Vistas of Special Functions
Vistas of Special Functions
SPECIAL FUNCTIONS
This page intentionally left blank
VISTAS OF
SPECIAL FUNCTIONS
World Scientific
N E W J E R S E Y • L O N D O N • S I N G A P O R E • B E I J I N G • S H A N G H A I • H O N G K O N G • TA I P E I • C H E N N A I
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher.
ISBN-13 978-981-270-774-1
ISBN-10 981-270-774-3
Printed in Singapore.
v
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Preface
This book is intended for aspirant readers who are eager to have basic
knowledge of special functions in an organic way. We have kept paying at-
tention to make an order in various equivalent statements on special func-
tions. A unique feature is that the reader can gain a grasp of (almost) all
existing (and scattered around) formulas in the theory of gamma functions
etc. in a clear perspective through the theory of zeta-functions. Thus,
this is a book of special functions in terms of the zeta-functions. Reading
through this book, the reader can master both fields efficiently. Here a
hunter looking for two rabbits gets two.
Here are some descriptions of the contents.
In Chapter 1, we present a unified theory of Bernoulli polynomials with
all equivalent conditions properly located. We have revealed that the dif-
ference equation (DE) satisfied by the Bernoulli polynomial corresponds
to differentiation while the Kubert identity (K) corresponds to integration
(the Riemann sum into equal division). This new view point makes the
whole theory very lucid.
In Chapter 2 we shall present rather classical and standard theory of the
gamma and related functions. Classical as it looks, we shall provide some
very unique features of the Euler digamma function from which we may
deduce the corresponding properties of the gamma function. Especially, we
shall give three proofs of the remarkable formula of Gauss on the values
of the digamma function at rational arguments. One is classical and is
presented in Chapter 2. Other two proofs are more original given in Chap-
ter 8, one is the limiting case (Theorem 8.2) of the Eisenstein formula in
its genuine form (a theorem due to H.-L. Li, L.-P. Ding and M. Hashimoto,
describing a basis element in terms of another basis of the space of periodic
Dirichlet series), the other is the theorem of M. Hashimoto, S. Kanemitsu
vii
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
and M. Toda about the equivalence between the finite form of the value of
the Dirichlet L-function at 1 and the formula of Gauss.
In Chapter 3, we shall present the theory of the Hurwitz zeta-function.
The main ingredient is the integral representation for its partial sum. This
is to the effect that once we have an integral representation as the one we
have, we may immediately draw information for the derivatives, i.e. we
have an inheritance of the information. The integral representation for
the partial sum is so informative that it contains all information we need
(Theorem 3.1). The versatility of this result will be developed in Chapter 5,
where through Lerch’s formula, we transfer the results on the Hurwitz zeta-
function to those on the gamma and related functions. Especially, the
asymptotic results established in Chapter 3 will immediately transfer to the
Stirling formula and other asymptotic formulas for relatives of the gamma
function.
In Chapter 4, we shall present the theory of Bernoulli polynomials
through the negative integer values ζ(−n, z) of the Hurwitz zeta-function.
Here we shall establish only three statements, i.e. the Fourier series (H),
the difference equation (DE) and the Kubert identity (K) from any of which
we may complete the theory following the logical scheme in Chapter 1.
In Chapter 5, first we shall reveal the power of theorems in Chapter 3
to exhibit what the Dufresnoy-Pisot type uniqueness theorem means. Then
we shall go on to presenting the first circle (krug p’iervyi) which connects
various identities between gamma and trigonometric functions to the func-
tional equations (zeta-symmetry) of the zeta-functions . Thus we shall show
that everything comes from the functional equation. A remarkable notice
is that such trigonometric identities like the infinite product for the sine
function or the partial fraction expansion for the cotangent function are
equivalent to the functional equation, thus revealing why Euler succeeded
in solving the Basler problem.
In Chapter 6, we shall further pursue this zeta-symmetry in relation to
the crystal symmetry through the Epstein zeta-function. We surpass the
preceding results by introducing the signs and giving the Chowla-Selberg
type formula (based on the Mellin-Barnes integrals) and provide a quick
means for computation of the Madelung constants.
In Chapter 7, we shall provide rudiments of the theory of Fourier series
and integrals to such an extent that is sufficient for applications and reading
through this book, for the sake of the reader who wants to learn it quickly.
Chapter 8 is, so to say, a discrete version of Chapter 7, i.e. the finite
Fourier series (transforms). Through this we make clear the orthogonality
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Preface ix
of characters and other bases of the space of Dirichlet series with periodic
coefficients, giving rise to the theorem mentioned above. We can naturally
extend our method to develop the similar theory for higher derivatives of
the Dirichlet L-function, including Kronecker’s limit formula. But because
of limitation of time, we cannot go further.
Appendix A gives the very basics of the theory of complex functions.
We present mostly results only, and the interested reader should consult a
standard book for their proofs. We shall give, however, some details on the
use of residue theorem.
Appendix B assembles summation formulas and convergence theorems
used in the book. Especially, the Fourier series for the first periodic
Bernoulli polynomial is so essential and important, we give two proofs,
one depending on ordinary Fourier theory (Chapter 7) and the other on
the polylogarithm function of degree 1, where we apply the theorem of
Abel and Dirichlet in place of Fourier theory.
As is explained above, Chapters 1 and 4 are parallel, so are Chapters 2
and 5. To understand Chapters 4 and 5, one should read Chapter 3 first.
If one finds some difficulties, then one is referred to Appindices A and B.
Chapters 7 and 8 can be read independently, but it will be more instructive
to read both in parallel. Chapter 6 can be read separately which requires
more knowledge of Bessel functions. Because of lack of time, we could not
state much about them.
This publication was supported by Kinki University Grant for Publica-
tion, No. GK04 in the academic year 2006. The authors are thankful to
Kinki University for their generosity of this support. They also would like
to thank Ms. Chiew Ying Oi who helped them all through the process with
her efficient editorial skills. And toward the end of the process Ms. Zhang
Ji supported us and we would like to express our heartily thanks to her.
The authors would like to express their hearty thanks to their close
friend Professor Y. Tanigawa for his constant support, encouragement, and
stimulating discussions. The first author would like to thank his close friend
Professor Heng Huat Chan for his enlightening remark on the equivalent
statements to the functional equation, thanks to which he got motivated
enough to start writing this book. The second author was naturally got
infected the passion of the first. Thanks are also due to Ms. L.-P. Ding and
Mr. M. Toda for their devoted endeavor, without their enthusiastic help,
the book would have not been risen out.
the authors
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Contents
Preface vii
xi
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Bibliography 207
Index 213
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Chapter 1
Abstract
1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
n!
Bn(k) (x) = Bn−k (x). (1.2)
(n − k)!
X
n−1
n
Bk (0) = 0, n ≥ 2. (1.3)
k
k=0
i.e. (1.7), where, by umbral calculus, we mean that after expanding the
binomial, the exponent of B is to be degraded to subscript.
Theorem 1.1 The defining conditions in Definition 1.1 are equivalent to
conditions (DE)-(H).
(DE) {Bn (x)} are (principal) solutions of the difference equation
As will be indicated in Remark 1.1, it follows from (1.5) and (1.6) that
Bn (x) is a polynomial of degree n given by
Xn
n
Bn (x) = Bn−k xk . (1.7)
k
k=0
where B n (x) = Bn (x − [x]) , [x] being the integral part of x, for n ∈ N (in
the case n = 1, we should have (7.9)).
Then
1
Z(n, r) = (Br+1 (n + 1) − Br+1 ) , (r ∈ N) (1.11)
r+1
and
r
1 X r+1
Z(n, r) = Bk · (n + 1)r+1−k , (r ∈ N). (1.12)
r+1 k
k=0
Taylor expansion in |z| < 2π, given by the Cauchy product (or sometimes
called Abel convolution)
∞
! ∞ !
z X B k
X xl
exz = zk zl ,
ez − 1 k! l!
k=0 l=0
Fig. 1.3
we have (A).
(A)⇒(D0 )
For y 6= 0 we have
n
Bn (x + y) − Bn (x) X n
= Bn−k (x) y k−1 ,
y k
k=1
(A)⇒(U )
We note that the umbral calculus formula (U ) is the special case of (A)
with y = 0.
(U )⇒(A)
We have by (1.4)
n
X n
Bn (x + y) = Bn−k (x + y)k
k
k=0
n
X k
X
n k
= Bn−k y k−l xl
k l
k=0 l=0
n
n X
X n k
= Bn−k y k−l xl ,
k l
l=0 k=l
we obtain
n
X n
X
n l n−l
Bn (x + y) = y Bn−k xk−l
l k−l
l=0 k=l
n
X n−l
X
n n−l
= yl Bn−l−k xk , (1.15)
l k
l=0 k=0
by the change of variable. Now the inner sum is Bn−l (x) by (1.7).
(U )⇒(D0 )
k
X
0 k
Bk (x) = Bk−r r xr−1
r=1
r
X
k−1
k
= (s + 1)Bk−1−s xs
s=0
s + 1
X k − 1
k−1
=k Bk−1−s xs
s=0
s
= kBk−1 (x).
Z 1
Proof of Bn (x) dx = 0 for n ∈ N.
0
Z 1 n
X Z 1
n
Bn (x) dx = Bn−k xk dx
0 k 0
k=0
n
1 X n+1
= Bn−k
n+1 k+1
k=0
n+1
1 X n+1
= Bn+1−k .
n+1 k
k=1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Pn+1 n+1
Now the sum is k=0 k Bn+1−k − Bn+1 , which is Bn+1 (1) − Bn+1
R 1 (1.4); this is in turn Bn+1 (0) − Bn+1 by (1.16) and is 0. Note that
by
0 B0 (x) dx = 1.
(A)⇒(DE)
By (A)
n
X n
Bn (x + 1) − Bn (x) = Bn−k (1) xk − Bn−k xk
k
k=0
n
= (B1 (1) − B1 ) xn−1
n−1
= n xn−1
on using (1.13).
(G)⇐⇒(U )
X∞
Bn (x) n
(U )⇒(G): We form the generating function z and substitute
n=0
n!
(1.4) to get
∞ ∞
!
X Bn (x) n X 1 X n!
z = B k xl zn
n=0
n! n=0
n! k! l!
k+l=n
∞
! ∞
!
X Bk k X xl l
= z z
k! l!
k=0 l=0
z
= exz ,
ez −1
(H)⇒(K)
Substituting (1.9) into the right-hand side of (1.8), we obtain
m−1
X X∞ m−1
e2πi m r n 1 X 2πi r k
x
n−1 x+k n!
m Bn =− m e m .
m (2πi)n r=−∞ rn m
k=0 k=0
r6=0
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
(G)⇒(1.11)
Consider the generating function
∞
X Z(n, r)
fn (z) = z r+1 .
r=0
r!
X∞ n n X∞
1 X r r X 1
fn (z) = z k z =z (kz)r ,
r=0
r! r=0
r!
k=1 k=1
whence
n
X
fn (z) = z ekz .
k=1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
e(n+1)z − ez ze(n+1)z z
z z
= z − z − z,
e −1 e −1 e −1
we have
∞
X ∞
X
zr zr
fn (z) = Br (n + 1) − Br − z
r=0
r! r=0
r!
by (G), whence
∞
X z r+1
fn (z) = (Br+1 (n + 1) − Br+1 ) + nz;
r=1
(r + 1)!
but for r = 0:
Z(n, 0) = B1 (n + 1) − B1 − 1 = n,
and so
∞
X z r+1
fn (z) = (Br+1 (n + 1) − Br+1 ) .
r=0
(r + 1)!
Xr
(−1)n r + 1
Br = Z(n, r). (1.18)
n=0
n+1 n+1
We shall deduce (1.18) from (1.12) with the help of (1.19) which gives the
closed form for the Stirling number of the second kind (cf. (1.20))
k
1 X j k
S(n, k) = (−1) (k − j)n (1.19)
k! j=0 j
k
1 X k−i k
= (−1) in
k! i=1 i
1 k n
= ∆ 0 , 1 ≤ k ≤ n,
k!
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
k
X Xk
k k
(−1)j (k − j)n = (−1)j E k−j 0n
j=0
j j=0
j
= (E − 1)k 0n
= ∆k 0n ,
nn = E n 0 n .
The left-hand side is the same as that of (1.20), while on the RHS, we apply
E = ∆ + 1 formally n-times to deduce that
n
X Xn
n 1 j n
E n 0n = (∆ + 1)n 0n = ∆j 0 n = ∆ 0 (n)j ,
j=0
j j=1
j!
Xr r
(−1)n r + 1 1 X r+1
the RHS = Bk (n + 1)r+1−k
n=0
n + 1 n + 1 r + 1 k
k=0
Xr X r
1 r+1 r+1
= Bk (−1)n (n + 1)r−k
r+1 k n=0
n + 1
k=0
Xr X r
1 r+1 r+j r + 1
= Bk (−1) (r + 1 − j)r−k
r+1 k j=0
j
k=0
and
r+1
X
j r+1
(−1) = (1 − 1)r+1 = 0,
j=0
j
Example 1.2
n
X 1
Z(n, 2) = k2 =
(B3 (n + 1) − B3 )
3
k=1
1 3 1
= (n + 1) (n + 1)2 − (n + 1) +
3 2 2
1
= (n + 1)(2n2 + 4n + 2 − 3n − 3 + 1)
6
1
= n(n + 1)(2n + 1).
6
Compared with (S), the following is less well-known ([Com, p.155]):
Proposition 1.1
r+1
X
n
Z(n, r) = (j − 1)! S(r + 1, j) . (1.21)
j=1
j
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
n+1−j +j j
we rewrite (n + 1)j as =1+ to obtain
n+1−j n+j−1
r+1
X r+1
X
1
S(r + 1, j)(n)j + S(r + 1, j)n · · · (n − j + 2),
j=1
j j=1
the first term is Z(n, r) by hypothesis and the second can be written as
r+1
1 X
S(r + 1, j)(n + 1)j ,
n + 1 j=1
1
which is (n + 1)r+1 = (n + 1)r , on applying (1.20).
n+1
The second proof. We first prepare auxiliary results. First,
X n
j n+1
= , 0 ≤ k ≤ n. (1.22)
k k+1
j=k
which becomes
X j
n X n
n X
X
j j
xk+1 = xk+1
j=0 k=0
k k
k=0 j=k
P
n
n+1
by changing the order of summation. Since (x + 1)n+1 − 1 = k+1 xk+1 ,
k=0
we obtain (1.22) by comparing the coefficients of xk+1 .
Secondly, we also need the triangular recurrence formula for S(r, j):
P
r+1
On one hand, it is S(r + 1, j)(x)j , and on the other, it is
j=0
r
X
x · xr = x S(r, j)(x)j .
j=0
Since
x (x)j = (x − j + j)x(x − 1) · · · (x − j + 1)
= (x)j+1 + j(x)j ,
we have
r
X
xr+1 = (S(r, j − 1) + j S(r, j)) (x)j ,
j=1
by (1.22). Hence
r+1
X
n n
Z(n, r) = S(r, j) j! + S(r, j − 1) (j − 1)!
j=1
j j
r+1
X
n
= (j S(r, j) + S(r, j − 1)) (j − 1)! ,
j=1
j
it follows that
n n
Z(n, 2) = S(3, 1)n + S(3, 2) + 2! S(3, 3)
2 3
3 1
= n + n(n − 1) + n(n − 1)(n − 2)
2 3
1 2
= n(2n + 3n + 1)
6
1
= n(n + 1)(2n + 1).
6
Exercise 1.2 Prove that (H) implies (U ) under Euler’s identity (5.66).
Solution Since B1 (x) = x − 12 for 0 < x < 1, the unique polynomial that
coincides with B 1 (x) is B1 (x) = x − 21 = x + B1 . Denoting the right-hand
side of (1.9) by bn (x), we obtain
1 d
bn (x) = bn−1 (x),
n dx
whence
1 dn
bn (x) = B1 (x), 0 < x < 1.
n! dxn
1 d
Integrating 2 dx b2 (x) = B1 (x), we deduce that
Z
1 1
b2 (x) = B1 (x) dx = x2 + B1 x + C,
2 2
where
∞
1 1 2 X 1 1 1
C = b2 (0) = = 2 ζ(2) = = B2
2 2 π 2 n=1 n2 π 6
by (5.66).
Repeating this procedure, each time using (5.66), we arrive at (U ).
Now we shall follow Lehmer [Leh2] to deduce some of the above defining
conditions from (K). First we state a lemma.
Lemma 1.1 (Lehmer) For a given n ∈ N there is a unique monic
polynomial of degree n satisfying (1.8).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Pn (x) = b0 xn + b1 xn−1 + · · · + bn , b0 6= 0
n(n+1)
Now, from Z(n, 1) = 2 , it follows that
n
b1 = B1 .
1
n
r
The product of binomial coefficients is r k , and so
X
r−1 X r − k + 1
r−k
−n n r k−1 1
RHS = −m Bk m Bl mr−k+1−l .
r k r−k+1 l
k=0 l=0
Hence
n n
RHS = −m−n+r S1 − m−m S2 ,
r r
where
r−1
X r 1
S1 = Bk (1.26)
k r−k+1
k=0
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
r−1
X r−k
X
r 1 r−k
S2 = Bk Bl mr−l . (1.27)
k l l−1
k=0 l=1
r
r−k
Rewriting the product k l−1 of binomial coefficients in S2 by
r
r−l+1
l−1 r−k−l+1 , we obtain
r
X r−l
X
1 r r−l r−l+1
S2 = Bl m Bk
l l−1 r−k−l+1
l=1 k=0
1
= (Br+1 (1) − Br+1 (0)) − Br = −Br .
r+1
Hence
−n+r n −n n
LHS = −m (−Br ) − m Br
r r
or
−n r −n −n+r n
m (m − 1)br = −(m −m ) Br ,
r
whence
n
br = Br ,
r
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Solution Let
t ext
F (x, t) =
et − 1
and expand it into the Taylor series in t:
X∞
1
F (x, t) = Gn (x)tn , |t| < 2π (1.29)
n=0
n!
say.
(n) 1
We may determine a0 as follows: Replacing x by y and t by ty in
(1.29), we get
X∞
yt t 1 n 1 n
e = y G n t ,
eyt − 1 n=0
n! y
which leads, as y → 0, to
X∞
t 1 (n) n
e = a0 t
n=0
n!
1 (n)
(since y n Gn y → a0 ):
(n)
Hence a0 = 1 and Gn is monic.
Now,
m−1 ∞ m−1
1 XX 1 x+k n 1 X x+k
Gn t = F ,t
m n=0
n! m m m
k=0 k=0
m−1
X x+k
1 t
= e m t
m et − 1
k=0
x
1 t emt
= ,
m e mt − 1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
which is equal to
X∞
t 1 1
F x, = Gn (x) n tn .
m n=0
n! m
1 n
Hence, comparing the coefficients of n! t , we-conclude that
m−1
1 X x+k
Gn = m−n Gn (x),
m m
k=0
Proof. From (1.30), f (z) must be of degree m and may be put in the
form
m
X m
f (z) = aj z m−j , a0 6= 0. (1.31)
j=0
j
X
m−j
m−j
(z + 1)m−j − z m−j = aj z m−j−r
r=1
r
Xm
m−j
= aj z m−k ,
k−j
k=j+1
we find that
m
X m
X
m m − j m−k
f (z + 1) − f (z) = aj z .
j=0
j k−j
k=j+1
Hence
m
X X
m−1
m k
f (z + 1) − f (z) = aj z m−k . (1.32)
k j=0
j
k=1
Comparing (1.30) and (1.32), we see that it is enough to show that the
following system of m+1 linear equations in m+1 unknowns a0 , a1 , · · · , am
and f (z) has a unique solution:
X
k−1
k
aj = b k , k = 1, 2, · · · , m (1.33)
j=0
j
and
X
m−1
m
− aj z m−j + f (z) = am . (1.34)
j=0
j
This is indeed the case because the coefficient matrix is lower trian-
gular,
so that its determinant is the product of all diagonal components
1 2 m
0 1 · · · m−1 = m! 6= 0. Hence f (z) is determined by b1 , · · · , bm and am .
Now, comparing (1.33) with k = m and (1.34) with z = 1, we find that
am = bm by the condition f (1) = 0. Hence f (z) is determined uniquely
by b1 , · · · , bm .
Theorem 1.3 For each m ∈ N there exists a unique polynomial fm (z)
satisfying the conditions
fm (1) = 0. (1.36)
we obtain
X∞ 0 X∞ X∞
fm (z) m x2 Bm (z) m+1 Bm−1 (z) m
x = x exz = x = x ,
m=0
m! e − 1 m=0
m! m=1
(m − 1)!
whence we have
1 0
f (z) = Bm−1 (z), m ∈ N,
m m
and
f00 (z) = 0,
for any q ∈ N.
We follow Böhmer [Böh] to add (1.35) in the form
to obtain
q−1
X q−1
X
m k+1 m k
q fm+1 z+ =q fm+1 z+ + (m + 1)q m z m . (1.40)
q q
k=0 k=0
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
to deduce that
q−1
X
m k+1
fm+1 (qz + 1) − q fm+1 z +
q
k=0
q−1
X (1.41)
k
= fm+1 (qz) − q m fm+1 z +
q
k=0
= F (z),
say, whence
1
F z+ = F (z)
q
i.e. F (z) is periodic of period 1q . But, F (z) being a polynomial, we must
have F (z) = constant.
Hence differentiating (1.41), we infer that
q−1
X
0 m 0 k
qfm+1 (qz) −q fm+1 z+ = 0,
q
k=0
or
q−1
X
m k
q(fm (qz) + Bm (1)) − q fm+1 z+ + Bm (1) = 0,
q
k=0
(D ′ ) ⇔ ( A) ⇔ (U ) ⇒ (H )
⇐
⇓⇓
⇓
⇔
(G ) ⇐ (K )
⇑
⇓
(S )
⇑
(DE )
Fig. 1.4 Logical scheme
and
Y
Gi (x) = gk (x), 1 ≤ i ≤ n. (1.43)
k6=i
Proof. We note the following facts. Since fi0 (x) = ki xki −1 , we have, for
a primitive ki -th root of 1 and 1 ≤ l ≤ n − 1,
l(k −1)
fi0 ζil = ki ζi i = ki ζi−l .
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Also Gi ζil = 0 for k 6= i, 6= 0 for k = i. Hence by (1.43),
n
X
Gi ζil φi ζil = Gi ζil φi ζil = 1.
k=1
and all summands on the right are conjugate one another, so that the right-
hand side is the trace.
This is essentially used to find the dual basis in a finite extension of Q.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Chapter 2
Abstract
First, we develop the theory of the gamma function defined by (2.1), one
of equivalent conditions to be discussed in Chapter 5. The gamma function,
being the Mellin transform of e−x to be mentioned in §7.4, is defined by
the Eulerian integral of the second kind
Z ∞
Γ(s) = e−x xs−1 dx (2.1)
0
Γ (s + n + 1)
Γ(s) = , n ∈ N ∪ {0}. (2.2)
s (s + 1) · · · (s + n)
29
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
whose right-hand side is analytic for σ > −1. From (2.2) we see that
s = −n, n ∈ N ∪ {0} are simple poles and that the residues at these poles
are
(−1)n
Res Γ(s) = . (2.4)
s=−n n!
Also if we put s = n ∈ N in
then we have Γ(n + 1) = n!, which means that the gamma function is a
function which interpolates the factorial n!.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Γ(z + 1) Γ(z + n + 1) (z + 1 − r) · · · (z + n)
=
Γ(z + 1 − r) (z + 1) · · · (z + n) Γ(z + n + 1)
(z + 1 − r) · · · (z + n − 1)
= ,
(z + 1) · · · (z + n − 1)
we see that
Γ(z + 1) (−n + 1 − r) · · · (−1)
→
Γ(z + 1 − r) (−n + 1) · · · (−1)
(−1)n+r−1 (n + r − 1)!
=
(−1)n−1 (n − 1)!
X∞ Z ∞
(−1)n 1
Γ(s) = + e−x xs dx. (2.8)
n=0
n! s + n 1
whence in particular
1 √
Γ = π, (2.11)
2
R∞ 2
√
π
or the value of the probability integral 0
e−x dx = 2 .
2+1
e−x (t +1) dt.
0 t
Solution Recalling the fundamental theorem of calculus, we obtain
Z x
0 −x2 2
f (x) = 2 e e−t dt.
0
f 0 (x) + g 0 (x) = 0,
we conclude that
Z ∞ 2
2 π
lim f (x) = e−t dt = ,
x→∞ 0 4
i.e.
Z ∞ √
−t2 π
e dt = .
0 2
Γ(α) Γ(β)
B(α, β) = (2.16)
Γ(α + β)
√
and the value of the probability integral Γ 12 = π to deduce Wallis’
formula
2
π 2 · 4 · · · (2n) 1
= lim . (2.17)
2 n→∞ 1 · 3 · · · (2n − 1) 2n + 1
say, and
1
Z π2
Γ Γ (n + 1) 1
2
= B , n + 1 = 2 cos2n+1 θ dθ = S2n+1 .
Γ(n + 32 ) 2 0
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1
S2n S2n−1 n+ 2 1
1< < = =1+ →1
S2n+1 S2n+1 n 2n
as n → ∞. Hence
S2n
lim = 1. (2.18)
n→∞ S2n+1
Since
S2n Γ n + 12 Γ n + 32
=
S2n+1 Γ(n + 1)2
!2
n − 12 n − 32 · · · 12 Γ 21 1
= n+
n · (n − 1) · · · 1 2
2
(2n − 1)(2n − 3) · · · 1 π
= (2n + 1) .
(2n)(2n − 1) · · · 2 2
N !
X 1 √
lim log n − N + log N + N = log 2π, (2.19)
N →∞
n=1
2
we obtain
Γ(2n + 1) √
log 2
= log 22n − log 2n + 1 + log 2 − c + o(1), (2.22)
Γ(n + 1)
whence
√
2n 2n + 1
log = log 2 − c + o(1). (2.23)
n 22n
Since
1 2n 2 · 4 · · · 2n · 1 · 3 · · · (2n − 1) 1 · 3 · · · (2n − 1)
2n
= n 2
= ,
2 n (2 n!) 2 · 4 · · · 2n
√ q
the LHS of (2.23) is log 1·3···(2n−1)
2·4···2n
2n+1
, whose limit is log 2
π by Wallis’
formula. Hence
r
2 √
c = log 2 − log = log 2π.
π
Recalling the more exact form of (2.21), we may deduce Stirling’s for-
mula from the above (cf. Corollary 5.1):
1 √ 1
log Γ(x) = x − log x − x + log 2π + O . (2.24)
2 x
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Xr
2n n log 2n
β= −2 l , r= . (2.28)
pl p log p
l=1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
2n
Now by (2.25), n 2n , so that
n 2n
log 2 ≤ log ≤ π(2n) log 2n.
n
Hence
n
π(2n),
log 2n
so that for x ≥ 2, using (2.27), we obtain
h x i x
2 x
π(x) ≥ π 2 .
2 log 2 x2 log x
To prove the other inequality in (2.26), we recall the remark after (2.28)
to the effect that those prime p, n < p ≤ 2n do not divide the denominator,
Q
and therefore 2n n must be divisible by p. Hence
n<p≤2n
Y
2n
nπ(2n)−π(n) ≤ p≤ .
n
n<p≤2n
2n
We deduce from this, on using (2.25) in the form log n n, that
whence that
x x
r
X 1
π(x) log x − π r
log r
x
2 2 2l
l=0
Γ(n + z)
lim = 1, 0 < Re z < 1 (2.31)
n→∞ Γ(n) nz
are equivalent.
Proof. Γ(s) being meromorphic over the whole plane, it suffices to con-
sider the case 0 < z = x < 1.
First we show that (2.1) implies (2.31). In (2.1) with s = n ∈ N put
t = nu to obtain
Z ∞
e−nu un−1 du = Γ(n) n−n . (2.32)
0
Note that
Z 1 Z 1 Z 1
1 d −nt n 1
e−nt tn−1 dt − e−nt tn dt = (e t ) dt = e−n . (2.34)
0 0 n 0 dt n
1
by (2.31). Expanding x(x+1)···(x+n) into partial fractions, we obtain
n
X
x n 1
k
Φn (x) = n (−1) ,
k x+k
k=0
Hence
Z n n
x−1 t
Φn (x) = t 1− dt.
0 n
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1
Now choose 0 < ε < x+2 and divide the interval [0, n] into [0, nε ]∪[nε , n].
For 0 ≤ t ≤ nε we have
t
n log 1 − = −t + O t2ε−1 ,
n
and so
Z nε n Z nε
t
tx−1 1 − dt = tx−1 e−t dt + O nε(x+2)−1 ,
0 n 0
while for nε ≤ t ≤ n,
t
n log 1 − ≤ n log 1 − nε−1 ≤ −c nε , c > 0,
n
so that
Z n
ε
= O nε e−cn .
nε
It follows that
Z nε
Φ(x) = tx−1 e−t dt + o(1),
0
R ∞ x−1 −t
which tends to 0 t e dt. Thus from (2.37) we conclude (2.1).
and that
Z ∞
lim f (x) e−nx dx = 0. (2.39)
n→∞ 0
Hence, in particular
Z ∞ Z ∞
f (t) e−nt dt = G(∞) = n F (t) e−nt dt,
0 0
which is (2.38).
Now, to prove (2.39), we note that F (t) is differentiable on R+ and, a
fortiori, continuous in the right neighborhood of 0. Hence, for any ε > 0,
there is a δ > 0 such that
Now, divide the interval (0, ∞) into (0, δ) and (δ, ∞) to obtain
Z ∞ Z δ ! Z ∞
−nt −nt
f (t) e dt = O n |F (t)| e dt + O n |F (t)| e−nt dt
0 0 δ
=O ε[−e−nt ]δ0 + O [−e−nt ]∞
δ
−nδ
= O(ε) + O(e ),
Exercise 2.10 Deduce from (2.39) the integral representation for Euler’s
constant defined by (5.16):
Z ∞
γ=− f (t, 1) dt, (2.40)
0
where
e−t e−zt
f (t, z) = − . (2.41)
t 1 − e−t
Prove the Gauss’ integral representation for the digamma function ψ
defined by (5.18)
Z ∞ Z ∞ −t
e e−zt
ψ(z) = f (t, z) dt = − dt, Re z > 0. (2.42)
0 0 t 1 − e−t
R∞
Solution We notice that r1 = 0 e−rt dt, r ∈ N. Hence
n
X Z n
∞X Z ∞
1
−t r 1 − e−nt
= e dt = dt. (2.43)
r=1
r 0 r=1 0 1 − e−t
whence
Z ∞
1
F (z + 1) − F (z) = e−zt dt = (2.46)
0 z
and
+ F (z + n) − F (1 + n)
X
n−1
1 1
= − + F (z + n) − F (1 + n) (2.48)
1+k z+k
k=0
by (2.46). Hence
which is the same as the Gaussian representation for ψ (cf. (5.17)), whence
(2.42) follows.
and
Z ∞ Z ∞
π π 1 − u1−z 1 − v z−1
tan z = − du = dv, −1 < Re z < 1,
2 2 0 1 − u2 0 1 − v2
(2.57)
by the interchange of z by 1 − z.
Adding (2.56) and (2.57), we deduce (2.51), whence (2.52) follows by
1
changing z by − z. This completes the proof.
2
We are now in a position to prove the following remarkable result of
Gauss (cf. Theorems 8.1 and 8.2):
where
up−1 − uq−1
f (u) = q . (2.60)
uq − 1
1
Let ε = e2πi q be the first primitive q-th root of 1. Then the denominator
decomposes into
q−1
Y
(u − 1) u − εl ,
l=1
and the factor u − 1 cancels that of the denominator. Hence the partial
fraction expansion for f (u) is of the form
q−1
X Ak
f (u) = ,
u − εk
k=1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
and
Ak = lim (u − εk )f (u)
u→εk
k p−1
(ε ) − (εk )q−1 qε−k εpk − 1
=q =
Q
q−1
k l
Q
q−1 Q
q−1
(ε − ε ) εk (1 − εl−k )
l=0,l6=k l=0,l6=k l=0,l6=k
q εpk − 1
= = εpk − 1,
(εq )k q
Hence
q−1
X Z
p 1 1
ψ +γ = εpk − 1 du
q
k=1 0 u − εk
or
q−1
X
p 1 − εk
ψ +γ = εpk − 1 log . (2.63)
q −εk
k=1
Noting that
Q
q−1
q−1
X (1 − εk )
1 − εk k=1 q
log k
= log q−1 = log q q−1
= log q,
−ε Q k −ε 2
k=1 (−ε )
k=1
and
Xq−1
p p 2πp k
ψ +ψ 1− = 2i sin k −πi ,
q q q q
k=1
or
q−1
p p p 2π X 2πp
π cot π = ψ −ψ 1− = k sin k. (2.67)
q q q q q
k=1
Chapter 3
Abstract
3.1 Introduction
We shall consider the partial sum defined by (3.5) of the Hurwitz zeta-
function defined by (3.1) and prove the integral representation which turns
out to hold true for ζ itself. The proof as presented here is quite simple,
but the result is far-reaching and we may even base the whole theory of the
gamma and related functions on our results (Theorem 3.1 and its corollar-
ies). We shall develop this aspect of our theory in Chapter 5. The special
feature of Theorem 3.1 is that the derivatives may be computed by differen-
tiating with respect to u and the whole results may be inherited (for more
details, cf. the introductory remark at the beginning of §3.2).
In §3.3, we are going to give the sixth proof of the far-reaching formula
51
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
deduce the Poisson summation formula, we might regard our proof more
fundamental.
Thus, we put the existing literature on the Hurwitz zeta-function in
their hierarchical and historical perspective, with our recent contributions
[KKaY], [KKSY], [KTTY3] as touchstones.
We define the Hurwitz zeta-function and its special case, the Rie-
mann zeta-function, by Dirichlet series absolutely convergent for σ > 1
in the first instance. Both are meromorphically continued over the whole
plane with a simple pole at s = 1 as we shall see below.
∞
X 1
ζ(s, a) = , Re s = σ > 1, a>0 (3.1)
n=0
(n + a)s
X∞
1
ζ(s) = ζ(s, 1) = , σ>1 (3.2)
n=1
ns
X∞
e2πina
ls (a) = , σ > 1, a ∈ R (or s = 1, 0 < a < 1) (3.3)
n=1
ns
– the incomplete gamma functions of the 1st and the 2nd kind (cf. (3.53),
(3.66)), which satisfy γ(s, a) + Γ(s, a) = Γ(s);
Γ0 (s) d
ψ(s) = = log Γ(s)
Γ(s) ds
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
– the n-th Bernoulli polynomial with Bk the k-th Bernoulli number defined
through
∞
X Bk
z
z
= zk |z| < 2π
e −1 k!
k=0
– the n-th periodic Bernoulli polynomial, with [x] and {x} signifying the
integral and fractional parts of x, respectively.
We use the following as known:
denote the partial sum of the Hurwitz zeta-function, where for negative
values of u, the possible value of n for which n + a = 0 is to be excluded.
We shall use the Euler-Maclaurin sum formula (Theorem B.5, i.e. under
Appell’s (D0 )) to prove an integral representation for Lu (x, a), which has
∂k
the following far-reaching features shared by the derivatives ∂u k Lu (x, a) as
well, i.e. all statements about the function in u (Lu (x, a) and ζ(−u, a)) are
valid for their derivatives as well in the form of (i) below.
(i) It gives an analytic expression for Lu (x, a), which entails an integral
∂k
P u k
representation for each derivative ∂u k Lu (x, a) = 0≤n≤x (n+a) log (n+a)
(the differentiation of the integral being carried out under the integral sign).
(ii) It gives an asymptotic formula for Lu (x, a) in x by estimating the
integral trivially, which is feasible for applications in the divisor problems.
(iii) It gives a generic definition of ζ(−u, a) for u 6= −1 (and for γ0 (a) :=
−ψ(a) for u = −1).
(iv) It gives an integral representation for the associated Hurwitz zeta-
∂k (k)
function ζ(−u, a) (and its derivatives ∂u k ζ(−u, a) = ζ (−u, a)) for u 6=
−1, and for u = −1, it gives an analytic expression for the generalized
Euler constant γk (a) (the k-th Laurent coefficient of ζ(s, a) at s = 1),
which follows by simply putting x = 0 in the integral representation.
(v) The integral representation for ζ(s, a) (or γk (a)) in (iii) yields an
asymptotic formula for the ζ(s, a + z) in z with Bernoulli polynomial coef-
ficients (Theorem 2 [Kat1]) as given by Theorem 5.2 below.
Γ(u+1)
Convention. We sometime use (ur ) r! and Γ(u+1−r) interchangeably,
where the former is suited for easier calculation and the latter for expected
differentiation with respect to u.
l
X Γ(u + 1) (−1)r
Lu (x, a) = B r (x) (x + a)u−r+1
r=1
Γ(u + 2 − r) r!
Z ∞
(−1)l Γ(u + 1)
+ B l (t) (t + a)u−l dt (3.6)
l! Γ(u + 1 − l) x
1 (x + a)u+1 + ζ(−u, a), u 6= −1,
+ u+1
log(x + a) − ψ(a), u = −1.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Xl
(−1)r u
Lu (x, a) = B r (x)(x + a)u−r+1 + O xRe (u)−l
r=1
r r−1
(3.7)
1 (x + a)u+1 + ζ(−u, a), u 6= −1
+ u+1
log(x + a) − ψ(a), u = −1
holds true as x → ∞.
Furthermore, the integral representation
Xl
1 uu+1 (−1)r u
ζ(−u, a) = a − a − Br au−r+1
u+1 r r − 1
Z ∞
r=1 (3.8)
l+1 u
+ (−1) B l (t)(t + a)u−l dt,
l 0
which follows from (3.6) by putting x = 0, holds true for all complex u 6=
−1, where l can be any natural number subject only to the condition that
l > Re u + 1.
we see that the terms in the Euler-Maclaurin sum formula (Theorem B.5)
with a = 0 become
Z x Z x 1 (x + a)u+1 − 1 au+1 , u 6= −1
f (t) dt = (t + a)u dt = u + 1 u+1
0 0 log(x + a) − log(a), u = −1,
(−1)r n o
Xl
Br (x)f (r−1) (x) − Br (0)f (r−1) (0)
r=1
r!
l
X Γ(u + 1) (−1)r
= Br (x)(x + a)u−r+1 − Br au−r+1 ,
r=1
Γ(u + 2 − r) r!
and
Z x Z x
(−1)l+1 (l) (−1)l+1 Γ(u + 1)
Bl (t)f (t) dt = Bl (t)(t + a)u−l dt,
l! 0 l! Γ(u + 1 − l) 0
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
P
respectively. Hence writing Lu (x, a) = au + (n + a)u , we obtain
0<n≤x
1 1
(x + a)u+1 −
u au+1 , u 6= −1
Lu (x, a) = a + u + 1 u+1
log(x + a) − log a, u = −1
l
X Γ(u + 1) (−1)r
+ Br (x) (x + a)u−r+1
r=1
Γ(u + 2 − r) r! (3.9)
l
X r
Γ(u + 1) (−1)
− Br au−r+1
r=1
Γ(u + 2 − r) r!
Z x
(−1)l+1 Γ(u + 1)
+ Bl (t) (t + a)u−l dt.
l! Γ(u + 1 − l) 0
Z ∞
Bl (t) (t + a)u−l dt
x
∞ Z
Bl+1 (t) u−l ∞
= (t + a)u−l − Bl+1 (t)(t + a)u−l−1 dt
l+1 l + 1 x
Z ∞x
Re u−l
u−l−1
=O x +O (t + a) dt = O xRe u−l ,
x
Z x Z ∞ Z ∞
u−l u−l
Bl (t)(t + a) dt = Bl (t)(t + a) dt − Bl (t)(t + a)u−l dt
0 0 x
Z ∞
= Bl (t)(t + a)u−l dt + O xRe u−l .
0
R∞
Hence we may replace the integral in (3.9) by 0 Bl (t)(t + a)u−l dt +
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
O xRe u−l to obtain
1 (x + a)u+1 − 1 au+1 , u 6= −1
u
Lu (x, a) = a + u + 1 u+1
log(x + a) − log a, u = −1
l
X Γ(u + 1) (−1)r
+ Br (x) (x + a)u−r+1
r=1
Γ(u + 2 − r) r!
l
X Γ(u + 1) (−1)r
− Br au−r+1
r=1
Γ(u + 2 − r) r!
Z ∞
(−1)l+1 Γ(u + 1)
+ Bl (t) (t + a)u−l dt + O xRe u−l .
l! Γ(u + 1 − l) 0
(3.10)
X l
1 Γ(u + 1) (−1)r
ζ(−u, a) = au − au+1 − Br au−r+1
u+1 r=1
Γ(u + 2 − r) r!
Z ∞ (3.11)
l
(−1) Γ(u + 1)
+ Bl (t) (t + a)u−l dt,
l! Γ(u + 1 − l) x
which is (3.8).
Now for any u 6= −1, take l ∈ N such that l > Re u + 1. Then the last
integral in (3.11) is absolutely convergent for Re u < l − 1 and represents an
analytic function in Re u < l − 1. Substituting (3.11) in (3.10), we deduce
(3.6) for u 6= −1.
In the case u = −1, the Euler-Maclaurin sum formula on the same lines
as above (cf. the proof of (5.39)) gives rise to
Z ∞
−1
L−1 (x, a) = (−1)l+1 B l (t) (t + a)−1−l dt
l 0
X l
−1 1
− (−1)r Br a−r
r=1
r − 1 r
X l (3.12)
−1 1
+ (−1)r B r (x) (x + a)−r
r=1
r−1 r
Z ∞
l −1
+ (−1) B l (t) (t + a)−1−l dt
l x
+ a−1 + log(x + a) − log a.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Substituting the constant term −ψ(a) in (3.12), we have the integral rep-
resentation for the partial sum.
Since the integrals appearing in Theorem 3.1 are analytic in the region
Re u < 1 − l, we may differentiate (3.6) and (3.8) in u there. We state the
counterpart of (3.7) as the following corollaries (the counterpart of (3.8) to
be read off from them by putting x = 0).
Corollary 3.1 For any complex u and a > 0,
d X
Lu (x, a) = (n + a)u log(n + a) (3.15)
du
0≤n≤x
l
X (−1)r
= B r (x)(x + a)u−r+1
r=1
r!
Γ(u + 1)
× {ψ(u + 1) − ψ(u + 2 − r) + log(x + a)}
Γ(u + 2 − r)
Z
(−1)l ∞
+ B l (t)(t + a)u−l
l! x
Γ(u + 1)
× {ψ(u + 1) − ψ(u + 1 − l) + log(t + a)} dt
Γ(u + 1 − l)
1 1
u+1 u+1
(x + a) log (x + a) − 2 (x + a)
u+1 (u + 1)
+ −ζ 0 (−u, a) , u 6= −1
1 {log (x + a)} 2 + γ (a) ,
1 u = −1.
2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
We note that Theorem 3.1 [(3.6), (3.7)] is first obtained by Mellin [Me]
by means of the integral transform under his name (§7.4) and is the most
∂
informative for Lu (x, a), so are Corollaries 3.1 and 3.2 for ∂u Lu (x, a) and
∂2
∂u2 Lu (x, a), respectively. Formula (3.6) with l = 1, u 6= −1, Re u < 0
appears as a prototype of the “approximate functional equation” in Landau
[Lan, 9–10]. Mikolaś [M1] used it with x = 1 to obtain (3.8) with l = 1.
Then he proceeds to deduce (3.8) with l = 2 by integration by parts.
Γ(u+1)
Care should be taken in interpreting the coefficients like Γ(u+1−l) (ψ(u +
1) − ψ(u + 1 − l)) when u is a negative integer; it is to be taken as one
without singularities (e.g. in deducing (5.20)).
Corollary 3.3 The k-th Laurent coefficient of the Hurwitz zeta-function
k
(at s = 1) is given by (−1)
k! γk (a), where
X logk (n + a) logk+1 (x + a)
γk (a) = lim − (3.17)
x→∞ n+a k+1
0≤n≤x
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Proof. The following is the simplest possible method known. The start-
ing point is Theorem 3.1 with l = 1 and −s (s 6= 1, σ > 0) for u:
Z ∞
(x + a)1−s B 1 (x) B 1 (t)
L−s (x, a) = + ζ(s, a) − +s dt. (3.19)
1−s (x + a)s x (t + a)s+1
Since both sides of (3.19) are analytic in σ > 0, we may compute the k-th
Taylor coefficient around s = 1. The k-th Taylor coefficient of the left-hand
side is
1 ∂k (−1)k X
k
L−s (x, a)|s=1 = (n + a)−1 logk (n + a) (3.20)
k! ∂s k!
0≤n≤x
logk+1 (x + a) B 1 (x)
− + logk (x + a) (3.22)
k+1 x+a
Z ∞
B 1 (t) k k−1
− log (t + a) − k log (t + a) dt.
x (t + a)2
We now note that (3.22), being valid for any x ≥ 0, implies both (3.17)
and (3.18) by letting x → ∞ and x = 0 respectively, (cf. Berndt [Ber3]).
In the case k = 0, we note that (3.13) and (3.14) correspond to (3.17) and
(3.18), respectively.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
In this section we are going to give the sixth proof of the fundamental
summation formula based on the use of finite differences, which has been
applied successfully in recent researches [KKY3], [KKY2].
X∞ λ
X
ζ(m, α) m+λ λ 0
z = ζ (−k, α − z) z λ−k
m=2
m + λ k
k=0
λ
X 1 (3.23)
− ζ 0 (−λ, α) − ζ(k − λ, α) z k
k
k=1
1
+ (ψ(α) − Hλ ) z λ+1 .
λ+1
Proof. Let ∆α f (α) = f (α + 1) − f (α) be the difference operator (intro-
duced in (DE) in Chapter 1). We apply this to the sum S on the LHS of
(3.23) to obtain
1 z m
X∞ X∞ ∞
X
ζ(m, α) m+λ α−m m+λ
∆α S = ∆ α z =− z = −αλ .
m=2
m+λ m=2
m+λ m α
m=λ+2
whence
λ+1
X αλ−m m
∆α S = αλ log(α − z) − αλ log α + z . (3.24)
m=1
m
Rewriting
the first term on the RHS of (3.24) in the form
Pλ λ λ−k
k=0 k z (α − z)k log(α − z) and telescoping (3.24), thereby noting
that
we get
λ
X λ
S= ζ 0 (−k, α − z) z λ−k − ζ 0 (−λ, α)
k
k=0
(3.25)
λ
X 1 ψ(α) λ+1
− ζ(k − λ, α) z k + z + f (z, α),
k λ+1
k=1
∆α f (z, α) = 0 (3.26)
and
f (0, α) = 0. (3.27)
Hλ λ+1
It remains to determine f (z, α) (to be − λ+1 z ). First note that
d 0 ∂ ∂
ζ (−k, α − z) = ζ(s, α − z) |s=−k
dz ∂s ∂z
∂
= s ζ(s + 1, α − z) |s=−k
∂s
(
ζ(1 − k, α − z) − k ζ 0 (1 − k, α − z), k ∈ N
=
−ψ(α − z), k = 0.
∂ X λ
λ−1
S= (λ − k) ζ 0 (−k, α − z) z λ−k−1 (3.28)
∂z k
k=0
Xλ Xλ
λ λ 0
+ ζ(1 − k, α − z) z λ−k − ζ (1 − k, α − z) z λ−k
k k
k=1 k=1
λ
X ∂
− ψ(α − z) z λ − ζ(k − λ, α) z k−1 + ψ(α) z −λ + f (z, α).
∂z
k=1
We note that the two sums on the RHS of (3.28) containing ζ 0 cancel
each other, while the second sum, say S2 , becomes, in view of the addition
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
formula,
λ
X λ 1
S2 = − Bk (α − z)z λ−k (3.29)
k k
k=1
Xλ λ−k X k λ λ−k
X
λ z k λ z
=− Bl (α) (−z)k−l − (−z)k
k k l k k
k=1 l=1 k=1
= S2,1 + S2,2 ,
say. Using (1.14) and changing the order of summation in S2,1 , we have
λ
X λ−l
X
λ λ − l (−1)k
S2,1 = − Bl (α) z λ−l .
l k k+l
l=1 k=0
we deduce that
λ
X λ
X λ
X
Bl (α)
S2,1 = − z λ−l = ζ(1−l, α) z λ−l = ζ(l−λ, α) z l−1 . (3.30)
l
l=1 l=1 l=1
to obtain
S2,2 = Hλ z λ . (3.31)
is first stated by Wilton [Wil1, Eq.(4)] and is a rather special case of Theo-
rem 3.2. As an asymptotic formula in a, this gives the Stirling formula and
is a special case of Corollary 5.1.
We shall give two proofs of [EM2], Theorem 4.3, which seems the highest
summit of the paper, and coincides with our Corollary 3 (i) [KKY3]; the
first proof depends on a modified form of Theorem 3.2, which we state
as Lemma 3.1 while the second depends on a more antecedent one, i.e.
the intermediate formula toward the proof of Proposition 1 [KKaY, p.10],
which we state as Lemma 3.2.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Z z
tλ ψ(t + a) dt
0
λ
X
λ Bk+1 (a + z)
= (−1)k z λ−k ζ 0 (−k, a + z) − Hk (3.36)
k k+1
k=0
Bλ+1 (a + z)
− (−1)λ ζ 0 (−λ, a) − Hλ .
k+1
Z z
tλ ψ(α + t) dt
0
λ
X
= (−1)λ Cλ (r, α) log Γr+1 (α + z)/Γr+1 (α)
r=0
λ
X
λ l λ 0 Bλ−l+1 (α) z λ+1
+ (−1) (−1) ζ (l − λ) + zl + Hλ .
l l(λ − l + 1) λ+1
l=1
(ii) For λ ∈ N
Z z
λ tλ−1 log Γ(α + t) dt
0
λ
X
= z λ log Γ(α + z) − (−1)λ Cλ (r, α) log Γr+1 (α + z)/Γr+1 (α)
r=0
λ
X
λ Bλ−l+1 (α) z λ+1
− (−1)λ (−1)l ζ 0 (l − λ) + zl − Hλ ,
l l(λ − l + 1) λ+1
l=1
X∞
ζ(m, α) m+λ
z
m=2
m+λ
Xλ
λ (3.37)
= {ζ 0 (−k, α − z) + Hk ζ(−k, α − z)} z λ−k
k
k=0
ψ(α) λ+1
− (ζ 0 (−λ, α) + Hλ ζ(−λ, α)) + z .
λ+1
Proof. (First proof of Theorem 3.3) We start from the Taylor expansion
(|z| < α)
X∞ X∞
ψ (n) (α) n
ψ(z + α) = z = ψ(α) + (−1)m ζ(m, α) z m−1 . (3.38)
n=0
n! m=2
Multiplying both sides of (3.38) by z λ and integrating over [0, z] with re-
spect to z, we deduce that
Z z
uλ ψ(α + u) du (3.39)
0
Z z ∞
X Z z
= uλ ψ(α) du + (−1)m ζ(m, α) uλ+m−1 du
0 m=2 0
X∞
ζ(m, α) z λ+1
= (−1)λ (−z)m+λ + ψ(α).
m=2
m+λ λ+1
which is (3.36).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
λ
1 1 λ+1
Rewriting k k+1 as λ+1 k+1 and writing k for k+1, we derive from (3.44)
that
λ+1
1 X λ+1 Bλ+1 (a)
F (1) = − Bk (a + z) (−z)λ+1−k + , (3.45)
λ+1 k λ+1
k=0
where we incorporated the last term in (3.39) in the first sum of (3.45).
Noting that the fist sum of (3.45) is nothing but the expansion of the
Bernoulli polynomial Bλ+1 (a + z − z) = Bλ+1 (a), we conclude F (1) = 0.
Hence, weR z may take the limit as s → 1 of (3.37). On the left side we
have (−1)λ 0 uλ ψ(a + u) du by the Laurent expansion of ζ(s, a + u), and
on the right-side we just differentiate F (s) with respect to s, thereby noting
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
the formula
0
Γ(2 − s)
Γ(k + 1 − s)
s=1 (3.46)
Γ(2 − s) 1
= (−ψ(2 − s) + ψ(k + 2 − s)) = Hk ,
Γ(k + 2 − s) s=1 k!
to obtain
λ
X λ
F 0 (1) = {ζ 0 (−k, a + z) + Hk ζ(−k, a + z)} (−z)λ−k
k (3.47)
k=0
and our Theorem 3.3 coincides with Theorem 4.3 of Espinosa and Moll.
Remark 3.2 (i) Espinosa and Moll [EM1] developed the Hurwitz trans-
form
Z 1
f (u)ζ(s, u) du
0
and deduced several results for special types of f (u) which can be expanded
into Fourier series as consequences of their Theorem 2.2, which in turn is
a consequence of the “Fourier series”:
∞
X πs
ζ(s, u) = 2 Γ(1 − s) (2πn)s−1 sin 2πnu + , (3.48)
n=1
2
or, more commonly known as the Hurwitz formula (cf. (5.56)). We note
that Mikolás’ [M3] gave the simplest proof of (3.48) as the Fourier series,
whereby he computed the Fourier coefficients
Z 1
Γ(1 − s)
ζ(s, u) e−2πiνu du = 1−s
, (3.49)
0 (2πiν)
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
for Re (1 − s) < 12 . We note that Mikolás [M1] obtained the result on the
basis of Fourier analysis (the Parseval formula):
Z 1
ζ(s, u) ζ(s0 , u) du
0
0
π
= 2 (2π)s+s −2 Γ(1 − s) Γ(1 − s0 ) cos (s − s0 ) ζ(2 − s − s0 )
2
for max{0, Re s} + max{0, Re s0 } < 1; the region of validity wider than that
of Espinosa and Moll who have s < 0, s0 < 0.
In [KKSY] statements were made about the proof of the functional equa-
tion, or the Hurwitz formula (3.48), for the Hurwitz zeta-function, using the
absolutely convergent Fourier series for B 2 (t) rather than the boundedly
convergent Fourier series for B 1 (t). Meanwhile the book of Laurinčikas and
Garunkštis [LG] has appeared which has rich contents about rather wide
P∞ e2πinξ
spectrum of the theory of the Lerch zeta-function φ(ξ, a, s) = n=0 (n+a) s
1 1 1
ζ(−u, a) = − au+1 + au − u au−1
u+1 2 12
Z (3.50)
u(u − 1) ∞
− B 2 (t)(t + a)u−2 dt,
2! 0
∞ ∞
1 X e2πint + e−2πint 1 X cos(2πnt)
B 2 (t) = = 2 (3.51)
2π 2 n=1 n2 π n=1 n2
in the integral in (3.8) in the same context as in Rademacher [R, p.83], Pan
and Pan [PP, p.125], Kanemitsu [Kan], and Ueno and Nishizawa [UN], we
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
d2 w dw
z + (c − z) − aw = 0 (3.56)
dz 2 dz
and is denoted by U (a, c; z) in Slater [Sla].
Using [Erd, p.257, Eq. 6.5(6)]
in (3.54), we get
ζ(−u, a)
1 1 1
=− au+1 + au − u au−1
u+1 2 12 (3.60)
∞
u(u − 1) au−1 X 1
− {Ψ (1, u; −2πina) + Ψ(1, u; 2πina)} .
2 2π 2 n=1 n2
∞ ∞
1 X e2πint − e−2πint 1 X sin (2πnt)
B 1 (t) = − =− , (3.62)
2πi n=1 n π n=1 n
Z ∞
B 1 (t)(t + a)u−1 dt
0
∞
1 X au −2πina (3.63)
=− e (−2πina)−u Γ(u, −2πina)
2πi n=1 n
− e2πina (2πina)−u Γ(u, 2πina) ,
and
1 1
ζ(−u, a) = − au+1 + au
u+1 2
u au X 1 (3.64)
− Ψ(1, u + 1; −2πina),
2πi n
n6=0
We invert the order of summation and integration in the last term and
P0 ∞ −2πina(u−1)
consider the series n=−∞ e as the Fourier series for δ(a(u −
1)) − 1. Then we are left with the integration (σ < 0)
Z 1 Z 1
−s 1 1
− δ(a(u − 1)) u du + u−s du = − − .
0 0 2a s − 1
1
Hence the last term is − 2a1s − s−1 1
as−1 , which cancels the second term and
we finally arrive at the Hurwitz formula
Γ(1 − s) 1−s πi − 1−s
ζ(s, a) = e 2 l (1 − a) + e 2 πi l (a) , (3.67)
1−s 1−s
(2π)1−s
which is equivalent to (3.48).
Finally, we introduce a class of functions γn (x), n ∈ N (due to Mil-
nor [Mi]) defined by
∂
γ1−t (x) = ζ(t, x) (3.68)
∂t
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Note that
Γ(x)
γ1 (x) = ζ 0 (0, x) = log √ .
2π
Exercise 3.1 Prove the Kubert identity
q−1
X
s−1 x+k
ζ(1 − s, x) = q ζ 1 − s, (3.69)
q
k=0
Solution By (8.12),
Φ(s, a, 1) = ζ(s, a). (3.71)
Hence (8.13) reduces to (3.69).
To prove (3.70), we differentiate
q−1
X
x+k
q s ζ(s, x) = ζ s, , (3.72)
q
k=0
Chapter 4
Abstract
1
ζ(1 − n, x) = − B n (x), n∈N (4.1)
n
from (3.8).
n
1 n X (−1)r n − 1
ζ(1 − n, a) = an−1 − a − Br an−r .
n r=1
r r − l
1
Br = (−1)r Br , r ≥ 2, B1 = − ,
2
77
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
into
n
1 n 1 n−1 1 X n
ζ(1 − n, a) = − a + a − Br an−r
n 2 n r=2 r
n
1X n
=− Br an−r ,
n r=0 r
Comparing these completes the proof of (1.9) in the case 0 < x < 1.
If, in general, n − 1 < x < n, n ∈ Z, then [x] = n and [1 − x] = −n, and
therefore
[1 − x] = −[x], / Z.
x∈ (4.6)
By (A),
n
X n
Bn (x) = Bn (x + [x]) = B n−k (x)[x]k .
k
k=0
/ Z.
Substituting (1.9) for 0 < x < 1 and (4.6). we conclude (1.9) for x ∈
For x ∈ Z, (1.9) follows from continuity.
Proposition 4.2 The functional equation (3.67) for the Hurwitz zeta-
function implies the Fourier expansion (H) for the Bernoulli polynomials.
which gives (H) for n ≥ 2, since, then, the series are absolutely convergent.
In the case n = 1, the sum is to be taken symmetrically:
N
X 2πikx N
X
e e−2πikx sin 2πkx
lim − = lim 2i ,
N →∞ k k N →∞ k
k=−N k=1
k6=0
whence
∞
1 X sin 2πkx
B 1 (x) = −
π k
k=1
as in (7.9).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Proposition 4.3 The Kubert identity (3.69) for the Hurwitz zeta-
function implies the Kubert identity (1.8) for Bernoulli polynomials.
Proof. This follows from (3.69) on substituting (4.1).
Now that we have established (DE), (H) and (K) for Bernoulli poly-
nomials, we may trace the logical path given at the end of Chapter 1 to
complete the theory of Bernoulli polynomials.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Chapter 5
Abstract
N
!
X 1
ψ(s) = − lim − log(N + s)
N →∞
n=0
n+s
81
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
γ = −ψ(1)
B k (t) = Bk (t − [t])
—the k-th Bernoulli polynomial (cf. (1.7)); Bk -the k-th Bernoulli number;
[t] —the integral part of t.
Theorem 5.1 If we suppose the integral representations for ζ(s, z)
and ψ(z):
Z ∞
1 1
ζ(s, z) = z 1−s + z −s − s B 1 (t) (t + z)−s−1 dt, σ > −1, (5.1)
s−1 2 0
Z ∞
1
ψ(z) = log z − z −1 + B 1 (t) (t + z)−2 dt, (5.2)
2 0
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Hence
Z ∞ √
1+ B 1 (t) t−1 dt = log 2π, (5.11)
1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
ζ 0 (s, x)
Z ∞
−1 x1−s 1
= 2
x1−s − log x − x−s log x − B 1 (t) (t + x)−s−1 dt
(s − 1) s−1 2 0
Z ∞
−s B 1 (t) (t + x)−s−1 (− log(t + x)) dt,
0
whence
Z ∞
1
ζ 0 (0, x) = −x + x log x − log x − B 1 (t) (t + x)−1 dt. (5.13)
2 0
Now
Z ∞
∂2 0 1 1 1
ζ (0, x) = + −2 B 1 (t) (t + x)−3 dt. (5.14)
∂x2 x 2 x2 0
Now the last integral on the right-hand side of (5.14) is the sum of the
terms
Z n+1
t − n − 21 (t + x)−3 dt
n
Z n+1
= (t + x) − n + x + 1
2 (t + x)−3 dt
n
Z n+1
= (t + x)−2 − n + x + 1
2 (t + x)−3 dt
n
n+1
1 1
= − + n + x + 12 (t + x)−2
t+x 2 n
1 1
= −
n+x n+x+1
1 1
+ n + 1 + x − 21 (n + 1 + x)−2 − n + x + 12 (n + x)−2
2 2
1 1
= −
n+x n+x+1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1 1 1 1 1 1
+ − − + .
2 n+1+x n+x 4 (n + 1 + x)2 (n + x)2
Hence
Z ∞
B 1 (t)(t + x)−3 dt
0
∞ ∞
1X 1 1 1X 1 1 1
= − − +
2 n=0 n + x n + x + 1 2 n=0 (n + x)2 4 x2
∞
11 1 1 1X 1
= + 2
− .
2x 4x 2 n=0 (n + x)2
First
ζ 0 (0, 1) = a + b
and
ζ 0 (0, 2) = 2a + b.
Recalling ζ(s, x + 1) = ζ(s, x) − x−s , we see that ζ 0 (s, x + 1) = ζ 0 (s, x) +
x log x, whence ζ 0 (0, 2) = ζ 0 (0, 1). Hence a = 0.
−s
for any z other than negative integers, the two definitions for ψ are equiv-
alent:
X∞
1 1
ψ(z) + γ = − , (5.17)
n=1
n z+n−1
we deduce that
N
X
1
ψ(z) + γ = γ − lim − log(N + z) ,
N →∞
n=0
z+n
whence (5.18).
On the other hand, (5.18) may be written as
N
X N
X !
1 1 1
ψ(z) = lim − − − log(N + z) ,
N →∞
n=1
n z + n −1 n=1
n
X∞
1 1
= − − γ,
n=1
n z+n−1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
i.e. (5.17).
Remark 5.1 In (5.16), z is usually taken to be 0, but can be any number
as in (5.16) because
log(N + z) − log(N + w) → 0, N → ∞.
The absolute
convergence of the series in (5.17) is clear because each term
is O n12 , and the existence of limits in (5.16) and (5.18) follows from the
comparison with the corresponding integral, or the Euler-Maclaurin formula
(cf. Chapter 2).
We shall illustrate the far-reaching power of Theorem 3.1 by the first
∂
derivative ( ∂u Lu (x, a) or −ζ 0 (−u, a)) in the special case of u = m, m ∈
N ∪ {0}. For N 3 l > m + 1, Corollary 3.1 eventually yields (cf. [KTTY3])
N
X
−ζ 0 (−m, a) = lim (n + a)m log(n + a) (5.19)
N →∞
n=0
1 1
− (N + a)m+1 log(N + a) + (N + a)m+1
m+1 (m + 1)2
1 X m Br
m+1
− (N + a)m log(N + a) −
2 r=2
r − 1 r!
!
1 1 m−r+1
· +··· + + log(N + a) (N + a) .
m m−r+2
and
ζ 0 (−m, a) (5.20)
1 1 1 1
= am+1 log a − am+1 − am log a + am−1 log a
m+1 (m + 1)2 2 12
m+1
X Br X r−2
+ (−1)j m 1
+
m
log a
r=4
r j=0
j r − 1 − j r −1
X l r−1
X
1 r−m−2 1 m−r+1
+ Br (−1)j a
m + 1 r=m+2 j=0
j r − j
Z ∞ X l−1
+ (−1)l+1 (−1)j l − m − 1 1
B l (t)(t + a)m−l dt,
0 j=0
j l − j
Exercise 5.2 Show that, in view of Lerch’s formula (5.4), (5.19) with
m = 0 gives Euler’s product formula (or Weierstrass’ canonical product of
genus 1) (5.21) for Γ(a).
∞
1 Y 1 −a a
=a 1+ 1+ , (5.21)
Γ(a) n=1
n n
N
X 1
−ζ 0 (0, a) = lim log(n + a) − log(N + a)
N →∞
n=0
2
! (5.22)
− (N + a) log(N + a) + N + a .
N !
X 1
0
−ζ (0) = lim log n − N + log N + N . (5.23)
N →∞
n=1
2
Substituting
1 1
log(N + a) = log N + o(1),
2 2
(N + a) log(N + a) = N log N + a log N + a + o(1),
Γ(a)
− log √
2π
N ! (5.24)
X 1
= lim log(n + a) − N + log N + N − a log N .
N →∞
n=0
2
N
!
X n+a
− log Γ(a) = lim log a + log − a log N . (5.25)
N →∞
n=1
n
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
PN −1
By expressing log N as n=1 log n+1
n , we may write
!
n + a n + 1 −a
N
X
− log Γ(a) = lim log a + log · (5.26)
N →∞
n=1
n n
1 −a a
X∞
= log a + log 1+ 1+ ,
n=1
n n
whence (5.21).
N ! (N + 1)a
Γ(a) = lim , (5.27)
N →∞ a(a + 1) · · · (a + N )
N N
!
X X
log Γ(a) = lim − log(n + a) + log n + a log(N + 1) (5.27)0
N →∞
n=0 n=1
and deduces Lerch’s formula by comparing (5.27)0 with (5.22) (and (5.23)).
Of course, we can cover (5.27)0 in the same way as above. Indeed, from
(5.23) and (5.24), we deduce that
N N
!
X X
0 0
−ζ (0, a) + ζ (0) = lim log(n + a) − log n − a log N ,
N →∞
n=0 n=1
(N + a){log2 (N + a) − 2 log(N + a) + 2}
1
= N log2 N − 2N log N + 2N + a log2 N + O .
N
Supposing further that |α| < |z|, we infer, by the binomial expansion,
0
that the right-hand side of (3.8) can be written as
S + O |z|Re (u)−l ,
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
where
l l
1 XX u+1 u+1−r
S := − Br αk−r z u−k+1 ,
u + 1 r=0 r k−r
k=r
Theorem 5.2 For any integer l ≥ 0 and any z in | arg (z) | < π,
l−1
X (−1)r+1
1
ζ(s, α + z) = z 1−s + Br+1 (α) (s)r z −s−r
s−1 r=0
(r + 1)! (5.35)
−Re (s)−l
+ O |z| .
Corollary 5.1 For any integer l and any fixed α > 0, the generalized
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Stirling’s formula:
Γ(z + α) 1
log √ = z+α− log z − z
2π 2
l−1
X (−1)r+1
+ Br+1 (α) z −r + O |z|−l
r=1
r(r + 1)
N !
a 1 X 1 1
ψ(a) = lim −z + − − − . (5.39)
N →∞ N a k+a k
k=1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Proof. Only (5.39) needs a proof, which goes on the similar lines as those
for the case u = −1 of Theorem 3.1. Formula (5.38) gives
X
1 1
ψ(a) = log(x + a) − +O ,
n+a x
0≤n≤x
f (x + 1) − f (x) = g(x), x ∈ R+ .
Proof. For curiosity, we prove (i) without assuming (5.39). With g(x) =
1
x , (5.40) is satisfied. (a) follows from the definition and (c) follows from
(5.18); only (b) remains.
We differentiate (5.2) to obtain
Z ∞
1 1
ψ 00 (a) = − 3 − 2 + 6 B 1 (t) (t + a)−4 dt. (5.43)
a a 0
To express the last integral inR closed form is an easy exercise. Indeed,
n+1
it is the sum of integrals of type n . Since
Z n+1 Z n+1
−4 1
B 1 (t)(t + a) dt = t+a− a+n+ (t + a)−4 dt
n n 2
Z n+1 Z n+1
1
= (t + a)−3 dt − a + n + (t + a)−4 dt
n 2 n
1
= − (n + 1 + a)−2 − (n + a)−2
2
1 1
+ n+a+ (n + 1 + a)−3 − (n + a)−3
3 2
1 1 1
= − (n + 1 + a)−2 − (n + a)−2 + (n + a + 1)−2 − (n + 1 + a)−3
2 3 2
1 1
− (n + a)−2 + (n + a)−3
3 2
1 1
= − (n + 1 + a)−2 − (n + a)−2 − (n + a + 1)−3 + (n + a)−3 ,
6 6
summing these for n = 0, 1, 2, 3, 4, · · · , we obtain
Z ∞
∞
1 −2 1 −3 1 X 1
B 1 (t) (t + a)−4 dt = a − a − .
0 6 6 6 n=0 (n + 1 + a)3
Hence
2
ψ 00 (a) = − − ζ(3, a + 1) < 0, (5.44)
a3
and (b) follows, whence uniqueness follows from Lemma 5.2.
Γ(z)
(ii) log √ 2π
is the unique solution (convex for large argument) of the DE
∞
X ∞
X ∞ ∞ k
1 1 XX z
z k−1 =
n=0
(n + a)k z n=0 n+a
k=2 k=2
2
∞ z ∞
1 X n+a X 1
= z =
z n=0 1 − n+a n=0
(n + a − z)(n + a)
X 1 1
= lim − log(x + a − z) + log(x + a) − ,
x→∞ n+a−z n+a
0≤n≤x
1 1 1
ψ(2z) = ψ(z) + ψ z + + log 2, (5.47)
2 2 2
and, a fortiori,
2z−1 − 21 1
Γ(2z) = 2 π Γ(z) Γ z + . (5.48)
2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1
√
which is ψ(z) − log 2. (5.48) follow from (5.47) if we use Γ 2 = π.
Remark 5.4 The property in Corollary 5.3 is a special case of the Ku-
bert identity (or distribution property) shared by a wide class of functions
(cf. (8.13), [Mi], [Su1]).
is a consequence of (5.45).
which is also a well-known formula (cf. e.g. [SC]). We need only the special
case of (5.50) with a = 1, z replaced by −z:
∞
X (−z)k
ζ(k) = log Γ(z + 1) + γ z. (5.51)
k
k=2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1 z k
X∞ X∞
− ,
n=1
k n
k=2
the inner sum of which can be summed by the elementary formula (|r| < 1)
∞
X 1 k
r = −r − log(1 − r).
k
k=2
Hence
∞ z
∞
X X
(−z)k z
ζ(k) = − log 1 + . (5.52)
k n=1
n n
k=2
Lemma 5.3 The asymmetric form (5.54) of the functional equation for
the Riemann zeta-function is a consequence of the functional equation for
the Hurwitz zeta-function (or the Hurwitz formula) (0 < a < 1)
π π
ζ(s, a) = −i (2π)s−1 Γ(1 − s) e 2 is l1−s (a) − e− 2 is l1−s (1 − a) , (5.56)
X∞
e2πina
where ls (a) = stands for the polylogarithm function (3.3), For-
n=1
ns
mula (5.56) (which already appeared as (3.48) and (3.67)) in the long run,
is a consequence of (5.1).
A recent proof of (5.56) based on the Fourier expansion of the Dirac
delta function can be found in [BKT] or [KTTY3] and is sketched in §3.5.
A more laborious but easier proof can be found in [R] (for the Riemann
zeta) and [PP] (for the general case). It amounts to completing the integral
in
1 1 1
ζ(s, a) = a1−s + a−s + s a−s−1
s−1 2 12
Z (5.57)
s(s + 1) ∞
− B 2 (t) (t + a)−s−2 dt, σ > −2,
2 0
R∞
in the from −a B 2 (t)(t + a)−s−2 dt, then using the absolutely converging
Fourier series for B 2 (t) and appealing to a formula for the Mellin transform.
We refer to the above references.
Proposition 5.3 The product representation for the sine function
∞
Y
sin πz z2
= 1− 2 , (5.58)
πz n=1
n
Proposition 5.4 The partial fraction expansion for the cotangent func-
tion
∞
1 2z X 1
cot πz = −
πz π n=1 n2 − z 2
∞
1 1X 1 1
= + − , (5.61)
πz π n=1 n + z n−z
is a consequence of (5.58).
Proof. This follows immediately by logarithmic differentiation.
Remark 5.6 Comparing (5.17) and (5.61), we cover formula (2.55)
again.
Proposition 5.5 The partial fraction expansion for the hyperbolic cotan-
gent function and (5.61) are equivalent:
∞
1 1 1 1 xX 1
coth πx = 2πx + = + , Re x ≥ 0. (5.62)
2 e −1 2 2πx π n=1 n2 + x2
The theory of the gamma and related functions via zeta-functions 101
Proposition 5.6 The functional equation (5.56) for the Hurwitz zeta-
function implies the partial fraction expansion (5.61) for the cot-function.
Proof. We remark that the functional equation (5.56) for ζ(s, x) may be
expressed on the basis of (5.53) as (5.56)0 .
First we assume that Im x > 0. Then the sum for l0 (x) converges for
every s ∈ C, and the left-hand side is
∞
X e2πix 1
l0 (x) = e2πinx = 2πix
= (−1 + i cot πx) . (5.64)
n=1
1−e 2
we get
(5.17) =⇒ (5.49) ⇒
(5.58) ⇔ (5.61) ⇔ (5.62) ⇔ (5.63) ⇔ (5.56)
(5.56) ⇒ (5.54) ⇒ (5.53) ⇒
m m
⇐
(5.65) ⇐ (5.56) (5.54)
Lemma 5.6 (Berndt) The functional equation (5.56) for the Hurwitz
zeta function implies Kummer’s Fourier series for log Γ(x), which reads
Γ(x) 1 1
log √ = − log(2 sin πx) + (γ + log 2π)(1 − 2x)
2π 2 2
X∞ (5.65)
1 log n
+ sin 2πnx,
π n=1 n
The theory of the gamma and related functions via zeta-functions 103
X∞
2πx 2 1
+ πx = 1 + 2x
e2πx − 1 n=1
n 2 + x2
1
= 1 − z + 2z 2 ϕ z 2
2
say, where
∞
X −1
ϕ(w) = w + 4π 2 n2 . (5.68)
n=1
Since
∞
X −r−1
ϕ(r) (w) = (−1)r r! w + 4π 2 n2 ,
r=0
we see that
X∞
ϕ(r) (0) 1 (−1)r
= (−1)r 2r+2
= ζ(2r + 2).
r! n=1
(2πn) (2π)2r+2
Hence
X∞
z 1 (−1)r
= 1 − z + 2 ζ(2r + 2) z 2r+2 ,
ez − 1 2 r=0
(2π) 2r+2
and so
X∞
z 1 2 (−1)m−1
= 1 − z + ζ(2m) z 2m .
ez − 1 2 m=1
(2π) 2m
we conclude (5.66).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Remark 5.8 As is proved above, (5.62) and the functional equation for
the Riemann zeta-function are equivalent, whence we see that Euler’s iden-
2
tity (5.66), and in particular the solution to the Basel problem ζ(2) = π6 ,
is a consequence of the functional equation.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Chapter 6
Abstract
105
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Definition 6.1 The n-th Bessel function Jn (z) is defined as the n-th
z 1
Laurent coefficient of the function exp 2 (w − w ) in w, viz.
X∞
z 1
exp w− = Jn (z) wn . (6.1)
2 w n=−∞
on noting that w − w1 = 2i sin θ. Dividing the interval [0, 2π] into [0, π] and
[π, 2π] and making the change of variable in the integral over [π, 2π], we
obtain
Z π
1
Jn (z) = e−i(z sin θ−nθ) dθ.
2π 0
which is a Fourier series converging to the left-hand side member (in view of
Theorem 7.2) and the Fourier coefficient Jn (z) may be computed by (7.2):
Z π
1
Jn (z) = ei(z sin θ−nθ) dθ,
2π −π
Z π
1
sin(z sin θ) sin(nθ) dθ = 0
π 0
Z π
1
sin(z sin θ) sin(nθ) dθ = Jn (z) (6.7)
π 0
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Solution Jn (z) + J−n (z) is 2Jn (z) for n even and 0 for n odd, and by
(6.2), this is
Z
1 π
(cos(z sin θ − nθ) + cos(z sin θ + nθ)) dθ,
π 0
whence the first and the second identities follow.
Considering Jn (z) − J−n (z), we deduce the third and the fourth identi-
ties.
/ Z, the Bessel functions Jν (x) and J−ν (x) are two indepen-
In case ν ∈
dent solutions to the Bessel differential equation
d2 y 1 dy n2
+ + 1 − 2 y = 0. (6.8)
dx2 x dx x
For n ∈ Z, the fundamental solutions to (6.8) are given by Jn (x) and
Yn (x), the Weber function, relevant to analytic number theory. Jν (z) and
Yν (z) are often referred to as the Bessel function of the first kind and
of the second kind, respectively.
Equally relevant to number theory are modified Bessel functions. The
modified Bessel function Iν (z) of the first kind is defined by
∞
X 1 z ν+2n
Iν (z) = , (6.9)
n=0
n! Γ(ν + n + 1) 2
whence
In (z) = i−n Jn (iz)
for n ∈ Z.
The modified Bessel function Kν (z) of the second kind is defined
by
π I−ν (z) − Iν (z)
Kν (z) = (6.10)
2 sin πν
(the limit is to be taken for ν ∈ Z), which has the integral representation
Z
1 ∞ − 12 z(t+ 1t ) ν−1 1 π
Kν (z) = e t dt, Re ν > − , | arg z| < . (6.11)
2 0 2 4
This appears in the proof of Theorem 6.1 in the context of Mellin inversion
(sometimes referred to as the inverse Heaviside integral)
Z
1 µ+ν µ−ν µ √
Γ s+ Γ s+ x−s ds = 2 x 2 Kν 2 x , (6.12)
2πi (c) 2 2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
for c + Re µ+ν
2 ≥ Re ν > 0.
Bessel functions are, in a sense, generalizations of exponential functions,
and they reduce to these functions for half-integral order, e.g.
r r
2 2
J 21 (z) = sin(z), J− 12 (z) = cos(z) (6.13)
πz πz
and
r
π −z
K 21 (z) = K− 21 (z) = e . (6.14)
2z
We now introduce the notation (from Terras [Ter1]) concerning the Epstein
zeta-functions, which will be used throughout in what follows.
Notation. Let g, h ∈ Rn be n-dimensional real vectors which (in the
first place) give rise to the perturbation and the (additive) characters, re-
spectively.
Let Y = (yij ) be a positive definite n × n real symmetric matrix. Define
the Epstein zeta-function associated to the quadratic form
n
X
Y [a] = a · Y a = t aY a = yij ai aj , (6.15)
i,j=1
where σ = Re s.
For g, h ∈ Rn define the general Epstein zeta-function (of Hurwitz-
Lerch type) by
X e2πih·a n
Z(Y, g, h, s) = , σ> , (6.17)
Y [a + g]s 2
a∈Zn
a+g6=0
which satisfies the functional equation of the form (5.63) with an additional
factor and replacement of parameters (proof given in §6.4):
1 n
Λ(Y, g, h, s) = p e−2πig·h Λ Y −1 , h, −g, − s . (6.19)
|Y | 2
In what follows we always denote the special vector t 12 , 21 , 21 by c0 :
1
2
c0 = 12 .
1
2
and
3 −1 −1
B = −1 3 −1 . (6.23)
−1 −1 3
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Hautot [Hautot], without giving any reasons, transforms (6.20) into the
form
2
√ MCsCl = 2MN aCl
3
X X X −1/2
2 2 2
+6 (2l) + (2m + 1) + (2n + 1) (6.24)
−1/2
− (2l)2 + (2m + 1)2 + (2n)2 ,
and then proceeds to transfer the triple sum using the Schlömilch series
technique (cf. [KTZ] also). Thus (6.24) suggests that there may be a rela-
tionship between MN aCl and MCsCl structure. This suggestion is strength-
ened by the comparison of numerical values
MN aCl = 1.74756459463 . . .,
(6.25)
MCsCl = 1.76267477307 . . ..
The real situation is the following duality relations (6.26) and (6.27),
which can be found only through the study of lattice structures.
Between
√ the Madelung constants MN aCl = −Z(I, 0, c0 , 21 ) and MCsCl
= − 3 Z(B, 0, c0 , 12 ), the duality relations hold (under the notation (6.23)):
1 2
MN aCl = −Z I, 0, c0 , = − {Z(B, 0, 0, 1) − Z(B, 0, c0 , 1)} (6.26)
2 π
and
√
√ 1 3
MCsCl = − 3 Z B, 0, c0 , =− {Z(I, 0, 0, 1) − Z(I, 0, c0 , 1)}
2 2π
(6.27)
(cf. Formula (1.8) on p.721 of [KTTY1]; proof given in Example 6.2 below).
In the case of MZnS , Hautot states another relation corresponding to
(6.24), again without giving any reason why the Madelung constants MZnS
and MCsCl should be related:
4 2 XXX 1
√ MZnS = √ MCsCl − 6 p ,
3 3 (2l) + (2m + 1)2 + (2n + 1)2
2
(6.28)
where we note that the Madelung constant MZnS is to be defined by
√ √
3 1 1 3 1
MZnS = Z A, c0 , 0, − Z A, 0, 0, , (6.29)
2 2 2 2 2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
where
21 1
A = 1 2 1 . (6.30)
11 2
Comparing (6.27) and (6.29) and the numerical values (6.25) and (6.31)
below does not give much to expect a relation between them;
Let f 1 = e2 + e3 , f 2 = e3 + e1 , f 3 = e1 + e2 , and
011
J = 1 0 1 . (6.35)
110
and
Λ(Y, 0, 0, s) = Λ tJ Y J, 0, 0, s + Λ tJ Y J, c0 , 0, s . (6.37)0
Now note that the inverse matrix (tJ Y J)−1 is the Gram matrix associ-
ated to the dual lattice L01 (∼
= Hom(L1 , Z)) or recall the functional equation
(6.19) to transform the right-hand side of (6.37) further into
1 3 3
p Λ (tJ Y J)−1 , 0, 0, −s − Λ (tJ Y J)−1 , 0, −c0 , −s ,
|tJ Y J| 2 2
so that
Λ(Y, 0, c0 , s)
1 t −1 3 t −1 3
=p Λ ( J Y J) , 0, 0, −s − Λ ( J Y J) , 0, −c0 , −s .
t
| J Y J| 2 2
(6.38)
This explains the reason why the proper definition (given in [KTTY2])
of the Madelung constant MN aCl as the value at s = 21 of
Z(A, c0 , 0, s) − Z(A, 0, 0, s)
n
absolutely convergent for σ > 2. Here we understand the meaning of q(x)
through
L0 ⊗ R ∼
= Hom(L, R) ∼
= HomR (L ⊗ R, R)
and the completion corresponding to (6.18):
where M = (e1 , . . . , en ).
Let φ be the canonical isomorphism
φ : Zn −→ L, x = φ(a) = a1 e1 + · · · + an en , (6.51)
φ : Rn −→ L ⊗ R, x = φ(a) = a1 e1 + · · · + an en ,
X e2πiq◦φ(a)
Z(L, p, q, s) = (6.53)
(φ(a + g), φ(a + g))sL⊗R
a∈Zn
a+g6=0
X e2πih·a
= = Z(Y, g, h, s).
Y [a + g]s
a∈Zn
a+g6=0
Thus, we have
0 0 100
Z 1 ⊕ Z 0 with Gram matrix I = 0 1 0. The zeta-function is
0 1 001
X 1
Z Z3 , 0, 0, s = Z(I, 0, 0, s) = . (6.54)
3
|a|2s
a∈Z
a6=0
1
Z(A, 0, 0, s) = (Z(I, 0, 0, s) + Z(I, 0, c0 , s)) . (6.55)
2
Lb = a ∈ Z3 |a2 + a3 , a3 + a1 , a1 + a2 ∈ 2Z
= a ∈ Z3 |(−1)a2 +a3 + (−1)a3 +a1 + (−1)a1 +a2 = 3
3 −1 −1
with Gram matrix B = −1 3 −1 ((6.23)). The zeta-function is
−1 −1 3
Since
1
1 1
2 0 2 2
3 Z I, 0, 12 , s = Z I, 0, 21 , s + Z I, 0, 0 , s + Z I, 0, 12 , s
1 1
0 2 2 0
X (−1)a2 +a3 + (−1)a3 +a1 + (−1)a1 +a2
= ,
3
I[a]s
a∈Z
a6=0
whence
1
1
21
Z(B, 0, 0, s) = Z(I, 0, 0, s) + 3 Z I, 0, 2 , s . (6.56)
4
0
−1 1 1
K = 1 −1 1 (6.57)
1 1 −1
Z(J Y J, g, h, s) (6.58)
1
2
1 −1 −1 1
= Z(Y, Jg, J h, s) + Z Y, Jg, J h + 2 , s
2 1
2
Z(KY K, g, h, s) (6.59)
0
1 −1 −1 1
= Z(Y, Kg, K h, s) + Z Y, Kg, K h + 2 , s
4 1
2
1
1
2 2
+ Z Y, Kg, K h + 0 , s + Z Y, Kg, K h + 12 , s
−1 −1
.
1
2 0
Example 6.4 For the notation and more details, cf. [KTTY2].
(i) The N aCl (Sodium Chloride) structure. Here the data is
n+ = n− = 1,
S++ = S−− = a ∈ Z3 |a1 + a2 + a3 ∈ 2Z (f.c.c.),
S+− = S+− = a ∈ Z3 |a1 + a2 + a3 ∈ 2Z + 1 (f.c.c.),
so that by (6.40)
1
MN aCl = −Z I, 0, c0 , (6.61)
2
X (−1)a1 +a2 +a3
=− = 1.7475645849 . . .
3
|a|
a∈Z
a6=0
n+ = n− = 1,
3
2
S++ = S−− = √ Z (s.c.),
3
3
2 1
S+− = S+− = √ Z+ (s.c.),
3 2
Hence
√ √
3 1 3 1
MCsCl = Z I, c0 , 0, − Z I, 0, 0, (6.63)
2 2 2 2
√ 1
= − 3 Z B, 0, c0 ,
2
n+ = n− = 1,
( )
2 3
S++ = S−− = √ a a ∈ Z , a1 + a2 + a3 ∈ 2Z (f.c.c.),
3
( 3 )
2 1 1
S+− = S+− = √ a a∈ Z+ , a1 + a 2 + a 3 ∈ 2 Z + (f.c.c.),
3 2 2
n+ = 1, n− = 2,
( )
2 3
S++ = √ a a ∈ Z , a1 + a2 + a3 ∈ 2Z (f.c.c.),
3
3
2 1
S+− = √ Z+ (s.c.)
3 2
3
2
S−− = √ Z (s.c.)
3
( 3 )
2 1 1
S−+ = √ a a ∈ Z + , a1 + a2 + a3 ∈ 2Z + (f.c.c.).
3 2 2
The zeta-function is
1 4 4
ZCaF2 (s) = Z I, c0 , 0, s − 2 Z A, 0, 0, s
2 3 3
1 4 1 4
+ 2Z A, c0 , 0, s − Z I, 0, 0, s
2 3 2 3
s s
1 3 1 3
= Z(I, c0 , 0, s) − Z(I, 0, 0, s)
2 4 2 4
s s
3 1 3
+ Z A, c0 , 0, s − Z(A, 0, 0, s),
4 2 4
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
and which we call the Mellin-Barnes type, being dependent on the Mellin-
Barnes integrals.
As a corollary to Theorem 6.1, we shall prove the Benson-Mackenzie
formula (Corollary 6.1), and for applications of Theorem 6.2, we refer to
[KTTY2].
1 Γ(s − n2 ) 1
+ δ(h) p n n , (6.69)
|Y | π s− 2 bs− 2
where Ks (z) signifies the modified Bessel function of the second kind defined
by (6.10).
Proof. This is Formula (1.25) [KTY7] with the term ε(g)(πb)−s Γ(s) in-
corporated in the left-side member. There the proof depended on the mod-
ular relation, i.e. the Poisson summation modified so as to suit the case. We
refer to Terras for a similar but subtler proof using the Poisson summation
formula.
We may deduce (6.69) from the functional equation (6.19) via the
Mellin-Barnes integral
Z
1 Γ(s − z) Γ(z) −z
(1 + x)−s = x dz (6.70)
2πi (c) Γ(s)
for x > 0, 0 < c < σ, which has been used extensively in various context
(cf. e.g. [KTZ], [KTTY1], [Matsumoto] and [PK]). The proof starts from
expressing the sum in the form of the integral (6.70), applying the functional
equation, and then finally appealing to (6.12).
have
r s
6 π s+1 X X I2 [a] p
2 b 2π
Z(I, 0, c0 , s) = b (−1) K s 2 I 2 [a]b
Γ(s + 1) 2
b2
a∈(Z+ 12 ) b∈Z
b6=0
(6.71)
and
∞
X ∞
X q 2
1
Z I, 0, c0 , = −12π sech a21 + a22 π , (6.72)
2
a1 = 21 a2 = 21
3
Proof. Since for σ > 2
we may write
XX (−1)a1 +a2 +b b2
Z(I, 0, c0 , s) = 3 ,
(a21 + a22 + b2 )s+1
b∈Z a∈Z2
b6=0
whence
!
X X e2πic1 ·a
2 b
Z(I, 0, c0 , s) = 3 b (−1) . (6.73)
(I2 [a] + b2 )s+1
b∈Z a∈Z2
b6=0
We apply Theorem 6.1 to the inner sum on the right of (6.73) to obtain
Z(I, 0, c0 , s)
r s !
X 2 π s+1 X I2 [a + c1 ] p
2 b 2
=3 b (−1) Ks 2 I2 [a + c1 ]b π ,
Γ(s + 1) 2
b2
b∈Z a∈Z
b6=0
(6.74)
which is (6.71).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
To state Theorem
6.2, we introduce new notation.
A B
Let Y = t be a block decomposition with A an n × n matrix
BC
and B an m × m matrix. Set
D = C − t BA−1 B.
In accordance
with this decomposition, we decompose the vectors g =
g1 h1
,h = , g 1 , h1 ∈ Z n , g 2 , h2 ∈ Z m .
g2 h2
Λ(Y, g, h, s) (6.76)
1 n
= δ(g 2 ) e−2πig2 ·h2 Λ(A, g 1 , h1 , s) + δ(h1 ) p Λ D, g 2 , h2 , s −
|A| 2
2e −2πig 1 ·h1 X X −1
+ p e2πi(−g1 ·a+h2 ·b) e−2πiA B(b+g2 )·(a+h1 )
|A| a∈Zn b∈Zm
a+h1 6=0 b+g 2 6=0
s s− n
A−1 [a+ h1 ]
2
p
× Ks− n2 2 A−1 [a + h1 ] D[b + g 2 ] π .
D[b + g 2 ]
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Λ(Y, g, h, s)
X e2πih1 ·a
= δ(g 2 ) π −s Γ(s) e−2πih2 ·g2 (6.79)
A[(a + g 1 ]s
a∈Zn
a+g 1 6=0
X X e2πih1 ·a
+ π −s Γ(s) e2πih2 ·b .
m (A[a+g1 +A−1 B(b+g 2 )]+D[b+g 2 ])
s
b∈Z a∈Zn
b+g 2 6=0
The first sum on the right of (6.79) is Λ(A, g 1 , h1 , s) and to the inner
sum in the second term, we apply Theorem 6.1. Then the second term on
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
b∈Zm ,b6=0
|A| a∈Z n
which are the third and second terms on the right of (6.76), whence the
result follows.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Chapter 7
Abstract
131
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
We then write
∞
X nπ
f (t) ∼ cn ei T t .
n=−∞
n
X kπ
Sn (t) = ck ei T t , (n ∈ N) (7.4)
k=−n
P 2
In particular, we have j∈J |cj | < ∞.
holds.
with the aid of Exercise 7.2 and give a direct proof of Bessel’s inequality in
Corollary 7.1.
and
n
X n
a0 X kπt kπt
Sn (t) = Ak (t) = + ak cos + bk sin . (7.8)
2 T T
k=0 k=1
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
This is called the Fourier cosine (respectively, the Fourier sine) series.
Z 1
1 1
bn = 2 t− sin(2πnt) dt = − , n ∈ N.
0 2 πn
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Since f is piecewise smooth, we should have the equality (by Theorem 7.1
below)
∞
1 X sin(2πnt)
f (t) = − , ∀t ∈ R.
π n=1 n
Fig. 7.1
Fig. 7.2
Fig. 7.3
Z b
lim f (t) eiRt dt = 0.
R→∞ a
π
We shall prove the first assertion. Putting t = u + R , then
Z b Z b−π/R
π
I := f (t) sin(Rt) dt = − f u+ sin(Ru) du.
a a−π/R R
Hence
Z
b−π/R
π
2I = I − f u+ sin(Ru) du
a−π/R R
Z b Z b−π/R π
= f (u) sin(Ru) du + f (u) − f u + sin(Ru) du
b−π/R a R
Z a π
− f u+ sin(Ru) du
a−π/R R
1 1
=O + o(1) + O = o(1), R → ∞.
R R
Exercise 7.7 (i) Divide [a, b] into subintervals to prove that the above
proof can be reduced to the case where f (t) is continuous on the whole
interval. (ii) Prove the remaining assertions of Proposition 7.1.
Theorem 7.1 If f is a periodic function of period 2T and is piecewise
smooth as well as continuous on any finite interval, then its Fourier series
(7.3) converges to f (t) uniformly on any finite interval.
Proof. First we show that {Sn } is convergent, where Sn = Sn (t) is the
n-th partial sum defined by (7.4):
For integers N > M > 0, we have
N
X 1 i kπ t
|SN − SM | = k ck e T
k
|k|=M +1
v v
u X u X
u N u N 1 2
≤t k |ck | t
kπ
2 2
2
ei T t
k
|k|=M +1 |k|=M +1
v
u N
T 0 t
u X 1
≤ kf k
π k2
k=M +1
1
π
sin n + 2 Tu sin n Tπ u cos Tπ u
2 π
= + cos n u,
sin Tπ u2 sin Tπ u2 T
we find that
Z T
1 π
Sn (t) − f (t) = g(u, t) sin n u du
2T −T T
Z T
1 π
+ (f (u + t) − f (u)) cos n u du,
2T −T T
where
(f (u + t) − f (u)) cos Tπ u
2
g(u, t) =
sin Tπ u2
whose possible discontinuity, save for those of f , is at u = 0.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
But
1 1
lim F (t) = f (0−) + f (0+) − f (0−) = f (0+) + f (0−)
t→0− 2 2
and
1 1
lim F (t) = f (0+) − f (0+) − f (0−) = f (0+) − f (0−) ,
t→0+ 2 2
and therefore F (t) is also continuous at 0.
Hence we may apply Theorem 7.1 to conclude that F has the Fourier
series which converges to F (t) everywhere. But g(t) has the Fourier series
∞
1 X sin Tπ nt
−
π n=1 n
Hence it follows that f (t) also has the Fourier series converging to it
at continuities and to F (0) = 21 (f (0+) − f (0−)), thereby completing the
proof.
and
n
X sin πnx cos πx − cos π(2n + 1)x
sin 2πkx = sin((n + 1)πx) = (7.14)
sin πx 2 sin πx
k=1
Solution We have
n
X e2πix e2πinx − 1 e2πinx − 1
2πikx
Sn = e = = .
eπix − 1 1 − e−2πix
k=1
We factor out eπinx (resp. e−πix ) from the numerator (resp. denominator)
to get
eπinx −e−πinx
Sn = eπi(n+1)x 2i
,
eπix −e−πix
2i
which proves (7.13). The imaginary part of (7.12) is the same as (7.14).
Another proof uses the formulae
1
cos α sin β = sin(α + β) − sin(α − β)
2
and
1
sin α sin β = − cos(α + β) − cos(α − β) .
2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
which is (7.13).
The second gives
n
X
sin 2πkx sin πx
k=1
n
1X
=− (cos(2k + 1)πx − cos(2k − 1)πx)
2
k=1
1
= − cos(2n + 1)πx − cos πx ,
2
which is (7.14).
f (x) = f (x + 2n), n ∈ Z.
x
-1 0 1 2 3
Fig. 7.4
1 1
Recalling B 1 (x) = x − [x] − , we have [x] = x − B 1 (x) − , so that
2 2
1 x+1 x+1 1
f (x) = x + 1 − B 1 (x + 1) − − 2 − B1 −
2 2 2 2
1 x+1
= + 2B 1 − B 1 (x + 1).
2 2
Applying the Fourier expansion (7.9) for B 1 (x) below, we see that
∞ ∞
1 2 X sin 2πn x+1
2 1 X sin 2πnx
f(x) = − +
2 π n=1 n π n=1 n
∞ ∞
1 1X2 1 X sin 2πnx
= + (−1)n−1 sin πnx +
2 π n=1 n π n=1 n
∞
1 2 X 2
= + (−1)2m−1 sin 2πmx
2 π m=1 2m
∞ ∞
1 X 2 1X1
+ (−1)2m−2 sin(2m − 1)πx + sin 2πnx
π m=1 2m − 1 π n=1 n
∞
1 2 X 1
= + / Z.
sin(2m − 1)πx, x ∈
2 π m=1 2m − 1
Fig. 7.5
Fig. 7.6
Fig. 7.7
the customary one for deducing (7.9), which appeals to Abel’s continuity
theorem.
We wish to apply this theorem to the Maclaurin expansion for
− log(1 − z).
y
3
x
-1 0 1 2 3
Fig. 7.8
Z z X∞ Z z X∞
1 1 n
− log(1 − z) = dz = z n dz = z ,
0 1−z n=0 0 n=1
n
∞
X 1 iθ n
− log 1 − eiθ = e , θ∈
/ 2πZ. (7.17)
n=1
n
X∞ X∞
1 1
cos nθ + i sin nθ,
n=1
n n=1
n
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
X∞
1
θ − π = −2 sin nθ. (7.19)
n=1
n
Thus we have not only recovered (7.10) again but obtained the Fourier
expansion of the log sin function.
(
0, −1 ≤ x < 0
Example 7.4 The function f (x) = can be expressed
x, 0 ≤ x < 1
as 12 (x + |x|), the positive part f + (x) of f (x) = x, where the positive part
of f (x) is defined to be
f (x) + |f (x)|
f + (x) = .
2
The function f (x) (for its graph, see Fig. 7.9) obtained from f (x) by
continuing it periodically with period 2 is
1 x+1 x+1
f (x) = x−2 + x−2 . (7.20)
2 2 2
x+1
Indeed, for 2n − 1 ≤ x < 2n + 1 (n ∈ Z), we have n ≤ < n + 1,
2
x+1 x+1
and so = n. Hence for 2n − 1 ≤ x < 2n, we have x − 2 =
2 2
x − 2n < 0,so that f (x) = 21 (x − 2n + 2n − x) = 0. For 2n ≤ x < 2n + 1, we
x+1
have x − 2 = x − 2n ≥ 0, and therefore f (x) = 2(x − 2n) = x − |x|.
2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
x
-1 0 1 2 3
Fig. 7.9
x+1
Noting that 2 2 = x − 2B 1 x+1
2 , we deduce that
1 x+1 x+1
f (x) = 2B 1 + 2B 1 (7.21)
2 2 2
x+1 x+1
= B1 + B1
2 2
x+1
We remark that the graph of y = B 1 is the directly connected
2
infinite tents (see Fig. 7.10).
x
-1 0 1 2 3
Fig. 7.10
Of course, if we recall this fact first, then the expression (7.21) would
follow by inspection.
We also note that if we replace the period 2 by 2π, we get
1 1 x + 21 1 x + 21
f 1 (x) = x−2 + x−2 . (7.22)
2 π 2π π 2π
Motivated by the above function, we shall digress here into an equivalent
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1
f2k (x) = tk+1 (x),
2k+1
where by an iterate we mean the successive composition of t:
tk (x) = t tk−1 (x) , t1 (x) = t(x), t0 (x) = 1.
Denoting the directly connected n tents by fn (x) with length n1 and height
1 k k
2n , we note that f2k (x) = t (x) in the case n = 2 .
The Farey sequence Fx of order [x] is defined to be the increasing se-
quence of irreducible fractions ρν between 0 and 1 (0 exclusive) with de-
nominators ≤ x. This may be constructed from the lower order one by
inserting the mediants of adjacent fractions until the denominator reaches
P
[x]. The total number of elements of Fx is n≤x φ(n) = Φ(x), say, where
ϕ(n) indicates the Euler function (cf. (8.27)).
For any even integrable core function f on [0, 1] we define the error term
Φ(x)
X Z 1
Ef (x) = f (ρν ) − Φ(x) f (x) dx,
ν=1 0
Then we have
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Lemma 7.3 ([BKY, Lemmas 2.1 and 2.2]) (i) For the directly con-
nected n tents fn (x), the associated Mellin transform F (s) is given by
1
F (s) = Fn (s),
12n
where
X∞
cn (m)
Fn (s) =
m=1
ms+1
and
where
X X µ(δ)
Cn (s) = d1−s .
δ2
d|n δ|d
d:odd
1 − 2(1−s)(1+k)
C2k (s) =
1 − 21−s
1
which does not vanish for 2 < δ < 1.
Z ∞
(Cf ) (s) = FC (s) = cos(xs) f (x) dx Fourier (cosine) transform
−∞
Z ∞
(Sf) (s) = FS (s) = sin(xs) f (x) dx Fourier (sine) transform
−∞
Z ∞
(LI f ) (s) = FLI (s) = e−xs f (x) dx (one-sided) Laplace transform
0
Z ∞
(LII f ) (s) = FLII (s) = e−xs f (x) dx (two-sided) Laplace transform
−∞
Z ∞
(M f) (s) = FM (s) = xs−1 f (x) dx Mellin transform
0
can be expressed as a Fourier series. If the annulus does not contain the
unit circle, then we cannot write g(θ) = f eiθ even if g(θ) can be expanded
into Fourier series). There is a theory of operators developed on the basis of
two-sided Laplace transforms (cf. [Pa]). It looks as if the Laplace transforms
have driven out Fourier transforms in applied analysis (on the ground that
the former seem to have a wider applicability than the latter), but they
are essentially the same. In comparison with the real Fourier transforms,
the condition that the improper integral for (LII f )(s) be convergent in the
strip β1 < Re s < β2 restricts the class of functions than the condition that
the improper integral for (F f )(y) be convergent for any y ∈ R.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
∞
X Z T /2
iλn t 1
f (t) = cn e , cn = f (x) e−iλn x dx,
n=−∞
T −T /2
2π T
λn = n, |t| < .
T 2
and therefore
∞
1 X
f (t) = T cn eiλn t (λn+1 − λn )
2π n=−∞
∞ Z ∞
1 X ˆ iλn t 1 ˆ eity dy.
→ f (λn ) e ∆λn ∼ f(y)
2π n=−∞ 2π −∞
Viewing this as
Z ∞ Z ∞
1
f (t) = f (x) e−ixy dx eiyt dy,
−∞ 2π −∞
this may be thought of as giving the motivation for the definition of fˆ(x).
Also using the defining equation
Z ∞
f (x) δ(t − x) dx = f (t)
−∞
for the delta function and one of its well known properties
Z ∞
1
eixt dx = δ(t),
2π −∞
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
−1 s
L (t) = cos(ωt). (7.30)
s2 + ω 2
The following proof is, however, much more concise and instructive.
Suppose first that s = σ > 0 and invoke Euler’s identity to deduce that
Z ∞
L[sin(ωt)](s) = Im e−st eiωt dt
0 ∞
1 −(s−iωt)
= Im − e
s − iω 0
1 s + iω ω
= Im = Im 2 = 2 .
s − iωt s + ω2 s + ω2
Now, L[sin(ωt)](s) is an analytic function in s for σ > 0 since the integral
ω
is absolutely convergent there, and so is the function 2 . Hence, by
s + ω2
the principle of analytic continuation (Theorem A.9), they must coincide
in the region σ > 0, and this proves Formula (7.26).
Note that the above argument also gives a proof of (7.29), since
Z ∞
s + iω s
L[cos(ωt)](s) = Re e−st eiωt dt = Re 2 2
= 2 (σ > 0)
0 s +ω s + ω2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
h i 1 √
L erfc at−1/2 (p) = e−2a p (7.31)
p
1 √ √
Γ ,z = π erfc z , (7.32)
2
or more generally,
h a i ν ν √
L Γ ν, (p) = 2a 2 p 2 −1 Kν (2 ap) , (7.33)
t
Kν (z) indicating the modified Bessel function of the second kind defined
by (6.10), which for ν = 21 reduces to (7.31) in view of (6.14) and
Γ(ν) 1
L[Γ(ν, at)](p) = 1− . (7.34)
p (1 + ap )ν
In this section, we shall show that the one-sided (complex) Fourier trans-
form and the Laplace transform have the same function by illustrating
Examples 7.6 and 7.7. We use the following data and scheme.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
y 00 + ω 2 y = a, (7.35)
y(0) = b, y 0 (0) = c,
1 1 1
(iz)2 ŷ(z) − √ izy(+0) + y 0 (+0) + ω 2 ŷ(z) = a √
2π 2πi z
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1
(iz)2 + ω 2 ŷ(z) = √ (b i z 2 + cz − ia)
2π z
1 −ibz 2 − cz + ia 1 A B C
ŷ(z) = √ =√ + +
2π z(z − ω)(z + ω) 2π z z−ω z+ω
−ibz 2 − cz + ia ia a
A = lim z ŷ(z) = lim = 2
= −i 2
z→0 z→0 (z − ω)(z + ω) −ω ω
−ibz 2 − cz + ia
B = lim (z − ω)ŷ(z) = lim
z→∞ z→∞ z(z + ω)
2
−ibω − cω + ia bω 2 − icω − a
= = −i
ω · 2ω 2ω 2
−ibz 2 − cz + ia
C = lim (z + ω)ŷ(z) = lim
z→∞ z→∞ z(z − ω)
2 2
−ibω + cω + ia −ibω + cω − a
= =
−ω · (−2ω) 2ω 2
1 1 1 1
y=√ AF+−1 + BF+−1 + CF+−1
2π z z−ω z+ω
1 √ √ √
=√ A 2π i + B 2π i eiωt + C 2π i e−iωt
2π
a b ω 2 − icω − a iωt b ω 2 + icω − a −iωt
= i · (−i) + e + e
ω2 2ω 2 2ω 2
a b ω 2 − a − icω iωt
= 2 + 2 Re e
ω 2ω 2
a 1
= 2 + 2 (b ω 2 − a) cos(ωt) + cω sin(ωt) ,
ω ω
which is the solution.
Example 7.7 We solve the same differential equation (7.35) under the
same initial conditions as in Example 7.6 by the Laplace transform method.
The above scheme reads.
L[y 00 ] + ω 2 L[y] = aL[1]
1
s2 Y (s) − s y(0) − y 0 (0) + ω 2 Y (s) = a
s
a
s2 + ω 2 Y (s) = b s + c +
s
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
bs2 + cs + a A B C
Y (s) = = + +
s(s2 + ω 2 ) s s − iω s + iω
bs2 + cs + a a
A = lim s Y (s) = lim = 2
s→0 s→0 s2 + ω 2 ω
bs2 + cs + a
B = lim (s − iω) Y (s) = lim
s→iω s→iω s(s + iω)
1 1 1
y = AL−1 + BL−1 + CL−1
s s − iω s + iω
= A + B eiωt + C e−iωt
a b ω 2 − a − icω iωt b ω 2 − a + icω −iωt
= 2
+ e + e
ω 2ω 2 2ω 2
a 1
= 2 + 2 (b ω 2 − a) cos(ωt) + cω sin(ωt) ,
ω ω
which is the solution.
and this can be easily proved by the Residue Theorem (Theorem A.11), the
convergence of the integral taken for granted (which can be checked by the
Stirling formula (Corollary 5.1))
∞
X π X∞
(−1)2m 2m
RHS = Res x−s cos s Γ(s) = cos mπ x
m=0
s=−2m 2 m=0
(2m)!
X∞
(−1)m 2m
= x = cos x,
m=0
(2m)!
Z ∞
1 1 π
M (s) = xs−1 dx = ,
1+• 0 1+x sin πs
Z
1 1 π
= x−s ds
1+x 2πi (c) sin πs
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Z ∞
−•
M e (s) = xs−1 e−x dx =Γ(s),
0
Z
1
e−x = x−s Γ(s) ds.
2πi (c)
The second pair appears in Corollary A.4, while the last pair is the most
well-known one appearing as (2.1) above. The improper integrals being all
absolutely convergent for large values of Re s for which the above formulas
hold. Other examples include (7.23) and (B.5).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Chapter 8
Abstract
Definition 8.1 The space C(q) of all arithmetic functions f (n) of pe-
riod q,
f : Z → C; f (n + q) = f (n), n ∈ Z,
165
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
forms a vector space over C. For f , g ∈ C(q), we define their inner product
X
(f, g) = f (a) g(a).
a mod q
Then C(q) becomes a metric vector space with respect to the norm
1
kf k = (f, f ) 2 .
The notions of orthogonality, ONS (orthonormal system), ONB (or-
thonormal basis) remain the same except that we now speak of the inner
product.
Exercise 8.1 Prove that (1) satisfies the defining properties of a scalar
product (cf. Problem 4.4).
1 1 1
Then ε0 , ε1 , · · · , εq−1 is an ONB of C(q).
q q q
Proof. This follows from the relation (cf. (8.5))
(
X 2πi j−k q, j = k
(εj , εk ) = e q =
a mod q
6 k.
0, j =
homomorphism into C× :
q−1
M
Since C(q) = Cεj , we automatically have the Fourier expansion of f :
i=0
q−1
X
f (n) = ˆ εj (n)
f(j) (8.2)
j=0
1 1 a
χˆj (a) = εa (j) = e−2πi q j (8.3)
q q
q−1 q−1
(
X X 2πi qj a q, j ≡ 0 (mod q)
εj (a) = e = (8.5)
a=0 a=0 0, j 6≡ 0 (mod q).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
E.g.
whence
q−1
X
j +a q
Φ(s, a, z) = q −s z j Φ s, ,z , (8.13)
j=0
q
which continues the left-hand side meromorphically over the whole plane.
Example 8.4 If we trace the above argument in the reverse direction,
we get
X ∞
a χa (n)
q −s ζ s, = .
q n=1
ns
whence we have
−s a 1 X j
q ζ s, = εa (j) ls , (8.18)
q q q
j mod q
being valid for all s 6= 1, whose ls (x) indicates the polylogarithm function
defined by (3.3) (cf. Ishibashi [Is, p.447]).
But, if we state (8.18) in the form (the genuine generalization of the
Eisenstein formula)
q−1
X
a k a
e−2πi q k ls = q 1−s ζ s, − ζ(s), (8.19)
q q
k=1
and interpret the case s = 1 as the limit as s → 1, then (8.19) is valid for all
s ∈ C. The limit interpretation of the right-hand side of (8.19) corresponds
to the normalization ls (0) = ζ(s) in Milnor [Mi], but even under this, (8.16)
is not valid and only (8.19) stands. We state the limiting case of (8.19) as
Theorem 8.1 The limiting case of (8.19) implies Gauss’ formula (8.29).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
of
q−1
X
φ(s, a, ξ) = φj (s, a, ξ), (8.21)
j=0
where φ(s, a, ξ) = Φ(s, a, e2πiξ ) (cf. (8.12)). Note that one expression is a
consequence of (8.4),
1 X 1 X 2πi m−a
q j.
χa (m) = a (j) j (m) = e (8.22)
q q
j mod q j mod q
∞
X e2πimξ
Solution Since φj (s, a, ξ) = , we may express it, on
m=0
(m + a)s
m≡j ( mod q)
writing m = j + lq, l = 0, 1, · · · , as
X∞
e2πi(lqξ+jξ)
φj (s, a, ξ) = q −s s
a+j
l=0 l + q
2πijξ −s a+j
=e q φ s, , qξ .
q
On the other hand, substituting (8.3), we derive that
X∞ q−1
e2πimξ 1 X 2πi m−j
q r
φj (s, a, ξ) = s q
e
m=0
(m + a) r=0
q−1
1 X −2πi qj r X e2πi(ξ+ q )m
∞ r
= e ,
q r=0 m=0
(m + a)s
being valid for any Dirichlet character mod q, not necessarily primitive.
Recall the Laurent expansion for ζ(s, x),
1
ζ(s, x) = − ψ(x) + O(s − 1), s → 1, (8.24)
s−1
Γ0 0
ψ(x) = (x) = (log Γ(x)) . (8.25)
Γ
Also recall the orthogonality of characters
q−1
(
X 0, if χ 6= χ0 ,
χ(a) = (8.26)
a=1 ϕ(q), if χ = χ0 ,
where χ0 and ϕ(q) stand for the principal character mod q and the Euler
function defined by
X
ϕ(q) = 1, (8.27)
1≤a≤q
(a,q)=1
respectively.
From (8.23), (8.24) and (8.26) we obtain
q−1
1X a
L(s, χ) = − χ(a) ψ + O(s − 1), s→1
q a=1 q
and a fortiori
q−1
1X a
L(1, χ) = − χ(a) ψ . (8.28)
q a=1 q
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
For the values of ψ pq , we have a formula of Gauss (2.58) which may be
stated by Lemma 8.1 below as
X
p π p 2pk k
ψ = −γ − log q − cot π + 2 cos π log 2 sin π , (8.29)
q 2 q q
q q
k≤ 2
q−1
π X a
L(1, χ) = χ(a) cot π (8.30)
2q a=1 q
q−1
1 X a
L(1, χ) = − √ b(a) log 2 sin π
χ (8.31)
q a=1 q
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1 X k
b(a) =
χ χ(k) e−2πi q a (8.32)
q
k mod q
q−1
πi X a
L(1, χodd ) = − χ(a) B1 (8.30)0
G(χ) a=1 q
and
q−1
1 X a
L(1, χeven) = − χ(a) log 2 sin π , (8.31)0
G(χ) a=1 q
q−1
πi X a
L(1, χodd ) = − √ b(a) B1
χ , (8.30)00
q a=1 q
the product extending over all prime divisors of q, where µ and Λ signify
the Möbius function and the von Mangoldt function, respectively.
f : Z → C; f (n + q) = f (n), n ∈ Z.
1 + (−1)q q
q−1
X X X
f (a) = 2 f (a) = 2 f (a) + f .
2 2
a=1 a≤ 2q a< 2q
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Proof. For q ≥ 3, the set {±1} forms a subgroup of the reduced residue
×
class group G = (Z/qZ) of index 2. Hence the factor group G/{±1} has
ϕ(q)
order . Since the set of all even characters coincides with the character
2
group of G/{±1}, it follows, from the orthogonality of characters, that
X X
χ(a) = χ(n)
χ even \
χ∈G/{±1}
0 if n 6= 1 in G/{±1},
= ϕ(q)
if n = 1 in G/{±1},
2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
which proves the first assertion. The second assertion follows from the first
and the orthogonality relation
(
X 0 if n 6≡ 1 (mod q)
χ(n) = (8.38)
b
ϕ(q) if n ≡ 1 (mod q).
χ ∈G
b(k)
= q χ(−1) χ
we conclude (8.30)0 .
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
whose inner sum is again G(k, χ). Therefore for χ primitive, we have
q−1
G(χ) X k
L(1, χ) = − χ(k) log 2 sin π ,
q q
k=1
Proof. That (8.29) implies (8.30) and (8.31) is immediate. Indeed, sub-
stituting (8.29) in (8.28) and using Lemma 8.1, we obtain (8.30) for χ odd
and (8.42) for χ even, which is the same as (8.31).
Now we are to prove the converse, i.e. we are to deduce (8.29) from
(8.30) and (8.31).
With p, (p, q) = 1, we multiply (8.28) by χ(p−1 ) and sum over χ mod
q, χ 6= χ0 to obtain
X q
1X a X
χ(p−1 ) L(1, χ) = − ψ χ(ap−1 ) (8.43)
q a=1 q
χ0 6=χ mod q χ0 6=χ mod q
= S1 + S2 ,
say, where
q X
1 X a
S1 = − ψ χ(ap−1 )
q a=1 q
χ mod q
and
q
1 X a
S2 = ψ χ0 (ap−1 ).
q a=1 q
The sum S2 is
q−1
1 X∗ a
S2 = ψ ,
q a=1 q
the star on the summation sign indicating the sum over all a’s, relatively
prime Xto q, (a, q) = 1, which condition may be replaced by introducing the
sum µ(d). Writing the condition d|(a, q) as d|q, a = a0 d ≤ q − 1, we
d|(a,q)
have
d −1 0 q
1 X X a
S2 = µ(d) ψ q
q 0 d
d|q a =1
X
q
d
a+1 q q q
whose inner sum is ψ q = − log − γ + γ by Lemma 8.2.
a=0 d d d d
Hence
log q 1 ϕ(q)
S2 = − ϕ(q) − log Nq − γ, (8.45)
q q q
X
ϕ(q) p 1
χ(p−1 ) L(1, χ) = −ψ − log q − log Nq − γ .
q q ϕ(q)
χ0 6=χ mod q
(8.46)
It remains to calculate the left-hand side of (8.46) by dividing the sum
into two parts:
X X
and
χ0 6=χ even χ odd
First, by (8.31),
X
χ(p−1 ) L(1, χ) (8.47)
χ0 6=χ mod q
q−1
1 X a X
= −√ log 2 sin π b(a)
χ(p) χ
q a=1 q
χ0 6=χ even
q−1 X
1 X a
= −√ log 2 sin π −χ0 (p) χ
c0 (a)
q a=1 q χ even
= T1 + T2 ,
say, where
q−1 X
1 X a 1 X k
T1 = − √ log 2 sin π χ(p) √ χ(k) e−2πi q a (8.48)
q a=1 q χ even
q
k mod q
and
q−1
1 X a 1 X k
T2 = √ log 2 sin π √ χ0 (k) e−2πi q a . (8.49)
q a=1 q q
k mod q
which is equal to
q
µ (a,q)
ϕ(q)
q
ϕ (a,q)
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Comparing (8.46) and (8.54), we see that the terms involving log Nq
cancel each other and (8.29) follows. This completes the proof.
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Appendix A
Complex functions
183
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Proof. We shall prove both Theorem A.3 and its Corollary A.1 at the
same time. By the Cauchy integral formula in Theorem A.7, we have for
any rectifiable simple curve C in D and any point z in C,
X∞ ∞ Z
1 X fn (w)
f (z) = fn (z) = dw.
n=1
2πi n=1 C
w−z
P
∞
But fn (w) is uniformly convergent on C, and so Theorem A.1 allows
n=1
us to integrate term by term after multiplying (w − z)−1 :
Z X ∞ Z
1 fn (w) 1 f (w)
f (z) = dw = dw.
2πi C n=1 w − z 2πi C w − z
Hence the Cauchy integral formula holds for f (z) and so it follows that
f (z) is analytic in C and that
Z
k! f (w)
f (k) (z) = dw, k ∈ N ∪ {0}.
2πi C (w − z)k+1
Let
n
X
Sn (z) = fk (z)
k=1
and take any bounded closed subset D 0 in D. Then take any simple closed
contour C ⊂ D of finite length containing D 0 and suppose dist (D 0 , C) =
δ > 0. The we have
Z
(k) k! Sn (w)
Sn (z) = dw, k ∈ N ∪ {0}.
2πi C (w − z)k+1
Hence, it follows that
k!
Λ(C) max |f (w) − Sn (w)| ,
2πδ k+1 w∈C
whence we have
uniformly on D0 .
Corollary A.2 (The Weierstrass double series theorem) Suppose
{fn (z)} are analytic in |z − z0 | < r and has the Taylor expansion
∞
X (n)
fn (z) = ak (z − z0 )k .
k=0
Then if
∞
X
fn (z) = f (z)
n=1
That is, the iterates of the double series coincide — the order of summation
being interchangeable —
∞ X ∞ ∞ ∞
!
X (n)
X X
k k
ak (z − z0 ) = fn (z) = f (z) = ak (z − z0 )
n=1 k=0 n=1 k=0
and
Z ∞ Z b
M (x) dx < ∞ (resp. M (x) dx < ∞),
a a
R∞ Rb
then a f (x, y) dx (resp. a M (x) dx < ∞) is absolutely and uniformly
convergent on Y .
Following Titchmarsh [Tit], we often refer to this as “by absolute con-
vergence.”
If the series or integrals are convergent but not absolutely convergent,
i.e. conditionally convergent, we need to appeal to more delicate conver-
gence tests such as Dirichlet’s (cf. §B.2).
Exercise A.1 Noting that the principal branch of the natural logarithm
log z (often denoted by Log z) may be defined by the Condition
d 1
log z = , log 1 = 0, (A.1)
dz z
prove the integral representation (Re z > 0)
Z ∞ −t
e − e−zt
log z = dt. (A.2)
0 t
e−(z−1)t − 1
Solution Since the integrand f (t) = f (tz) = −e−t → z−1
Z ∞ t
as t → 0, the improper integral f (t) dt is absolutely convergent. Hence
0
we may differentiate under the integral sign to get
Z ∞ Z ∞
d 1
f (t) dt = e−zt dt = .
dz 0 0 z
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
R∞
Since f (t, 1) = 0, we have log 1 = 0
f (t, 1) dt = 0, and Condition (A.1) is
satisfied.
Example A.1 (Power series) Series in the form of an infinite degree
polynomial
∞
X
f (z) = an (z − z0 )n ,
n=0
X
cl = a m bn
m+n=l
(cf. Remark 1.1); the division is carried out exactly as we do with ordinary
polynomials: e.g. to check the numerical values of Bernoulli numbers in
1 2 1 3
Example 1.1, we may divide z by z + 2! z + 3! z + · · · to obtain
1 1
z 1 6 2 30 4
= 1 − z + z − z +···
ez − 1 2 2! 4!
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
is called the Dirichlet series. Contrary to the case of power series, the region
of absolute convergence of Dirichlet series is a right half-plane and we may
speak about the abscissa of absolute convergence, often denote by σa .
We have a counterpart of Theorem A.5 for Dirichlet series.
Theorem A.6 Let σa denote the abscissa of absolute convergence of
P
∞
a Dirichlet series f (s) = λ−s
n . Then in the region σ > σa , f (s) is
n=1
analytic, integrable and differentiable term by term. Two Dirichlet series
P∞ P∞
f (s) = am m−s and g(s) = bn n−s may be multiplied by the Dirich-
m=1 n=1
let convolution:
∞
X X
f (s) g(s) = cl l−s , cl = a m bn .
l=1 mn=l
Remark A.1 By Theorem A.11, the value of the integral may be de-
termined by computing the residues, which, as stated above, amounts to
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
clearing the denominator. The vacuous case n = 0 implies the most fun-
damental Cauchy Integral Theorem which asserts that the integral along a
closed curve contained in the region of analyticity is 0, which in turn orig-
inates from the fact that in this case the region can be made to shrink into
a point, a topological feature of analytic functions in a region.
Then,
Z ∞
2πi X
xa f (x) dx = Res (z a f (z)) , (A.7)
0 1 − e2πia
z6=0
where the power function is defined by z a = exp (a Log z), Log z signifying
the principal branch.
Proof. First note that the improper integral (A.7) is absolutely conver-
gent both at 0 and ∞ by the Weierstrass M-test. Since f (z) has only finitely
many poles, we may choose 0 < r < R such that all the poles of f (z) other
than the origin lies in the annulus r < |z| < R. Let D denote this annulus
with branch cut along the positive real axis, i.e. its boundary consisting
of the curves C1 : starting from r and moving along the upper edge of the
positive real axis to R, moving along the bigger circle CR and returning
back to the point R, then moving along the lower edge of the positive real
axis to r, moving along the smaller circle cr , and returning to the start-
ing point r. C1 : z = x, 0 ≤ x ≤ R; CR : z = Reiθ , 0 < θ < 2π;
C2 : z = xe2πi , x : R → r; cr : z = re−iθ , 0 < θ < 2π. Then we apply
the residue theorem to this cut region. Since the argument of z increases
from 0 to 2π, we have
Z Z r
a
z f (z) dz = xa e2πia f (x) dx,
C2 R
(with branch cut along the positive real axis as above) along CR and cr :
Z Z
g(z) dz, g(z) dz.
CR cr
By dividing the annulus by any ray starting from the origin and lying
inside the second and the third quadrants, we introduce two regions D1
and D2 with branch cut along the negative and positive imaginary axis,
respectively. For concreteness’ sake, we choose the negative real axis (any
ray can do if there are no poles of the integrand on it):
L : z = xeπi , x : R → r
where CR,1 (resp. cr,1 ) signifies the upper half of CR , (resp. cr ,) . Also
integrating the branch
1 5
g2 (z) = z a f (z), z 6= 0, π<z< π
2 2
along
to get
Z R Z Z Z
− ea(Log x+i0) f (x) dx + g2 + g2 + g2
r CR,2 −L cr.2
where CR,2 (resp. cr,2 ) signifies the lower half of CR , (resp. cr , ). Now note
that except on the positive real axis,
g1 (z) = g(z), D1 ∪ ∂D1
and
g2 (z) = g(z), D2 ∪ ∂D2 ,
so that
Z Z Z R X
+ + 1 − e2πia ea log x f (x) dx = Res z a f (a),
C cr r z∈D
and
Z Z 0 Z 2π
= ra eiaθ f reiθ ireiθ dθ = O Ra+1 f Reiθ dθ .
cr 2π 0
Appendix B
X Z x
a(n) f (n) = A(x) f (x) − A(t) f 0 (t) dt. (B.1)
a<n≤x a
Proof. (The first proof due to Arhipov and Chubarikov.) For x > a(≥ 0),
put
X
F (x) = a(n) f (n) − A(x) f (x),
a<n≤x
Z x
G(x) = − A(t) f 0 (t) dt.
a
197
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
whence it follows that F (x) and G(x) are continuous on [a, x] and piecewise
differentiable on [a, x]. Furthermore
F 0 (x) = G0 (x), / Z.
x∈
Hence, Lemma B.1 applies, and we have F (x) = G(x) + C. And since
F (a) = G(a) = 0, it follows that F (x) = G(x), i.e. (B.1) ensues.
The second proof uses a special case of the formula for integration by
parts in the theory of Stieltjes integrals stated in the following theorem:
Theorem B.2 Suppose f (t) is continuous Ron [a, b] and that α(t) is of
b
bounded variation. Then the Stieltjes integral a f (t) dα(t) exists. Further,
Rb
if f (t) is of bounded variation and α(t) is continuous, then a α(t) df (t)
exists and the formula for integration by parts
Z b h ib Z b
f (t) dα(t) = f (t)α(t) − α(t) df (t) (B.2)
a a a
holds true.
P
Proof. (The second proof.) Putting α(t) = A(t) = a(n), α(t) is
a<n≤t
P
a step function and so of bounded variation. Since a(n) f (n) =
a<n≤x
Rx
a f (t) dα(t), it follows from (B.2) that
X Z x h ix Z x
a(n) f (n) = f (t) dA(t) = f (t) A(t) − A(t) df (t). (B.3)
a a a
a<n≤x
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Rx
Since f ∈ C 1 , the last integral may be written as a
A(t)f 0 (t) dt, and (B.3)
leads to (B.1).
Rb
The Stieltjes integral a f (t) dα(t) exists under weaker conditions than
those stated in Theorem B.2.
Theorem B.3 If f (t) and α(t) are both of bounded variation on [a, b]
and have no common discontinuity, then each one is integrable with respect
to the other from a to b.
X X
LHS = A(n) f (n) − A(m) f (m + 1)
a<n≤x a−1<m≤x−1
[x]−1
X
= −A([a]) f ([a]) + A(n) (f (n) − f (n + 1)) + A(x) f ([x]) ,
n=[a]+1
X Z
[x]−1 n+1
LHS = − A(t) f 0 (t) dt + A(x) f ([x])
n=[a]+1 n
Z [x]
=− A(t) f 0 (t) dt + A(x) f ([x]) .
[a]+1
Z x Z [a]+1 Z x
− A(t) f 0 (t) dt + A(t) f 0 (t) dt + A(t) f 0 (t) dt + A(x) f ([x])
a a [x]
Z x
= A(x) f (x) − A(t) f 0 (t) dt.
a
Proof. (The fourth proof.) We transform the integral on the RHS of (B.1)
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
and
Z x
G(x) = f (t) + B1 (t) f 0 (t) dt,
a
we shall show that F (x) = G(x). Clearly F (a) = G(a) = 0. G(x), being a
function of the upper limit of integration. is continuous and so is F (x) for
x∈/ Z by definition. And for x = m ∈ Z, we note that
X 1
lim F (x) = f (n) − B1 (a) f (a) − f (m) (= F (m)) = lim F (x),
x→m+0 2 x→m−0
a<n≤m
Proof. (The second proof.) (Similar to the second proof of Theorem B.4)
Putting
(−1)r n o
X Xl
F (x) = f (n) − Br (x) f (r−1) (x) − Br (a) f (r−1) (a)
r=1
r!
a<n≤x
and
Z x
(−1)l (l)
G(x) = f (t) + Bl (t) f (t) dt,
a l!
X Z x h ix Z x
n−s = u−s du − B1 (u) u−s + B1 (u) −s u−s−1 du
1 1 1
1<n≤x
Z x
x1−s 1 1
= + − B 1 (x) x−s − − s B 1 (u) u−s−1 du,
1−s s−1 2 1
X Z x
1 1 x1−s
n−s = + + − B 1 (x) x−s − s B 1 (u) u−s−1 du
s−1 2 1−s 1
n≤x
Z ∞
1 1 x1−s
= + + −s B 1 (u) u−s−1 du + O x−σ . (B.4)
s−1 2 1−s 1
1
The first term is−
R∞2 B2 and the second term is absolutely convergent for
σ > −1, whence 1 B 1 (u) u−s−1 du is analytic for σ > −1 and it follows
that (B.5) holds and ζ(s) is analytic for σ > −1 except for s = 1, where it
has a simple pole with residue 1.
Now if we restrict to −1 < σ < 0, then
Z 1 Z 1
−s−1 1 1 1
s B 1 (u) u du = s u− u−s−1 du = − − .
0 0 2 s−1 2
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
1 π X ∞
1 s
s
ζ(s) = (2π) s Γ(−s) sin − s n
π 2 n=1 n (B.9)
π
= 2 (2π)s−1 Γ(1 − s) sin s ζ(1 − s)
2
for −1 < σ < 0, where we used (2.5). For σ < 0, ζ(1 − s) has the Dirichlet
series expression. Formula (B.9) is the asymmetric form of the functional
equation (5.54) for the Riemann zeta-function.
and we are led to find the value of ζ(−1), which we may find with the aid
of (4.3).
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
∞
X
lim f (z) = an z0n .
z→z0
n=0
Proof. By the formula for partial summation (cf. the third proof of The-
orem B.3.), we have for integers M , N , M < N ,
X X
an (x) bn (s) = An (x) (bn (s) − bn+1 (s))
M <n≤N M <n≤N
which can be made < ε for any ε > 0 uniformly in x and s provided that
M and N are sufficiently large. Hence the Cauchy criterion applies and
uniform convergence follows.
For fixed x ∈ R, analyticity of the sum function is a consequence of
Theorem A.3
Corollary B.1 (Dirichlet’s test for uniform convergence) If bn → 0
N
X
as n → ∞, and an (x) = O(1) uniformly in x ∈ R, then the series
n=1
∞
X
bn an (x) is uniformly convergent on R.
n=1
∞
X
Example B.2 If bn → 0, then bn sin 2πnx is uniformly convergent
n=1
in any interval not containing an integer. This follows from Exercise 7.8,
(7.14). In particular, the case bn = n1 establishes the uniform convergence of
∞
1 X sin 2πnx
the Fourier series − for B1 (x) in any interval not containing
π n=1 n
an integer. Therefore, Lebesgue’s theorem allows to integrate it term by
term to obtain B2 (x) (and higher order periodic Bernoulli polynomials).
This example is a special case of the following.
Proposition B.1 The series for the polylogarithm function ls (x) defined
in the first instance for σ > 1 by
X∞
e2πinx
ls (x) =
n=1
ns
Bibliography
[Ap1] T. M. Apostol, Remark on the Hurwitz zeta function, Proc. Amer. Math.
Soc. 2 (1951), 690–693.
[Ap2] T. M. Apostol, On the Lerch zeta-function, Pacific J. Math. 1 (1951),
161–167.
[Ap3] T. M. Apostol, Addendum to ‘On the Lerch zeta-function’, Pacific J. Math.
2 (1952), 10.
[Ap4] T. M. Apostol, Introduction to analytic number theory, Springer, 1976.
[Ber6] B. C. Berndt, Identities involving the coefficients of a class of Dirichlet
series VI, Trans. Amer. Math. Soc. 160 (1971), 157–167.
[Ber3] B. C. Berndt, On the Hurwitz zeta-function, Rocky Mount. J. Math. 2
(1972), 151–157.
[Ber4] B. C. Berndt, Two new proofs of Lerch’s functional equation, Proc. Amer.
Math. Soc. 32 (1972), 403–408.
[Ber2] B. C. Berndt, The gamma function and the Hurwitz zeta-function, Amer.
Math. Monthly 92 (1985), 126–130.
[Böh] P. E. Böhmer, Differenzengleichungen und Bestimmte Integrale, Koecher
Verlag, Berlin, 1939.
[Bor] J. M. Borwein and P. B. Borwein, Pi and the AGM: A study in analytic
number theory and computational complexity, Wiley, 1987.
[BG] P. Bateman and E. Grosswald, On the Epstein zeta-function, Acta Arith. 9
(1964), 365–373.
[BKT] R. Balasubramanian, S. Kanemitsu and H. Tsukada, Contributions to the
theory of Lerch zeta-functions, to appear.
[BKY] R. Balasubramanian, S. Kanemitsu and M. Yoshimoto, Euler products,
Farey series, and the Riemann hypothesis II, Publ. Math. (Debrecen) 69
(2006), 1–16.
[Ca] R. Campbel, Les intégrales Eulériennes et leurs applications. Étude appro-
fondie de la fonction gamma, Dunod, Paris 1966.
[Car] L. Carlitz, A note on the Dedekind sums, Duke Math. J. 21 (1954), 399–403.
[Com] L. Comtet, Advanced Combinatorics: The Art of Finite and Infinite Ex-
pansions, Reidel, Dordrecht, Holland 1974.
[CS] S. Chowla and A. Selberg, On Epstein’s zeta-function (I), Proc. Nat. Acad.
207
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Bibliography 209
Bibliography 211
Index
213
March 27, 2007 17:14 WSPC/Book Trim Size for 9in x 6in vista
Index 215