This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
where to runs through fj(N) lies between boimded multiples o f N (Hardy and Wright, 1960); but there are too many terms in the sum (5.21) to allow (5.17) to be deduced from (5.22).
CHARACTERS
1.3
11
a complete set o f residues m od#. Since (r, #) = 1, mr also runs through a complete set o f residues m od#, and tlie sum is unchanged. This contradicts the choice o f r with x{f ) ^ lifth e s u m is n o tz e r o . Similarly, in eqn (3.3) if m is not congruent to unity m od# there is a character xi with Xi(m) 1- W hen we multiply the left-hand side o f (3.3) bj^ 2 x(m)xi(m) is a sum over all characters m od#, and again this sum must be zero, for we have multiplied b y a constant that is not xuiity but have succeeded only in permuting the terms o f the sum. O f course, eqns (3.1), (3.2), and (3.3) are essentially special cases o f a theorem on dual finite Abelian groups. An important notion is the propriety o f characters. Let #2be a multiple o f qv and xi a character m od#x. The group o f reduced residue classes mod #2 maps homomorphically onto the corresponding group m od#x, and we define a character x% mod #2 b y the equation (3.4) W e note that xi and x% are different arithmetical functions. I f #x = 3, #a = 6, and xi takes the values 1, - 1, 0, 1, - 1, 0
(3.5)
for m = 1, 2, 3, 4, 5, 6, then %2 takes the values 1, 0 , 0 , 0 , — 1 , 0 ,
(3.6)
since Xz is zero when m is a multiple o f 2 as well as when m = 0 (mod 3). W hen Xi is constructed b y eqn (3.4), we say that ^1 m o d #1 induces Xz mod #2, and, if #2 ^ #1( that y2 m o d #2 is improper. A. proper character mod # (also called a primitive character) is one that is not induced by a character mod d for any divisor d o i q other than # itself. The smallest / for which a character Xi m o d / induces x mod # is called the conductor o f x ■ The customary letter / is the initial o f a German word for a tram conductor. W e shall now discuss Gauss's sum r(x), defined by t( x)
=
2
mmo clg
x(m)eg(m).
(3.7)
In this curious, expression, the factors correspond to the multiplicative group o f reduced residues mod # and the additive group o f all residues m od#. The absolute value o f r(x) is found below. Gauss (preceding Dirichlet) considered onty characters x for which x(m) takes the values rb 1 and zero only. In this case r 2 is real, but it is still not easy to find the argument o f r(x).
IN T R O D U C T O R Y RESULTS
12
1.3
W e shall use t ( x ) to remove characters from a summation. W ith x ( m ) denoting the complex conjugate o f x(m) (the inverse o f x hi the group o f characters), eqn (3.7) gives r(x)x(m) =
2
x(aK {am )
(3.8)
flm odg
whenever (to,#) = 1. I f r(x) is non-zero, we can use (3.8) to change a summation over x(m) to a summation over the exponential maps ea(am), which are easier to manipulate. A defect in eqn (3.8) is the condition (to, q) = 1. If, however, x is proper m od q eqn (3.8) holds for all integers m. W e must show that the sum on the right-hand side o f (3.8) is zero when
m = tn,
In this case,
2
a m od a
q = tr,
xia)eq(am) =
2
am o d g
t > 1.
(3.9)
xia)er{an)-
(3.10)
Since % is not induced by any character m odr, there is an integer b with (b,q) = 1, b = 1 (m odr), but x(b) ^ 1- Our standard p roof now applies. Multiplication by x(b) permutes the residue classes in the sum on the left o f eqn (3.10), but multiplies the value o f the sum by a constant that is not unity. The sums in (3.10) and (3.8) are thus zero if x is proper m od q and (to, q) > 1. For eqn (3.8) to be o f use we must be sure that t ( % ) is non-zero. When X is proper mod q there is an elegant demonstration. For each m m od q, |x(to)t(x)|2 =
2
2
x («)x (6)ea(«m —M -
(3.11)
2
2
x(a)x(b) 2
(3.12)
a m o d s 6 m od s
EE0IIC6 2
w m od ff
\x(m)?\T( x ) ? =
a m o d g &mod
m m odg
ea(am—bm).
The inner sum is zero b y eqn (3.1), unless a = b, when it is q, so that H x ) l2
2
mmodtz
\x{™)\2 = q
2
am od g
lx ( « ) l2-
( 3 . i3 )
Since x is proper m od# if and only if its inverse x is proper m od#, we deduce that , , f \r \X)\ q i°r X proper m o d q. (3.14) Equation (3.8) thus applies when x is proper m od#, the factor t (x ) being non-zero. W e conclude with the case when x m od# has con du ctor/, and # = fg where g > 1. I f the lowest common multiple h = [/, g] o f f and g is not # itself, x is not merely induced b y some character Xa, mo(i but is actually equal to X2> since any integer prime to h is already prime to fg — q. Replacement o f m in the sum in eqn (3.7) b y m-\-h permutes the residue classes m od#, but multiplies r(x) b y eq(h), which is not unity. In this case therefore r(x) is zero.
CHARACTERS
1.3
13
W hen (f,g) = 1 we invoke the Chinese remainder theorem. eqn (1.2) there are integers u, v with f u + g v = 1.
By
(3.15)
Residue classes a (m odfg) correspond to pairs o f classes b (m o d /), c (m od#) according to the relation a = cfu-^bgv (modfg).
(3.16)
In (3.16), a (m odfg) is a reduced class if and only if both b (m od /) and c (m od#) are reduced, and thus H x) =
2 *
amodtf
x ( a ) ea (a )
= bm2* 2* X{cfu+bgv)ea(cu)ef{bv) od f cnioclf/ =
2 *
b mod/
x ( b )ef( b v )
2 *
cmodtf
e f, M
= Xi(v)r(xiK(u)}
(3.17)
where xi m od/ is the character inducing % mod #, and cy(u) is Ram a nujan’ s sum, defined in eqn (1.11). W e now proceed to compute Ram anujan’s sum. B y eqns (1.11) and ^ ' 27^’
c„(?0 = =
2
amod{/
eu(au) X ^ ( d) d\a
mu
2 P'W bmo&gld 2 %ld{bu),
d\g
(3.18)
where we have written a = bd. From eqn (3.1) the inner sum is zero unless u is a multiple o f gjd. Writing h = g/d, we have °o(u ) = 2 ¥ ( # ) •
Z
(3' 19)
It is possible to continue and to express cu{u) in terms o f Euler’s cp function. In our application, eqn (3.15) ensures that (g,u) = 1 and thus that ca(u) is /u.(#), which is itself zero i f # has a repeated prime factor. Ram anujan’s sum with (#, %) = 1 can be regarded as Gauss’ s sum for the trivial character %0 m od#, for which Xoim) — 1 whenever (g,m) = 1. In this chapter, we have shown that r(x) is zero unless # = # // is composed solely o f those primes whose squares do not divide #, in which case
|r(X)|2 = / .
(3.20)
4 PO LY A ’S THEOREM
F r o m eqn (3.2) we see that the sum function
x (v) = 2 xim)
(4.1)
is bounded when x is a non-trivial character m odg. Since the sum over any q consecutive integers is zero, the absolute value o f X (x) can be at most \q. Polya (1918) proved the following sharper result. T h e o re m .
Let x be a non-trivial character m.odq with conductor f. (4.2)
Then where the term o (1) is to be interpreted as f -> oo.
P olya ’s theorem was discovered independently by I. M. Vinogradov (1955, chapter 3, example 12), with a different constant in the upper bound. Later proofs have been given b y Linnik and Renyi (1947) and by Knapowski in an unpublished manuscript. W e shall follow P olya’s argument, as it is the most precise and can easily be adapted to show that the sum in (4.2) is frequently ! > / 3. First we introduce some notation. I f a is a real number, we write [ct] for the largest integer not exceeding a, and ||a|| for the distance from a to the nearest integer, so that [a] = max m,
(4.3)
|a|| = min|m—a|,
(4.4)
where the maximum in (3.3) and the minimum in (3.4) are over all integers m. W e now state a lemma. Lemma. The Fourier series
(4.5) — oo
m¥=0
1.4
P O L Y A ’S
converges to
15
t h e o r e m
a __[a] _ i
if a
no%an integer,
0
i f a is an integer,
(4.6)
and the partial sums satisfy the relation M e(mct) -M
_
1
(4.7)
2 ttw i
m 7^=0
when a is not an integer. Proof. W e prove (4.7). Since H(a) has period 1 and H( — a) = —H (a), we suppose that 0 < a < -J. Now (2TTim)-1e(m<x) — (2TTbn)~1( ~ l ) m = J e(mt) At, £ M
and thus
2
(4.8)
* M
_-M nr in # 0
{2nim)-1e(ma.)-\-^ — a =
j
J _ i\j
e(mt) di
e(M t+%t)—e(—M t —lt) ^ e ( ¥ ) ~ e( ~ ¥ ) f sin(2ili"-|-l)7rf d t, sin 771
(4.9)
J
which by one o f the mean-value theorems for integrals does not exceed in modulus
0
sin(2Jf-(-l)77i
sin 77a
At
1 2a (2M + \ ) t t
1
(4.10)
2 ttM cx
for some j3 between a and This has proved (4.7), which gives (4.6) if a is not an integer; if a is an integer every term in H (a) is zero. W e now return to the sum in eqn (4.1). Since x is non-trivial, we have rq+q 2 x(w) = o,
(4.11)
rq + l
and so we can replace the sum X (x) by a sum o f the same form with 0 ^ x < q. I f now x > \q, we can replace m by — q-\-m and obtain a sum with 0 ^ x ^ \q, multiplied by —x ( — 1). Thus we can suppose that 0 < x ^ \cq in (4.1). W e now use H (a) to construct 0(a), where n 0(®) = I \ \0
if 0 < a < xjq, if a = 0 or a = xjq, if xjq < a < 1,
(4.12)
16
IN T R O D U C T O R Y RESULTS
1.4
using the equation 0(a) = xjq-\-H(a.—xjq)—H(ot).
(4.13)
W e now have X(l) + x(2) + ...+ l* (a 0 =
I
x(m)0(mlq).
(4.14)
m —1
W hen we use eqn (4.7) to truncate the sums for H (a —xjq) and H(a) in (4.14), the total error in modulus is at most
2J
M
[„ ,/„« )-! <
1 dh?
4
m=l
< 277-~1i l f - 1#logg.
(4.15)
W e now have finite sums to manipulate. Writing CO
0(a) = 2 a(m)e(ma),
(4.16)
— CO
we have to consider q
M
2 X(r)
J .= l
2
—M
es(TOr)-
(4.17)
Supposing first that x is proper m od#, we have from eqn (3.8) a
m
2 a (m) 2
-M
.
M
x0')e9(™') =
>■=1
2 a(m)x(m)r(x).
(4.18)
—M
Since x is non-trivial a(0)%(0) is zero, and by eqns (4.5) and (4.13), for \a(m)\ < \TTin\-1.
(4.19)
The modulus o f the expression in (4.18) is now /
M
< ? i 2 | (nm)-1 m=l
W e choose so that (4.20) is
< 2 7 T ^ ^ (lo g J f+ 0 (l)).
(4.20)
M = qi+s,
(4.21)
77-^1gi lo g g (l-l-0 (S)),
(4.22)
and after (4.15) the tails o f the series give < ( n M ^ q l o g q = o(qHogq).
(4.23)
Finally the omitted term \xix ) is 0(1), and we have P olya’s theorem when x is proper mod q.
P 6 L Y A ’S
1.4
IT
t h e o r e m
I f x is induced by Xi proper m o d /, where / < q, we could complete the proof similarly, but it is easier to deduce this case. 2
x (™ ) =
= <
2
m^cc
X iM
X
d\q d\m
M d)
2
^(d)Xi(d)
2
/ } l o g '/ { ^ + o ( l ) } ,
d\q
2
m ^xfd
Xi(»») (4.24)
d\q
(d, f) = 1
since xi is proper m od /. The number o f terms in the sum is at most d(qjf), and we have completed the proof o f P olya’s theorem. As a simple corollary we prove that for each prime p > 2 there is an integer m with
m < p i\ogp
(4.25)
u2 = m (m od#)
(4.26)
for which the congruence
has no solution. Since 1, 4, 9,..., £(#— l )2 are distinct m od#, there are \(p — 1) reduced residue classes that are congruent to squares m od#, and these form a subgroup o f index 2. There is therefore a character X m od # with x(m) — 1 when m is congruent to the square o f a reduced residue and — 1 when (m,q) = 1 but m is not congruent to a square. W e choose . i, ,. x> p -\ og p (4.27) in (4.2). N ot all terms in the sum (4.1) can be non-negative if p is sufficiently large, and so there is an m < # - l o g # with xim) = — !• The exponent \ in (4.25) can be improved, but the conjecture that the asymptotic inequality
< #s
(4.28)
for each 8 > 0 holds in place o f (4.25) has not yet been proved or confounded.
5 D IR IC H LE T
SERIES
’ Well,’ said Owl, ‘ the customary procedure in such cases is as follows.’ ‘What does Crustimoney Proseedcake mean ? ’ said Pooh. ‘ For I am a Bear of Very Little Brain, and long words bother me.’ ‘ It means the Thing to Do.’ I. 48
A Dirichlet series is an analytic function o f the complex variable s = c j+ it defined b y a series / ( « ) = 1 a{m)m~s,
(5.1)
or a generalization thereof. All the Dirichlet series that we need are special cases o f (5.1). I f eqn (5.1) converges at s0 = cr0-(-ii0, then \a(m)m~8< ‘ \ is bounded. This simple observation is the basis for the theory o f convergence o f Dirichlet series. B y partial summation (5.1) converges whenever a >
4 (® )= 2 a (m )
(5.2)
is the sum function o f the coefficients in (5.1), then CO
f(s ) = j s x - ^ A i x ) da;.
(5.3)
i
Formula (5.3) can be inverted: fro m /(s) we can recover the sum function A(x) o f the coefficients. Let a and u be positive real numbers. Then ar-f- i oo
23
(0
/
1
a- ic0
\1
if
0
if« =
1,
(5.4)
if U > 1.
For tt ^ 1 we consider the integral from a —i2\ to a-f-iT%. W hen u > 1, this is equal to the integral round the three remaining sides o f the
1.5
D IR IC H L E T SER IE S
19
rectangle whose other corners are Rjlogu^iT^, R jlo g u —iT^ modulus o f the integral in eqn (5.4) is thus o~R R ______ -^ e -K . ir ___________________________ 2tt t 277^ log u 1 2irT2i\ogti’
The
( 5 gx
a number which tends to zero as R, Tx, and T2 tend to -|-oo. W hen u < 1, Rflogu is negative, and we must add the residue from the pole o f s-1 at s = 0 ; this gives unity. Finally when u = 1 we define the value o f the integral (5.4) to be the limit o f the integral from a —i T to a-\-iT when T -> oo. This reduces to an inverse tangent integral. T
Rj logu
Of.
1' 1 F ig .
1
I f f(s ) defined b y eqn (5.1) converges uniformly in t on the line a — a, then for x > 0 term-by-term integration gives oi-J-ico
f J
a —ico
x ^ f i s ) ds = 2 a(m)+±a(x), m<x
(5.6)
where the last term occurs only if x is an integer. There are many integral transforms from Dirichlet series to their coefficient sums, all proved b y the same method. The simplest one after (5.6) itself is a-f ico
h J
2 ( - “)«<»>•
a —ico
A special class o f Dirichlet series occurs naturally in number theory. The functions are known as L-functions after Dirichlet’s functions
Lis>x) = 2
(5.8)
m —1
where x is a Dirichlet’s character to some modulus q, or as zeta functions after Riem ann’s function £(«) =
2 m ~s> m —1
(5.9)
IN TR O D U C TO R Y RESULTS
20
1.5
which is the special case o f Dirichlet’s definition when q = 1 and x is trivial. ^-functions are defined by two properties. First, the coefficients a(m) are multiplicative, so that Euler’s product identity a(m ) _
-|— r (
a(p )
a ( p 2)
m=1 holds in a half-plane a ^ a in which one side o f eqn (5.10) converges absolutely. I f the product in (5.10) converges, f(s ) can be zero only when one o f the factors on the right-hand side o f (5.10) is zero. The convergence o f the left-hand side o f (5.10) alone does not im ply that o f the product; L(s, x) with x non-trivial has a series (5.8) converging for a > 0, but the function itself has zeros in a > preventing the product from converging in 0 < a < J-. The second defining property is that f(s ) should have a functional equation
f{s)G (s) = f * ( r - s ) G * ( r - s ) ,
(5.11)
where r is a positive integer, G(s) is essentially a product o f gamma functions, and the operation * has ( /* )* = / and ((?*)* = G. As an example, in the functional equation for L (s ,x ) in Chapter 11, L*(s, x) is L (s ,x ). An important conjecture about Z-functions is the Biemann hypothesis that if f(s ) satisfies eqns (5.10) and (5.11) then all zeros o f f(s)G (s) have real part Jr. The truth or falsity o f this hypothesis is not settled for any Z-function. Two generalizations that are often called zeta functions are CO
1 ( m + 8) - s, m=l
(5.12)
where 8 is a fixed real number, and 00 2 r(m)m~s, m=1
(5.13)
where r(m) is the number o f representations o f to by a positive definite quadratic form. E xcept in special cases these fail to have a product formula o f the form (5.10), and not all o f their zeros lie on the appropriate line. Some authors even use ‘ zeta function’ as a synonym for ‘Dirichlet series’ . In Chapter 11 we shall obtain analytic continuations o f £(s) and other .L-functions over the whole plane. Since the sum function X (x ) formed
D IR IC H L E T SE R IE S
1.5
21
with a non-trivial character %is bounded, b y partial summation (5.8) con verges for ct > 0 except when x is trivial. Similarly, the function co
X ( — l ) m-1TO~~s = (1 —21-S)£(s) m—1
(5.14)
converges for a > 0 and provides an analytic continuation for £(s). W hen we make s 1 in (5.14), we see that £(s) has a pole o f residue 1 at s = 1. W hen we p u t /(s ) = £(s) in (5.6), the integrand has a simple p o le a ts = 1 with residue®. The value o f the right-hand side o f eqn (5.6) is between x — 1 and x. I f we deform the contour in (5.6) so that it passes to the left o f the pole, the residue makes the main contribution, and the contour integral left over is bounded. Let
M (x) =
2 A{m)
(5.15)
2 Mm)>
(5.16)
m <x
m < (G
which are the coefficient sums o f — £'(s)/£(s) and o f l/£(s) (we shall prove this below). The function — £'(s)/£(s) also has a pole o f residue 1 at s = 1, but l/£(s) does not. I f the corresponding contour integrals were negligible, we should have t(x ) = x + o ( x ) ,
(5.17)
M (x) = o(x).
(5.18)
These are forms o f the prime-number theorem, which we shall prove in Part II. Writing m = f g , we have co
,
CO
\
i
co
N
2 m~s 2 « ( / W W / ) = ( 2 « ( / ) / “ *) ( 2 H g ) r s)-
m= 1
f\m
'/= 1
' ' ( 7=1
'
(5’ 19)
I f b(g) = 1 and a (f) = /x(/) for each pair o f integers / , g, co
£(5) 2 K m)m~a = m =l
CO
2 w “ s 2 M /) = 1
m =l
/|m
(5'2°)
from eqn (1.27). Since eqns (5.3) and (5.6) im ply that expansions in Dirichlet series are unique, we have shown that the series on the left-hand side o f (5.20) represents l/£(s) wherever it converges. Similarly, using eqns (5.19) and (1.32) we can check that — £'(s)/£(s) has a Dirichlet series with coefficients A(m). For fifty years (1898-1948) the only proofs known o f eqns (5.17) and (5.18) used contour integration and other complex-variable techniques. In 1948, Selberg and Erdos gave a real p roof o f (5.18) (see Hardy and
22
IN TR O D U CTO R Y RESULTS
1.5
W right (1960), Chapter 22). The real-variable approach is not so well understood, and the strongest forms o f (5.17) and (5.18) (those in which the error term is smallest) have been obtained b y analytic methods. The form (16.22) in which we shall prove (5.17) is a little stronger than the best so far obtained b y Selberg’s method. Aj)art from the analytic arguments, studjr o f log(A'r!) suggests the form (5.17) as a conjecture. B y eqn (1.32),
=
2 MN/e),
(5.21)
e < iV
where weJiave written m = tie in the first sum. On the other hand, by expression (2.5),
= N logN ~ N -}-0(logN )
(5.22)
which agrees with the result o f substituting (5.17) into (5.21). In this way, Gauss was led to conjecture that (5.23) 2
which is another form o f the prime-number theorem. B y consideration o f the binomial coefficient 2lvCjv> ^ can be shown that
6 SC H IN ZE L ’S H Y PO TH E SIS And all the good things which an animal lilies Have the wrong sort of swallow or too many spikes. II. 30 M a n y problems in prime-number theory follow a similar pattern.
Various constraints are laid on a set o f integer unknowns, and we ask whether the integer unknowns can all be jsrime simultaneously, and whether this happens infinitely often. Many o f these problems are subsumed under SchinzeVs hypothesis: i£ f1(x1,...,xn) ,...,fm(x1,...,xn) are polynomials (with integer coefficients) irreducible over the integers, and there is no prime p for which J J / { = 0 (m od#) for all sets xn o f residues m od# , then there are infinitely many sets x v ..., xn o f integers for which the absolute values o f / 1;..., f m are all prime. There is a conjectiired asymptotic formula for the number o f sets ik-l,..., x n with 0 < xi ^ N for each i with each o f f m prime, and some bold authors have conjectured that the asymptotic formula can be stated with an error term smaller than the main term b y a factor for each e greater than zero. Thus the simplest case is one p oly nomial, f ( x ) = x, and the conjecture now states that tt(N), the number o f primes up to N, satisfies the relation ( 6 . 1) 2
a very strong form o f the prime-number theorem (5.23). The accuracy o f (6.1) seems unattainable. W e shall prove later that the hypothesis is true for one linear polynom ial/(a;) = qx-\-a\ this is the prime-number theorem for arithmetical progressions, but the error term in the asym ptotic formula will only be shown to be slightly smaller than the leading term. The next simplest case concerns two linear polynomials, f ^ x ) = x, f 2(x) = x — 2. Here the conjectured formula is Ar
2 XJ {1 — 853518X
— l ) " 2} I (log £)“ 2
-f-error term,
(6.2)
24
IN T R O D U C T O R Y RESU LTS
1.6
W e shall now describe how to write down the conjectured asymptotic formulae. Let
a/ \ S (a )=
v 2, p^N
n.
m o\ (6>3)
such an expression is called an exponential sum or a trigonometric sum. B y the fundamental relation / / » t (1 J e(m«) * . = ( „
i f m — 0, jj
(6-4)
we see that the number o f primes p ^ N for which p — 2 is also prime is l J S(a)S(—a)e(— 2a) da. o
(6.5)
W e cannot, o f course, work out this integral, but we can suggest a plausible value for it. Writing n (N ;q ,b )=
2
1.
(6.6)
P
p = b (modg)
we have
S(a/q) =
2
ea(ab)ir(N;q,b).
(6.7)
fcniodg
N ow the sum in eqn (6.6) is 0 or 1 i f b has a common factor with q. I f we make the approximation that the primes are divided equally between the
(6.8)
where cq(a) is Ram anujan’s sum (1.11). I f (a, q) = 1 the Ram anujan’s sum is just /x(g), from eqn (3.19). This argument suggests that S(a) has a ‘ spike’ at a/q o f height proportional to ii(q)l
2V'"17r2(iV*)-{-error term,
(6.9)
and this can be proved to be true. I f we assume further that all the spikes at rational points are the same shape, the spikes at rational points ajq contribute N^1tt2(N) 2 lJ'2(q){(q))~z 2 *
amodg
Q
ea(2a)+error term,
(6.10)
and the sum over q in (6.10) converges to 2n { i - ( ? - n P> 2
which gives the main term in eqn (6.2).
(6. n )
1.6
SC H IN ZE L ’ S H YP O TH E SIS
25
For small q the argument above can be made rigorous; but then part o f the range o f integration in (6.5) does not support spikes. Away from a spike we cannot estimate S(a) except b y replacing it b y its absolute l value; and the spikes with small q contribute very little to J |$(a) |2 da. o In the integral o f |$(a)|3 the spikes do dominate, and b y this method I. M. Vinogradov was able to prove that every large odd number is the sum o f three primes. The approach to Schinzel’s hypothesis through exponential sums does lead to an upper bound for the number o f sets x 1}..., xn o f integers not exceeding N for w h ic h /l v .., f m are all prime. To explain the method we shall take n = 1, so we are considering integers * in the range 1 ^ x N for w liich /1(a;),...,/m(a;) are all prime. W e now work modulo a prime # . Apart from the finite number o f x for which one o f f t(x),..., f m(x) is p, x must be such that none o f fx(x),...,fm(x) ~ 0 (m odp). This means that x must be confined to certain residue classes m od#. W e therefore divide the residue classes m od # into a set H (#) o f /(# ) forbidden classes and a set K (p ) o f g(p) = # —/ ( # ) permitted classes; h m od # is forbidden if and only if one o f the p o ly n o m ia ls /^ ) is a multiple o f # . I f x falls into a forbidden class for any prime # smaller than each o f the fi(x), then one at least o f t h e / i(x) cannot be prime. The values o f x that make f x(x),..., f m(x) primes greater than some bound Q form a sifted sequence, in the following sense. The increasing sequence J f o f positive integers % , n 2>... is sifted b y the primes # < Q if for each prime # < Q there is a set H(p) (possibly empty) o f /( # ) residue classes m od # into which no member o f J r falls. W e shall show in Chapter 8 that, i f J/' satisfies the above condition, the number o f members o f J f in any interval o f N consecutive integers is N v ... ■>,...wi, w ..+ error t e r m >
I /**(?)/(?)/?(?)
( 6 - 12)
a
where
f{q) = q H f ( p ) l p ,
(6-!3)
g(i) = ? IT { l - f i v ) h ? } v\a
(6-14)
W e shall work out examples o f this upper bound in Chapter 8 ; in each case the leading term is a multiple o f the leading term in the conjectured formula. Upper bounds o f the right order o f magnitude were first found by Viggo Brun using combinatorial arguments. Rosser used Brun’s method
26
IN T R O D U C T O R Y RESULTS
1.6
to obtain expression (6.12), which was found in a different way b y Selberg. A n outline o f Selberg’ s method follows. It rests on the construction of an exponential sum T(a) with the same spikes at rational points as S («) =
2
eK «)
« !< N
(6-15)
is conjectured to have. Let K (q ) be the set o f g(q) residue classes that are in K (p ) (permitted classes) for each prime factor p o f q. A t a/q we expect a spike o f height proportional to K {a,q) =
2
k eK (q )
ea(ak)l9(q)-
(6' 16)
I f p is a repeated prime factor o f q, then the classes k-\-qjp are also in K{q), and since replacement o f k in eqn (6.16) b y k + q j p multiplies the right-hand side b y ep(a), which is not unity, K(a, q) is zero if q has any repeated prime factor. N o av the simplest exponential sum i s W
=
2 e(ma), m=l
(6.17)
which has a spike at a = 0, since ]F (a) | = |sin ttN a/sin tt<x \.
(6.18)
W e therefore compare 8(a) with y («) = 2
2*
q^ Q omodg
K (a ,q )F (a -cilq ).
(6.19)
The coefficient o f e(ma) in T(a) is
2
g«CQ
^
2fteK (g) aTm o d q
= g2« :Q ®
2
k eK (q )
(6 .2 0 )
using the definition (1.11) o f Ram anujan’s sum. W e know the sum over 1c must be zero if q has a repeated prime factor; i f q is square-free then i \
=
2
e K(q n) k eK
2d\n d\q
d\(k-m)
M 'I
!
g\g(g) d\q meK(d)
w M )
= pte)9te)M ftei)lgtei),
(6.21)
where qx is the largest factor o f q for which m e K ( q 1). W e write 2 = so m e H ( c l ) then d divides q2. The expression in
(6.21) becomes
A
(6.22)
1.6
SC H IN ZE L ’S H Y PO TH ESIS
N ow
27
=
(6.23)
and so the coefficient o f e(ma) in eqn (6,19) is
d=C_Q
meH(d)
ii{d)d
^
fid )
Z-,
J
'
^{q)f{q)
g = o(m ocld )
(6>24)
glq)
s
Selberg’s argument proceeds as follows. I f m is in only when d = 1, so that b y eqn (6.4)
jo W
nm^N 'a^QT
W e obtain (6.12) b y Cauchy’s inequality, l i i J S{<x)T(-~a) d a 2 < j ^ ( a ) ! 2 0 noting that, from eqn (6.4), 1
and
'
<6'26>
J |T(/3)|2 dj3, 0 0
(6.26)
1
f | £ (a )| 2 d a =
0
da
then m e H(d)
f j S ( a ) /S ( -a ) d a =
2
1’
( 6 '2 7 )
0J
f |T (a)|2 da = N ^ o s«Q
+ error term.
(6.28)
W e can prove eqn (6.28) either directly from (6.19) or b y writing ^ ( a )!2 as T(a)T( — a) and using (6.4) again and the expression (6.24) for the coefficients. Selberg used the second method, which can be expressed entirely in terms o f the manipulation o f the coefficients (6.24), with the integrations (6.25) and (6.28) occurring only implicitly. The first proof uses eqn (6.18) to show that the spikes o f T(a) provide the main contribu tion to the integral in (6.28), and the observation that \K(a,qW = ^(q)f(q)lg(q).
(6.29)
am odg
The usual account o f Selberg’s method in terms o f coefficient manipu lations can be found in Halberstam and R oth (1966, Chapter 4). The sketch above is presented in terms o f exponential sums for comparison with the ‘large sieve’ o f the next two chapters.
7 THE
LARGE
SIEVE
But whatever his weight in pounds, shillings and ounces, He always seems bigger because of his bounces. II.30
W e have seen that the behaviour o f a given sequence o f integers con sidered m odg is reflected in the behaviour o f the sums 8(a/q) where JV S(ot) = (7.1) 1 in which am is 1 if m is in the given sequence, and 0 if m is not. An ujjper bound for the sum ^ ^ |(g(a/?)|a (7 2) q ^ Q amodg
(or for some related expression) is called a large sieve for residue classes. Other large sieves will appear in Chapter 18. In this chapter we prove what is probably the simplest o f the upper bounds for the sum (7.2), and in the next chapter we shall use it to prove an upper bound o f the form (6.12) for the number o f elements o f a sifted sequence in a bounded interval. The p roof does not require that am in eqn (7.1) takes only the values 0 or 1, and it treats (7.2) as a special case o f the sum
f |S(*,)la (7.3) }.=i where 0 ^ x x < x 2 < ... < xR ^ 1. Since ajq occurs in the sum (7.2) only i f it is in its lowest terms, the points x r are distinct in our proposed application. W e write S = min{*2—a^Xg—x z,...,xx+ l — x ^
(7.4)
and suppose that S > 0. Before proving an upper bound for (7.3), we consider what form it might possibly take. Certainly there will exist sequences o f coefficients am and points xr for which b 1 2 |S(® ,)|»>8- i f |S(«)|»d« »•= ! J
= 8-1 £ K P 1
(7-5)
THE LARG E SIEVE
1.7
29
On the other hand, any one term in the sum (7.3) may be as large as ( 2 K n!)2> ari(l ^ ail the
are equal in modulus this is t f f h J 81
(7.6)
An optimistic conjecture is that the inequality |
\S(xr) \ * ^ ( N + 8 ~ i ) f \ a m\*
)■=i
(7.7)
i
always holds. Surprisingly, the right-hand side o f (7.7) has the correct order o f magnitude: we shall prove that f
\S(x,.)\^ ( i V + p - W 3 + 0 ( l ) ) i
(7.8)
■ml2;
1
r=l
this result is due essentially to Bombieri (1972). To prove the relation (7.8) we use the language o f ^-dim ensional vectors over the complex numbers. The inner product (g, h) o f two vectors g = {g-s_,---,gN) and h — (hv ...,hN) is given by N
and the norm ||g|[ b y
(g, h) = % g j i ns x
(7.9)
||g|| = ((g,g))*.
(7.10)
W e can now state a fundamental lemma. Lemma. Let u, f (1),..., f(fi) be N-dimensional vectors, and cv ..., cR be any complex coefficients. Then
f c ,( u ,f W ) r —1
< ||u||(| | c,| f ( max ' 1
'
f
|(fM,fM)|)*.
1,...,R 8=1
(7.11)
'
Proof. The left-hand side o f (7.11) is R
(u , | c,fW)
2 M W
(7.12)
The square o f the second factor on the right-hand side o f (7.12) is
i
f crcs(w,f®) < 1 f
j’ = l s = l
r=l
=
2
r=l
| (Icri«+Kia)i(fw,fw)i s==l
K l2 2
s=1
l(f(,,)>f(8))l»
(7.13)
from which the result follows. In applying the lemma we choose c,, so that each summand on the left o f (7.11) is real and positive. In Chapter 27 we shall apply the lemma
30
IN T R O D U C T O R Y RESULTS
1.7
with cr o f unit modulus to obtain Halasz’s method for estimating the number o f times 8(a) is large. To prove (7.8) we take cr given b y cr = (fMf<s>)
(7.14)
and deduce the corollary. We have
C o ro lla ry .
2
|(u,fW)|2 < jjujj3 m ax
r=1
^
|(f(,,),f (s))|.
(7.15)
1<9'
W e write
M S(u.) = e(M<xJr a) 2 bmc(ffla),
(7,16)
-M
where 2M = N or N — 1, whichever is even. W e choose = &me(—m®,),
(7.17)
um = b j k m,
(7.18)
where k_M,..., kM are given by km = 1
for \m\ ^ M (7.19)
k.m = 0
for \m\ ^ M-\-L
)
here L is a parameter which we shall choose below. W e now have M
2 b me(mx,) = \8(xr)\.
l ( « , f (,,))l =
Moreover
—M
M !l«l!2 =
and
2
-M
(7 .2 0 )
N IKWK, = 2 K l 2 i
( f « fW) = K (x s~ x r),
(7.21) (7.22)
M + L
where
K(a) =
e(ma) -M -L
sin2(ilf -f- L )n a —sin2ilf7ra
(7.23)
Hence we have J*
2
S= 1
l-^r(®s' '*'»') I
2-ZI4"-f--Zv— }—2
i(B -l)
^
1
Z -1 coseo27TW§ CO
< 2 i ¥ + JL + 2 L - 17T2 |;TO-2S--2+ 0 ( L - i ) l
< 2 i f + Z + ( 3 i > § 2) - 1+ 0 ( i v - i ) .
(7,24)
1.7
THE LARG E SIEVE
31
When we choose L to be an integer close to (SV3)-1, (7.24) becomes < 2 iY + f S - 1V3 + 0(1).
(7.25)
W e substitute (7.20), (7.21), and (7.25) into (7.15) to obtain (7.8). Our inequality (7.8) represents an improvement o f an inequality o f R oth (1965), which has led to much recent work. The best upper bounds known for the sum (7.3) at the time o f writing are
S-1(l + 270Ar3S3) f
\am\2,
(7.27)
i
and
2max(AT, S^1) 2
i
(7.28)
Of these, (7.2 6) is the result o f this chapter, appropriate when N > f 8 _1V3, and (7.27) and (7.28) are results o f Bombieri and Davenport (1969,1968). (7.27) is appropriate when S_1 > 3(10)i/V, and (7.28) for the inter mediate range. Note added in proof. H. M ontgomery and R . C. Vaughan have now proved the conjecture (7.7). This supersedes (7.26) and (7.28) but not (7.27).
8 THE
U P P E R -B O U N D
SIEVE
1It’s a comfortable sort of thing to have said Christopher Robin, folding up the paper and putting it into his pocket. II. 170
I n this chapter we obtain the upper bound (6.12) as an application o f the large sieve. The notation is that o f Chapter 6. J r is a sequence o f positive integers, and for each q ^ Q there are sets H(q) and K(q) o f residue classes m od#. The f(q ) classes o f H(q) are precisely those that are not congruent to any member o f J r m od# for any prim e# dividing q, so that, if h is in a class o f H(q) and n e J f, (n—h,q) = l.
(8.1)
The g(q) classes o f K(q) are those that for each # dividing q are congruent m o d # to some member o f J r \their union contains all members o f the sequence J f. W e work with the exponential sum o f eqn (6.15): #(“ ) =
2
« ( « » “ ).
(8 '2 )
where the sum is over members nv n2,... o f the sequence J r . I f h is a class o f H(q), 2*
S(alq)ea{ - a h ) =
am odg
c ^ —h)
2 U; < N
= p{q)M,
(8.3)
where M is the number o f members o f J r , and we have used eqn (3.19) in the sjiecial case (8.1). Hence
2* 2 8{alq)eq(-ah).
(8.4)
a m o d g heH (q)
Cauchy’s inequality now gives ,i*(q)f*(q)M* ^
(
2 *
' am odg
| S ( « / ? ) l a) (
2
' ' am odg
2
^
heH ( q) hef-T(n\
)
1
'
>
( 8 -5 )
TH E U P P E R -B O U N D SIE V E
1.8
33
The second sum over a on the right-hand side o f (8.5) can be rearranged as follows. 2*
2
amodg heH(q)
ea( —ah) 2
geH(q)
%iacJ)
=
2
2
2
2
geH(q) JisHiq)
=
05H(q) heH(q)
c fl( 0 — A)
2 M v l d) d\(g-h)
= 'ZMqld)P(q)lf(d).
(8.6)
d\q
In eqn (8.6) we have a sum that is a Mobius inverse (in the sense o f (1.27)) o f that in
' /*2(}’)fl,(r) _ rlm
f (r )
m
(8.7)
m y
an equation that expresses the fact that for a prime modulus p, every residue class is either in H(p) or in K{p). The terms involving d in (8.6) therefore come to p?(q)g(q)lf(q), and we can put (8.5) into the form 'V '*
j^ / J 2
""
q lq ) Jw
Z -i amodg
S‘a
( 8 .8)
W e apply the large sieve (7.8) with the rationals ajq with q ^ Q and (a ,q ) = 1 as the points x x,..., xR, so that § = (Q (Q -l))-1
(8.9)
in eqn (7.4). The upper bound (7.8) now gives 2
2*|-S(«/?)|2 < ( ^ + 0 ( e 2) ) ^ .
q<,Q amodg
(8-10)
Combining (8.8) and (8.10), we have M 2 2 [i2(q)f(q)lg(q) < (N + 0 { Q * ) ) M ,
(8.11)
q^Q
or
M ^ ^ + ^ ( ^ 2) M ^ W i i i ' q
fo i9\ ( ' ’
which is the relation (6.12) with an explicit error term. W e have stated our result for the interval 1 ;;C n $ ' N o f the sequence J /\ but (8.12) holds as an upper bound for the number o f members o f J r in any interval LA-1 < %
34
IN T R O D U C T O R Y RESU LTS
1.8
Tw o worked examples follow. First we consider the perfect squares not exceeding N. These are a sifted sequence: the set H(p) contains the residue classes m od # that do not contain squares, and thus / ( 2) = 0,
g( 2) = 2
f ( p ) = U p - 1)>
9(P) = U P + !)
fo r # > 3
I t is not difficult to sIioav that '2,p2te)fte)lg(q) =
(< H -o (i))Q ,
(s.i4 )
where c is a constant, as Q -> oo. Choosing Q — N -, we have shown that the number o f perfect squares not exceeding N is 0(2V*). This is very encouraging, since the sieve upper boim d is ‘ sharp’ , differing only by a constant factor from the actual number o f squares. It is surprising that we have not lost the correct order o f magnitude in combining so many inequalities. Our second, less trivial example concerns the primes between Q and N ; these form a sifted sequence with f ( p ) = 1> and thus a
v(p
g(q)
)
=
#
- 1 =
( 8 -1 5 )
Z-t
(8.16)
when worked out as in Chapter 2. W e can obtain a lower bound for the sum in (8.16) with less effort, since
2 ^ = 2 fir n K + ? + ~ )
q^Q r
q^Q
P\q '
‘
= 2 iK
t8-1?)
the sum being over all m whose prime factors are at most Q. All m ^ Q are included in this sum, and thus
2 ^T W '> 2m<<3s > ' ^ for q > 1, by (2.5). When we choose Q a little smaller than that the primes between 1 and N number < Q + { N + O m ) / l o g Q < (2 + o (l))N/logN.
<8'I8> we see (8.19)
The right-hand side o f (8.19) is just double the true value (5.23). M ore over, (8.19) is also an upper bound for the number o f primes in any interval o f length N.
1.8
TH E U P P E R -B O U N D SIE V E
35
W e derived the inequality (8.12) from Cauchy’s inequality; the difference between the two sides o f the inequality (8.12) is a measure o f how closely the values o f S(a/q) are proportional to those o f O /ta))'1 2
lx K (q )
e«(cl&)’
(8'2°)
and this in its turn measures how evenly J f is distributed among the g(q) residue classes m od# into which it is allowed to fall. W e could add an explicit term on the right o f (8.8) to measure the unevenness (what statisticians might term a variance). The inequality (7.8) gives a strong upper bound for this variance as well as for the main term. W hen we use (7.8) to prove Bom bieri’s theorem, it is the variance bound that is important, not the bound for the main term.
9 FR A N E L ’S THEOREM ‘ It’s just a thing you discover’, said Christopher Robin carelessly, not being quite sure himself. I. 109
T h e Farey sequence o f order Q consists o f the fractions ajq in their lowest terms (i.e. (a, q) = 1), with q Q ancl 0 < a q. W e name them f r = ar/qr in increasing order, so that f x = 1IQ ,f2 = l/(Q — l),...,fF = 1. For notational convenience we may refer to f F+r] this is to be interpreted as 1+ / r Here F is the number o f terms in the Farey sequence, so that
F = 2 ?(?) = 377-*Q*+0(Q log Q)
(9.1)
from eqn (2.11). The properties o f the Farey sequence are discussed by H ardy and W right (1960, Chapter 3). W e shall sketch a p roof that fr+l
fr =
( M m ) - 1-
(9.2 )
Let us represent rational numbers ajq (not necessarily in their lowest terms) by points (a, q) o f two-dimensional Euclidean space. Since/,, and f r+1 are consecutive, the only integer points in the closed triangle with vertices 0 (0, 0), Pr (ar, qr), and Pr+1 {ar+1, qr+1) are its three vertices. B y symmetry, the only integer points in the parallelogram OPr TPr+1 are its vertices, where T is (ar-\-ar+1, qr-\-qr+1). W e can now cover the plane with the translations o f this parallelogram in such a way that integer points occur only at the vertices o f parallelograms. It follows that OPr TPr+1 has unit area, which is the assertion (9.2). Before stating FranePs theorem we introduce some notation. For 0 < a < 1 we write E{a) =
2
1- a F ,
(9.3)
so that E(a) is the excess number o f Farey fractions in (0, a] beyond the expected number aF. Franel considered the sum S Im fr) I2
r= l
(9.4)
•>
F R A N E L ’S TH EO REM
1.9
37
and showed that it is o (Ql) i f and only i f eqn (5.18) holds; and an upper bound for (9.4) with 3 < A < 4 is valid if and only if \M{x)\
(9.5)
W e can also connect M (x) with l J |2?(o:)|2 da. o
(9-6)
In fact, the ratio o f (9.4) and (9.6) lies between bounded multiples o f F ; but this fact requires proof, as the Riemann sum corresponding to the integral (9.6) and the p o in t s /i,...,/F is f
)•=1
(fr+ l-fr) i m W ,
(9.7)
and from eqn (9.2) we see that the difference f r+1—f r varies from Q -1 almost down to Q~2. Franel (1924) produced a curious identity for the sum (9.4). W e shall show in (9.20) that (9.4) is less than a bounded multiple o f Franel’s expression involving M (x), and deduce the ‘only i f ’ clause o f Franel’s result b y a method o f Landau (1927, Vol. II, pp. 16977). W e use the function H(a) o f eqns (4.5) and (4.6): tt, \ # (« ) =
M —\ n (0
i f a is not an integer, •+ it-fa is■ an integer.
9-8
W e have £
_ (F(a.) i f a is in the Farey sequence, “ \ J ? ( « ) - i if not,
. . (J' J)
for 0 < a -
-■ J] m# 0
e(mot) 277-1TO
(9.10)
Q2INI
we have from (9.9) F
t=i
V, »rt# 0
I
38
IN TRO D U C TO RY RESULTS
1.9
B y (9.2), the consecutive Farey fractions are at least Q~2 apart, and so the sum of the error terms in (9.11) is < Q~z X Q2lt < log -P < log Q,
(9.12)
t= i
where we have used expressions (2.5) and (9.1). We can now replace the term t = r (since H ( 0) = 0) and rearrange the first term on the right o f (9.11) as F
Q2
X
X
t~ 1
Q2
{27nm)-1e{mfr—mfi)
=
X
2 —Q2
—Q1 m# 0
to #
3
0
X*
e (m /^ )e ( — a m )
amodg
2 (2wim)-1e(TO/1.) X cq(-™). S o a<°1 (9.13)
=
We apply the large sieve (7.8) with (9.13) as the exponential sum and /x,..., fff as the points. By eqn (9.2), S-1 is here Q(Q— 1). We have F Q2 Q X (2 7r i m ) - 1e (m /).) X c a( — m ) 2 cg(—m) X r= 1 -Q2 to#0 TO TO##0 Q2 2
2
ra=l
2
*
2 = 0 (m o d d )
(9.14)
where we have used eqn (3.19) for Ramanujan’s sum. The coefficient of Q 2 in (9.14) is now <
2 d^Q
dM(Qld) X f M (Qlf) f^Q
X
m = 0 (m o d tf) w = 0 (m o d /)
m~2>
(9 J 5 )
where we have taken the sum over to from m = 1 to infinity, since the coefficient o f to~2 in (9.14) is non-negative. The inner sum in (9.15) is 772
1
_ tr2 ( d J Y
6 \ d ,ff ~
6 dT
‘
1 '
;
Let u(m) be the arithmetical function with 2
u(d) = to2.
(9.17)
d\m
By Mobius’s inversion (1.28), we can verify that 0 < u(t) < t2.
(9.18)
We can now write (9.15) as Itt2 2
dM(Qld) X m Q l f ) d ~ y - * X Mt) < X /< Q
tf
« (0 (
2
\ 4 f r S o a ,)
d -'M (Q ld )y. 1
(9.19)
1.9
FR A N E L’S THEOREM
39
Squaring (9.11) and substituting (9.12) and (9.19), we have 2 W ) I 2 < Q2 2 «(*)( r=1
'
t^ Q
2
c i-m (Q id )Y + F iog *Q
2
< Q 2l H
'
d
KO v d^Q
d -i M ( Q l d ) Y + Q * l o g * Q .
'
d = 0(m od t)
(9.20)
The leading term on the right of (9.20) is essentially the sum involving Mobius functions o f Franel’s identity. The inequality from (9.8) to M (x ) is less involved. Since, by (3.19), 2 * ea(a) = c9(l) = rfq),
(9.21)
am odg
we have Now
M (Q ) = 2 e(fr)-
(9-22)
r—1
\ e (f^ e {r lF ) \^ 2 n \ f - r l F \
< 2 ^ 1 E ( f r)\. Since from (3.1), we have
2 ^(rjF) = 0
(9.23) (9.24)
r=l
|jf(<3)|< 2 ^ 2 TO O! r=l
< Q - i( z m
\r=i
r) \ f ,
/
(9-25)
and so a bound for (9.8) of the form o(Q 4) implies the prime-number theorem in the form (5.18).
853518
t)
P A R T II
T h e P r im e -N u m b e r T h e o r e m
10 A MODULAR
RELATION
Winnie-the-Pooh read the two notices very oarefully, first from left to right, and afterwards, in case he had missed some of it, from right to left.
I. 46 I n this second part we study £(s) and L(s, x) as functions o f the complex
variable s, and work towards the prime-number theorem. Our investiga tions are based on the functional equations for 'C(s) and for L(s, x). The first step is therefore to prove these. W e need a lemma from the theory of Fourier series. (Poisson’s summation formula.) Let Tc, I be integers. L e t f ( x )
Lem m a.
be a differentiable ftmction of a real variable tvith \ m \ < A on [k, I]. Then
«>
ki l
f m e(rn x)d x = i m + f ( } c + l ) + ...+ f ( k + l ~ l ) ^ f ( l c + l )
X m = -co
£
(10 .1 )
and moreover the -partial sums satisfy the inequality M
l,e + 1
f(x )e (m x )d x -y (k )-...-y (k + l) M
k
^ lA M -H o g M . ( 10 . 2 )
Proof. Equation (10.1) is additive on intervals: its truth for [/<:,r] and for j>, t] implies its truth for [k, t]; so we may suppose 1 = 1. By
2.10
A M ODULAR RE LA T IO N
41
a change of variable we may suppose also Jc = 0. We use the function H(x) of eqn (4.4): H (x)=
y M = H Z-t —27rim 10 m——co v m^O
if o < * < i, if x ~ 0 or I,
and in particular if 0 < x < 1 we have (4.6):
^ > + 2M. 27rim 5 —
7)1#0
Hence /
M
V2
M
H ( x ) Jr 2 (277im)_1e(m3;) da;
2 m -1 d a ;+
o
0
1'
da; 1I'M
M -1 log M .
(10.5)
We now have jif, } 2 - m
I f(x)e(mx) da; J
m#0 0 =
[/(%)
2
(277-iTO)_1e(?wa;)j J— jV'O®) 2
m^0
(27rim)_1e(OTa;) d*.
(10.6)
0
The first term on the right of eqn (10.6) is zero, since the terms for m and —m cancel, and by (10.5) the second is 1
J f ( x ) H ( x ) d x - { - 0 ( A M ^ 1log M ) . 0
(10.7)
The integral in (10,7) is 1
j / '( * ) ( * - * ) da = i / ( 0 ) + i / ( l ) ~ J /(* ) d®. o o
(10.8)
Combining (10.6), (10.7), and (10.8), we have M
}
2 f f{x)e(mx)
dx = | /( 0 ) + i /( l ) + 0 (^ ilf-1log Jf),
(10.9)
and letting M tend to infinity we have the case k = 0, 1 = 1 of (10.1). A general method of proving functional equations is to write the required function as an infinite series, apply Poissozi’s summation formula to a partial sum, and then let the length of the sum tend to infinity. This entails a change in the order of summation in a double infinite series. We could prove the functional equations for ’( (s) and
42
THE P R IM E -N U M B E R TH EOREM
2.10
L(s, x) directly by this method, but it is more troublesome to justify the interchange o f summations and more difficult to identify the functions that arise. We shall therefore prove the identity co
2
oo
exp{—(to-|-8)27t2:-1} = x i
m = — co
2
m= —
exj)( — i7i2Trx)e(m8),
(10.10)
co
where x is real and 0 s ' S <1 1. Rather than apply (10.2) directly, we vise eqn (10.9), so that M
}
2
| exp{—(r + i+ S )2^ - 1} di
-M
J
=
| e x p { -(r + 8 )2} + i e x p { - ( r + l + § ) 2} + + 0[ne_1exp{— ( r - ^ S y n x ^ j M ^ l o g Jf],
(10.11)
and the double series 00
OO
2
2
m = — co
e x i ^ { t + d f n x ^ } e { m t ) At
(10.12)
r = — co ^
converges and is equal to co
2 y =
e x p { - ( r + 8 ) 277a;-1}.
(10.13)
— CO
A typical term in (10.12) is CO
J exp{— (t-\-i)‘lTTX^1Jr 2nimt} At — CO CO
=
J exp{— TTXZ2Jr2nT\m(xz— 8)}* dz, — co co
= icexp{— nxm?— 2mmZ) J e x p {— rrx(z— im)zJAz, —
(10.14)
CO
where we have put t-\-8 = xz. The integral in (10.14) can naturally be regarded as the contour integral of e~z2 along the line Ixnz = to. To evaluate the integral, we can move the line o f integration to the real axis, when (10.14) becomes CO
a exp(—■ rrxm2— 2mm§)
J exp(—irxt2) At — CO
= C77~sa;=exp(—nxm2— 2irim§),
(10.15)
co
where
c=
J e x p ( — t2) At.
(10.16)
A M ODULAR R E LA T IO N
2.10
43
I f we combine (10.15) and (10.12) we have proved that CO
CO
2 ?•= —
exp{— (»'-)- 8)27rai-1} = CO
j 7)1— —
C7r~%*exp(—7rm2x)e(mS).
(10.17)
co
Putting S = 0, x = 1 verifies that c = ttI, (10.18) and we have proved (10.10). Equation (10.17) should not be let pass without some comment. Let us put m ®(w) = 2 exp(7rmaai) (10.19) m= —oo for values of to for which the series in (10.19) converges, that is, for complex to with positive real part. Then we have 6(to + 2 ) =
8 ( w ),
(1 0 .2 0 )
and (10.10) with 8 = 0 gives us 02( - l / « ) = to82{to),
(10.21)
provided w is pure imaginary. However, since eqn (10.21) holds along the imaginary axis, the two sides o f (10.21) have the same derivatives at points iy with y > 0, and, since power-series expansions of regular functions are unique, (10.21) must hold whenever 8(to) and 6( — l/to) are both defined, which is whenever the imaginary part of oj is positive. From eqns (10.20) and (10.21) we see that 04H dw
(10.22)
is invariant under the group of transformations o f the upper half o f the complex plane generated by co -> w + 2 and w -> — 1/to. We are now in the realm of the elliptic modular functions. A modular function is one that is invariant under the group of transformations generated by a> ->oj-|-1 and oj -> — 1/co or under a subgroup o f finite index in this group. The derivatives of a modular function are not invariant under these transformations, since dw itself is not invariant; functions that satisfy the transformation law for a power of a derivative o f a modular function are called modular forms. The name ‘ elliptic modular functions’ arises as follows. The periods o f an elliptic function form a free Abelian group on two generators tox and to2. A modular form corresponds to a function of two complex variables and w2 which is homogeneous and whose value does not change when we replace ojl and to2 by another pair of generators o f the same free Abelian group (or, more generally, which takes a finite set of different values when toJ
44
THE P R IM E -N U M B E R TH EOREM
and a)2 are generators of the same Abelian group). Here Thus we may consider f{a>i,co2) = io^e^iojcox),
w ith
and
/(o j
2, —
OJjJ
=
U)2~1 02( — W j/o jg ) =
2.10 co
= w2/wi(10.23)
162(co2/co1)
= / K . w8) /(oij, a)2+2co1) — cox 102(co2/a)1-)-2) = /(aij, co2).
(10.24) (10.25)
11 THE
FUNCTIONAL
EQUATIONS
When he awoke in the morning, the first thing he saw was Tigger, sitting in front of the glass and looking at himself. ‘ H a llo !’ said Pooh. ‘ H allo! ’ said Tigger. ‘ I ’ve found somebody just like me. I thought I was the only one of them.’
II. 21
W e return to £(s) and L ( s , x )• The definition CO
r (¥) = / e - v y ^ d y ,
(11.1)
0 valid for cr > 0 (we recall s — cr+itf, cr and t real), becomes A i s)
CO
=
J
e~~m°7rxx is~1
da;
(H -2)
o
when we write y = -nmPx. Summation over m gives co
=
7 t - }‘ sr ( \ s ) i ( s )
«>
2
q -™ 2itxx Is - 1
da;.
( 11 .3 )
m= 1 0
Since the sum and integral in eqn (11.3) each converge absolutely, we can rearrange the right-hand side of (11.3) as co
CO
J 2
o
dx =
1
J 4 (0 (i® )-l)* * « -1 da;,
(11.4)
o
where 6(a>) is the function of eqn (10.19). Next we write
1
«>
| i ( 0 (i» )-l) ® * » -1 d a != J (0(i/f) — l)$ -ls-1 dt
0
(11.5)
1
and use eqn (10.10) to put (11.5) into the form CO
co
f ^ ( ^ ( i t J -lJ H 8- 1 dt =
f I0(it ) t ~ ^ i dt - K 2 S - 1)
i
i CO
= J 4 ( 0 ( i i ) - l ) H » - * d i - s - 1- ( l - a ) - 1.
(11.6)
46
THE P R IM E -N U M B E R TH EOREM
2.11
The left-hand side of (11.3) has now been expressed as CO
J £(0(ia;) — l ) ( » J « - i + a r + ' - * ) d® — { « ( i — a) } - i
(1 1 .7 )
i
The expression (11.7) was obtained under the assumption a > 1, but, Since
0 (i® )-l< e -™
(1 1 .8 )
for x > 1, the integral in (11.7) converges for all complex s. Since r(|s) is a known function, and ( m s ) ) - 1 is integral (single-valued and regular over the whole s-plane), we can take (11.3) with (11.7) as the CO
definition of £(s), knowing that X m ~s agrees with our new definition i
when the series converges. We have now continued l(s) over the whole plane. Further, (11.7) is unchanged when we replace s by 1—s, so that ^ r ( i s ) U s ) = t ^ - ^ T ( ! - 1 S) £ ( 1 - S),
(11.9)
the promised functional equation. Since 'M 1 I2
= 21^7r^T(s)cosis7r,
(11.10)
2 °)
an alternative form o f (11.9) is £(1—s) = 21- s7T~sr(s)cos %stt t,(s).
(11-11)
We now list some properties of F(s) (see for example Jeffreys and Jeffreys 1962, Chapter 15). The product 1 r(s+ i)
where y is tbe constant of (2.5), converges for all s, and defines F(s) as a function tliat is never zero and has simple poles at 0, — 1, — 2,.... Using this information in (11.7) we see that the pole of (11.7) at 1 comes from £(s), the pole at 0 from -T(^s), and that £(s) must have zeros at s = — 2 , —4,..., to cancel the other poles of F(^s). From eqn (11.12), T (s + 1 ) = sF(s),
and
Jrr( 1-)—s)_Z^(1—s) =
its cosec
(11.13)
77s,
(11.14)
where we have used the product formula for sin 7rs. We can verify eqn ( 11. 10) by showing that the ratio of the two sides is a constant. Equation (11.1) is obtained by evaluation of the limit of N
C *10+1( 1—t/N )N dt
(11.15)
2.11
THE F U N C TIO N A L E Q U A TIO N S
47
in two ways as N tends to infinity. We can also obtain from (11.12) the asymptotic formulae log ics) = (s -P o g s -s + llo g 2 7 r + 0 (l/| s | ), and
! » / / »
= lo g s+0(l/|s|)
(11.16) (11.17)
which hold as |s| co uniformly in any angle —7i+S < args < tt— S for any 8 > 0. Next we consider an L-function L(s, x) with x a proper character mod#. There are two cases. I f x(~~l) is 1> we argue as above up to 00
f
Tr-*sqissr(^ s)L(s, x) =
x(m)e~m*n!>:lQ
a;is-1 2
o
m=1
00 =
± j
(11.18)
x i ^ '? (x ,x )d x ,
0 co
where
2 x ( m ) e ~m*’nx^a'
( p ( x ,x ) =
(11.19)
— CO
W e approach cp(.T, x ) through (10.10): CO
CO
2 e -(» n + 8 )M * =
xh 2 e ~ mS7r-r e ( m S ) .
— co
( 1 1 .2 0 )
— co
W e put S = cijq and use eqn (3.8): x (m)T(x)
=
2*
(H.21)
x ( a )ea (a m )>
am o d g
so that
r(xMx>x)
=* 2* r n 2 co = 2* x{a)(
m = — co
am o d g
m ~ — co
which we may rearrange as co
2
co
2*
= (?/*)*
— co a m od g
=
2
?•= — oo
x ( r ) ^ ,m
(q lx M l/ x ,x ).
(H .23)
This will play the part of the modular relation (10.10). As before, we split up the range of integration in (11.18) and find that 1
CO
J x ^ c p ix , x) da; = J ^ is- 1(r(x))“ 1(g<)i(^ x) da;.
(11.24)
48
THE P R IM E -N U M B E R TH E O R E M
2.11
The analogue of (11.7) is now seen to be co
co
TT~lsqisr(^s)L{s, x ) = i J
x)
~
1
x~^-^(x, x) d a .
1
(11.25)
As before, the right-hand side of (11.25) converges for all s, so that L(s, x) has an analytic continuation over the whole plane, with no singularities. Moreover, L ( s ,x ) must have zeros at 0, —2, —4,... to cancel the poles of F(\s). We proceed to deduce the functional equation. We have t (x )
=
2
mmodg
X(m)eq(m ) =
2
mmodg
x ( - ™ ) e g{— m)
(11.26)
= r(x),
since it was assumed that x( —1) = 1. By eqn (3,14), since x is proper g*/r(x) = r ( x ) l r .
m0dq’
(11.27)
We now see that the right-hand side of (11.25) is r(x)g,_i times the corresponding expression with s replaced by 1-—s and x by x> which gives the functional equation ir-bqi*r($s)L(8,x) = T(x)q~lTT-l+*sq ^ sr ( ± ~ ± s ) L ( l - s , jf).
(11.28) We now consider characters xim) proper modg with %( —1) = — 1. Since we want to consider a sum from —co to co, we use mx(m) in place o f x(m). Writing s + 1 for s in (11.2), we have 7T-Ks+I)^i(s+I)p (i(s-|-I))i(s, x) =
”
OD
2 me-”'!“ % !s- “ da: o m=1 co
co
where
p(x, x) —
= iJ p (a ? ,X)®**-*da!, 0
2
(11.29) (11.30)
— co
We find a functional equation for p(x, x) by differentiating (10.10) with respect to 8. We get t (x )p (x >X)
= i ^ x - i p ( l f x , x).
(11.31)
Arguing as before, we find w-*«-tgi»+ *r(i(*+ l))i(s,x) CO
= i | />(*, x ) ^ d x + ^ ) f
CO
P(X>x)*-** d®.
(11.32)
THE F U N C TIO N A L E Q U A TIO N S
2 .II
49
Again, the integrals on the right of eqn (11.32) converge for all s, so that L (s ,x ) has an analytic continuation; it must have zeros at —1, —3, — 5,... to cancel the poles of r (| (s + l)), and satisfies the functional equation „is
lq i
i s r ( l — % 8)L(l— s,x) =
^ -« s+1¥ (s+1)r(| (S+ l ) ) L ( S, x).
(11.33)
To check this, we note that when %( — 1) = — 1 (11.34) = —^(x)There is also an analytic continuation of L(s, y) when % mod q is not proper. I f Xi proper m o d / induces % modg, then for cr > 1 t (x )
(11.35)
when we write m = dr. The sum over r in (11.35) is L(s, ^i), which has an analytic continuation since Xi m o d / is proper, and the sum over d is defined for all /. The corresponding functional equation for L(s, x) contains the sum over d explicitly. We shall not need this case again. A number of proofs of the functional equation can be found in Chapter 2 of Titchmarsh (1951).
12 HAD AM ARD ’S PRODUCT
FORMULA
Suddenly Christopher Robin began to tell Pooh about some of the things: People called Kings and Queens and something called Factors. II. 174
In proving the prime-number theorem, Hadamard studied integral functions of finite order, that is, functions f(s ) regular over the whole
plane, with
log|/(s)| < |«|^
(12.1)
for some constant A , as |s| -s*oo. The order o f/(s ) is the lower bound o f those A for which an inequality o f the form (12.1) holds. Hadamard showed that an integral function o f finite order can be written as an infinite product containing a factor s — p corresponding to each zero p o f the function. This generalizes the theorem that a polynomial can be written as a product of linear factors. Weierstrass’ s definition (11.12) of the gamma function is an example. The product is especially simple w h en /(s) has order at most unity. The order of l/r(s-\-l) is unity, from (11.16). We shall obtain the product formulae for £(s) an(l £(s>x) g iv e n b y and
m
= 8(l-8)n-*r(i8)Z(a)
£(«,*) = ( q M ^ r { U s + a ) ) L ( s , x ),
( 12.2) (12.3)
where % is proper modg and a = 0 or 1 accoi’ding to the relation %( — 1) = ( — 1)“. Note that (11.9) is just the assertion that £(1—s) is equal to £(s), and (11.28) or (11.33) implies that lf(l-« .X )l =
\£(S,X)\-
(12.4)
First we show that |(s, x) has order one. By eqn (5.3), if a > 0, L(s, x ) =
f j
2 x (m) dx.
(12.5)
By Polya’s theorem (4.2), the sum over m is bounded, and thus 00 !-£(*> X)I < 2 ^ ogc| J Is la -^ d o : < a-i|s|gUogg, (12.6)
2.12
H A D A M A R D ’S PRODUCT FORMULA
61
and for cr ^ § we have i o g l ^ x ) !
(12.7)
and we have
logics, X)( < ls llog?l«l (12'8) uniformly in q and in a From (12.4), inequality (12.8) is also true for a ^ -J. To bound £(s) we recall eqn (5.14): co
(1 -2 1 -)£ (S) = 2 { - v r - h n , - , (12.9) m=1 valid for a > 0. As in the derivation of (12.7) from (12.5), we deduce that the right-hand side of (12.9) is |s| for cr 5; Now when a ^ ■§ w e lm v e
( 1 — s ) / ( l — 2 i- s ) <
and so
|a |,
(1 2 .1 0 )
log|(l—s)£(s)| <^log|s| + l,
(12.11)
and Stirling’ s formula gives logics)I < Is l(log|«| + 1)
( 12.12)
uniformly in cr Js Since £ (l—s) = £(s), (12.12) is also valid for cr i. We have now shown that £(s) and £(s, %) have order at most one. I f s is real and positive then (12.7) and (12.12) are best possible, by Stirling’s formula, so that the order of £(s) and £(s , x) exactly unity. We now discuss the zeros p of the functions £(s) and |(s, %). First we prove a lemma. Lem m a.
(Jensen’s formula.)
Let f( s ) be a function of the complex
variable s = rcis# which is regular in |s| ^ R with no zeros on |s| = R, fo r which f ( 0 ) — 1. Then 2
w
B
(2-rr)-1 J log|/(i?cis0)| dd — J r_1iV(>’) dr,
o
(12.13)
o
where N(r) is the number of zeros o f f(s ) inside the circle G(r): |s| = r. Proof. We can write the left-hand side of (12.13) as 27T R
27T R
(2tt)-1 J JRe(d/ / / ) dr d 0 = (Zir)-1J
0 i-=0
J R e ® cis 6 drdB
0 r=0 R
27T 27T
= Re f i f W J
»'=o
and the inner integral is 2niN(r).
d
2niJ f(s) 6=0
A
r’
(12.14)
>
THE P R IM E -N U M B E R TH EO R EM
52
2.12
Now £(0) = 1, so we can apply (12.13) to £(s) at once. We shall prove later that £(0, x) is non-zero, so that (12.13) can be applied to £(s, x)/£(0, x). To avoid a circular argument, we choose a 8 for which £(S, x) is non-zero and apply Jensen’s formula to g(s, x)/£(&> x)(12.7) or (12,12) we have B J r~xN(r) dr < ^ .B l o g B , (12.15) o and as T oo N (T ) T log T. (12.16) Here, N ( T ) is the number o f zeros o f £(s) or o f £(s, x) with |s| ^ T. When we examine the formulae (12.2) and (12.3) for £(s) and |(s, x), we see that any zero of £(s, x) must be a zero of L(s, x), and similarly for £(s ). The converse is not true, because L(s, x) has extra zeros at negative integer values to cancel the poles o f the gamma function in (12.3). I f s = o-\-rl with a > I, Euler’s product formula ^ . X ) = = n { l - X ( ^ - S} - 1
(12.17)
p
converges absolutely and so is non-zero. By the functional equation, £(s, x) is therefore non-zero for cr < 0, since g(s, x) is non-zero for a > 1. Thus all zeros p = /3+iy o f £(s, x) have 0 ^ ^ 1, and the same is true for £(s) by a similar argument. Riemann’s hypothesis is that /3is always -J. Riemann stated the hypothesis for £(s), but it is difficult to conceive a proof o f the hypothesis for £(s) that would not generalize to £(s, x). We shall prove later that 0 < j8 < 1: this statement is equivalent to the prime-number theorem in the form (5.17) in the sense that each can be derived from the other. For later use we now prove a result more precise than (12.16). Lemma. The number o f zeros p = /3-f-iy of £(s, x) in the rectangle B , 0 < j8 < l, is at most
T ^ ys^ T + 1 ,
(12.18)
^log(g(|y|-fe)),
(12.19)
< l o g ( m + e).
(12.20)
and of i(s) in B is at most
Proof. Let s0 be the point 2 + i(T + | ). Then
|£(*o>x)l > i - i - i - A - - > i We apply Jensen’s formula (12.13) with B = 3 and
( 12 .21 )
/(« ) = £ ( « - * „ , x)/£(* 0.X). ( 12.22) By (12.7) and Stirling’s formula (11.16) the left-hand side o f eqn (12.13) is < l o g g+log(|T|+e).
(12.23)
2.13
H A D A M A R D ’S PRODUCT FORMULA
53
The circle radius 3 and centre s0 clears the box B by a distance at least \, so that, if N is the number of zeros required, the right-hand side of eqn (12.13) is
> t flo g 6 /5 .
(12.24)
This proves (12,19), and (12.20) follows similarly. We shall not need the following more accurate formula. I f T ^ e, the number of zeros of £(s, x) with 0 -
(12.25)
from which (12.19) and (12.16) follow readily. We can now prove the product formula. Let f (s) be £(s, x) or £(s), and let p run through all zeros of f(s). By (12.19) or (12.20) the series 2 Iph2 converges, and so therefore does the product
IT ( ! —s/p)exP(«//>)> (12.26) P# o where if 0 is a zero we add a factor s at the beginning. P(s) is a regular function with the same zeros as f(s ) . We should like f(s)jP (s) to be a constant or some other simple function. Certainly p (s) =
g(s) = log {f(s)lP(s))
(12.27)
can be defined to be single-valued and regular over the whole s-plane. We shall prove that g{g) = A + B s _ (12 28) By (12.19) or (12.20), there is a sequence of B tending to infinity with i?-|p| > ( l o g i ? ) - i
(12.29)
for each zero p. W e want a lower bound for log P(s) on the circle |s] = B . Now
2
-
log|(l—«//>)exp(«/p)| < B
0<\p\
2
0<|p\
l/>l_1 < ^ l o g 2E, (12.30)
by (12.19) or (12.20), and -
V
a)exp(s/p)| log|(1 —s/p)exp(s/ p) | <
T
(1+logIogU )
B log B loglog B ,
(12.31)
by (12.29). Finally -
2
lpl>2B
log'l(1—*(p)exp(s//>))| < B 2
2
lp|>2B
l/>l~2 < R ^ o g B .
(12.32)
On such a circle we have log|/(5)/p (5)l < Rlog^B.
(12.33)
The left-hand side of (12.33) is the real part of g(s). I f we write s = B cis d,
T H E P R IM E -N U M B E R T H E O R E M
54
2.12
the power-series expansion of g(s) makes g(R cis 9) a Fourier series in 6: g(R cis 6) =
so that
B,eg(Rcmd) =
and
ra(n)Rn =
(12.34)
(a(n)Jr ib(n))Rn cisnd> R n(a(n)cosn6— b(n)sinnd)
(12.35)
2tt f cos(—?i0)Reg{R cis 6) |dd.
(12.36)
Hence *7/
J |Re<7(i?cis 0) d0
\a(n)\Bn
o 27T
(\Reg(R cis d j l + K e g i R cis 0) - a(0)) dd 27T
1+
J max{R>e g(R cis 6), 0} dd o
< 1 -f RlogZR.
(12.37)
Since (12.37) holds for an infinite sequence o f R, a(2), tt(3),... must be zero, and similarly so must 6(2), 6(3),..., and we have proved eqn (12.28). W e have therefore £(s) = eBs JJ (1—s//>)exp(s/p),
(12.38)
p C(x)eB{x)s I I ( ! — slp)exp(slp), (12.39) p with the modification mentioned above if p = 0 occurs as a zero. These i(s, x) =
products converge for all s. By taking logs and differentiating, we get ^ 1 = B ----- j— + i l o g 2n — I F 'j f S + 1) + y ( ~ — h -V £(s) s— 1 2 r ( 2s + l ) ^ \S~ P pj
(12.40)
and (S>X) _ R/,,\
110" ^
l-^ '(2 (S+ ffl)) | X ' / 1
I
fl^ 41)
where p runs over all zeros of ((s) or L(s, x) that do not coincide with the gamma-function poles. We can substitute the equation
- l m
= ir +
(12'42)
which follows from the corresponding product formula (11.13) for F(s), and obtain a sum over all zeros (except p — 0) of £(s) or L(s, x). It can be shown that
= I0g 2 + i l o g w - l - i y ,
(12.43)
but no simple expression for G(x) and B(x) is known in general.
13 ZEROS
O F f ( s)
You remember how he discovered the North P ole; well, he was so proud of this that he asked Christopher Robin if there were any other Poles such as a Bear of Little Brain might discover.
I. 131 W e saw in the last chapter that i(s) and g(s, x) have 110 zeros p — /3+iy with /3 > 1. The prime-number theorem corresponds to the fact that no zeros of £(5) (these include the zeros o f £(s)) have j8 = 1. All the direct proofs that j8 < 1 at a zero are based on the following argument. The function —£'(s)/£(s) has poles of residue 1 at the poles o f £(s) and poles of residue — 1 at the zeros. Now
p
2 log p i l —p - * ) - 1 (13.1) where A(m) is the function o f eqn (1.31). At s = 1, the series diverges to + 00, corresponding to the pole of residue -(-1. Now, if 1 + iy were a zero of £(s), we should expect the series in (13.1) to diverge to —co, and the partial sums to be as large as those for the case s = 1, but negative. To achieve these, the numbers to-1? must be predominantly near — 1. The values of to2i>/ are therefore predominantly near + 1 , and there is a pole at l + 2iy with residue -(-1) aild so a simple pole o f £(s) at s = l + 2iy, which we know does not occur. To make this argument rigorous, we use (13.1) with s = a + ii where a > 1, so that the series converges. For all real 9,
Since
3 + 4cos0 + cos 29 = 2(l + cos0)2 ^ 0. 00 Re(£'(s)/£(«)) = 2 A(m )m ~aco&(it\ogm),
(13.3) (13.4)
we have 853518 X
(13.2)
E
56
TH E P R IM E -N U M B E R TH EO R EM
2.13
We now make cx+if tend to a zero j8+iy. Since £(s) has a pole at 1, there is a circle centre 1 and some radius r, within which £(s) is non-zero. (Calculation shows that r = 3 has this property.) I f we suppose that j3 > 1 - f r ,
(13.5)
then |y| ^ § r, and so is bounded away from zero. In eqn (12.40),
t'W m
P,
1
‘ W l )
V / 1 ,1)
= ~ B + — i " ll0 8 2 ’ - i T ( i i + i ) - 2 , 1 ^ + W
(13'6)
we shall assume 1 < a < 2, \t\ ^ fr > 0. Here the sum is over zeros p o f £(s), not over all zeros o f £(s), and, since s — p and p have positive real part, we have / -, ^ Re _ i _ + ± > 0 (13.7) \S~~-P
PI
whenever a > 1 and p = /3+iy has 0 ^ / 3 ^ 1. By (11.17) the term in ^ ( i « + l ) is
(13.8)
At s = cr-j-iy we omit all terms in (13.6) except those from the particular zero with which we are concerned; by (12.7) this gives us the inequality _£ '(cr+ iy)/£(ff+ i y) < - {a- p ) - i + 0 ( l o g ( \ y \ + e ) ) . Similarly,
_ £ '( ff+2iy)/£(a+2iy) < 0(log(|y|+e)).
(13.9) (13.10)
When we substitute (13.8), (13.9), and (13.10) into (13.4), we have 4 ( a - j 8 ) - i - 3 ( a - l ) - i < 0(log(|y|+c)),
(13.11)
valid as a -> 1 from the right. By giving a a suitable value, we see that j3 < 1—,— log(|y|+e)
(13. 12)
where c is an absolute constant. I f we constrain c to be less than f r then, by (13.5), (13.12) also holds when |y| < |r and is thus valid for all zeros p = /3+iy o f £(s). For twenty years, (13.12) remained the best known upper bound; in 1922 J. E. Littlewood proved that c2loglog(|y|+e2)
P < 1—
k O T + ^ T ’
no
ion
(13J3)
2.13
ZE R O S 03?
£(s)
57
for some constant c2 (Titchmarsh 1951, theorem 5.17). The latest result is C3(e) ft < 1 —,— nTT T IVfTe’ (13' 14) log(|y|+e)^ where the constant c3 depends on e. This was proved in 1958 by Korobov and by I. M. Vinogradov independently. Intermediate improvements on (13.13) used the intricate methods o f I. M. Vinogradov (1954).
14 ZEROS
O F |(s, x )
‘ W hat do you think you’ll answer ? ’ ‘ I shall have to wait till I catch up with it, ’ said Winniethe-Pooh. I. 34
In this chapter and the next we prove results like (13.12) for the zeros p = /3+iy o f L(s, x), with uniform constants in the upper bounds. We actually work with £(s, %), where % is proper mod#; the zeros of £(s, x) are those zeros of L(s, x) that are not at negative integers and so are not cancelled by gamma-function poles. We use the product formula in the differentiated form (12.41), L ( s ,x )
TT
Z-4\s— p
2 i (!(« + « ))
pj
where the sum is over zeros p of £(s, x)> the term 1/p being omitted if p = 0. We shall assume throughout the chapter that s = a-\-it with 1 < (7 < f. The first complication is the elimination o f B(x) from eqn (14.1). We subtract from (14.1) its value at s — 2, noting that CO
\L'(2, x )\I\L(2>x)\ <
2 (14.2) m=1 which is bounded independently o f x • Estimating the gamma-function term from (11.17), we have
-
K
e
tM
5 5
~
2 R6( ^ , - ^ ) +0(log(|,1+e))’
( 1 4 '3 )
where the term yo = 0, if it occurs, is now included in the sum. We now note that
^ ^ - f ) - 1 = 2 ( 2 ^ ) \ 2 ~ p \ -2 < lo g ?, p
p l(t) = log{q(\t\+e)},
by (12.19). Writing we have
—Be ^
^ < — "V Re —-— \-0(l(t))
JjyS} ^ j
S
’p
for 1 < a ^ |, the implied constant being absolute.
(14.4) (14.5) (14.6)
2.14
ZE R O S OF
{{s,
59
x )
When we apply (13.2) we obtain the relation _ 3
o) - 4 Re 7; ' (<7 : ] 1’ y) - Re L 'ja+ 2^ X2) L i°>Xo)
L ( a Jr it, x )
p,
( 14.7)
L (a -\ -2it,x)
validforcr > 1, as the analogue of (13.4), Here, Xo is the trivial character modg, whose value xim ) is 1 when (m, q) = 1 and 0 otherwise, and x 2 is the character whose value at m is (x(m)}2. Although we have supposed x to be proper mod q, x 2 might be trivial and certainly need not be proper mod <7. The trivial character xo is not proper mod <7. However, if Xi proper mod/induces x 2 mod 5, th e n i'(s, Xz)l^is >X2) and _Z7(s, Xi)!L{s, X-d differ only by terms involving powers of those primes that divide q but n o t /. For a > 1, these terms give at most
m=2? p\q
_
p \Xf
^ 2 los p K p — 1 ) < log ? p\q
( u -8)
The inequality (14.8) applies also f o r / = 1, X2 = Xoconclude that (14.6) is valid for any non-trivial xm odg, possibly with a different O-constant, and that _ R e ^ ^
® Z jL _ R e V _ i — |- 0 (l (t)) (14.9) Is 1| for the trivial character Xo mod q. I f xa is non-trivial, substitution of (14.6) and (14.9) into (14.7) gives L(s,Xo)
implying that
4(u—j8)_1 < 3(ct— 1 )-!+ O(Z(0),
(14-10)
/? < 1 -cJ H y )
(14.11)
for some absolute constant cx when we choose a appropriately. I f x 2 is trivial, then 4(ct—jS) - 1 < 3 ( a - l ) - 1+ ( a - l ) / { ( a - l ) 2+ 4 y 2}+0{Z(y)}, (14.12) which is consistent with j8 = 1 when a -» 1. However, if \y\ > cjl(y)
(14.13)
for some positive c2, then by choice of a in (14.12) we can show that 0 < 1- c 8/Z(y),
(14.14)
with a smaller absolute constant c3. We have now shown that either (14.14) is true or
|y| < 8/log?,
(14.15)
where § is an absolute constant. The absolute constant c3 in (14.14) depends on the choice of 8 in (14.15). When (14.15) is satisfied with
THE P RIM E -N U M B E R THEOREM
60
2.14
y ^ 0 we can still deduce an upper bound for f3, but it is very close to unity, and tends to 1 as y 0. The third and greatest difficulty is to deal with zeros close to unity when x2 is trivial. First we show that there is at most one. We have _ R e ^ 4
L(a, x)
= Re y ^ A*
^
- )
co
^ — 2 A(m )m ~a i
> — {a— 1)-X+ 0 (1 ).
(14.16)
I f px = /Sx-(-iYi and p2 = j82+ iy 2 are two zeros satisfying (14.15), then —Re y ^
L(a,x)
^ —Re —------- -Re —------\- 0(logq) a— Pl
a
^ “ (cr-jSi)*+yf ~
p2
+ 0(l0g q)
< (7 -j3 1)*+ 8 ?iogg)-a+ '9(l0g^)’
(14J7)
if /32 ^ j8x ^ 1—S/(logg). I f 8 is small enough, this implies that Pi ^ 1
cJ l(Yi)-
(14.18)
Clearly we can choose c4 ^ c3, and so (14.18) is true for every zero fti+iyi ° f x) except (possibly) p2, and the possible exception p2 occurs only if ^(m) is always real, so that x %is trivial. Since p2 is also a zero when L(s, x) has real coefficients, if p2 fails to satisfy (14.18) we conclude that p2 is real. We devote the next chapter to the case of an exceptional zero ft on the real axis.
15 THE
EXCEPTIONAL
ZERO
Piglet said that the best place would be somewhere where a Heffalump was, just before he fell into it, only about a foot farther on.
I. 57 Ik this chapter we consider real characters, that is, characters for which x(m) is always real and thus is trivial. As far as we know, the corre sponding -L-functions may have real zeros /3 with \ ^ /3 1. Just as before we saw that L (s ,x ) cannot have two zeros both close to 1, we shall now see that two functions L(s, x) corresponding to different proper characters cannot both have zeros close to I. Suppose Xi is proper m od?!, x% is proper mod<j2, and the corresponding i-functions vanish at j8j and jS2. In place of (13.2) we use (1+ Xi(m)) (1+ X2(™)) > 0,
(15.1)
which implies that I'(a)
L'(o, Xi)
L ' ( a , x 2)
L'{a,
Xss) ^
£(ct)
L{a,X\)
L (a>X2)
L(a>X 1 X2)
^15 2j
where X1 X2 denotes the character modg'1g'2 whose value at m is Xi(m )Xz(m )' When xi and Xz are different, the character X1 X2 non trivial, and (14.6) gives — L '(o > X iX 2 )m ‘J>XiXz) < ° ( l ° g W h ) ,
(15-3)
— L '(
(15.4)
and for L(a, xi)
and similarly for (a
and if
In place of (14.17) we have fti)-1~MCT fi%)~1
{a
I)-1
O ilogq ^),
(15.5)
^ /?2 then /32 at any rate satisfies the relation As < i-C iA logff^g),
(15.6)
where cx is an absolute constant. We deduce a uniform zero-free region.
THE P R IM E -N U M B E R THEOREM
62
2.15
By (15.6) and (14.18) there is a constant c2with the following property. Let Q > 1. Then no i-function formed with a character x mod q with q ^ Q has a zero p = /3-)-iy with p >
1—ca/log{Q(|y|+e)}
(15.7)
except possibly at a point J81 on the real axis, where L(s, x) has at worst a simple zero. All ^modr? with q ^ Q for which L (pv X ) — 0 are induced by the same real character. To prove the prime-number theorem for an arithmetic progression with common difference q, we need to know that neither £(s) nor any .L-function formed with a character x m o d q has a zero p = fi-j-iy with P = 1. The proof is simpler if we have p explicitly bounded away from unity. We have to deal only with the case x real, p real. One method is to interpret L( 1, x) as the density of ideals in a quadratic number field. This gives a very weak bound. We shall prove Siegel’s theorem. Theorem. For each e > 0 there is a constant c(e) such that if L Wi> X) = °> ivhere x is a character mod q then
^ < 1—c(e)
(15.8)
The constant c(e) in Siegel ’ s theorem is ineffective; that is, the proof does not enable us to calculate it. All previous constants in upper bounds, such as c2 in (15.7), have been ones we could calculate, given a table of values o f £(s) for |s| 3 and standard inequalities such as Stirling’ s formula. Following Estermann’s account (1948) of Siegel’s theorem, we con sider the function F(s) =
£(s)L(s, xi)L(s, Xz)L (s , Xi Xz)>
(15-9)
where Xi and Xz are real characters proper m o d ^ and mod <72 respec tively. By (15.1), the Dirichlet series CO
log F{s) =
J {i + Xi(m ) + Xz(m ) + XiXz(m )}A (m )m ^s
7)1—1
(15-10)
has non-negative coefficients; it converges for a > 1. For a > 1 we can take the exponential of (15.10): CO
F(s) =
2 m=1
(15.11)
with non-negative coefficients. F(s) has (at worst) a simple pole at s = 1 of residue
A = L (l, Xl)L (l, Xz)L(l, XlXz),
(15-12)
2.15
THE E X C E P T IO N A L ZERO
63
and no pole if A = 0. Moreover, F(s) has a power-series expansion, CO
(15.13)
F(s) = 2 b ( r ) ( 2 - 8y ,
where
b(r) =
(-l)'\ F «(2 )/r! (15.14)
In particular, b(0) = F ( 2), which is at least unity. The function F ( s ) - X / ( s - l ) = 2 {&(>') “ A}(2— 8)r
0
(15.15)
has no singularities, and so its power series converges everywhere. If A = 0, the series on the right of eqn (15.15) is positive on the negative real axis and represents F(s). Since F(s) is zero at 0, —2, —4,..., we conclude that A ^ 0. We now have < 1. To prove the more precise result (15.9), we use Cauchy’s formula, integrating round a circle, centre 2 and radius f, to obtain the coefficients b(r). On the circle, £(s) is bounded, and for the ^-functions we use (12.6), (15.16)
|£(«>x)l < c r “ 1|s|^logg,
which was proved from eqn (5.3) and Polya’s theorem, and is thus true for any non-trivial x mod?, proper or not. We have
< ( i Y i l l log V i M log q2)(qlq\ log qxq2)
(15.17)
< (t)r?l?2 1°g8?l?2>
the constant implied in the sign being absolute. We restrict ourselves to the range t9q ^ a < 1 and estimate the tail of the series (15.13) for F{s). CO
2 r =--R R+1
|6(»*)—A|(2—cr)*- < 2
< q 1qalog'q1qs 2 (tt)r
2?-}"1
< q 1q2logaq1q2{ ^ ) It,
(15.18)
64
T H E P R IM E -N U M B E R T H E O R E M
2.15
and thus (taking only the first term in the series for F(s)) F ( a ) ----- > 1 - A f ( 2 - Cr)’ - 0 ( g 1g2log3g1g'2(ii)iJ)
a— 1
o
> 1- A( (2 - a)5^1-1 ) /( 1 - a) - 0 {qxq2log3?! ga(ii)B). (15.19) We choose R so that the error term in (15.19) is less than possible with a choice of R with R = A log qx q2,
this is (15.20)
where A is an absolute constant. Now F (°)
> i -A ( 2 -a ) * /( l-( T )
> i —A(l—o-)-1exp{i?log(l + ( l —a))} >
— A ( l— a)^ 1e x p { i? ( l — cr)}
> i - £ A ( l - a ) - i ( ?1?a)^-">,
(15-21)
where B is another absolute constant. We now choose xt so that L(s, x z) has a real zero j8a with l - e f t A + l) < j8 2 <
1;
(15.22)
if the choice is impossible then (15.8) certainly holds. Since/(/3a) = 0, we have
^<
(15.23)
where the constant depends on the choice of ^2 and so on e. This is a lower bound for the product A of the three L-functions at 1. Since -£(l,X'a) is a known constant, we seek an upper bound far L ( l , x x Xt)From eqn (5.3), m L (S, X ) =
Let
f sa;-*-1 2
x M d:K-
r = gHogg,
(15.24) (15.25)
the upper bound in Polya’ s theorem. Then r
|L(«,X)| ^
J
oo
da; +
1
J \s\x~a-h' da;.
I f a ^ l - ( l o g q ) - 1, the first integral is < | s| ,a ild s°
(15.26)
r
jsjlogq and the second is
| i(«,x )l< l* | lo g 2 .
(15.27)
When we integrate round a circle radius i(lo g g )-1 to find L '(s ,x ), we have for
® > i-^ lo g ? )"1
(15.28)
the bound
\L'(a>x)\ < l o g 2q.
(15.29)
2.15
T H E E X C E P T I O N A L ZE R O
65
Hence, if L(s, xi) has a zero j8x in the range (15.28), A = L ( l , X i ) L ( 1 > Xi ) L ( l > Xi X z ) < 1° g ? i i (1.Xi)
< l o g q 1{ l - P 1)L,{a,Xl)
(15-30)
for some a in /3X < a < 1, and so by (15.23) and (15.29) we have 1 < ( l - / W (1-^ lo g 3
(15.31)
where the constants implied depend on xz and so on e. I f /3X does not satisfy (15.28), then (15.31) certainly holds, and Siegel’s theorem (15,8) follows from (15.31). '
16 THE
PRIM E-NUM BER
THEOREM
The clock slithered gently over the mantelpiece, collecting vases on the way. II. 135
WE can now prove the prime-number theorem in the form (5.17). Com bining (5.4) and (5.5), we have
f
J
277-1
if u
'0 ( u a{T\\ogu\)-x)
a + iT
-A s
S
cc—iT
•i + 0 ( a /T )
<
1,
if u = 1,
(16.1)
,l + 0 ( « o:(?, log?()~1) if u > 1.
On the line a — a, where a > 1, the series — Z'(s)lt{s) = 2
( 16.2 )
A(m)m-S
m= l
converges absolutely. We now suppose that x is an integer plus one-half. Then a-f-i T
J_
}
27rt J
a_ iT
„
^
Z-, rn=l
/
2
ms s
m
A(m) +
o k
\T
'
no
f
\
J -V log x m\ m a
Z-i m=1 1 a '
1
(16.3)
/
To estimate the series in the error term of (16.3), we first note that 2 A (m )< 2/,
m
(16.4)
by the sieve upper bound (8.19). Partial summation gives V A (m ) | V A (m ) ^ 1 Z., m a\\ogxjm\ Z-t a:“ log xlm a —l ’
(16.5)
In the remaining range for m we write on == x-\-^r, where r is an odd integer. Thus
ix < m < 2 x
— A ^ —-<^ ® -“ loga: T |loga:(l-f h'lx) h 1 x alog\xm\ & r=T-x 2
2x
a;-“ loga; 2 xlr )•=! ■ ^ ^ -“ log2®.
(16.6)
2.16
T H E P R IM E -N U M B E R T H E O R E M
67
We can now write the right-hand side o f eqn (16.3) as xa
x lo g 2x\
T (a —1) 1
( 16. 7)
T
where ^{x), as usual, is the sum function of the coefficients A(m) of
-mm*
\T
-i r F ig . 2. The contour G
We must now find an approximate value for the integral on the left of eqn (16.3). By (12.20), there is a value of T in each interval t < T < t+ 1
for which
i
^ n __m\~i \v\ — T\ > (log T)-
(16.8)
for each zero p = /3+iy of £(s)- With such a value of T we move the integral to the contour 0 consisting of C^: the line segment [a—i T, Gz: the line segment [ —
-in
iT , —| + i T \
Os : the line segment [—J + i ? 1, a-f- iT].
In eqn (12.40), ^ 1
= 1
we subtract the value at 2 + i t from that at
using the fact that
THE P R IM E -N U M B E R TH EOREM
68
2,16
£'(s)/£(s) is bounded on a = 2, where its Dirichlet series converges uniformly. £'(*) _.p m £(«)
1 r'(tts + i))
1 r (f+ ijQ
y
2 -r d + ii^ )
^ -^'(i('s+ l ) )
/ 1 ^
Vs — p
.1
\
2+ i i — p )
2—a (a + ii—p)(2-(-i—tp) (16.10) where we have used (12.20) and (16.8) on the sum over zeros, and (11.17) on the gamma-function terms. We can now estimate the integral round C. First -1 f S ^ d . 27ri m a
.
iog2y
:
T
x a do-
x a log2T j (T log x ),
(16.11)
and similarly for the integral along C3, whilst 1 f V(s)xs ds 27ri J £(s) s
: x ~ -l o g zT
J
ds s
o2
« _ ilog ST.
(16.12)
At this stage it is convenient to make the choice of a a =
l-f(log.'K)-1,
(16.13)
whereupon we have
1 r 277-i , £(«) 5 0
. *log 2T j-ai^log3! 1. T log x
(16.14)
The contour has been moved past several poles of the integrand. The residues of £'(s)/£(s) are + 1 at zeros of £(s), —1 at poles, and so the integral in (16.3) differs from that in (16.14) by xP
-
2 Iy \ < T
£'(0)
(16.15)
£(0)’
where the sum is over zeros p = /3+i y of £(s) with )3 > 0. These are all zeros o f £(s), and by inequality (13.12) each satisfies the relation (16.16)
2.16
THE P R IM E -N U M B E R TH EOREM
69
By (12.20), the sum over zeros in (16.15) is in modulus
2
rsSalogr < a ;1- slog2T ) where
w lyl<2r 21
8 = cJlogiT-^e).
(16.17) (16.18)
We have now shown that >p(x) = 3!+ o (* l° y f X + a;1" 8log2y j ,
(16.19)
and when we choose lo g ? 1 = (lo g * )* + 0 (l),
(16.20)
where the 0(1) is to allow us to find a T for which (16.8) holds, the error terms in (16.19) are bounded by expressions of the form * e x p {—c^loga:)4}
(16.21)
with different values of c. Hence there exists an absolute constant c2 such that
^
= a;_|_o|-;Cexp(—c2(loga:)*)}.
(16.22)
We can quickly deduce from (16.22) that 2 logp = » ;+ 0 { a e x p ( - c 8(loga;)J)},
(16.23)
p<x
and integration by parts gives X
A x) =
2 1 = f — + 0 { ® e x p ( - c 4(loga;)i )},
p<x
J
U
(16.24)
2
the classical form of the prime-number theorem. Clearly, only the O-constants in (16.22), (16.23), and (16.24) are affected when we drop the restriction that x be an integer plus one-half. We have approached the prime-number theorem by the classical route of Riemann’s functional equation and Hadamard’s product formula. Neither o f these is necessary for the proof o f eqn (16.22). There are elementary proofs of eqn (5.17), ones which do not use the complex variable at all (see for example Hardy and Wright 1960, Chapter 22). The elementary proofs are disappointing in that much extra effort is needed to prove the prime-number theorem with an explicit error term, rather than as an asymptotic equality. An analytic proof that does not use the functional equation is given by Titchmarsh (1951, Chapter 3).
17 THE
PRIM E-NUM BER
AN ARITHM ETIC
THEOREM
FOR
PROGRESSION
’P(x>x) = mX, A ('m)x(m)<x
L et
C17’ 1)
When x is proper mod q, we apply the method o f the last chajjter to obtain <"•*» ipl>« ° where <x is given by (16.13), and T and R are chosen so that ||y[-T| X l o g g T ) - 1,
(17.3)
IIH—i?| X l o g g ) - 1
(17.4)
for each zero p = jS+iy of |(s, x). By (12.19), we can suppose that A < R < J. The contour C consists of Gx: the line segment [a—iT, — f —i? 1], C2: the line segment [ —
\T, —| + i T],
C3: the line segment [ —| + i T, a + iT ], Ga: the circle centre 0 , radius R described negatively.
The circle Gi avoids the pole of x*/s and a possible Z-function zero at s = 0. As in (16.10), the inequality (17.5)
\L'(S, x )IL(S, x ) \ ^ \ o g \ T
holds on the contour G, the constant being absolute. Thus 1 f L ' {S’ X) X8ds 2771 , L ( s , x ) 8
(17.6)
a
We now treat the sum in (17.2). By (14.18), non-exceptional zeros satisfy the relation
ft ^
(17.7)
1— c J lo g q T
with an absolute constant cv I f q satisfies an inequality log
< c2(log i# ,
.
(17.8)
2.17
A R IT H M E T IC P R O G R E S S IO N
71
with c2 < 1, we can choose T so that (17.3) holds, and log q T = (loga;)J+ 0 ( l ) .
(17.9)
As in the last chapter, we deduce
‘/'(a, x)
f —3^7/?!+ 0 { x exp(—c3(log x )i )}
,0 {* e x p (—c3(loga:)*)}
if there is an exceptional zero Ji> (17.10) if not.
The value o f c3 depends on that of c2 in (17.8), but can be effectively calculated.
We can absorb the term *&//?! only if q satisfies q < logiVa;
(17.11)
for some N > 0. By Siegel’ s theorem with e = ^ N -1,
and
ft > 1—c4g~e > 1—c4(loga;)-^
(17.12)
a;A/ft <§; &exp{—c^logx)*}.
(17.13)
W e have now shown that the inequality \i/j { x ,
x)| < ®exp{— c6(loga:)*},
(17.14)
where c5 is an absolute constant, holds if (17.11) does, or if (17.8) holds and there is no exceptional zero. 853518X
THE P R IM E -N U M B E R TH EO REM
72
2.17
Our characters can at last abandon propriety. I f x niod q is induced b y Xi proper mod/ then <(>(x >Xi)-~>l)(x >x) = 2 logP 2
Xi(Pr)>
(17.15)
an expression whose modulus does not exceed lo g ^ f l 2 i> la
log ,t\ < log a; lo g ?. k>g'W
(17.16)
The estimate (17.16) also holds if x is trivial mod#, <jj(x, Xi) being tfi(x). This error term absorbs easily into that o f (17.14), which thus holds whenever % is non-trivial mod q. For the trivial character Xo mo(i q> (17.16) gives 1#^ Xo)
x \ < x exp(—c6(loga) *) + log x 2 1, vVi
and again the first error term will absorb the second. Let ? ,« )= 2 A(m). m
(17.17)
(17.18)
w = a(mod<2)
There are two cases. I f (a, q) is not unity, only the powers o f primes that divide q can occur in the sum (17.18). If, however, (a,q) = 1, we have from eqn (3.3) if>(x;q,a) =
2 Xia°d9
X) =
0 { * e x p ( - c 7(loga:)*)}, (17.19)
the prime-number theorem for the arithmetic progression with first term a and common difference q. We have proved (17.19) if (17.8) holds and no character m od? has an exceptional zero in its i-function, or if the stronger'condition (17.11) on q holds. The result (17.19) subject only to (17.11) is called the Siegel-Walfisch theorem. When (a,q) — 1 the arithmetic progression a, a+g,... contains infinitely many primes, but we have not proved that the first xjq terms include a prime unless (17.11) holds, so that x is very much larger than q. For a given q we may have to choose x larger still to ensure that the second term in (17.19) is smaller than the first, and t/i(x‘, q, a) is non-zero. Thus (17.19) is useless for numerical work, since we do not know the value o f c7. From the proof o f Siegel’ s theorem it follows that, if we knew one L-function with an exceptional zero, we should be able to find the c7 corresponding to one particular value of N in (17.11). I f no Z/-function has an excep tional zero, then (17.19) holds subject only to (17.8), and we can find c7.
PART
III
The Necessary Tools
18 A SURVEY
OF SIE VE S
It wasn’t what Christopher Robin expected, and the more he looked at it, the more he thought what a Brave and Clever Bear Pooh was. I. 140
S u p p o s e that we have a classification of the integers into certain
classes, and a system o f functions a from the classes to the complex numbers with the properties (i) and (ii) below, which generalize those that the functions , * = ea(am) , \ 0 \ ct(to) (18.1) possess for the classification of integers
to into
residue classes modg.
(i) Apart from the trivial function a for which a(m) is always 1, the sum (or integral as appropriate) of ct(to) over a complete set of classes is zero. (ii) I f to and n are in different classes there is at least one function a for which a (to) ^ a(n). Then a sequence m i of integers is uniformly distributed among the classes if and only if | °t% ) = o ( X |a(TOf)|)
(18.2)
for each a except the trivial one. This is known as W eyl’s criterion for uniform distribution. As an example, the prime numbers form a sequence that is uniformly distributed among those residue classes a mod q for which (a,q) = 1, in the sense that a proportion (l/
74
T H E N E C E S S A R Y TOOLS
3.18
deduced this from the fact that j/r(x, %) is of smaller order than x when X is a non-trivial character mod?. Franel’s theorem o f Chapter 9 fits this scheme. We were concerned there with the distribution of the F Farey fractions in the interval [0,1]. The appropriate functions a(m) are e(bfm), where b is an integer, and a calculation with Ramanujan sums shows that F
1
(18.3)
e(bfm) = 2 d M ( Q l d ) , d\b
in the notation of Chapter 9. Franel’s theorem relates a measure of the uniform distribution of Farey fractions in the interval to the mean square of the sums in eqn (18.3) for varying b. Many results in number theory rest on the uniform distribution of some sequence among certain classes. This is the final step in the proof of the prime-number theorem for arithmetic progressions. Often the uniformity result is needed at an intermediate stage, and in these cases it is usually sufficient to know (in a quantitative sense) that (18.2) holds most o f the time. An assertion that, for a given sequence m t , and set o f classes, eqn (18.2) holds for almost all functions cr is called a largesieve result. A common form of large-sieve result is that the mean square o f the left-hand side o f (18.2) is less than the mean square of the right. The sieve (8.10) of Chapter 8 is such a result with a(m) given by (18.1). In this chapter we shall also consider the characters x(m ) an(l the powers m~s as functions a(m). These correspond to distribution among reduced residue classes, and distribution of the sequence within intervals. The proof o f a large-sieve result depends on an upper bound for an average of some convenient auxiliary function. The simplest example is a result for all characters % to a fixed modulus q. Let u(m) be any complex coefficients and (18.4) In Chapter 7, we introduced the sum M +N
$ (“ ) =
Z
m=M + l
(18.5)
u(m)e(mtx);
and we obtain the result by comparing the identities 2
(18.6)
3.18
A S U R V E Y OF S I E V E S
75
wliicii follows from eqn (3.1), and iU+iV X mo clq
b moclg'
(18.7)
m.—M + 1
m = 6(modg)
which follows from eqn (3.3). The stun on the left of (18.6) can be bounded by the large sieve for exponential sums (7.8), and we have M+iV I \8X\‘ < r W s H x + o i r 1)) I M™)l2(18-8) X mod q
m —M + 1
When we restrict our attention to proper characters we can sum over q on the left of (18.8); by eqn (3.8) for % proper mod?, A x)x(m ) =
and we have
I*
am odg
(18.9)
x ( a ) e a( a m ) ,
M +N r(x)>s x =
2 2 * X ( a ) u { m ) e q{ a m ) ■m=M + 1 amodg
(18.10) * qn * • > «© -t-t-i/-»/! ''-L'■ a2 mod By (3.3) again we have
2
*
^modg
x{a)S X mod a
< ? (?) 2 ;l
8
(18.11)
am odg
where the asterisk on the sum over % indicates a restriction to proper characters. When % is proper m od? then |t ( x )|2 = ?, and we have y jL j
jl y ap(q) Z-*
2
g^Q TX-t/ ^m odg
g< 0
r
8
am odg M +N
< {N + o m )
2
m=M+l
k » * )is
(18.12)
by (8.10), that is, by (7.8) again. A third sieve result for characters is possible if the coefficients u(m) are 0 whenever to has any factor smaller than Q; for then (to,?) = 1, and (18.9) holds for all characters m od?. For example, if M = 0, and u(m) is 1 when to is a prime greater than Q and 0 otherwise, we have
2 2
a=SQ x mocla
«p(
< { N + 0 ( Q 2) ) X |m(to)|2 1 < (A ^+0(Q 2))( tx(iY) + 0(Q )).
(18.13)
76
3.18
T H E N E C E S S A R Y TOOLS
The left-hand side of (18.13) can be expanded as
f^Q
2 xmocl/
(18.14)
*
j s O fm o d /)
s=s Q
By eqn (8.16), the term / = 1 gives (77(iV))2l o g Q (l+ o (l)) ,
(18.15)
which is approximately half the right-hand side of (18.13) when Q is almost N$. I f for some % L(s, x) has an exceptional zero in the sense o f (15.7), the term in |^|2 can be as big as (18.15). This give sanother proof that ‘ at most one x has an exceptional zero’ , blit the range for q is not as large as in (15.7) when we work out the details. To obtain a sieve result for a(m) = m~a we return to first principles and use what we shall refer to as Gallagher’s first lemma (1967). Lem m a. Let f(x )
be a differentiable function of a real variable x.
Suppose that a^,..., xR are at least § apart and
Z . + i S < xr < Z a- J S for r = 1,..., R.
(18.10)
Then
it
Z I l/(® r )la < g J r=1
!/(•■») 12 da; + 1 J jf(x)j2 d x j i j \f'(x)\*dx X, W ' 'Xi (18.17)
Proof. We integrate the identity
!/(*,■) I2 = \f(x)\2~
Ax
(18.18)
j/(a;)|2 da;
over an interval of length S and obtain a,v+i8
J
S|/(a;,)|2 <
9,v+*8 V \f(x)\2 da; -f-
!’'f- id
J
Xr-iS
j - ^ \ f ( x ) \ 2 dx
dy
X,
x r+ tS
J
A
\f(x)\2 da da; -f-
Xr—
da;
da;.
(18.19)
i» r “ i S
By (18.16), these intervals are disjoint and subintervals of [X 1; X 2], and r
§ 2 I /M 2<
f l/(*)l2 da; + i § f |2/(a;)/'(a;)I da;,
and (18.17) follows from Cauchy’s inequality.
(18.20)
3.18
A S U R V E Y OF S I E V E S
77
To obtain large-sieve results from Gallagher’s lemma, we arrange the sum on the left o f eqn (18.2) to b e f ( x ) , different xr giving the different functions cr. For instance, for S(<x) given by (18.5) and ct(to) by (18.1), the points x r are the rationals ajq in their lowest terms, and Gallagher’s lemma gives M +N
s la
< (Q2+
Q amodg 1
ttN)
2
K(™)!2-
m =M +l
(18-21)
Like the relation (7.8), (18.21) is an approach to the conjecture (7.7). Whereas (7.8) gives N the conjectured coefficient unity, (18.21) obtains the conjectured coefficient of Q2 but not that of N , The application of Gallagher’s first lemma to m rs was made by Davenport. Let N f(t) = 2 (18.22) ?n=1 where a is fixed. The first integral on the right of (18.22) is
n
2'i
Tt
T2
V i ^ l L 2 f dt + V
Z-t
m 2a
J
rp
V
Z-t
f
m ana
J \mj
11
1
rp
d*.
(18-23)
1
The second term in (18.23) is
22 m
2 2
u ( m) u ( n ) (n j m ) iT2 — («,/w)iTi
ma?ia
n m^n
ilo &(nlm)
|2 \(n— m)lm\ \2
?w=l ]nK)K2m x
, V / V
1 V
m2a \
+ mZ=1 n>2m Z + ?KZ£?n. 7
1 1-1/7
n2a
1 A l«(™)l2 , 1 M l 2'!
+
no t>/n
(18'24)
' where we have used the geometric-mean inequality. The coefficient of \u{m)\hn~iCL in (18.24) is 0 { N \ o g N ) , and thus we can write the righthand side of (18.23) as (T2— T y -}-0 (N lo g N )) 2 \u{m) |2m_2“ . m=i
(18.25)
N
Since
f'(t) =
2 iiog m u (m )m ~ a~il,
(18.26)
and log to ^ logiV, the upper bound for the integral o f \f'{t)\2 is at
78
T H E N E C E S S A R Y TOOLS
3.18
most log2lY times that for (18.23). We have now proved that if tR satisfy the relations , , 2 /,., /, > § (18.27) for r = 1,..., R — 1 and T y + l 8 < tr < T2- 18
(18.28)
for r = 1,..., R, then |u(m) |2 m 2a 1
(18.29)
which is Davenport’s sieve, first published by Montgomery (1969a).
19 THE
HYBRID
SIEVE
Kanga saicl to Roo, ‘ Drink up your milk first, dear, and talk afterwards.’ So Roo, who was drinking his milk, tried to say that he could do both at once . . . and had to be patted on the back and dried for quite a long time afterwards. I. 151
In the last chapter we obtained sieve results in the form of inequalities with two terms on the right-hand side, one of which corresponded to the maximum size of the summands, the other to the mean square of the function times the number of summands. In the series for L(s, %) we can sieve either over % or over s. Montgomery (1969a) sieved over both x and s and obtained a hybrid result, which is better than the one we should obtain by sieving over the one and summing over the other. P. X . Gallagher has found a simple method of obtaining hybrid sieve results, which we describe here. The function f(t) of a real variable t is said to be the Fourier transform of f { x ) when „ f(t)=
\ f(x)e(xt) dx.
(19.1)
— CO
Under suitable conditions (these include the continuity of / at x), oo
J / W e ( - ^ ) di.
/(« ) = —
(19.2)
CO
Equations (19.1) and (19.2) imply formally that co
00
CO
J 1/(0 r d* = J f(t) J j( x ) e ( — xt) clad/, —
oo
—
co
—
co
co
co
= J /> ’) j f(t)e(— xt) dtdx —
CO
—
CO
CO
=j —co
l/(*)lada,
(19,3)
80
T H E N E C E S S A R Y TOO LS
3.19
Plancherel’s identity. We shall use eqns (19.1) and (19.2) w ith /a linear combination o f the values of
* H r £&!>$ Here
F{t) = “ 2 ^ ,
(19.5)
and we can check by means of a contour integral that (19.2) holds except at the points o f discontinuity x — ±| 8 . We can verify from first principles that the interchange o f integrations in the proof of (19.3) is valid. We now deduce Gallagher’s second lemma. Lem m a.
S(t) = 2 o(v)e(vt),
Let
(19.6)
V
where the sum is over a finite set of real numbers v. Let D{x) =
o{v).
2
T
(19.7)
oo
J |^(<)|2d / < 772T 2 f |D(*)|2da;.
Then —
T
—
(19.8)
co
Proof. Let S = ( 2 T ) - 1 in eqn (19.5). Then D{x) = S 2 c { v ) F { x - v )
(19.9)
V
and
B(i) = § J c(v)e(rf)l(i) = §S(t)_F(i).
(19.10)
By (19.3), co
co
T
J |X>(*)|a d® = Sa J \S(t)\*\F(t)\* dt ^ 8* f |£(/)|2|jfy)|2 dZ, —
co
—
co
— *27
(19.11) and
\F(t)\^2/TT
(19.12)
for |f| T , which gives (19.8). We apply the lemma to a Dirichlet polynomial, that is, a finite Dirichlet series N $ (s>x) = X, a(m)x(m)m~a~it> (19.13) m=l
so that the numbers v are of the form —logm, and D ( — x) =
where
Ae* 2 a(m)x(m)> A-ie*
A = exp((2T )-1).
(19.14) (19.15)
3.19
T H E H Y B R I D S IE V E
81
The large sieve (18.8) gives 2
.2
XinodQ a~1gx
«(™)x(®)
< (( A -A -1)e *+ ? ) f
\a{m)\
A_1ez
< ( 2 1- 1ea:+ ? ) X
|«(m)|2.
(19.16)
A ~'e*
i (U>«
xyl * - ( l o f > ' xY)-1
0
--
11
1*
1 + (lo g A r)_1
— i(log A ’ ) - 1
F ig . 4. T h e contour C
By the lemma, T
f J T
logfiiA
N
2
|S(s,x)la tlf < T 2 X
xm oda
I
m=l
( T - V + g ) dx
log(m./A)
X {m-\-qT)\a(m)\2m ~Za.
=1
(19.17)
A more realistic application is the following. For each character X mod q there is a set of points s(r, x) = o(r, x) + i t(r>x) within a rectangle «
—T s ^ t s ^ T ,
(19.18)
satisfying the condition t(r + l,x )-t(r ,x )> 8 ,
(19.19)
where § is a given real number, not necessarily the S of equation (19.4). Clearly we need Gallagher’s first lemma (18.17), but it is not directly applicable, since a is also a function of r. We turn to Cauchy’s identity. Let C be the rectangle (Fig. 4) whose vertices are a—(log iV J -^ i^ og A7) - 1,
1-)- (log N ) _1± i(lo g N )-1.
82
T H E N E C E S S A R Y TOOLS
3.19
Then if s = cr+i/ satisfies (19.18) we have S *(s,x ) = ™ f 2wi
and hence
du,
|S(s, %)|2 < logi\r j |$(t(4~i/> x) I2 \&u \>
(19.20) (19.21)
a
uniformly in cr. We sum over r and apply (18.17) under the integral sign. Next we sum over x mod?. One of the terms on the right of (18.17) is a geometric mean, but Cauchy’s inequality allows us to sum over X hi each factor before taking the geometric mean. We have now to estimate two sums over x of integrals, which are of the form (19.17). In each, a and T have been replaced by A and f + f S , where %i = A+ir, A and r real; in the first a(m) has been replaced by a(m)m~iT, and in the second by —ia(m)logmro~iT. It is simpler to describe this inequality in words than to write it out! After applying (19.17) we have R(x)
X
.
n
2 \s {u+ it(r>x)> x)\2 < ( 3 - 1+loglY ) X (m+qT)\a(m)\2m -* x.
vmodtf i’==l
m —1
(19.22)
Finally we integrate round 0 , noting that o i+ (lo g iV )_1
f
\du\ <^TO-2 a (log.ZV’) - 1 -|-
O
J a -(lo g iV )-1
m - 2 '' dA
m_2am in{l—a+ClogiV)-1, (log to)-1} (m2“ log(TO-f e))_1.
(19.23)
Hence x 2 "
f
(19.24) where for each X the points sr — s(r,x) satisfy (19.18) and (19.19). We can use (18.12) in place of (18.8) and show that
f 2 w ) 2 * 1S(*’ * )l' d‘ < 2 _rp (7
<19'26»
and R(X)
2 ® T'
2* 2 Xmocla r=l
W
’ -'* * '* )1’ N
< ( S - . + l„gJ 0 2 ( » + « 13’) E M
^ J ^ l !.
m= 1 &v 7 where the points $(?*,%) satisfy (19.18) and (19.19) for each
(19.26)
3.19
T H E H Y B R I D S IE V E
83
The point of these hybrid sieve results is that we have N-\-Q 2T (in fact, m-\-Q2T) in (19.25) etc. in place o f (N-\-Q2) T or (N -\ -T ) Q 2 which we would have from sieving one parameter and summing the other. A useful mnemonic for large-sieve results is that N 2 \a(m)\2m ~2a corre sponds to the square o f the trivial term t = 0, % trivial, and that there are 0 ( Q 2T) non-trivial terms, whose mean square is 2 |«(to)|2to“ 2ct. We have certainly shown that eqn (18.2) holds for almost all pairs s,
20 AN A P P R O X I M A T E F U N C T I O N A L E Q U A T I O N (I) In a corner of the room, the tablecloth began to wriggle. Then it wrapped itself into a ball and rolled across the room. Then it jumped up and down once or twice, and put out two ears. It rolled across the room again, and unwound itself. II. 135
W e now have efficient tools for averaging finite Dirichlet series. We would like to average powers of L(s, x) over points s and characters xIn this chapter and the next, we express powers of L(s, x) as finite sums plus error terms. The best results of this type contain partial sums for L(s, x ), as one would expect, but also partial sums for L ( l — s, x). They have thus earned the nickname of approximate functional equations. There are many ways of proving approximate functional equations; we give a straightforward method of H. L. Montgomery’s which uses the functional equation explicitly. Our application leads us to consider L 2(s, x), where x is proper m od?. This has the functional equation L 2( l — u , x) = (qlir)2'<-1G( u)L*(u, x ),
(20.1)
in which we have put
where a is 0 if x( — 1) is 1, 1 if x( — 1) is — 1. We shall use u for a complex variable u = A-fir, where A and r are real. By Stirling’s formula (11.16), when u tends to infinity in a region A |r|5 we have | I »| = (27r)i|«pe-»'W (l + 0(|T|-^)),
(20.3)
and so under the same condition on u, \{ql’rr)‘lu~1G{u) |~
(ferM/TT-)2*-1.
(20.4)
We use the ideas of Chapter 5 to get a partial sum for the Dirichlet series of L 2(s, x). In the fundamental equation (5.4) the convergence of
3.20
A N A P P R O X I M A T E F U N C T I O N A L E Q U A T I O N (I)
85
the integral is only conditional, so we introduce a kernel K(co) to make the convergence absohite. Let i£i(«) = ^7r4e " { ( « —§77-i)(o) —■|7Ti)(oj + l7Ti)(co+|7ri)}-1
and
(20.5)
K(w) = Z ^ w J + ^ - o * ) = ^ c o s h w f l (<«—(r—
-l
( 2 0 . 6)
We have normalized so that K ( 0) is 1. In partial fractions, K t {aj)
1 e“
'r '
u>
2 a)
/ w 3 2 ( 2 — ?’ ) ! ( 1 - ) - ? ') ! ai — (?'—
9
e“
(20.7)
For real positive x, repeated use of eqn (5.4) gives 2+ + ico lco
2rri 'tt i
J
m w
co
\x j
2 —ico
where c(v) = 0
if v > e
c(w) = -J—3^ sin(^77-logw)—3^ sin(f77-log-y) c(v) = 1
if e -1 < v ^ e
J. (20.9)
if 0 < v < e-1
Lets = cr+ii be a fixed complex number with J ^ a < f . The restriction on a is not essential for the proof, but it is sufficient for our application. Then 2 —cr+ ico
y
d{'m] ^ m') ct~~) = 2~1
f
L%s+aj,x)daj.
(20.10)
By construction K(to) is an integral function, and the only pole of the integrand in (20 . 10) is at co = 0, with residue L 2{s,x)- The difference between L 2(s, x) and the sum in (20.10) is therefore given by an integral along a contour passing to the left o f this pole, which we shall take as the vertical line Reco = —i —cr. Equations (20.1) and (20.4) provide bounds for the integrand that justify moving the contour: for example, they imply that | £ ( - i + i r , x )|2 (20. 11) since L ( |—ir, %) is bounded, and that |Z(A+ir)| < r - 4
(20. 12)
in — 1 ^ A ^ 2, |r| ^ 10, so that the integral along Reco = —a converges absolutely. We change the variable of integration to u
T H E N E C E S S A R Y TOOLS
86
3.20
( = A+ir) where s+a> = \— u, and use the functional equation (20.1) to write
ox |+ico
1 2771
f
) ( z Y u~1 Q(u ) L n u>x ) d u .
J
(20.13)
s + tt— 1
f —i c o
Because of the factor K(s-\-u — 1), the integrand in (20.13) is largest when is close to 1—s. Now when it is near 1—s the terms (20.4) approxi mate the (2A— l)th power of the constant \ c[\s \Itt. In modulus at any rate the main contribution to (20.13) is an integral of the form (20.10) with s and x replaced by 1— s and x , and x replaced by y, where the integer y is given by y = (20.14) More precisely, provided |Im(s-|-tt— 1) |< 1
a\21,-1
'
we have *\2\A—1
G(%)xl- Uy x~n
47r2x y
?ki
< (l + 0(|/|^) + 0(|2/|-1))2i'l* < 1,
provided that
j£|-'.
y
I f we choose y by
4772x y = q2t2
(20.15) (20.16) (20.17)
exactly, so that y is not necessarily an integer, the term 0 ( y ^ 1) is not present and (20.15) is true in any case; but in this application we shall g^) and (20.16) will hold easily. have y The series for L2(f + ir , x) converges absolutely uniformly, so that we may integrate it term by term. By analogy with the term-by-term integration o f (20.10), we break the series up into 91+93 or into (according to whether we take K-^s-^-u— 1) or 1^(1—s — u)), where cp4 are given by 9i(u) ~ 2(u ) =
2
d(m)x(m)m-u N
2
d (m )x{m )m -,‘
2
d(m)x (m )m -v
m<ev m <e~ iy
(Pz(u ) —
(20.18)
m >ey ? i(u )
=
2
d{m )x{m)m-u
We write the integral in (20.13) as the sum o f four integrals numbered correspondingly / 4. First we consider the integral I3 involving
3.20
A N A P P R O X I M A T E F U N C T I O N A L E Q U A T I O N (I)
87
cp3('ii)Z1(s+% — 1). When (20.15) is valid, the integrand decreases in modulus as we move the contour to the right. We employ the contour 0 consisting (in the case \t\ ^ 10) o f Gy. the line segment (2—ioo, 2—it— i|£|}], G2: the semicircle centre 2—it, radius 111*to the right o f the line A = 2, Gs : the line segment [2—ii-f-i|i|*, 2+ioo),
2
l-.s
Pig. 5. The contour G
which is the line A = 2 with an indentation around the poles of 1). I f \t\ < 10 we take G to be the line A = 2; the estimates below will then hold with (if)-1replaced by 1 and with different implied constants. For A ^ 2 we have
2
l<Pa(“ )l <
m >oy
d{m )m ~x
2 m ~ \ D (m )— D { m — 1)} m>ey < {ey^ log y, —
(20.19) by partial summation and the estimate (2.13) for the sum function D(m) of the divisor function. By (20.19) and (20.4), T .'B1- s~“X 1(s+'M— 1) j g^2" - 1 ( G(u)
<J Ci
,.1 -a -A
Jxyir y ~ *
< ?»-
y^logy dr
{l+\t+r\f\2rr j
■< x - at~2
853518 X
/^|t |\2A-1
4qH2) _1iog y-
q\t\logy ( 20 . 20 )
88
T H E N E C E S S A R Y TOOLS
3.20
The same estimate holds for the integral along Cs. On the semicircle C2, (20.4) and (20.15) are valid, and we have the upper bound < x - a(l+ \ t\ i) - 5q\t\logy(ir\t\i) <^.qx~a\t\-1\ogy.
(20.21)
Hence / 3 and similarly / 4 are bounded by the expression in (20.21).
21 AN A P P R O X I M A T E F U N C T I O N A L E Q U A T I O N (II) ‘ Wliy, what’s happened to your tail ? ’ he said in surprise. ‘ Wliat lias happened to it ? ’ said Eeyore. ‘ It isn’t there! ’ I. 43
W e have now shown that I 3 and I i can be regarded as error terms. We treat (px and cp2 by moving the contour the other way. The first case considered is |i| > 10. I f t > 10 we use the contour D given by D x: the line segment (—
ioo, —J— it— i|£|*],
D 2: the semicircle, centre — \ — it, radius \t^!, to the left o f A = — D s : the line segment [—J—i£-|-i|i|% —-J-—i],
D 4: the line segment [ —
i, J—i],
Z>5: the line segment [J—i, J + i], D 6: the line segment [J + i, — i + i ] , D 1\ the line segment [ —1 + i, —J —ico).
The contour D for t < — 10 is the reflection in the real axis of that described above. The indentation about the origin avoids a possible double pole of G(u) at u = 0. The analogue of (20.19) is that, uniformly l
(21 . 1)
C x 1~s~uK 1(s-}-u— 1)
!7ii J D
Using (21.1) in place of (20.19) we see that |ii| <4qx-°\t\~nogy,
(21.3)
and the right-hand side of (21.3) is similarly an upper bound for / 2 in which K x( l - ~ s — u)(p2(u) replaces K x(s-\-u — l)tp1(it).
90
T H E N E C E S S A R Y TOOLS
3.21
Between the contours G and D lie the poles of w - 1K ( w ) at iv = 0 and ibO'—IV i, where w = s-\-u— 1. The residues at these can be calculated term by term; in total they give
w =
2
d(m)x(m) ,lm c “ m 1—s
(21.4)
Fig . 6. The contour D
where c'(v) = 0
if v ^ e
+2
1 9 32(1+J*)!(2—?•)! \n2vxyi
X ( ? ( l — 5 + ( r — l)7ii)
X
(21.5)
if e_1 ^ v < e
1—2s
c'{v)
G ( l — s)
if 0 < v < e_1
If is then Q{u ) is regular at 0. For \t\ ^ 10 we take the contour D to be the line A = — f . The bound for Jt and / 2 on this contour
3.21
A N A P P R O X I M A T E F U N C T I O N A L E Q U A T I O N (II)
91
is the same, but without the factor |£|-1 and with a different implied constant. I f x( —1) is + 1> however, there is a double pole. I f a is bounded away from 1 we can run the contour D between the origin and the poles of /^ (s+ m — l)/(s-\-u— 1). I f a < 1 we take D as D x\ the line segment (—
D 2: the line segment D 3: the line segment Z>4: the line segment D h\ the line segment
ioo, — i], [—1 —i, 1(1 — cr)—i], [1(1 —cr)—i , l ( l —cr)+i], [1(1 —cr)+i, —1 + i], [ —1 + i, —1+ioo).
The upper bound for the integral along I)3 will be (21.6)
qx~°(l — a)~zlo g y.
I f all else fails, we can take I) as the vertical line X = — \ and get an explicit but very complicated extra term in the approximate functional equation from the residue at n — 0 of the multiple pole. When |£| < 10, (20.14) is interpreted as q 2 < ^ x y < g 2. (21.7) We have now shown that an equation of the form 7jV jx) _ 2
- M
A ») + 2
™
,(5 )+
H ( 2 1 .8 )
holds both in \ ^ a ^ f, \t\ > 10, in which case each of / 1;..., / 4 is
( 2 1 .9 )
/ 4 is
(1— o ) - 2qx~ulo g y,
(21.10)
the estimates being uniform in the ranges stated. Although we have assumed q > 1, so that L(s, %) is not £(s), we can treat £2(s) by the same method. We assume \t\ > 10, since it would be unprofitable to try to approximate near the double at s = 1. When we move the integral in (20.13) to the line Re?o = —1—cr, this double pole gives a residue 2Aiv~1K{iv) + A
[to^ K iiv))
(21.11)
evaluated at to = 1—s, where A is the second coefficient (in fact, A is Euler’s constant y) in the Laurent expansion £(«) = ( a - l ) - H - 4 + 0 ( « - l )
(21.12)
92
T H E N E C E S S A R Y TOOLS
3.21
of £(s) about the pole at s = 1. We must add to eqn (21.8) the additional terms (21.11); they are ^ (21.13) uniformly i n | / | ^ 1 0 , j ^ c 7 ^ f . There are many approximate functional equations in the literature. The one we have proved is easy to sieve, but has more complicated coefficients. Approximate functional equations for £(s) and for £2(s) are given by Titchmarsh (1951, Chapter 4), and generalized by Chandra sekhar an and Narasimhan (1963). A. F. Lavrik has a number of papers on the subject in the Doklady and Izvestia of the USSR Academy of Sciences. Fogels (1969) has a form rather like that of Montgomery’s given here.
F O U R T H P O W E R S OF L - F U N C T I O N S They were out of the snow now, but it was very cold, and to keep themselves warm they sang Pooh’s song right through six times, Piglet doing the tiddley-poms and Pooh doing the rest of it, and both of them thumping on top of the gate with pieces of stick at the proper places. II. 7
T h e approximate functional equation for L 2(s, x) is used to obtain upper bounds for |L(s, x) |at individual values of s, or on average. Our inspiration is the Lindelof hypothesis, ( 22 .1)
for each e > 0, with a constant depending on e. In this chapter, we show that the mean fourth power o f L(s, x) satisfies the inequality (22.1). Beyond the fourth power, the sums in the approximate functional equation are too long for us to prove (22.1) even on average. We use the hybrid sieve (19.26),
( 2 2 .2 )
Although (22.2) allows us to vary a, we take a — \ throughout, since \{ql'rr)1^2s(xle)^r7TiG (l— s + > '7 7 i ) |
=
(22.3)
1
for a = J, the gamma functions in 0 being evaluated at conjugate complex points. The proof is no simpler, but the form of the upper bounds is less complicated. Cauchy’s inequality applied to eqn (21.8) gives \L(*> X)l4 < 6|2d(ra)x(m)m-«c(m/®)|a+ 4
3.22
T H E N E C E S S A R Y TOOLS
94
and we have a similar result for £(s), with 7 instead of 6 and an extra term 0 ( |/|~10). The integers x and y in (22.4) are connected by (20.14): y = [qH*l( 4 A ) ] .
(22.5)
When we fix x and average over x and t, y is varying. For a good upper bound, x and y must be of the same order of magnitude. We restrict ourselves for the moment to P < q < 2P and U < \t\ < 2 U , and average the right-hand side of (22.4) over all integer values of a; between \ P U j n and \ P U jit . For each fixed x and t ^he corresponding values of y given bjr (22.5) are distinct and lie between ^PUj-rr— 1 and 1 6 P U / tt. The average of the square of the terms involving y taken over all integers y in this range is at least times the corresponding average over the values o f y that actualfy occur in the sum. This device allows us to sum a Dirichlet series of fixed length ey over varying x and t, provided that x does not occur in the coefficients (and vice versa). For each value of x in the above range, (22.2) gives R(X)
u
n
# < < K 2P T'-t/ ymoclg
d ( m )x (m )
hn'
TO^’-X)
\«
cx
r —1
(S_1+loga;) ^
(m +P ^U )
m—1
cZ2(to) log x mlog (m +e)
< P 2£/log5(P ?7+e),
(22.6)
where we have used (2.24) for ^ d2(m)jm, and partial summation. We have assumed
8> 1
(22 7)
It is only slightly harder to average over the reflected series. For e ^ y < to < ey we must break the coefficients c'(m/y) up into five terms
corresponding to the five poles of fl\
»)
I-
We take out a factor
/ —.2/^A —Y n i
(i)
(22'8)
from the term corresponding to the x^ole of K - ^ w ) ^ at ttt\, where r == 0, + 1 . By (22.3), the modulus o f the expression (22.8) is fixed independently of x, q, t, and the parameter a in the definition of 0(u), which is 1 if x( — 1 ) = - - 1 and 0 if p^(— 1) = + 1 . After this factor has been removed, a sum like that in (22.6) remains, with no gamma-function factors and no concealed dependence on *. For each r, the right-hand side of (22.6) is an upper estimate for this sum, possibly with a different 0 -constant.
3.22
F O U R T H P O W E R S O F L -F U N C T I O N S
95
The contour integrals Iv ..., / 4 present complications. First we fix t and average only over characters. We have for any function F(u) for which the integrals exist 2 F(u) d u
''£> Ti
D
71 '' '' D
'
(22.9) and similarly for integrals along the contour G. The first factor here is We shall find an estimate for the average of Iv Here, F(u) = x1- s- u(sJr u — l ) - lK 1(sJr u — \)(qlir)'in- 1G(u,x)qii{w)( 2 2 . 10 )
We wish to average this over % with the length of the sum cp^w) fixed, that is, with y fixed. By (22.4), x is approximately varying as q2, and is bounded on the contour D by a function o f t and y alone, by (20.14). Similarly, (20.4) permits us to take out (q'/tt)2 1 G?(^<-) at its maximum. B y (18.12),
1
2
2
*
xmods
msgey < ( y + P 2)(e«/)i-2Mog!ty,
(22.11)
where we have used the estimate 2 d2(m) < l o g 3i¥,
(22.12)
which can be deduced from (2.24). The sum of the integrals involving F(u) on the right-hand side of (22.9) is now yy,\ 2—2 u/^y\4u —2
< max max \ t \ l ar (q,x) med (|£|T
>
m
G*{u) ( y + P i){ey)i--*x \ogzy
(max&)1_2l7|i|~*(?/-j-P2)log3?/ < (P 2+P|i|)|i|-Jlog3(P(|<|+e)),
(22.13)
where we have used (20.15). Hence q ]T * l^l2 < ( P 2+P|<|)|
(22.14) and similarly the same upper bound holds for the corresponding sum involving / 2.
96
T H E N E C E S S A R Y TO OLS
3.22
To apply a Gallagher sieve we need a similar bound for a sum involving The contour D varies with t, but this was only to help estimation, and we can keep D fixed and differentiate the integrand. A similar calculation will now give
(22.15) Gallagher’s lemma (18.17) gives 277 i it kj+ T"2
i
,/ ^ 2d'~ V/~a +i
x\ ,1 /, 2U + i
-
J i/ii2^+ J i-^rdiW J
U < t r^ W
U~i
'u - t
'
'77-i
i j dt 1
2
. jl
\t
dtj
(22.16) We sum (22.16) over % (using Cauchy’s inequality on the second term) to find , „ _ mx) 2 Jo ) Z 2 IAI8 < ( - P + ^ ^ - 8log4(P (?7+ e)) (22.17) P
x m od2 r = 1
when we substitute in (22.14) and (22.15). The same upper bound holds for the sum with Ia. The integrals I 3 and / 4 present a further complication in that
19 d (m )x (m )m ~ u\
>i>ey
I
/ J"?
\/ ^
I
( 2 ?l_2)( 2 ,l2 V=1
' 'n=l
,_i 2
leM y<m<eM +1j/
\ d (m )x (m )m ~ u 2) '
(22.
By (18.11) we have g cp(<7)
d(m)x (m)
'V *
—. ,
P
in
(e“ +1« /+ P 2)(e,li/) 1_2;'(ii+log?/)3,
(22.19)
and the sum of terms of the form (22.19) from n = 1 to infinity converges uniformly in A to
^ ( y + P a)(ey)i-^ 1og 8y.
( 22.20)
When we use (22.20) in place of (22 . 11) we show that the sums over I 3 and J4 also satisfy the bound (22.17).
3.22
F O U R T H P O W E R S OF ^ -F U N C T I O N S
97
We have now shown that
< P 2C71og6(P (?7+e)) + (P + £ 7 )P £ /-3log4(P (?7+e)).
(22.21)
By (21.13), if the zeta function is included in the sum there is an extra term TJ~9, which is easily absorbed in the second term on the right of (22.21). When we sum over values of P and U running through powers of two, we can show that
provided that for r =
1 ,...,
R(x) we have
(22.23) and for r = 1,..., R ( x ) ~ 1 we have t(r + l,x )-t{r ,x ) >
1,
(22.24)
and for q = 1 \t(r, x)| ^ 1 . We could replace (22.24) by a weaker condition t { r + l , x ) ^ t ( r>X) > (log Q T ) - 1,
(22.25)
and still have the upper bound (22.22), so that the root mean fourth power of L(s, %) is log Q T , Without the sieving over T this method proves that the root mean fourth power at afixedsis <^log(Q(|s|+e)). A simpler proof of a fourth-power moment is in Gallagher (1967). Another result is given by Linnik (1964, section 41). These relate to summing over x- The situation is much easier if we merely integrate over f; there are examples in Titchmarsh (1951, Chapter 7). The hybrid moment we have proved here is sketched in Montgomery’s paper (19696), but with a larger logarithm power.
P A R T IV
Zeros and Prime Numbers
23 INGHAM’S THEOREM He tried Counting Slieep, which is sometimes a good way of getting to sleep, and, as tliat was no good, he tried counting Heffalumps. I. 62
W e shall estimate the number of zeros p — j 3 + i y of L(s,x) or o f £(s) in a rectangle
« < 0 < 1,
—T < y < T
(23.1)
where re > -J, T > J. Let X be a large integer and M (s , x) =
2
(23.2)
m<:X
I f Riemann’s hypothesis were true, then for a > \ M (s , x) would tend to {£(s, x)}_1 as tended to infinity. We write L ( S, x ) M ( s , x ) = l + f ( s , x ) ,
(23.3)
w here/(s, x) lias the Dirichlet series
2
a(m)x{'in)m~s,
(23.4)
m >X
a(m) =
2 M
(23-5)
d^X
The series (23.4) converges without any hypothesis for a > 1. Until A. I. Vinogradov’s work (1965), the number of zeros was customarily estimated by integration o f the logarithm of 1+/(«> x) round the boundary of the rectangle (23.1). As we shall see below, f(s , x ) is less than unity in root mean square, whether averaged over t or over x °r
4.23
IN G H A M ’ S THEOREM
99
over both, provided a > J. At a zero, f(p , x) is — 1. A. I. Vinogradov revived an alternative approach: to count directly the number of times f ( s , x) has modulus at least unity. While no simpler in detail, this method is the more flexible, and we follow it here. The integral transform 2 +ico
J
I f ™ 2t t i
JY\W r ( w ) ^ j W dIV = e - m'r
(23.6)
2 —ico
is easily verified by moving the line of integration to Be vo = -J— R, where R is a large positive integer, and letting R tend to infinity: the residues at poles of F(w) give the terms o f the exponential series. Hence m >X
2 + ico
J
277i
L i p + t v , x ) M ( p Jr w, x ) Y wr (w ) div.
(23.7)
2 —ico
Here p is a zero of L (s , x) so that the zero of L(p-\-w, x) cancels the pole o f F(w) at w = 0. We take the integral back to the line Rew = (3. I f x is not trivial, the right-hand side is equal to i-p+im
1 27ri
J
L (p Jr iv, x ) M ( p Jr iv, x ) Y wF(w) dw,
(23.8)
and if x mod q is trivial there is an extra term 9(q)q-1M ( l , x ) Y ^ P F ( l - p )
(23.9)
from the pole at p .\-w = 1. We write
I — log Q T
and suppose that
logAr .
lo g !7
(23.10) 10Z.
(23.11)
The terms on the left of (23.7) with to > 100IY contribute less than , if I is sufficiently large. We recall Stirling’s formula in the form (20.3), valid if A ^ |t |j : |r(A+ir)| = e-^W|A+ir|;‘-i{(277^ + 0(|T|^)}.
(23.12)
The term (23.9) is now seen to be at most ^ when |y| ^ 100Z, again provided I is sufficiently large. Apart from the zeros o f £(s) with |y|^ 100Z, all zeros fall into one or both o f the following classes. Class (i). Zeros p with y
a(m)x('>n)m-Pe~m!Y > ■§.
(23.13)
ZEROS A N D PRIME NUMBERS
100
4.23
Class (ii). Zeros p with - ; 8 +ico
j
L(p~>r'W, x)iY(/>+ti>, x ) Y l°r(w ) dtv >■§ 77.
(23.14)
i-fl-ic o
We subdivide class (i) by writing the range (X , lOOZF] as the union of intervals/,.: 2rY < m ^ 2,,+1I r (the first and last intervals having instead the end points X and 100IT). A zero is of class (i,r) if 2 a(m)x(ni)m~Pe~m,r > {20(r2+ l)} ^ 1.
(23.15)
nisi?
By (12.19) or (12.20) the number of zeros of a fixed function L(s, x) in a subrectangle t y ^ t-\-1, a ^ j8 ^ 1 of (23.1) is 0(1). From each class of zeros in (23.1) we can pick out a sequence of zeros whose imaginary parts differ by at least unity, in such a way that the sequence contains a proj)ortion o f at least ;> l~x of the zeros of L(s, x) of that class in (23.1). It is possible that our sequence contains all p at which L (p ,x ) has a zero of the given class, but these p are multiple zeros of order 0(1). We write Nt(x) for the number of class (i) zeros of L(s, x) in (23.1), N2(x) for class (ii) zeros, and N (x ) for the total number. We call our sub sequence the representative zeros. Summing over representative zeros p of class (i, r) we have Y
JL
V * V
Z-t m(q) Z-i
<1
xmoda
Z-, p
V /-i
m elr
a(m)x(m ) e -,„ir 2 mP
^ Z 2 (™ +Q 2y)^2(™)™~2a:exp( —2'’) melr
£(2'T)1~2“(Q2I 7+ 2 ,T )log4r e x p ( — 2r) < Z6exp(-2'')(<92y (2 'T )1- 2“ + (2 'T )2- 2a),
(23.16)
by (19.26) with 8 = 1 ; we have used (23.3) to estimate a(m) by a divisor function, then (2.24) and partial summation. Comparing (23.15) and (23.16) and adding 12 for the zeros of £(s) with |y|< 100Z, we have 2
2 " Nlix) < ^ 2+ / ° i ; ( r 2+ 1)exP( ~ 2 0 x
q^Q ''■-i-' ^mod(2
X{<922,(2,T ) 1- 2“ + (2 ,T ) 2- 2“}
(23.17)
where we have assumed X > Q2T. We now choose X = Q2Tl.
(23.18)
There are several ways of treating zeros of class (ii). The ingenious work of Montgomery (19696) makes much use of this flexibility. To
4.23
IN G H A M ’S THEOREM
101
obtain Ingham’s theorem we raise the expression on the left o f (23.14) to the four-thirds power and sum over representative zeros p — /3-f-iy of class (ii). By Holder’s inequality we have i - 0 + icO
k
V JL V* V (pftf) jL-4 Z-4
q ^ Q T SJ-' xmod q
L(p-\-w, x)M(p-\-io, x )Y wr(iv) dto 3
p
) — 1 CO
i +ico 2 ^ 2 * 2 f a
j
X
2 ^ ' e=so
2 *
)
\2
J
2
xm°as p
| d«| r x
l ^ + i y - x)\z\r iu ~P)\* |d»lj •
(23.19)
By (23.12) and (22.22) the first integral on the right o f (23.19) is i
<
f Q2T l5m & x . { t + B - t ) - t d t + o P CO
+ j" Q2( T Jr t)(lJr \ogt)5msbx(e-27rllH~iPls) di <
1
(23.20)
When we apply the hybrid sieve (19.26), the second integral on the right of (23.19) becomes x ( m + Q 2T) lo g Z m log(m +e) m=1
<^i(a- i ) - n x + Q m )
(23.21)
<
by (23.18). The right-hand side of (23.19) is (23.22)
<
and thus
^ a
We now choose
2 * xmoda
^ {ot— l ) - * Q iTY*a-*<Mii.
Y — (Q2T l - 2)8l(-i - 2oc\
(23.23) (23.24)
so that (23.17) and (23.23) together give 2 “ H 2 * iY(x) ^ (a -4 )-l( Q 82,)a(1- fl«,(S!- “)ZW-oO 2*SQ ‘ ^morta
(23.25)
102
Z E R O S A N D P R IM E N U M B E R S
4.23
We have assumed that 100ZY is greater than X ; this is certainly true if
a > $ + fZ -ilo g Z.
(23.26)
I f a is less than this bound we quote (12.19) or (12.20), which give an upper bound
N (x) < T l ,
so that
^ a^Q
T iV(/Y) ^ Q2Tl xvaoaq
(23.27) (23.28)
in any case. In fact, (23.27) and (23.28) give
2m2
*
< min{Q 2Tl, ( « - JJ-^QaTJsa-aJKa-^s/OJ-a)}.
(23.29)
Ingham proved his theorem for the zeta function only, with no sum over his result (1940) was N < ysa-oo/ta-oO^
(23.30)
where N refers to the zeros of the zeta function satisfying (23.1). (23.25) is stronger than (23.30) for a near but weaker for a > f . We have lost a logarithm factor by taking representative zeros instead of counting according to multiplicity.
24 BOMBIERFS THEOREM ‘ I see, I see,’ said Pooh, nodding his head. ‘ Talking about large somethings,’ he went on dreamily, ‘ I generally have a small something about now— about this time in the morning. ’ I. 48
W e proved the prime-number theorem for arithmetic progressions a modr/ uniformly in q <, Q, where Q =
(log x)N,
(24.1)
in the notation of Chapter 15. In 1965, Bombieri (1965) and A. I. Vinogradov (1965) provedindejDendentlythat the prime-number theorem holds uniformly for almost all progressions with Q a little smaller than Bombieri’s result was the following theorem. T h e o r e m . Given A > 0 toe can find B > 0 such that for all large x
2
. am odg y ^ x
Hy> >«)-
q
y
9 (g)
^ (log®)^’
Q = x*Q.ogx)-B
where
(24'2) (24.3)
and the asterisk indicates a restriction to reduced residtie classes.
The right-hand side of (24.2) is smaller by a factor (loga;)^+1 than the sum o f the corresponding main terms. The proportion of progressions for which the error term in the prime-number theorem for the interval 1 ^ P ^ x is greater than (logx)~A times the main term is O((loga:))-1. In calculations we often apply the prime-number theorem to several progressions, and its failure in a small number of cases affects only the error term in the eventual asymptotic formula. An example is the proof that every large even integer is the sum o f a prime and an integer with at most three prime factors. It may be that (24.2) holds with some larger constant in place of but ?, is the limit of present methods. An upper bound for the left-hand side of (24.2) is
q ^ Q TV- " 863518 X
2
xm odg
X non-trivial
2
"
« r
?
l!% ' j;“>~ * '!'
, 2 4 -4 )
ZEROS A N D P R I M E N U M B E R S
104
4.24
When we use eqn (17.15) to pass to proper characters, the sum in (24.4) does not exceed
1 « /« Q
x m o c l/
V<X
s s O t m o d /)
a
+max|i/i(2/ ) - 2/l S ' - ^ r + A2Q,
(24.5)
A = log*.
(24.6)
v^o;
where we have written
(q)
The coefficient of the term in % mod/ in (24.5) is V
ffl(ff)
Z -!
g= 0(m od/) Tv~i '
a
ro(f) Z -, m(rn) Tw '
'
cp(/)
(24.7)
rw '
Relations (16.22) and (24,3) assure us that the second and third terms on the right of (24.5) are ^ x^_A ^24 ^ for x sufficiently large. Since / -
4>{y, x)
= ~ 2 ^+
i+ ^l)log2? Ty\’
(24.9)
where the sum is over zeros o f £(s, %) with |p| > R and |y| < T , R and T being chosen for each % so that (17.3) and (17.4) hold; in particular the O-constant in (24.9) is independent of T, or y and uniform in £^ R J. When we choose T to be a little smaller than x we have
I•p{y>x)\<
y
Y^+O{x^logzx)
(24.10)
whenever y < x, with an absolute O-constant. The error term in (24.10) contributes O( Q x Ho g ^ ) (24.11) to the sum (24.4). We now divide the rectangle 0 ^ / 3 < 1, ~ T ^ y < T into smaller rectangles, so that /3 is in one o f the ranges [0, J+(log® )-1],..., [i+ r ^ o g a O -^ l+ ^ ’-flX lo g * )-1],..., and |y| in one of the ranges [0,1], [1,2],..., [2r, 2>’+1],.... Within each rectangle, xP{\p\ varies by a bounded factor. Again, we divide the range (A45, Q\ for q into intervals
4.24
B O M BIB RI’ S THEOREM
105
P < q ^ 2P. We use Ingham’s theorem: the number of zeros p = |3+iy counted with weight q[
(24.12)
by (23.25) and (23.28). Each of these zeros contributes at most ^
_A_ <
\P \ (q) ^
— JL P U
(24.13) K
1
to the sum in (24.5). We multiply the expression in (24.12) by Xxa/ P U and sum over the appropriate sequence of values of P U . First, we have 2
£ / 3 ( l - a ) / ( 2 - a ) —1
^ ^
(24.14)
V
when a > |- and the sum is over powers of two not exceeding x. In the sum over P we distinguish two cases. The exponent o f P in (24.12) is greater than unity for a < f and less than unity for a. > f. The values of P are the powers of two between JA4B and Q; they are 0(A) in number. Hence for a ^ f we have 2 p8(l-o0/(2-a)-l
^^4B)6(l-a)/(2-a)-l
P
^ ^ l+ 4 B(0(l- a )- l)
< a ;i“ “A1- 4B,
(24.15)
when we assume x to be sufficiently large. For \ < a ^ § we have 2 _p6(l—a)/(2— a)—1
AQ6a-«)/(2-a)-l
P a. 3 ( l - a ) / ( 2 - a ) - ^ l - J ?
< a ;1“ “A1- I?,
(24.16)
where we have substituted Q = a^A~B from (24.3). The terms from zeros o f L-functions formed with characters whose conductors exceed A43 are now seen to be
^ aA10--5
(24.17)
by (24.12), (24.13), (24.14), (24.15), and (24.16). The other terms in (24.5) are of the form (24.17), by (24.8) and (24.11), when we choose B = ,4 + 10.
(24.18)
We have thus proved (24.2) with B given by (24.18), which clearly can be improved slightly, since (24.15) and (24.16) can be improved. There are two ways o f proving Bombieri’ s theorem. The shorter, due to Gallagher (1968), is to perform the sieving directly on L'(s, x)jL(s, x)
106
ZEROS A N D PRIM E NUM BERS
4.24
and related functions in contour integrals. The proof is an anagram of the one we used; multiplication by M (s , X) and fourth-power averages o f L(s, x) occur in it. The longer way is to prove a zero-density theorem and deduce Bombieri’ s theorem from that, as in this chapter. In either case, the Siegel-Walfisch form of the prime-number theorem has to be used, and the constants are non-elfective. Without using the hybrid sieve or the approximate functional equation, we should obtain a weaker zero-density result than (23.29), but one still powerful enough to enable (24.2) to be deduced.
>
25 I. M. V I N O G R A D O V ’ S E S T I M A T E I f we are to capture Baby Roo, we must get a Long Start, because Kanga runs faster than any of Us, even Me.
I. 93 T h e sum
8{oc) = S(y, a) =
X logpe(pa)
(25.1)
is of the type considered in Chapter 6, with local maxima near rational points ajq (if y is sufficiently large). In this chapter, we adapt the proof o f Bombieri’s theorem to show rigorously that 8 (a ) is small except possibly when a is close to a rational point with small denominator. More precisely, if B is fixed and x is large, then (25.2) where
(25.3)
and ajq is in its lowest terms with (25.4) Inequality (25.2) is I. M. Vinogradov’s famous ‘ minor arcs estimate’ , with 8 in place of Vinogradov’s His proof is elementary, making no use of Dirichlet ’ s series; an analytic proof was published by Linnilc (1945), We remark first that S (y,a) =
2 A(m)e(ma) + 0 ( y n o g y ) ,
(25.5)
the error term arising from squares and higher powers of primes. When ajq is in its lowest terms,
the error term arising from powers of the O(logg) primes that divide q; these are not congruent to reduced residues b mod#. The sum of the terms involving b in (25.6) is x(a)T(x)> and by ecln (3.20) the modulus
ZEROS AND PRIME NUMBERS
108
4.25
o f this expression is at most g*. We combine (24.10) and (17.15) into (25-7) IPI > 7,
1 /5 1
lyl<» whenever y + x, the O-constant being absolute. The stun is over zeros o f £(s, Xj), where Xi proper mod/ induces x modg. Since every zero of i ( s >Xi) i,s a zero x)> we shall take the sum in (25.7) to be over zeros of £(s, x). From (25.5), (25.6), and (25.7), we have
2
2
<*•«)
TW/ xmod« ip|>‘
where y ^ a; and q ^ a:. We do not use Bombieri’s theorem itself, but analogous results for zeros o f the functions L(s, x) where x induces a character to the fixed modulus q. This argument is not as natural as that leading to Bombieri’s theorem, and the various inequalities do not fit together so well. First we give a form of Ingham’s theorem for the set of all characters % mod q, where q is fixed. I f x modg is induced by xi proper m od/, we have L(s, X) = L(s, Xl)
(l ~
~
p \ <j
(25.9)
\ '
p/ f
The method of Chapter 22 with (19.24) instead o f (19.26) gives
2*
R(x)
2
\L(^+it(r>x))\l < c P(/)T lo g !/ T ,
(25.10)
Xmod/ r=l
and hence
2 2 xmoda >•=!
m + it(r ,
x))l4 <
T lo g tq T
Y
II \
p\
PXf
T l o g 5q T J ! (p — l + l+^J- *) p\a
qT\og5qT,
(25.11)
since the product over primes dividing q is at most g£(f). Here we sirppose that the points t(r, x) satisfy conditions (22.23) and (22.24). When we use (18.8) and (25.11) in place of (18.12) and (22.22), Ingham’s theorem becomes: the number o f zeros p = jS-f-iy of functions L (s ,x ) formed with characters x mod q in the rectangle a ^ / J ^ l , —T ^ y ^ ^ 1 is at most max{
A = log qT.
(25.12) (25.13)
4.25
I. M. V I N O G R A D O V ’ S E S T I M A T E
109
As in the proof of Bombieri’s theorem, we sum zeros over rectangles. For a ^ the analogue of (24.15) is ^3(1-00/(2-00-* <^(^3(1-00/(2-00-* ^ x l-ocl-iB (25.14) when x is sufficiently large, and for a < f the analogue of (24.16) is ^3(l-a)/(2-a0-.} ^ (xl~B)3(l-a0/(2-a)-} <^ ^1-^-5 B_ (25.15) Since the primes less than q and not dividing it are reduced mod q, we see from the prime-number theorem that y{q)X6jq is greater than unity for q satisfying (25.4). By (25.14), (25.15), and (24.14) the sum over x and p in (25.8) is < (25.16) and we have proved (25.2) for /3 = 0. We now show that (25.2) holds (with a different O-constant) when j8 is different from zero. Since m j 2irifie(fiy) dy = e(m/3) — 1, (25.17) o we have the identity S (x ,a l q + P ) X
= 8(x,alq) + 2iril3 [ { S(x,alq) — S(y,a/q)}e(Py) dy. o The right-hand side of eqn (25.18) is in modulus < (l + 47r|/3|a;)max|$(2/,a/q)\, and we have proved (25.2).
(25.18)
(25.19)
26 I. M. V I N O G R A D O V ’ S T H R E E - P R I M E S THEOREM W i t h S ( y } a) given
eqn (25.1) and x an integer we have
x f S 3(x, ot)e(-xot) da = 2 2 2 lo g ^ jlo g ^ lo g ^ , (j
(26.1)
Pi P 2 Pa
where the sum is over tiiples (Pi,P 2,P s) of primes with P i+P z+P a = x -
(26.2)
We shall find an asymptotic formula for the left-hand side of eqn (26.1), and deduce that every sufficiently large odd integer is a sum of three primes. Since we have to use the prime-number theorem for arithmetical progressions, the constant in the error term of the asym ptotic formula depends on Siegel’s theorem and cannot be stated; how ever, there are weaker results than Siegel’s theorem in which the con stants are effective, and in principle a finite set could be given in which any exceptional x must lie. Let
Q = [xl~2S],
(26.3)
where I = log a:. Since S(x,u) is periodic, we can take the range of integration in (26.1) to be the unit interval [1/(Q +1), (Q-t-2)/(Q-|-l)]. We divide this interval into arcs - i
Ir
A r + ir -l
a,+a,.+1
(26.4)
5V+2V + 1.
corresponding to the fractions ar/qr of the Farey sequence of order Q. By eqn (9.2) if ar/qr~{-/3 is on then IPI < M ) - 1-
(26.5)
The intervals I r are known as Farey arcs (if we think o f e(a) as a com plex variable, the integration is round the unit circle). We divide the Ir into major arcs, on which we work out the integral over the spike at arlqr explicitly, as promised in Chapter 6, and minor arcs, on which we
4.20
I. M. V I N O G R A D O V ’ S T H R E E - P R I M E S T H E O R E M
111
use an upper estimate for the integrand. We call Ir a minor are if qr > Z26, when (25.2) with B — 26 gives \S(x, ar/qr+ P ) | < (1 +®Z-aeQ -1)a:Z-8 < x
(26.6)
l
for ar/qr lying on Ir. On the major arcs we approximate 8(a). First we calculate the size of the spike at ajq. From (17.19) we have V
log® — ?(?
p itv ,,
3?= &(modg)
(26.7)
xl~
uniformly in q < 126 and in b reduced mod q. We recall that (26.8)
2 * ea(a&) = cq(a) = fx(q) b mode
when a is reduced modg. Hence \ S (y,a J g )— n(q)\jf]/(p(q)\
< 1+
y log «-— ., 8 ?(?) 35=vDk (mvodg)
2 * bm odq
(26.9) cp(q)xl~32.
< The simplest sum with a spike is F(a) =
By the identity (25.18),
(26.10)
^ « (« 4
m^x
h
w
y^x
) m
l2Sq - 1cp(q)xl-3i
(26.11)
xh*.
We now have /A\?) F ( m , S » ( a l q + P ) - V M F 3(j3) < x l ^ [ \ S ( a l q + P ) \ * + ^ ? 3(q) ? a(?) (26.12) and the integral in eqn (26.1) is , 2
(Q+2)/(Q+D
+
f
Ir majore Ir
.
|*%)|2da) +
I K Q + l)
+ °{
2
\ I r m ajor
I,
'( -? J
dot .
(26.13)
The integral in the first error term in (26.13) is 2 log2jJ
2J^x
(26.14)
112
ZEROS A N D PRIME N U M B ERS
4.26
by the prime-number theorem or by (8.19), and the integral in the second error term is ,2 < f |P(j8)|2 d/3 = x. (26.15) Using (8.16) for 2
?(q)> we see that the error terms in (26.13) are
both
(26.16)
Finally we must work out the first integral in (26.13). Since ^ (a )= e j ^ + a ) - l < e(a) — 1
1 ^ ||a||
we have * eqr( - a rx)
J JP8(i8)c(-®j8)dj8-Ji,8|a-|:Je(-®a[)da
when qr < 126. The integral over /3 on the left of (26.18) is (26.19)
%x{x— l),
and the integral in eqn (26.1) is now seen to be 2
'* eg( - ^ V ( f f ) ! 0 (x*\ 2*
< »■ »>
Q^V20 amodg
We can replace the sum over q ^ W by one over q from 1 to infinity with an error term 0(xH~2&). Now
U—1 am odg
s(q)
/-<
22 g
d\x
d\x d\q
dix{qld)iJ?{q) '
* ■*'
2
g^O(modfZ) r
1
\ n l-i
n + ( p» - wny
\
p
(p - 1)3
( p - i ) 3 ( ^ - i ) 3+ i
4.26
I. M. V I N O G R A D O V ’ S T H R E E - P R I M E S T H E O R E M
113
which is zero if x is even and positive if x is odd. We put A {x ) for the constant in (26.21), so that the left-hand side o f (26.1) is \ A (x)x2-\-0(xH~z),
(26.22)
which is non-zero for x odd and sufficiently large. Since primes less than xl ~2 can contribute at most xH” 1 to the sum on the right of eqn (26.1), and I ^ lo g p ^ I— 2log?
(26.23)
if x ^ p ^ xl~2, we can assert that the number of solutions of (26.2) is i A ( x ) x H - s+ 0 ( x H ~ i logl),
(26.24)
ail expression which again is seen to be non-zero for x odd and sufficiently large. The formula (26.24) is the famous theorem of I. M. Vinogradov.
27 H A L A S Z’ S METHOD ‘And I know it seems easy,’ said Piglet to himself, ‘ but it isn’t everyone who could do it.’ II. 18
W e have been concerned with estimating how often various sums S are large. The large-sieve results that we have had gave an upper bound for the sum o f |$|2 over different values of a parameter. The upper bounds we obtained contained two terms; the first was the maximum o f \S\2 and the second the mean square o f |S| multiplied by the number of values of the parameter. Thus if N
S(t) =
2 m=1
(27.1)
then (18.29) gives
J
r= l
|£(y|3 < iY lo g 2V 2 |o(m)|a+ (2 1 ,-!7 1 )lo g ^ 2 M m)l2»
(27-2)
where the points tr are at least (log N )^ 1 apart and lie between and 2\, The second term in (27.2) corresponds to the mean value o f |$(£)'|2 and the first to its maximum. Halasz’ s method sometimes enables us to count the number o f values o f r for which \S(tr) \ is large, without the second term’s being present on the right of (27.2). Halasz’s method is based on the lemma of Chapter 7, f c,(u,fW) < ||u||( % |c,|2W max f |(fW,f
(27.3)
In the proof of (27.3), we assumed that the vectors u, had finite dimension N , but the proof remains valid when the dimension is infinite, provided that each vector involved has finite norm. We choose the coefficients cr so that ^ .(u ,^ ) is real and positive. For (7.15) we took c},to be the complex conjugate of (u, fw), but here we give c,, unit modulus,
4.27
H A L A S Z ’ S M ETH OD
115
so that the second factor on the right of (27.3) is R*. I f each term on the left o f (27.3) is at least V we can square (27.3) to give i?2T/2 < P||u||2 max f |(fW,ffe>)| < -B||u||2(max||f(,')||2+ ( P —l)max|(fW,ffe))|], \ r } r,g
(27.4)
r^q
We can now deduce Halasz’ s lemma. L e m m a . Let
u,
for r = 1, 2,,.., R .
f(R) be vectors of finite norm with |(u,fM| > V
(27.5)
R < 2F~2||u||2 m ax ||fW||2 l
(27.6)
F2 ^ 2||u||2max|(fw,f (9))|.
(27.7)
Then
provided that
We shall apply (27.6) with (u,fW) = where sr = or-\-itr, 0 ^ ar <
2
Following Montgomery (1969a), we take
um = 0
iiff 1 1
fm = e - mlNm - ar<--iir
Here,
(27.8)
m^N
||u||2 < e2 f
l
(27.9)
for all m ^ 1.
(27.10)
|a(m)|2
(27.11)
CO
and
||f«||a < 2 e_2m,iV = W + 0 ( 1). 1
(27.12)
Writing a for o-,.+crg and t for tq— tr, we have 00 (f(r), f(^) =
^
0— 2/«/iV^—cr—it
m —1 2 + ico
=
f 2771
J 2 —ico
by the integral transform (23.6).
r(w)(%N)w£(w-{-<j-\-it) dw
(27.13)
116
ZE RO S A N D PRIME N U M B E R S
4.27
Before estimating the integral in eqn (27.13), we move the line of integration to the contour 0 consisting of C'1: the C2' the the Gs : the
line segment (—ico, —i(logiV)-1], semicircle, centre the origin, radius (logA7)-1, to the right of imaginary axis, line segment [i^og-A^'+ico).
A residue
(27.14)
r ( l — cr~ i t ) ( ^ N ) 1~'J- it
accrues from the pole of formula in the form (20.3):
a tw + ij+ ii = 1. We recall Stirling’s
!r( A+k)| = e-^H|A+ir|^{(27r)i+0(|r|^)},
valid when A
(27.15)
|r|-. Hence if \t\ > logA7
(27.16)
the residue (27.14) is bounded. Next we need a bound for |£(A+ir)|. Erom the approximate func tional equation of Chapter 21 we have for 0 ^ A ^ |r| ^ 1 0 |£(1— A—ir)|2 <
2
m<0(|r|)
d { m ) m x~1+
-H K A + iT + ^ i)}
+ max
r {i(l-A - i- r - f 77i)} -uv■'
M<2
m^o('|r|)
< M Alog2M,
(27.17)
where we have used (20.4) and (2.14), and partial summation. The functional equation and the estimate (20.4) now give |£(A+ir)|2 < M ^ lo g ^ r l,
(27.18)
and (27.17) and (27.18) together give m + i r ) \ <|Tp-A>logM
(27.19)
uniformly i n O ^ A ^ l , |t | ^ 10. We have now shown that
|(f«,ffa>)| < 1 + J | r^ )| | ?(ty + A + ir)| | P 7|,u |dw| C CO
< l-j-|£|*log|£|logiV+ J e_-7I"rr “ “(l + jr—£|=l og(j r—
dr
2
< |i|*logaJV|«|,
(27.20)
4.27
HALASZ’S
m ethod
117
where we have used (27.15) for the gamma function and (27.19) for the zeta function. We can now restate the condition (27.8) as T < T0)
(27.21)
where Ta is such that equality holds in (27.8) when we use (27.20) in substituting for the scalar product. Hence F2 > f
|a(m)|a2 * lo g O T .
i
(27.22)
Clearly when T > Ta we must divide up the range for T into intervals of length at most Tg. Repeated application o f the inequality (27.7) gives us
IT \ R < (m +A y-*
n
2
\a{m)\m.
v*0 / m= 1 When we substitute for T0 we have the result which follows: T h e o r e m . Let
If
(27.23)
N
< 3 = 2 \a(m)\2. m=1
(27.24)
2
(27.25)
^ F
m —1
for s =
sR, where sr = ar-\-itr with 0 < cr,,. ^ J and T >
\ K - t a\ > lo g iV
(27.26)
for q ^ r, then R <^ G N V -*+G P N T V -*lo ffN T ,
(27.27)
the implied constants being absolute.
The form of the second term in (27.27) arises from our choice o f functions f^ ; it is larger than the first term unless (27.22) holds with T in place o f T0. A plausible conjecture is that R < G N V -2
whenever
F2
GTS
(27.28) (27.29)
for any fixed 8 > 0. The use of the zeta function to prove (27.27) is a curious feature of Halasz’s method. I f Lindelof’s hypothesis is true, we can take the line of integration in (27.13) to Re w — l . with the effect o f replacing T\ in (27.22) by JV'Syg for any e > 0. This is an improvement for T > N (and if T < N then (27.28) follows trivially from (27.2)), but it is still a long way from weakening the condition on T0 to (27.29).
28 GAPS B E T W E E N P R I M E N U M BERS ‘ I shall do i t ’, said Pooh, after waiting a little longer, ‘ by means of a trap. And it must be a Cunning Trap, so you will have to help me, Piglet.’ I. 56
F i r s t we jn’o v e a theorem on the zeros of £(s), replacing the large sieve
(19.26) by Halasz’s method in the work of Chapter 23. We shall use the notation of that chapter with Q = 1, so that only zeros of the zeta function are considered. The definition o f class (i) and class (ii) zeros remains as before. We pick representatives o f each class o f zeros in such a way that their imaginary parts differ by at least 21, where l = \ogT,
(28.1)
but the representatives are in number ;> l ~2times the zeros in that class. We suppose a > f , since the result (28.19) which we obtain below improves on Ingham’s theorem only for a > f . The parameters X and Y will satisfy
x ^ y2;
im Y ^ T 2
(28.2)
In the definition (23.14) o f a class (ii) zero p = j8+i y, l - f l + im
J
■
l[pJr io)M(pJr w)Ywr(iv) dw > §7t,
(28.3)
ito
the parts of the integrand with |Imt«| > 100? give less than ^ (if I is sufficiently large). The integral of |.T(|-+ii)| converges rapidly so, for (28.3) to hold, there must be some t with \t— y\ sj 100? for which m + i t ) M ( i + i t ) \ > c Y P -i,
(28.4)
where c is an absolute constant. We pick as representatives o f the class (ii) zeros a sequence o f values o f t satisfying (28.4). By (22.22) the number of these t with |£(f+if)| > U,
(28.5)
where we choose U below, is < T U - H 5.
(28.6)
4.28
GAPS B E T W E E N P R I M E N U M B E R S
119
Otherwise we have |i¥(i+ii)| > V = c U - lY « - ! ,
(28.7)
and by (27,28) the number of such t is
We choose
jj
< X V - H + X T V - 9? .
(28.8)
__ ^Y-i/ioys(2a-i)/io;
(28.9)
V = eX1'10^ 2*-1)'5,
(28.10)
and on adding (28.6) and (28.8) and multiplying by I2we see that class (ii) zeros number
^ jf2/5y-e(2a~i)/5^9^_jy4/5y-2(2a-i)/5^3;
(28.11)
the second term in (28.11) being less than the first provided Z 2r « 2“- « < TH30.
(28.12)
A zero is o f class (i,r) if j J a(m)m~Pe~mir j > {20(?-a+ l ) } - 1. I m elr
(28.13)
I
We pick representatives and apply (27.28) with G — 2 \a(m)\^m~2ae~2mlY melr
(2>'Y)1- 2aexp( — 2'')l3.
(28.14)
The number of representatives is thus < »*4(2'T)2- 2aexp(-2'-+1)P + f12(2’T )4- 6“ 21e x p ( - 3 . 2r+1)l13. (28.15) Summing over r and multiplying by I2, we see that there are at most < r 2- toi6+ Z 4- 6“ T P
(28.16)
class (i) zeros. Choosing X = 21^a“_1)^“4+“_i), Y
=
y « 5 o;-3)|{aa+a:-l))
(28.17) (28.18)
we find (28.12) is satisfied for 0 ^ a ^ 1, and that the number N(a, T) of zeros p = jS+i y o f £(s) with j8 ^ ot and |y( ^ T satisfies the relation N(a, T) < T«5<x-3Xl-M ))Kas+a-l)^7 (28.19) for f ^ a ^ 1. The result (28.19) is also true for \ ^ a ^ | b y Ingham’ s theorem. We now sketch tlie proof o f our theorem on gaps between prime numbers. Theorem. Let e be a real number greater than
Then whenever a: is
sufficiently large, there is a prime p with x < p ^ x - ^ x 0. 853518 x
i
(28.20)
120
ZEROS AN D PRIME NUM BERS
4.28
Such a result was first proved by Hoheisel (1930) with c a little less than one. Ingham (1937) obtained the result with c > f and indicated how to replace f with a smaller number by improving an upper bound for |£(-§-f-i?)|. Several authors achieved this by means of intricate arguments. Recently Montgomery (19696) obtained the result for c > § by the method given here, but with a less efficient use o f the Halasz lemma; the improvement to was seen by the author in preparing the present exposition. As we have seen, Montgomery’ s method rests on the Halasz lemma, and thus on bounds for |£( 1+i?) |. As with Ingham’s result c > §, improvements at -|+i? improve the constant, in that a good estimate for the mean of a higher power than |£(-J+it) |4 would decrease the estimate for class (ii) zeros, both in (28.19) and in Ingham’ s theorem. However, even if we knew Lindel5f’s hypothesis, we should only be able to deduce (28.20) for c > It has long been conjectured (Cramer 1936) that for large x there is always a prime p with x < p ^ a;+ 0(log2a;),
(28.21)
but there seems no chance of approaching this conjecture by present methods. There are two essentials for a proof of (28.20) with c < 1: a zerodensity theorem such as (28.19) and a result on zeros of £(s) with /3 close to 1. We shall assume that £(s) has no zeros p = /3+iy with > l-4 {lo g (| y | + c)}-* ,
(28.22)
where B < 1. The inequality (28.22) is proved by Hadamard’s double height method just as (13.12) was, but the proof uses such inequalities as |£(l+tt)| < l o g c(|*|+e)
(28.23)
with c < 1. No better way to prove bounds for |£(l-f~i?) |is known than to replace exp(—iilogm ) by exp{—itP(m)}, where P(m) is a polynomial arising from the first few terms in the expansion of the logarithmic series. Since m runs through integer values, the resulting sum depends only on the fractional part of ^ / tt. More direct arguments fail, because t is much larger than any other parameter involved. After this transformation we must use the intricate methods of I. M. Vinogradov; W eyl’s simpler approach gives only l£(£+i*)l ^ io g '^ l+ e y io g lo g ^ l+ e 2),
(28.24)
where any higher power o f loglog t than the first would suffice for the application. The proof of (28.22) and (28.23) occupies one and a half chapters of Titchmarsh (1951).
4.28
GAPS B E T W E E N P R IM E N U M B E R S
121
While proving the prime number theorem in Chapter 16 we saw that # * )_ * -
J
-+ o K ^ ),
Iy \ < T
'
P
(28.25)
'
where T is chosen less than x with the property that each zero p = /3+iy of l(s ) has
\y— T\ > l o g 2 \
(28.26)
the sum in (28.25) being over all zeros p of £(s, x) with \y\ < T . Hence ftx+li)-t(x)=h+
2
xp~ { x + h ) p + o ( XP j ,
Iy\ < T
where we have written
'
P
A = logo:.
(28.27)
'
(28.28)
By the mean-value theorem the sum over zeros in (28.27) is in modulus <
2
H x + B h f-1
2
xP-1
(28.29)
Iy \ < T
\y\
for some 9 in 0 < 9 < 1. To estimate the sum in (28.29) we divide the interval [0,1] into ranges [0, i], [|, i+ A ^ 1],..., [| + rA -\ \ + { r + 1)A-1] ,.... The number of zeros p = /3+i y with a ^ j3 ^ a+A -1 is y*.*
(28.30)
this being by (28.19) for a ^ f and by Ingham’s theorem (23.29) for \ ^ a jSC f . The sum in (28.29) is thus < 7iA28 max (ay-M/ejP-i,
(28.31)
\y\< T
the maximum being over zeros p with |y| < T . With T = x5/12- 8,
(28.32)
where 8 > 0, the expression in (28.31) is < 7jA28 max
< 7iA28expfSA^A-*)},
(28.33)
Iy\
which is o (1) as x tends to infinity; here we have used (28.22). The error term in (28.27) is also less than h when h > a:7'12+8A2.
(28.34)
I f (28.34) holds with a sufficiently large constant then tf,(x+h)— ^ x ) > {£—o(l)}A.
(28.35)
122
ZE R O S A N D P R IM E N U M B E R S
4.28
Finally we note that the prime powers up to x-\-h contribute (28.36) to the sum >/i(x-\-h)— tfi(x), and so for sufficiently large x 2
x
loEP >
(28.37)
and we have proved (28.20) when we choose § = i(c
tt )'
(28.38)
NOTATION 2 .n V
indicate a stim ox1 a product over primes only (Chapter 1)
V
(m , n )
highest common factor of the integers m and n (Chapter 1)
m = n (mod g) m —n is a multiple of q (Chapter 1) e(a) = expZwia. ea(a) = exp(27ria/g) (1.4) a
1
sum over a set of representatives of residue classes modg
m od q
2* amodg 9(m) cq(m) s = cr+ ii x (m )
d(m) A(m) >fi(x) fi x ) < g(x)
fix) t(x ) X o (m )
M IN I
H(a) L (s ,X)
£(«) S(a) 7t ( x )
a/q r(s) £(s)> £(s’ X) p
= /S+iy
< A (« . x) ip(x;q,a) Sx X* Xiiiodg u = A+ ir 0(u) M( 8 , x ) N ( X )
(Chapter 1) sum over a set of representatives of reduced residue classes mod q (Chapter 1) Euler’s function (1.10) Ramanujan’s sum (1.11) complex variable (1.10) a Dirichlet’s character (Chapter 1) the number of divisors of m (1.23) Mobius’s function (1.25) logp if m is a prime power p a, otherwise 0 (1.31) sum function of A(m) (2.1) \f(x)\ = 0(g(x)) (2.3) the conductor of a character (Chapter 3) Gauss’s sum (3.7) a trivial character (Chapter 3) largest integer not exceeding a (4.3) distance of a from the nearest integer (4.4) the saw-tooth Fourier series (4.5) Dirichlet’s £-function (5.8) Riemann’s zeta function (5.9) an exponential sum (Chapter 6) number of primes up to x (Chapter 6) a rational number in its lowest terms (Chapter 6) Euler’s gamma function (11.1), (11.12) functions occurring in function equations for £(s) and L{s, x) (12.2), (12.3) a zero of £(s) or of £(s,x) (Chapter 12) sum function of A(m)x(m) (17.1) sum function of A(m) in the arithmetic progression a (modg) (17.18) a character sum corresponding to the exponential sum S(a) (18.4) a sum over proper characters modg (Chapter 18) auxiliary complex variable (Chapter 20) all the ‘ junk’ in the functional equation (20.2) partial stun for the inverse of L(s, x) (23.2) number of zeros of L (s ,x ) m a rectangle (Chapter 23)
BIBLIOGRAPHY The epigraphs are from I. Winnie the Pooh II. The House at Pooli Corner B o m b i e r i , E. (1965). On the large sieve. Matliematika 12, 201-25. ------- (1972). A note on the large sieve. To appear. ------- and D a v e n p o r t , H . ( 1 9 6 8 ) . On the large sieve method. Abhandlungen aus Zahlentheorie und Analysis zur Erinnerung an Edmund Landau. Berlin. ----------------(1969). Some inequalities involving trigonometric polynomials. Annali Scu. norm, sup., Pisa 23, part 2, 223-41. C h a n d r a s e k h a r a n , K . and N a r a s i m h a n , R. (1963). The approximate functional equation for a class of zeta-functions. Math. Annaln 152, 30-64. C r a m e r , H . (1936). On the order of magnitude of the difference between con secutive prune numbers. Acta arith. 2, 23-46. D a v e n p o r t , H. (1967). Multiplicative number theory. Markham, Chicago. E s t e r m a n n , T. (1948). On Dirichlet’s ^-functions. J. Lond. math. Soc. 23, 275-9. F o g e l s , E . (1969). Approximate functional equation for Hecke’s ^-functions of quadratic field. Acta arith. 16, 161-78. F r a n e l , J. (1924). Les suites de Farey et les problemes des nombres premiers. Nachr. Qes. Wiss. Gottingen, 198-201. G a l l a g h e r , P. X . (1967). The large sieve. Matliematika 14, 14-20. ------- (1968). Bombieri’s mean value theorem. Ibid. 15, 1-6. H a l A s z , G. and T u r A n , P. (1969). On the distribution of the roots of Riemann Zeta and allied functions I. J. Number Theory 1, 121-37. H a l b e r s t a m , H . and R o t h , K . F. (1966). Sequences, vol. 1. Oxford. H a p i d y , G. H. and W r i g h t , E. M. (1960). A n introduction to the theory of numbers. 4th edn. Oxford. H o h e i s e l , G. (1930). Primzahlprobleme in der Analysis. Sber. berl. math. Qes. 580-8. I n g h a m , A. E. (1937). On the difference between consecutive primes. Q. Jl Math. 8, 255-66. ------- (1940). On the estimation of N(a, T). Ibid. 11, 291-2. J e f f r e y s , H . and J e f f r e y s , B. (1962). Methods of mathematical physios. Cambridge. L a n d a u , E. (1927). Vorlesungen iiber Zahlentheorie. Leipzig. L i n n i k , Y u . V . (1945). On the possibility of a unique method in certain prob lems of additive and multiplicative number theory. Doklady Alcad. Nauh SSSB,, ser. mat. 49, 3-7. ------- (1964). All large numbers are sums of a prime and two squares (a problem of Hardy and Littlewood) II. Am. math. Soe. Transl. (2) 37, 197-240. --------- and R e n y i , A. (1947). On some hypotheses in the theory of Dirichlet charactex’s. Izv. Alcad. Nauk SSSB, ser. mat. 11, 539-46. M o n t g o m e r y , H . L . (1968). A note on the large sieve. J. Lond. math. Soc. 43, 93-8. ------- (1969a). Mean and large values of Dirichlet jjolyxiomials. Invent, math. 8, 334-45. ------- (19696). Zeros of i-functions. Ibid. 8, 346-54. ------- (1971). Lectures on multiplicative number theory. Springer.
B IBLIO G RAPH Y P o ly a ,
125
G. (1918). Uber die Verteilung der quadratischen Reste und Nichtreste.
Nachr. Ges. Wiss. Gottingen, 21-9. P b a c h a b , , K . (1957). Primmhlverteilung. Springer. R o th , K . F. (1965). On the large sieves of Linnik and Renyi. Mathematika 12, 1-9. T i t c h m a b s h , E. C. (1951). The theory of the Riemann zeta-function. Oxford. V i n o g r a d o v , A. I. (1965). On the density hypothesis for Dirichlet i-functions. Izv. Akad. Nauk SSSB, ser. mat. 29, 903-34. V i n o g b a d o v , I. M. (1954). The method of trigonometric sums in the theory of numbers (transl. A . Davenport and K . F. Roth). Intersoience, New York. ------- (1955). A n introduction to the theory of numbers (translation). Pergamon, Oxford.
INDEX abscissa o f covergence, 18 additive function, 5 approxim ate functional equation, 84-93, 106, 116 arithm etical functions, 2 additive, 5 m ultiplicative, 2 totally m ultiplicative, 2 B om bieri’s theorem , 35, 103-9 ch aracters: con du ctor of, 11 defined, 4 induced, 11 propriety of, 11 sieved, 74-6, 79-82 trivial, 3, 13 classification o f zeros, 99-100 congruence 1 D irich le t: characters, see characters divisor problem , 8 polynom ials, 80, 84 series, defined, 18 divisor fu n ctio n : averaged, 7 -9 defined, 4 elem entary proofs, 69, 107 E uler’s : function defined, 3 product., 20, 52 exceptional zero, 60-5, 70—72, 76 exponential maps, 1 exponential s u m : defined, 24 sieved, 28-31, 77 spikes of, 24-6, 110-11 F arey: arcs, 110 sequence, 36, 74, 110 F o u rie r: series, 14, 37, 40-1 transform, 79 Franel’s theorem, 36-9, 74 functional equation, 20, 41, 45-50, 69, 84, 116 Gallagher’s : first lemma, 76-8, 81—2 second lemma, 80
Gauss, 11, 22 Gauss’s sum, 11—13 H adam ard’s product, 50-4, 69 H alasz’s m ethod, 30, 114-18 ineffective constants, 62, 72, 110 Ingham ’s theorem , 98-102, 105, 118-19, 121 integral functions, 50, 85
108,
Jensen’s form ula, 51—2 _L-functions defined, 19-20 L in delof hypothesis, 93, 117, 120 m ajor arcs, 110—12 m inor arcs, 107, 110-11 m od u la r: functions, 43 relation, 43 M ob iu s: fun ction defined, 4 inversion, 5, 33, 38 m ultiplicative fun ction defined, 2 Planoherel’s identity, 79-80 P oisson ’s sum m ation form ula, 40-42 P o ly a ’s theorem , 14-17, 50, 63 prime num ber th eorem : for an arithm etic progression, 23, 62, 71-2, 74, 103, 110 form s of, 21-2, 23, 39, 55, 62 proved, 66-72 propriety o f characters, 11 R am an ujan ’s sum, 3, 13, 24, 26, 38, 111 representative zeros, 100, 119 residues: classes of, 1 -2 com plete set of, 2 , 10—11 reduced, 3, 10 R iem ann hypothesis, 20, 52 Schinzel’s hypothesis, 23-5 Selberg, A ., 21-2, 26-7 Siegel’s theorem , 62-5, 71-2, 110 Siegel—W alfisch theorem , 72, 104, 106 s ie v e : for characters, 74-6 for exponential sums, 28-31, 77 hybrid, 79-83, 93, 101
128
IN DEX
sieve (em it.) large, 27, 28-35, 38, 74-8, 114, 118 upper bound, 25-7, 32—5 sifted sequence, 25, 32-4 spikes o f an exponential sum, 24-6, 110-11 Stirling’ s formula, 47, 51, 52, 84, 99, 116 totally m ultiplicative functions, 2 trigonometries sums, see exponential sums uniform distribution, 73-4 V in ogradov, A. I., 98-9, 103
V inogradov, I. M., 8, 14, 25, 57, 107, 113. 120 estimate, 107—9 m ethod, 57, 120 theorem, 113 W ey l’s criterion, 73 zero-density theorems, 98-102, 105, 108-9, 118-20 zero-free region, 56-65, 120 zeta-function defined, 19-20
r