-
Notifications
You must be signed in to change notification settings - Fork 1
/
Copy pathgrowth_conditions.tex
357 lines (281 loc) · 12.3 KB
/
growth_conditions.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
\documentclass {amsart}
\usepackage{amsmath, amsthm, amsfonts, amssymb}
%\usepackage{tikz} \usetikzlibrary{positioning} % for commutative diagrams
%%%% Theorem-type MACROS
\theoremstyle{plain}
\newtheorem{proposition}{Proposition}
\newtheorem{theorem}[proposition]{Theorem}
\newtheorem{corollary}[proposition]{Corollary}
\newtheorem*{observe}{Observation}
\newtheorem{lemma}[proposition]{Lemma}
\theoremstyle{definition}
\newtheorem{definition}{Definition}
\newtheorem{exercise}{Exercise}
\newtheorem{axiom}{Axiom}
\newtheorem{declaration}[axiom]{Declaration}
\theoremstyle{remark}
\newtheorem{example}{Example}
\newtheorem*{remark}{Remark}
\newtheorem*{note}{Note}
\newtheorem*{assumption}{Background Assumption}
%%%% Character MACROS
\newcommand{\defeq}{\stackrel{\mathrm{def}}{\, = \,}}
\newcommand{\bN}{{\mathbb{N}}}
\newcommand{\bZ}{{\mathbb{Z}}}
\newcommand{\Z}{{\mathbb{Z}}}
\newcommand{\bR}{{\mathbb{R}}}
\newcommand{\bC}{{\mathbb{C}}}
\newcommand{\bQ}{{\mathbb{Q}}}
\newcommand{\bF}{{\mathbb{F}}}
\newcommand{\bA}{{\mathbb{A}}}
\newcommand{\bP}{{\mathbb{P}}}
\newcommand{\h}{{\mathcal{H}}}
%%%%%%%%%% END OF MACROS %%%%%%%%%%%%
\begin{document}
\title{A Condition for Being Holomorphic at All Cusps}
\author{Modular Form Study Group}
\date{August 2014, modified \today}
\maketitle
%%%%%%%%%%%%%%%%%%%%
\section {Introduction}
In this document we elaborate on Proposition~1.2.4 of Diamond, Shurman (and
the companion Exercise 1.2.6).
In particular we prove that if the Fourier coefficients of a candidate modular form (for some modular group~$\Gamma$) has polynomial bounds, then it is holomorphic at all of the cusps. In other words, it is a modular form for~$\Gamma$.
%%%%%%%%%%%%%%%%%%%%
\section {Removable singularities at infinity}
The key to the above result is to bound the growth of our candidate function as $z$
approaches the cusps. By transforming the cusp to infinity, we prove our result by
bounding the growth of the transformed function $f$ as $z$ goes to infinity. In this section we show that if $f$ is bounded
by an expression of the form $A y^r$, then $f$ is indeed holomorphic at infinity. In later sections
we will show that the transformed candidate has such growth for all cusps.
Let $f(z)$ be a periodic function on~$\h$ (or some region of $\bC$ containing all $z = x + yi$
with $y$ sufficiently large). Assume the period is a positive real number~$N$.
Then $f(z)$ has a Fourier expansion (Ahlfors Chapter 7, Section 1). In other words, $f(z) = g(q(z))$
where $q (z) = e^{2 \pi i z / N}$, and where
$$
g(q) = \sum_{n\in\bZ} a_n q^n
$$
is the Fourier expansion of~$f$.
The function $f$ is said to be holomorphic at infinity if and only if $g(q)$ is holomorphic
at zero. This occurs if and only if $q g(q)$ approaches zero as $q$ approaches zero.
In other words, $f$ is holomorphic at infinity if and only if $q(z) f(z)$ approaches zero as $z$ approaches infinity.
(See Theorem~7 of Ahlfors Chapter 4, Section 3).
In what follows we will consider functions $f$ such that for $z = x + yi$ with $y \ge B$ (where $B$ is some fixed bound),
$$
|f(z)| \le A y^t
$$
for some constant~$A$ and nonnegative integer~$t$.
In this case
$$
|f(z) q(z)| \le A \frac{y^t} { e^{ 2 \pi y / N}}.
$$
The limit of ${y^t} / { e^{ 2 \pi y / N}}$ is zero as $y$ goes to infinity (L'Hopital's Rule). So we conclude that
\emph{since $f(z)$ is so bounded by $A y^t$, then it must be holomorphic at infinity}.
%%%%
\section {Comparing $\sum f(n)$ to $\int f(x) dx$}
We want to begin our investigation by translating a growth condition on Fourier coefficients to
a growth condition of $f$ near its cusps. This can be done via the familiar
trick of comparing a sum $\sum f(n)$ to the associated integral~$\int f(x) dx$.
We will describe how to get bounds in the case where $f$ has a single maximum.
Suppose that $f(x)$ is a nonnegative continuous function on $[0, \infty)$ that is
increasing on $[0, m]$ and decreasing on~$[m, \infty)$. Assume also
that
$$
\int_{0}^\infty f(x) dx \; < \; \infty.
$$
Then we have the following upper bound:
$$
\sum_{n=0}^\infty f(n) \le f(m) + \int_{0}^\infty f(x) dx.
$$
To establish this bound, let $M$ be a positive integer chosen so that $m$ is in the interval $[M - 1, M]$.
If $0 \le i \le M -2$ is an integer, consider the rectangle $A_i$ with base $[i, i + 1]$
and height~$f(i)$.
If $i \ge M + 1$ is an integer, then consider the rectangle $B_i$ with base $[i - 1, i]$
and height $f(i)$.
Finally, let $C$ be the rectangle with base $[M, M+1]$ and height the minimum
of $f(M)$ and $f(M+1)$.
Note that the area of the union of these rectangles is equal to the integral of a step
function that is bounded above by $f$. So the sum of the areas of these rectangles
is bounded above by~$\int_{0}^\infty f(x) dx$. This gives us
$$
\sum_{n \in I} f(n) \le \int_{0}^\infty f(x) dx
$$
where $I$ is $\bN$ with one element (either $M$ or $M+1$) removed.
The result follows.
%%%
\section {Example: bounding $\sum n^r e^{-\lambda n}$}
We are particularly interested in the function $f(x) = x^r e^{-\lambda x}$ where $r \in \bN$
and $\lambda$ is a positive real constant. Note that for large $x$, $f(x) \le e^{-\lambda x / 2}$
so,
$$
\int_{0}^\infty f(x) dx \; < \; \infty.
$$
Also, note that $f$ is increasing on $[0, r/\lambda]$ and decreasing on~$[r/\lambda, \infty)$.
So the above result gives the estimate
$$
\sum_{n=0}^\infty n^r e^{-\lambda n} \le (r/\lambda)^r e^{-r} + \int_{0}^\infty x^r e^{-\lambda x} dx.
$$
With a simple change of variables of the integral ($u = \lambda x$), we can rewrite this
inequality as
$$
\sum_{n=0}^\infty n^r e^{-\lambda n} \le \frac{C_1}{\lambda^r} + \frac{C_2}{\lambda^{r+1}}
$$
where $C_1$ and $C_2$ are positive constants dependent on~$r$, but not on~$k$.
If we fix a bound $B >0$, then we can split the inequality into two simpler inequalities
(where $C_1, C_2$ are possibly different positive constants that depend on $B$ and $r$).
If $\lambda \ge B$ we have
$$
\sum_{n=0}^\infty n^r e^{-\lambda n} \le \frac{C_1}{\lambda^r} .
$$
If $0 < \lambda \le B$ we have
$$
\sum_{n=0}^\infty n^r e^{-\lambda n} \le \frac{C_2}{\lambda^{r+1}}.
$$
Observe that in the second case, the case that we will be needed in what follows,
the integral gives the estimate of the discrete sum.
\section{Alternative proof of bound on $\sum n^r e^{-\lambda n}$}
\label{sec:altern-proof-bound}
The sum $\sum n^r e^{-\lambda n}$ can be evaluated exactly, which leads to an alternative argument for the bound shown in the previous section. Let $L$ denote the differential operator $x \frac{d}{dx}$. Let $f(x) = \dfrac{1}{1-x}$, and let $f_r = L^r f$. We observe that
\[
\sum n^r e^{-\lambda n} = f_r(e^{-\lambda}).
\]
In principle, one can thus compute each sum explicitly; however, there is no obvious formula for the $f_r$. But we can come up with a sufficiently good description to bound the summation.
\begin{lemma}
If $r \geq 1$, then there exists $p_r(x) \in \Z[x]$ of degree $\leq r$ such that $p_r(0) = 0$, $p'_r(0) = 1$, and
\[
f_r(x) = \frac{p_r(x)}{(1-x)^{r+1}}.
\]
\end{lemma}
\begin{proof}
Set $p_0(x) = 1$. Via the product rule, one checks that if $p_r(x)$ is defined by
\[
p_r(x) = (1-x)^{r+1} f_r(x),
\]
then the recursion
\[
p_r = x(1-x)p'_{r-1} + rxp_{r-1}
\]
holds for $r \geq 1$. The claim follows from an induction argument.
\end{proof}
The induction can also show that $p_r$ is monic of degree $r$, but this conclusion is not necessary for the below.
\begin{proposition}
For $\lambda$ a sufficiently small positive number, there exists a constant $C$ such that
\[
\sum n^r e^{-\lambda n} < \frac{C}{\lambda^{r+1}}.
\]
\end{proposition}
\begin{proof}
We use the fact that the summation equals $f_r(e^{-\lambda})$. Let $p_r(x)$ be as in the last lemma. Note that $p_r(x) = x + \cdots$, where the dots indicate higher order terms. Then $p_r(e^{-\lambda}) < k e^{-\lambda} < k$ for some $k > 0$ and sufficiently small $\lambda$. (Here the $k$ depends on $p_r$, and hence on $r$.) By Taylor's Theorem, $1-e^{-\lambda} > \ell \lambda$ for some $\ell > 0$ and small enough $\lambda$. The claim now follows.
\end{proof}
%%%
\section {Application to $q$-expansions}
Suppose that $f \colon \h \to \bC$ is a holomorphic function with real period~$N$.
Suppose that its $q$-expansion has the form
$$
\sum_{n=0}^\infty a_n q^n
$$
where $q = e^{2 \pi i z / N}$ and where for positive~$n$
$$
|a_n| \le C n^r
$$
for some real constant~$C$ and some nonnegative integer~$r$.
From this we get the estimate
$$
|f(z)| \le |a_0| + C \sum_{n=0}^\infty n^r \left|q^n \right| \le
|a_0| + C \sum_{n=0}^\infty n^r e^{- (2 \pi y / N) n}
$$
where $z = x + y i$.
So if we assume $z$ is such that $y \le B$ for some bound~$B$, the estimate of the
previous section can be applied. It gives us the estimate
$$
|f(z)| \le \frac{C_0}{y^{r+1}}.
$$
for some constant $C_0$ independent of~$z$.
(The constant $|a_0|$ gets absorbed by choosing $C_0$ sufficiently large.
So $C_0$ does depends on $C$, $r$, $B$, and $|a_0|$).
%%%
\section {Behavior of Growth Conditions at Cusps under Transformation}
Let $f: \h \to \bC$ be a holomorphic function on the upper half plane that is
invariant under the weight $k$ action of some modular group~$\Gamma$.
In other words, $f [\alpha]_k = f$ for all $\alpha \in\Gamma$.
Let $a/c$ be a cusp
of~$f$.
Choose $\alpha \in SL(2, \bZ)$ so that $\alpha$ has form
$$ \alpha = \begin{bmatrix}
a & b \\
c & d
\end{bmatrix} \in SL (2, \bZ).$$
(This is possible since $a$ and $c$ are relatively prime).
Then $f [\alpha]_k$ is not necessarily equal to~$f$, but it is periodic with
some positive integer period. Furthermore, by definition $f$ is holomorphic at
the cusp~$a/c$ if and only if $f [\alpha]_k$ is holomorphic at infinity.
We want to investigate how growth conditions at cusps translates into growth conditions
of $f [\alpha]_k$ at infinity. So assume that there is a bound $B$, and constants $C_0 \ge 0$
and $s \in \bN$ so that
$$
|f(z)| \le \frac{C_0}{y^{s}}
$$
for all $z = x + yi$ with $y \le B$.
As we will see, this translates into bounds for $f [\alpha]_k$.
First we consider the behavior of $f(\alpha(z))$ near~$\infty$. Write $z = x + y i$
and $\alpha(z) = x' + y' i$.
By continuity, there is a bound $B_0 > 0$ such that if $y \ge B_0$ then $y' \le B$.
If $y \ge B_0$ then
$$
|f(\alpha(z))| \; \le \; \frac{C_0}{(y')^{s}}.
$$
However, we know that $y' = y / |c z + d|^2$, so
$$
|f(\alpha(z))| \; \le \; C_0 \frac{ |c z + d|^{2 s}}{(y)^{s}}.
$$
By definition, $f [\alpha]_k (z) = f(\alpha (z)) (c z + d)^{-k}$.
Thus
$$
\left| f [\alpha]_k (z) \right| \; \le \; C_0 \frac{ |c z + d|^{2 s - k}}{(y)^{s}}.
$$
Since $ f [\alpha]_k $ is periodic with rea period, we can assume~$|x|$
is bounded and so, if $y \ge B_0$,
$$|c| |y| \le |c z + d| \le |c x + d| + |c y| \le D |y|$$
for some constant~$D$ independent of~$y$.
Thus, if $y \ge B_0$,
$$
\left| f [\alpha]_k (z) \right| \; \le \; A y ^{s-k}
$$
for some constant~$A$ (independent of~$y$).
%%%
\section {Proof of main theorem}
We start with a candidate modular form $f$
of weight $k$ for a modular group~$\Gamma$.
In other words,
$f: \h \to \bC$ is a holomorphic function on the upper half plane
such that $f [\alpha]_k = f$ for all $\alpha \in\Gamma$.
Suppose that the Fourier coefficients $a_n$ of $f$ (with $n\ge 1$) have
bounds of the form.
$$
|a_n| \le C n^r.
$$
We saw about how this translates into bounds for~$f$ of the form
$$
|f(z)| \le \frac{C_0}{y^{r+1}}
$$
for $y$ sufficiently small. So this behavior applies when we are sufficiently
close to a cups $a/c$. If $\alpha \in SL(2, \bZ)$ maps $a/c$ to $\infty$
we saw that $f [\alpha]_k$ is bounded (for $y$ sufficiently large) by bounds of the form
$$
|f [\alpha]_k (z)| \le A y^t
$$
where (without loss of generality) we assume $t$ is a nonnegative integer.
We saw above that this implies, in turn, that $f [\alpha]_k$ has a removable singularity at infinity.
In other words, $f$ is holomorphic at the cusp $a/c$.
%%%
\section {Other Notes}
This proof follows the outline on page 22, but of course with slightly different notation.
I believe the proof on page 22 - 23 needs a correction.
I believe on page 22 the conclusion of 1.2.6 (a) should be
$$
|f(\tau)| \le C_0 + C/ y^{r+1}
$$
(not $y^r$) as $y$ approaches~$0$ (not as $y$ approaches $\infty$).
Also I would use the fact that exponentials dominate polynomials (not that polynomials dominate logarithms).
The book promises a converse in Section 5.9.
\end{document}