-
Notifications
You must be signed in to change notification settings - Fork 0
/
5_representations.tex
595 lines (527 loc) · 29.8 KB
/
5_representations.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
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
% !TEX root = Daniel-Miller-thesis.tex
\chapter{Pathological Galois representations}\label{ch:construct-Galois}
\section{Notation and supporting results}
In this section we loosely summarize and adapt the results of
\cite{khare-larsen-ramakrishna-2005,pande-2011}. Throughout, if $F$ is a field
and $M$ a $G_F$-module, we write $\h^\bullet(F,M)$ in place of
$\h^\bullet(G_F,M)$. All Galois representations will take values in
$\GL_2(\bZ/l^n)$ or $\GL_2(\bZ_l)$ for $l$ a (fixed) rational prime, and
all deformations will have fixed determinant. So we consider the cohomology of
$\Ad^0\bar\rho$, the induced representation on trace-zero matrices by
conjugation.
If $S$ is a set of rational primes, $\bQ_S$ denotes the largest extension of
$\bQ$ unramified outside $S$. So $\h^i(\bQ_S,-)$ is what is usually written as
$\h^1(G_{\bQ,S},-)$. If $M$ is a $G_\bQ$-module and $S$ a finite set of primes,
denote the corresponding Tate--Shafarevich group by
\[
\Sha^i_S(M) = \ker\left( \h^i(\bQ_S,M) \to \prod_{p\in S} \h^i(\bQ_p,M)\right) .
\]
If $l$ is a rational prime and $S$ a finite set of primes containing $l$, then
for any $\bF_l[G_{\bQ_S}]$-module $M$, write $M^\vee=\hom_{\bF_l}(M,\bF_l)$
with the obvious $G_{\bQ_S}$-action, and write $M^\ast = M^\vee(1)$ for the
Cartier dual of $M^\vee$. By \cite[Th.~8.6.7]{neukirch-schmidt-winberg-2008},
there is an isomorphism $\Sha^1_S(M^\ast) \simeq \Sha_S^2(M)^\vee$. As a
result, if $\Sha_S^1(M)$ and $\Sha_S^2(M)$ are trivial, and $S\subset T$, then
$\Sha_T^1(M)$ and $\Sha_T^2(M)$ are also both trivial.
\begin{definition}
A \emph{good residual representation} is an odd, absolutely irreducible,
weight-$2$ representation $\bar\rho\colon G_{\bQ_S} \to \GL_2(\bF_l)$, where
$l\geqslant 5$ is a rational prime.
\end{definition}
Recall that $\bar\rho$ is weight-$2$ if $\det\bar\rho$ is the mod-$l$
cyclotomic character. Similarly, $\rho\colon G_\bQ \to \GL_2(\bZ_l)$ is
weight-$2$ if $\det\rho$ is the $l$-adic cyclotomic character.
Roughly, ``good residual representations'' have enough properties that we can
prove meaningful theorems about their lifts without assuming the modularity
results of Khare--Wintenberger.
\begin{theorem}[{\cite[Th.~1.3]{taylor-2003}}]\label{thm:always-can-lift}
Let $\bar\rho\colon G_\bQ \to \GL_2(\bF_l)$ be a good residual
representation. Then there exists a finitely ramified weight-$2$ lift of
$\bar\rho$ to $\bZ_l$.
\end{theorem}
\begin{definition}
Let $\bar\rho\colon G_{\bQ_S} \to \GL_2(\bF_l)$ be a good residual
representation. A prime $p\not\equiv \pm 1\pmod l$ is \emph{nice} if
$\Ad^0\bar\rho\simeq \bF_l \oplus \bF_l(1)\oplus \bF_l(-1)$, i.e.~if the
eigenvalues of $\bar\rho(\frob_p)$ have ratio $p$.
\end{definition}
Taylor allows $p\equiv -1\pmod l$, but the results of \cite{pande-2011}
require $p\not\equiv -1\pmod l$. The following theorem gives a complete
description of the versal deformation ring for
$\left.\bar\rho\right|_{G_{\bQ_p}}$ when $p$ is nice.
\begin{theorem}[\cite{ramakrishna-1999}]
Let $\bar\rho$ be a good residual representation and $p$ a nice prime. Then
any deformation of $\left.\bar\rho\right|_{G_{\bQ_p}}$ is induced by
$G_{\bQ_p} \to \GL_2(\bZ_l\pow{a,b} / \langle a b\rangle)$, sending
\[
\frob_p \mapsto \smat{p(1+a)}{}{}{(1+a)^{-1}} \qquad \tau_p \mapsto \smat{1}{b}{}{1} ,
\]
where $\tau_p\in G_{\bQ_p}$ is a generator for tame inertia.
\end{theorem}
We close this section by introducing some new terminology and notation to
condense the lifting process used in \cite{khare-larsen-ramakrishna-2005}.
Fix a good residual representation $\bar\rho$. We will consider weight-$2$
deformations of $\bar\rho$ to $\bZ/l^n$ and $\bZ_l$. Call such a deformation a
``lift of $\bar\rho$ to $\bZ/l^n$ (resp.~$\bZ_l$).'' We will often restrict the
local behavior of such lifts, i.e.~the restrictions of a lift to $G_{\bQ_p}$
for $p$ in some set of primes. The necessary constraints are captured in the
following definition.
\begin{definition}
Let $\bar\rho$ be a good residual representation,
$h\colon \bR^+ \to \bR_{\geqslant 1}$ an
increasing function. An \emph{$h$-bounded lifting datum} is a tuple
$(\rho_n,R_n,U_n,\{\rho_p\}_{p\in R_n\cup U_n})$, where
\begin{enumerate}
\item
$\rho_n\colon G_{\bQ_{R_n}} \to \GL_2(\bZ/l^n)$ is a lift of $\bar\rho$.
\item
$R_n$ and $U_n$ are finite sets of primes, $R_n$ containing $l$ and all primes at
which $\rho_n$ ramifies.
\item
$\pi_{R_n}(x)\leqslant h(x)$ for all $x$.
\item
Both $\Sha_{R_n}^1(\Ad^0\bar\rho)$ and $\Sha_{R_n}^2(\Ad^0\bar\rho)$ are
trivial.
\item
For all $p\in R_n\cup U_n$,
$\rho_p\colon G_{\bQ_p} \to \GL_2(\bZ_l)$ satisfies
$\rho_p\equiv \left. \rho_n\right|_{G_{\bQ_p}}\pmod{l^n}$.
\item
For all $p\in R_n$, $\rho_p$ is ramified.
\item
$\rho_n$ admits a lift to $\bZ/l^{n+1}$.
\end{enumerate}
\end{definition}
If $(\rho_n,R_n,U_n,\{\rho_p\})$ is an $h$-bounded lifting datum, we call
another $h$-bounded lifting datum $(\rho_{n+1},R_{n+1},U_{n+1},\{\rho_p\})$ a
\emph{lift} of $(\rho_n,R_n,U_n,\{\rho_p\})$ if $U_n\subset U_{n+1}$,
$R_n\subset R_{n+1}$, and for all
$p\in R_n\cup U_n$, the two possible $\rho_p$ agree.
\begin{theorem}\label{thm:lifting-datum}
Let $\bar\rho$ be a good residual representation,
$h\colon \bR^+ \to \bR_{\geqslant 1}$
increasing to infinity. If $(\rho_n,R_n,U_n,\{\rho_p\})$ is an $h$-bounded lifting
datum, $U_{n+1}\supset U_n$ is a finite set of primes disjoint from $R_n$, and
$\{\rho_p\}_{p\in U_{n+1}}$ extends $\{\rho_p\}_{p\in U_n}$, then there exists an
$h$-bounded lift $(\rho_{n+1},R_{n+1},U_{n+1},\{\rho_p\})$ of
$(\rho_n,R_n,U_n,\{\rho_p\})$.
\end{theorem}
\begin{proof}
By \cite[Lem.~8]{khare-larsen-ramakrishna-2005}, there exists a finite set
$N$ of nice primes such that the map
\begin{equation}\label{eq:h1-isom}
\h^1(\bQ_{R_n\cup N},\Ad^0\bar\rho) \to \prod_{p\in R_n} \h^1(\bQ_p,\Ad^0\bar\rho) \times \prod_{p\in U_{n+1}} \h_\nr^1(\bQ_p,\Ad^0\bar\rho)
\end{equation}
is an isomorphism. In fact,
$\# N = \dim\h^1(\bQ_{R_n\cup U_n},\Ad^0\bar\rho^\ast)$, and the primes in $N$ are
chosen, one at a time, from Chebotarev sets. Since $\pi_{R_n}(x)$ is eventually
constant and $h(x)$ increases to infinity, $h(x) \geqslant \pi_{R_n}(x)+1$ for all
$x\geqslant C_1$ for some $C_1$. Choose the first prime $p$ in $N$ to be
$\geqslant C_1$; then $\pi_{R_n\cup\{p\}}(x) \leqslant h(x)$ for all $x$. Repeat
this process for for all the other primes in $N$. We can ensure
that the bound $\pi_{R_n\cup N}(x) \leqslant h(x)$ continues to hold. We also
choose the primes in $N$ to be larger than any prime in $U_{n+1}$.
By our hypothesis, $\rho_n$ admits a lift to $\bZ/l^{n+1}$; call one such lift
$\rho^\ast$. For each $p\in R_n\cup U_{n+1}$, $\h^1(\bQ_p,\Ad^0\bar\rho)$ acts
transitively on lifts of $\left.\rho_n\right|_{G_{\bQ_p}}$ to $\bZ/l^{n+1}$. In
particular, there are cohomology classes $f_p\in \h^1(\bQ_p,\Ad^0\bar\rho)$
such that $f_p\cdot \rho^\ast \equiv \rho_p\pmod{l^{n+1}}$ for all
$p\in R_n\cup U_{n+1}$. Moreover, for all $p\in U_{n+1}$, the class $f_p$ is
unramified. Since the map \eqref{eq:h1-isom} is an isomorphism, there exists
$f\in \h^1(\bQ_{R_n\cup N},\Ad^0\bar\rho)$ such that
$\left.f\cdot \rho^\ast\right|_{G_{\bQ_p}}\equiv \rho_p\pmod{l^{n+1}}$ for all
$p\in R_n\cup U_{n+1}$.
Clearly $\left. f\cdot \rho^\ast\right|_{G_{\bQ_p}}$ admits a lift to $\bZ_l$
for all $p\in R_n\cup U_{n+1}$, but it does not necessarily admit such a lift for
$p\in N$. By repeated applications of \cite[Prop.~3.10]{pande-2011}, there
exists a set $N'\supset N$, with $\# N'\leqslant 2\# N$, of nice primes and
$g\in \h^1(\bQ_{R_n\cup N'},\Ad^0\bar\rho)$ such that
$(g+f)\cdot \rho^\ast$ still agrees with $\rho_p$ for $p\in R_n\cup U'$, and
$(g+f)\cdot \rho^\ast$ is nice for all $p\in N'$. As above, the primes in $N'$
are chosen one at a time from Chebotarev sets, so we can continue to ensure the
bound $\pi_{R_n\cup N'}(x)\leqslant h(x)$ and also that all primes in $N'$
are larger than those in $U_{n+1}$. Let $\rho_{n+1} = (g+f) \cdot \rho^\ast$. Let
$R_{n+1} = R_n\cup \{p\in N' : \rho_{n+1}\text{ is ramified at }p\}$. For each
$p\in R_{n+1}\smallsetminus R_n$, choose a lift $\rho_p$ of
$\left. \rho_{n+1}\right|_{G_{\bQ_p}}$ to $\bZ_l$.
Since $\left.\rho_{n+1}\right|_{G_{\bQ_p}}$ admits a lift to $\bZ/l^{n+2}$ (in
fact, it admits a lift to $\bZ_l$) for each $p$, and
$\Sha_{R_{n+1}}^1(\Ad^0\bar\rho)$, $\Sha_{R_{n+1}}^2(\Ad^0\bar\rho)$ are
trivial, the deformation $\rho_{n+1}$ admits a lift to $\bZ/l^{n+2}$. The tuple
$(\rho_{n+1},R_{n+1},U_{n+1},\{\rho_p\})$ is the desired lift of
$(\rho_n,R_n,U_n,\{\rho_p\})$ to $\bZ/l^{n+1}$.
\end{proof}
\section{Galois representations with specified Satake parameters}
Fix a good residual representation $\bar\rho$, and consider weight-$2$
deformations of $\bar\rho$. The final deformation,
$\rho\colon G_\bQ \to \GL_2(\bZ_l)$, will be constructed as the inverse limit
of a compatible collection of lifts $\rho_n\colon G_\bQ \to \GL_2(\bZ/l^n)$. At
any given stage, we will be concerned with making sure that there exists a
lift to the next stage, and that there is a lift with the necessary properties.
Fix a sequence $\bx=(x_1,x_2,\dots)$ in $[-1,1]$. The set of unramified primes
of $\rho$ is not determined at the beginning, but at each stage there will be
a large finite set $U$ of primes which we know will remain unramified.
Reindexing $\bx$ by these unramified primes, we will construct $\rho$ so that
for all unramified primes $p$, $\tr\rho(\frob_p)\in \bZ$, satisfies the Hasse
bound, and has $\frac{\tr\rho(\frob_p)}{2\sqrt p} \approx x_p$.
Moreover, we can ensure that the
set of ramified primes has density zero in a very strong sense (controlled by a
parameter function $h$) and that our trace of Frobenii are very close to
specified values.
Given any deformation $\rho$, write $\pi_{\ram(\rho)}(x)$ for the function
which counts $\rho$-ramified primes $\leqslant x$. Since we will have
$\pi_{\ram(\rho)}(x)\ll h(x)$ and bounds of this form are only helpful
if $h(x) = o(\pi(x))$, we will usually assume $h(x) \ll x^\epsilon$,
e.g.~$h(x) = \log x$ or something which grows even slower (for example, the
inverse of the Ackermann function). In \cite{khare-rajan-2001}, it is proved
that for \emph{any} continuous semisimple $\rho\colon G_\bQ \to \GL_2(\bZ_l)$,
we will have $\pi_{\ram(\rho)}(x) = o(\pi(x))$. That is, any continuous
Galois representation we consider will be ramified at a density zero set of
primes. However, by \cite[Th.~19]{khare-larsen-ramakrishna-2005}, it is
possible for $\pi_{\ram(\rho)}(x)$ to be
$\Omega(\frac{x}{\log(x)^{1+\epsilon}})$. This means the ability to bound
$\pi_{\ram(\rho)}(x)$ by slow-growing functions like $\log(x)$ in the
following result is non-trivial.
\begin{theorem}\label{thm:master-Galois}
Let $l$, $\bar\rho$, $\bx$ be as above. Fix a function
$h\colon \bR^+ \to \bR_{\geqslant 1}$ which increases to infinity. Then there
exists a weight-$2$ deformation $\rho$ of $\bar\rho$, such that:
\begin{enumerate}
\item
$\pi_{\ram(\rho)}(x) \ll h(x)$.
\item
For each unramified prime $p$, $a_p=\tr\rho(\frob_p)\in \bZ$ and satisfies the
Hasse bound.
\item
For each unramified prime $p$,
$\left| \frac{a_p}{2\sqrt p} - x_p\right| \leqslant \frac{l h(p)}{2\sqrt p}$.
\end{enumerate}
\end{theorem}
\begin{proof}
Begin with $\rho_1= \bar\rho$. By Theorem \ref{thm:always-can-lift}, each
$\left.\rho_1\right|_{G_{\bQ_p}}$ admits a lift to $\bZ_l$.
By \cite[Lem.~6]{khare-larsen-ramakrishna-2005},
there exists a finite set $R$, containing the set of primes at which $\bar\rho$
ramifies, such that $\Sha_R^1(\Ad^0\bar\rho)$ and $\Sha_R^2(\Ad^0\bar\rho)$ are
trivial.
Let $R_1$ be the union of $R$ and all primes $p$ with
$\frac{l}{2\sqrt p} > 2$. Since $\frac{l}{2\sqrt p} \to 0$ as $p\to \infty$,
the set $R_1$ is finite. For all $p\notin R_1$ and any $a\in \bF_l$, there
exists $a_p\in \bZ$ satisfying the Hasse bound with $a_p\equiv a\pmod l$. In
fact, given any $x_p\in [-1,1]$ and $a\in \bF_l$, there exists $a_p\in \bZ$
satisfying the Hasse bound, congruent to $a$ modulo $l$, such that
$\left| \frac{a_p}{2\sqrt p} - x_p\right| \leqslant \frac{l}{2\sqrt p}$.
Choose, for all primes $p\in R_1$, a ramified
lift $\rho_p$ of $\left. \rho_1\right|_{G_{\bQ_p}}$. Let $U_1$ be the set of
primes $p$ not in $R_1$ such that
$\frac{l^2}{2\sqrt p} > \min\left(2, \frac{l h(p)}{2\sqrt p}\right)$; this is
finite because $\frac{l^2}{2\sqrt p} \to 0$ and also eventually
$h(p) \geqslant l$. If $U_1$ is empty, then the next few sentences of the
proof are superfluous, but the theorem still holds.
For each $p\in U_1$, there exists $a_p\in \bZ$, satisfying the
Hasse bound, such that
\[
\left| \frac{a_p}{2\sqrt p} - x_p\right| \leqslant \frac{l}{2\sqrt p} \leqslant \frac{l h(p)}{2\sqrt p} ,
\]
and moreover $a_p\equiv \tr\bar\rho(\frob_p)\pmod l$. For each $p\in U_1$, let
$\rho_p$ be an unramified lift of $\left.\bar\rho\right|_{G_{\bQ_p}}$ with
$\tr\rho_p$ being the desired $a_p$. It may not be that
$\pi_{R_1}(x) \leqslant h(x)$ for all $x$. Let
$C = \max\left\{\pi_{R_1}(x)\right\}$; this is finite because
$R_1$ is and $\pi_{R_1}(x)$ is constant past the largest prime in $R_1$. Then
for $h^\ast = C h$, we have $\pi_{R_1}(x) \leqslant h^\ast(x)$ for all $x$.
We have constructed our first $h^\ast$-bounded lifting datum
$(\rho_1,R_1,U_1,\{\rho_p\})$. We proceed to construct
$\rho = \varprojlim \rho_n$ inductively, by constructing a new $h^\ast$-bounded
lifting datum for each $n$. We ensure that $U_n$ contains all primes for which
$\frac{l^{n+1}}{2\sqrt p} > \min\left(2, \frac{l h(p)}{2\sqrt p}\right)$, so
there are always integral $a_p$ satisfying the Hasse bound which satisfy any
mod-$l^{n+1}$ constraint, and that can always choose these $a_p$ so as to
preserve statement 2 in the theorem.
The base case is complete, so suppose we have
$(\rho_{n-1},R_{n-1},U_{n-1},\{\rho_p\})$. We may assume that $U_{n-1}$
contains all primes for
which $\frac{l^n}{2\sqrt p} > \min\left(2, \frac{l h(p)}{2\sqrt p}\right)$. Let
$U_n$ be the set of all primes not in $R_{n-1}$ such that
$\frac{l^{n+1}}{2\sqrt p} > \min\left(2, \frac{l h(p)}{2\sqrt p}\right)$. For
each $p\in U_n\smallsetminus U_{n-1}$, there is an integer $a_p$, satisfying
the Hasse bound, such that $a_p\equiv \rho_n(\frob_p)\pmod{l^n}$, and moreover
$\left|\frac{a_p}{2\sqrt p} - x_p\right| \leqslant \frac{l^n}{2\sqrt p}$.
Since $p\notin U_{n-1}$, we know that $l^n\leqslant l h(p)$,
s the bound in the previous sentence implies
$\left|\frac{a_p}{2\sqrt p} - x_p\right| \leqslant \frac{l h(p)}{2\sqrt p}$.
For $p\in U_n\smallsetminus U_{n-1}$, let $\rho_p$ be an unramified lift of
$\left. \rho_n\right|_{G_{\bQ_p}}$ such that $\tr\rho_n(\frob_p)$ is the
desired $a_p$. By Theorem \ref{thm:lifting-datum}, there exists an
$h^\ast$-bounded lifting datum $(\rho_n,R_n,U_n,\{\rho_p\})$ extending and
lifting $(\rho_{n-1},R_{n-1},U_{n-1},\{\rho_p\})$. This completes the inductive
step.
\end{proof}
The implied constant in the bound $\pi_{\ram(\rho)}(x)\ll h(x)$ depends on
$\bar\rho$ (and hence $l$) but not on $h$.
We will apply this theorem to construct Galois representations with specified
Sato--Tate distributions in the next section, but for now here is a small
consequence, which addresses the results in \cite{sarnak-2007}. Sarnak, assuming
the generalized Riemann hypothesis along with linear independence of the
zeros of $L(\sym^k E,s)$, proves
that for $E_{/\bQ}$ a non-CM elliptic curve of rank $r$, the partial sums
$\frac{\log x}{\sqrt x} \sum_{p\leqslant x} \frac{a_p}{\sqrt p}$ approach a
limiting distribution with mean $1 - 2 r$.
\begin{corollary}
Let $L \in [-\infty,\infty]$ and $\epsilon>0$ be given. Then there exists a
weight 2 Galois representation $\rho\colon G\to \GL_2(\bZ_l)$, such that
each $a_p = \tr\rho(\frob_p)\in \bZ$ satisfies the Hasse bound,
\[
L = \lim_{N\to \infty} \frac{\log N}{\sqrt N}\sum_p \frac{a_p}{\sqrt p}
\]
and $\pi_{\ram(\rho)}(x) \ll \log(x)$.
\end{corollary}
\begin{proof}
Begin with a sequence $(x_p)$ in $\left[-\frac 1 2,\frac 1 2\right]$ such that
$\lim_{N\to \infty} \frac{\log N}{\sqrt N}\sum_{p\leqslant N} x_p = L$. If
$L=\pm \infty$, we can choose $x_p = \pm \frac 1 2$. By Theorem
\ref{thm:master-Galois}, there exists $\rho\colon G_\bQ \to \GL_2(\bZ_l)$ with
$\pi_{\ram(\rho)}(x) \ll \log(x)$, and such that for each unramified $p$,
$a_p = \tr\rho(\frob_p)\in \bZ$, satisfies the Hasse bound. Moreover, after
re-indexing $(x_p)$ by the unramified primes of $\rho$, we can have
$\left| \frac{a_p}{2\sqrt p} - x_p\right| < \frac{l \log p}{\sqrt p}$. In fact,
by inspecting the proof of Theorem \ref{thm:master-Galois}, we can even ensure
that the partial sums
$\sum_{p\leqslant N} \left(\frac{a_p}{2\sqrt p} - x_p\right)$ are bounded.
How? When choosing $a_p$, all that Theorem \ref{thm:master-Galois} requires is
for $a_p$ to be sufficiently close to $x_p$. Since
$x_p \in \left[-\frac 1 2,\frac 1 2\right]$, if the partial sum up to (but not
including $p$) is $<0$, choose $a_p$ so that $\frac{a_p}{2\sqrt p}$ is to the
right of $x_p$ (hence $\frac{a_p}{2\sqrt p} - x_p$ is positive). If the
partial sum is $>0$, choose $a_p$ so that $\frac{a_p}{2\sqrt p}$ is to the left
of $x_p$ (hence $\frac{a_p}{2\sqrt p}-x_p$ is negative). This is possible for
all $p$ such that $\frac{l\log p}{2\sqrt p} < \frac 1 2$, i.e.~all but finitely
many $p$. So we cannot control the partial sums
$\sum_{p\leqslant N} \left(\frac{a_p}{2\sqrt p} - x_p\right)$ for a bounded
set of $N$, but then as $N\to \infty$, the sum can change by at most
$\frac{l\log p}{2\sqrt p}$ at each step. Moreover, once $N$ is sufficiently
large, the partial sum decreases (by at most $\frac{l\log p}{2\sqrt p}$)
whenever it is $>0$ and increases (by at most $\frac{l\log p}{2\sqrt p}$)
whenever it is $<0$. Thus the partial sums are bounded.
Write $A_N = \frac{\log N}{\sqrt N}\sum_{p\leqslant N} \frac{a_p}{2\sqrt p}$
and $B_N = \frac{\log N}{\sqrt N} \sum_{p\leqslant N} x_p$, both sums tacitly
taken over $\rho$-unramified primes. Then
\[
\left| A_N - B_N\right| \leqslant \frac{\log N}{\sqrt N} \left| \sum_{p\leqslant N} \left(\frac{a_p}{2\sqrt p} - x_p\right)\right| ,
\]
which converges to zero because the partial sums
$\sum_{p\leqslant N} \left(\frac{a_p}{2\sqrt p} - x_p\right)$ are bounded, and
$\frac{\log N}{\sqrt N} \to 0$. The proof isn't quite
complete, because we only know that
$\lim_{N\to \infty} \frac{\log N}{\sqrt N}\sum_{p\leqslant N} x_p = L$ when
$x_p$ is indexed by \emph{all} the rational primes, not just by the
$\rho$-unramified ones. We need to prove that $B_N$ converges to $L$.
Let $C_N = \frac{\log N}{\sqrt N} \sum_{p\leqslant N} x_p$, where here the sum
is taken with $x_p$ indexed by all primes. Write
$M_N = \pi^{-1}(\pi(N) - \pi_{\ram(\rho)}(N))$; then the number of
$\rho$-unramified primes $\leqslant N$ is the same as number of all primes
$\leqslant M_N$. It follows that
$B_N = \frac{\log N}{\sqrt N} \frac{\sqrt{M_N}}{\log{M_N}}C_{M_N}$, so to prove
$B_N \to L$, it suffices to prove that
$\frac{\log N}{\sqrt N} \frac{\sqrt{M_N}}{\log{M_N}} \to 1$. Convergence
$\frac{M_N}{N} \to 1$ follows from the prime number theorem, so we show that
this implies $\frac{\log N}{\log M_N} \to 1$. Since
$\frac{\log M_N}{\log N} = \log_N(M_N)$, we want to prove
that $\log_N(M_N) \to 1$. Write $M_N = N^{\alpha_N}$; then
$\frac{M_N}{N} = N^{\alpha_N - 1}$. Since $N\to \infty$, the only way for
$N^{\alpha_N - 1} \to 1$ is for $\alpha_N \to 1$,
i.e.~$\frac{\log N}{\log{M_N}} \to 1$.
When $L\ne \pm\infty$, this shows that the limit in question exists and is $L$.
When $L=\pm \infty$, this shows that the the sums in question diverge to $L$.
\end{proof}
\section{Galois representations with specified Sato--Tate distributions}
For $k\geqslant 1$, let
\[
U_k(\theta) = \tr\sym^k\smat{e^{i\theta}}{}{}{e^{- i \theta}} = \frac{\sin((k+1)\theta)}{\sin\theta} .
\]
Then $U_k(\cos^{-1} t)$ is the $k$-th Chebyshev polynomial of the 2nd kind.
Moreover, $\{1\}\cup\{U_k\}$ forms an orthonormal basis for
$L^2([0,\pi],\ST) = L^2(\SU(2)^\natural)$.
This section has two parts. First, for any reasonable measure $\mu$ on
$[0,\pi]$ invariant under the same ``flip'' automorphism as the Sato--Tate
measure, there is a sequence $(a_p)$ of integers satisfying the Hasse
bound $|a_p|\leqslant 2\sqrt p$, such that for
$\theta_p = \cos^{-1}\left(\frac{a_p}{2\sqrt p}\right)$, the discrepancy
$\D_N(\btheta,\mu)$ behaves like $\pi(N)^{-\alpha}$ for predetermined
$\alpha\in \left(0,\frac 1 2\right)$, while for any odd $k$, the strange
Dirichlet series $L_{U_k}(\btheta,s)$, which we will write as
$L(\sym^k \btheta,s)$, satisfies the Riemann hypothesis. In the second part of
this section, we associate Galois representations to these fake Satake
parameters.
\begin{definition}
Let $\mu = f(\theta)\, \dd \theta$ be an absolutely continuous probability
measure on $[0,\pi]$. If $f(\theta) \ll \sin(\theta)$ on $[0,\pi]$, then $\mu$
is a \emph{Sato--Tate compatible measure}.
\end{definition}
Recall that $\cos_\ast\mu = \frac{f(\cos^{-1} t)}{\sqrt{1-t^2}}\, \dd t$. So
the Radon--Nikodym derivative of $\cos_\ast\mu$ is bounded if and only if
$\frac{f(\cos^{-1} t)}{\sqrt{1-t^2}}$ is bounded. Plugging in
$t = \cos\theta$, we see that $\cos_\ast\mu$ has bounded Radon--Nikodym
derivative if and only if $\frac{f(\theta)}{\sin\theta}$ is bounded,
i.e.~$f(\theta) \ll \sin\theta$. So we could rephrase the definition of a
Sato--Tate compatible measure to be ``an absolutely continuous measure $\mu$
such that $\cos_\ast\mu$ has bounded Radon--Nikodym derivative.''
Since $\ST = \frac 2 \pi \sin^2\theta\, \dd\theta$ clearly satisfies this
definition, the Sato--Tate measure is itself Sato--Tate compatible.
If $\mu$ is Sato--Tate compatible, then $\cos_\ast\mu$
satisfies the hypotheses of Theorem \ref{thm:discrepancy-arbitrary}, so
there are ``$N^{-\alpha}$-decaying van der Corput sequences'' for
$\cos_\ast\mu$, and also that since $\cos\colon [0,\pi] \to [-1,1]$ is
strictly decreasing, we know that for any sequence $\bx$ on $[-1,1]$,
$\D_N(\bx,\cos_\ast\mu) \approx \D_N(\cos^{-1}\bx,\mu)$, with the
difference being $O(N^{-1})$. Finally, the
Radon--Nikodym derivative of $\mu$ (and also $\cos_\ast\mu$) is bounded , so
Lemma \ref{lem:disc-of-two-seq} applies to both $\mu$ and $\cos_\ast\mu$.
Recall that for deceasing functions
$\varphi_1,\varphi_2$, we write $\varphi_1(N) = \Theta(\varphi_2(N))$ if
there exists constants $0 < C_1 < C_2$ such that
$C_1 \varphi_2(N) \leqslant \varphi_1(N) \leqslant C_2 \varphi_2(N)$.
\begin{theorem}\label{thm:integral-a_p-alpha}
Let $\mu$ be a Sato--Tate compatible measure, and fix
$\alpha\in \left(0,\frac 1 3\right)$.
Then there exists a sequence of integers $a_p$ satisfying the Hasse bound,
such that if we set $\theta_p = \cos^{-1}\left(\frac{a_p}{2\sqrt p}\right)$,
then $\D_N(\btheta,\mu) = \Theta(\pi(N)^{-\alpha})$.
\end{theorem}
\begin{proof}
Apply Theorem \ref{thm:discrepancy-arbitrary} to find a sequence $\bx$ such
that $\D_N(\bx,\cos_\ast \mu) = \Theta(\pi(N)^{-\alpha})$. For each prime
$p$, there exists an integer $a_p$ such that $|a_p|\leqslant 2\sqrt p$ and
$\left| \frac{a_p}{2\sqrt p} - x_p\right| \leqslant \frac{1}{2\sqrt p}$. Let
$y_p = \frac{a_p}{2\sqrt p}$, and apply Corollary \ref{cor:close-seq-disc}.
We obtain
\[
\left| \D_N(\bx,\cos_\ast \mu) - \D_N(\by, \cos_\ast \mu)\right| \ll \pi(N)^{-\frac 1 3} ,
\]
which tells us that $\D_N(\by,\cos_\ast\mu) = \Theta(\pi(N)^{-\alpha})$.
Now let $\btheta = \cos^{-1}(\by)$. Apply Lemma \ref{lem:push-discrepancy} to
$\btheta = \cos^{-1}(\by)$, and we see that
$\D_N(\btheta,\mu) = \Theta(\pi(N)^{-\alpha})$.
\end{proof}
We can improve this example by controlling the behavior of the sums
$\sum_{p\leqslant N} U_k(\theta_p)$ for odd $k$. Let $\sigma$ be
the involution of $[0,\pi]$ given by $\sigma(\theta) = \pi-\theta$. Note that
$\sigma_\ast \ST = \ST$. Moreover, note that for any odd $k$,
$U_k\circ\sigma = - U_k$, so $\int U_k\, \dd\ST = 0$. Of course,
$\int U_k \,\dd\ST = 0$ for the reason that $U_k$ is the trace of a
non-trivial unitary representation, but we will directly use the ``oddness''
of $U_k$ in what follows.
\begin{theorem}\label{thm:int-flip-seq}
Let $\mu$ be a $\sigma$-invariant Sato--Tate compatible measure. Fix
$\alpha\in \left(0,\frac 1 3\right)$. Then there is a sequence of integers
$a_p$, satisfying the Hasse bound, such that for
$\theta_p =\cos^{-1}\left( \frac{a_p}{2\sqrt p}\right)$, we have
\begin{enumerate}
\item
$\D_N(\btheta,\mu) = \Theta(\pi(N)^{-\alpha})$.
\item
For all odd $k$,
$\left| \sum_{k\leqslant N} U_k(\theta_p)\right| \ll \pi(N)^{1/2}$.
\end{enumerate}
\end{theorem}
\begin{proof}
The basic ideas is as follows. Enumerate the primes
\[
p_1 = 2, q_1 = 3, p_2 = 5, q_2 = 7, p_3 = 11, q_3 = 13, \dots .
\]
Consider the measure $\left.\mu\right|_{[0,\pi/2)}$. This is supported on
$[0,\pi/2)$, but we extend it by zero to $[0,\pi]$. An argument
nearly identical to the proof of Theorem \ref{thm:integral-a_p-alpha} shows
that we can choose $a_{p_i}$ satisfying the Hasse bound so that
\[
\D_N\left( \left\{\theta_{p_i}\right\},\left.\mu\right|_{[0,\pi/2)}\right) = \Theta(N^{-\alpha}) .
\]
Since $\frac{a_{p_i}}{2\sqrt{p_i}}\in \frac{1}{2\sqrt{p_i}} \bZ$ and
$\frac{a_{q_i}}{2\sqrt{q_i}}\in \frac{1}{2\sqrt{q_i}}\bZ$, we cannot obtain
$\frac{a_{p_i}}{2\sqrt{p_i}} = - \frac{a_{q_i}}{2\sqrt{q_i}}$, but we can get
quite close to equality. That is, we can also choose the $a_{q_i}$ such that
$\frac{a_{q_i}}{2\sqrt{q_i}}\in [-1,0)$ and
$\left| \frac{a_{p_i}}{2\sqrt{p_i}} + \frac{a_{q_i}}{2\sqrt{q_i}}\right| \ll \frac{1}{\sqrt{p_i}}$.
Let $\bx$ be the sequence of the $\frac{a_{p_i}}{2\sqrt{p_i}}$ and $\by$
the corresponding sequence with the $q_i$-s. Then Lemma
\ref{lem:flip-discrepancy} with $\sigma(t) = -t$ tells us that the discrepancy
of $\by$ decays at the same rate as $-\by$, and then Corollary
\ref{cor:close-seq-disc} with $\alpha = \frac 1 2$ tells us that the
discrepancy of $-\by$ decays at the same rate (within $O(N^{-1/3})$) as the
discrepancy of $\bx$. Thus the discrepancies of both $\bx$ and $\by$ decay as
$\Theta(N^{-\alpha})$. Finally, Theorem \ref{thm:wreath-seq} tell us that
$\D_N(\bx\wr\by,\mu) = \Theta(N^{-\alpha})$.
The function $U_k(\cos^{-1} t)$ is an odd polynomial in $t$, so for
$t_1,t_2\in [-1,1]$,
\[
|U_k(\cos^{-1} t_1) + U_k(\cos^{-1} t_2)| = |U_k(\cos^{-1} t_1) - U_k(\cos^{-1}(-t_2))| \ll |t_1 - (-t_2)|.
\]
It follows that since
$\left|\frac{a_{p_i}}{2\sqrt{p_i}} - \left(- \frac{a_{q_i}}{2\sqrt{q_i}}\right)\right| \ll p_i^{-1/2}$,
then $|U_k(\theta_{p_i}) + U_k(\theta_{q_i})|\ll p_i^{-1/2}$.
We can then bound
\[
\left| \sum_{i\leqslant N} \left(U_k(\theta_{p_i}) + U_k(\theta_{q_i})\right)\right| \ll \sum_{p\leqslant N} p^{-1/2} \ll \pi(N)^{1/2} .
\]
\end{proof}
Note that this proof actually shows that for any $f\in C([0,\pi])$ such
that $f\circ \cos^{-1}$ is Lipschitz, and $f(\pi-\theta) = -f(\theta)$, the
estimate $\left| \sum_{p\leqslant N} f(\theta_p)\right| \ll \pi(N)^{1/2}$
holds.
\begin{theorem}\label{thm:bad-Galois}
Let $\mu$ be a Sato--Tate compatible $\sigma$-invariant measure on $[0,\pi]$.
Fix $\alpha\in \left(0,\frac 1 3\right)$ and a good residual representation
$\bar\rho\colon G_\bQ \to \GL_2(\bF_l)$. Then there exists a weight-$2$ lift
$\rho\colon G_\bQ \to \GL_2(\bZ_l)$ of $\bar\rho$ such that
\begin{enumerate}
\item
$\pi_{\ram(\rho)}(x) \ll \log(x)$.
\item
For each unramified prime $p$, $a_p = \tr\rho(\frob_p)\in \bZ$ and satisfies
the Hasse bound.
\item
If, for unramified $p$ we set
$\theta_p = \cos^{-1}\left(\frac{a_p}{2\sqrt p}\right)$, then
$\D_N(\btheta,\mu) = \Theta(\pi(N)^{-\alpha})$.
\item
For each odd $k$, the function $L(\sym^k \rho,s)$ satisfies the Riemann
hypothesis.
\end{enumerate}
\end{theorem}
\begin{proof}
Let $\bx$ be an $N^{-\alpha}$-decay van der Corput sequence for
$\cos_\ast \left.\mu\right|_{[0,\pi/2)}$, so that $\bx$ is contained in
$(0,1]$. Let $\by = -\bx$ (contained in $[-1,0)$), and put
$\bz = \bx\wr\by$, reindexed by the prime numbers. We have
$\D_N(\bz,\cos_\ast\mu) = \Theta(\pi(N)^{-\alpha})$ just as in the proof
of Theorem \ref{thm:int-flip-seq}. Set $h(x) = \log(x)$. By Theorem
\ref{thm:master-Galois}, there is a $\rho\colon G_\bQ \to \GL_2(\bZ_l)$ lifting
$\bar\rho$ such that $\pi_{\ram(\rho)}(x) \ll \log x$, the $\tr \rho(\frob_p)$
are integral, satisfy the Hasse bound, and
$\left| \frac{a_p}{2\sqrt p} - z_p\right| \leqslant \frac{l \log p}{2\sqrt p}$.
This implies, just as in the proof of Theorem \ref{thm:master-Galois}, that
the discrepancy of the sequence $\left\{\frac{a_p}{2\sqrt p}\right\}$ decays
as $\Theta(\pi(N)^{-\alpha})$ and by Lemma \ref{lem:push-discrepancy} with
$f(t) = \cos^{-1}(t)$, the
discrepancies of $\left\{\frac{a_p}{2\sqrt p}\right\}$ and $\{\theta_p\}$
decay at the same rate.
We've proved statements 1--3 in the theorem, which follow essentially for free
from Theorem \ref{thm:master-Galois} and its proof. All that remains is to prove
the Riemann hypothesis for odd symmetric powers. The proof of Theorem
\ref{thm:int-flip-seq} gives us an estimate
$\left| \sum_{p\leqslant N} U_k(\theta_p)\right| \ll N^{\frac 1 2+\epsilon}$,
and this combined with Theorem \ref{thm:AT->RH:gp} yields the result.
\end{proof}
This entire discussion works with absolutely continuous measure $\mu$. For
example, let $I$ be an arbitrarily small subinterval of $[0,\pi]$
(e.g.~$I = \left[\frac \pi 2 - \epsilon,\frac \pi 2 + \epsilon\right]$), let
$B_I(t)$ be a bump function for $I$, normalized to have total mass one. Then
Theorem \ref{thm:bad-Galois} gives Galois representations with empirical
Sato--Tate distribution converging at an arbitrarily slow rate to
$\mu_I = B_I(t)\, \dd t$. This is a strictly stronger result than
\cite[Th.~5.2]{pande-2011}. Moreover, the proof of Theorem
\ref{thm:int-flip-seq} shows that in fact for any $f\in C([0,\pi])$ with
$f\circ \cos^{-1}\in C^1([-1,1])$ and $f(\pi-\theta) = -f(\theta)$, the
Dirichlet series
$L_f(\rho,s) = \prod \left( 1 - f(\theta_p) p^{-s}\right)^{-1}$ satisfies the
Riemann hypothesis.