Contact
CoCalc Logo Icon
StoreFeaturesDocsShareSupportNewsAboutSign UpSign In
| Download

William Stein -- Talk for Mathematics is a long conversation: a celebration of Barry Mazur

Views: 2342
1
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
2
\providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
3
% \MRhref is called by the amsart/book/proc definition of \MR.
4
\providecommand{\MRhref}[2]{%
5
\href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
6
}
7
\providecommand{\href}[2]{#2}
8
\begin{thebibliography}{10}
9
10
\bibitem{Note01}
11
{\bf How not to factor the numerator of a Bernoulli number:} \par As mentioned
12
in Chapter~\ref {ch:envision}, the coefficient $B_k$ of the linear term of
13
the polynomial $$ S_k(n)= 1^k+2^k+3^k+\dots +(n-1)^k $$ is (up to sign) the
14
$k$-th {\bf Bernoulli number}. These numbers are rational numbers and,
15
putting them in lowest terms, their numerators play a role in certain
16
problems, and their denominators in others. (This is an amazing story, which
17
we can't go into here!) \par One of us (Barry Mazur) in the recent article
18
{\em How can we construct abelian Galois extensions of basic number fields?}
19
(see \url
20
{http://www.ams.org/journals/bull/2011-48-02/S0273-0979-2011-01326-X/}) found
21
himself dealing (for various reasons) with the fraction $-B_{200}/400$, where
22
$B_{200}$ is the two-hundredth Bernoulli number. The numerator of this
23
fraction is quite large: it is---hold your breath---\ $$389\cdot 691\cdot
24
5370056528687 \ \ \ \ \text {times this 204-digit number:} $$\vskip 10pt
25
$N:=3452690329392158031464109281736967404068448156842396721012$\newline
26
$\mbox {}\,\,\,\qquad
27
9920642145194459192569415445652760676623601087497272415557$\newline $\mbox
28
{}\,\,\,\qquad
29
0842527652727868776362959519620872735612200601036506871681$\newline $\mbox
30
{}\,\,\,\qquad 124610986596878180738901486527$ \vskip 10pt and he {\it
31
incorrectly asserted} that it was prime. Happily, Bartosz Naskr\c {e}cki
32
spotted this error: our $204$-digit $N$ is {\it not} prime. \par How did he
33
know this? By using the most basic test in the repertoire of tests that we
34
have available to check to see whether a number is prime: we'll call it the
35
``{\bf Fermat $2$-test.}'' We'll first give a general explanation of this
36
type of test before we show how $N$ {\it fails the Fermat $2$-test.} \par
37
\par The starting idea behind this test is the famous result known as {\em
38
Fermat's Little Theorem} where the ``little'' is meant to alliteratively
39
distinguish it from you-know-what. \begin {theorem}[Fermat's Little Theorem]
40
If $p$ is a prime number, and $a$ is any number relatively prime to $p$ then
41
$a^{p-1} - 1$ is divisible by $p$. \end {theorem} A good exercise is to try
42
to prove this, and a one-word hint that might lead you to one of the many
43
proofs of it is {\em induction}.\footnote {Here's the proof: $$(N+1)^p \equiv
44
N^p +1 \equiv (N+1)\ \mod p,$$ where the first equality is the binomial
45
theorem and the second equality is induction.} \par Now we are going to use
46
this as a criterion, by---in effect---restating it in what logicians would
47
call its {\it contrapositive}: \begin {theorem}[The Fermat $a$-test] If $M$
48
is a positive integer, and $a$ is any number relatively prime to $M$ such
49
that $a^{M-1} - 1$ is {\it not} divisible by $M$, then $M$ is {\it not} a
50
prime. \end {theorem} \par Well, Naskr\c {e}cki computed $2^{N-1}-1$ (for the
51
$204$-digit $N$ above) and saw that it is {\it not} divisible\footnote {The
52
number $2^{N-1}-1$ has a residue after division by $N$ of\\
53
$3334581100595953025153969739282790317394606677381970645616725285996925$\newline
54
$
55
6610000568292727335792620957159782739813115005451450864072425835484898$\newline
56
$ 565112763692970799269335402819507605691622173717318335512037457$.} by $N$.
57
Ergo, our $N$ fails the Fermat $2$-test so is {\it not} prime. \par But then,
58
given that it is so ``easy'' to see that $N$ is not prime, a natural question
59
to ask is: what, in fact, is its prime factorization? This---it turns
60
out---isn't so easy to determine; Naskr\c {e}cki devoted $24$ hours of
61
computer time setting standard factorization algorithms on the task, and that
62
was not sufficient time to resolve the issue. The factorization of the
63
numerators of general Bernoulli numbers is the subject of a very interesting
64
website run by Samuel Wagstaff (\url
65
{http://homes.cerias.purdue.edu/~ssw/bernoulli}). Linked to this web page one
66
finds (\url {http://homes.cerias.purdue.edu/~ssw/bernoulli/composite}) which
67
gives a list of composite numbers whose factorizations have resisted all
68
attempts to date. The two-hundredth Bernoulli number is 12th on the list.
69
\par The page \url
70
{http://en.wikipedia.org/wiki/Integer_factorization_records} lists record
71
challenge factorization, and one challenge that was completed in 2009
72
involves a difficult-to-factor number with 232 digits; its factorization was
73
completed by a large team of researchers and took around 2000 years of CPU
74
time. This convinced us that with sufficient motivation it would be possible
75
to factor $N$, and so we asked some leaders in the field to try. They
76
succeeded! \begin {quote}\sf Factorisation of B200 \par by Bill Hart on 4 Aug
77
05, 2012 at 07:24pm \par We are happy to announce the factorization of the
78
numerator of the 200th Bernoulli number: \par \begin {align*} N &= 389 \cdot
79
691 \cdot 5370056528687 \cdot c_{204}\\ c_{204} &= p_{90} \cdot p_{115}\\
80
p_{90} &= 149474329044343594528784250333645983079497454292\\ &=
81
838248852612270757617561057674257880592603\\ p_{115} &=
82
230988849487852221315416645031371036732923661613\\ &=
83
619208811597595398791184043153272314198502348476\\ &= 2629703896050377709
84
\end {align*} \par The factorization of the 204-digit composite was made
85
possible with the help of many people: \begin {itemize} \item William Stein
86
and Barry Mazur challenged us to factor this number. \item Sam Wagstaff
87
maintains a table of factorizations of numerators of Bernoulli numbers at
88
\url {http://homes.cerias.purdue.edu/~ssw/bernoulli/bnum}. According to this
89
table, the 200th Bernoulli number is the 2nd smallest index with unfactored
90
numerator (the first being the 188th Bernoulli number). \item Cyril Bouvier
91
tried to factor the c204 by ECM up to 60-digit level, using the TALC cluster
92
at Inria Nancy - Grand Est. \item {\sf yoyo@home} tried to factor the c204 by
93
ECM up to 65-digit level, using the help of many volunteers of the
94
distributed computing platform \url {http://www.rechenkraft.net/yoyo/}. After
95
ECM was unsuccessful, we decided to factor the c204 by GNFS. \item Many
96
people at the Mersenne forum helped for the polynomial selection. The best
97
polynomial was found by Shi Bai, using his implementation of Kleinjung's
98
algorithm in CADO-NFS: \url
99
{http://www.mersenneforum.org/showthread.php?p=298264##post298264}. Sieving
100
was performed by many volunteers using {\sf NFS@home}, thanks to Greg
101
Childers. See \url {http://escatter11.fullerton.edu/nfs} for more details
102
about {\sf NFS@home}. This factorization showed that such a distributed
103
effort might be feasible for a new record GNFS factorization, in particular
104
for the polynomial selection. This was the largest GNFS factorization
105
performed by {\sf NFS@home} to date, the second largest being $2^{1040}+1$ at
106
183.7 digits. \item Two independent runs of the filtering and linear algebra
107
were done: one by Greg Childers with msieve (\url
108
{http://www.boo.net/~jasonp/qs.html}) using a 48-core cluster made available
109
by Bill Hart, and one by Emmanuel Thom\'{e} and Paul Zimmermann with CADO-NFS
110
(\url {http://cado-nfs.gforge.inria.fr/}), using the Grid 5000 platform.
111
\item The first linear algebra run to complete was the one with CADO-NFS,
112
thus we decided to stop the other run. \end {itemize} \par Bill Hart \end
113
{quote} \par We verify the factorization above in SageMath as follows: \\
114
\par {\sf sage: p90 = 1494743290443435945287842503336459830794974542928382\\
115
48852612270757617561057674257880592603\\ sage: p115 =
116
230988849487852221315416645031371036732923661613619\\
117
2088115975953987911840431532723141985023484762629703896050377709\\ sage: c204
118
= p90 * p115\\ sage: 389 * 691 * 5370056528687 * c204 ==
119
-numerator(bernoulli(200))\\ True\\ sage: is\_prime(p90), is\_prime(p115),
120
is\_prime(c204)\\ (True, True, False) }.
121
122
\bibitem{Note02}
123
\label {bibnote:factor} Given an integer $n$, there are many algorithms
124
available for trying to write $n$ as a product of prime numbers. First we can
125
apply {\em trial division}, where we simply divide $n$ by each prime $2, 3,
126
5, 7, 11, 13, \ldots $ in turn, and see what small prime factors we find (up
127
to a few digits). After using this method to eliminate as many primes as we
128
have patience to eliminate, we typically next turn to a technique called {\em
129
Lenstra's elliptic curve method}, which allows us to check $n$ for
130
divisibility by bigger primes (e.g., around 10--15 digits). Once we've
131
exhausted our patience using the elliptic curve method, we would next hit our
132
number with something called the {\em quadratic sieve}, which works well for
133
factoring numbers of the form $n=pq$, with $p$ and $q$ primes of roughly
134
equal size, and $n$ having less than 100 digits (say, though the 100 depends
135
greatly on the implementation). All of the above algorithms---and then
136
some---are implemented in SageMath, and used by default when you type {\tt
137
factor(n)} into SageMath. Try typing {\tt factor(some number, verbose=8)} to
138
see for yourself. \par If the quadratic sieve fails, a final recourse is to
139
run the {\em number field sieve} algorithm, possibly on a supercomputer. To
140
give a sense of how powerful (or powerless, depending on perspective!) the
141
number field sieve is, a record-setting factorization of a general number
142
using this algorithm is the factorization of a 232 digit number called
143
RSA-768 (see \url {https://eprint.iacr.org/2010/006.pdf}): \par $ n =
144
12301866845301177551304949583849627207728535695953347921973224521\\
145
517264005072636575187452021997864693899564749427740638459251925573263\\
146
034537315482685079170261221429134616704292143116022212404792747377940\\
147
80665351419597459856902143413 $ \par \noindent {}which factors as $pq$,
148
where\par \noindent {} $ p =
149
334780716989568987860441698482126908177047949837137685689124313889\\
150
82883793878002287614711652531743087737814467999489 $ \par \noindent {}and\par
151
\noindent {} $ q =
152
367460436667995904282446337996279526322791581643430876426760322838\\
153
15739666511279233373417143396810270092798736308917. $ \par \noindent {}We
154
encourage you to try to factor $n$ in SageMath, and see that it fails.
155
SageMath does not yet include an implementation of the number field sieve
156
algorithm, though there are some free implementations currently available
157
(see \url {http://www.boo.net/~jasonp/qs.html}).
158
159
\bibitem{Note03}
160
We can use SageMath (at \url {http://sagemath.org}) to quickly compute the
161
``hefty number'' $p = 2^{43,112,609}-1$. Simply type {\tt p = 2\^{}43112609 -
162
1} to instantly compute $p$. In what sense have we {\em computed} $p$?
163
Internally, $p$ is now stored in base $2$ in the computer's memory; given the
164
special form of $p$ it is not surprising that it took little time to compute.
165
Much more challenging is to compute all the base 10 digits of $p$, which
166
takes a few seconds: {\tt d = str(p)}. Now type {\tt d[-50:]} to see the last
167
50 digits of $p$. To compute the sum $58416637$ of the digits of $p$ type
168
{\tt sum(p.digits())}.
169
170
\bibitem{Note04}
171
In contrast to the situation with factorization, testing integers of this size
172
(e.g., the primes $p$ and $q$) for primality is relatively easy. There are
173
fast algorithms that can tell whether or not any random thousand digit number
174
is prime in a fraction of second. Try for yourself using the SageMath command
175
{\tt is\_prime}. For example, if $p$ and $q$ are as in endnote~\ref
176
{bibnote:factor}, then {\tt is\_prime(p)} and {\tt is\_prime(q)} quickly
177
output True and {\tt is\_prime(p*q)} outputs False. However, if you type {\tt
178
factor(p*q, verbose=8)} you can watch as SageMath tries forever and fails to
179
factor $pq$.
180
181
\bibitem{Note05}
182
In Sage, the function {\tt prime\_range} enumerates primes in a range. For
183
example, {\tt prime\_range(50)} outputs the primes up to $50$ and {\tt
184
prime\_range(50,100)} outputs the primes between $50$ and $100$. Typing {\tt
185
prime\_range(10\^{}8)} in SageMath enumerates the primes up to a hundred
186
million in around a second. You can also enumerate primes up to a billion by
187
typing {\tt v=prime\_range(10\^{}9)}, but this will use a large amount of
188
memory, so be careful not to crash your computer if you try this. You can see
189
that there are $\pi (10^9) = 50{,}847{,}534$ primes up to a billion by then
190
typing {\tt len(v)}. You can also compute $\pi (10^9)$ directly, without
191
enumerating all primes, using the command {\tt prime\_pi(10\^{}9)}. This is
192
much faster since it uses some clever counting tricks to find the number of
193
primes without actually listing them all. \par In Chapter~\ref
194
{sec:tinkering} we tinkered with the staircase of primes by first counting
195
both primes and prime powers. There are comparatively few prime powers that
196
are not prime. Up to $10^8$, only $1{,}405$ of the $5{,}762{,}860$ prime
197
powers are not themselves primes. To see this, first enter {\tt
198
a~=~prime\_pi(10\^{}8); pp~=~len(prime\_powers(10\^{}8))}. Typing {\tt
199
(a,~pp,~pp-a)} then outputs the triple {\tt (5761455, 5762860, 1405)}.
200
201
\bibitem{Note06}
202
Hardy and Littlewood give a nice conjectural answer to such questions about
203
gaps between primes. See Problem {\bf A8} of Guy's book {\em Unsolved
204
Problems in Number Theory} (2004). Note that Guy's book discusses counting
205
the number $P_k(X)$ of pairs of primes up to $X$ that differ by a fixed even
206
number $k$; we have $P_k(X)\geq \Gap _k(X)$, since for $P_k(X)$ there is no
207
requirement that the pairs of primes be consecutive.
208
209
\bibitem{Note07}
210
If $f(x)$ and $g(x)$ are real-valued functions of a real variable $x$ such that
211
for any $\epsilon >0$ both of them take their values between $ x^{1-\epsilon
212
}$ and $ x^{1+\epsilon }$ for $x$ sufficiently large, then say that $f(x)$
213
and $g(x)$ are {\bf good approximations of one another} if, for any positive
214
$\epsilon $ the absolute value of their difference is less than $ x^{{1\over
215
2}+\epsilon }$ for $x$ sufficiently large. The functions $\Li (X)$ and $R(X)$
216
are good approximations of one another.
217
218
\bibitem{Note08}
219
This computation of $\pi (X)$ was done by David J. Platt in 2012, and is the
220
largest value of $\pi (X)$ ever computed. See \url
221
{http://arxiv.org/abs/1203.5712} for more details.
222
223
\bibitem{Note09}
224
In fact, the Riemann Hypothesis is equivalent to $|\Li (X) - \pi (X)| \leq
225
\sqrt {X}\log (X)$ for all $X\geq 2.01$. See Section 1.4.1 of Crandall and
226
Pomerance's book {\em Prime Numbers: A Computational Perspective}.
227
228
\bibitem{Note10}
229
For a proof of this here's a hint. Compute the difference between the
230
derivatives of $\Li (x)$ and of $x/\log x$. The answer is $1/\log ^2(x)$. So
231
you must show that the ratio of $\int _2^X dx/\log ^2(x)$ to $\Li (x)= \int
232
_2^Xdx/\log (x)$ tends to zero as~$x$ goes to infinity, and this is a good
233
calculus exercise.
234
235
\bibitem{Note11}
236
See \url {http://www.maths.tcd.ie/pub/HistMath/People/Riemann/Zeta/} for the
237
original German version and an English translation.
238
239
\bibitem{Note12}
240
We have $$\psi (X) =\sum _{p^n \le X}\log p$$ where the summation is over prime
241
powers $p^n$ that are $\le X$.
242
243
\bibitem{Note13}
244
See \url {http://en.wikipedia.org/wiki/Fast_Fourier_transform}.
245
246
\bibitem{Note14}
247
\url {http://en.wikipedia.org/wiki/Distribution_\%28mathematics\%29} contains a
248
more formal definition and treatment of distributions. Here is Schwartz's
249
explanation for his choice of the word {\it distribution:} \begin
250
{quote}``Why did we choose the name distribution? Because, if $\mu $ is a
251
measure, i.e., a particular kind of distribution, it can be considered as a
252
distribution of electric charges in the universe. Distributions give more
253
general types of electric charges, for example dipoles and magnetic
254
distributions. If we consider the dipole placed at the point $a$ having
255
magnetic moment $M$, we easily see that it is defined by the distribution
256
$-D_M \delta _{(a)}$. These objects occur in physics. Deny's thesis, which he
257
defended shortly after, introduced electric distributions of finite energy,
258
the only ones which really occur in practice; these objects really are
259
distributions, and do not correspond to measures. Thus, distributions have
260
two very different aspects: they are a generalization of the notion of
261
function, and a generalization of the notion of distribution of electric
262
charges in space. [...] Both these interpretations of distributions are
263
currently used.''\end {quote}.
264
265
\bibitem{Note15}
266
David Mumford suggested that we offer the following paragraph from\\ \url
267
{http://en.wikipedia.org/wiki/Dirac_delta_function}\\ on the Dirac delta
268
function: \begin {quote} An infinitesimal formula for an infinitely tall,
269
unit impulse delta function (infinitesimal version of Cauchy distribution)
270
explicitly appears in an 1827 text of Augustin Louis Cauchy. Sim\'{e}on Denis
271
Poisson considered the issue in connection with the study of wave propagation
272
as did Gustav Kirchhoff somewhat later. Kirchhoff and Hermann von Helmholtz
273
also introduced the unit impulse as a limit of Gaussians, which also
274
corresponded to Lord Kelvin's notion of a point heat source. At the end of
275
the 19th century, Oliver Heaviside used formal Fourier series to manipulate
276
the unit impulse. The Dirac delta function as such was introduced as a
277
``convenient notation'' by Paul Dirac in his influential 1930 book {\em
278
Principles of Quantum Mechanics}. He called it the ``delta function'' since
279
he used it as a continuous analogue of the discrete Kronecker delta. \end
280
{quote}.
281
282
\bibitem{Note16}
283
As discussed in\\ \url
284
{http://en.wikipedia.org/wiki/Distribution_\%28mathematics\%29},
285
``generalized functions'' were introduced by Sergei Sobolev in the 1930s,
286
then later independently introduced in the late 1940s by Laurent Schwartz,
287
who developed a comprehensive theory of distributions.
288
289
\bibitem{Note17}
290
If the \RH {} holds they are precisely the {\it imaginary parts of the
291
``nontrivial'' zeroes of the Riemann zeta function}.
292
293
\bibitem{Note18}
294
{\it The construction of $ \Phi (t)$ from $\psi (X)$:} \par Succinctly, for
295
positive real numbers $t$, $$\Phi (t): \ = \ e^{- t/2}\Psi '(t),$$ where
296
$\Psi (t) = \psi (e^t)$ (see Figure~\ref {fig:psi_200}), and $\Psi '$ is the
297
derivative of $\Psi (t)$, viewed as a distribution. We extend this function
298
to all real arguments $t$ by requiring $\Phi (t)$ to be an even function of
299
$t$, i.e., $\Phi (-t)= \Phi (t)$. But, to review this at a more leisurely
300
pace, \begin {enumerate} \par \item Distort the $X$-axis of our staircase by
301
replacing the variable $X$ by $e^t$ to get the function $$\Psi (t):= \psi
302
(e^t).$$ No harm is done by this for we can retrieve our original $\psi (X)$
303
as $$\psi (X) = \Psi (\log (X)).$$ Our distorted staircase has risers at ($0$
304
and) all positive integral multiples of logs of prime numbers. \par \par \par
305
\par \item Now we'll do something that might seem a bit more brutal: {\it
306
take the derivative of this distorted staircase $\Psi (t)$.} This derivative
307
$\Psi '(t)$ is a {\it generalized} function with support at all nonnegative
308
integral multiples of logs of prime numbers. \par \vskip 20pt \ill
309
{bigPsi_prime}{1}{$\Psi '(t)$ is a (weighted) sum of Dirac delta functions at
310
the logarithms of prime powers $p^n$ weighted by $\log (p)$ (and by $\log
311
(2\pi )$ at $0$). The taller the arrow, the larger the weight.} \par \item
312
Now---for normalization purposes---multiply $\Psi '(t)$ by the function
313
$e^{-t/2}$ which has no effect whatsoever on the support. \end {enumerate}
314
\vskip 20pt\ill {psi_200}{.4}{Illustration of the staircase $\psi (X)$
315
constructed in Chapter~\ref {sec:tinkering} that counts weighted prime
316
powers.\label {fig:psi_200}} \par \par \par In summary: The generalized
317
function that resulted from the above carpentry: $$\Phi (t) = e^{-t/2}\Psi
318
'(t),$$ retains the information we want (the placement of primes) even if in
319
a slightly different format.
320
321
\bibitem{Note19}
322
A version of the Riemann--von Mangoldt explicit formula gives some theoretical
323
affirmation of the phenomena we are seeing here. We thank Andrew Granville
324
for a discussion about this. \par \ill {granville}{0.25}{Andrew Granville}
325
\par \par Even though the present endnote is not the place to give anything
326
like a full account, we can't resist setting down a few of Granville's
327
comments that might be helpful to people who wish to go further. (This
328
discussion can be modified to analyze what happens unconditionally, but we
329
will be assuming the \RH {} below.) The function ${\hat \Phi }_{\le C}(\theta
330
)$ that we are graphing in this chapter can be written as: $${\hat \Phi
331
}_{\le C}(\theta )= \sum _{n\le C}\Lambda (n)n^{-w}$$ where $w = {\frac
332
{1}{2}}+i\theta $. This function, in turn, may be written (by Perron's
333
formula) as \begin {align*} &{\frac {1}{2\pi i }}\lim _{T \to \infty }\int
334
_{s=\sigma _o-iT}^{s=\sigma _o+iT}\sum _{n}\Lambda (n)n^{-w}\left ({\frac
335
{C}{n}}\right )^{s}{\frac {ds}{s}}\\ &= {\frac {1}{2\pi i }}\lim _{T \to
336
\infty }\int _{s=\sigma _o-iT}^{s=\sigma _o+iT}\sum _{n}\Lambda
337
(n)n^{-w-s}C^{s}{\frac {ds}{s}}\\ &= -{\frac {1}{2\pi i }}\lim _{T \to \infty
338
}\int _{s=\sigma _o-iT}^{s=\sigma _o+iT}\left ({\frac {\zeta '}{\zeta
339
}}\right )(w+s){\frac {C^{s}}{s}}ds. \end {align*} \par Here, we assume that
340
$\sigma _o$ is sufficiently large and $C$ is not a prime power. \par One
341
proceeds, as is standard in the derivation of Explicit Formulae, by moving
342
the line of integration to the left, picking up residues along the way. Fix
343
the value of $w= {\frac {1}{2}}+i\theta $ and consider $$K_w(s,C):=\ \ {\frac
344
{1}{2\pi i }}\left (\frac {\zeta '}{\zeta }\right )(w+s){\frac {C^{s}}{s}},$$
345
which has poles at $$s= 0, \ \ \ 1-w,\ \ \ {\rm and }\ \ \ \rho -w, $$ for
346
every zero $\rho $ of $\zeta (s)$. We distinguish five cases, giving
347
descriptive names to each: \par \begin {enumerate} \item {\it Singular pole:}
348
$s= 1-w$. \item {\it Trivial poles:} $s= \rho -w$ with $\rho $ a trivial zero
349
of $\zeta (s)$. \item {\it Oscillatory poles:} $s= \rho -w = i(\gamma -\theta
350
) \ne 0$ with $\rho = 1/2 + i \gamma (\ne w)$ a nontrivial zero of $\zeta
351
(s)$. (Recall that we are assuming the \RH {}, and our variable $w= {\frac
352
{1}{2}}+i\theta $ runs through complex numbers of real part equal to ${\frac
353
{1}{2}}$. So, in this case, $s$ is purely imaginary.) \item {\it Elementary
354
pole:} $s=0$ when $ w$ is not a nontrivial zero of $\zeta (s)$---i.e., when
355
$0=s\ne \rho -w$ for any nontrivial zero $\rho $. \item {\it Double pole:}
356
$s=0$ when $ w$ is a nontrivial zero of $\zeta (s)$---i.e., when $0=s=\rho
357
-w$ for some nontrivial zero $\rho $. This, when it occurs, is indeed a
358
double pole, and the residue is given by $m\cdot \log C +\epsilon $. Here $m$
359
is the multiplicity of the zero $\rho $ (which we expect always---or at least
360
usually---to be equal to $1$) and $\epsilon $ is a constant (depending on
361
$\rho $, but not on $C$). \par \end {enumerate} \par The standard technique
362
for the ``Explicit formula'' will provide us with a formula for our function
363
of interest ${\hat \Phi }_{\le C}(\theta )$. The formula has terms resulting
364
from the residues of each of the first three types of poles, and of the {\it
365
Elementary} or the {\it Double} pole---whichever exists. Here is the shape of
366
the formula, given with terminology that is meant to be evocative: \par \par
367
$$ {\bf (1)}\ \ \ {\hat \Phi }_{\le C}(\theta ) = {\Sing }_{\le C}(\theta ) +
368
{\Triv }_{\le C}(\theta ) + {\Osc }_{\le C}(\theta ) + {\Elem }_{\le
369
C}(\theta )$$ \par Or: \par $${\bf (2)}\ \ \ {\hat \Phi }_{\le C}(\theta ) =
370
{\Sing }_{\le C}(\theta ) + {\Triv }_{\le C}(\theta ) + {\Osc }_{\le
371
C}(\theta ) + {\Double }_{\le C}(\theta ),$$ \par \noindent the first if $w$
372
is not a nontrivial zero of $\zeta (s)$ and the second if it is. \par The
373
good news is that the functions ${\Sing }_{\le C}(\theta ), {\Triv }_{\le
374
C}(\theta )$ (and also ${\Elem }_{\le C}(\theta )$ when it exists) are smooth
375
(easily describable) functions of the two variables $C$ and $\theta $; for
376
us, this means that they are not that related to the essential
377
information-laden {\it discontinuous} structure of $ {\hat \Phi }_{\le
378
C}(\theta )$. Let us bunch these three contributions together and call the
379
sum ${\Smooth }(C,\theta )$, and rewrite the above two formulae as: $$ {\bf
380
(1)}\ \ \ {\hat \Phi }_{\le C}(\theta ) ={\Smooth }(C,\theta ) + {\Osc }_{\le
381
C}(\theta ) $$ \par Or: \par $${\bf (2)}\ \ \ {\hat \Phi }_{\le C}(\theta ) =
382
{\Smooth }(C,\theta )+ {\Osc }_{\le C}(\theta ) + m\cdot \log C +\epsilon ,$$
383
\par depending upon whether or not $w$ is a nontrivial zero of $\zeta (s)$.
384
\par \par We now focus our attention on the Oscillatory term, ${\Osc }_{\le
385
C}(\theta ) $, approximating it by a truncation: $$\Osc _w(C,X):= 2\sum
386
_{|\gamma | < X} {{\frac {e^{i\log C\cdot (\gamma -\theta )}}{i(\gamma
387
-\theta )}}}.$$ Here if a zero has multiplicity $m$, then we count it $m$
388
times in the sum. Also, in this formula we have relegated the ``$\theta $" to
389
the status of a subscript (i.e., $w = {\frac {1}{2}} +i\theta $) since we are
390
keeping it constant, and we view the two variables $X$ and $C$ as linked in
391
that we want the cutoff ``$X$" to be sufficiently large, say $X \gg C^2$, so
392
that the error term can be controlled. \par At this point, we can perform a
393
``multiplicative version'' of a Ces\`aro summation---i.e., the operator $F(c)
394
\mapsto ({\Ces } F)(C):= \int _1^CF(c) dc/c.$ This has the effect of forcing
395
the oscillatory term to be bounded as $C$ tends to infinity. \par This
396
implies that for any fixed $\theta $, \begin {itemize} \item ${\Ces }{\hat
397
\Phi }_{\le C}(\theta )$ is bounded independent of $C$ if $\theta $ is not
398
the imaginary part of a nontrivial zero of $\zeta (s)$ and \item ${\Ces
399
}{\hat \Phi }_{\le C}(\theta )$ grows as ${\frac {m}{2}}\cdot (\log C)^2
400
+O(\log C)$ if $\theta $ is the imaginary part of a nontrivial zero of $\zeta
401
(s)$ of multiplicity $m$, \end {itemize} giving a theoretical confirmation of
402
the striking feature of the graphs of our Chapter~\ref
403
{ch:prime-to-spectrum}.
404
405
\bibitem{Note20}
406
A reference for this is: \par \vskip 10pt {\bf [I-K]: }H. Iwaniec; E. Kowalski,
407
{\bf Analytic Number Theory}, {\em American Mathematical Society Colloquium
408
Publications} {\bf 53} (2004). \par (See also the bibliography there.) \vskip
409
10pt Many ways of seeing the explicit relationship are given in Chapter 5 of
410
{\bf [I-K]}. For example, consider Exercise 5 on page 109. \par $$\sum _{\rho
411
}{\hat \phi }(\rho ) = -\sum _{n \ge 1}\Lambda (n)\phi (n) + I(\phi ),$$ \par
412
where \begin {itemize} \item $\phi $ is any smooth complex-valued function on
413
$[1,+\infty )$ with compact support,\item ${\hat \phi }$ is its Mellin
414
transform $${\hat \phi }(s):= \int _0^{\infty }\phi (x)x^{s-1}dx,$$ \item the
415
last term on the right, $I(\phi )$, is just $$I(\phi ):= \int _1^{\infty
416
}(1-{\frac {1}{x^3-x}})\phi (x)dx$$ (coming from the pole at $s=1$ and the
417
``trivial zeroes''). \item The more serious summation on the left hand side
418
of the equation is over the nontrivial zeroes $\rho $, noting that if $\rho $
419
is a nontrivial zero so is~${\bar \rho }$. \end {itemize} \par \par \par Of
420
course, this ``explicit formulation'' is not {\it immediately} applicable to
421
the graphs we are constructing since we cannot naively take ${\hat \phi }$ to
422
be a function forcing the left hand side to be $G_C(x)$. \par See also
423
Exercise 7 on page 112, which discusses the sums $$ x - \sum _{|\theta | \le
424
C} {\frac {x^{{\frac {1}{2}} +i\theta }-1}{{\frac {1}{2}} +i\theta }}. $$.
425
426
\bibitem{Note21}
427
You may well ask how we propose to order these correction terms if RH is false.
428
Order them in terms of (the absolute value of) their imaginary part, and in
429
the unlikely situation that there is more than one zero with the same
430
imaginary part, order zeroes of the same imaginary part by their real parts,
431
going from right to left.
432
433
\bibitem{Note22}
434
Bombieri, {\em The Riemann Hypothesis}, available at \url
435
{http://www.claymath.org/sites/default/files/official_problem_description.pdf}.
436
437
\end{thebibliography}
438
439