Skip to main content

Section 2 Duke's theorem

Recall

1.2 1.9.

We will now prove this, not via Linnik's proof, but via an analogue of Weyl's criterion. Strategy:

  1. Weyl's criterion \(\leadsto\) bounding exponential sums.
  2. To bound these sums we will use automorphic methods.

We will begin with 1.2.

So we will be working on \(S^2\text{.}\) Recall 1.28 which was for \(\lb 0,1)\text{.}\) There we used exponentials, why? Because they are a dense, convenient basis, fourier theory.

To replace this on the sphere \(S^2\) we use the spherical harmonics, these are homogeneous harmonic polynomials. Analogous with the \(S^1\) case.

\begin{equation*} \left( \frac{x+iy}{|x+iy|}\right)^m = e(m\theta) \end{equation*}

the same spherical harmonic construction works for \(S^n\text{.}\)

We will show

\begin{equation*} \frac{1}{\# \Omega_n} \sum_{x\in \Omega_n} P(x) \to 0 \end{equation*}

for \(\deg P(x) \gt 0\text{.}\) i.e.

\begin{equation} \sum_{\alpha \in \ZZ^3,|\alpha|^2 = n} P\left(\frac\alpha{|\alpha|}\right) = o( r_3(n))\label{eqn-spherical-harmonic-bd}\tag{2.1} \end{equation}

where

\begin{equation*} r_3(n) = \# \{ a^2 + b^2 + c^2 = n: a,b,c\in \ZZ\}\text{.} \end{equation*}
Connection to automorphic forms

\(\theta\) functions, \(P\) spherical harmonics as before.

Definition 2.1
\begin{equation*} \theta_P(z) = \sum_{\alpha \in \ZZ^3} P(\alpha) e ( |\alpha|^2 z) \end{equation*}

for \(z\in \HH\) this converges.

\begin{equation*} \theta_P(z) = \sum_{n=0}^\infty r_3(n,P) e(nz) \end{equation*}
\begin{equation*} r_3(n,P) =\sum_{|\alpha|^2 = n} P (\alpha)\text{.} \end{equation*}
Observation 2.3

If \(\deg P \not \equiv 0 \pmod 2\) this function is \(0\text{.}\)

To show (2.1):

  1. Show that
    \begin{equation*} r_3(n) \gg_\epsilon n^{1/2 - \epsilon} \end{equation*}
    (Gauss-Siegel)
  2. Show that
    \begin{equation*} r_3(n,P) \ll_\delta n^{k/2 - 1/4 - \delta} \end{equation*}
    for some \(\delta \gt 0\text{.}\)

Why?

\begin{equation*} \sum_{|\alpha|^2 = n} P\left( \frac{\alpha}{|\alpha|} \right) = n^{-\deg(P)/2} \sum_{|\alpha|^2 = n} P(\alpha) \end{equation*}
\begin{equation*} = n^{-\deg(P)/2} r_3(n,P) \end{equation*}

note:

\begin{equation*} \frac k2 - \frac14 - \delta = \frac 12 + \frac{\deg(P)}{2} - \delta\text{.} \end{equation*}

So if \(r_3(n,P) \ll n^{(1+\deg(P))/2 - \delta}\) which implies

\begin{equation*} \sum_{|\alpha|^2 = n} P\left( \frac{\alpha}{|\alpha|} \right) \ll n^{\frac 12 - \delta}\text{.} \end{equation*}

If we knew a half-integral weight Ramanujan conjecture:

\begin{equation*} f \in S_k(N) ,\, k \in \frac 12 + \ZZ_{\gt 0} \end{equation*}

for squarefree \(n\)

\begin{equation*} a_f(n) = O(n^{(k-1)/2 +\epsilon})\text{.} \end{equation*}
Digression (\(\theta\)-functions)

Recall: Integral weight modular forms

\begin{equation*} \Gamma = \SL_2(\ZZ) = \Gamma(1) \end{equation*}
\begin{equation*} \Gamma_\infty = \left\{ \begin{pmatrix} 1\amp n \\ 0 \amp 1\end{pmatrix} : n \in \ZZ\right\} \end{equation*}

In general \(\Gamma\) congruence subgroup \(\Gamma(N) \subseteq \Gamma \subseteq \Gamma(1)\) for some \(N\text{.}\)

\begin{equation*} f\colon \HH \to \CC \end{equation*}

is called a modular form of weight \(k\) for \(\Gamma\) if \(f\) is holomorphic everywhere, including at the cusps.

\begin{equation*} f(\gamma z) = (cz+ d)^k f(z) \,\forall \gamma \in \Gamma\text{.} \end{equation*}

If \(f (z) = \sum _{n=1}^\infty a_nq^n\) then \(f\) is cuspidal, the space of such is \(S_k(\Gamma)\) where S stands for the German Spitzenform. Spitze means cusp, kinda like a pointy spit?

Example 2.4
\begin{equation*} \Delta(q) = q\prod_{n=1}^\infty (1-q^n)^{24} \end{equation*}
\begin{equation*} E_k(z)= \sum_{(c,d) = 1} \frac{1}{(cz +d)^{2k}} \end{equation*}

Classically

\begin{equation*} \widetilde \theta(z) = \sum_{m\in \ZZ} e^{i \pi m^2z} \end{equation*}

converges absolutely on \(z\in \HH\)

\begin{equation*} \widetilde \theta(z + 2) =\widetilde \theta(z) \end{equation*}
\begin{equation*} \widetilde \theta(-1/z ) = \sqrt{-iz} \widetilde \theta(z) \end{equation*}

in general for \(\gamma \in \Gamma_0(4)\)

\begin{equation*} \widetilde \theta(\gamma z ) = j(\gamma ; z) \widetilde \theta(z) \end{equation*}

where \(j(\gamma; z) = \legendre cd \epsilon_d^{-1} (cz+d)^{1/2}\) where \(\epsilon_d = 1\) if \(d\equiv 1 \pmod 4\text{,}\) \(-1\) if \(d\equiv 3 \pmod 4\) sign of Gauss sum. \(\legendre cd\) is a sibling of Legendre form.

Recall: \(P\) spherical harmonic degree \(l \text{.}\) Aim to show:

\begin{equation*} \sum_{|x|^2 = n} P\left( x \right) = o( r_3(n)) \end{equation*}

where

\begin{equation*} r_3(n) = \# \{ a^2 + b^2 + c^2 = n: a,b,c\in \ZZ\}\text{.} \end{equation*}

Took

\begin{equation*} \theta(z; P) = \sum_{n\in \ZZ} r(n; P) e(nz)\text{.} \end{equation*}

A modular form of weight \(3/2 + l\) for \(\Gamma_0 (4)\text{.}\) Cusp form if \(l \gt 0\text{.}\)

The strategy is then to show

\begin{equation*} r_3(n) \gg n^{1/2- \epsilon}\,\forall \epsilon \gt 0 \end{equation*}
\begin{equation*} r_3(n) \ll n^{k/2 -1/4 - \delta}\,\text{ for some}\,\delta \gt 0\text{.} \end{equation*}
Definition 2.6 Half-integral weight modular forms

Let \(N \equiv 0 \pmod 4\text{.}\) A modular form of half-integral weight \(k \in \frac 12 + \ZZ_{\ge 0}\text{.}\) For \(\Gamma_0(N)\) is a holomorphic function on \(\HH\) s.t.

  1. \begin{equation*} f(\gamma z ) = j(\gamma; z)^{2k} f(z) \end{equation*}
  2. \(f\) is holomorphic at the cusps

if \(\chi\begin{pmatrix}a\amp b \\c \amp d \end{pmatrix} = \chi (d)\text{,}\) \(f(\gamma z) = \chi(d) j(\gamma ; z)^{2k} f(z)\) gives the space \(M_k(\Gamma_0(N), \chi)\text{.}\)

Where do these things come from?

A construction due to Schoenberg 1939, Pfetzer 1953, Shimura 1973 is as follows:

\(A\) an \(n \times n\) positive definite integral matrix, \(N \in \ZZ\) s.t. \(N A\inv\) is integral, \(P\) a spherical harmonic relative to \(A\) i.e. for \(P\) homogeneous of degree \(v\text{.}\)

\begin{equation*} \sum_{i,j} \tilde a_{ij} \frac{\partial ^2 P}{\partial x_ix_j} = 0 \end{equation*}
\begin{equation*} [a_{ij}] = A \inv \end{equation*}
Definition 2.7

Let \(h\in \ZZ^n\text{,}\) set

\begin{equation*} \tilde \theta_P (z,h,N) = \sum_{m \equiv h \pmod N} P(m) e\left( \frac{(m^\transpose A m)z}{2N^2} \right)\text{.} \end{equation*}
Example 2.9

\(n = 1, N = 1\) \(P(m) = m^v\) \(v= 0,1\)

\begin{equation*} \tilde \theta_P(z) = \sum_{m\in \ZZ} m^v e(m^2z/2) \end{equation*}

for \(v=0\) this is classical \(\theta\) for \(v = 1\) cusp form on \(\Gamma_0(8)\) of weight \(3/2\text{,}\) we could also twist by a character mod 4.

Example 2.10

\(A = I_{n \times n}\) \(P\) spherical harmonic of degree \(v\)

\begin{equation*} \tilde \theta_P(z; 0, 1) = \sum _{m \in \ZZ^n} P(m) e(|m|^2 z/2) \end{equation*}

set \(z=2z\)

\begin{equation*} \theta_P(z) = \sum_{m\in \ZZ} P(m) e(|m|^2 z)\text{.} \end{equation*}
Note 2.11
\begin{equation*} \theta_P (z) \in S_k(\Gamma_0(4)) \end{equation*}
\begin{equation*} k = \frac h2 + v,\,v \gt 0 \end{equation*}
\begin{equation*} \theta_P (z)\in M_{h/2}(\Gamma_0(4)) \end{equation*}

not cuspidal.

Example 2.12

\(A = 4\times 4\) integral positive definite

\begin{equation*} Q = Q_A(x),\, x^\transpose A x \end{equation*}
\begin{equation*} r_Q(n) = \# \{ x \in \ZZ^4:Q(x) = n\} \end{equation*}
\begin{equation*} \theta_Q(z) = \sum_{n=0}^\infty r_Q(n) e(nz) \in M_2(\Gamma_0(N)) \end{equation*}

for some \(N = 4\det(A)\text{.}\)

Half-integral weight Ramanujan conjecture (Metaplectic)

Naively we would like to mimic the integral case and say

\begin{equation*} a_f(n) = O(n^{(k-1)/2 + \epsilon}) \end{equation*}

where \(f\in S_k(\Gamma)\text{.}\)

But this is not true as stated here!:

Example 2.13

Let

\begin{equation*} \theta(z; \chi) = \sum_{m\in \ZZ} m \chi(m) e(m^2 z) \end{equation*}

for odd \(\chi\) this is a cusp form in \(S_k(\Gamma, \chi)\text{.}\) For \(k=3/2\text{,}\) let \(f =\theta\) and

\begin{equation*} a_f(m^2) \sim m\chi(m) \implies a_f(n) = O(\sqrt{n}) \end{equation*}

for \(n\) square. But \((k -1)/2= 1/4\) but \(O(\sqrt n)\) is not \(O(n^{1/4 + \epsilon})\) for any \(\epsilon \ge 1/4\text{.}\)

So we avoid these or stick to squarefree \(n\text{.}\)

Why this exponent? For integral weight: Representation theory gives \(a_f(m)\)'s which correspond to Hecke eigenvalues which under Langlands are tempered for \(\GL_N\text{.}\)

Digression:

\(y = \Im(z), y(\gamma z) = |cz+d| ^{-2} y\) so

\begin{equation*} F(z) = y^{k/2} |f(z)| \end{equation*}

Invariant on \(\Gamma\) bounded at the cusps \(|F(z)| \lt M\text{.}\) So

\begin{equation*} \int_0^1 f(z) e(nx) = e^{-2 \pi i ny }a_f(n) \end{equation*}
\begin{equation*} \le \int_0^1 y^{-k/2} M = O(y^{-k/2}) \end{equation*}

so \(a_f(n) = O(y^{-k/2} e^{2\pi i n y})\) with \(y = 1/n\text{.}\)

Another digression:

What is the distribution of rational points on \(S^2\text{?}\)

\begin{equation*} \{x_1^2 + x_2^2 + x_3^2 = 1: x_i \in \QQ\} \end{equation*}
\begin{equation*} \iff r_3(n^2) \end{equation*}

Prove this!

David Fried's proof of Hurwitz on \(r_3(n^2)\)

Without proof Hurwitz stated

\begin{equation*} r_3(N^2) = 6 P \prod_{q|Q} ( q^a + 2q^{a-1} + \cdots + 2q + q) \end{equation*}

where \(N = 2^k PQ\) and each prime factor of \(P\) is \(\equiv 1 \pmod 4\) and each prime factor of \(Q\) is \(\equiv -1 \pmod 4\) and

\begin{equation*} q^a || Q \end{equation*}

he suggested a proof along the lines of his published note on

\begin{equation*} r_5(N^2)\text{,} \end{equation*}

here is such a proof:

We denote \(P = P(N), Q = Q(N)\) We may suppose \(k = 0\) since

\begin{equation*} r_3(4n) = r_3(n) \end{equation*}

(each solution of \(4n = x^2 + y^2 + z^2\) has \(x,y,z\) all even). So \(N\) is odd and each solution of \(N^2 = x^2 + y^2 + z^2\) has two of \(x,y,z\) even. Hence

\begin{equation*} r_3(N^2) = \frac 32 \sum_{x \text{ even}} r_2( N^2- x^2) = \frac 22 \sum_{a+b = 2N,a,b\text{ odd}} r_2(ab) \end{equation*}

now \(r_2(n)\) is the number of Gaussian integers \(z\) with \(z \bar z = n\text{.}\) As \(\ZZ \lb i \rb\) is a PID with 4 units \(\pm 1, \pm i\text{,}\) the function \(\rho(n) = \frac 14 r_2(n)\) is multiplicative!

\begin{equation*} \rho (ab) = \rho(a)\rho(b) \text{ if } a,b \text{ coprime} \end{equation*}

clearly

\begin{equation*} r_3(N^2) = 6 \sum_{a+b = 2N,\,a,b\,\text{odd}}\rho (ab)\text{.} \end{equation*}

To evaluate \(\rho(ab)\) we need spme standard functions,

\begin{equation*} \tau(n)= \# \text{divisors }d\text{ of }n,\,\sigma(n) = \sum d \end{equation*}

the Möbius function \(\mu(n)\text{.}\) and

\begin{equation*} \square (n) = \frac 12 r_1(n) \end{equation*}
\begin{equation*} w(n) = \mu(P)|\mu(Q)| \end{equation*}

The multiplicativity of \(\rho\) generalises as follows:

Each \(p = a^2 + b^2\) with \(a\gt b \gt 0\) uniquely (Euler). Let \(z(p) = a+bi\text{.}\) For \(n\) odd \(n = PQ\) and \(r_2(n) = 0\) unless \(Q = \square\text{.}\) In which case the solutions of \(n = z \bar z\) are

\begin{equation*} z = i^k \sqrt Q \prod_p z(p)^j \overline{z(p)}^{\alpha- j} \end{equation*}

where \(p^\alpha || P\text{,}\) \(k \in \ZZ/4\text{.}\) Thus

\begin{equation*} \rho(n) = \tau(p) \square (Q)\text{.} \end{equation*}

When \(\rho(ab) = 0\) every term on the right vanishes. For \(Q(ab) \ne \square\) so

\begin{equation*} Q\left( \frac ad \right),\,Q\left( \frac bd \right) \end{equation*}

cannot both be squares.

When \(\rho(ab) \ne 0\) we have \(Q(ab) = m^2\) for some\(m\text{.}\) There is exactly one squarefree \(\gamma\) dividing \(m\) and \(g\) such that

\begin{equation*} Q\left(\frac a\gamma \right),\, Q\left(\frac b\gamma \right) \end{equation*}

are squares. The term

\begin{equation*} w(d) \rho(a/d) \rho(b/d) \end{equation*}

is nonzero just when \(\alpha = \gamma\delta\) where \(\delta\) is squarefree and \(\delta|P(g)\) so the sum reduces to

\begin{equation*} \sum_\delta \mu(\sigma) \tau(P(a/\gamma\delta)) \tau(P(b/\gamma\delta)) \end{equation*}
\begin{equation*} = \sum_\delta \mu(\delta) \tau(P(a)/\delta) \tau(P(b)/\delta) \end{equation*}

but

\begin{equation*} \sum_{d|r,d|s} \mu(d) \tau(r/d) \tau(r/s) = \tau(rs) \end{equation*}

by multiplicativity of \(\tau\) this reduces to the case where \(r = \pi^k,s= \pi^l\) for prime \(\pi\) but \(1+ k + l = (1+k)(1+l) - kl\text{.}\) Here now \(r=P(a), s=P(b),\gcd(r,s) = P(g)\text{.}\) So the sum over \(\delta\) equals

\begin{equation*} \tau(P(a)P(b)) = \tau(P(ab)) \square (Q(ab)) = \rho(ab) \end{equation*}

as desired.

Using this we find

\begin{equation*} r_3(n^2) = 6 \sum_{d|N} w(d) S \left(\frac Nd\right) \end{equation*}

where for odd \(n\)

\begin{equation*} S(n) = \sum_{a+b = 2n,\,a,b\,\text{odd}} \rho(a) \rho(b) \end{equation*}

but

\begin{equation*} S(n) = \frac{1}{16}\#(2n = \underbrace{w^2 + x^2}_{\text{odd}}+ \underbrace{y^2 + z^2}_{\text{odd}}) \end{equation*}
\begin{equation*} = \frac{1}{4}\#(2n = 4w^2 + 4x^2+y^2 + z^2) \end{equation*}
\begin{equation*} = \sigma(n) \end{equation*}

by Jacobi.

So

\begin{equation*} r_3(N^2)= 6\sum_{d|N} w(d) \sigma\left(\frac Nd\right) \end{equation*}
\begin{equation*} = 6(w \ast \sigma) (N) \end{equation*}

where \(\ast\) denotes Dirichlet convolution.

Pasing to Dirichlet series

\begin{equation*} \sum_{N\text{ odd}} \frac{ r_3 (N^2)} {N^s} = 6 \sum _{d\text{ odd}} \frac{w(d)}{d^s } \sum _{m\text{ odd}} \frac{\sigma(m)}{m^s } \end{equation*}
\begin{equation*} = 6\underbrace{ \prod_p(1- p^{-s})\prod_q (1+ q^{-s})}_{L(s, \legendre{-4}{\cdot})\inv}\underbrace{\sum_{k \text{ odd}} \frac{1}{k^s}}_{\zeta_{\text{odd}}(s-1)}\underbrace{\sum_{l \text{ odd}} \frac{l}{l^s}}_{\zeta_{\text{odd}}(s-1)} \end{equation*}
\begin{equation*} = 6\prod_p \sum_a \frac{p^a}{p^{as}} \prod_q \sum_a \frac{q^a + 2q^{a -1} + \cdots + 2q + 2} {q^{as}} \end{equation*}

which gives

\begin{equation*} r_3(N^2) = 6 \prod_{p^a||N} p^a \prod_{q^a||N} (q^a + 2q^{a-1} + \cdots 2q + 2) \end{equation*}

as in Hurwitz and

\begin{equation*} \sum_{n=1}\infty \frac{r_3(n^2)}{n^s} = 6 L(s, \legendre{-4}{\cdot}) \inv \zeta(s) \zeta_{\text{odd}}(s-1) \end{equation*}

as in Duke.

Expand the formula, take \(n=p\) an odd prime

\begin{equation*} \zeta(s) \zeta(1-s) = \sum \frac{\sigma(n)}{n^s} \end{equation*}

(exercise)

\begin{equation*} L(s, \chi_{-4})\inv = \prod_p(1- \frac{\chi_{-4}(p)}{p^s}) \end{equation*}

so

\begin{equation*} \sum \frac{r_3(n^2)}{n^s} = \sum \frac{\sigma(m_1)}{m_1^s} \prod (1 - \chi_{-4}(m_p)/p^s) 6 (1-2^{1-s}) \end{equation*}
\begin{equation*} = \sum \frac{\sigma(m_1)}{m_1^s} \sum \mu(m_) \chi_{-4}(m_2)/m_2^s 6 (1-2^{1-s}) \end{equation*}

so

\begin{equation*} r_3(p^2) = (\sigma(p) - \chi_{-4}(p))6 \ge 6p \end{equation*}

The next big breakthrough came much later, showing something similar for \(r_3(n, P)\text{,}\) but now we hope to see cancellation.

Recall:

\begin{equation*} \theta_P(z) = \sum r_3(n; P) e(nz) \in S_{3/2 + l} (\Gamma_0(4)) \end{equation*}

is a weight \(3/2 +l\) form if \(l= \deg P\text{.}\)

Shimura gives us

\begin{equation*} F_P(z) \in S_{2l+2}(\Gamma_0(2)) \end{equation*}

then we get

The Hecke bound implies \(A_P(d) \ll d^{l +1}\text{.}\)

Detour: bounding Fourier coefficients

There are various approaches, such as Poincaré series, Kloosterman sums, and bound these, to get a bound on the Fourier coefficient.

If \(f\in S_{k_1} (N_1)\) we will write a spanning set for \(S_{k_1}(N_1)\text{,}\) then bound fourier coefficients on each of these guys, which will suffice.

Poincaré series:

If we were just looking for a modular form we might try Eisenstein series

\begin{equation*} E_k(z) = \sum_{c,d} \frac{1}{(cz+d)^{2k}} \end{equation*}

these are not cuspidal. But if we twist this a little bit

\begin{equation*} P_m(z; k) = \sum_{\gamma \in \Gamma_\infty \backslash \Gamma} j(\gamma; z)^{-2k} e(m \gamma z)\text{.} \end{equation*}

Where \(\Gamma_\infty = \left\{ \begin{pmatrix} 1\amp n \\ 0 \amp 1 \end{pmatrix}\right\}\text{.}\)

\begin{equation*} j(\gamma; z) = \left( \frac cd \right) \epsilon_d \inv (cz+d)^{1/2} \end{equation*}

they converge absolutely for \(k \gt 2\text{.}\)

Some properties:

Exercise, \(j\) satisifes a cocycle relation

\begin{equation*} j(\gamma ; \tau z) j(\tau; z) = j(\gamma \tau; z)\text{.} \end{equation*}

To follow.

\begin{equation*} \pair{P_m(z; k) }{f}_{\mathrm{Pet}} = \int_{\Gamma \backslash \HH} P_m(z; k) \overline f(z) y^k \frac{\diff x \diff y }{y^2} \end{equation*}
\begin{equation*} = \int_{\Gamma\backslash \HH} \sum_{\gamma \in \pm \Gamma_\infty \backslash \Gamma} j(\gamma; z)^{-2k} e(m \gamma z) \overline f(z) y^k \frac{\diff x \diff y }{y^2} \end{equation*}
\begin{equation*} = \sum_{\gamma \in \pm \Gamma_\infty \backslash \Gamma}\int_{\Gamma\backslash \HH} j(\gamma; z)^{-2k} e(m \gamma z) \overline f(z) y^k \frac{\diff x \diff y }{y^2} \end{equation*}
\begin{equation*} = \int_{\Gamma_\infty\backslash \HH} e(m z) \overline f(z) y^k \frac{\diff x \diff y }{y^2} \end{equation*}
\begin{equation*} = \int_0^1 \int_0^\infty e(m z) \overline f(z) y^k \frac{\diff x \diff y }{y^2} \end{equation*}
\begin{equation*} = \overline a_f(m) \int_0^\infty e(2 m iy ) y^k \frac{\diff x \diff y }{y^2} \end{equation*}
\begin{equation*} = \frac{\overline a_f(m) }{(4 \pi m)^{k-1}}\Gamma(k-1) \end{equation*}

If \(f \perp P_m(z, k)\) for all \(m\) then \(f= 0\text{.}\)

Observation 2.26

The corollary implies that bounding fourier coefficients of \(f\in S_k(\Gamma)\) is equivalent to bounding fourier coefficients of \(P_m(z,k)\text{.}\)

\begin{align*} \int_0^1 P_m(z,k) e(-nz) \diff z \amp= \int_0^1 \left( \sum_{\pm \Gamma_\infty \backslash \Gamma} j(\gamma ; z) ^{-2k} e(m\gamma z)\right) e(-nz) \diff z\\ \amp= \int_0^1 \underbrace{j(1; z)^{-2k} e((m-n)z)}_{\delta_{m,n}} \diff z + \text{rest}\\ \amp= \int_0^1 j(1; z)^{-2k} e((m-n)) \diff z + \text{rest} \end{align*}

Note:

\begin{equation*} e(\begin{pmatrix} 1\amp n \\ 0 \amp 1 \end{pmatrix} z) = e(z)\text{.} \end{equation*}

So

\begin{equation*} \int_0^1\sum_{\Gamma_\infty\backslash \Gamma} j(\gamma; z)^{-2k} e(m \gamma z) \diff z \end{equation*}
\begin{equation*} = \int_0^1sum_{\Gamma_\infty\backslash \Gamma /\Gamma_\infty} \sum_{\gamma_\infty \in \Gamma_\infty} j(\gamma \gamma_\infty z)^{-2k} e(m\gamma \gamma_\infty z) e(-nz) \diff z \end{equation*}
\begin{equation*} \int_0^1 \sum_{\Gamma_\infty\backslash \Gamma /\Gamma_\infty} \sum_{\alpha \in \ZZ} j(\gamma ; z+ \alpha )^{-2k} e(m\gamma (z + \alpha)) e(-nz) \diff z \end{equation*}
\begin{equation*} \int_0^1 \sum_{\Gamma_\infty\backslash \Gamma /\Gamma_\infty} j(\gamma ; z)^{-2k} e(m\gamma z-nz) \diff z \end{equation*}

Note:

\begin{equation*} m\gamma z = m \frac{az+ b}{cz+d} = m \left( \frac ac - \frac{1}{c(cz+d)}\right) \end{equation*}
\begin{equation*} \pm \Gamma_\infty \backslash \Gamma / \Gamma_\infty =\{ (a,d,c) : c \gt 0 ,\,a,d \in (\ZZ/c)^\times,\,ad \equiv 1 \pmod c\} \end{equation*}

So our main integral is

\begin{equation*} \sum_{c\gt 0 ,\,d\in \ZZ/c^\times} \epsilon_d^{-2k} \left(\frac cd \right)^{2k} \int_{-\infty}^\infty \frac{1}{(cz+d)^k} e\left(m\frac ac - \frac{m}{c(cz+d)} -nz\right) \diff z \end{equation*}
\begin{equation*} \sum_{c\gt 0 ,\,d\in \ZZ/c^\times,\,ad\equiv 1 \pmod c} \frac{e((ma+nd)/c)}{c^k} \epsilon_d^{-2k} \left(\frac cd \right)^{2k} \int_{-\infty}^\infty \frac{1}{z^k} e\left( - \left(zn + \frac{m}{c^2z} \right)\right) \diff z \end{equation*}

Then use the following integral representation

\begin{equation*} \int_{-\infty + iA}^{\infty + iA} w^{-k} e^{-(\mu_1 w + \mu_2 m\inv)} \diff w \end{equation*}
\begin{equation*} = 2\pi \left(\frac {\mu_1}{\mu_2}\right)^{(k-1)/2} e^{i k \pi / 2} J_{k-1} (4 \pi \sqrt{\mu_1\mu_2}) \end{equation*}

To recap:

\begin{equation*} P_m(z) = \sum_{n=1}^\infty \hat P_m(n) e(nz) \end{equation*}
\begin{equation*} \hat P_m(n) = \left( \frac nm \right)^{(k-1)/2} \left\{ \delta_{m,n} + \frac{2\pi}{i^k} \sum_{c\equiv 0 \pmod N,\,c \gt 0} J_{k-1} \left( \frac {4\pi \sqrt{mn}}{c} \right) \frac{K(m,n; c)}{c} \right\} \end{equation*}

Digression: Let \(\Gamma = \SL_2(\ZZ)\text{,}\) \(k= 12\) then \(S_k(\Gamma) = \langle \Delta \rangle\)

\begin{equation*} \Delta(z) = (2\pi)^{12} \sum_{n=1}^\infty \tau(n) e(nz)\text{.} \end{equation*}

As \(\Delta = \eta^{24}\) we have \(\tau(n) \in \ZZ\text{.}\)

As

\begin{equation*} P_{m}(z; 12) \in S_{12} (1) = \langle \Delta\rangle \end{equation*}
\begin{equation*} \implies P_{m}(z; 12) = \kappa(m) \Delta \end{equation*}

to calculate \(\kappa(m)\)

\begin{equation*} \pair{P_m(z)}{\Delta} = \frac{\overline a_m}{(4\pi m )^{k-1}} \Gamma(k-1) = \frac{2 \pi \overline{\tau (m)}}{(4 \pi m)^{k-1}} \Gamma(k-1) = \frac{2\pi 10!}{(4\pi m )^{11}} \end{equation*}

so

\begin{equation*} P_m(z) = \frac{2\pi 10!}{(4\pi m )^{11}} \frac{\Delta(z)}{\| \Delta\|^2_{\mathrm{Pet}}} \tau(m)\text{.} \end{equation*}

Another consequence:

Let \(P_m(z) = \sum_f \pair{P_m(z)}{f}f\) then

\begin{equation*} \pair{P_m(z)}{P_n(z)} = \sum_f \overline {a_f(m)} a_f(n) \frac{\Gamma(k-1)^2}{(4 \pi m)^{k-1} (4\pi n )^{k-1}} \end{equation*}

on the other hand

\begin{equation*} \pair{P_m(z)}{P_n(z)} = \frac{\Gamma(k-1)}{(4\pi m)^{k-1}} \left(\frac mn\right)^{(k-1)/2} \left\{ \cdots \right\}\text{.} \end{equation*}

Recall Shimura gave us

\begin{equation*} \sum_{n=1}^\infty \frac{r_3(n^2, P)}{n^s} = \frac{\sum_{n=1}^\infty \frac{A_P(n)}{n^s}}{L(s-l, \chi_{-4})} \end{equation*}

where the numerators is the \(L\)-function of a modular form of weight \(2l + 2\) where \(l = \deg P\text{.}\)

We want to show \(A_P(n) \ll n^{1+l-\delta}\) for some \(\delta \gt 0\text{.}\)

Idea: Poincaré series span cusp forms, so its enough to bound Fourier coefficients of these.

We know

\begin{equation*} \hat P_m(n) = \left(\frac nm \right)^{(k-1)/2} \left( \delta_{m,n} + 2 \pi i^{-k} \sum_{c \equiv 0 \pmod N,\, c \gt 0}\overbrace{\frac{K(m,n,c)}{c}}^{\sum_{x \pmod{^* c}} e((x m + \overline x n)/c)} J_{k-1} \left(\frac {4 \pi \sqrt{mn}}{c}\right) \right)\text{.} \end{equation*}

Try: Input 1:

\begin{equation*} J_{k-1}(x) \ll \min\{x^{k-1},1/\sqrt{x}\}\text{.} \end{equation*}

Trivial input:

\begin{equation*} | K(m, n; c) | \lt c\text{.} \end{equation*}

For

\begin{equation*} c \gg \sqrt{mn} \leadsto J_{k-1}\left( \frac{4 \pi \sqrt{mn}}{c} \right) \sim \left( \frac{\sqrt{mn}}{c} \right)^{k-1} \end{equation*}
\begin{equation*} \lt \left(\frac nm \right)^{(k-1)/2} \sum_{c \gg \sqrt{nm}} \frac{1}{c^{k-1}} (mn)^{(k-1)/2} \end{equation*}
\begin{equation*} \lt n^{k-1} \frac{1}{\sqrt{mn}^{k-2}} = O(n^{k-1 - (k-2)/2}) \end{equation*}

for

\begin{equation*} c \ll \sqrt{mn} \leadsto J_{k-1}\left( \frac{4 \pi \sqrt{mn}}{c} \right) \sim \frac{\sqrt c}{(mn)^{1/4}} \end{equation*}
\begin{equation*} \lt \left( \frac nm \right)^{(k-1)/2} \sum_{c\ll \sqrt{nm}} \frac{\sqrt{c}}{(mn)^{1/4}} = O (n^{k/2})\text{.} \end{equation*}

Note \(k/2 = l+1\text{:}\) So we are just short, we have \(O(n^{l+1})\) rather than \(O(n^{l+1-\delta})\) for all \(\delta \gt 0\text{,}\) just by using a stupid bound for \(K(m,n; c)\text{.}\)

More major input (Weil bound):

\begin{equation*} |K(m,n \chi ) | \ll_\epsilon c^{1/2 +\epsilon} \end{equation*}
\begin{equation*} \implies | \hat P_m(n) | = O(n^{k/2 - 1/4 + \epsilon}) \end{equation*}

so

\begin{equation*} \forall \epsilon \lt \frac 14 \end{equation*}

this does the job.

So we have finished proving Linnik for \(n^2\text{,}\) it wasn't too bad, the major inputs were the Shimura correspondence.

When we move to the general case things will be totally different.

Remark 2.29

Weil bound comes from the Riemann hypothesis for curves over finite fields.

Reparameterise \(xy =1\text{,}\) \(x= a+b,y= a-b\text{.}\) Exercise.

\begin{equation*} \sum_{x \pmod p} \legendre{x^2 - n \overline m}{p} e\left( \frac {2m x}{p}\right) = \sum_{x\pmod p} \left(\legendre{x^2 - n \overline m}{p} + 1 \right) e\left( \frac {2m x}{p}\right)\text{.} \end{equation*}
Note 2.32

LHS is calculating a Fourier coefficient. i.e.

\begin{equation*} h_{m,n} (\alpha) = \legendre{\alpha^2 - \overline m n }{p} \end{equation*}
\begin{equation*} \text{LHS} = \hat h(2m) \end{equation*}

the above two lemmas imply

\begin{equation*} K(m,n ; p) \end{equation*}

is the Fourier coefficient \(\hat h_{m,n}(m)\text{.}\)

So

\begin{equation*} h(\QQ(\sqrt{u^2 - 4\overline m n})) \sim L\left( 1, \legendre{u^2 - 4 \overline m n}{\cdot} \right) \end{equation*}
\begin{equation*} \text{Selberg trace formula} \leftrightarrow_{\text{Fourier transform}} \text{Petersson trace formula}\text{.} \end{equation*}
Remark 2.33

In the non-square case we want \(|r_3(n,P)| \ll n^{k/2 - 1/4 - \delta}\text{.}\)

This is because of the subtlety of the Shimura correspondence.

More precisely for \(n\) squarefree \(n^{k/2 - 1/4 - \delta}\)

\begin{equation*} \theta(z; P) \text{ weight } \frac 32 + \deg P = k \end{equation*}
\begin{equation*} r_3(n ; P) \gg n^{1/2 - \delta} \end{equation*}

want

\begin{equation*} r_3(n; P ) \ll n^{k/2 - 1/4 - \delta} \end{equation*}
\begin{equation*} \sum_{|\alpha|^2 = n} P\left( \frac{\alpha}{|\alpha|}\right) = |n|^{-\deg P / 2} r_(n , P) \end{equation*}
\begin{equation*} r_3(n, P ) \lt n^{k/2 - 1/4 - \delta} = n^{3/4 + \deg P/2 - 1/4 - \delta} \end{equation*}

for \(n^2\)

\begin{equation*} r_3(n^2) \gg n \end{equation*}

have Shimura

\begin{equation*} \theta(z; P) \leadsto \text{weight }k \end{equation*}
\begin{equation*} A_p(n^2) \end{equation*}

are fourier coefficients of a weight \(2 \deg P + 2\) form.

Back to \(r_3(n,P)\)

\(n\)-squarefree, we don't have Shimura, but we still have the Fourier expansion of \(P_m(z)\text{.}\)

Strategy: We will first exploit a certain form of these Kloosterman sums for \(\Delta\) squarefree, Salié sums. Major detail

\begin{equation*} \sum_c \frac{K(m, n ; c)}{c} \end{equation*}

will bound these not individually, but by showing that the angles of these cancel. We show other equidistribution results from equidistribution of Kloosterman sums essentially.

Main issue
\begin{equation*} \sum_c \frac{K(m,n ; c)}{c} J_{k-1} \left(\frac{4\pi \sqrt{mn}}{c}\right) \end{equation*}

saw that bounding each individual \(K(m, n; c)\) falls short of what is needed.

Detour (Salié sums)

Recall for \(k = \frac 32 + l\) the sums we are getting are

\begin{equation*} K_k(m, n ; c) = \sum_{d \pmod{^* c}} \epsilon_d^{-2k} \left(\frac cd \right) e\left( \frac{ md + n \overline d}{d} \right) \end{equation*}

where

\begin{equation*} \legendre{c}{d} \text{ Kronecker symbol} \end{equation*}
\begin{equation*} \epsilon_d = \begin{cases} 1 \amp d \equiv1 \pmod 4,\\ i \amp d \equiv -1 \pmod 4\end{cases} \end{equation*}

It will turn out that these sums can be calculated in elementary terms. Analogy

\begin{equation*} K_k(m, n ; c) \leftrightarrow K_k \end{equation*}

for \(k \in \frac 12 + \ZZ\text{,}\) \(J_k(x)\) is expressible in terms of elementary functions, e.g.

\begin{equation*} J_{1/2}(x) = \sqrt{\frac{2}{\pi x}} \cdot \sin(x) \end{equation*}
\begin{equation*} J_{3/2}(x) = \sqrt{\frac{2}{\pi}} \cdot\left(\frac{\sin(x)}{x^{3/2}} -\frac{\cos(x)}{\sqrt x} \right) \end{equation*}

Calculation: Reductions:

Prove this.

  1. \begin{equation*} x = \overline m y \end{equation*}
  2. \begin{equation*} x = m \overline y \end{equation*}
  3. \begin{equation*} S(1, n^2; q) = \sum_{x\pmod{^* q}} \legendre{x}{q} e\left( \frac{x + n^2 \overline x}{q}\right) \end{equation*}
    recall DFT
    \begin{equation*} f(u) = \frac 1q \sum_{\alpha \pmod q} \hat f(\alpha) e\left( \frac{\alpha u}{q}\right) \end{equation*}
    \begin{equation*} \hat f(\alpha) = \sum _{u\pmod q} f (u) e\left( \frac{-\alpha u}{q}\right) \end{equation*}
    \begin{equation*} = \sum_{q\pmod q} \sum _{x\pmod{^* q}} \legendre xq e \left( \frac{x + u^2 \overline x}{q} \right) e\left(\frac{-\alpha u}{q} \right) \end{equation*}
    \begin{equation*} = \sum_{u \pmod q,x\pmod{^* q}} \legendre xq e\left(\frac xq\right) e\left( \frac{y^2 \overline x - \alpha u}{q}\right) \end{equation*}
    \begin{equation*} = \sum_{x\pmod{^* q}} \legendre xq e\left( \frac{x(1- \overline 4x^2)}{q}\right)\underbrace{ \sum_{u \pmod q} e\left(\frac{\overline x (u - 2\alpha x)^2}{q}\right)}_{\legendre xq G(q)} \end{equation*}
    \begin{equation*} G(q) \sum_{x \pmod{^* q}} e\left( \frac{ x (1 - \overline 4 \alpha^2)}{q}\right) \end{equation*}
    we need some more results to conclude!
\begin{equation*} c_q(r) = \sum_{x\pmod{^* q }}e\left( \frac{xr}{q} \right) \end{equation*}
\begin{equation*} =\sum_{d|q} \sum_{a \pmod {^* q/d}} e\left( \frac{ar}{q/d}\right) \end{equation*}
\begin{equation*} =\sum_{d|q} c_{q/d} (r) \end{equation*}
\begin{equation*} =\sum_{d|q} c_d(r) \end{equation*}

let

\begin{equation*} F(q)=\sum_{d|q} c_d(r) \end{equation*}

then by Möbius inversion

\begin{equation*} c_q(r) = \sum_{d|q} \mu\left( \frac dq\right) F(d) \end{equation*}
\begin{equation*} F(d) = \sum_{x\pmod d} \left(\frac{xr}{d} \right) = \begin{cases} d \amp \text{ if } d|r,\\0 \amp \text{otw} \end{cases} \end{equation*}
Note 2.38
\begin{equation*} c_q(r) = \mu \left( \frac{q}{\gcd(q,r)}\right) \frac{\phi (q)}{\phi(q/\gcd(q,r))} \end{equation*}

This lemma implies that in the proof above we have

\begin{equation*} G(q) c_{q} (q- \overline 4 \alpha^2) = G(q) \sum_{d|\gcd(q,(1-\overline 4 \alpha^2))} \mu \left(\frac qd \right) d \end{equation*}

so putting everything together

\begin{equation*} f(u) = \frac 1q \sum_{\alpha \pmod q} \left( G(q) \sum_{d|\gcd(q,(1-\overline 4 \alpha^2))} \mu \left(\frac qd \right) d \right) e\left(\frac{\alpha u}{q} \right) \end{equation*}
\begin{equation*} = \frac{G(q)}q \sum_{d|q} \sum_{\alpha \pmod q,\overline 4 \alpha^2 - 1 \equiv 0 \pmod d} d e\left(\frac{\alpha u}{q} \right) \end{equation*}

If \(d \ne q\) then \(de = q\text{,}\) \(e \ne 1\text{,}\) \(\alpha = \alpha_1 + d\alpha_2\) so

\begin{equation*} \sum_\alpha = \sum_{\alpha_1,d,\overline 4 \alpha_1 ^2 = 1 \pmod d} e\left( \frac{\alpha_1 u}{q}\right) \sum_{\alpha_2} e\left( \frac{\alpha_2 u}{q}\right) = 0\text{.} \end{equation*}

This finishes the proof for Salié sums.

Final input: Lower bound (Gauss, Siegel): Gauss showed

\begin{equation*} r_3(n) = \frac{24}{w(d)} h(d) \left( 1- \legendre d2 \right) \end{equation*}

where \(h(d)\) is the class number \(\QQ(\sqrt{-n})\) where \(d = \disc (\QQ{-n})\) and \(w(d)\) is the number of roots of unity in this field.

Show this

Question 2.43

How large is \(h(d)\text{?}\)

Euler, Siegel

Last time \(f\in S_k(\Gamma), k \in \frac 12 + \ZZ\) for \(n \) squarefree Iwaniec implies \(|a_f(n)| \ll_\epsilon n^{k/2 - 1/4 - 1/28 + \epsilon}\text{.}\) This gives \(r_3(n; P) \ll n^{k/2 - 1/4 - \delta}\text{.}\)

Finding the argument \(P = \) spherical harmonic, \(\deg (P) = l\) and \(h= l + 3/2\text{.}\)

\begin{equation*} r_3(n; P) = \sum_{|\alpha|^2 = n} P(\alpha)\text{.} \end{equation*}

Today:

\(P \equiv 1\) main contribution.

\(P \not \equiv 1\) error.

\(P \equiv 1\)
\begin{equation*} r_3(n ) = \sum_{|\alpha|^2= n }1 \end{equation*}

(with no local obstructions, i.e. \(n \) squarefree and \(n \not \equiv 7 \pmod 8\)).

Note 2.46

\(\tilde n \sim n\) as we are squarefree so off by at most a factor of 4. Then \(r_3(n) \sim h(\tilde n)\)

Prove the theorem.

We now wish to demonstrate a lower bound on \(h(\tilde n)\text{.}\)

Prove the corollary (Analytic class number formula)

Note 2.51

The constant in Siegel's theorem is ineffective, meaning one cannot actually calculate it (without proving there are no Landau-Siegel zeroes).

Finally Siegel + Iwaniec implies

\begin{equation*} \frac{1}{\sqrt n} \# \{ \alpha \in \ZZ^3 : |\alpha|^2 = n\} \end{equation*}

gets equidistributed on \(S^3\) as \(n \to \infty\text{.}\)

Recall this above was problem 1.

Problem 2

\(\Gamma = \SL_2(\ZZ)\)

\begin{equation*} \frac{1}{\# \Lambda_d} \sum_{z_Q \in \Lambda_d} f(z_Q) \to \int_{\Gamma \backslash \HH} f(z) \diff \mu (z) \end{equation*}

where for \(Q = ax^2 + bxy + cy^2\text{,}\) positive definite.

\begin{equation*} z_Q = \frac{-b + \sqrt{b^2 - 4ac}}{2a} \in \HH\text{.} \end{equation*}
Duke's proof
  1. What to replace the exponential sums, i.e. the \(P\)'s?
  2. Follow the same strategy and bound the nontrivial sums from above (\(P \not \equiv 1\)), trivial sums from below (\(P \equiv 1\)).

Recall \(\lb 0 ,1) \leadsto P = e(nx)\text{,}\) \(S^2 \leadsto P\) spherical harmonics, \(\Gamma \backslash \HH \leadsto\) automorphic forms \(E(z,s)\) or Maass cusp forms.

Digression \(\Delta_{\text{hyp}} \acts \Gamma\backslash \HH\) fourier transform on \(\RR\) then \(f(x) = \int_{i\RR} \hat f(\alpha) e(\alpha x) \diff x\)

\begin{equation*} S^1 \implies f(x) = \sum_{n\in \ZZ} \hat f (n) e(nx) \end{equation*}

i.e.

\begin{equation*} L^2( \Gamma\backslash \HH) = \oint_\RR E( z, \frac 12 + iit) \diff \mu (t) \end{equation*}
\begin{equation*} \bigoplus \sum_\lambda \phi_\lambda \end{equation*}

\(\phi_\lambda\) a Maass cusp form of eigenvalue \(\lambda\text{.}\)

Detour: Maass forms and Eisenstein series

Definition 2.53 Maass form

\(f\colon \HH \to \CC\) is a Maass form if \(\forall \gamma \in \Gamma = \SL_2(\ZZ)\text{.}\)

\begin{equation*} f(\gamma z) = f(z) \end{equation*}

\(f\) is an eigenfunction for

\begin{equation*} \Delta_{\text{hyp}} = -y^2 \left( \frac{\partial^2}{\partial x^2 } \frac{\partial^2}{\partial y^2 }\right) \end{equation*}
\begin{equation*} f(z) = O(y^N) \end{equation*}

for some \(N\text{.}\)

Cusp form if

\begin{equation*} \int_0^1 f(x + iy) \diff x = 0\text{.} \end{equation*}
Example 2.54 Eisenstein series
\begin{equation*} E(z,s) = \sum_{(c,d)=1} \frac{y^s}{|cz+d|^{2s}} \end{equation*}
\begin{equation*} = \sum_{\gamma \in \Gamma_\infty\backslash \Gamma} \im(\gamma z)^s \end{equation*}

Back on track

\begin{equation*} \sum_{|\alpha|^2 = n} P\left( \frac {\alpha}{|\alpha|}\right) \end{equation*}
\begin{equation*} \sum_{z_Q \in \Lambda_d} 1 \end{equation*}
\begin{equation*} \sum_{z_Q \in \Lambda_d} E(z_Q, \frac 12 + it) \end{equation*}
\begin{equation*} \sum_{z_Q \in \Lambda_d} \phi_\lambda(z_Q) \end{equation*}
\begin{equation*} \sum_{z_Q} \zeta(2s) E(z_Q,s) =\sum_{z_Q} \sum_{(u,v)\in \ZZ^2 \smallsetminus (0,0)} \frac{y_Q^2}{|uz_Q + v|^2s} \end{equation*}

\(\Lambda_d\) the class group of binary quadratic forms of discriminant \(d\text{.}\)

\begin{equation*} |uz_Q + v|^2 = |u\left(\frac{-b + \sqrt{d}}{2a}\right) + v|^2 \end{equation*}
\begin{equation*} = \frac{av^2 - uv b + cv^2}{a} \end{equation*}

so we get in the above

\begin{equation*} = \left(\frac{|d|}{4a}\right)^{s/2} \sum _{Q\in \Lambda_d} \frac{1}{Q(u,v)^s}\text{.} \end{equation*}

This gives

  1. \begin{equation*} \sum_{z_Q} 1 \sum |d|^{1/2} L(1,\chi_d) \gg_\epsilon |d|^{1/2 - \epsilon} \end{equation*}
    by Siegel.
  2. \begin{equation*} \sum_{z_Q} E(z_Q, \frac 12 + it ) \sim \frac{|\zeta(\frac 12 + it) L( \frac 12 + it , \chi_d)|}{|\zeta(1 + 2 it)|} |d|^{\frac 14} \end{equation*}
    want
    \begin{equation*} L(\frac 12 + it, \chi_d) \ll |d|^{\frac 14 - 8}\text{.} \end{equation*}

We also need:

Phragmén-Lindelöf principle gives convexity bound on the \(d\)-aspect

\begin{equation*} L(\frac 12 + it, \chi_d ) \ll_\epsilon | d|^{\frac 14 + \epsilon}\text{.} \end{equation*}

Subconvexity bound

What about

\begin{equation*} \sum \phi_\lambda (z_Q)\text{?} \end{equation*}

Harder even!

Last time

\begin{equation*} \frac{1}{|\Lambda_d|} \sum_{z_Q \in \Lambda_d} f(z_Q) \to \int_{\Gamma \backslash \HH} \diff \mu (z) \end{equation*}

Siegel's theorem gives a lower bound

\begin{equation*} \frac{1}{|\Lambda_d|} \sum_{z_Q \in \Lambda_d} 1 \end{equation*}

Subconvexity bound on \(L(\frac 12 + it, \chi_d)\) and non-vanishing \(\zeta(1 + 2i t)\text{.}\)

\begin{equation*} \frac{1}{|\Lambda_d|} \sum_{z_Q \in \Lambda_d} E(z_Q, \frac 12 + it) \end{equation*}

is

\begin{equation*} \frac{1}{|\Lambda_d|} \sum_{z_Q \in \Lambda_d} \phi(z_Q) \end{equation*}

a cusp form?

Waldspurger's formula

Roughly Waldspurger says

\begin{equation*} \frac{|a_{|D|}|}{\pair ff} = \frac{(k - \frac 32)!}{\pi} \frac{|D|^{k-1}}{\pair gg} L(g\otimes \chi_D, k- \frac 12) \end{equation*}

\(f\in S_k(\Gamma_0(4)\) half integral weight, \(f\mapsto g\) Shimura lift, \(a_n\) fourier coeffs of \(F\text{.}\)

Remark 2.59

Shimura lifting takes weight \(k\) to weight \(2k-1\) so \(k- \frac 12\) is the center of the critical strip of

\begin{equation*} L(g \otimes \chi_D, s)\text{.} \end{equation*}
\begin{equation*} L(g\otimes, \chi_d, s) = \sum \frac{b_n \chi_D(n)}{n^s}\text{.} \end{equation*}
\begin{equation*} \sum \phi (z_Q) \leadsto a_{\theta_\phi}(|D|) \end{equation*}

Katok-Sarnak, compare with

\begin{equation*} P\left(\frac{\alpha}{|\alpha|}\right) \leadsto \sum r_3(P, n) e(nz) = \theta_P(z) \end{equation*}

Then “all we need to do” is to show that

Subconvexity bound
\begin{equation*} L(g \otimes \chi_D, k- \frac 12) \ll |D|^{\frac 12 - \delta} \end{equation*}

this would be enough

Why? Recall

\begin{equation*} \sum_{|\alpha|^s = n} P(\alpha) \leadsto \theta_P(z) \end{equation*}
\begin{equation*} \sum_{|\alpha| = n} P\left( \frac{\alpha}{|\alpha|}\right) = \frac{1}{n^{\deg P/2}} r_3(P; n) \end{equation*}

\(\theta_P(z)\) of weight \(\deg P + \frac 32\text{.}\)

Want

\begin{equation*} \frac {|a_{|D|}|^2}{|D|^{\deg P}} \ll |D|^{1-s} \end{equation*}
\begin{equation*} g =\text{weight } 2k-1 - 2\deg P + 2 \end{equation*}
\begin{equation*} \sim \frac{|D|^{k-1}}{|D|^{k- \frac 32}} L(g\otimes \chi_D, k - \frac 12) \end{equation*}
\begin{equation*} \sim |D|^{\frac 12}L(g\otimes \chi_D, k - \frac 12) \end{equation*}

final touch:

\begin{equation*} \sum_{z_Q \in \Lambda_d} 1 \gg |d|^{1/2 - \epsilon} \end{equation*}

so if we can get the subconvex bound

\begin{equation*} \frac{|a_{|D|}|^2}{|D|^{k-3/2}} \ll |D|^{1-\delta} \end{equation*}

it is enough.

This follows from GRH.

\begin{equation*} |\zeta(1+ \epsilon + it)| \le \zeta(1+ \epsilon ) \le c_{\epsilon \gt 0}\text{.} \end{equation*}

Functional equation + Stirling gives

\begin{equation*} |\zeta( -\epsilon - it)| = O_\epsilon( |t|^{1/2 + \epsilon}) \end{equation*}
\begin{equation*} |\zeta(\sigma+ it)| \sim | \zeta(1- \sigma-it) \left| \frac{t}{2\pi} \right|^{\frac 12 - \sigma}\text{.} \end{equation*}
\begin{equation*} f(s) = \zeta(s) \zeta(1-s) \end{equation*}

+ Phragmén-Lindelöf, bound it on \(a = -\epsilon, b= 1+ \epsilon\text{:}\)

\begin{equation*} |f(s)| \le |t|^{1/2 +\delta}\text{.} \end{equation*}