pi t} {T} + \frac{\pi \alpha } {T}\right )\right )dt \\ & =:& c_{12}W_{g}(X). {}\end{array}$$
(106)

As

$$\displaystyle{ \sin \left ( \frac{\pi t} {T} + \frac{\pi \alpha } {T}\right )\ =\ \sin \left ( \frac{\pi \alpha } {T}\right ) + \frac{\pi t} {T}\cos \left ( \frac{\pi \alpha } {T}\right ) - \frac{\pi ^{2}t^{2}} {2T^{2}}\sin \left ( \frac{\pi \alpha } {T}\right ) + O\left ( \frac{t^{3}} {T^{3}}\right ), }$$
(107)

we see that

$$\displaystyle\begin{array}{rcl} & & W_{g}(X) \\ & =& c_{13}T\sum _{\vert \alpha \vert <\frac{T} {2} }e\left (Y \sin \left ( \frac{\pi \alpha } {T}\right )\right )\int _{-\infty }^{\infty }\hat{g}''(t)e\left (\frac{\pi Y t} {T} \cos \left ( \frac{\pi \alpha } {T}\right )\right )dt \\ & & \ \ -\ c_{14}\frac{\pi ^{2}Y } {T} \sum _{\vert \alpha \vert <\frac{T} {2} }e\left (Y \sin \left ( \frac{\pi \alpha } {T}\right )\right )\sin \left ( \frac{\pi \alpha } {T}\right )\int _{-\infty }^{\infty }t^{2}\hat{g}''(t)e\left (\frac{\pi Y t} {T} \cos \left ( \frac{\pi \alpha } {T}\right )\right )dt \\ & & \ \ +\ O\left (\frac{Y } {T} + \frac{Y ^{2}} {T^{2}} \right ) {}\end{array}$$
(108)
$$\displaystyle\begin{array}{rcl} & =& c_{15}T\sum _{\vert \alpha \vert <\frac{T} {2} }e\left (Y \sin \left ( \frac{\pi \alpha } {T}\right )\right )\tilde{\tilde{g}}\left (\frac{\pi Y } {T}\cos \left ( \frac{\pi \alpha } {T}\right )\right ) \\ & & \ \ -\ c_{16}\frac{Y } {T}\sum _{\vert \alpha \vert <\frac{T} {2} }e\left (Y \sin \left ( \frac{\pi \alpha } {T}\right )\right )\sin \left ( \frac{\pi \alpha } {T}\right )\tilde{\tilde{g}}''\left (\frac{\pi Y } {T}\cos \left ( \frac{\pi \alpha } {T}\right )\right ) \\ & & \ \ +\ O\left (\frac{Y } {T} + \frac{Y ^{2}} {T^{2}} \right ). {}\end{array}$$
(109)

Now bound the second term trivially to get the claim.

Appendix 3: 2-Level Calculations

The purpose of this appendix is to provide additional details to the 2-level computation of Sect. 4.4. Letting

$$\displaystyle\begin{array}{rcl} S_{1}(u_{j},\phi _{i})&:=& \sum _{p}\frac{2\lambda _{p}(u_{j})\log p} {p^{1/2}\log R} \hat{\phi }_{i}\left ( \frac{\log p} {\log R}\right ) \\ S_{2}(u_{j},\phi _{i})&:=& \sum _{p}\frac{2\lambda _{p^{2}}(u_{j})\log p} {p\log R} \hat{\phi }_{i}\left (\frac{2\log p} {\log R} \right ){}\end{array}$$
(110)

and summing over the family, a standard calculation reduces the determination of the 2-level density to understanding

$$\displaystyle\begin{array}{rcl} \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \prod _{i=1}^{2}\Bigg[\widehat{\phi _{ i}}(0)\frac{\log (1 + t_{j}^{2})} {\log R} & -& S_{1}^{c(i)}(u_{ j},\phi _{i}) - S_{2}^{c(i)}(u_{ j},\phi _{i}) \\ & +& O\left (\frac{\log \log R} {\log R}\right )\Bigg]^{2} {}\end{array}$$
(111)

(the other terms are straightforward consequences of the combinatorial book-keeping from the inclusion–exclusion argument), where c(1) is the identity map and c(2) denotes complex conjugation.

We can move the factor O(loglogR∕logR) outside the product at the cost of an error of the same size outside all the summations. To see this, since O(loglogR∕logR) is independent of u j these terms are readily bounded by applying the Cauchy-Schwarz inequality. Letting \(\mathcal{S}\) represent any of the factors in the product over i in (111), the product involving this is O(loglogR∕logR):

$$\displaystyle\begin{array}{rcl} & & \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \mathcal{S}\cdot O\left (\frac{\log \log R} {\log R}\right ) \\ & & \ll \left [ \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} O\left (\left (\frac{\log \log R} {\log R}\right )^{2}\right )\right ]^{1/2} \cdot \left [ \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \vert \mathcal{S}\vert ^{2}\right ]^{1/2} \\ & & \ll O\left (\frac{\log \log R} {\log R}\right ) \cdot \left ( \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \vert \mathcal{S}\vert ^{2}\right )^{1/2}. {}\end{array}$$
(112)

We now analyze the four possibilities for the sum involving \(\vert \mathcal{S}\vert ^{2}\). If \(\mathcal{S}\) is either \(\widehat{\phi _{i}}(0)\frac{\log (1+t_{j}^{2})} {\log R}\) or \(O\left (\frac{\log \log R} {\log R}\right )\), this sum is trivially O(1), and thus the entire expression is \(O\left (\frac{\log \log R} {\log R}\right )\). We are left with the non-trivial cases of \(\mathcal{S} = S_{1}\) or \(\mathcal{S} = S_{2}\). To handle these cases, we rely on results that we will soon prove in lemmas below: for sufficiently small support | S 1 | 2 = O(1) and | S 2 | 2 = o(1). Note that we use these lemmas for ϕ = ϕ 1 ϕ 1 instead of the usual ϕ = ϕ 1 ϕ 2, but this does not affect the proofs, and thus we may move the \(O\left (\frac{\log \log R} {\log R}\right )\) factor outside the product.

By symmetry, it suffices to analyze the following terms to determine the 2-level density:

$$\displaystyle\begin{array}{rcl} & & \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \widehat{\phi _{1}}(0)\widehat{\phi _{2}}(0)\left (\frac{\log (1 + t_{j}^{2})} {\log R} \right )^{2} \\ & & \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \widehat{\phi _{1}}(0)\frac{\log (1 + t_{j}^{2})} {\log R} S_{k}(u_{j},\phi _{2}),\ \ k \in \{ 1,2\} \\ & & \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} S_{k}(u_{j},\phi _{1})\overline{S_{\ell}(u_{j},\phi _{2})}\ \ k,\ell\in \{ 1,2\}. {}\end{array}$$
(113)

We now analyze these terms. For small support, a similar analysis as performed earlier in the paper shows that the first term in (113) satisfies

$$\displaystyle{ \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \widehat{\phi _{1}}(0)\widehat{\phi _{2}}(0)\left (\frac{\log (1 + t_{j}^{2})} {\log R} \right )^{2} =\widehat{\phi _{ 1}}(0)\widehat{\phi _{2}}(0) + O\left ( \frac{1} {\log \log R}\right ). }$$
(114)

We next handle the terms where we have exactly one S-factor.

Lemma 7.1.

For sufficiently small support, the \(\widehat{\phi _{1}}(0)S_{1}(u_{j},\phi _{2})\) and \(\widehat{\phi _{1}}(0)S_{2}(u_{j},\phi _{2})\) terms are O(log log R∕log R), and thus do not contribute.

Proof.

The proof is almost identical to the application of the Kuznetsov trace formula to prove similar results for the 1-level density, the only change being that now we have the modified weight function h T (t j )log(1 + t j 2).

We now turn to the more interesting terms in (113). We first handle the second term.

Lemma 7.2.

For sufficiently small support,

$$\displaystyle{ \frac{1} {\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} } \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} S_{1}(u_{j},\phi _{1})\overline{S_{1}(u_{j},\phi _{2})} = 2\int _{-\infty }^{\infty }\vert z\vert \hat{\phi }_{ 1}(z)\hat{\phi }_{2}(z)dz + O\left (\frac{\log \log R} {\log R}\right ). }$$
(115)

Proof.

$$\displaystyle\begin{array}{rcl} & & \sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} S_{1}(u_{j},\phi _{1})\overline{S_{1}(u_{j},\phi _{2})} \\ & & =\ 4\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \sum _{p_{1},p_{2}} \frac{\lambda _{p_{1}}(u_{j})\overline{\lambda _{p_{2}}(u_{j})}} {p_{1}^{1/2}p_{2}^{1/2}} \frac{\log p_{1}\log p_{2}} {\log ^{2}R} \hat{\phi }_{1}\left (\frac{\log p_{1}} {\log R} \right )\hat{\phi }_{2}\left (\frac{\log p_{2}} {\log R} \right ) \\ & & =\ 4\sum _{p_{1},p_{2}} \frac{1} {p_{1}^{1/2}p_{2}^{1/2}} \frac{\log p_{1}\log p_{2}} {\log ^{2}R} \hat{\phi }_{1}\left (\frac{\log p_{1}} {\log R} \right )\hat{\phi }_{2}\left (\frac{\log p_{2}} {\log R} \right )\sum _{j}\frac{h_{T}(t_{j})} {\|u_{j}\|^{2}} \lambda _{p_{1}}(u_{j})\overline{\lambda _{p_{2}}(u_{j})}.{}\end{array}$$
(116)

As before, we apply the Kuznetsov formula to the inner sum. Since the formula has a \(\delta _{p_{1},p_{2}}\), we need to split this sum into the case when p 1 = p 2 and the case when p 1p 2. For small support one easily finds the case p 1p 2 does not contribute; however, the case p 1 = p 2 does contribute. Arguing as in the 1-level calculations, after applying the Kuznetsov formula we are left with

$$\displaystyle\begin{array}{rcl} 4\sum _{p}\frac{1} {p} \frac{\log ^{2}p} {\log ^{2}R}\hat{\phi }_{1}\left ( \frac{\log p} {\log R}\right )\hat{\phi }_{2}\left ( \frac{\log p} {\log R}\right ) = 2\int _{-\infty }^{\infty }\vert z\vert \hat{\phi }_{ 1}(z)\hat{\phi }_{2}(z)dz + O\!\left (\frac{\log \log R} {\log R}\right )\!\!.& &{}\end{array}$$
(117)

(the equality follows from partial summation and the Prime Number Theorem, see [42] for a proof).

Lemma 7.3.

For small support, the contribution from the \(S_{k}(u_{j},\phi _{1})\overline{S_{\ell}(u_{j},\phi _{2})}\) terms is O(log log R∕log R) if \((k,\ell) = (1,2)\) or (2,2).

Proof.

The proof is similar to the previous lemma, following again by applications of the Kuznetsov trace formula. The support is slightly larger as the power of the primes in the denominator is larger.

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Alpoge, L., Amersi, N., Iyer, G., Lazarev, O., Miller, S.J., Zhang, L. (2015). Maass Waveforms and Low-Lying Zeros. In: Pomerance, C., Rassias, M. (eds) Analytic Number Theory. Springer, Cham. https://doi.org/10.1007/978-3-319-22240-0_2

Download citation

Publish with us

Policies and ethics