-
Notifications
You must be signed in to change notification settings - Fork 0
/
3_Dirichlet.tex
304 lines (266 loc) · 14.5 KB
/
3_Dirichlet.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
% !TEX root = Daniel-Miller-thesis.tex
\chapter{Dirichlet series with Euler product}\label{ch:Dirichlet-series}
\section{Definitions and motivation}
We start by considering a very general class of Dirichlet series: those that
admit a product formula with degree 1 factors. The motivating example was
suggested to the author by Ramakrishna. Let $E_{/\bQ}$ be an elliptic
curve and let
\[
L_{\sgn}(E,s) = \prod_p \frac{1}{1-\sgn(a_p) p^{-s}} .
\]
How much can we say about the behavior of $L_{\sgn}(E,s)$? For example, does it
admit analytic continuation to $\Re = 1$? Yes, by
\cite[A.2]{serre-1989}. We will see later that the Akiyama--Tanigawa
conjecture implies the existence of a non-vanishing analytic continuation of
$L_{\sgn}(E,s)$ to $\Re > \frac 1 2$. Can the rank of $E$ be
found from $L_{\sgn}(E,s)$? Theoretically yes, by the following result, which
the author learned from Harris.
\begin{theorem}
If $E_1$ and $E_2$ are non-CM elliptic curves over $\bQ$ with
$\sgn a_p(E_1) = \sgn a_p(E_2)$ for all $p$, then $E_1$ and $E_2$ are
isogenous.
\end{theorem}
\begin{proof}
Assume by way of contradiction that $E_1$ and $E_2$ are non-isogenous, non-CM
elliptic curves over $\bQ$ with $\sgn a_p(E_1) = \sgn a_p(E_2)$ for all $p$. By
\cite[5.4]{harris-2009}, the pairs $(\theta_p(E_1),\theta_p(E_2))$ are
equidistributed with respect to
$\ST\times \ST = \frac{4}{\pi^2} \sin^2\theta_1\sin^2\theta_2\, \dd\theta_1\, \dd\theta_2$
on $[0,\pi]\times [0,\pi]$.
Recall that if $f(\theta) = 1_{[0,\pi/2)}(\theta) - 1_{(\pi/2,\pi]}(\theta)$,
then $\sgn(a_p) = f(\theta_p)$. Moreover,
$g(\theta_1,\theta_2) = |f(\theta_1) - f(\theta_2)|$ is non-negative and
continuous almost everywhere, and $g(\theta_p(E_1),\theta_p(E_2)) = 0$ if and
only if $\sgn a_p(E_1) = \sgn a_p(E_2)$. It is clear that
$\int g\, \dd \ST\times \ST > 0$. Harris' equidistribution result tells us that
\[
\int g\, \dd \ST\times\ST = \lim_{N\to \infty} \frac{1}{\pi(N)} \sum_{p\leqslant N} g(\theta_p(E_1),\theta_p(E_2)) = 0 ,
\]
which is a contradiction.
\end{proof}
It follows that if $L_{\sgn}(E_1,s) = L_{\sgn}(E_2,s)$ for all $s$ in some
right half-plane, then $E_1$ and $E_2$
are isogenous, so in particular they have the same rank. Can we recover the
rank of $E$ from the behavior of $L_{\sgn}(E,s)$ at $s = \frac 1 2$? For
$k\geqslant 1$, let $r_k$ be the order of vanishing of $L(\sym^k E,s)$ at
$s = \frac 1 2$. Also, for $f\in L^1([0,\pi],\ST) = L^1(\SU(2)^\natural)$, let
$\widehat f(\sym^k) = \int_{\SU(2)^\natural} f(x) \tr\sym^k(x)\, \dd x$ be the
$\sym^k$-Fourier coefficient of $f$. The heuristics in \cite{sarnak-2007}
suggest that if the Akiyama--Tanigawa (and other natural conjectures) hold,
then $L_f(E,s)$ has a zero of order
$\sum_{k\geqslant 1} \widehat f(\sym^k) \left(2 r_k + (-1)^k\right)$ at
$s = \frac 1 2$. If $f = \tr\sym^1$ (this is called $U_1$ in
\cite{sarnak-2007}), then $\widehat f(\sym^k)$ vanishes for $k\geqslant 2$, so
the order of vanishing of $L_{U_1}(E,s)$ at $s = \frac 1 2$ is $2 r_1 - 1$,
which, if we assume the Birch and Swinnerton-Dyer conjecture, allows us to
``read'' the rank of $E$ from the behavior of $L_{U_1}(E,s)$ at
$s = \frac 1 2$. However, for $f = 1_{[0,\pi/2)} - 1_{(\pi/2,\pi]}$, we have
$\widehat f(\sym^k) = 0$ when $k$ is even, and
$\widehat f(\sym^k) = \frac 2 \pi (-1)^{\frac{k-1}{2}} \left(\frac 1 k + \frac{1}{k+2}\right)$ when $k$ is odd. So we should expect that
\[
\ord_{s = \frac 1 2} L_{\sgn}(E,s) = \frac 2 \pi\sum_{k\text{ odd}}(-1)^{\frac{k-1}{2}} \left(\frac 1 k + \frac{1}{k+2}\right)\left(2 r_k + (-1)^k\right) .
\]
In light of this, it does not seem like the rank of $E$ can be directly read
from the behavior of $L_{\sgn}(E,s)$ at $s = \frac 1 2$.
\begin{definition}
Let $\bx=(x_2,x_3,x_5,\dots)$ be a sequence of complex numbers indexed by the
primes. The associated Dirichlet series is
$L(\bx,s) = \prod_p \left(1- x_p p^{-s}\right)^{-1}$.
\end{definition}
If $x_p$ is defined only for a subset of the primes, we tacitly set $x_p = 0$
(so the Euler factor is $1$) at all primes for which $x_p$ is not defined.
\begin{lemma}
Let $\bx$ be a sequence with $|\bx|_\infty \leqslant 1$. Then $L(\bx,s)$
defines a holomorphic function on the region $\Re > 1$. On that region,
$\log L(\bx,s) = \sum_{p^r} \frac{x_p^r}{r p^{r s}}$.
\end{lemma}
\begin{proof}
Expanding the product for $L(\bx,s)$ formally, we have
$L(\bx,s) = \sum_{n\geqslant 1} \frac{\prod_p x_p^{v_p(n)}}{n^s}$.
An easy comparison with the Riemann zeta function tells us that this sum
is holomorphic on $\Re > 1$. By \cite[Th.~11.7]{apostol-1976}, the
product formula holds in the same region. The formula for $\log L(\bx,s)$
comes from \cite[11.9 Ex.~2]{apostol-1976}.
\end{proof}
Abel summation is a commonly-used result that will allow us to turn questions
on the analytic continuation and non-vanishing of $L(\bx,s)$ into questions
about the asymptotics of $\sum_{p\leqslant N} x_p$.
\begin{lemma}[Abel summation]\label{lem:abel-sum}
Let $\bx=(x_2,x_3,x_5,\dots)$ be a sequence of complex numbers, $f$ a smooth
$\bC$-valued function on $\bR$. Then
\[
\sum_{p\leqslant N} f(p) x_p = f(N) \sum_{p\leqslant N} x_p - \int_2^N f'(t) \sum_{p\leqslant t} x_p\, \dd t .
\]
\end{lemma}
\begin{proof}
If $p_1,\dots,p_n$ is an enumeration of the primes $\leqslant N$, we have
\begin{align*}
\int_2^N f'(t) \sum_{p\leqslant t} x_p\, \dd t
&= \sum_{p\leqslant N} x_p \int_{p_n}^N f'(t)\, \dd t + \sum_{i=1}^{n-1} \sum_{p\leqslant p_i} x_p \int_{p_i}^{p_{i+1}} f'(t)\, \dd t \\
&= \left(f(N) - f(p_n)\right) \sum_{p\leqslant N} x_p + \sum_{i=1}^{n-1} \left(f(p_{i+1}) - f(p_i)\right) \sum_{p\leqslant p_i} x_p \\
&= f(N) \sum_{p\leqslant N} x_p - \sum_{p\leqslant N} f(p) x_p ,
\end{align*}
as desired.
\end{proof}
\begin{theorem}\label{thm:AT->RH}
Let $|\bx|_\infty \leqslant 1$, and assume
$|\sum_{p\leqslant N} x_p| \ll N^{\alpha+\epsilon}$ for some
$\alpha\in [\frac 1 2,1]$. Then the series for $\log L(\bx,s)$ converges
conditionally to a holomorphic function on $\Re > \alpha$.
\end{theorem}
\begin{proof}
Formally split the sum for $\log L(\bx,s)$ into two pieces:
\[
\log L(\bx,s) = \sum_p \frac{x_p}{p^s} + \sum_p \sum_{r\geqslant 2} \frac{x_p^r}{r p^{r s}} .
\]
For each $p$, we have
\[
\left| \sum_{r\geqslant 2} \frac{x_p^r}{r p^{r s}}\right| \leqslant \sum_{r\geqslant 2} p^{- r \Re s} = p^{-2 \Re s} \frac{1}{1-p^{-\Re s}} .
\]
Elementary analysis gives
$1 \leqslant \frac{1}{1-p^{-\Re s}} \leqslant 2 + 2\sqrt 2$, so the second
piece of $\log L(\bx,s)$ converges absolutely on $\Re >\frac 1 2$. We could
simply cite \cite[II.1 Th.~10]{tenenbaum-1995} to finish the proof; instead we
prove directly that $\sum \frac{x_p}{p^s}$ converges absolutely to a
holomorphic function on the region $\Re > \alpha$.
By Lemma \ref{lem:abel-sum} (Abel summation) with $f(t) = t^{-s}$, we have
\begin{align}
\sum_{p\leqslant N} \frac{x_p}{p^s}
&= N^{-s} \sum_{p\leqslant N} x_p + s \int_2^N \sum_{p\leqslant t} x_p\, \frac{\dd t}{t^{s+1}} \label{eq:log-L-sum} \\
&\ll N^{-\Re s + \alpha + \epsilon} + |s| \int_2^N t^{\alpha+\epsilon} \frac{\dd t}{t^{\Re s+1}} . \nonumber
\end{align}
Since $\alpha-\Re s < 0$, the first term is converges to zero. Since
$\Re s+1-\alpha > 1$ and
$\epsilon$ is arbitrary, the integral converges absolutely, and the proof is
complete.
\end{proof}
The proof of Theorem \ref{thm:AT->RH} actually gives an absolutely
convergent expression for $\log L(\bx,s)$ on the region $\Re >\alpha$. Since
the term $N^{-s} \sum_{p\leqslant N} x_p$ in \eqref{eq:log-L-sum}
converges to zero, we get
\[
\log L(\bx,s) = s \int_2^\infty t^{-s-1}\left(\sum_{p\leqslant t} x_p\right) \dd t + \sum_p \sum_{r\geqslant 2} \frac{x_p^r}{r p^{r s}} .
\]
Let $X$ be a topological space, $f\colon X\to \bC$ a function with
$|f|_\infty\leqslant 1$, and $\bx=(x_2,x_3,\dots)$ a sequence in $X$. Write
\[
L_f(\bx,s) = \prod_p \frac{1}{1-f(x_p) p^{-s}} ,
\]
for the associated Dirichlet series. In the remainder, we will
exclusively focus on Dirichlet series of this type.
\section{Automorphic and motivic \texorpdfstring{$L$}{L}-functions}
Suppose $G$ is a compact group, $G^\natural$ the space of conjugacy classes in
$G$. If $\bx = (x_2,x_3,x_5,\dots)$ is a sequence in $G^\natural$ and $\rho$ is
a finite-dimensional unitary representation of $G$, put
\[
L(\rho(\bx),s) = \prod_p \frac{1}{\det(1-\rho(x_p) p^{-s})} .
\]
Clearly $L((\rho_1\oplus \rho_2)(\bx),s) = L(\rho_1(\bx),s) L(\rho_2(\bx),s)$.
Now, suppose $G$ is a compact connected Lie group, let $T\subset G$ be a
maximal torus, and recall that
$T\twoheadrightarrow G^\natural$ \cite[IX.5 Prop.~5]{bourbaki-2005}.
The representation
$\left.\rho\right|_T$ decomposes as $\bigoplus \chi^{\oplus m_\chi}$, where
$\chi$ ranges over characters of $T$ and the entire expression is
$W$-invariant. We may regard the $x_p$ as lying in $T/W$, so we have
\[
L(\rho(\bx),s) = \prod_\chi L(\chi(\bx),s)^{m_\chi} .
\]
If the trivial representation appears in $\left.\rho\right|_T$, this product
formula will include a copy (possibly several) of $\zeta(s)$. Since
$\chi(x_p) \in S^1$, the above formula decomposes $L(\rho(\bx),s)$ into a
product of Dirichlet series of the type considered above. For $G = \SU(2)$, the
trivial representation occurs in $\left.\sym^k \right|_T$ if and only if $k$ is
even, and the resulting $\zeta(s)$ in the product decomposition of
$L(\sym^k \bx,s)$ may explain why some of the results in this thesis
(e.g.~Theorem \ref{thm:bad-Galois}) only apply
to \emph{odd} symmetric powers.
\begin{theorem}\label{thm:AT->RH:gp}
Let $G^\natural$ be the space of conjugacy classes in a compact group,
$\bx = (x_2,x_3,x_5,\dots)$ a sequence in $G^\natural$. If $\rho$ is a
nontrivial unitary representation of $G$ and
$\left| \sum_{p\leqslant N} \tr \rho(x_p)\right| \ll N^{\alpha+\epsilon}$ for
some $\alpha\in \left[\frac 1 2,1\right]$, then $L(\rho(\bx),s)$ admits a
non-vanishing analytic continuation to $\Re > \alpha$.
\end{theorem}
\begin{proof}
On $\Re > 1$, we have
$\log L(\rho(\bx),s) = \sum_{r\geqslant 1} \sum_p \frac{\tr \rho(x_p)^r}{r p^{r s}}$.
Just as in the proof of Theorem \ref{thm:AT->RH}, we can split the sum into
two terms, $\sum_p \frac{\tr \rho(x_p)}{p^s}$ and
$\sum_{r\geqslant 2} \sum_p \frac{\tr \rho(x_p)^r}{r p^{r s}}$. Analytic
continuation and nonvanishing remain the same when we omit finitely many Euler
factors, so we may ignore all primes for which
$\dim(\rho) \geqslant \sqrt p$. Then the second sum converges on
$\Re > \frac 1 2$, and assuming
$\left| \sum_{p\leqslant N} \tr \rho(x_p)\right| \ll N^{\alpha+\epsilon}$, an
argument identical to the one used in the proof of Theorem \ref{thm:AT->RH},
using Abel summation, shows that
$\sum_p \frac{\tr \rho(x_p)}{p^s}$ converges to a holomorphic function on
$\Re > \alpha$. This yields the desired non-vanishing analytic continuation.
\end{proof}
\section{Discrepancy and the Riemann hypothesis}
\begin{definition}
We say the \emph{Riemann hypothesis} for $L(\bx,s)$ holds if the function
$\log L(\bx,s)$ admits analytic continuation to $\Re > \frac 1 2$.
\end{definition}
Under reasonable analytic hypotheses, namely conditional convergence of
the Dirichlet series for
$\log L(\bx,s)$ on $\Re > \frac 1 2$, the result
\cite[II.1 Th.~10]{tenenbaum-1995} gives
an estimate $|\sum_{p\leqslant N} x_p| \ll N^{\frac 1 2 + \epsilon}$.
\begin{theorem}
Let $(X,\mu)$ be a probability space in which discrepancy and Koksma--Hlawka
make sense (i.e., Theorem \ref{thm:koksma-hlawka} applies), and let
$\bx=(x_2,x_3,x_5,\dots)$ be a sequence in $X$ with
$\D_N(\bx,\mu) \ll N^{-\frac 1 2+\epsilon}$. For any function $f$ on $X$ of
bounded variation with $\int f\, \dd\mu = 0$, $L_f(\bx,s)$ satisfies
the Riemann hypothesis.
\end{theorem}
\begin{proof}
By the Koksma--Hlawka inequality (Theorem \ref{thm:koksma-hlawka}), the bound
on discrepancy yields the estimate
$\left| \sum_{p\leqslant N} f(x_p)\right| \ll N^{\frac 1 2+\epsilon}$.
By Theorem \ref{thm:AT->RH}, the Riemann hypothesis holds for $L_f(\bx,s)$.
\end{proof}
The same proof shows that if $\D_N(\bx,\mu)\ll N^{-\alpha+\epsilon}$, then
$\log L_f(\bx,s)$ conditionally converges to a holomorphic function on
$\Re > 1 - \alpha$. This theorem applied to the function $L_{\sgn}(E,s)$ shows
that the Akiyama--Tanigawa conjecture implies the Riemann hypothesis for
$L_{\sgn}(E,s)$. The author is unaware of any results, conditional or
otherwise, that suggest $L_{\sgn}(E,s)$ has analytic continuation past
$\Re = \frac 1 2$ or has any kind of functional equation. Also, if
$\int f\, \dd\mu \ne 0$, the function $L_f(\bx,s)$ will have a singularity at
$s = 1$, but the author is not aware of a way to continue the function
$L_f(\bx,s)$ past $\Re = 1$, even if $\D_N(\bx,\mu)$ decays rapidly.
Let $F = \bF_q(t)$ be a function field, $E_{/F}$ a generic elliptic curve.
There is, for every prime $\fp$ of $F$, a Satake parameter
$\theta_\fp \in [0,\pi]$, defined in the usual way. It is known
\cite[Ch.~3]{katz-1988} that
\begin{equation}\label{eq:katz-char-bound}
\left|\sum_{\N(\fp) \leqslant x} \tr\sym^k\smat{e^{i \theta_\fp}}{}{}{e^{- i \theta_\fp}} \right| \ll k \sqrt{x} .
\end{equation}
Briefly, let $G$ be a compact Lie group, $\rho$ an irreducible unitary
representation of $G$. For $f\in L^1(G)$, the Fourier coefficient of $f$ at
$\rho$ is $\widehat f(\rho) = \int f(x) \overline{\tr\rho(x)}\, \dd x$. If
$f\in L^1(G^\natural)$, then $f = \sum_\rho \widehat f(\rho) \tr \rho$. When
$G = \SU(2)$, the nontrivial irreducible unitary representations are
$\sym^k$ for $k\geqslant 1$.
Equation \eqref{eq:katz-char-bound} tells us that for any
$f\in C(\SU(2)^\natural)$ with
$\sum_{k\geqslant 1} |\widehat f(\sym^k)|<\infty$ and $\widehat f(\sym^0) = 0$,
the strange Dirichlet series $L_f(\btheta,s)$ satisfies the Riemann hypothesis.
The best estimate on discrepancy is found in
\cite{niederreiter-1991}, where it is shown that $D_x \ll N^{-\frac 1 4}$
by applying a generalization of the Koksma--Hlawka inequality to
$\SU(2)^\natural$. Namely, for any odd $r$, we have
\[
\D_x(\btheta,\ST) \ll \frac 1 r + \sum_{k=1}^{2 r -1} \frac 1 k \left| \frac{1}{\pi_F(x)} \sum_{\N(\fp)\leqslant x} \tr\sym^k\smat{e^{i \theta_\fp}}{}{}{e^{- i \theta_\fp}}\right| .
\]
Using the estimate \eqref{eq:katz-char-bound} on character sums, Niederreiter
is able to derive $\D_x \ll x^{-\frac 1 4}$. This fits well with the results of
\cite{bucar-kedlaya-2015,rouse-thorner-2016} in the number field case, both of
which derive estimates of the form $\D_N \ll N^{-\frac 1 4+\epsilon}$, assuming
both the generalized Riemann hypothesis and the functional equation for all
symmetric-power $L$-functions associated to the (non-CM) elliptic curve in
question.