(Dorian Goldfeld) Automorphic Forms and L-Function
(Dorian Goldfeld) Automorphic Forms and L-Function
DORIAN GOLDFELD
Columbia University
© D. Goldfeld 2006
Cambridge University Press has no responsibility for the persistence or accuracy of urls
for external or third-party internet websites referred to in this publication, and does not
guarantee that any content on such websites is, or will remain, accurate or appropriate.
Dedicated to Ada, Dahlia, and Iris
Contents
Introduction page xi
1 Discrete group actions 1
1.1 Action of a group on a topological space 3
1.2 Iwasawa decomposition 8
1.3 Siegel sets 15
1.4 Haar measure 19
1.5 Invariant measure on coset spaces 23
1.6 Volume of S L(n, Z)\S L(n, R)/S O(n, R) 27
2 Invariant differential operators 38
2.1 Lie algebras 39
2.2 Universal enveloping algebra of gl(n, R) 42
2.3 The center of the universal enveloping algebra of gl(n, R) 46
2.4 Eigenfunctions of invariant differential operators 50
3 Automorphic forms and L–functions for S L(2, Z) 54
3.1 Eisenstein series 55
3.2 Hyperbolic Fourier expansion of Eisenstein series 59
3.3 Maass forms 62
3.4 Whittaker expansions and multiplicity one for G L(2, R) 63
3.5 Fourier–Whittaker expansions on G L(2, R) 67
3.6 Ramanujan–Petersson conjecture 68
3.7 Selberg eigenvalue conjecture 70
3.8 Finite dimensionality of the eigenspaces 71
3.9 Even and odd Maass forms 73
3.10 Hecke operators 74
3.11 Hermite and Smith normal forms 77
3.12 Hecke operators for L2 (S L(2, Z))\h2 80
vii
viii Contents
The theory of automorphic forms and L-functions for the group of n × n invert-
ible real matrices (denoted G L(n, R)) with n ≥ 3 is a relatively new subject.
The current literature is rife with 150+ page papers requiring knowledge of a
large breadth of modern mathematics making it difficult for a novice to begin
working in the subject. The main aim of this book is to provide an essentially
self-contained introduction to the subject that can be read by someone with a
mathematical background consisting only of classical analysis, complex vari-
able theory, and basic algebra – groups, rings, fields. Preparation in selected
topics from advanced linear algebra (such as wedge products) and from the
theory of differential forms would be helpful, but is not strictly necessary for
a successful reading of the text. Any Lie or representation theory required is
developed from first principles.
This is a low definition text which means that it is not necessary for the reader
to memorize a large number of definitions. While there are many definitions,
they are repeated over and over again; in fact, the book is designed so that a
reader can open to almost any page and understand the material at hand without
having to backtrack and awkwardly hunt for definitions of symbols and terms.
The philosophy of the exposition is to demonstrate the theory by simple, fully
worked out examples. Thus, the book is restricted to the action of the discrete
group S L(n, Z) (the group of invertible n × n matrices with integer coefficients)
acting on G L(n, R). The main themes are first developed for S L(2, Z) then
repeated again for S L(3, Z), and yet again repeated in the more general case of
S L(n, Z) with n ≥ 2 arbitrary. All of the proofs are carefully worked out over
the real numbers R, but the knowledgeable reader will see that the proofs will
generalize to any local field. In line with the philosophy of understanding by
simple example, we have avoided the use of adeles, and as much as possible
the theory of representations of Lie groups. This very explicit language appears
xi
xii Introduction
particularly useful for analytic number theory where precise growth estimates
of L-functions and automorphic forms play a major role.
The theory of L-functions and automorphic forms is an old subject with roots
going back to Gauss, Dirichlet, and Riemann. An L-function is a Dirichlet series
∞
an
n=1
ns
where the coefficients an , n = 1, 2, . . . , are interesting number theoretic func-
tions. A simple example is where an is the number of representations of n as a
sum of two squares. If we knew a lot about this series as an analytic function of
s then we would obtain deep knowledge about the statistical distribution of the
values of an . An automorphic form is a function that satisfies a certain differ-
ential equation and also satisfies a group of periodicity relations. An example
is given by the exponential function e2πi x which is periodic (i.e., it has the
same value if we transform x → x + 1) and it satisfies the differential equa-
2
tion ddx 2 e2πi x = −4π 2 e2πi x . In this example the group of periodicity relations
is just the infinite additive group of integers, denoted Z. Remarkably, a vast
theory has been developed exposing the relationship between L-functions and
automorphic forms associated to various infinite dimensional Lie groups such
as G L(n, R).
The choice of material covered is very much guided by the beautiful paper
(Jacquet, 1981), titled Dirichlet series for the group G L(n), a presentation of
which I heard in person in Bombay, 1979, where a classical outline of the theory
of L-functions for the group G L(n, R) is presented, but without any proofs. Our
aim has been to fill in the gaps and to give detailed proofs. Another motivating
factor has been the grand vision of Langlands’ philosophy wherein L-functions
are akin to elementary particles which can be combined in the same way as
one combines representations of Lie groups. The entire book builds upon this
underlying hidden theme which then explodes in the last chapter.
In the appendix a set of Mathematica functions is presented. These have
been designed to assist the reader to explore many of the concepts and results
contained in the chapters that go before. The software can be downloaded by
going to the website given in the appendix.
This book could not have been written without the help I have received from
many people. I am particularly grateful to Qiao Zhang for his painstaking read-
ing of the entire manuscript. Hervé Jacquet, Daniel Bump, and Adrian Diaconu
have provided invaluable help to me in clarifying many points in the theory.
I would also like to express my deep gratitude to Xiaoqing Li, Elon Linden-
strauss, Meera Thillainatesan, and Akshay Venkatesh for allowing me to include
their original material as sections in the text. I would like to especially thank
Introduction xiii
Dan Bump, Kevin Broughan, Sol Friedberg, Jeff Hoffstein, Alex Kontorovich,
Wenzhi Luo, Carlos Moreno, Yannan Qiu, Ian Florian Sprung, C. J. Mozzochi,
Peter Sarnak, Freydoon Shahidi, Meera Thillainatesan, Qiao Zhang, Alberto
Perelli and Steve Miller, for clarifying and improving various proofs, defini-
tions, and historical remarks in the book. Finally, Kevin Broughan has provided
an invaluable service to the mathematical community by creating computer
code for many of the functions studied in this book.
Dorian Goldfeld
1
Discrete group actions
The genesis of analytic number theory formally began with the epoch making
memoir of Riemann (1859) where he introduced the zeta function,
∞
ζ (s) := n −s , ((s) > 1),
n=1
together with precise knowledge of the analytic behavior of ζ (s) could be used
to obtain deep information on the distribution of prime numbers.
One of Riemann’s original proofs of the functional equation is based on the
Poisson summation formula
f (ny) = y −1 fˆ(ny −1 ),
n∈Z n∈Z
1
2 Discrete group actions
If f (y) and fˆ(y) have sufficient decay as y → ∞, then the integral above
converges absolutely for all complex s and, therefore, defines an entire function
of s. Let
∞
dy
f˜(s) = f (y)y s
0 y
denote the Mellin transform of f , then we see from the above integral rep-
resentation and the fact that fˆˆ(y) = f (−y) = f (y) (for an even function f )
that
Choosing f (y) = e−π y , a function with the property that it is invariant under
2
This identity is at the heart of the functional equation of the Riemann zeta
function, and is a known transformation formula for Jacobi’s theta function
∞
2
θ(z) = e2πin z ,
n=−∞
1.1 Action of a group on a topological space 3
a b
where z = x + i y with x ∈ R and y > 0. If is a matrix with integer
c d
coefficients a, b, c, d satisfiying ad − bc = 1, c ≡ 0 (mod 4), c = 0, then the
Poisson summation formula can be used to obtain the more general transfor-
mation formula (Shimura, 1973)
az + b
= d−1 χc (d)(cz + d) 2 θ(z).
1
θ
cz + d
Here χc is the primitive character of order ≤ 2 corresponding to the field exten-
1
sion Q(c 2 )/Q,
1 if d ≡ 1 (mod 4)
d =
i if d ≡ −1 (mod 4),
1
and (cz + d) 2 is the “principal determination” of the square root of cz + d, i.e.,
the one whose real part is > 0.
It is now well understood that underlying the functional equation of the
Riemann zeta function are the above transformation formulae for θ (z). These
transformation
formulae are induced from the action of a group of matrices
a b
on the upper half-plane h = {x + i y | x ∈ R, y > 0} given by
c d
az + b
z
→ .
cz + d
The concept of a group acting on a topological space appears to be absolutely
fundamental in analytic number theory and should be the starting point for any
serious investigations.
Example 1.1.2 Let G denote the additive group of integers Z. Then it is easy
to verify that the group Z acts continuously on the real numbers R with group
4 Discrete group actions
action ◦ defined by
n ◦ x := n + x,
+
1.1.3 Let G = G L(2, R) denote the group of 2 × 2 matrices
Example
a b
with a, b, c, d ∈ R and determinant ad − bc > 0. Let
c d
h := x + i y x ∈ R, y > 0
a b
denote the upper half-plane. For g = ∈ G L(2, R)+ and z ∈ h define:
c d
az + b
g ◦ z := .
cz + d
Since
to be the equivalence class or orbit of x, and let G\X denote the set of equiva-
lence classes.
y
<
(cx + d)2
if the following inequalities hold:
|d| > |c|r + (y −1 ) 2 ≥ |c|r + (δ)− 2 .
1 1
and the total number of such pairs (not counting (c, d) = (0, ±1), (±1, 0)) is at
most 4(δ)−1 (r + 1).
6 Discrete group actions
(1) it is enough to show that for any compact subset A ⊂ h there are only
finitely many g ∈ S L(2, Z) such that (g ◦ A) ∩ A = φ;
(2) every compact subset of A ⊂ h is contained in a rectangle Rr,δ for some
r > 0 and 0 < δ < δ −1 ;
(3) ((αg) ◦ Rr,δ ) ∩ Rr,δ = φ, except for finitely many α ∈ ∞ , g ∈ ∞ \.
To prove (3), note that Lemma 1.1.6 implies that (g ◦ Rr,δ ) ∩ Rr,δ = φ except
for finitely many g ∈ ∞ \. Let S ⊂ ∞ \ denote this finite set of such ele-
ments g. If g ∈ S, then Lemma 1.1.6 tells us that it is because Im(gz) < δ for all
z ∈ Rr,δ . Since Im(αgz) = Im(gz) for α ∈ ∞ , it is enough to show that for each
g ∈ S, there are only finitely many α ∈ ∞ such that ((αg) ◦ Rr,δ ) ∩ Rr,δ = φ.
This last statement follows from the fact that g ◦ Rr,δ itself
lies in some other
1 m
rectangle Rr ,δ , and every α ∈ ∞ is of the form α = (m ∈ Z), so
0 1
that
−1
α ◦ Rr ,δ = x + i y − r + m ≤ x ≤ r + m, 0 < δ ≤ δ ,
Example 1.1.9 A fundamental domain for S L(2, Z)\h can be given as the
region D ⊂ h where
1 1
D = z − ≤ Re(z) ≤ , |z| ≥ 1 ,
2 2
with congruent boundary points symmetric with respect to the imaginary axis.
1.1 Action of a group on a topological space 7
-1 -1/2 0 1/2 1
√
Note that the vertical line V := − 12 + i y y ≥ 23 is equivalent to the
√
vertical line V := 12 + i y
y ≥ 23 under the transformation
z
→ z + 1.
z − 2 ≤ Re(z) < 0, |z| = 1 is equivalent to
1
Furthermore, the arc A
:=
the reflected arc A := z 0 < Re(z) ≤ 12 , |z| = 1 , under the transformation
z
→ −1/z. To show that D is a fundamental domain, we must prove:
We first prove (1). Fix z ∈ h. It follows from Lemma 1.1.6 that for every
> 0, there are at most finitely many g ∈ S L(2, Z) such that g ◦ z lies in the
strip
1 1
D := w − ≤ Re(w) ≤ , ≤ Im(w) .
2 2
Let B denote the finite set of such g ∈ S L(2, Z). Clearly, for sufficiently small
, the set B contains at least one element. We will show that there is at least
one g ∈ B such that g ◦ z ∈ D. Among these finitely many g ∈ B ,choose one
0 −1
such that Im(g ◦ z) is maximal in D . If |g ◦ z| < 1, then for S = ,
1 0
8 Discrete group actions
1 1
T = , and any m ∈ Z,
0 1
−1 Im(g ◦ z)
Im(T Sg ◦ z) = Im
m
= > Im(g ◦ z).
g◦z |g ◦ z|2
This is a contradiction because we can always choose m so that T m Sg ◦ z ∈ D .
So in fact, g ◦ z must be in D.
To complete the verification that D is a fundamental
domain, it only remains
a b
to prove the assertion (2). Let z ∈ D, g = ∈ S L(2, Z), and assume
c d
that g ◦ z ∈ D. Without loss of generality, we may assume that
y
Im(g ◦ z) = ≥ Im(z),
|cz + d|2
(otherwise just interchange z and g ◦ z and√ use g −1 ). This implies that
|cz + d| ≤ 1 which implies that 1 ≥ |cy| ≥ 23 |c|. This is clearly impossi-
ble if |c| ≥ 2. So we only have to consider the cases c = 0, ±1. If c = 0
then d = ±1 and g is a translation by b. Since − 12 ≤ Re(z), Re(g ◦ z) ≤ 12 ,
this implies that either b = 0 and z = g ◦ z or else b = ±1 and Re(z) = ± 12
while Re(g ◦ z) = ∓ 12 . If c = 1, then |z + d| ≤ 1 implies that d = 0 unless
z = e2πi/3 and d = 0, 1 or z = eπi/3 and d = 0, −1. The case d = 0 implies
that |z| ≤ 1 which implies |z| = 1. Also, in this case, c = 1, d = 0, we
must have b = −1 because ad − bc = 1. Then g ◦ z = a − 1z . It follows that
a = 0. If z = e2πi/3 and d = 1, then we must have a − b = 1. It follows that
g ◦ e2πi/3 = a − 1+e12πi/3 = a + e2πi/3 , which implies that a = 0 or 1. A similar
argument holds when z = eπi/3 and d = −1. Finally, the case c = −1 can be
reduced to the previous case c = 1 by reversing the signs of a, b, c, d.
on G L(n, R) and g
t
is the orthogonal group. Here I denotes the identity
matrix
y x
denotes the transpose of the matrix g. The matrix in the decomposition
0 1
α β
(1.2.1) is actually uniquely determined. Furthermore, the matrices
γ δ
d 0 ±1 0
and are uniquely determined up to multiplication by .
0 d 0 ±1
Note that explicitly,
± cos t − sin t
O(2, R) = 0 ≤ t ≤ 2π .
± sin t cos t
We shall shortly give a detailed proof of (1.2.1) for G L(n, R) with n ≥ 2.
The decomposition (1.2.1) allows us to realize the upper half-plane
h = x + i y x ∈ R, y > 0
where
⎧⎛ ⎞ ⎫
⎪ ⎪
⎨ d 0 ⎬
⎜ .. ⎟
Zn = ⎝ . ⎠ d ∈ R, d = 0
⎪
⎩ ⎪
⎭
0 d
is the center of G L(n, R), and O(2, R), Z 2 denotes the group generated by
O(2, R) and Z 2 .
The isomorphism (1.2.2) is the starting point for generalizing the classical
theory of modular forms on G L(2, R) to G L(n, R) with n > 2. Accordingly,
we define the generalized upper half-plane hn associated to G L(n, R).
10 Discrete group actions
u · g · tg = · d (1.2.7)
with
⎛ ⎞
d1
⎜ .. ⎟
d=⎝ . ⎠, d1 , . . . , dn > 0.
dn
a b
For example, consider n = 2, and g = . Then
c d
a b a c a 2 + b2 ac + bd
g· g= t
· = .
c d b d ac + bd c2 + d 2
1 t
If we set u = , then u satisfies (1.2.7) if
0 1
2
1 t a + b2 ac + bd ∗ 0
· = ,
0 1 ac + bd c2 + d 2 ∗ ∗
solve n(n − 1)/2 equations to satisfy (1.2.7). This system of linear equations
has a unique solution because its matrix g · tg is non–singular.
It immediately follows from (1.2.7) that u −1
d = g · tg = d · t
(t u)−1 , or
equivalently
The above follows from the fact that a lower triangular matrix can only equal an
upper triangular matrix if it is diagonal, and that this diagonal matrix must be
d by comparing diagonal entries. The entries di > 0 because g · tg is positive
definite.
Consequently
d = d(t u)−1 . Substituting this into (1.2.7) gives
u · g · tg · t u = d = a −1 · (t a)−1
for
⎛ − 12
⎞
d1
⎜ ⎟
a=⎜
⎝
..
. ⎟.
⎠
− 12
dn
Hence aug · (tg · t u · t a) = I so that aug ∈ O(n, R). Thus, we have expressed
g in the form
g = (au)−1 · (aug),
from which the Iwasawa decomposition immediately follows after dividing and
−1
multiplying by the scalar dn 2 to arrange the bottom right entry of (au)−1 to
be 1.
It only remains to show the uniqueness of the Iwasawa decomposition.
Suppose that zkd = z k d with z, z ∈ hn , k, k ∈ O(n, R), d, d ∈ Z n . Then,
since the only matrices in hn and O(n, R) which lie in Z n are ±I where I is the
identity matrix, it follows that d = ±d. Further, the only matrix in hn ∩ O(n, R)
is I . Consequently z = z and k = ±k .
We shall now work out some important instances of the Iwasawa decompo-
sition which will be useful later.
1.2 Iwasawa decomposition 13
Proposition 1.2.8 Let I denote the identity matrix on G L(n, R), and for every
1 ≤ j < i ≤ n, let E i, j denote the matrix with a 1 at the {i, j}th position and
zeros elsewhere. Then, for an arbitrary real number t, we have
⎛ ⎞
1
⎜ . ⎟
⎜ .. ⎟
⎜ ⎟
⎜ ⎟
⎜ 1
··· 2 t
⎟
⎜ ⎟
1 1
(t 2 + 1) 2 (t + 1) 2
⎜ ⎟
I + t E i, j = ⎜ ..
.
.. ⎟ mod (O(n, R) · R× ) ,
⎜ . ⎟
⎜ 1 ⎟
⎜ (t 2 + 1) 2 ⎟
⎜ ⎟
⎜ .. ⎟
⎝ . ⎠
1
1
where, in the above matrix, 1
1 occurs at position { j, j}, (t 2 + 1) 2 occurs
(t 2 + 1) 2
at position {i, i}, all other diagonal entries are ones, t
1 occurs at position
(t 2 + 1) 2
{ j, i}, and, otherwise, all other entries are zero.
t2
u · g · tg · t u = I + t 2 E i,i − E j, j ,
t2 +1
t
u −1 = I + E j,i ,
t2 + 1
1 &
a −1 =I+ √ − 1 E j, j + t 2 + 1 − 1 E i,i .
t +1
2
Therefore,
1 & t
u −1 a −1 = I + √ − 1 E j, j + t 2 + 1 − 1 E i,i + √ E j,i .
t +1
2 t +1
2
the n ×n identity matrix except for the ith and (i + 1)th rows where we
to be
0 1
have on the diagonal. Then
1 0
⎛ ⎞⎛ ⎞
1 x1,2 x1,3 ··· x1,n y1 y2 · · · yn−1
⎜ 1
x2,3 ···
x2,n ⎟⎜ y1 y2 · · · yn−2
⎟
⎜ ⎟⎜ ⎟
⎜ .. .. ⎟⎜ .. ⎟
ωi z ≡ ⎜ . . ⎟·⎜ . ⎟
⎜ ⎟⎜ ⎟
⎝ 1
xn−1,n ⎠⎝ y1 ⎠
1 1
mod (O(n, R) · R× ) , where yk = yk except for k = n − i + 1, n − i, n − i − 1,
in which case
yn−i '
yn−i = 2 , yn−i±1 = yn−i±1 · 2
xi,i+1 + yn−i
2
,
xi,i+1 + yn−i
2
and xk,
= xk,
except for
= i, i + 1, in which case
2
xi− j,i yn−i + xi− j,i+1 xi,i+1
xi− j,i = x i− j,i+1 − x i− j,i x i,i+1 , xi− j,i+1 = ,
2
xi,i+1 + yn−i
2
for j = 1, 2, . . . , i − 2.
Proposition 1.3.2 Fix an integer n ≥ 2. For any z ∈ hn there are only finitely
many g ∈ n such that g ◦ z ∈ √3 , 1 . Furthermore,
2 2
(
G L(n, R) = g ◦ √3 , 1 . (1.3.3)
2 2
g∈ n
√
Remarks The bound 23 is implicit in the work of Hermite, and a proof can
be found in (Korkine and Zolotareff, 1873). The first part of Proposition 1.3.2
is a well known theorem due to Siegel (1939). For the proof, we follow the
exposition of Borel and Harish-Chandra (1962).
∗
where t,u denotes the subset of matrices t,u · Z n which have determinant 1
∗ ∗
and ◦ denotes the action of S L(n, Z) on 0,∞ . Note that every element in a,b
16 Discrete group actions
is of the form
⎛ ⎞ ⎛ ⎞
1 x1,2 x1,3 ··· x1,n dy1 y2 · · · yn−1
⎜ 1 x2,3 ··· x2,n ⎟ ⎜ dy1 y2 · · · yn−2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ .. .. ⎟ ⎜ . ⎟
⎜ . . ⎟·⎜ .. ⎟
⎜ ⎟ ⎜ ⎟
⎝ 1 xn−1,n ⎠ ⎝ dy1 ⎠
1 d
(1.3.5)
where the determinant
⎛ ⎞
dy1 y2 · · · yn−1
⎜ dy1 y2 · · · yn−2 ⎟
⎜ ⎟
⎜ .. ⎟
Det ⎜ . ⎟ = 1,
⎜ ⎟
⎝ dy1 ⎠
d
so that
−1/n
n−1
d= yin−i .
i=1
φ : S L(n, R) → R>0
from S L(n, R) to the positive real numbers. For all g = (gi, j )1≤i, j≤n in S L(n, R)
we define
'
φ(g) := ||en · g|| = gn,12
+ gn,2
2
+ · · · + gn,n
2 .
To verify the claim, note that for k ∈ S O(n, R), and v ∈ Rn , we have
& √ √
||v · k|| = (v · k) · t (v · k) = v · k · t k · t v = v · t v = ||v||.
This immediately implies that φ(gk) = φ(g), i.e., the claim is true.
∗
Note that if z ∈ 0,∞ is of the form (1.3.5), then
−1/n
n−1
φ(z) = d = yi(n−i) . (1.3.6)
i=1
∗
Now, if z ∈ 0,∞ is fixed, then
en · S L(n, Z) · z ⊂ Ze1 + · · · + Zen − {(0, 0, . . . , 0)} · z, (1.3.7)
⎛ ⎞
1 u 1,2 u 1,3 ··· u 1,n
⎜ 1 u 2,3 ··· u 2,n ⎟
⎜ ⎟
⎜ .. .. ⎟
u=⎜ . . ⎟ ∈ S L(n, Z)
⎜ ⎟
⎝ 1 u n−1,n ⎠
1
to arrange that the minimum of φ lies in ∗√3 1 . This does not change the
2 ,2
value of φ because of the identity φ(u · z) = ||en · u · z|| = ||en · z||. We shall
use induction on n. We have already proved a stronger statement for n = 2
in Example 1.1.9. Fix γ ∈ S L(n, Z) such that φ(γ ◦ z) is minimized. We set
γ ◦ z = x · y with
⎛ ⎞ ⎛ ⎞
1 x1,2 x1,3 · · · x1,n dy1 y2 · · · yn−1
⎜ 1 x2,3 · · · x2,n ⎟ ⎜ dy1 y2 · · · yn−2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ .. . ⎟ ⎜ .. ⎟
x =⎜ . .. ⎟ , y = ⎜ . ⎟,
⎜ ⎟ ⎜ ⎟
⎝ 1 xn−1,n ⎠ ⎝ dy1 ⎠
1 d
18 Discrete group actions
Since |xn−1,n | ≤ 12 we see that φ(αγ z)2 ≤ d 2 (y12 +14 ). On the other hand, the
assumption of minimality forces φ(γ z)2 = d 2 ≤ d 2 y12 + 14 . This implies that
√
y1 ≥ 2
3
.
g 1
Step 2 Let g ∈ S L(n − 1, Z), g = . Then φ(gγ z) = φ(γ z).
0 1
This follows immediately from the fact that en · g = en .
√
Step 3 yi ≥ 3
for i = 2, 3, . . . , n − 1.
2
z ·d ∗
Let us write γ ◦ z = with z ∈ S L(n − 1, R) and d ∈ Z n−1
d
a suitable diagonal matrix. By induction, there exists g ∈ S L(n − 1, Z) such
that g ◦ z = x · y ∈ ∗√3 1 ⊂ hn−1 , the Siegel set for G L(n − 1, R). This is
2 ,2
equivalent to the fact that
⎛ ⎞
an−1
⎜ an−2 ⎟
⎜ ⎟
y = ⎜ .. ⎟
⎝ . ⎠
a1
and
√
a j+1 3
≥ for j = 1, 2, . . . , n − 2. (1.3.9)
aj 2
g 0
Define g := ∈ S L(n, Z). Then
0 1
g 0 z ·d ∗ g ◦ z · d ∗
g◦γ ◦z = ◦ = = x · y ,
0 1 d 0 d
1.4 Haar measure 19
yd 0 x ∗
where y = , x = . The inequalities (1.3.9) applied to
0 d 0 1
⎛ ⎞
y1 y2 · · · yn−1 d
⎜ .. ⎟
yd 0 ⎜ . ⎟
y = =⎜ ⎟,
0 d ⎝ y1 d ⎠
d
√
imply that yi ≥ 3
for i = 2, 3, . . . , n − 1. Step 2 insures that multiplying by
2 √
g on the left does not change the value of φ(γ z). Step 1 gives y1 ≥ 2
3
.
(g, h) → g · h −1
Let gl(n, R) denote the Lie algebra of G L(n, R). Viewed as a set, gl(n, R)
is just the set of all n × n matrices with coefficients in R. We assign a topology
20 Discrete group actions
must be an open set since {0} is closed. Also, the operations of addition and
multiplication of matrices in gl(n, R) are continuous maps from
given by Inv(g) = g −1 for all g ∈ G L(n, R), is also continuous since each entry
of g −1 is a polynomial in the entries of g divided by Det(g). Thus, G L(n, R)
is a topological subspace of gl(n, R) and we may view G L(n, R) × G L(n, R)
as the product space. Since the multiplication and inversion maps: G L(n, R) ×
G L(n, R) → G L(n, R) are continuous, it follows that G L(n, R) is a topological
group.
By a left Haar measure on a locally compact Hausdorff topological group
G, we mean a positive Borel measure (Halmos, 1974)
µ : {measurable subsets of G} → R+ ,
which is left invariant under the action of G on G via left multiplication. This
means that for every measurable set E ⊂ G and every g ∈ G, we have
µ(g E) = µ(E).
In a similar manner, one may define a right Haar measure. If every left invariant
Haar measure on G is also a right invariant Haar measure, then we say that G
is unimodular.
1.4 Haar measure 21
Given a left invariant Haar measure µ on G, one may define (in the usual
manner) a differential one-form dµ(g), and for compactly supported functions
f : G → C an integral
f (g) dµ(g),
G
for every measurable set E. We shall also refer to dµ(g) as a Haar measure.
The fundamental theorem in the subject is due to Haar.
We shall not need this general existence theorem, because in the situations
we are interested in, we can explicitly construct the Haar measure and Haar
integral. For unimodular groups, the uniqueness of Haar measure follows easily
from Fubini’s theorem. The proof goes as follows. Assume we have two Haar
measures µ, ν on G, which are both left and right invariant. Let h : G → C be
a compactly supported function satisfying
h(g) dµ(g) = 1.
G
= h(g ) f (g)dν(g)dµ(g )
G G
=c· f (g ) dµ(g )
G
* −1
where c = G h(g )dν(g).
22 Discrete group actions
⎛ ⎞
⎜ ) ⎟
because gs, j also occurs in the product ⎝ dgi, j ⎠ and dgs, j ∧ dgs, j = 0.
1≤i< j≤n
i=r
Consequently, the measure is invariant under left multiplication by x̃r,s .
then f (gh) = f (g) for all h ∈ H. Thus, f H is well defined on the coset
H H
the coset space. For any measurable subset E ⊂ G/H , we may easily choose
a measurable function δ E : G → C so that
1 if g H ∈ E,
δ E (g) = δ EH (g H ) =
0 if g H ∈ E.
and
f (g) dµ(g) = f H (g H ) d µ̃(g H ),
G G/H
Remarks There is an analogous version of Theorem 1.5.1 for left coset spaces
H \G. Note that we are not assuming that H is a normal subgroup of G. Thus
G/H (respectively H \G) may not be a group.
d ∗ (gz) = d ∗ z
y x
For example, for n = 2, with z = , we have d ∗ z = d xd y
y2
, while for
0 1
n = 3 with
⎛ ⎞
y1 y2 x1,2 y1 x1,3
z= ⎝ 0 y1 x2,3 ⎠ ,
0 0 1
we have
dy1 dy2
d ∗ z = d x1,2 d x1,3 d x2,3 .
(y1 y2 )3
Note that the Weyl group Wn has order n! and is simply the symmetric group on
n symbols. It is clear that d ∗ (gz) = d ∗ z if g is an upper triangular matrix with
1s on the diagonal. This is because the measures d xi, j (with 1 ≤ i < j ≤ n) are
all invariant under translation. It is clear that the differential d ∗ z is Z n -invariant
where Z n ∼= R× denotes the center of G L(n, R). So, without loss of generality,
26 Discrete group actions
Then
az = ax y = (axa −1 ) · ay
⎛ ⎞
1 an−1 x1,2 an−1 an−2 x1,3 · · · an−1 · · · a1 x1,n
⎜ 1 an−2 x2,3 · · · an−2 · · · a1 x2,n ⎟
⎜ ⎟
⎜ .. .. ⎟
=⎜ . . ⎟
⎜ ⎟
⎝ 1 a1 xn−1,n ⎠
1
⎛ ⎞
a1 y1 · · · an−1 yn−1
⎜ .. ⎟
⎜ . ⎟
×⎜ ⎟.
⎝ a1 y1 ⎠
1
n−1
∗ −1
)
Thus d (axa ) = akk(n−k) d ∗ x. It easily follows that
k=1
where ζ (s) is the Riemann zeta function. The fact that the special values (taken
at integral points) of the Riemann zeta function appear in the formula for the
volume is remarkable. Later, Weil (1946) found another method to prove such
results based on a direct application of the Poisson summation formula. A vast
generalization of Siegel’s computation of fundamental domains for the case of
arithmetic subgroups acting on Chevalley groups was obtained by Langlands
(1966). See also (Terras, 1988) for interesting discussions on the history of this
subject.
The main aim of this section is to explicitly compute the volume
d ∗ z,
S L(n,Z)\S L(n,R)/S O(n,R)
∗
where d z is the left–invariant measure given in Proposition 1.5.3. We follow
the exposition of Garret (2002).
to be the left S L(n, R)–invariant measure on hn = S L(n, R)/S O(n, R). Then
n
ζ (
)
d ∗ z = n 2n−1 · ,
S L(n,Z)\hn
=2
Vol(S
−1 )
where
√
2( π)
Vol(S
−1 ) =
(
/2)
+
∞
denotes the volume of the (
− 1)–dimensional sphere S
−1 and ζ (
) = n −
n=1
denotes the Riemann zeta function.
28 Discrete group actions
Proof for the case of S L(2, R) We first prove the theorem for S L(2, R). The
more general result will follow by induction. Let K = S O(2, R) denote the
maximal compact subgroup of S L(2, R). We use the Iwasawa decomposition
which says that
,
1
1 x y2 0
S L(2, R)/K ∼
= z = x ∈ R, y > 0 .
y− 2
1
0 1 0
F(z) := f ((m, n) · z).
(m,n)∈Z2
a b
If γ = ∈ S L 2(Z), then
c d
a b
F(γ z) = f (m, n) · ·z
c d
(m,n)∈Z2
= f (ma + nc, mb + nd) · z
(m,n)∈Z2
= F(z).
- .
{(m, n) ∈ Z2 } = (0, 0) ∪
· (0, 1) · γ 0 <
∈ Z, γ ∈ ∞ \S L(2, Z) ,
(1.6.2)
where
1 r
∞ = r ∈Z .
0 1
The factor 2 occurs because −1 −1 acts trivially on h2 . We easily observe
that
1
y2 0 1
f
(0, 1) · z = f
(0, 1) · − 12
= f 0,
y − 2 .
0 y
y → 2 y, y → y −2
that
∞
d xd y
F(z) 2 = f ((0, 0)) · Vol(\h2 ) + 22 ζ (2) f ((0, y)) ydy. (1.6.3)
y
\h2 0
To complete the proof, we make use of the Poisson summation formula (see
appendix) which states that for any z ∈ G L(2, R)
1
F(z) = f ((m, n)z) = fˆ (m, n) · (t z)−1
(m,n)∈Z2
|Det(z)| (m,n)∈Z2
= fˆ (m, n) · (t z)−1 ,
(m,n)∈Z2
y− 2 x
1 1
y2
since z = and Det(z) = 1. We now repeat all our computations
y− 2
1
0
with the roles of f and fˆ reversed. Since the group is stable under transpose–
inverse, one easily sees (from the Poisson summation formula above), by letting
z
→ (t z)−1 , that the integral
d xd y
F(z) 2
y
\h2
If we combine (1.6.5) and (1.6.6) and solve for the volume, we obtain
2ζ (2)
f ((0, 0)) − fˆ((0, 0)) · vol(\h2 ) = f ((0, 0)) − fˆ((0, 0)) · .
π
Since f is arbitrary, we can choose f so that f ((0, 0)) − fˆ((0, 0)) = 0. It
follows that
2ζ (2) π
Vol(\h2 ) = = .
π 3
Proof for the case of S L(n, R) We shall now complete the proof of
Theorem 1.6.1 using induction on n.
and that d x1 d x2 · · · d xn = t n−1 dt dµ(θ). Then the volume of the unit sphere,
Vol(S n−1 ), is given by
√
2( π)n
Vol(S ) =
n−1
dµ(θ ) = .
S n−1 (n/2)
Since f is a rotationally invariant function, it follows that
1
f (0, . . . , 0, t) = f (x1 , . . . , xn ) dµ(θ)
Vol(S n−1 ) S n−1
1.6 Volume of S L(n, Z)\S L(n, R)/S O(n, R) 33
n−1 −1/n
)
with t = Det(y)−1/n = yin−i .
i=1
In analogy to the previous proof for S L(2, R) we let f : Rn /K n → C be an
arbitrary smooth compactly supported function. We shall also define a function
F : S L(n, R)/K n → C by letting
F(z) := f (m · z).
m∈Zn
denote the set of all n × n matrices in S L(n, Z) with last row (0, 0, . . . , 0, 1).
Let n = S L(n, Z). Then we have as before:
F(z) = f (0) + f (
en · γ · z),
0<
∈Z γ ∈Pn \n
The factor 2 occurs because −In (In = n × n identity matrix) acts trivially on
hn . The computation of the integral above requires some preparations.
We may express z ∈ hn in the form
⎛ ⎞
y1 y2 · · · yn−1 t
⎜ y1 y2 · · · yn−2 t ⎟ − 1
⎜ ⎟ 1
⎜ .. ⎟ t n − 1 · In−1 t n − 1 · In−1
z = x ·⎜ . ⎟· · ,
⎜ ⎟ t −1 t
⎝ y1 t ⎠
t
⎜ y1 y2 · · · yn−2 · t n − 1
n
⎟
⎜ ⎟ 1
⎜ .. ⎟ t − n − 1 · In−1
×⎜ . ⎟·
⎜ ⎟ t
⎝ n
y1 · t n − 1 ⎠
1
⎛ ⎞
1 x1,n
⎜ 1 x2,n ⎟
− 1
⎜ ⎟
⎜ .. .. ⎟ z t n − 1 · In−1
=⎜ . . ⎟· · ,
⎜ ⎟ 1 t
⎝ 1 xn−1,n ⎠
1
(1.6.12)
1.6 Volume of S L(n, Z)\S L(n, R)/S O(n, R) 35
where
⎛ ⎞
1 x1,2 x1,3 ··· x1,n−1
⎜ 1 x2,3 ··· x2,n−1 ⎟
⎜ ⎟
⎜ .. .. ⎟
z =⎜ . . ⎟
⎜ ⎟
⎝ 1 xn−2,n−1 ⎠
1
⎛ n ⎞
y1 y2 · · · yn−1 · t n−1
⎜ n
y1 y2 · · · yn−2 · t n−1 ⎟
⎜ ⎟
×⎜ .. ⎟.
⎝ . ⎠
n
y1 · t n−1
n−2
−(k+1)(n−1−k)−1
= d xi, j yk+1 dyk ,
1≤i< j≤n k=0
we see that
n−1
dy1
d ∗ z = d ∗ z d x j,n t n . (1.6.13)
j=1
y1
−(n−1)/n
n−1
−(n−i)/n
t = y1 yi ,
i=2
we see that
dt n − 1 dy1
=− + ,
t n y1
where is a differential form involving dy j for each j = 2, 3, . . . , n − 1, but
36 Discrete group actions
f (
en · z) = f
en · = f (
t en ). (1.6.15)
t
The last thing we need to do is to construct a fundamental domain for the
action of Pn on hn . Every p ∈ Pn can be written in the form
γ b
p= , (with γ ∈ S L(n − 1, Z), b ∈ Zn−1 ).
1
By (1.6.12), we may express z ∈ hn in the form
⎛ ⎞
x1,n
− 1
⎜ ⎟
z u t n−1 · In−1 ⎜ x2,n ⎟
z= · , u=⎜ .. ⎟ ∈ Rn−1 .
1 t ⎝ . ⎠
xn−1,n
It follows that
− 1
γ z γ ·u+b t n−1 · In−1
p·z = · ,
1 t
from which one deduces from Lemma 1.6.7 that
Pn \hn ∼
= S L(n − 1, Z)\hn−1 × (R/Z)n−1 × (0, ∞),
With these preliminaries, we can now continue the calculation of (1.6.11).
It follows from (1.6.14), (1.6.15), and Lemma 1.6.8 that
2 f (
en · z) d ∗ z
>0
Pn \hn
⎛ ⎞ ⎛ ⎞
∞
2n ⎜ ∗ ⎟
⎜
n−1
⎟ dt
= ⎝ d z⎠ ⎝ d x j,n ⎠ f (
t en ) t n
n − 1
>0 j=1
t
n−1 \hn−1 (R/Z)n−1 0
∞
2n dt
= ζ (n) Vol n−1 \hn−1 · f (t en ) t n
n−1 t
t=0
2n fˆ(0)
= ζ (n) Vol n−1 \hn−1 · . (1.6.16)
n−1 Vol(S n−1 )
1.6 Volume of S L(n, Z)\S L(n, R)/S O(n, R) 37
which holds for Det(z) = 1. Since the group n is stable under transpose–
inverse, we can repeat all our computations with the roles of f and fˆ reversed,
and the integral
F(z) d ∗ z
n \hn
It has been shown in the previous chapter that discrete group actions can give
rise to functional equations associated to important number theoretic objects
such as the Riemann zeta function. Thus, there is great motivation for studying
discrete group actions from all points of view. Let us explore this situation in
one of the simplest cases. Consider the additive group of integers Z acting on
the real line R by translation as in Example 1.1.2. The quotient space Z\R is just
the circle S 1 . One may study S 1 by considering the space of all possible smooth
functions f : S 1 → C. These are the periodic functions that arise in classical
Fourier theory. The Fourier theorem says that every smooth periodic function
f : S 1 → C can be written as a linear combination
f (x) = an e2πinx
n∈Z
where
1
an = f (x)e−2πinx d x,
0
for all n ∈ Z. In other words, the basic periodic functions, e2πinx with n ∈ Z,
form a basis for the space L2 (Z\R) . It is clear that a deeper understanding of
this space is an important question in number theory. We shall approach this
question from the viewpoint of differential operators and obtain a fresh and
illuminating perspective thereby.
A basis for the space L2 (Z\R) may be easily described by using the Laplace
2
operator ddx 2 . One sees that the basic periodic functions e2πinx (with n ∈ Z) are
all eigenfunctions of this operator with eigenvalue −4π 2 n 2 , i.e.,
d 2 2πinx
e = −4π 2 n 2 · e2πinx .
dx2
38
2.1 Lie algebras 39
[a, b] = a ◦ b − b ◦ a, ∀ a, b ∈ A.
L ⊆ Lie(U (L)).
V = ⊕i K vi .
V ⊕W
2.1 Lie algebras 41
⊕ V
vi ⊗ w j , (i = 1, 2, . . . , j = 1, 2, . . .)
and define I (L) to be the two–sided ideal of T (L) generated by all the ten-
sors (linear combinations of tensor products), X ⊗ Y − Y ⊗ X − [X, Y ], with
X, Y ∈ L. The universal enveloping algebra U (L) of L is defined to be
Example 2.1.4 Let L = M(3, R), the Lie algebra of 3 × 3 matrices with
coefficients in R and with Lie bracket
[X, Y ] = X · Y − Y · X
where · denotes matrix multiplication. We shall now exhibit two examples of
multiplication ◦ in U (L). First:
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0 1 0 1 0 0 0 1 0
⎝ 0 0 0 ⎠ ◦ ⎝ 0 0 0 ⎠ = ⎝ 0 0 0 ⎠ ⊗ ⎝ 0 0 0 ⎠ (mod I (L)).
0 0 0 0 0 0 0 0 0 0 0 0
42 Invariant differential operators
X ◦Y −Y ◦ X = X ·Y −Y · X (mod I (L)),
[α, β] = α · β − β · α
for all α, β ∈ gl(n, R), and where · denotes matrix multiplication. We shall
find an explicit realization of the universal enveloping algebra of gl(n, R) as an
algebra of differential operators. We shall consider the space S consisting of
smooth (infinitely differentiable) functions F : G L(n, R) → C.
+
Remark Recall that exp(tα) = I + ∞ k=1 (tα) /k!, where I denotes the
k
identity matrix on gl(n, R). Since we are differentiating with respect to t and
then setting t = 0, only the first two terms in the Taylor series for exp(tα)
matter.
Dα+β = Dα + Dβ ,
Dα ◦ Dβ − Dβ ◦ Dα = D[α,β] ,
where [α, β] = α · β − β · α, denotes the Lie bracket in gl(n, R), i.e., · denotes
matrix multiplication.
Consequently,
n
∂F
(Dβ ◦ Dα − Dα ◦ Dβ )F(g) = (g · (β · α − α · β))i, j ·
i, j=1
∂gi, j
= D[β,α] F(g),
Define
⎛ ⎞
−1
⎜ 1 ⎟
⎜ ⎟
δ1 := ⎜ .. ⎟.
⎝ . ⎠
1
Proposition 2.2.6 For n ≥ 2, let f : G L(n, R) → C be a smooth function
which is left invariant by G L(n, Z), and right invariant by the center Z n . Then
for all D ∈ Dn , D f is also left invariant by G L(n, Z), right invariant by Z n ,
and right invariant by the element δ1 .
∂ ∂ ∂ t1 α1 +···+tm αm
= ··· f ge
∂t1 ∂t2 ∂tm t1 =0,...,tm =0
= D f (g).
∂ ∂ ∂ t1 α1 +···+tm αm
= ··· f ge δ
∂t1 ∂t2 ∂tm t1 =0,...,tm =0
∂ ∂ ∂
= ··· f get1 α1 +···+tm αm
∂t1 ∂t2 ∂tm t1 =0,...,tm =0
= D f (g).
Finally, we must show that D f is right invariant by δ1 . Note that for all
g ∈ G L(n, R),
The associative algebra Dn can also be made into a Lie algebra by defining
a bracket [D, D ] = D ◦ D − D ◦ D for all D, D ∈ Dn . There is a useful
identity given in the next proposition.
i.e.,
(D f )(γ · g · k · δ) = D f (g),
Proof Proposition 2.2.6 proves Proposition 2.3.1 for the left action of
G L(n, Z) and the right action by the center Z n . It only remains to show that
(D f )(g · k) = D f (g) for k ∈ O(n, R) and g ∈ G L(n, R).
Fix the function f , the differential operator D ∈ Dn ,the matrix
g ∈ G L(n, R), and, in addition, fix a matrix h ∈ gl(n, R) which satisfies
h + t h = 0. Given f, D, g, h, we define a function {φ f,D,g,h = φ} : R → C as
follows:
φ(u) := D f (g · exp(uh)) − (D f ) g · exp(uh) .
Clearly φ(0) = 0. We will now show that dφ/du = 0, which implies by ele-
mentary calculus that φ(u) is identically zero. The proof that φ (u) = 0 goes as
follows:
∂
φ (u) = φ(u + t)
∂t t=0
∂
= D f (g · exp((u + t) · h) − (D f ) g · exp((u + t) · h)
∂t t=0
∂
= D f (g · exp(uh) · exp(th) − (D f ) g · exp(uh) · exp(th)
∂t t=0
= (D ◦ Dh )( f (g · exp(uh))) − ((Dh ◦ D) f )(g · exp(uh))
= 0,
because D ◦ Dh = Dh ◦ D.
It follows, as explained before, that φ(u) = 0. Now, the elements exp(uh)
with h + t h = 0 generate O(n, R)+ (the elements of the orthogonal group with
2.3 The center of the universal enveloping algebra of gl(n, R) 47
Thus, D f is invariant on the right by O(n, R)+ . On the other hand, we already
know by Proposition 2.2.6 that D f is invariant on the right by the special
element δ1 of determinant −1. It follows that D f must be right–invariant by
the entire orthogonal group O(n, R). This proves the proposition.
[E i, j , E i , j ] = E i, j · E i , j − E i , j · E i, j (2.3.2)
=δ i , j E i, j −δ
i, j E i , j ,
1 if i = j
where δi, j = is Kronecker’s delta function.
0 otherwise,
Proof Let
n
n
n
D= ··· Di1 ,i2 ◦ Di2 ,i3 ◦ · · · ◦ Dim ,i1 .
i 1 =1 i 2 =1 i m =1
n
n
[Dr,s , D] = [Dr,s , Di1 ,i2 ] ◦ Di2 ,i1 + Di1 ,i2 ◦ [Dr,s , Di2 ,i1 ]
i 1 =1 i 2 =1
n
n
= δi1 ,s Dr,i2 − δr,i2 Di1 ,s ◦ Di2 ,i1 + Di1 ,i2 ◦ δi2 ,s Dr,i1 − δr,i1 Di2 ,s
i 1 =1 i 2 =1
n
n
n
n
= Dr,i2 ◦ Di2 ,s − Di1 ,s ◦ Dr,i1 + Di1 ,s ◦ Dr,i1 − Dr,i2 ◦ Di2 ,s
i 2 =1 i 1 =1 i 1 =1 i 2 =1
= 0.
Example 2.3.4 (The Casimir operator for gl(2, R)) We use the notation
y x
of Proposition 2.3.3 in the case n = 2, and let z = ∈ G L(2, R). By
0 1
Definition 2.2.1 we have the following explicit differential operators acting on
smooth functions f : h2 → C.
∂ y x y x 1 0
D1,1 f (z) := f +t
∂t 0 1 0 1 0 0
t=0
∂ y(1 + t) x
= f
∂t 0 1
t=0
∂
= y f (z).
∂y
∂ y x y x 0 1
D1,2 f (z) := f +t
∂t 0 1 0 1 0 0
t=0
∂ y x + ty
= f
∂t 0 1
t=0
∂
= y f (z).
∂x
∂ y x y x 0 0
D2,1 f (z) := f +t
∂t 0 1 0 1 1 0
t=0
∂ y + tx x
= f
∂t t 1
t=0
∂ y xt 2
+yt+x
= f t +1
2 t +1
2
∂t 0 1
t=0
∂
= y f (z)
∂x
2.3 The center of the universal enveloping algebra of gl(n, R) 49
∂ y x y x 0 0
D2,2 f (z) := f +t
∂t 0 1 0 1 0 1
t=0
∂ y x + tx
= f
∂t 0 1+t
t=0
y
∂ x
= f 1+t
∂t 0 1
t=0
∂
= −y f (z).
∂y
∂ ∂ y x 1 t1 1 0
D2,1 ◦ D1,2 f (z) = f · ·
∂t1 ∂t2 0 1 0 1 t2 1
t1 =0, t2 =0
∂ ∂ t2 x + y(1 + t1 t2 ) x + t1 y
= f
∂t1 ∂t2 t2 1
t1 =0, t2 =0
t1 t22 +t2 +t1
∂ ∂ y
x+( )y
= f t22 +1 t22 +1
∂t1 ∂t2 0 1 t1 =0, t2 =0
∂2
= y2 f (z).
∂x2
∂ ∂2
D2,2 ◦ D2,2 f (z) = y + y2 2 f (z).
∂y ∂y
∂2 ∂2
D1,1 ◦ D1,1 + D1,2 ◦ D2,1 + D2,1 ◦ D1,2 + D2,2 ◦ D2,2 = 2y 2 + 2y 2 2 .
∂y 2 ∂x
The following proposition is the basic result in the subject. As pointed out in
(Borel, 2001), it was first proved by Capelli (1890).
D f (z) = λ D f (z)
2.4 Eigenfunctions of invariant differential operators 51
n−1
n−1 b s
Is (z) := yi i, j j ,
i=1 j=1
where
ij if i + j ≤ n,
bi, j =
(n − i)(n − j) if i + j ≥ n.
The coefficients bi, j are incorporated into the definition because they make
later formulae simpler. Note that since Is (z) is defined on the generalized upper
half-plane, hn , it must satisfy
Is (z · k · a) = Is (z)
Example
(Eigenfunction for G L(2, R)) In this example, we may take
2.4.2
y x
z= , s ∈ C, and Is (z) = y s . Then, we have shown in Example 2.3.4
0 1
2
∂ ∂2
that = y 2 + is a generator of D2 . Clearly,
∂x2 ∂ y2
of any of these with respect to t and then set t = 0 you must get zero as an
1
answer. So the only contribution comes from the off diagonal entry t/(t 2 + 1) 2 .
Consequently
∂
Di, j Is (y) = y1 y2 · · · yn− j Is (y) = 0.
∂ x j,i
GL(n)pack functions The following GL(n)pack functions, described in the
appendix, relate to the material in this chapter:
ApplyCasimirOperator GetCasimirOperator IFun.
3
Automorphic forms and L–functions for
S L(2, Z)
54
3.1 Eisenstein series 55
∂ ∂2
function of the hyperbolic Laplacian = −y ∂ x 2 + ∂ y 2 , with eigenvalue
2
s(1 − s). If (s) ≥ 12 , the function Is (z) is neither automorphic for S L(2, Z)
nor is it square integrable with respect to the G L + (2, R) invariant measure
d xd y/y 2 (over the standard fundamental domain S L(2, Z)\h given in Exam-
ple 1.1.9). The fact that is an invariant differential operator does imply that
az+b a b
(2) E cz+d
, s = E(z, s) for all ∈ S L(2, Z).
c d
(3) E(z, s) = s(1 − s)E(z, s).
The second statement of Proposition 3.1.3 follows easily from the fact that for
every γ ∈ S L(2, Z), we have γ (∞ \S L(2, Z)) = (∞ \S L(2, Z)) . The third
statement is an easy consequence of (3.1.1) and the definition of E(z, s).
The Ramanujan sum can be explicitly evaluated using the Moebius function
µ(n) which is defined by the conditions:
µ(1) = 1, (3.1.5)
1 if n = 1
µ(d) = (3.1.6)
d|n 0 if n > 1,
or equivalently by the identity
1 ∞
µ(n)
= = (1 − p −s ).
ζ (s) n=1
n s
p
d|c m=1
c c
= µ(d) =
µ .
d
d |c, d |n
c
|n,
|c
+
q nm
Here, we have use the fact that the sum e2πi q is zero unless q|n, in which
m=1
case it is q.
For the second part of the proposition, we calculate
S(n; c)
∞ ∞ c ∞
µ(m)
−s
= c
µ =
=
1−s · ζ (s)−1 .
c=1
c s
c=1
|n,
|c
|n m=1
(m
) s
|n
It is easily verified that all the above sums converge absolutely for
(s) > 1.
58 Automorphic forms and L–functions for S L(2, Z)
y x
Theorem 3.1.8 Let (s) > 1 and z = ∈ h2 . The Eisenstein series
0 1
E(z, s) has the Fourier expansion
√
2π s y 1
E(z, s) = y s + φ(s)y 1−s + σ1−2s (n)|n|s− 2 K s− 12 (2π |n|y)e2πinx
(s)ζ (2s) n=0
where
√ s − 12 ζ (2s − 1)
φ(s) = π ,
(s) ζ (2s)
σs (n) = ds,
d|n
d>0
and
∞
1 du
e− 2 y (u+ u ) u s
1 1
K s (y) = .
2 0 u
∞ ∞
−2s
c
2πinr y s e−2πinx
= ζ (2s)y δn,0 + s
c e c d x.
c=1 r =1
(x 2 + y 2 )s
−∞
Since
c
2πinr c c|n
e c =
r =1 0 c |n,
3.2 Hyperbolic Fourier expansion of Eisenstein series 59
it is clear that
1 ∞
e−2πinx y
ζ (2s) E(z, s)e−2πinx d x = ζ (2s)y s δn,0 + σ1−2s (n)y 1−s d x,
0 −∞ (x 2 + 1)s
with the understanding that σ1−2s (0) = ζ (1 − 2s). The proof of the theorem
now immediately follows from the well-known Fourier transform:
⎧√
∞ −2πi x y ⎨ π (s− 2 )
1
if y = 0,
e (s)
dx = (3.1.9)
−∞ (x + 1)
2 s ⎩ 2π |y| K 1 (2π|y|)
s s− 21
if y =
0.
(s) s− 2
Theorem 3.1.10 Let z ∈ h2 and s ∈ C with (s) > 1. The Eisenstein series
E(z, s) and the function φ(s) appearing in the constant term of the Fourier
expansion of E(z, s) can be continued to meromorphic functions on C satisfying
the functional equations:
(1) φ(s)φ(1 − s) = 1;
(2) E(z, s) = φ(s)E(z, 1 − s).
The modified function E ∗ (z, s) = π −s (s)ζ (2s)E(z, s) is regular except
for simple poles at s = 0, 1 and satisfies the functional equation
E ∗ (z, s) = E ∗ (z, 1 − s). Furthermore the residue of the pole at s = 1 is given
by
3
Res E(z, s) =
s=1 π
for all z ∈ h .2
α β
Let ρ = ∈ S L(2, Z) with trace |α + δ| > 2 and γ > 0. Such a
γ δ
matrix is termed hyperbolic. Set D = (α + δ)2 − 4. Then, a point w ∈ C is
termed fixed under ρ if (αw + β)/(γ w + δ) = w. It is easily seen that ρ has
exactly two real fixed points
√ √
α−δ+ D α−δ− D
ω= , ω = .
2γ 2γ
1 −ω −1
Now, define κ = . Then κρκ = is a diagonal
1 −ω −1
matrix with action on z ∈ h2 given by 2 z. Since conjugation preserves the √trace,
we see that + −1 = α + δ. Consequently,
√ = γ ω
+ δ = (α + δ − D)/2
is a unit in the quadratic field
√ Q( D). We shall assume that it is a fundamental
unit, i.e., every unit in Q( D) is, up to ±1, an integral power of > 0. The
Eisenstein series E(κ −1 z, s) is invariant under z → 2 z. This is because
E(κ −1 z, s) = E(ρκ −1 z, s) = E(κ −1 ( 2 z), s).
Therefore, on the positive imaginary axis (i.e., choosing z = iv), the Eisenstein
series ζ (2s) · E(κ −1 z, s) (for (s) > 1) has a Fourier expansion
πin
ζ (2s) · E(κ −1 (iv), s) = bn (s) v log , (3.2.1)
n∈Z
with
2
1 πin dv
bn (s) = ζ (2s) · E(κ −1 (iv), s) v − log .
2 log 1 v
A direct computation shows that
v s · (ω − ω )s
ζ (2s) · E(κ −1 (iv), s) = s .
c,d ∈ Z (cω + d)2 v 2 + (cω + d)2
{c,d}={0,0}
The reason for multiplying by ζ (2s) on the left is to have the sum go over all
c, d ∈ Z ({c, d} = {0, 0}) and not just coprime pairs of c, d. Thus,
2 β
− πin · β 2 s
(ω − ω ) s β log v πin dv
bn (s) = N (β)−s v − log ,
2 log β=0 β v 2 +1 v
β
β
√
integers β1 , β2 ∈ Q( D) satisfy (β1 ) = (β2 ) if and only if β1 = m β2 for some
integer m. Consequently
√ − πin
(N (b) D)s log
−s β
bn (s) = N (β)
2 log b|(β)=0 β
m
2 · ββ m
s
v2 πin dv
× v − log
m∈Z
v +1
2 v
β m
β m
√ − πin
(N (b) D)s β log
= N (β)−s
2 log b|(β)=0 β
−2m+2 · ββ
s
v2 πin dv
× v − log
m∈Z
v +1
2 v
−2m · ββ
√ − πin ∞ s
(N (b) D)s log
−s β v2 πin dv
= N (β) v − log
2 log b|(β)=0 β v2 + 1 v
0
πin πin
s− log s+ log
√
2 2 (N (b) D)s
= ·
(s) 2 log
− πin
β log
× N (β)−s .
β
b|(β)=0
√
For a principal ideal (β) of Q( D), we define the Hecke grössencharakter
− πi
β log
ψ((β)) := ,
β
The space L (S L(2, Z)\h2 ) is actually a Hilbert space with inner product given
2
by
d xd y
f, g := f (z)g(z) 2
y
S L(2,Z)\h2
for all f, g ∈ L2 (S L(2, Z)\h2 ). This inner product was first introduced by
Petersson.
Definition 3.3.1 Let ν ∈ C. A Maass form of type ν for S L(2, Z) is a non–
zero function f ∈ L2 (S L(2, Z)\h2 ) which satisfies:
r f (γ z) = f (z), for all γ ∈ S L(2, Z), z = y x ∈ h2 ;
r f = ν(1 − ν) f ; 0 1
r * 1
0 f (z)d x = 0.
Proposition 3.3.2 Let f be a Maass form of type ν for S L(2, Z). Then ν(1 − ν)
is real and ≥ 0.
Proof The proof is based on the fact that the eigenvalues of a symmetric
operator on a Hilbert space are real. We have by Green’s theorem (integration
3.4 Whittaker expansions and multiplicity one for G L(2, R) 63
by parts) that
ν(1 − ν) f, f = f, f
2
∂ ∂2
= − + 2 f (z) · f (z) d xd y
∂x2 ∂y
2
S L(2,Z)\h
∂ f 2 ∂ f 2
=
∂ x + ∂ y d xd y
S L(2,Z)\h2
= f, f = ν(1 − ν) f, f .
The positivity of f, f and the inner integral above implies that ν(1 − ν) is
real and non–negative.
the unit circle, is characterized by the fact that ψ(x + x ) = ψ(x)ψ(x ), for all
x, x ∈ R. Formally, we have the following definition.
Definition 3.4.2 A Whittaker function of type ν associated to an additive
character ψ : R → U is a smooth non–zero function W : h2 → C which
satisfies the following two conditions:
W (z) = ν(1 − ν)W (z),
1 u
W · z = W (z)ψ(u).
0 1
Proposition 3.4.6 Let ψm (u) = e2πimu , and let W (z, ν, ψm ) be the Whittaker
function (3.4.5). Then we have
√ (π |m|)ν− 12 &
W (z, ν, ψm ) = 2 2π y K ν− 12 (2π |m|y) · e2πimx ,
(ν)
where
∞
1 du
e− 2 y (u+ u ) u ν
1 1
K ν (y) =
2 u
0
where
∞ ν
y
W (y, ν, ψm ) = e−2πium du
u + y2
2
−∞
∞
e−2πiuym
=y 1−ν
du.
(u 2 + 1)ν
−∞
and
∞
π I−ν (y) − Iν (y) 1 dt
e− 2 y (t+ t ) t ν
1 1
K ν (y) = · =
2 sin πν 2 0 t
are classical Bessel functions which have the following asymptotic behavior:
√ 1
lim y Iν (y)e−y = √ ,
y→∞ 2π
2
√ π
lim y K ν (y)e y = .
y→∞ 2
The assumption that (y) has polynomial growth at ∞ forces
&
(y) = a 2π|m|y · K ν− 12 (2π|m|y), (a ∈ C)
Proposition 3.5.1 Let f be a non–constant Maass form of type ν for S L(2, Z).
Then for z ∈ h2 we have the Whittaker expansion
&
f (z) = an 2π y · K ν− 12 (2π|n|y) · e2πinx
n=0
Proof Recall (3.4.1) which says that the fact that f (z) is periodic in x implies
that f has a Fourier expansion of type
f (z) = An (y)e2πinx .
n∈Z
it easily follows that An (y)e2πinx must have polynomial growth at ∞. The proof
of Proposition 3.5.1 is now an immediate consequence of the multiplicity one
Theorem 3.4.8.
for all n = 1, 2, 3, . . . The function is a cusp form of weight 12 for S L(2, Z).
This means that
az + b
= (cz + d)12 (z)
cz + d
a b
for all ∈ S L(2, Z). Petersson generalized Ramanujan’s conjecture to
c d
holomorphic cusp forms f of weight k which satisfy
az + b
f = (cz + d)k f (z)
cz + d
a b
for all ∈ S L(2, Z). Petersson conjectured that the nth Fourier coef-
c d
ficient of a weight k cusp form is bounded by O n (k−1)/2 d(n) . This explains
3.6 Ramanujan–Petersson conjecture 69
the 11
2
in Ramanujan’s conjecture. We remark that Petersson’s conjecture has
been proved by Deligne for all holomorphic cusp forms (of even integral weight)
associated to congruence subgroups of S L(2, Z). Deligne’s proof is based on the
very deep fact that the coefficients of holomorphic cusp forms can be expressed
in terms of the number of points on certain varieties defined over certain finite
fields, and that optimal error terms for the number of points on a variety over a
finite field are a consequence of the Riemann hypothesis (proved by Deligne)
for such varieties.
Non–constant Maass forms for S L(2, Z) can be thought of as non–
holomorphic automorphic functions of weight zero. One might be tempted,
by analogy with the classical theory of holomorphic modular forms, to make
a Petersson type conjecture about the growth of the Fourier coefficients in the
Fourier–Whittaker expansion in Proposition 3.5.1. Remarkably, all evidence
points to the truth of such a conjecture.
We have shown in Proposition 3.5.1 that every non–constant S L(2, Z)–
Maass form of type ν has a Fourier expansion of type
&
f (z) = an 2π y · K ν− 12 (2π|n|y) · e2πinx . (3.6.1)
n=0
|an | = O (d(n))
+
where d(n) = d|n 1 denotes the number of divisors of n, and the O–constant
depends only on the Petersson norm of f .
Proof It follows from Proposition 3.5.1 that for any fixed Y > 0, and any
fixed n = 0,
∞ 1 ∞
d xd y dy
| f (z)|
2
2
= 2π |am | ·
2
|K ν− 12 (y)|2
Y 0 y m=0 2π|m|Y y
∞
dy
|an |2 · |K ν− 12 (y)|2 . (3.6.4)
2π |n|Y y
If we now choose Y = |n|−1 and combine this estimate with (3.6.4) the propo-
sition immediately follows.
We shall now prove this conjecture for S L(2, Z) with a much better lower
bound than 14 . It should be remarked that the eigenvalue 14 can occur for con-
gruence subgroups (see (3.15.1)).
Theorem 3.7.2 (M–F Vigneras) Let f be a Maass form of type ν for
S L(2, Z). Then ν(1 − ν) ≥ 3π 2 /2.
Proof Let D = S L(2, Z)\h2 be the standard fundamental domain (see
Example 1.1.9) for the action of S L(2, Z) on h2 . We also let D ∗ denote the
transform by z → −1/z of D. Now, f is a Maass form of type ν with Fourier–
Whittaker expansion given by (3.6.1):
&
f (z) = an 2π y · K ν− 12 (2π|n|y) · e2πinx .
n=0
2 2
1
∞ 2
∂ f ∂ f
≥ d xd y
∂ x d xd y ≥ ∂x
√
D∪D ∗ 3 − 12
2
∞ & 2
= |an |2 · 4π 2 n 2 · 2π y · K ν− 12 (2π |n|y) dy
√ n=0
3
2
1
∞ 2
3 d xd y
≥ · 4π 2 | f (z)|2
4 √
y2
3 − 12
2
≥ 3π 2 f, f .
Hence, it follows that λ ≥ 3π 2 /2.
is a Maass form of type ν for S L(2, Z). Then, if n 0 is sufficiently large, it follows
that f (z) = 0 for all z ∈ h2 .
Proof Without loss of generality, we may assume that the Petersson norm
f, f = 1. We have
d xd y
1= | f (z)|2
y2
S L(2,Z)\h2
∞ 2 & 2
1
d xd y
≤ an 2π y · K ν− 12 (2π|n|y)e2πinx
√
|n|>n 0
y2
3 − 12
2
∞
& 2 dy
= |an |2 · 2π y · K ν− 12 (2π |n|y) 2
|n|>n 0 √
y
3
2
∞
2 dy
|an |2 · K ν− 12 (2π|n|y) . (3.8.4)
|n|>n 0 √
y
3
2
√
We now make use of Proposition 3.6.3 which says that an is bounded by |n|.
We also make use of the well-known asymptotic formula
√ √π
lim y K ν (y)e y = 2
y→∞
for the K –Bessel function. It immediately follows from these remarks and
3.9 Even and odd Maass forms 73
(3.8.4) that
∞
dy √
1 |n| · e−4π |n|y |n|e −2 3π|n|
e−2n 0 ,
|n|>n 0 √
y2 |n|>n 0
3
2
Finally, we remark that if f is an arbitrary Maass form of type ν for S L(2, Z),
then
13 4 13 4
f (z) = f (z) + T−1 f (z) + f (z) − T−1 f (z)
2 2
3 4 3 4
where 12 f (z) + T−1 f (z) is an even Maass form and 12 f (z) − T−1 f (z) is
an odd Maass form.
(
d
= (g −1 g) ∩ δi , (3.10.3)
i=1
(
d
g = gδi . (3.10.4)
i=1
3.10 Hecke operators 75
Tg : L2 (\X ) → L2 (\X )
by the formula
d
Tg ( f (x)) = f (gδi x),
i=1
d
Tg ( f (γ x)) = f (gδi γ x).
i=1
d
d
Tg ( f (γ x)) = f δi gδσ (i) x = f (gδi x) = Tg ( f (x)),
i=1 i=1
so Tg is well defined.
We also consider for any integer m, the multiple mTg , which acts on
f ∈ L2 (\X ) by the canonical formula (mTg )( f (x)) = m · Tg ( f (x)). In this
+
manner, one constructs the Hecke ring of all formal sums m k Tgk with
k∈Z
m k ∈ Z.
We now define a way of multiplying Hecke operators so that the product
of two Hecke operators is a sum of other Hecke operators. For g, h ∈ C G (),
consider the coset decompositions
( (
g = αi , h = β j , (3.10.6)
i j
as in (3.10.4). Then
( ( ( (
(g) · (h) = gβ j = αi β j = w = w.
j i, j w ⊂ gh w ⊂ gh
76 Automorphic forms and L–functions for S L(2, Z)
One may then define, for g, h ∈ C G (), the product of the Hecke operators,
Tg Th , by the formula
Tg Th = m(g, h, w)Tw , (3.10.7)
w ⊂ gh
Hermite found a canonical or normal form for right or left equivalent matrices
while Smith found a normal form for equivalent matrices. The Hermite and
Smith normal forms play such an important role in Hecke theory that we have
decided to give a self-contained exposition at this point.
Proof We first prove that every matrix in G L(n, Z), with n ≥ 1, is left equiva-
lent to an upper triangular matrix. The result is obvious if n = 1. We proceed by
induction on n. Let t (a1 , a2 , . . . , an ) denote the first column of A. Either every
ai = 0, (i = 1, 2, . . . , n) or some element is non-zero. Suppose the latter and set
δ = 0 to be the greatest common divisor of the elements ai , (i = 1, 2, . . . , n).
+
n
Then there exist coprime integers γ1 , γ2 , . . . , γn such that ai γi = δ. Thus,
i=1
there exists a matrix γ ∈ S L(n, Z) with first row (γ1 , γ2 , . . . , γn ) such that the
matrix γ A has δ in the (1, 1) position and all the remaining elements of the
first column are multiples of δ. By successively multiplying γ A on the left by
matrices which have 1s on the diagonal, zeros everywhere else except for one
element in the first row one may easily show that there exists γ ∈ S L(n, Z)
such that γ A takes the form
⎛ ⎞
δ ∗ ∗ ··· ∗
⎜0 ⎟
⎜ ⎟
⎜ ⎟
γ A = ⎜0 A ⎟
⎜. ⎟
⎝ .. ⎠
0
3.11 Hermite and Smith normal forms 79
Suppose that the submatrix B contains an element bi, j which is not divisible
by bn,n . If we add column j to column n, then the new column n will be of the
form
t
(b1, j , b2, j , . . . , bi, j , . . . , bn−1, j , bn,n ).
We may then repeat the previous process and replace bn,n with a proper
divisor of itself. Continuing in this manner, we obtain a matrix B of the form
(3.11.3) which is equivalent to A and where bn,n divides every element of the
matrix B .
The entire previous process can then be repeated on B . Continuing induc-
tively, we prove our theorem. Again, we leave the proof of uniqueness to the
reader.
g → t g, g ∈ ,
where t g denotes the transpose of the matrix g. Since diagonal matrices are
always invariant under transposition, one immediately sees that the conditions
of Theorem 3.10.10 are satisfied so that the Hecke ring R, is commutative.
Now, two double cosets are either the same or totally disjoint and different.
Thus a union of double cosets can be viewed as an element in the Hecke ring
as in Definition 3.10.8. It follows that for each integer n ≥ 1, we have a Hecke
operator Tn acting on the space of square integrable automorphic forms f (z)
with z ∈ h2 . The action is given by the formula
1 az + b
Tn f (z) = √ f . (3.12.3)
n ad=n d
0≤b<d
Clearly, T1 is just the identity operator. Note that we have introduced the nor-
√
malizing factor of 1/ n to simplify later formulae.
The C-vector space L2 (\h2 ) has a natural inner product (called the
Petersson inner product), denoted , , and defined by
d xd y
f, g = f (z)g(z) ,
y2
\h2
Tn f, g = f, Tn g,
Since the left union above is invariant on the right by any σ ∈ , it follows that
for any σ1 , σ2 ∈ :
( (
m0m1 0
= σ1 ασ2 .
0 m0 α∈S
m 0 m 1 =n
2 n
Further, since diagonal matrices are invariant under transposition, we see that
( (
m0m1 0
= σ1 · t α · σ2 .
0 m0 α∈S
m 0 m 1 =n
2 n
But the action of the Hecke operator is independent of the choice of right coset
decomposition. Consequently
1 d xd y
Tn f, g = √ f (σ1 · t α · σ2 z)g(z)
n α∈Sn
y2
\h2
1 d xd y
=√ f (z) g σ2−1 · t α −1 · σ1−1 z , (3.12.5)
n α∈Sn
y2
\h2
Proof We have already pointed out at the very beginning of this section that
matrix transposition satisfies the conditions of Theorem 3.10.10 which implies
that the Hecke operators commute with each other, i.e.,
Tm Tn = Tn Tm , ∀m, n ≥ 1.
That the Hecke operators also commute with T−1 and , and T−1 and com-
mute with each other is a fairly straightforward calculation which we leave to
the reader.
Tn f = λn f, (3.12.7)
for all n = 1, 2, . . .
Tn f = a(n) · f, ∀n = 1, 2, . . .
It follows that
2
2πdy dy d x+a
Tn f (z) = a(M) · K ν− 12 2π |M| · e2πi M b .
M=0 bd=n 0≤a<b
nb b
The sum over 0 ≤ a < b is zero unless b|M, in which case the sum is b.
Consequently, if we let M = bm , and afterwards m = dm , then we see that
2
2πdby
Tn f (z) = a(bm ) · K ν− 12 (2π|m |dy)e2πidm x
m =0 bd=n
n
mb &
= a · 2π y · K ν− 12 (2π |m|y) e2πimx
m=0 bd=n
d
d|m
mn &
= a · 2π y · K ν− 12 (2π|m|y) e2πimx .
m=0 d|m, d|n
d2
for all values of n. Consequently, if a(1) = 0 the Maass form f would have to
vanish identically. Thus, we may assume that if f = 0, then it is normalized
so that a(1) = 1. In this case, the identity (3.12.10) shows that λn = a(n). The
other claims of Theorem 3.12.8 follow easily.
of the Riemann zeta function, the major difference being that there are two
Euler factors for each prime instead of one. We shall show that (3.13.2) also
86 Automorphic forms and L–functions for S L(2, Z)
y 0
On the other hand, since f is invariant under y
→ y −1 , it follows
0 1
that
∞
y 0 dy
f ys
0 1 y
0
1 ∞
y −1 0 dy y 0 dy
= f y s
+ f ys
0 1 y 0 1 y
0 1
∞ dy
y 0
= f y s + y −s . (3.13.7)
0 1 y
1
y 0
Since f has exponential decay in y as y → ∞, the above integral
0 1
is easily seen to converge for all s ∈ C, and thus defines an entire function. It
is also invariant under the transformation s
→ −s. This gives the functional
equation for even Maass forms.
We now consider the case when f is an odd Maass form. The above argu-
ment does not work because by Proposition 3.9.2, you would have an = −a−n ,
+
and, therefore, an · |n|−s = 0 which implies by the calculation (3.13.6)
n=0
that
∞
y 0 dy
f ys = 0.
0 1 y
0
Hence,
∂ &
an 2π y · K ν− 12 (2π|n|y) · e 2πinx
∂x
n=0 x=0
&
= 2πi an · n · 2π y · K ν− 12 (2π |n|y).
n=0
88 Automorphic forms and L–functions for S L(2, Z)
∞
∂ & dy
an 2π y · K ν− 12 (2π|n|y)e2πinx ys
∂x y
n=0 x=0
0
√ s+ν 1+s−ν
= 2 i · π 1−s L f (s − 12 ). (3.13.8)
2 2
y −x
= f x 2 +y 2 x 2 +y 2 .
0 1
Consequently
∂ y x ∂ y −x
f = f x 2 +y 2 x 2 +y 2
∂x 0 1 ∂x 0 1
x=0 x=0
−1
1 ∂ y x
=− · f
y2 ∂x 0 1
x=0
−1
1 ∂f y 0
=− .
y2 ∂x 0 1
y x
Here, if we let f = f (y, x, z, w), then d
f = f (0,1,0,0) . It then
z w dx
∞
∂f y 0 dy
ys
∂x 0 1 y
0
1 ∞
∂f y −1 0 dy ∂f y 0 dy
=− y s−2 + ys
∂x 0 1 y ∂x 0 1 y
0 1
∞ 6 7 dy
∂f y 0
= y s − y 2−s . (3.13.9)
1 ∂x 0 1 y
The following elementary computation shows that L E(∗,w) (s) is just a product
of two Riemann zeta functions at shifted arguments.
∞
σ1−2w (n) · n w− 2 −s
1
L E(∗,w) (s) =
n=1
∞
∞
∞
n w−s− 2 (md)w−s− 2
1 1
= d 1−2w = d 1−2w
n=1 d|n d=1 m=1
= ζ s + w − ζ s − w + 12 .1
2
Consequently, if we define
s+w− 21 s + w − 1
E(∗,w) (s) := π − 2 2
ζ s + w − 12
2
s−w+ 1
− 2 2
s − w + 12
×π ζ s − w + 12
2
s + w − 1
s − w + 1
= π −s 2
2
ζ s + w − 12 ζ s − w+ 12 ,
2 2
90 Automorphic forms and L–functions for S L(2, Z)
Note that this matches perfectly the functional equation of an even Maass form
of type w (as in Proposition 3.13.5) as it should be.
We shall now show directly that the Eisenstein series is an eigenfunction
of all the Hecke operators. This explains why the L-function associated to the
Eisenstein series has an Euler product.
Then the set Sn given in Lemma 3.12.1 is just a set of coset representatives for
1 \n . If R is any set of coset representatives for ∞ \1 then naturally R Sn is
a set of coset representatives for ∞ \n . On the other hand, Sn R is also a set
of coset representatives for ∞ \n . It follows that
1
Tn E(z, s) = √ E(αz, s)
n α∈∞ \n
1
=√ (γ αz)s
n γ ∈∞ \1 α∈∞ \n
1
=√ (αγ z)s
n γ ∈∞ \1 α∈∞ \n
1 1−s n s
=√ d E(z, s)
n d|n d
1
= n s− 2 σ1−2s (n)E(z, s).
3.15 Converse theorems for SL(2, Z) 91
+
∞
Theorem 3.15.3 (Hecke–Maass converse theorem) Let L(s) = (a(n)/n s )
n=1
(with a(n) ∈ C) be a given Dirichlet series which converges absolutely for (s)
sufficiently large. Assume that for fixed ν ∈ C, L(s) satisfies the functional
equation
ν −s
s + − 12 + ν s + + 12 − ν
(s) := π L(s) = (−1) ν (1 − s),
2 2
with = 0 (respectively = 1), where ν (s) is EBV. Then
&
a(n) 2π y · K ν− 12 (2π|n|y)e2πinx
n=0
must be an even (respectively odd) Maass form of type ν for S L(2, Z), where
we have defined a(n) = (−1) · a(−n) for n < 0.
+ √
Proof For z ∈ h2 , define f (z) = a(n) 2π y · K ν− 12 (2π |n|y)e2πinx , which
n=0
by our assumptions is an absolutely convergent series. Then clearly
f (z) = ν(1 − ν) f (z),
√
because the Whittaker function 2π y K ν− 12 (2π y)e2πi x is an eigenfunction of
with eigenvalue ν(1 − ν). We also get for free the fact that f (z) is periodic
in x. This implies that
1 1
f z = f (z).
0 1
Since S L(2, Z) is generated by the two matrices
1 1 0 −1
,
0 1 1 0
it follows that all that is left to be done to prove the converse theorem is to check
that
y x 0 −1 y x
f = f
0 1 1 0 0 1
y −x
= f x 2 +y 2 x 2 +y 2 . (3.15.4)
0 1
0 = F(z) − λF(z)
∞
3 4
= − (n + 2)(n + 1)y 2 bn+2 (y) − y 2 bn (y) − λbn (y) x n .
n=0
Now, the initial conditions (3.15.6) tell us that b0 (y) = b1 (y) = 0, and it imme-
diately follows from the recurrence relation
y 2 bn (y) + λbn (y)
bn+2 (y) = −
(n + 1)(n + 2)y 2
that bn (y) = 0 for all integers n ≥ 0.
Lemma 3.15.5 implies that we can prove (3.15.4) if the function
y −x
y x x +y x +y
F(z) = f − f
2 2 2 2
0 1 0 1
and similarly
−2 ∂f y −1 0
y = 0.
∂x 0 1
In an analogous manner, if f is odd, it is enough to prove (3.15.8).
94 Automorphic forms and L–functions for S L(2, Z)
We shall prove (3.15.7), (3.15.8) for even and odd f , respectively, using the
Mellin inversion formulae
∞
dy
h̃(s) = h(y)y s (3.15.9)
y
0
σ+i∞
1
h(y) = h̃(s)y −s ds. (3.15.10)
2πi
σ −i∞
These formulae hold for any smooth function h : R+ → C and any fixed real
σ provided h̃(s) is EBV.
We apply (3.15.7), (3.15.8) with h̃(s) = ν (s). It follows from our previous
calculations that we have the Mellin transform pair:
∞ 1
ν ∂ 1 dy
(s) = f (z) y s− 2 ,
0 ∂x x=0 y
1 σ+i∞
∂ 1 dy
ν (s) y 2 −s
1
f (z) = . (3.15.11)
∂x x=0 2πi y
σ −i∞
the constant term in the Fourier expansion around any cusp (a real number
equivalent to ∞ under the discrete group) be zero, and also because the
Eisenstein series is in the continuous spectrum of the Laplace operator. The
latter means that E(z, s) = s(1 − s)E(z, s), or that s(1 − s) is an eigenvalue
of for any complex number s.
Let η j (z), ( j = 1, 2, . . . ) be an orthonormal basis of Maass forms for
S L(2, Z). We may assume as in Theorem 3.12.8 that each η j is an eigen-
function of all the Hecke operators, so that its L-function has an Euler product.
We shall also adopt the convention that
2
3
η0 (z) = ,
π
is the constant function of norm 1. The Selberg spectral decomposition is given
in the following theorem.
∞
1
f (z) = f, η j η j (z) + f, E(∗, s) E(z, s) ds,
j=0
4πi
2 −i∞
1
where
d xd y
f, g = f (z)g(z)
y2
S L(2,Z)\h2
denotes the Petersson inner product on L2 S L(2, Z)\h2 .
We shall not give a complete proof of this theorem, but will only sketch one
of the key ideas of the proof which is contained in the following proposition.
Proposition 3.16.2 Let f (z) ∈ L2 S L(2, Z)\h2 be orthogonal to the
constant function, i.e.,
d xd y
f, 1 = f (z) 2 = 0.
y
S L(2,Z)\h2
where f 0 (z) is automorphic for S L(2, Z) with constant term in its Fourier
*1
expansion equal to zero, i.e., f 0 (z) d x = 0 or f 0 ∈ L2cusp .
0
Proof The main idea of the proof is based on Mellin inversion. Recall that if
h(y) is a smooth complex valued function for y ≥ 0 then the Mellin transform
of h is
∞
dy
h̃(s) = h(y)y s .
0 y
The transform is well defined provided there exists c ∈ R such that the integral
converges absolutely for (s) ≥ c, and in this case, h̃(s) is analytic for (s) ≥ c.
The inverse transform is given by
c+i∞
1
h(y) = h̃(s)y −s ds.
2πi
c−i∞
The proof of Proposition 3.16.2 consists of two steps. In the first step it is
shown that the inner product f, E(∗, s) is the Mellin transform of the constant
term of f (z). In the second step, it is shown that the constant term of
2 +i∞
1
1
f, E(∗, s)E(z, s) ds
4πi
2 −i∞
1
is the inverse Mellin transform of f, E(∗, s) which brings you back precisely
to the constant term of f (z). Thus,
2 +i∞
1
1
f (z) − f, E(∗, s)E(z, s) ds
4πi
2 −i∞
1
The assumption that f is orthogonal to the constant function implies that the
residue at s = 1 of f, E(∗, s̄) is zero. Further, E(z, s) = E ∗ (z, s)/ζ (2s) which
implies (by the fact that ζ (1 + it) = 0 for real t) that Ã0 (s − 1) is holomorphic
for (s) ≥ 12 . The functional equation (Theorem 3.1.10) of the Eisenstein series
tells us that
or equivalently that
Since Ã0 (s − 1) is holomorphic for (s) ≥ 12 , this implies that we may shift
the above line of integration to (s) = 12 . It follows from the transformation
s → 1 − s that
2 +i∞ 2 +i∞
1 1
1 1
A0 (y) = Ã0 (s − 1) y 1−s ds = Ã0 (−s) y s ds.
2πi 2πi
2 −i∞ 2 −i∞
1 1
98 Automorphic forms and L–functions for S L(2, Z)
Maass forms for S L(2, Z) were introduced in Section 3.3. An important objec-
tive of this book is to generalize these functions to the higher-rank group
S L(n, Z) with n ≥ 3. It is a highly non-trivial problem to show that infinitely
many even Maass forms for S L(2, Z) exist. The first proof was given by Selberg
(1956) where he introduced the trace formula as a tool to obtain Weyl’s law,
which in this context gives an asymptotic count (as x → ∞) for the number
of Maass forms of type ν with |ν| ≤ x. Selberg’s methods were extended by
Miller (2001), who obtain Weyl’s law for Maass forms on S L(3, Z) and Müller
(2004), who obtained Weyl’s law for Maass forms on S L(n, Z).
A rather startling revelation was made by Phillips and Sarnak (1985) where it
was conjectured that Maass forms should not exist for generic non-congruence
subgroups of S L(2, Z), except for certain situations where their existence is
ensured by symmetry considerations, see Section 4.1. Up to now no one has
found a single example of a Maass form for S L(2, Z), although Maass (1949)
discovered some examples for congruence subgroups (see Section 3.15). So it
seemed as if Maass forms for S L(2, Z) were elusive mysterious objects and the
non-constructive proof of their existence (Selberg, 1956) suggested that they
may be unconstructible.
Recently, Lindenstrauss and Venkatesh (to appear) found a new, short, and
essentially elementary proof which shows the existence of infinitely many
Maass forms on G(Z)\G(R)/K ∞ where G is a split semisimple group over
Z and K ∞ is the maximal compact subgroup. Lindenstrauss and Venkatesh
were also able to obtain Weyl’s law in a very broad context. Their method
works whenever one has Hecke operators. Although in the case of S L(2, Z),
the proof is perhaps not much simpler than the trace formula, it has the advan-
tage of being much more explicit (it allows one to write down an even cus-
pidal function). However, a much bigger advantage is that it generalizes in a
99
100 Existence of Maass forms
relatively straightforward way to higher rank (unlike the trace formula, where
one encounters formidable technical obstacles). I would very much like to thank
Elon and Akshay for preparing and allowing me to incorporate a preliminary
manuscript which formed the basis of this chapter.
Proposition 4.1.1 There are infinitely many odd Maass forms for S L(2, Z).
defined by
The rest of this chapter will be devoted to showing that the space of even
Maass forms for S L(2, Z) is also infinite dimensional. The only other known
4.2 Integral operators 101
proof of this fact uses the trace formula (Selberg, 1956), see also (Hejhal, 1976),
as we already mentioned earlier.
Proof First differentiate under the integral sign, and then convert to polar
coordinates, to obtain
∞ ∞ ∞
η2 ξ2 η2
F x+ dη = f x + + dξ dη
2 2 2
−∞ −∞ −∞
2π ∞
r2
= f x+ r dr dθ
2
0 0
∞
= 2π f (x + w) dw,
0
Let g(x) be an even smooth function of compact support on the real line with
Fourier transform
∞
h(t) = g(x)eit x d x.
−∞
Following Selberg (1956), we define a variation of the Abel transform, denoted
k, as follows.
∞
1
(v − u)− 2 dq(v),
1
k(u) = − (4.2.4)
π
u
where
1 √ √
q(v) := g 2 log v + 1 + v .
2
The relation between k and h is called the Selberg transform. It is clear from
the definitions that k is compactly supported and continuous. Indeed, if g is
supported in [−M, M], then k(u) vanishes for u > sinh2 (M/2).
For z, w ∈ h2 , let u(z, w) be given by (4.2.1). Then u(z, w) is real valued
and positive. Define
u := u(z, w). (4.2.5)
Definition 4.2.6 (Point pair invariant) Let g : R → C be an even smooth
function of compact support. The point pair invariant K : h2 × h2 → C
associated to g is the function defined by
K (z, w) = k(u(z, w)) = k(u), (for all z, w ∈ h2 ),
where k is given by (4.2.4) and u is given by (4.2.5).
4.2 Integral operators 103
Remark It is a basic fact (although we will not need it for the proof, it
provides valuable intuition) that, if φλ is an eigenfunction of the hyperbolic
Laplacian with eigenvalue λ = 14 + r 2 , then
K ∗ φλ = h(r )φλ . (4.2.9)
On the other hand, by the point pair property, and the fact that d ∗ w is an
S L(2, R)-invariant measure, we have for any α ∈ S L(2, R) that
K (i, w) d w = K ( j) (αi, αw) d ∗ w
( j) ∗
h2 h2
= K ( j) (αi, w) d ∗ w.
h2
4.3 The endomorphism ♥ 105
Now, for any z ∈ h2 , we may choose α ∈ S L(2, R) such that z = αi. It follows
that
K ( j) (z, w) d ∗ w −→ 1
h2
for any z ∈ h2 , as j → ∞.
Consequently, for any continuous function f , we have
K ( j) ∗ f (z) = K ( j) (z, w) f (w) d ∗ w
h2
= K ( j) (z, w) f (z) d ∗ w + K ( j) (z, w)( f (w) − f (z)) d ∗ w
h2 h2
−→ f (z), (4.2.10)
K ( j) ∗ f (z) → f (z)
satisfies
1 1 1
E z, + ir = + r · E z, + ir
2
(4.3.2)
2 4 2
1 −ir 1
T p E z, + ir = ( p + p ) · E z, + ir .
ir
(4.3.3)
2 2
We proceed formally for now. From (4.3.2) and (4.3.3), the operator
√1 √1
♥ := T p − p 4 − − p − 4 −
Let us pause to explain the connection of this with the vague idea that
we described at the start of this section. Work formally for a moment and
suppose that g0 were the “delta-function at 0.” Then, formally speaking, the
Fourier transform h 0 is the constant function 1, whereas the Fourier trans-
form h p (r ) = pir + p −ir . Using (4.2.9) we see (formally speaking – this is not
intended to be a rigorous proof!) that the map f
→ K 0 ∗ f is just the identity
endomorphism
√ √and that the operator f
→ K p ∗ f is essentially the operator
p − 4 + p − − 4 . So f
→ K p ∗ f − T p (K 0 ∗ f ), formally speaking, gives
1 1
Proof For any function F on S L(2, Z)\h2 , we define for y > 0, the constant
term (denoted FC T ):
FC T (y) = F(x + i y) d x.
R/Z
holds for any F on S L(2, Z)\h2 . To prove Lemma 4.4.1, we need to check that,
for any y > 0, (K p ∗ f )C T = (T p (K 0 ∗ f ))C T . So we just need to check that
where we have used the fact that f (w) = f (w + 1) to unfold the integral over
x ∈ R/Z, at the cost of restricting the w-integration from h2 to h2 /{w
→ w + 1}.
108 Existence of Maass forms
Now, using the fact that K (gz, gw) = K (z, w) for g ∈ S L(2, R), we see that
that if w = xw + i yw , then
K p (x + i y, w) d x = K p (x + i y, i yw ) d x (4.4.5)
x∈R x∈R
= yw K p x + i yyw−1 , i d x
x∈R
= (yyw )1/2 g p log yyw−1 .
So we get
∞
dy
y0 y −1/2 g p (log(y0 y −1 )) f C T (y)
1/2
(K p ∗ f )C T (y0 ) = .
y
0
From this and the fact that g p (x) := g0 (x + log( p)) + g0 (x − log( p)), (4.4.3)
follows by a simple computation.
be the Siegel set as in Definition 1.3.1. Then the natural projection: S(T ) to
S L(2, Z)\h2 is a homeomorphism onto an open subset. We can, therefore,
regard Cc∞ (S(T )), the space of smooth compactly supported functions on
S(T ), as a subset of Cc∞ (S L(2, Z) \h2 ); similarly L2 (S(T )) is a subset of
4.5 There exist infinitely many even cusp forms for S L(2, Z) 109
L2 (S L(2, Z)\h2 ). We will make these identifications throughout the rest of this
argument.
If f ∈ C ∞ (S(T )), we define for n ∈ Z, the nth Fourier coefficient an, f (y) to
*1
be the function on (T, ∞) defined by the rule an, f (y) = f (x + i y)e−2πinx d x.
0
Now, let R be so large that k0 (z, w) and k p (z, w) are supported in d(z, w) ≤
R, and let Y ≥ pe R . Then, one sees from (4.3.2) and Definition 4.2.7 that ♥
maps C ∞ (S(Y )) into C ∞ (S(1)). Indeed, it is enough to check that this is true
for f
→ K p ∗ f and f
→ T p (K 0 ∗ f ); we deal with the first and leave the
second to the reader. It is clear from Definition 4.2.7 that K p ∗ f is supported
in an R-neighborhood of the support of f . But an R-neighborhood of S(Y ) is
contained in S(Y/R), thus the claim.
Moreover, if the nth Fourier coefficient an, f (y) vanishes identically, then so
does an,K p ∗ f (y). This follows from (4.4.4):
−2πinx
an,K p ∗ f (y) := e K p (x + i y, w) f (w) d ∗ w d x
x∈Z\R w∈h2
= e−2πinx K p (i y, w − x) f (w) d ∗ w d x
x∈Z\R w∈h2
= K p (i y, w) f (w + x)e−2πinx d x d ∗ w,
w∈h2 x∈Z\R
T p (K 0 ∗ f )(z) = K 0 ∗ f ( pz),
Proof (Sketch only) We follow the notations of the previous section. Fix a
non-zero smooth function h on the real line, supported in (0, 1). Fix an integer
N ≥ 1. For each pair of positive integers j, k satisfying 1 ≤ j, k ≤ N and so that
p does not divide j, we put f jk (x + i y) = h ((N (y − Y )/Y ) − k) cos(2π j x),
regarded as an element of C ∞ (S(Y ))+ ⊂ L2cusp,+ . Let W be the span of f jk , so
an N 2 − N [N / p] dimensional subspace of L2cusp,+ . Also, let V = ♥(W ), where
we take the function g0 entering in the definition of ♥ to be an approximation
to a δ function. Now apply the previous lemma to V .
4.8 Interpretation via wave equation: higher rank case 111
u = u(x + i y, t)
to (4.8.1) may be regarded as describing the amplitude (at time t and position
x + i y) of a wave propagating in the hyperbolic plane. The low order term of
u/4 is natural for the hyperbolic Laplacian (see (Lax and Phillips, 1976)).
For every t ∈ R we can define a linear endomorphism Ut of
L2 S L(2, Z)\h2 ∩ C∞ S L(2, Z)\h2
112 Existence of Maass forms
hn = G L(n, R)/(O(n, R) · R× )
which is invariant under the discrete group S L(n, Z) and which is also an eigen-
function of every invariant differential operator in Dn , the center of the universal
enveloping algebra as defined in Section 2.3. A cuspidal function (or cuspform)
on h2 was defined by the condition that the constant term in its Fourier expansion
vanishes which in turn is equivalent to the condition that φ(z) has exponential
decay as y → ∞. These notions are generalized in the formal Definition 5.1.3.
Harish-Chandra was the first to systematically study spaces of automor-
phic forms in a much more general situation than G L(n). He proved (Harish-
Chandra, 1959, 1966, 1968) that the space of automorphic functions of a cer-
tain type (characterized by a cuspidality condition, eigenfunction condition,
and good growth) is finite dimensional. Godement (1966) explains why Maass
forms on G L(n) are rapidly decreasing. It was not at all clear at that time if
a theory of L-functions, analogous to the G L(2) theory could be developed
for G L(n) with n > 2. The first important breakthrough came in (Piatetski-
Shapiro, 1975), and independently in (Shalika, 1973, 1974), where the Fourier
expansion of a Maass form for S L(n, Z) was obtained for the first time. The
Fourier expansion involved Whittaker functions. In his thesis, Jacquet intro-
duced and obtained the meromorphic continuation and functional equations
of Whittaker functions on an arbitrary Chevalley group (see (Jacquet, 1967)).
These papers provided the cornerstone for an arithmetic theory of L-functions
in the higher-rank situation.
114
5.1 Maass forms 115
with
ij if i + j ≤ n,
bi, j =
(n − i)(n − j) if i + j ≥ n,
is an eigenfunction of every D ∈ Dn . Let us write
D Iν (z) = λ D · Iν (z) for everyD ∈ Dn . (5.1.2)
The function λ D (viewed as a function of D) is a character of Dn because it
satisfies
λ D1 ·D2 = λ D1 · λ D2
for all D1 , D2 ∈ Dn . It is sometimes called the Harish–Chandra character.
Definition 5.1.3 Let n ≥ 2, and let ν = (ν1 , ν2 , . . . νn−1 ) ∈ Cn−1 . A Maass
form for S L(n, Z) of type ν is a smooth function f ∈ L2 (S L(n, Z)\hn ) which
satisfies
(1) f (γ z) = f (z), for all γ ∈ S L(n, Z), z ∈ hn ,
= λ D f (z), for all D ∈ D , with λ D given by (5.1.2),
n
(2) D f (z)
(3) f (uz) du = 0,
(S L(n,Z)∩U )\U
One might reasonably expect that φ̃ m (z) is a Fourier coefficient of φ and that
φ might be recoverable as a sum of such Fourier coefficients. Unfortunately,
the fact that Un (R) is a non–abelian group (for n > 2) complicates the issue
enormously, and it is necessary to go through various contortions in order to
5.2 Whittaker functions associated to Maass forms 117
obtain a useful Fourier theory. We shall study this issue carefully in the next
section. For the moment, we focus on the integral (5.2.3).
Proposition 5.2.4 For n ≥ 2 and ν = (ν1 , ν2 , . . . , νn−1 ) ∈ Cn−1 , let φ be a
Maass form of type ν for S L(n, Z). Let m = (m 1 , . . . , m n−1 ) ∈ Zn−1 , and let
ψm be an additive character as in (5.2.2). Then the function φ̃ m (z) defined in
(5.2.3) satisfies the following conditions:
(1) φ̃ m (u · z) = ψm (u) · φ̃ m (z) (for all u ∈ Un (R)),
(2) Dφ̃ m = λ D φ̃ m , (for all D ∈ Dn ),
(3) |φ̃ m (z)|2 d ∗ z < ∞,
√3 1
2 ,2
where √3 , 1 denotes the Siegel set as in Definition 1.3.1, and d ∗ z is the left
2 2
invariant measure given in Proposition 1.5.3.
Remark Any smooth function: hn → C which satisfies conditions (1), (2),
(3) of Proposition 5.2.4 will be called a Whittaker function. A more formal
definition will be given in Section 5.4.
Proof First of all, the integral on the right-hand side of (5.2.3) is an integral
over Un (Z)\Un (R). Since both φ and ψm are invariant under Un (Z), the integral
is independent of the choice of fundamental domain for Un (Z)\Un (R).
Every z ∈ hn can be written in the form z = x · y, as in the beginning
of Section 5.1. In the integral (5.2.3), we make the change of variables
u → u · x −1 . It follows that
1 1
φ̃ m (z) = ··· φ(u · y) ψm (u · x −1 ) du i, j .
0 0 1≤i< j≤n
where
1
φ̃ m (x) = φ(u + x)e−2πimu du.
0
If φ satisfies conditions (1), (2), but does not satisfy condition (3) of
Definition 5.1.3, then the Fourier expansion takes the form
∞
∞
∞
γ
φ(z) = ··· φ̃ (m 1 ,...,m n−1 ) z .
γ ∈Un−1 (Z)\S L(n−1,Z) m 1 =−∞ m 2 =0 m n−1 =0
1
Proof It is clear that if δ ∈ B\A and δ ∈ C\B then δ δ ∈ C\A. On the other
hand, every γ ∈ C\A can be written in the form γ = ca for c ∈ C, a ∈ A. If
we now set δ = c · 1 ∈ C\B and δ = 1 · a ∈ B\A, then we have expressed
γ = δ δ.
commute with each other and generate the abelian group of all matrices of the
form
⎧⎛ ⎞ ⎫
⎪ 1 m1 ⎪
⎪
⎪ ⎪
⎪
⎪
⎪ ⎜ 1 m ⎟ ⎪
⎪
⎨⎜ 2 ⎟ ⎬
⎜ . . ⎟
⎜ . . .
. ⎟ m 1 , . . . , m n−1 ∈ Z .
⎪
⎪ ⎜ ⎟ ⎪
⎪
⎪
⎪ ⎝ 1 m n−1 ⎠ ⎪
⎪
⎪
⎩ ⎪
⎭
1
120 Maass forms and Whittaker functions for S L(n, Z)
1
define
1 1
φ̂ m (z) := ··· φ(vz)e−2πiv,m d ∗ v,
0 0
+n−1
where v, m = i=1 vi m i . Note the difference between φ̂ and φ̃, for
example, φ̃ involves integration with repect to all the variables u i, j with
1 ≤ i < j ≤ n − 1. With this notation, (5.3.4) becomes
φ(z) = φ̂ m (z). (5.3.5)
m∈Zn−1
The Fourier expansion (5.3.5) does not make use of the fact that φ is auto-
morphic for all S L(n, Z). To proceed further, we need the following lemma.
Lemma 5.3.6 Let n > 2. Fix an integer M = 0, and let γ ∈ S L(n − 1, Z).
Then
γ 0
φ̂ Men−1 γ (z) = φ̂ (0,...,0,M) ·z ,
0 1
Proof Let
⎛ ⎞
a1,1 a1,2 ··· a1,n−1
⎜ .. .. .. ⎟
⎜ ⎟
γ =⎜ . . . ⎟,
⎝ an−2,1 an−2,2 · · · an−2,n−1 ⎠
γ1 γ2 ··· γn−1
and
⎛ ⎞ ⎛ ⎞
1 v1 1 v1
⎜ 1 v2⎟ ⎜ 1 v2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ .. ⎟
.. ⎜ .. .. ⎟
v=⎜ . ⎟,
. v = ⎜ . . ⎟.
⎜ ⎟ ⎜ ⎟
⎝ 1 vn−1 ⎠ ⎝ 1
vn−1 ⎠
1 1
5.3 Fourier expansions on SL(n, Z)\hn 121
For n > 2, the group S L(n − 1, Z) acts on Zn−1 with two orbits:
where Z+ denotes the positive integers. The second orbit above is a consequence
of the fact that every non-zero m ∈ Zn−1 can be expressed in the form
Remark 5.3.8 The above argument breaks down when n = 2. In this case,
the orbit consists of all integers M = 0.
It now follows from this discussion and Lemma 5.3.6 that we may rewrite
(5.3.5) as follows:
∞
φ(z) = φ̂ (0,...,0) (z) + φ̂ M·en−1 ·γ (z)
M=1 γ ∈Pn−1 (Z)\S L(n−1,Z)
∞
γ 0
= φ̂ (0,...,0) (z) + φ̂ (0,...,0,M) ·z .
M=1 γ ∈Pn−1 (Z)\S L(n−1,Z)
0 1
(5.3.9)
The fact that φ is a Maass form implies that φ̂ (0,...,0) (z) = 0. Replacing M by
m n−1 , and setting Pn−1 = Pn−1 (Z), S L n−1 = S L(n − 1, Z), we may, therefore,
rewrite (5.3.9) in the form
∞
1 1
γ
φ(z) = ··· φ u· ·z
m n−1 =1 γ ∈Pn−1 \S L n−1 0 0 1
where
⎛ ⎞
1 u 1,n
⎜ 1 ⎟ u 2,n
⎜ ⎟
n−1
⎜ .. ⎟ ..
u=⎜ . ⎟, . d ∗u = du j,n .
⎜ ⎟
⎝ 1 u n−1,n ⎠
j=1
where
⎛ ⎞
1 v1
⎜ 1 ⎟ v2
⎜ ⎟
n−1
⎜ .. ⎟ ..
v=⎜ . ⎟, . d ∗v = dvi ,
⎜ ⎟
⎝ 1 vn−1 ⎠
i=1
1
is invariant under left multiplication by matrices of the form
⎛ ⎞
m 1,1 m 1,2 · · · m 1,n−1 0
⎜ m · · · m 2,n−1 0⎟
⎜ 2,1 m 2,2 ⎟
⎜ . .. .. .. ⎟
⎜ . ⎟
m=⎜ ⎜
. . . . ⎟ ∈ S L(n, Z).
⎟
⎜ m n−2,1 m n−2,2 · · · m n−2,n−1 0 ⎟
⎜ ⎟
⎝ 0 0 ··· 1 0⎠
0 0 ··· 0 1
Proof We have
1 1
φ̂ (0,...,0,M) (m · z) = ··· φ(v · m · z)e−2πi Mvn−1 d ∗ v
0 0
1 1
= ··· φ(m · v · z)e−2πi Mvn−1 d ∗ v
0 0
1 1
= ··· φ(v · z)e−2πi Mvn−1 d ∗ v,
0 0
where
⎛ ⎞
1 v1
⎜ 1 v2 ⎟
⎜ ⎟
⎜ .. .. ⎟
v =⎜ . . ⎟
⎜ ⎟
⎝
1 vn−1 ⎠
1
is chosen so that v · m = m · v . A simple computation shows that we may take
v1 = m 1,1 v1 + m 1,2 v2 + · · · + m 1,n−1 vn−1
v2 = m 2,1 v1 + m 2,2 v2 + · · · + m 2,n−1 vn−1
..
.
vn−2 = m n−2,1 v1 + m n−2,2 v2 + · · · + m n−2,n−1 vn−1
vn−1 = vn−1 .
124 Maass forms and Whittaker functions for S L(n, Z)
We remark here that in the derivation of (5.3.10), we only used the left
invariance of φ(z) with respect to Pn−1 (Z). In view of Lemma 5.3.11, we may
then reiterate all previous arguments and obtain, instead of (5.3.10), the more
general form
∞ ∞ 1 1
φ(z) = ···
m n−2 =1 m n−1 =1 γn−2 ∈Pn−2 \S L n−2 γn−1 ∈Pn−1 \S L n−1 0 0
⎛ ⎛ ⎞ ⎞
γn−2
γn−1
× φ ⎝u · ⎝ 1 ⎠· · z⎠
1
1
× e−2πi [m n−2 u n−2,n−1 +m n−1 u n−1,n ] d ∗ u, (5.3.12)
where
⎛ ⎞ ⎛ ⎞
1 u 1,n 1 u 1,n−1 0
⎜ ⎟ ⎜ 0⎟
⎜ 1 ⎟ ⎜
u 2,n 1 u 2,n−1 ⎟
⎜ ⎟ ⎜
.. .. .. ⎟
⎜ .. ⎟ ⎜ .. ⎟
u=⎜
⎜
. ⎟·⎜
.
⎟ ⎜
. . .⎟
⎟
⎜ ⎟ ⎜ u n−2,n−1 0⎟
⎜ ⎟ ⎜ ⎟
⎝ 1 u n−1,n ⎠ ⎝ 1 0⎠
1 1
⎛ ⎞
1 u 1,n−1 u 1,n
⎜ ⎟
⎜ 1 u 2,n−1 u 2,n ⎟
⎜ .. .. ⎟
⎜ .. ⎟
=⎜
⎜
. . . ⎟,
⎟
⎜ u n−2,n−1 u n−2,n ⎟
⎜ ⎟
⎝ 1 u n−1,n ⎠
1
and
d ∗u = du i,n−1 · du j,n .
1≤i≤n−2 1≤ j≤n−1
5.3 Fourier expansions on SL(n, Z)\hn 125
Note that by Remark 5.3.8, the sum over m n−2 in formula (5.3.12) will range
over all m n−2 = 0 when n = 3.
The lemma follows after one notes that a set of coset representatives for the
quotient P̃ n,r +1 \ P̃ n,r is given by
(
γ
.
γ ∈Pn−r \S L n−r
Ir
We now apply Lemma 5.3.13 (in the form Pn−1−r \S L n−1−r with r = 1) to
the inner sum
in (5.3.12). We can replace γn−2 ∈ Pn−2 \S L n−2 by γn−2 ∈ P̃ n−1,2 \Pn−1 , where
⎛ ⎞
γn−2
⎝ ⎠= γn−2
1 ,
1
1
and where P̃ n−1,2 is defined as in Lemma 5.3.13. If we then apply Lemma 5.3.3
to the sums
,
γn−2 ∈ P̃ n−1,2 \Pn−1 γn−1 ∈Pn−1 \S L n−1
it follows that
∞
∞
1 1
γ
φ(z) = ··· φ u· ·z
m n−2 =1 m n−1 =1 γ ∈ P̃ n−1,2 \S L n−1 0 0 1
where
⎧⎛ ⎞⎫
⎪ ∗ ∗ ··· ∗ ∗ ⎪
⎪
⎪ ⎪
⎪⎜ .. ..
⎪ . . ⎟⎪ ⎪
⎨⎜ . . · · · .. .. ⎟⎪⎬
⎜ ⎟
P̃ n−1,2 = ⎜ ∗ ∗ · · · ∗ ∗ ⎟ ⊂ S L(n − 1, Z),
⎪⎜
⎪ ⎟⎪⎪
⎪
⎪ ⎝ 1 ∗ ⎠⎪ ⎪
⎪
⎩ ⎪
⎭
1
5.3 Fourier expansions on SL(n, Z)\hn 127
and
⎛ ⎞
1 u 1,n−1 u 1,n
⎜ 1 u 2,n−1 u 2,n ⎟
⎜ ⎟
⎜ .. .. .. ⎟
u=⎜
⎜
. . . ⎟,
⎟
⎜ u n−2,n−1 u n−2,n ⎟
⎝ 1 u ⎠
n−1,n
1
∗
d u= du i,n−1 · du j,n .
1≤i≤n−2 1≤ j≤n−1
All steps previously taken can be iterated. For example, after one more
iteration equation (5.3.14) becomes
∞ ∞ ∞ 1 1
γ
φ(z) = ··· φ u· ·z
m =1 m =1 m =1 0 0 1
n−3 n−2 n−1 γ ∈ P̃ n−1,3 \S L n−1
−2πi [m n−3 u n−3,n−2 +m n−2 u n−2,n−1 +m n−1 u n−1,n ]
×e d ∗ u, (5.3.15)
where
⎧⎛ ⎞⎫
⎪
⎪ ∗ ∗ ··· ∗ ∗ ⎪ ⎪
⎪
⎪⎜. ⎪
⎪
⎪⎜ .. .. .. .. ⎟ ⎪
⎪
⎪
⎨⎜
⎪ . · · · . . ⎟⎪⎟ ⎪
⎪
⎜ ⎟⎬
P̃ n−1,3 = ⎜ ∗ ∗ · · · ∗ ∗ ⎟ ⊂ S L(n − 1, Z),
⎪⎜ ⎟
⎪
⎪⎜ 1 ∗ ∗ ⎟⎪ ⎪
⎪
⎪⎜ ⎟⎪⎪
⎪
⎪
⎪
⎝ 1 ∗ ⎠⎪ ⎪
⎪
⎪
⎩ ⎭
1
and
⎛ ⎞
1 u 1,n−2 u 1,n−1 u 1,n
⎜ ⎟
⎜ 1 u 2,n−2 u 2,n−1 u 2,n⎟
⎜ .. .. .. ⎟
⎜ .. ⎟
⎜ . . . . ⎟
u=⎜
⎜ 1 u n−3,n−2 u n−3,n−1
⎟
u n−3,n ⎟ ,
⎜ ⎟
⎜ 1 u n−2,n−1 u n−2,n ⎟
⎜ ⎟
⎝ 1 u n−1,n ⎠
1
d ∗u = du i,n−2 · du j,n−1 · du k,n .
1≤i≤n−3 1≤ j≤n−2 1≤k≤n−1
Theorem 5.3.2 follows from (5.3.15) after continuing this process inductively
for n − 2 steps, and taking into account Remark 5.3.8.
128 Maass forms and Whittaker functions for S L(n, Z)
and
ij if i + j ≤ n,
bi, j =
(n − i)(n − j) if i + j ≥ n,
is an eigenfunction of every S L(n, R)–invariant differential operator in Dn .
These are not the only possible eigenfunctions, however. For example, we have
shown in Section 3.4 that the functions
√ √
y ν , y 1−ν , y K ν− 12 (2π|m|y)e2πimx , y Iν− 12 (2π|m|y)e2πimx ,
2
(with m ∈ Z, m = 0) are all eigenfunctions of = −y 2 ∂∂x 2 + ∂∂y 2 with
2
given in Proposition 5.2.4. Here φ(z) is a Maass form for S L(n, Z). This example
shows that Whittaker functions occur naturally in the Fourier expansion of
Maass forms. The importance of Whittaker functions cannot be underestimated.
They are the cornerstone for the entire theory of L–functions.
We shall show in the next section that it is always possible to explic-
itly construct one non–trivial Whittaker function. Remarkably, this special
Whittaker function has good growth properties and is the only Whittaker func-
tion that appears in the Fourier expansion of Maass forms (multiplicity one
theorem).
where
⎛ cν,m = 0 (depends
⎞ only on ν, m) and M=
|m 1 m 2 ···m n−1 |
⎜ . .. ⎟
⎝ |m 1 m 2 | ⎠.
|m 1 |
1
+
n−1
)
n−1 bi, j ν j −i(n−i)
Remark The reader may verify that cν,m = |m i | j=1 . For example
i=1
when the dimension n = 2, we have c2,ν = |m 1 |ν1 −1 , whereas for “n = 3” the
coefficient is c3,ν = |m 1 |ν1 +2ν2 −2 |m 2 |2ν1 +ν2 −2 .
Proof We shall defer the proof of the convergence and the meromorphic con-
tinuation of the integral until later. At this point, we show that WJacquet (z; ν, ψm )
satisfies Definition 5.4.1 (1) and (2) of a Whittaker function. First of all, note
5.5 Jacquet’s Whittaker function 131
that if
⎛ ⎞
1 a1,2 a1,3 ··· a1,n
⎜ 1 a2,3 ··· a2,n ⎟
⎜ ⎟
⎜ .. .. ⎟
a = ⎜ . . ⎟ ∈ Un (R),
⎜ ⎟
⎝ 1 an−1,n ⎠
1
Second, using the fact that every differential operator D ∈ Dn is invariant under
left multiplication by S L(n, R), it follows from the definition of λ D given in
Definition 5.4.1, that
Consequently,
D WJacquet (z; ν, ψm ) = D (Iν (wn · u · z)) ψm (u) d ∗ u
Un (R)
= λD Iν (wn · u · z) ψm (u) d ∗ u
Un (R)
= λ D · WJacquet (z; ν, ψm ).
We have thus proved Proposition 5.5.2 under the assumption that the integral
(5.5.1) converges absolutely and uniformly on compact subsets of hn to an L2
function on the Siegel set √3 , 1 .
2 2
Next, we prove the identity
WJacquet (z; ν, ψm ) = cν,m · WJacquet M z; ν, ψ1 ,2 ,...,n−1 ,
that
WJacquet M z; ν, ψ1 ,2 ,...,n−1 = Iν (wn · u · M z)e−2πi [1 u 1 +···+n−1 u n−1 ] d ∗ u
Un (R)
n−1
= |m i | Iν (wn · Mu · z)e−2πi [|m 1 |1 u 1 +···+|m n−1 |n−1 u n−1 ] d ∗ u
i=1 Un (R)
n−1
= |m i | Iν (wn Mwn · wn uz)e−2πi [m 1 u 1 +···+m n−1 u n−1 ] d ∗ u
i=1 Un (R)
Note that this identity holds because M y is a diagonal matrix with positive
entries. To prove it, consider (for j = 1, 2, . . . , n − 1) the (n − j + 1)th row:
of the matrix u. It is easy to see that for each 1 ≤ j ≤ n − 1, we can make the
transformation
u j → j u j ,
by letting
u → δ j · u · δ j ,
If we make the above transformations in the integral for the Whittaker func-
tion, then that integral takes the form:
WJacquet (M y; ν, ψ1 ,...,n−1 ) = Iν (wn · u · M y)e−2πi [1 u 1 +···+n−1 u n−1 ] d ∗ u
Un (R)
= Iν (wn · δ j uδ j · M y)e−2πi [1 u 1 +···+u j +···+n−1 u n−1 ] d ∗ u
Un (R)
= Iν (δ j wn · u · M yδ j )e−2πi [1 u 1 +···+u j +···+n−1 u n−1 ] d ∗ u
Un (R)
= Iν (wn · u · M y)e−2πi [1 u 1 +···+u j +···+n−1 u n−1 ] d ∗ u
Un (R)
∞
0 −1 1 u y x
|WJacquet (z; ν, ψm )| ≤ IRe(ν) · · du
1 0 0 1 0 1
−∞
∞ Re(ν)
y
= du
(x + u)2 + y 2
−∞
∞
du
=y 1−Re(ν)
, (5.5.3)
(u 2 + 1)Re(ν)
−∞
which converges absolutely for Re(ν) > 1/2. For n = 2, the meromorphic
continuation of (5.5.1) is obtained by direct computation of the integral
134 Maass forms and Whittaker functions for S L(n, Z)
1 u
WJacquet (z; ν, ψm ). With the choice ψm = e2πimu , we have
0 1
∞ ν
y
WJacquet (z; ν, ψm ) = e−2πimu du
−∞ (x + u) + y
2 2
∞ −2πimuy
e
=e 2πimx 1−ν
y ν
du
−∞ (u + 1)
2
2|m|ν− 2 π ν √
1
v1 ⊗ v2 ⊗ · · · ⊗ v ,
with vi ∈ Rn for i = 1, 2, . . . ,
.
We define
v1 ⊗ v2 ⊗ · · · ⊗ v
,
5.6 The exterior power of a vector space 135
One easily checks that the map (5.6.4) is well defined and shows that Λ
(Rn )
is isomorphic to a subspace of ⊗
(Rn ) generated by
ei1 ⊗ ei2 ⊗ · · · ⊗ ei ,
v · k, w · k = (v · k) · t (w · k) = v · k · t k · t w = v · t w = v, w,
for all v, w ∈ ⊗
(Rn ). Finally, the invariance of the action by k on the inner
product can be extended to Λ
(Rn ) by (5.6.5).
5.6 The exterior power of a vector space 137
is well known on the tensor product space ⊗
(Rn ). It extends to Λ
(Rn ) by the
identity (5.6.5).
To prove the second Cauchy–Schwartz type inequality consider
v= ai1 ,i2 ,...,i
ei1 ⊗ ei2 ⊗ · · · ⊗ ei
∈
(Rn ),
i 1 ,i 2 ,...,i
w= b j1 , j2 ,..., j
e j1 ⊗ e j2 ⊗ · · · ⊗ e j
∈
(Rn ).
j1 , j2 ,..., j
and
||v||2
||w||2
= |ai1 ,i2 ,...,i
|2 |b j1 , j2 ,..., j
|2 .
i 1 ,i 2 ,...,i
j1 , j2 ,..., j
Now
v∧w
= ai1 ,i2 ,...,i
b j1 , j2 ,..., j
ei1 ⊗ ei2 ⊗ · · · ⊗ ei
∧ e j1 ⊗ e j2 ⊗ · · · ⊗ e j
i 1 ,i 2 ,...,i
j1 , j2 ,..., j
1
= 2
ai1 ,i2 ,...,i
b j1 , j2 ,..., j
Sign(σ ) Sign(σ )
(
!) i1 ,i2 ,...,i
σ ∈S
σ ∈S
j1 , j2 ,..., j
× eσ (i1 ) ⊗ · · · ⊗ eiσ (
) ⊗ eσ ( j1 ) ⊗ · · · ⊗ eσ ( j
) .
The result now follows because ||u + u ||2⊗
≤ ||u||2⊗
+ ||u ||2⊗
for all
u, u ∈ ⊗
(Rn ) .
Lemma 5.7.2 For n ≥ 2, let z be given by (5.7.1) and let || || denote the norm
on
(Rn ) as in Lemma 5.6.6. Then for ν = (ν1 , . . . , νn−1 ) ∈ Cn−1 , we have
the identity
n−2
+
n−1
iν
Iν (z) := (en−i ∧ · · · ∧ en−1 ∧ en ) ◦ z −nνn−i−1 · |Det(z)| i=1 n−i .
i=0
5.7 Construction of the Iν function using wedge products 139
Furthermore,
+
n−1
+ i νn−i
n−1
i νn−i n−1 i=1
|Det(z)| i=1 = y
n−
,
=1
⎝ ⎠ ⎝ ⎠
1 1
140 Maass forms and Whittaker functions for S L(n, Z)
We compute
⎛ ⎞⎛ ⎞⎛ ⎞
−1 1 u2 u3 y1 y2
e3 w3 uy = (0, 0, 1) ⎝ 1 ⎠⎝ 1 u1 ⎠ ⎝ y1 ⎠
1 1 1
⎛ ⎞
y1 y2 u 2 y1 u3
= (1, 0, 0) ⎝ y1 u1 ⎠
1
= y1 y2 e1 + u 2 y1 e2 + u 3 e3 .
Hence
3 41
||e3 w3 uy|| = y12 y22 + u 22 y12 + u 23 2 .
Similarly,
⎛ ⎞⎛ ⎞⎛ ⎞
−1 1 u2 u3 y1 y2
e2 w3 uy = (0, 1, 0) ⎝ 1 ⎠⎝ 1 u1 ⎠ ⎝ y1 ⎠
1 1 1
⎛ ⎞
y1 y2 u 2 y1 u3
= (0, 1, 0) ⎝ y1 u1 ⎠
1
= y1 e2 + u 1 e3 ,
so that
Consequently
3 41
||(e2 w3 uy) ∧ (e3 w3 uy)|| = y1 y12 y22 + (u 1 u 2 − u 3 )2 + u 21 y22 2 .
The result follows upon substituting the results of the above computations into
(5.7.4).
the identity
||en · wn uy|| = ||e1 · uy||
3 41
= (y1 · · · yn−1 )2 + (u 1,2 y1 · · · yn−2 )2 + · · · + (u 1,n−1 y1 )2 + (u 1,n )2 2 ,
it immediately follows from (5.8.2) that (5.8.1) converges absolutely for
Re(νi ) > 1/n (1 ≤ i ≤ n − 1) if the following two integrals converge
absolutely:
∞ ∞
3 4−nV /2
n
··· (y1 · · · yn−1 )2 + (u 1,2 y1 · · · yn−2 )2 + · · · + (u 1,n )2 du 1,k ,
k=1
−∞ −∞
(5.8.3)
∞ ∞
n−2
−nνn−i−1
··· (en−i ∧ · · · ∧ en−1 ) ◦ wn uy du i, j .
i=1 1<i< j≤n
−∞ −∞
(5.8.4)
Clearly, the first integral converges absolutely if Re(νi ) > 1/n, for
1 ≤ i ≤ n − 1. Furthermore, if 1 < i < n, then ei · wn = e
for some
= 1.
This immediately implies that
ei · wn uy = ei · wn u y
where
⎛ ⎞ ⎛ ⎞
1 0 0 ··· 0 1
⎜ 1 u 2,3 ··· u 2,n ⎟ ⎜ y1 y2 · · · yn−2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ .. .. ⎟ ⎜ .. ⎟
u = ⎜ . . ⎟, y = ⎜ . ⎟.
⎜ ⎟ ⎜ ⎟
⎝ 1 u n−1,n ⎠ ⎝ y1 ⎠
1 1
It follows that the second integral (5.8.4) may be rewritten as
∞ ∞
n−2
··· (en−i ∧ · · · ∧ en−1 ) ◦ wn u y n−i−1
−nν
du i, j .
i=1 1<i< j≤n
−∞ −∞
(5.8.5)
The remarkable thing is that the integral in (5.8.5) can be interpreted as a Jacquet
Whittaker function for S L(n − 1, Z). This allows us to apply induction from
which the absolute convergence of (5.8.1) follows.
Recall the definition of wn given just after (5.5.1). To see that the integral
(5.8.5) is a Jacquet Whittaker function for S L(n − 1, Z), we use the matrix
5.8 Convergence of Jacquet’s Whittaker function 143
identity:
⎛ ⎞ ⎛ ⎞
0 1 1 0
⎜ 0 1 ⎟ ⎜ !n/2"⎟
0 (−1) ⎟
⎜ ⎟ ⎜
⎜ .. .. ⎟ ⎜ . . ⎟
wn = ⎜ . . ⎟·⎜ .. .. ⎟.
⎜ ⎟ ⎜ ⎟
⎝ 0 1⎠ ⎝ 0 1 ⎠
1 0 0 1
Further,
⎛ ⎞
0 1
⎜ 0 1 ⎟
⎜ ⎟
⎜ .. .. ⎟
e
· ⎜ . . ⎟ = e
+1
⎜ ⎟
⎝ 0 1⎠
1 0
for 1 ≤
≤ n − 1. It immediately follows from these remarks that (5.8.5) may
be rewritten as
∞ ∞
n−2
··· −nνn−i−1
(en+1−i ∧ · · · ∧ en ) ◦ w u y du i, j ,
n
i=1 1<i< j≤n
−∞ −∞
(5.8.6)
where
⎛ ⎞
1 0
⎜ 0 (−1)!n/2" ⎟
⎜ ⎟
⎜ . . ⎟ 1
wn =⎜ .. .. ⎟= .
⎜ ⎟ wn−1
⎝ 0 1 ⎠
0 1
Furthermore, we may write
1 1
u = , y = ,
µ η
with µ ∈ Un−1 (R) and η a diagonal matrix in G L(n − 1, R). With these obser-
vations, one may deduce that
n−2
(en+1−i ∧ · · · ∧ en ) ◦ w u y −nνn−i−1
n
i=1
n/(n−1)
(n−1)−2
= (en−1−i ∧ · · · ∧ en−1 ) ◦ wn−1 µη−(n−1)νn−2−i .
i=0
144 Maass forms and Whittaker functions for S L(n, Z)
It follows that
∞ ∞
n−2
−nνn−i−1
··· (en+1−i ∧ · · · ∧ en ) ◦ w u y
du i, j
n
i=1 1<i< j≤n
−∞ −∞
∞ ∞
··· I(nν/(n−1)),n−1 (wn−1 µη) dµ,
y
−∞ −∞
where Iν,n−1 denotes the Iν function (as defined in Section 5.4) for
G L(n − 1, R). By induction, we obtain the absolute convergence in the region
Re(ν) > 1/n.
where
j−1
nνn−k+i − 1
v j,k = ,
i=0
2
and WJacquet (z; ν, ψ) denotes Jacquet’s Whittaker function (5.5.1).
5.9 Functional equations of Jacquet’s Whittaker function 145
+
n−1
2πi u
,
+1
ψ(u) = e
=1 ,
∗
for u ∈ Un (R). Then the Whittaker function WJacquet (z; ν, ψ) has a holomorphic
continuation to all ν ∈ C . For each w ∈ Wn , let ν, ν satisfy (5.9.1). Then
n−1
∗ ∗
WJacquet (z; ν, ψ) = WJacquet (z; ν , ψ).
∗ ∗
Remarks It is clear that WJacquet (z; ν, ψ) and WJacquet (z; ν , ψ) are both
Whittaker functions of type ν and character ψ. It follows from Shalika’s multi-
plicity one theorem (Shalika, 1974) that the functional equation must hold up to
a constant depending on ν. The assumption that ψ(u) = ψ1,1,...,1 is not restric-
tive because Proposition 5.5.2 tells us that there is a simple identity relating
∗ ∗
WJacquet (z; ν, ψm ) and WJacquet (M z; ν, ψ1,1,...,1 ).
Before giving the proof of the functional equation, we will obtain explicit
versions of the functional equation ν
→ ν given by (5.9.1). The Weyl group
Wn is generated by the simple reflections
⎛ ⎞
In−i−1
⎜ 0 −1 ⎟
σi = ⎜
⎝
⎟,
⎠ (i = 1, 2, . . . , n − 1),
1 0
Ii−1
where Ia denotes the a × a identity matrix. We adopt the convention that I0 is the
⎛ ⎞ ⎛ ⎞
In−2 0 −1
empty set so that σ1 = ⎝ 0 −1 ⎠ and σn−1 = ⎝ 1 0 ⎠.
1 0 In−2
Since σi (i = 1, . . . , n − 1) generate Wn , it is enough to give the functional
equation ν
→ ν for the simple reflections w = σi . Fix an integer i with
1 ≤ i ≤ n − 1. In this case, ν is defined by the equation
Iν− n1 (y) = Iν − n1 (σi y) = Iν − n1 σi yσi−1 = Iν − n1 (y ) (5.9.4)
146 Maass forms and Whittaker functions for S L(n, Z)
where
⎛ ⎞
y1 · · · yn−1
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
⎜ y1 · · · yi ⎟
⎜ ⎟
⎜ ⎟
y = ⎜ y1 · · · yi yi+1 ⎟
⎜ ⎟
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
⎝ y1 ⎠
1
is the diagonal matrix y with the (n − i)th and (n − i + 1)th rows interchanged.
It follows that if we put y in Iwasawa form
⎛ ⎞
y1 · · · yn−1
⎜ .. ⎟
⎜ . ⎟
y = ⎜ ⎟,
⎝ y1 ⎠
1
⎧
⎪
⎪ y
if
= i − 1, i, i + 1,
⎪
⎪
⎨y y if
= i − 1,
+1
y
=
⎪ y
−1
⎪ if
= i,
⎪
⎪
⎩
y
−1 y
if
= i + 1,
n−1
n−1
1 1
b
, j νj − = b
, j ν j − , (for
= i),
j=1
n j=1
n
(5.9.5)
n−1
n−1
1 1
bi, j νj − = (bi−1, j − bi, j + bi+1, j ) · ν j − .
j=1
n j=1
n
5.9 Functional equations of Jacquet’s Whittaker function 147
1
νn−i−1 =− + νn−i−1 + νn−i , (if i = n − 1),
n
2
νn−i = − νn−i ,
n
(5.9.6)
1
νn−i+1 =− + νn−i + νn−i+1 , (if i = 1),
n
ν
= ν
(
= n − i − 1, n − i, n − i + 1).
Example 5.9.7 For n = 2, 3 the functional equations (5.9.6) take the explicit
form:
ν1 = 1 − ν1 , (n = 2),
2 1
ν2 = − ν2 , ν1 = − + ν1 + ν2 , (n = 3, i = 1),
3 3
2 1
ν1 = − ν1 , ν2 = − + ν1 + ν2 , (n = 3, i = 2).
3 3
∗ ∗
WJacquet (z; ν, ψ) = WJacquet (z; ν , ψ),
nνn−i
π − 2 νn−i
n
2
WJacquet (z; ν, ψ) = − n ν · nνn−i · WJacquet (z; ν , ψ)
π 2 n−i 2
−(1− n2 νn−i ) 1 − nνn−i
π
= · 2
· WJacquet (z; ν , ψ).
π − 2 νn−i
n
nν2n−i
148 Maass forms and Whittaker functions for S L(n, Z)
wi := σi−1 wn .
WJacquet (z; ν, ψ) = Iν (wn · u · z) ψ(u) d ∗ u
U (R)
0 1
n
The inner integral above (over the region Ni ) is a Whittaker function for the
group S L(2, R), and has a functional equation of type (5.5.5). This functional
equation is independent of the choice of wi n i z, and is precisely what is needed
to complete the proof of Theorem 5.9.8. We shall find the functional equation
by examining the case when wi n i z is the identity matrix.
We first compute Iν (σi n i ) for
⎛ ⎞ ⎛ ⎞
In−i−1 In−i−1
⎜ 0 −1 ⎟ ⎜ 1 u ⎟
σi = ⎜
⎝
⎟,
⎠ ni = ⎜
⎝
⎟.
⎠
1 0 0 1
Ii−1 , Ii−1 ,
150 Maass forms and Whittaker functions for S L(n, Z)
It follows that
⎛ ⎞
In−i−1
⎜ ⎟
(u 2 + 1)− 2
1 1
⎜ −u/(u 2 + 1) 2 ⎟
σi n i = ⎜ 1 ⎟·k
⎝ 0 (u 2 + 1) 2 ⎠
Ii−1 ,
for some orthogonal matrix k. If we then put σi n i in standard Iwasawa form,
we have
⎛ ⎞ ⎛a ···a ⎞
In−i−1 1 n−1
⎜ 1 −u/(u 2 + 1) ⎟ ⎜ .. ⎟
σi n i = ⎜ ⎟·⎜⎜ . ⎟
⎟
⎝ 0 1 ⎠ ⎝ ⎠
a1
Ii−1 , 1
with
⎧
⎪
⎪ 1 if
= i − 1, i, i + 1
⎪
⎪
⎨(u 2 + 1) 12 if
= i − 1
a
= .
⎪
⎪(u 2 + 1)−1 if
= i
⎪
⎪
⎩ 2 1
(u + 1) 2 if
= i + 1.
Consequently
+
n−1
(bi−1, j −2bi, j +bi+1, j ) ν2j
Iν (σi n i ) = (u + 1)
2 j=1
.
The functional equation now follows from (5.5.5).
0 0 1 0
5.10 Degenerate Whittaker functions 151
Let
⎛ ⎞ ⎛ ⎞
1 u 1,2 u 1,3 u 1,4 y1 y2 y3
⎜0 1 u 2,3 u 2,4 ⎟ ⎜ y1 y2 ⎟
u=⎜
⎝0
⎟, y=⎜ ⎟.
0 1 u 3,4 ⎠ ⎝ y1 ⎠
0 0 0 1 1
With a brute force computation, one sees that for ν = (ν1 , ν2 , ν3 ) ∈ C3 we have
ν1 +2ν2 +ν3
y22 u 23,4 + y12
Iν (wuy) = 2
(u 2,4 − u 2,3 u 3,4 )2 + u 23,4 + y12 y22
(ν1 +2ν2 +3ν3 )/2
y12 (u 2,4 − u 2,3 u 3,4 )2 + u 23,4 + y12 y22
× 2 2
u 3,4 + y12
(3ν +2ν +ν )/2
× (u 2,4 − u 2,3 u 3,4 )2 + u 23,4 + y12 y22 y3 1 2 3 .
Clearly, the function does not involve the variables u 1,2 , u 1,3 , u 1,4 , so it is not
possible to integrate the function Iν (wuy) over the entire u space. We may only
consider some type of partial integral which does not involve all the u-variables.
We leave it to the reader to work out a general theory of degenerate Whittaker
functions and only briefly indicate how to define these objects. Let U = Un (R)
denote the group of upper triangular n × n matrices with real coefficients and
1s on the diagonal (upper triangular unipotent matrices). For each element w
in the Weyl group of S L(n, R) define
Uw := (w −1 · U · w) ∩ U, Ū w = (w−1 ·t U · w) ∩ U.
For example, if
⎛ ⎞
1 0 0 0
⎜0 0 0 1⎟
w=⎜
⎝0
⎟
1 0 0⎠
0 0 1 0
then
⎛ ⎞ ⎛ ⎞
1 ∗ ∗ ∗ 1
⎜ 1 ∗ ⎟ ⎜ 1 ∗⎟
Uw = ⎜
⎝
⎟,
⎠ Ū w = ⎜
⎝
⎟.
1 1 ∗⎠
1 1
⎛ ⎞
⎜
⎜
1 ⎟
⎟
⎜ ⎟
⎜ 1 u 2,4 ⎟⎟⎟ ∗
form u = ⎜
⎜
⎜
⎜
⎟, with u 2,4 , u 3,4 ∈ R. The natural measure, du , on
⎜
⎜
1 u 3,4 ⎟⎟⎟
⎝ ⎠
1
the space Ū w is du ∗ = du 2,4 du 3,4 .
Definition 5.10.1 For n ≥ 2, let z ∈ hn , ν = (ν1 , ν2 , . . . , νn−1 ) ∈ Cn−1 , and
let ψm be a character as in (5.5.1). Then the degenerate Whittaker function
associated to w is defined to be
Iν (wuz)ψm (u) d ∗ u,
Ū w
∗
where d u is the natural measure on Ū w .
Remark It may be shown that the degenerate Whittaker function can be
meromorphically continued and satisfies the same group of functional equations
as Jacquet’s Whittaker function as given in Theorem 5.9.8.
GL(n)pack functions The following GL(n)pack functions, described in the
appendix, relate to the material in this chapter:
FunctionalEquation IFun ModularGenerators
Wedge d Whittaker
WhittakerGamma WMatrix.
6
Automorphic forms and L-functions for
S L(3, Z)
(6.1.1)
∂2 2 ∂
2
∂2
+ y12 + y + 2y 2
x 1,2 ,
∂ x2,3
2 2
∂ x1,2
2 1
∂ x2,3 ∂ x1,3
∂3 ∂3 ∂3 ∂3
2 = −y12 y2 + y1 y2
2
− y 3 2
y
1 2 + y1 y2
2
∂ y12 ∂ y2 ∂ y1 ∂ y22 ∂ x1,3
2
∂ y1 ∂ x1,2
2
∂ y1
∂ 3 ∂3
− 2y12 y2 x1,2 + −x1,2 2
+ y22 y12 y2 2
∂ x2,3 ∂ x1,3 ∂ y2 ∂ x1,3 ∂ y2
∂3 ∂3 ∂3
− y12 y2 + 2y 2 2
y + 2y 2 2
y x 1,2
∂ x2,3
2
∂ y2 1 2
∂ x2,3 ∂ x1,2 ∂ x1,3 1 2
∂ x1,2 ∂ x1,3
2
∂2 ∂2 ∂2 2 ∂2
+ y12 2 − y22 2 + 2y12 x1,2 + x1,2 + y22 y12 2
∂ y1 ∂ y2 ∂ x2,3 ∂ x1,3 ∂ x1,3
∂2 2 ∂
2
+ y12 − y2 .
∂ x2,3
2
∂ x1,2
2
153
154 Automorphic forms and L-functions for S L(3, Z)
where
⎛ ⎞ ⎛ ⎞
1 1 u 1,2 u 1,3
w3 = ⎝ −1 ⎠, u=⎝ 1 u 2,3 ⎠ ,
1 1
and
It was shown in Section 5.8, 5.9 that WJacquet (z, ν, ψ1,1 ) has meromorphic con-
tinuation to all ν1 , ν2 ∈ C and satisfies the functional equations:
∗ ∗
WJacquet (z, (ν1 , ν2 ), ψ1,1 ) = WJacquet z, ν1 + ν2 − 13 , 23 − ν2 , ψ1,1
∗
= WJacquet z, 23 − ν1 , ν1 + ν2 − 13 , , ψ1,1 ,
where
∗ 3ν1 3ν2
2 −3ν1 −3ν2
1
WJacquet (z, (ν1 , ν2 ), ψ1,1 ) =π
2 2
3ν1 + 3ν2 − 1
× WJacquet (z, ν, ψ1,1 ).
2
Vinogradov and Takhtadzhyan (1982) and Stade (1990) have obtained the
following very explicit integral representation
∗
WJacquet (y, (ν1 , ν2 ), ψ1,1 )
∞ &
1+(ν −ν )/2 1−(ν −ν )/2
= 4y1 1 2 y2 1 2 K 3ν1 +3ν2 −2 2π y2 1 + u −2
2
0
& 3ν1 −3ν2 du
× K 3ν1 +3ν2 −2 2π y1 1 + u 2 u 2 . (6.1.3)
2 u
Using the above representation, or alternatively following Bump (1984), one
6.1 Whittaker functions and multiplicity one for SL(3, Z) 155
where
s1 +α s1 +β s1 +γ s2 −α s2 −β s2 −γ
2 2 2 2 2
G(s1 , s2 ) = 2
, (6.1.5)
s1 +s
2
2
and
Proof It is enough to prove the theorem for the case when ψ(x) = eπi (x2,3 +x1,2 )
since, in general, ν (z) = ν (y)ψ(x) for z = x y ∈ h3 . For s = (s1 , s2 ) ∈ C2 ,
consider the double Mellin transform
∞ ∞
dy1 dy2
˜ ν (s) =
y1s1 y2s2 ν (y)
y1 y2
0 0
which is well defined for (s1 ), (s2 ) sufficiently large by the assumption in
our theorem. Define
for z = x y ∈ h3 .
Define the inner product, , , on L2 (U (Z)\h3 ) by
∞ ∞ 1 1 1
dy1 dy2
f, g = f (z)g(z) d x1,2 d x1,3 d x2,3 ,
(y1 y2 )3
0 0 0 0 0
for all f, g ∈ L2 U (Z)\h3 . Taking f (z) = ν (z) and g(z) = Js̄ (z), it follows
that
∞ ∞
dy1 dy2
ν , Js̄ = ν (y) · y1s1 −2 y2s2 −2 ˜ ν (s ∗ )
=
y1 y2
0 0
for some λν (D) ∈ R. Since D is a self-adjoint operator with respect to the above
inner product, it follows that
where the sum ranges over the finite set S = {(0, 0), (0, 2), (2, 0)}, and
c(s, (0, 0), i ) = λs (i ), (i = 1, 2),
c(s, (2, 0), 1 ) = c(s, (0, 2), 1 ) = −(2πi)2 ,
c(s, (2, 0), 2 ) = (2πi)2 (1 − s2 ),
c(s, (0, 2), 2 ) = (2πi)2 (s1 − 1),
where s = 2s23−s1 , 2s13−s2 .
The proof of Lemma 6.1.9, first obtained by Friedberg and Goldfield
(1993), is given by a simple brute force computation which we omit. Note
that the map s
→ s denotes the linear transformation of C2 such that
y1s1 y2s2 = Is (z) with Iw (z) = y1w1 +2w2 y22w1 +w2 for all w = (w1 , w2 ) ∈ C2 .
for s1 = σ1 + it1 , s2 = σ2 + it2 and with |t1 | → ∞ and s2 fixed. We also have
a similar estimate for s1 fixed and |t2 | → ∞.
Let us define the quotient function
˜ ν (s)
Fν (s) := .
W̃Jacquet (s, ν)
If we fix s2 and let (s1 ), (s2 ) be sufficiently large, then (6.1.12) implies that
π
Fν ((s1 , s2 )) e 2 |t1 | (6.1.13)
as t2 → ∞.
Note that the shift equations imply that
Since Fν (s) is holomorphic for (s1 ), (s2 ) sufficiently large, it follows that
Fν (s) is entire. If we fix s2 and consider Fν ((s1 , s2 )) as a function of s1 , it is an
immediate consequence of (6.1.15) that Fν ((s1 , s2 )) will be periodic (of period
2) in s1 . Thus, for fixed s2 , Fν ((s1 , s2 )) will be a function of z 1 = eπis1 and will
have a Laurent expansion in the variable z 1 of the form
∞
Fν ((s1 , s2 )) = cn (s2 )z 1n ,
n=−∞
for |t1 | → ∞ and s2 fixed. This implies that ck (s2 ) = 0 for k = ±1, ±2, . . .
Similarly,
1
|Fν ((s1 , s2 ))|2 ds2 ≥ |a0, j |2 e−2π jt2
(s2 )=0
for |t2 | → ∞ and s1 fixed. It follows that a0, j = 0 for j = ±1, ±2, . . . Thus
Fν ((s1 , s2 )) must be a constant. This completes the proof of Theorem 6.1.6.
with
⎛ ⎞ ⎛ ⎞
1 u 1,2 u 1,3 1 u2 u 1,3
u=⎝ 1 u 2,3 ⎠ = ⎝ 1 u 1 ⎠ ∈ U3 (R)
1 1
and d ∗ u = du 1 du 2 du 1,3 . Note that we have relabeled the super diagonal
elements u 1 = u 2,3 , u 2 = u 1,2 as in Proposition 5.5.2.
Now, we have shown that φ̃ (m 1 ,m 2 ) (z) is a Whittaker function. Further, φ̃ (m 1 ,m 2 )
will inherit the growth properties of the Maass form φ and will satisfy the
conditions of Theorem 6.1.6. The multiplicity one Theorem 6.1.6 tells us that
only the Jacquet Whittaker function (6.1.2) can occur in the Fourier expansion
of a Maass form for S L(3, Z), and that φ̃ (m 1 ,m 2 ) must be a constant multiple of the
160 Automorphic forms and L-functions for S L(3, Z)
Jacquet Whittaker function. It follows from Theorem 5.3.2 and Proposition 5.5.2
that if φ is a Maass form of type ν = (ν1 , ν2 ) ∈ C2 for S L(3, Z) then
∞
A(m 1 , m 2 )
φ(z) =
γ ∈U2 (Z)\S L(2,Z) m 1 =1 m 2 =0
|m 1 m 2 |
⎛⎛ ⎞ ⎞
|m 1 m 2 |
×WJacquet ⎝⎝ m1 ⎠ γ z, ν, ψ1, |mm2 | ⎠ , (6.2.1)
1 2
1
⎛⎛ ⎞⎞
1 u2 u 1,3
where A(m 1 , m 2 ) ∈ C and ψ1 ,2 ⎝⎝ 1 u 1 ⎠⎠ = e2πi(1 u 1 +2 u 2 ) .
1
The particular normalization A(m 1 , m 2 )/|m 1 m 2 | is chosen so that later
formulae are as simple as possible.
Lemma 6.2.2 (Fourier coefficients are bounded) Let φ be a Maass form for
S L(3, Z) as in (6.2.1). Then for all integers m 1 ≥ 1, m 2 = 0,
A(m 1 , m 2 )
= O(1).
|m 1 m 2 |
⎛ ⎞⎛ ⎞
1 x2 x1,3 y1 y2
Proof Let z = ⎝ 1 x1 ⎠ ⎝ y1 ⎠ . A simple computation
1 1
shows that
⎛⎛ ⎞ ⎞
|m 1 m 2 |y1 y2
A(m 1 , m 2 )
· WJacquet ⎝⎝ m 1 y1 ⎠, ν, ψ1, m2 ⎠
|m 1 m 2 | |m | 2
1
1 1 1
= φ(z) e−2πi [m 1 x1 +m 2 x2 ] d x1 d x2 d x1,3 .
0 0 0
A(m 1 , m 2 ) = O |m 1 m 2 |) .
6.3 The dual and symmetric Maass forms 161
to pick off the (m 1 , m 2 )th Fourier coefficient, then because x1 and x2 are
interchanged we will actually get A(m 2 , m 1 ).
In the S L(2, Z) theory, the notions of even and odd Maass forms
(see Section 3.9) played an important role. If a(n) is the nth Fourier coeffi-
cient of an S L(2, Z) Maass form then a(n) = ±a(−n) depending on whether
the Maass form is even or odd. We shall see that there is a quite different situa-
tion in the case of S L(3, Z) and that there are no odd Maass forms in this case.
The cognoscenti will recognize that there are no odd Maass forms on S L(3, Z)
because our definition of Maass form requires a trivial central character.
Consider a diagonal matrix δ of the form
⎛ ⎞
δ 1 δ2
δ := ⎝ δ1 ⎠
1
where δ1 , δ2 ∈ {+1, −1}. We define an operator Tδ which maps Maass forms
to Maass forms, and is given by
Tδ φ(z) := φ(δzδ).
Note that
⎛⎛ ⎞ ⎛ ⎞⎞
1 x2 x1,3 y1 y2
Tδ φ ⎝⎝ 1 x1 ⎠ · ⎝ y1 ⎠⎠
1 1
⎛⎛ ⎞ ⎛ ⎞⎞
1 x2 δ2 x1,3 δ1 δ2 y1 y2
= φ ⎝⎝ 1 x 1 δ1 ⎠ · ⎝ y1 ⎠⎠ . (6.3.3)
1 1
Clearly (Tδ )2 is the identity transformation, so the eigenvalues of Tδ can only
be ±1.
Definition 6.3.4 A Maass form φ of type ν = (ν1 , ν2 ) ∈ C2 for S L(3, Z) is
said to be symmetric if Tδ φ = ±φ for all Tδ as in (6.3.3).
We shall now show that every Maass form φ for S L(3, Z) is even, i.e.,
⎛ ⎞
δ1 δ2
Tδ φ = φ, for all Tδ = ⎝ δ1 ⎠ with δ1 , δ2 ∈ {+1, −1}. The reason is
1
6.4 Hecke operators for SL(3, Z) 163
⎛ ⎞
−1
Proof Let Tδ = ⎝ 1 ⎠. Then since δzδ transforms x2 → −x2 and
1
x1,3 → −x1,3 it easily follows that
1 1 1
Tδ φ(z)e−2πim 1 x1 e−2πim 2 x2 d x1 d x2 d x1,3
0 0 0
picks off the A(m 1 , −m 2 ) coefficient of φ(z); and this equals A(m 1 , m 2 ) because
Tδ φ(z) = φ(z).
where f ∈ L2 (\X ), x ∈ X, and αi are given by (6.4.1). The Hecke ring con-
sists of all formal sums
ck Tgk
k
g → t g, g ∈ ,
where t g denotes the transpose of the matrix g. It is again clear that the conditions
of Theorem 3.10.10 are satisfied so that the Hecke ring is commutative.
The following lemma is analogous to Lemma 3.12.1, which came up in the
S L(2, Z) situation.
6.4 Hecke operators for SL(3, Z) 165
Proof First of all we claim the decomposition is disjoint. If not, there exists
⎛ ⎞
γ1,1 γ1,2 γ1,3
⎝ γ2,1 γ2,2 γ2,3 ⎠ ∈
γ3,1 γ3,2 γ3,3
such that
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
γ1,1 γ1,2 γ1,3 a b1 c1 a b1 c1
⎝ γ2,1 γ2,2 γ2,3 ⎠ · ⎝ 0 b c2 ⎠ = ⎝ 0 b c2 ⎠ . (6.4.4)
γ3,1 γ3,2 γ3,3 0 0 c 0 0 c
Since 0 ≤ b1 , b1 < b and 0 ≤ c1 , c2 , c1 , c2 < c, one concludes that
γ1,2 = γ1,3 = γ2,3 = 0, and the decomposition is disjoint as claimed.
Now, by Theorem 3.11.2, every element on the right-hand side of (6.4.3) can
be put into Smith normal form, so must occur as an element on the left-hand
side of (6.4.3). Similarly, by Theorem 3.11.1, every element on the left-hand
side of (6.4.3) can be put into Hermite normal form, so must occur as an element
on the right-hand side of (6.4.3). This proves the equality of the two sides of
(6.4.3).
By analogy with the S L(2, Z) situation (see (3.12.3)), it follows that for
every integer n ≥ 1, we have a Hecke operator Tn acting on the space of square
166 Automorphic forms and L-functions for S L(3, Z)
Note the normalizing factor of 1/n which was chosen to simplify later formulae.
Clearly, T1 is just the identity
operator.
The C-vector space L2 \h3 has a natural inner product, denoted , , and
defined by
f, g = f (z)g(z) d ∗ z,
\h3
⎛ ⎞ ⎛ ⎞
1 x1,2 x1,3 y1 y2 0 0
for all f, g ∈ L2 \h3 , z = ⎝ 0 1 x2,3 ⎠ · ⎝ 0 y1 0 ⎠ ∈ h3 ,
0 0 1 0 0 1
and where
dy1 dy2
d ∗ z = d x1,2 d x1,3 d x2,3
(y1 y2 )3
denotes the left invariant measure given in Proposition 1.5.3.
In the case of S L(2, Z), we showed in Theorem 3.12.4 that the Hecke oper-
ators are self-adjoint with respect to the Petersson inner product. For S L(n, Z)
with n ≥ 3, it is no longer true that the Hecke operators are self-adjoint. What
happens is that the adjoint operator is again a Hecke operator and, therefore,
the Hecke operator commutes with its adjoint, which means that it is a normal
operator.
Theorem 6.4.6 (Hecke operators are normal operators) Consider the
Hecke operators Tn , (n = 1, 2, . . . ) defined in (6.4.5). Let Tn∗ be the adjoint
operator which satisfies
Tn f, g = f, Tn∗ g
for all f, g ∈ L2 (\h3 ). Then Tn∗ is another Hecke operator which commutes
with Tn so that Tn is a normal operator. Explicitly, Tn∗ is associated to the
following union of double cosets:
⎛ 2 2 ⎞
( m0m1m2
⎝ m 20 m 1 m 2 ⎠ . (6.4.7)
m 30 m 21 m 2 =n 2
m0m1
6.4 Hecke operators for SL(3, Z) 167
Proof It follows from (6.4.3), and also from the fact that transposition is an
antiautomorphism (as in the proof of Theorem 3.10.10), that
⎛ ⎞
( m0m1m2 ( (
⎝ m0m1 ⎠ = α = α. (6.4.8)
m 30 m 21 m 2 =n m0 α∈Sn α∈Sn
Since the action of the Hecke operator is independent of the choice of right
coset decomposition, we obtain
1
Tn f, g = f (αz) g(z) d ∗ z
n α∈Sn
\h3
1
= f (z) g(α −1 z) d ∗ z
n α∈Sn
\h3
⎛⎛ ⎞ ⎞
n
1
= f (z) g ⎝⎝ n ⎠α −1 z ⎠ d ∗ z, (6.4.9)
n α∈Sn
\h3 n
⎛ ⎞−1
( ( m0m1m2
α −1 = ·ω⎝ m0m1 ⎠ ω−1 ·
α∈Sn m 30 m 21 m 2 =n m0
⎛ ⎞
( m −1
0
= ·⎝ (m 0 m 1 )−1 ⎠ · .
m 30 m 21 m 2 =n −1
(m 0 m 1 m 2 )
(6.4.10)
168 Automorphic forms and L-functions for S L(3, Z)
Tn f = A(n, 1) · f, ∀ n = 1, 2, . . .
1 1 1
f (z)e−2πi(m 1 x1 +m 2 x2 ) d x1,3 d x1 d x2
0 0 0
⎛⎛ ⎞ ⎞
|m 1 m 2 |y1 y2
A(m 1 , m 2 )
= · WJacquet ⎝⎝ m 1 y1 ⎠, ν, ψ1, m2 ⎠ .
|m 1 m 2 | |m | 2
1
(6.4.12)
Next, if we let
⎛ ⎞⎛ ⎞ ⎛ ⎞⎛ ⎞
a b1 c1 1 x2 x1,3 1 α2 α1,3 a
⎝ 0 b c2 ⎠ ⎝ 0 1 x1 ⎠ = ⎝ 1 α1 ⎠ ⎝ b ⎠,
0 0 c 0 0 1 1 c
It follows that the right-hand side of (6.4.13) can be expressed in the form
n n n
1
n 4 abc=n 0≤c1 ,c2 <c
0≤b1 <b 0 0 0
⎛⎛ ⎞ ⎞
ax2 + b1 ax1,3 + b1 x1 + c1
⎛ ⎞
⎜⎜1 b c ⎟ ay1 y2 ⎟
⎜⎜ ⎟ ⎟
× f⎜⎜
⎜⎜0 1
bx1 + c2 ⎟·⎝
⎟ by1 ⎠⎟
⎟
⎝⎝ c ⎠ c ⎠
0 0 1
× e−2πi(m 1 x1 +m 2 x2 ) d x1,3 d x2 d x1 ,
In view of the fact that the integrand above is periodic and does not change
under transformations of the form
But the above integral vanishes unless b|m 1 c and a|m 2 b. Furthermore, in the
case that b|m 1 c and a|m 2 b, we have
m c m b
2πi 1b 2 + 2a 1 c2 b if b|m 1 , a|m 2 ,
e =
0≤c1 ,c2 <c 0 otherwise.
0≤b1 <b
Consequently, it follows that our triple integral (6.4.14) may be written in the
form:
1 c2
· ab2 · a 2 c · a 2 b · c2 b
n 4 abc=n a 2
b|m 1 , a|m 2
⎛⎛ ⎞ ⎛ ⎞⎞
1 1 1 1 x2 x1,3 ay1 y2
× f ⎝⎝ 0 1 x1 ⎠ · ⎝ by1 ⎠⎠
0 0 0 0 0 1 c
cx bx
−2πi m 1 b1 +m 2 a2
×e d x1,3 d x2 d x1 .
Here we have used the fact that if F : R → C is a periodic integrable function
satisfying F(x + 1) = F(x), then for any integer M ≥ 1, we have
M 1
F(x) d x = M · F(x) d x.
0 0
Finally, the triple integral above can be evaluated with (6.4.12) and has the value
⎛⎛ ⎞ ⎞
|m 1 m 2 |y1 y2 c
A mb1 c , ma2 b
m c m b · WJacquet ⎝⎝ m 1 y1 c ⎠, ν, ψ1, m2 ⎠
1 · 2 |m 2 |
b a c
⎛⎛ ⎞ ⎞
|m 1 m 2 |y1 y2
A mb1 c , ma2 b
= m 1 c m 2 b · WJacquet ⎝⎝ m 1 y1 ⎠, ν, ψ1, m2 ⎠ ,
· a |m 2 |
b 1
from which it follows from (6.4.13) and (6.4.14) that
⎛⎛ ⎞ ⎞
|m 1 m 2 |y1 y2
A(m 1 , m 2 )
λn · WJacquet ⎝⎝ m 1 y1 ⎠, ν, ψ1, m2 ⎠
|m 1 m 2 | |m | 2
1
1 c2 A mb1 c , ma2 b
= 4 · ab · a c · a b · c b · m 1 c m 2 b
2 2 2 2
n a2 · a
abc=n b
b|m 1 , a|m 2
⎛⎛ ⎞ ⎞
|m 1 m 2 |y1 y2
× WJacquet ⎝⎝ m 1 y1 ⎠, ν, ψ1, m2 ⎠ .
|m | 2
1
172 Automorphic forms and L-functions for S L(3, Z)
If we cancel the Whittaker functions on both sides of the above identity and
simplify the expressions, we obtain
m1c m2b
λn A(m 1 , m 2 ) = A , . (6.4.15)
abc=n
b a
b|m 1 , a|m 2
It follows that
p pk
A( p, 1)A(1, p ) =k
A , = A(1, p k−1 ) + A( p, p k )
d| p
d d
p k+1 d0
A(1, p)A(1, p ) =
k+1
A d2 , = A(1, p k+2 ) + A( p, p k ),
d0 d2 = p
d2
d2 | p k+1
Remark It is clear that the L-function associated to the dual Maass form f˜
takes the form
∞ −1
L f˜ (s) = A(n, 1)n −s = 1 − A( p, 1) p −s + A(1, p) p −2s − p −3s .
n=1 p
so that the super diagonal elements of the matrix x are x1 , x2 . We also define
z 1 := x1 + i y1 , z 2 := x2 + i y2 .
It follows from Definition 5.4.1 (1), that that for any integers 1 , 2 , the Jacquet
Whittaker function satisfies
where
az 2 + b
z 2 = x2 + i y2 = , y1 = |cz 2 + d|y1 ,
cz 2 + d
x1 = cx1,3 + d x1 ,
x1,3 = ax1,3 + bx1 .
⎛ ⎛ ⎞ ⎞
y1 y2
|cz 2 +d|
⎜ ⎜ ⎟ ⎟
× WJacquet ⎝ M ⎝ y1 · |cz 2 + d| ⎠, ν, ψ1 ,2 ⎠ . (6.5.5)
1
176 Automorphic forms and L-functions for S L(3, Z)
Finally, we obtain the following theorem, which is the basis for the
construction of the L-function L f (s) (given in Definition 6.5.2) as a Mellin
transform.
Theorem 6.5.7 Let f (z) be a Maass form of type ν for S L(3, Z) as in (6.2.1).
Then we have the representation
6 7
∞
A(m 1 , m 2 ) 2πi m 1 (cx1,3 +d x1 )+m 2 azcz 2+d
+b
f (z) = ·e 2
m 1 =1 m 2 =0
|m 1 m 2 |
ab ∈ U2 (Z)\S L(2,Z)
cd
⎛⎛ ⎞⎛ y1 y2 ⎞ ⎞
|m 1 m 2 | |cz 2 +d|
⎜ ⎠⎜ ⎟ ⎟
× WJacquet ⎝⎝ m1 ⎝ y1 · |cz 2 + d| ⎠, ν, ψ1,1⎠.
1 1
6.5 The Godement–Jacquet L-function 177
Proof The proof follows from Theorem 6.4.11, (6.5.5) and Lemma 6.5.6.
Corollary 6.5.8 Let f (z) be a Maass form of type ν for S L(3, Z) as in (6.2.1).
Then
⎛⎛ ⎞ ⎞
1 1 1 u3
f ⎝⎝ 1 u 1 ⎠ · z ⎠ e−2πiu 1 du 1 du 3
0 0 1
⎛⎛ ⎞ ⎞
A(1, m 2 ) |m 2 |y1 y2
= e2πi(x1 +m 2 x2 ) · WJacquet ⎝⎝ y1 ⎠, ν, ψ1,1 ⎠ .
m 2 =0
|m 2 |
1
The integrals
1 1
e 2πim 1 cu 3
du 3 , e2πi(m 1 d−1)u 1 du 1 ,
0 0
vanish unless c = 0 and m 1 d = 1, in which case they take the value 1. The
proof of Corollary 6.5.8 follows immediately from this.
∞ −1
L f (s) = A(1, n)n −s = 1 − A(1, p) p −s + A( p, 1) p −2s − p −3s .
n=1 p
which occurs in the S L(2, Z) theory in Section 3.13. The functional equa-
tion in Z) case arises from the symmetry f (z) = f (ω · z), where
the S L(2,
−1
ω= .
1
One is, thus, highly motivated to try to generalize this idea to S L(3, Z) by
considering symmetries f (z) = f (ω · z) where ω is in the Weyl group. Curi-
⎛ ⎞
−1
ously, the choice ω = ⎝ 1 ⎠ does not work in the S L(3, Z) situation,
1
⎛ ⎞
1
but fortunately the choice ω = ⎝ 1 ⎠ does.
1
We shall now prove Lemma 6.5.9 which contains the symmetry required
to obtain the functional equation of the Godement–Jacquet L-function. A new
feature which does not appear in the G L(2) theory is the unbalanced nature of
this symmetry. One side has a double integral, while the other side has a triple
integral!
For the following Lemma 6.5.9, we define for any function f : h3 → C, its
dual function
⎛ ⎞
t −1 1
f˜(z) = f w (z )w , w=⎝ −1 ⎠,
1
6.5 The Godement–Jacquet L-function 179
and
⎛ ⎞ ⎛ ⎞
1 1
z∗ = ⎝ −1 ⎠ · t (z −1 ) · ⎝ 1⎠.
1 −1
⎛ ⎞ ⎛ −1 ⎞
y1 y2 x3 y2
If z = ⎝ y1 x1 ⎠ , then z ∗ = y1−1 ⎝ x3 y2−1 y1 x1 ⎠ .
1 1
Proof By assumption
f (z) = f˜ w · t (z −1 ) · w .
Consequently
⎛ ⎛ ⎞ ⎞ ⎛ ⎛ ⎞ ⎞
1 u3 1 u1 −u 3
⎝
f w1 ⎝ ⎠ ⎠ ⎝
1 u 1 · z = f w1
˜ ⎝
1 ⎠ · w · t (z −1 ) · w ⎠ .
1 1
(6.5.10)
Integrating both sides of (6.5.10) with respect to u 1 , u 3 gives the first identity
in Lemma 6.5.9.
180 Automorphic forms and L-functions for S L(3, Z)
⎛ ⎛ ⎞ ⎞
1 u3
f ⎝w1 ⎝ 1 u1 ⎠ · z⎠
1
⎛ ⎛ ⎞ ⎞
1 u 1 −u 3
= f˜ ⎝w1 ⎝ 1 ⎠ w1 −1 · w1 · w · t (z −1 ) · ww1 −1 ⎠
1
⎛⎛ ⎞ ⎞
1
−1
= f˜ ⎝⎝ u 3 1 u 1 ⎠ · w1 w · t (z −1 ) · ww1 ⎠ . (6.5.11)
1
⎛ ⎞
1
Note that w1 w = ⎝ −1 ⎠. Recall that z ∗ = w1 w · t (z −1 ) · ww1 −1 . If
1
we integrate both sides of (6.5.11) then
⎛⎛ ⎞ ⎞
1 1 1 u3
f ⎝⎝ 1 u 1 ⎠ · z ⎠ e−2πiu 1 du 1 du 3
0 0 1
⎛ ⎛ ⎞⎛ ⎞ ⎞
1 1 1 1
= f˜ ⎝⎝ 1 u1 ⎠ ⎝ u3 1 ⎠ · z ∗ ⎠ e−2πiu 1 du 1 du 3 .
0 0 1 1
Finally, the proof may be completed by applying the following Lemma 6.5.12
to the integral above.
⎛⎛ ⎞⎛⎞ ⎞
1 1 1 1
f ⎝⎝ 1 u1 ⎠ ⎝ u 1 ⎠ · z ⎠ e−2πiu 1 du 1 du
0 0 1 1
⎛⎛ ⎞⎛ ⎞ ⎞
∞ 1 1 1 u3 1
= f ⎝⎝ 1 u1 ⎠ ⎝ u 1 ⎠ · z ⎠ e−2πiu 1 du 1 du 3 du.
−∞ 0 0 1 1
6.5 The Godement–Jacquet L-function 181
Proof For any integer m note that since f is automorphic for S L(3, Z), we
have
⎛⎛ ⎞⎛ ⎞ ⎞
1 1 1 u3 1
f ⎝⎝ 1 u1 ⎠ ⎝ m 1 ⎠ · z ⎠ e−2πiu 1 du 1 du 3
0 0 1 1
⎛ ⎛ ⎞ ⎛ ⎞ ⎞
1 1 1 1 u3
= f ⎝⎝ m 1 ⎠⎝ 1 u 1 − mu 3 ⎠ · z ⎠ e−2πiu 1 du 1 du 3
0 0 1 1
⎛⎛ ⎞ ⎞
1 1 1 u3
= f ⎝⎝ 1 u 1 ⎠ · z ⎠ e−2πiu 1 e−2πimu 3 du 1 du 3 . (6.5.13)
0 0 1
⎛ ⎞
1
Replacing z by ⎝ u 1 ⎠ · z and then integrating over u in the above identity
1
yields
⎛⎛ ⎞ ⎛ ⎞ ⎞
1 1 1 1
f ⎝⎝ 1 u1 ⎠ · ⎝ u 1 ⎠ z ⎠ e−2πiu 1 du 1 du
0 0 1 1
⎛⎛ ⎞ ⎛ ⎞ ⎞
1 1 1 1 u3 1
= f ⎝⎝ 1 u 1 ⎠ · ⎝ u + m 1 ⎠z ⎠ e−2πiu 1 du 1 du 3 du
m∈Z
0 0 0 1 1
⎛ ⎛ ⎞ ⎛ ⎞ ⎞
∞ 1 1 1 u3 1
= f ⎝⎝ 1 u1 ⎠ · ⎝ u 1 ⎠ z ⎠ e−2πiu 1 du 1 du 3 du.
−∞ 0 0 1 1
where
−3s/2 s + 1 − 2ν1 − ν2 s + ν 1 − ν2 s − 1 + ν1 + 2ν2
G ν (s) = π ,
2 2 2
−3s/2 s + 1 − ν1 − 2ν2 s − ν1 + ν2 s − 1 + 2ν1 + ν2
G̃ ν (s) = π .
2 2 2
Since we are assuming that f is a Maass form, then we know that f (w1 z) =
f (z), and Lemma 6.5.9 takes the form
⎛⎛ ⎞⎞
1 1 y1 y2 u3
f ⎝⎝ y1 u 1 ⎠⎠ e−2πiu 1 du 1 du 3
0 0 1
⎛ ⎛ ⎞ ⎛ −1 ⎞⎞
∞ 1 1 1 u3 y2
= f˜ ⎝⎝ 1 u 1 ⎠ ⎝ uy2−1 y1 ⎠⎠ e−2πiu 1 du 1 du 3 du,
−∞ 0 0 1 1
(6.5.16)
Note that the right-hand sides of (6.5.17) and (6.5.18) are identical. Then
the double Mellin transforms in y1 , y2 of the right-hand sides of (6.5.17) and
(6.5.18) must be the same. For (s1 ), (s2 ) sufficiently large, the double Mellin
184 Automorphic forms and L-functions for S L(3, Z)
Similarly, for −(s1 ), −(s2 ) sufficiently large, the double Mellin transform
of the right-hand side of (6.5.18) converges absolutely and equals:
⎛⎛ ⎞⎞
∞ ∞ 1 1 y1 y2 u3
dy1 dy2
f ⎝⎝ y1 u 1 ⎠⎠ e−2πiu 1 y1s1 −1 y2s2 −1 du 1 du 3
y1 y2
0 0 0 0 1
⎛ ⎛⎛ −1 ⎞ ⎞ ⎞
∞ ∞ ∞ y2
= 2L f˜ (1 − s2 ) · ⎝ WJacquet ⎝⎝ u y1 ⎠, (ν2 , ν1 ), ψ1,1 ⎠ du ⎠
0 0 −∞ 1
dy1 dy2
× y1s1 −1 y2s2
y1 y2
⎛ ∞ ⎛⎛ ⎞ ⎞ ⎞
∞ ∞ y2
= 2L f˜ (1 − s2 ) · ⎝ WJacquet ⎝⎝ u y1 ⎠, (ν2 , ν1 ), ψ1,1 ⎠ du ⎠
0 0 −∞ 1
dy1 dy2
× y1s1 −1 y2−s2 . (6.5.20)
y1 y2
is precisely equal to
G ν (s2 )
.
G̃ ν (1 − s2 )
(6.5.23)
where
α = −ν1 − 2ν2 + 1,
β = −ν1 + ν2 ,
γ = 2ν1 + ν2 − 1.
A number of years later Stade (1990) found another method to obtain (6.5.23).
Finally, we may complete the proof of Lemma 6.5.21 by evaluating the
numerator of the expression in Lemma 6.5.21 using (6.1.4) and then explicitly
computing the ratio of double Mellin transforms given in Lemma 6.5.21.
is a non-zero Maass form for S L(3, Z), normalized so that A(1, 1) = 1, which is
a simultaneous eigenfunction of all the Hecke operators as in Theorem 6.4.11,
then the double Dirichlet series
∞ ∞
A(m 1 , m 2 )
(6.6.2)
m 1 =1 m 2 =1
m s11 m s22
∞ ∞
A(m 1 , m 2 ) L f˜ (s1 )L f (s2 )
s1 s2 = .
m 1 =1 m 2 =1
m1 m2 ζ (s1 + s2 )
It follows that
∞ ∞
A(m 1 , 1) A(1, m 2 ) ∞ ∞ m m
m −s −s2
1 2
s1 s2 = A , 1
1 m2
m 1 =1 m 2 =1
m 1 m 2 m 1 =1 m 2 =1 d|m 1 d|m 2
d d
∞
∞
∞
A(m 1 /d, m 2 /d)
=
d=1 m1 = 1 m 2 =1
m s11 m s22
m 1 ≡ 0 (mod d) m 2 ≡ 0 (mod d)
∞
∞
∞
A(m 1 , m 2 )
=
d=1 m =1 m =1
(m 1 d)s1 (m 2 d)s2
1 2
∞ ∞
A(m 1 , m 2 )
= ζ (s1 + s2 ) · .
m 1 =1 m 2 =1
m s11 m s22
1 (s1 , s2 )
⎛⎛ ⎞ ⎞
∞ ∞ y1 y2
4L f˜ (s1 )L f˜ (1 − s2 )
= WJacquet ⎝⎝ y1 ⎠, ν, ψ1,1 ⎠
((s1 + s2 )/2)ζ (s1 + s2 )
0 0 1
(s −s2 −1)/2 dy1 dy2
× y1s1 −s2 y2 1 K s1 +s2 −1 (y2 ) .
2 y1 y2
and d = ±1 (see below), and kills the sum over S L(2, Z). We have
(s1 , s2 )
∞ A(m 1 , m 2 )
=
m 1 =1 m 2 =0
|m 1 m 2 |
ab ∈U2 (Z)\S L(2,Z)
cd
∞ ∞ 1 6
ai y +b
7
2πi m 1 cu 3 +m 2 ci y 2+d
× e 2 du 3
0
⎛
0 ⎛0 ⎞ ⎞
y1 y2 m 1 |m 2 |
|ci y2 +d|
⎜⎜ ⎟ ⎟
× WJacquet ⎝⎝ y1 m 1 · |ci y2 + d| ⎠, ν, ψ1,1 ⎠
1
dy1 dy2
× y1s1 −1 y2s2 −1
y1 y2
⎛⎛ ⎞ ⎞
∞ ∞∞ y1 y2 m 1 |m 2 |
A(m 1 , m 2 )
=2 WJacquet ⎝⎝ m 1 y1 ⎠, ν, ψ1,1 ⎠
m 1 =1 m 2 =0
|m 1 m 2 |
0 0 1
dy1 dy2
× y1s1 −1 y2s2 −1
y1 y2
⎛⎛ ⎞ ⎞
∞ ∞ ∞ ∞ y1 y2
A(m 1 , m 2 ) ⎝ ⎝ ⎠, ν, ψ1,1 ⎠
=4 W Jacquet y1
m 1 =1 m 2 =1
m s11 |m 2 |s2
0 0 1
dy1 dy2
× y1s1 −1 y2s2 −1 .
y1 y2
where the last identity above is obtained after making the successive transforma-
tions y2
→ y1−1 , y1 y2
→ y2 , and u
→ −u. We now substitute the Whittaker
expansion of Theorem 6.5.7 into the above. In this case the sum over S L(2, Z)
is not killed. Remarkably, however, the integral can be significantly simplified
after several clever transformations. We have
A(m 2 , m 1 )
∞ ∞ 0
∞
1 (s1 , s2 ) =
m 1 =1 m 2 =0
|m 1 m 2 |
ab ∈ U2 (Z)\S L(2,Z)
0 0 −1
cd
⎛⎛ ⎞ ⎞
y1 y2 m 1 |m 2 |
|c(u+i y2 )+d|
2πim 2 (
a(u+i y2 )+b
⎜⎜ ⎟ ⎟
×e c(u+i y2 )+d
· WJacquet ⎝⎝ y1 m 1 ·|c(u+i y2 )+d| ⎠, ν, ψ1,1 ⎠
1
dy1 dy2
× y1s1 −s2 y2s1 −1 du .
y1 y2
Consequently
∞
|c|
2πim 2 r̄ A(m 2 , m 1 )
1 (s1 , s2 ) = e c
m 1 =1 m 2 =0 q∈Z c=0 r =1
|m 1 m 2 |
(r,c)=1
r
∞ ∞
q+ c
2πim 2 u
−
× e |c(u+i y2 )|2
0 0 q+ rc −1
⎛⎛ ⎞ ⎞
y1 y2 m 1 |m 2 |
|c(u+i y2 )|
⎜⎜ ⎟ ⎟
× WJacquet ⎝⎝ y1 m 1 ·|c(u+i y2 )| ⎠, ν, ψ1,1 ⎠
1
dy1 dy2
× y1s1 −s2 y2s1 −1 du . (6.6.6)
y1 y2
6.6 Bump’s double Dirichlet series 191
q+ c
∞
= . (6.6.7)
q∈Z
q+ rc −1 −∞
Next, we combine (6.6.6) and (6.6.7) and then make the successive transfor-
mations
y1 u y2
y1
→ , u
→ 2 , y2
→ 2 .
|c(u + i y2 )| c c
We obtain
∞ |c| 2πim 2 r̄ ∞ ∞ ∞ 2πim u
e c A(m 2 , m 1 ) − 2
1 (s1 , s2 ) = e |u+i y2 |2
m 1 =1 m 2 =0 c=0 r =1
|c|s1 +s2 |m 1 m 2 |
(r,c)=1 0 0 −∞
⎛⎛ ⎞ ⎞
y1 y2 m 1 |m 2 |
|u+i y2 |2
⎜⎜ ⎟ ⎟ s −s s −1
× WJacquet ⎝⎝ y1 m 1 ⎠, ν, ψ1,1 ⎠ y11 2 y21
1
dy1 dy2
× |u + i y2 |s2 −s1 du . (6.6.8)
y1 y2
We now successively make the transformations
|m 2 | y1
u
→ u · y2 , y2
→ y2 · , y1
→
u2 + 1 m1
in (6.6.8). It follows that
∞ |c| 2πim 2 r̄ ∞ ∞ ∞
e c A(m 2 , m 1 ) −
2πim 2 u
1 (s1 , s2 ) = +s
e y2 (u2 +1)
m 1 =1 m 2 =0 c=0 r =1
|c| 1 2 |m 1 m 2 |
s
(r,c)=1 0 0 −∞
⎛⎛ ⎞ ⎞
y1 m 1 |m 2 |
y2 (u 2 +1)
⎜⎜ ⎟ ⎟ s −s s
× WJacquet ⎝⎝ y1 m 1 ⎠, ν, ψ1,1 ⎠ y11 2 y22
1
dy1 dy2
× (u 2 + 1)(s2 −s1 )/2 du
y1 y2
∞ |c| 2πim 2 r̄
e c A(m 2 , m 1 )
= s1 +s2 1+s1 −s2
m 1 =1 m 2 =0 c=0 r =1
|c| m1 |m 2 |1−s2
(r,c)=1
⎛⎛ ⎞ ⎞
∞ ∞ ∞ m2
y1 y2−1
− 2πiu
y2 · |m 2 |
× e WJacquet ⎝⎝ y1 ⎠, ν, ψ1,1⎠
0 0 −∞ 1
y1s1 −s2 y2s2 dy1 dy2
× du . (6.6.9)
(u 2 + 1)(s1 +s2 )/2 y1 y2
192 Automorphic forms and L-functions for S L(3, Z)
Note that the Dirichlet series completely separates from the triple integral of the
Whittaker function in (6.6.9). To complete the proof, we need two lemmas. The
first lemma evaluates the Dirichlet series in (6.6.9) while the second evaluates
the triple integral of the Whittaker function.
Proof To prove Lemma 6.6.10, we apply Proposition 3.1.7. Since every Maass
form is even, we obtain
∞ |c|
e(2πim 2 r̄ )/c A(m 2 , m 1 )
|c|s1 +s2 1+s1 −s2
m 1 =1 m 2 =0 c=0 r =1 m1 |m 2 |1−s2
(r,c)=1
4 ∞ ∞
A(m 2 , m 1 )
= 1 −s2
ζ (s1 + s2 ) m 1 =1 m 2 =1 d|m 2 d s 1 +s2 −1 m 1+s
1 m 1−s
2
2
4 ∞ ∞ ∞
A(m 2 d, m 1 )
=
ζ (s1 + s2 ) m 1 =1 m 2 =1 d=1 m 1 1 −s2 m 1−s
1+s
2
2 s1
d
4 ∞ ∞ ∞
A(m 2 d, m 1 )
=
ζ (s1 + s2 ) m 1 =1 m 2 =1 d=1 (m 1 m 2 )1−s2 (m 1 d)s1
4 ∞ ∞ A m 2 d/m 21 , m 1
=
ζ (s1 + s2 ) m 2 =1 d=1 m 1 |(d,m 2 ) m 1−s
2
2 s1
d
6.6 Bump’s double Dirichlet series 193
4 ∞ ∞
A(d, 1) A(m 2 , 1)
= ·
ζ (s1 + s2 ) m 2 =1 d=1 d s1 m 1−s
2
2
4L f˜ (s1 )L f˜ (1 − s2 )
= .
ζ (s1 + s2 )
Finally, to prove Lemma 6.6.11, we use the identity
∞
−s1 −s2
e±2πiu/y2 (u 2 + 1) 2 du
−∞
s +s
2π 2
1 2
1−s1 −s2
= s1 +s2 y2 2 K s1 +s2 −1 2π y2−1 .
2 2
194
7.1 Converse theorem for S L(3, Z) 195
where τ (χ) is the Gauss sum. This allows us to represent χ (n) as a linear
combination of qth roots of unity. Note that in (7.1.1), the condition (n, q) = 1
can be dropped if χ is primitive. We may also represent each qth root of unity
as a linear combination of characters by the formula
1
χ̄ (
)τ (χ ) = e2πi
/q . (7.1.2)
φ(q) χ (mod q)
converge absolutely for (s) sufficiently large, and for fixed ν = (ν1 , ν2 ) ∈ C2 ,
satisfy the functional equation
τ (χ )2
q 2 s G ν (s + k)L χ (s) = i −k
3 3
√ · q 2 (1−s) G̃ ν (1 + k − s) L̃ χ̄ (1 − s),
τ (χ̄ ) q
196 The Gelbart–Jacquet lift
⎛ ⎞
1 0 0
for all z ∈ h3 , k ∈ Z, A = ⎝ α 1 0 ⎠ with α ∈ Q. Our aim is to prove a
0 0 1
converse to (7.1.6).
Basic Lemma 7.1.7 Let f, f˜ be defined by (7.1.4), (7.1.5), respectively.
Assume
1 1 1 1
−2πiqu 1
f (Auz)e du 1 du 3 = f˜ w2 · t (Auz)−1 e−2πiqu 1 du 1 du 3
0 0 0 0
⎛ ⎞
1 0 0
for all z ∈ h3 , h, q ∈ Z, A = ⎝ h/q 1 0 ⎠, with q = 0. Then f is a Maass
0 0 1
form of type ν for S L(3, Z) and f˜ is its dual form.
Proof of Lemma 7.1.7 First of all, we claim that
⎛ ⎞
∗ ∗ ∗
f ( pz) = f (z), f˜( pz) = f˜(z), ∀ p ∈ P = ⎝ ∗ ∗ ∗ ⎠ ⊂ S L(3, Z).
0 0 1
(7.1.8)
since the sum over γ is permuted by p. A similar argument holds for f˜.
198 The Gelbart–Jacquet lift
⎛ ⎞
1 r
It only remains to show that (7.1.8) holds for p = ⎝ 1 s ⎠ . To show
1
this, we just use the identity
⎛ ⎞⎛ ⎞ ⎛ ⎞
|m 1 m 2 | a b 1 r
⎝ m1 ⎠⎝c d ⎠·⎝ 1 s⎠
1 1 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 r |m 1 m 2 | a b
= ⎝ 1 s ⎠ · ⎝ m1 ⎠⎝c d ⎠,
1 1 1
where
r = |m 1 m 2 |(ar + bs), s = m 1 (cr + ds).
It follows that
∞
A(m 1 , m 2 )
f ( pz) =
γ ∈U2 (Z)\S L(2,Z) m 1 =1 m 2 =0
|m 1 m 2 |
⎛⎛ ⎞ ⎛ ⎞ ⎞
1 r |m 1 m 2 |
γ
× WJacquet ⎝⎝ 1 s ⎠·⎝ m1 ⎠ z, ν, ψ1, |mm2 | ⎠
1 2
1 1
= f (z),
because
⎛⎛ ⎞ ⎞
1 r
WJacquet ⎝⎝ 1 s ⎠ · z, ν, ψ ⎠ = WJacquet (z, ν, ψ)
1
for all z, ν, ψ, if r , s ∈ Z. Again, a similar argument applies for f˜.
Lemma 7.1.9 We have
f˜ w2 · t z −1 = f˜ w2 · t ( pz)−1
⎛ ⎞
1 r
for all p = ⎝ 1 s ⎠ , with r, s ∈ Z.
1
Proof By (7.1.8), we may, without loss of generality replace w2 by
⎛ ⎞
1 0 0
w2 = ⎝ 0 0 1 ⎠.
0 1 0
7.1 Converse theorem for S L(3, Z) 199
We compute:
⎛ ⎞ ⎛ ⎞
1 1
−1 ⎠ · w2 −1
w2 · t p −1 · w2 = w2 · ⎝ 1 = ⎝ −r 1 −s ⎠ ∈ P.
−r −s 1 1
Thus,
−1
f˜ w2 · t ( pz)−1 = f˜ w2 · t p −1 · t z −1 = f˜ w2 · t p −1 · w2 · w2 · t z −1
⎛⎛ ⎞ ⎞
1
= f˜ ⎝⎝ −r 1 −s ⎠ · w2 · t z −1 ⎠ = f˜ w2 · t z −1 .
1
Here we have used the fact that f˜ is invariant under left multiplication by
elements in P.
⎛ ⎞
1 0 0
Now, if (7.1.6) holds for all z ∈ h3 and all h, q ∈ Z, A = ⎝ h/q 1 0 ⎠,
0 0 1
−1
with q = 0, then on choosing z = A z , we must have
1 1
f (Au A−1 z )e−2πiqu 1 du 1 du 3
0 0
1 1
= f˜(w2 · t (Au A−1 z )−1 )e−2πiqu 1 du 1 du 3
0 0
1 1
3 4
f (uz) − f˜ w2 · t (uz)−1 e−2πiqu 1 e−2πi hu 3 du 1 du 3 = 0, (7.1.10)
0 0
*1 3 4
Lemma 7.1.11 We have f (uz) − f˜(w2 · t (uz)−1 ) du 1 = 0, wher e
⎛ ⎞ 0
10 0
u = ⎝ 0 1 u 1 ⎠.
00 1
1
3 4
f (uz) − f˜(w2 · t (uz)−1 ) du 1
0
⎡ ⎛ ⎛ ⎞ ⎞
∞ 0 1 0
⎣ A(m , m ) ⎝ ⎝ 0 ⎠ z⎠
1 2
= WJacquet M 1 0
m 1 =1 m 2 =0
m 1 |m 2 |
0 0 1
⎤
A(m 2 , m 1 )
− W̃Jacquet Mw2 · t z −1 ⎦ ,
m 1 |m 2 |
⎛ ⎞
m 1 |m 2 |
where M = ⎝ m1 ⎠ , and WJacquet is given as in Theorem 6.4.11.
1
But
⎛ ⎞
1
W̃Jacquet (z) = WJacquet w3 · t z −1 · w3 −1
, w3 = ⎝ 1 ⎠.
1
with γ = w2 , for all z ∈ h3 . Finally, we may complete the proof of Lemma 7.1.7.
Note that f (z) = f˜(t (γ z)−1 ) must hold for all γ ∈ S L(3, Z) because it will hold
with γ replaced by p1 γ p2 and p1 , p2 ∈ P. To see this p1 ∈ P implies
note that
γ γ γ
that p1 is either of the form or of the form w2 with
1 1 1
γ , γ ∈ S L(2, Z). In the former case, the invariance due to multiplication by p1
on the left follows from (7.1.8), while in the latter case we have already proved
the invariance of w2 on the left, so we are reduced to the first case. The invariance
due to multiplication by p2 on the right follows by letting z
→ p2−1 · z. So, we
have shown that f is automorphic.
⎛ ⎞
1 0 0
Proof of the converse theorem, Step II Fix A = ⎝ h/q 1 0 ⎠ with
0 0 1
h, q ∈ Z, and q = 0. Let f, f˜ be defined as in (7.1.4) and (7.1.5) and set
1 1
F(z, h, q) = f (Auz)e−2πiqu 1 du 1 du 3 ,
0 0
1 1
F1 (z, h, q) = f˜ w2 · t (Auz)−1 e−2πiqu 1 du 1 du 3 ,
0 0
⎛ ⎞ ⎛ ⎞
u3 1 1
where we recall that u = ⎝ 1 u 1 ⎠ , w2 = ⎝ 1 ⎠ . If we can show
1 1
that F(z, h, q) = F1 (z, h, q) for all z, h, q, k then the basic Lemma 7.1.7 tells
us that f is a Maass form for S L(3, Z). It remains to show that the functional
equations for L χ (s) stated in Theorem 7.1.3 imply that F(z, h, q) = F1 (z, h, q)
for all z, h, q. We shall do this in two remaining steps. First we show that the
functional equations for L χ (s) imply that
k
k
∂ ∂ ∂ ∂
F(z, h, q) = F1 (z, h, q) .
∂ x1 ∂ x2 x1 =x2 =0
∂ x1 ∂ x2 x1 =x2 =0
In the final step 3 of the proof, we extend the above result and prove the basic
Lemma 7.1.7 in all cases.
202 The Gelbart–Jacquet lift
The key point is that the functional equations for L χ (s) imply that
k ∞ ∞
∂ ∂ dy1 dy2
F(z, h, q) y1s1 −1 y2s2 −1
∂ x1 ∂ x2 y1 y2 x1 =x2 =0
0 0
k ∞ ∞
∂ ∂ dy1 dy2
= F1 (z, h, q)y1s1 −1 y2s2 −1 ,
∂ x1 ∂ x2 y1 y2 x1 =x2 =0
0 0
and then by taking inverse Mellin transforms we obtain the desired identity.
The above idea is exemplified in the following two lemmas.
= e qδ e2πiq x1 e qδ
2
m 2 =0
δ · |m 2 |
⎛⎛ ⎞ ⎞
δ|m 2 |y1 y2
∞ ∞
⎜⎜ qδ ⎟ ⎟ s −1 s −1 dy1 dy2
× WJacquet ⎝⎝ δy1 · qδ ⎠, ν, ψ1, |mm2 | ⎠ y11 y22 .
2 y1 y2
0 0 1
The lemma follows after taking partial derivatives and then performing a simple
variable change.
and
∞ ∞
dy1 dy2
F1 (z, h, q) y1s1 −1 y2s2 −1
y1 y2
0 0
∞∞1 1
dy1 dy2
= f˜ w2 · (t (Auz)−1 ) · w2−1 e−2πiqu 1 du 1 du 3 y1s1 −1 y2s2 −1 .
y1 y2
0 0 0 0
(c,d)=1 m 1 =1 m 2 =0
m 1 |m 2 |
0 0 0 0
x2 az 2 +b
−2πim 1 d q + x 2 +y 2
h
2πim 2 cz 2 +d
×e 2 2 e e−2πiqu 1 du 1
⎛⎛ ⎞ ⎞
√ m2 1 |m22 |y1 y2
⎜⎜ x2 +y2 · |cz 2 +d| ⎟ ⎟
× WJacquet ⎜
⎝⎝
⎜ m 1 y2 |cz 2 +d| ⎟ , (ν2 , ν1 ), ψ1,1 ⎟
⎠ ⎠
x22 +y22
1
dy1 dy2
× du 3 y1s1 −1 y2s2 −1 ,
y1 y2
where
'
z 2 = −u 3 − x3 + i y1 x22 + y22 .
m 1 |q m 2 =0
m 1 |m 2 |
∈Z
r =1
r, mq =1
1
7.1 Converse theorem for S L(3, Z) 205
r m1
+ +1
∞ ∞ q u 3 +x3 −1
−2πiq x2 2πim 2
c2 z 2
× e x22 +y22
e
r m1
0 0
+ q
⎛⎛ ⎞ ⎞
√m 12|m 2 |y1 y2
⎜⎜ x2 +y2 · |cz 2 | ⎟ ⎟
2
× WJacquet ⎜
⎝⎝
⎜ m 1 y2 |cz 2 | ⎟, (ν2 , ν1 ), ψ1,1 ⎟
⎠ ⎠
x22 +y22
1
dy1 dy2
× du 3 y1s1 −1 y2s2 −1 ,
y1 y2
which after summing over
∈ Z, and,'then successively making the transfor-
mations: u 3
→ u 3 − x3 , u 3
→ u 3 · y1 x22 + y22 , becomes
A(m 2 , m 1 ) 1
q/m
2πim 1 r h 2πim 1 m 2 r̄
= e2πiq x1 e q e q
m 1 |m 2 |
m 1 |q m 2 =0
r =1
r, mq =1
1
∞ ∞ ∞ √2
−2πiq x2 y1 u 3
√
2
2πim 1 m 2 u 3
0 0 −∞
⎛⎛ m 21 |m 2 |y2
⎞ ⎞
√
⎜⎜ q x2 +y22
2
u 23 +1
√ ⎟ ⎟
⎜⎜ ⎟ ⎟
× WJacquet ⎜⎜ qy1 y2
√ u 23 +1 ⎟, (ν2 , ν1 ), ψ1,1 ⎟
⎝⎝ x22 +y22 ⎠ ⎠
1
'
dy1 dy2
× du 3 y1s1 y2s2 −1 x22 + y22 .
y1 y2
If we now take partial derivatives with respect to x1 , x2 , set x1 = x2 = 0, and
make the substitutions
y1 m 2 |m 2 |y2
y1
→ ' , y2
→ '1 ,
q u 23 + 1 q u 23 + 1
it follows that
∞ ∞
∂
∂ k dy1 dy2
F1 (z, h, q) y1s1 −1 y2s2 −1
∂ x1 ∂ x2 y1 y2 x1 =x2 =0
0 0
(2πiq)
1
q/m
A(m 2 , m 1 ) 2πim 1 r h 2πim 1 m 2 r̄
= e q e q
q s1 +s2 m 1 |q m 2 =0 m 1−2s 2
|m 2 |1−s2
1
r =1
r, mq =1
1
206 The Gelbart–Jacquet lift
⎛⎛ ⎞ ⎞
∞ ∞ k y2−1
2πiqy1 u 3
× WJacquet ⎝⎝ y1 ⎠, (ν2 , ν1 ), ψ1,1 ⎠
m 21 |m 2 |y2
0 0 1
∞ 2πiu 3 m 2 −(s1 +s2 )/2 dy1 dy2
× e y1 y2 |m2 | u 23 + 1 du 3 y1s1 y2s2 .
y1 y2
−∞
+s 1−2s
e2πim 1 q
q s1 2
m 1 |q m 2 =0 m 1
2
|m 2 |1−s2 r =1
r, mq =1
1
⎛⎛ ⎞ ⎞
∞ ∞ k y1 y2
2πiqy12 y2
× WJacquet ⎝⎝ y1 ⎠, (ν2 , ν1 ), ψ1,1 ⎠
m 21 |m 2 |
0 0 1
⎛ ⎞
∞
−2πiu 3 y2 |m 2 | 2 −(s1 +s2 )/2 k dy1 dy2
m
×⎝ e 2 u3 +1 u 3 du 3 ⎠ y1s1 −s2 y2−s2 .
y1 y2
−∞
Note The identity (7.1.16) is a formal identity. It is understood that all series
are well defined by analytic continuation. We present (7.1.16) and its proof in
this form to simplify the exposition of the ideas.
The result follows after several routine computations using Lemmas 7.1.12,
7.1.13, and formula (7.1.1). For example, on the left-hand side we will have
from Lemma 7.1.12 the character sum
qδ
qδ
χ (h δ )e2πim 2 h̄ δ /qδ = χ̄ (h δ ) e2πim 2 h δ /qδ = τ (χ̄ ) · χ (m 2 ),
h δ =1 h δ =1
alternatively, on the right-hand side we will get from Lemma 7.1.13 (for
(r, qδ ) = 1), the character sum
qδ
χ(h δ )e−2πim 1 r ·h δ /qδ = τ (χ )χ̄ (−m 1r ) = τ (χ )χ̄ (−m 1r ),
m 1 |q h δ =1 m 1 |q m 1 |δ
since χ (m 1 ) = 0 if (m 1 , qδ ) > 1.
208 The Gelbart–Jacquet lift
3 τ (χ )2 3
q 2 s G̃ ν (s)L χ (s) = √ · q 2 (1−s) G̃ ν (1 − s) L̃ χ̄ (1 − s),
τ (χ̄ ) q
where
∞
A(n, 1)χ (n) ∞
A(1, n)χ̄ (n)
L χ (s) = , L̃ χ̄ (s) = ,
n=1
ns n=1
ns
and where G ν (s), G̃ ν (s) are products of Gamma functions as defined in Theorem
6.5.15. Note that if we take k = 1 in the above situation, then the identity (7.1.16)
holds automatically because each side will simply vanish since the sum over
m 2 with positive and negative terms will cancel out. We may also consider odd
primitive characters χ . In this case, the identity (7.1.16) automatically holds if
k = 0 and we get a functional equation
3 τ (χ )2 3
q 2 s G̃ ν (s + 1)L χ (s) = i √ · q 2 (1−s) G̃ ν (2 − s) L̃ χ̄ (1 − s),
τ (χ̄ ) q
χ then
∂
∂k ∂
∂k
F(z, h, q) =
F (z, h, q)
k 1
, (7.1.18)
∂ x1 ∂ x2
k
x1 =x2 =0 ∂ x1 ∂ x2 x1 =x2 =0
∞
∞
∂ 2 ∂ ci, j ∂ 2 ci, j
2 2
j j ci, j
1 x1i x2 ci, j = x1i x2 y12 + y2 − y1 y2
i=0, j=0 i=0, j=0 ∂ y12 ∂ y22 ∂ y1 ∂ y2
∞
j
+ y12 (i + 2)(i + 1)x1i x2 ci+2, j
i=0, j=0
∞
j
+ y22 ( j + 2)( j + 1)x1i x2 ci, j+2 .
i=0, j=2
Now assume for fixed j, we have ci, j = 0, this is certainly true for all i
provided j = 0 or 1. Then from (7.1.20), it follows that ci, j+2 = 0. Further,
from (7.1.19) and induction on j, we see that ci, j = 0 for all i and j. This
completes the proof of the converse Theorem 7.1.3.
∞
a(n)b(n)
L f ×g (s) = ζ (2s) . (7.2.3)
n=1
ns
They introduced, for the first time, the bold idea that L f ×g (s) can be constructed
explicitly by taking an inner product of f · ḡ with an Eisenstein series. This
beautiful construction has turned out to be extraordinarily important and has had
many ramifications totally unforeseen by the original discoverers. The Rankin–
Selberg convolution (for the case of Maass forms on S L(2, Z)) and its proof
are given in the following theorem.
7.2 Rankin–Selberg convolution for GL(2) 211
Proof Let E(z, s) denote the Eisenstein series given in Definition 3.1.2. We
compute, for (s) sufficiently large, the inner product
d xd y
ζ (2s) f ḡ, E(∗, s̄) = ζ (2s) f (z) g(z) · E(z, s̄)
y2
S L(2,Z)\h2
ζ (2s) d xd y
= f (z) g(z) Is (γ z)
2 γ ∈ \S L(2,Z) y2
∞
S L(2,Z)\h2
ζ (2s) d xd y
= f (z) g(z) y s
2 y2
∞ \h2
∞ 1
ζ (2s) d xd y
= f (z) g(z) y s
2 y2
00
= πζ (2s) a(n)b(n)
n=0
∞
dy
× K ν f − 12 (2π|n|y)K νg − 12 (2π |n|y) y s
y
0
∞
dy
= (2π) 1−s
L f ×g (s) · K ν f − 12 (y)K νg − 12 (y) y s ,
y
0
212 The Gelbart–Jacquet lift
where
∞
dy
K ν f − 12 (y)K νg − 12 (y) y s
y
0
s+1−ν f −νg s+ν f −νg s−ν f +νg s−1+ν f +νg
2 2 2 2
=2 s−3
.
(s)
This computation gives the meromorphic continuation of L f ×g (s). By
Theorem 3.1.10, the Eisenstein series E(z, s) has a simple pole at s = 1 with
residue 3/π. It follows that L f ×g (s) has a simple pole at s = 1 if and only if
f, g = 0. Finally, the functional equation for f ×g is a consequence of the
functional equation for the Eisenstein series
E ∗ (z, s) = π −s (s)ζ (2s)E(z, s) = E ∗ (z, 1 − s)
given in Theorem 3.1.10.
Finally, it remains to prove the Euler product representation for L f ×g (s). We
shall actually prove a more general result (see (Bump, 1987)). Let u, v : Z −→C
be functions. Let z be a complex variable. Assume that
∞
u(n)z n = (1 − αz)−1 (1 − α z)−1 ,
n=0
∞
v(n)z n = (1 − βz)−1 (1 − β z)−1 ,
n=0
and for |z| sufficiently small, the above series converge absolutely. We will
show that
∞
1 − αα ββ z 2
u(n)v(n)z n = . (7.2.5)
n=0
(1 − αβz)(1 − α βz)(1 − αβ z)(1 − α β z)
In fact, let us now show that (7.2.5) implies the Euler product representation of
L f ×g in Theorem 7.2.4. Recalling (7.2.3), it follows that
∞
+ n n ) p −ns
n=0 a( p )b( p
L f ×g (s) = . (7.2.6)
p (1 − p −2s )
One may now easily check that our result follows from (7.2.6) and (7.2.5) if we
choose
u(n) = a( p n ), v(n) = b( p n ), z = p −s .
To prove (7.2.5), let us define
∞
∞
U (z) := u(n)z n , V (z) := v(n)z n ,
n=0 n=0
7.3 Statement and proof of the Gelbart–Jacquet lift 213
somewhere in the G L(2) theory which have the functional equation (7.3.1). If
one could prove that all the twists by Dirichlet characters of such a candidate
L-function satisfy the functional equations and other conditions given in the
converse Theorem 7.1.3, then the candidate L-function would have to be asso-
ciated to a Maass form on S L(3, Z). The fact that this phenomenon occurs is
the substance of the Gelbart–Jacquet lift. In fact, if f is a Maass form of type
ν f for S L(2, Z), then by Theorem 7.2.4, and the functional equation
s
−s/2 −(1−s)/2 1−s
π ζ (s) = π ζ (1 − s)
2 2
of the Riemann zeta function, we see that
s + 1 − 2ν f s s − 1 + 2ν f L f × f (s)
π −3s/2
2 2 2 ζ (s)
2 − s − 2ν 1 − s −s + 2ν f L f × f (1 − s)
= π − 2 (1−s)
3 f
,
2 2 2 ζ (1 − s)
which exactly matches the functional equation of a self-dual Maass form on
S L(3, Z) of type (2ν f /3, 2ν f /3). We now state and prove the Gelbart–Jacquet
lift from S L(2, Z) to S L(3, Z).
Theorem 7.3.2 (Gelbart–Jacquet lift) Let f be a Maass form of type ν f for
S L(2, Z) with Fourier expansion (7.2.1). Assume that f is an eigenfunction
of all the Hecke operators. Let L f × f (s) be the convolution L-function as in
(7.2.3). Then
L f × f (s)
ζ (s)
is the Godement–Jacquet L-function of a self-dual Maass form of type
(2ν f /3, 2ν f /3) for S L(3, Z).
Proof Let us define, for (s) sufficiently large,
L f × f (s) ∞
L(s) := = A(n, 1)n −s . (7.3.3)
ζ (s) n=1
We need to show that if we define A(n, 1) = A(1, n) for all n = 1, 2, 3, . . . ,
then for (s) sufficiently large,
−1
L(s) = 1 − A( p, 1) p −s + A(1, p) p −2s − p −3s , (7.3.4)
p
and, in addition, we also must show that for every primitive Dirichlet character
χ , the twisted L-function
∞
L χ (s) := A(n, 1)χ (n)n −s
n=1
7.3 Statement and proof of the Gelbart–Jacquet lift 215
satisfies the EBV condition and functional equation given in the converse
Theorem 7.1.3. The proof will be accomplished in three independent
steps.
Lemma 7.3.5 For (s) sufficiently large, the L-function, L(s), (defined in
(7.3.3)) has a degree three Euler product
−1
−1
α 2p α p α p −1 α 2p
L(s) = 1− s 1− 1− s .
p p ps p
satisfies the functional equation specified in the converse Theorem 7.1.3. Such
functional equations, in some special cases of holomorphic modular forms,
were first considered by Rankin (1939). This method was later generalized in
(Li, 1975), (Atkin and Li, 1978), (Li, 1979). Also, Manin and Pančiškin (1977)
obtained the required functional equations for the case of χ (mod q) where q is
216 The Gelbart–Jacquet lift
a prime power. An adelic version of the Rankin–Selberg method for G L(2) was
given in (Jacquet, 1972), but the requisite functional equations are not given
in explicit form. We briefly sketch the method for obtaining such functional
equations.
Proof Since χ is primitive, the Gauss sum τ (χ̄ ), given by (7.1.1), cannot be
zero. It also immediately follows from (7.1.1) that
q
−1
Fχ (z) = τ (χ̄ ) χ̄ (
) · F z + .
=1
q
a b
Assume that γ = ∈ 0 (N q 2 ). We have the following matrix identity:
c d
⎛ ⎞
a +
c b − (ad − 1) d
− cdq 2
1 d 2
/q
2 2
1
/q a b
=⎝ ⎠
q q
.
0 1 c d d − cd
0 1
2
c q
It follows that
q
d 2
Fχ (γ z) = τ (χ̄ )−1 χ̄ (
) · F z +
=1
q
q
d 2
= τ (χ̄ )−1 χ̄ (
d 2 )χ (d)2 · F z +
=1
q
= χ (d)2 Fχ (z).
For the last part, suppose that (
, q) = 1. Then there exist integers r, s such
that rq − s
= 1. Further, we have the matrix identity
1
/q 0 −1 0 −q q −s 1 s/q
= .
0 1 q2 0 q 0 −
r 0 1
0 −q 0 −1
Note that F z =F z = F(z) for all z. Setting s =
¯
q 0 1 0
where
¯ ≡ 1 (mod q), it follows from the prior matrix identity that
q
−1 1
/q 0 −1
τ (χ̄ )Fχ 2 = χ̄ (
)F z
q z
=1
0 1 q2 0
q
1
/q
¯
= χ̄ (
)F z
=1
0 1
q
1
/q
= χ (
)F z
=1
0 1
= τ (χ )Fχ̄ (z).
Lemma 7.3.7 Let χ be an even primitive Dirichlet character (mod q). The
Eisenstein series
1 ys
E(z, s, χ) = χ (d)
2 (c,d)=1 |cz + d|2s
c ≡ 0 (mod q 2 )
+
Let w = u + iv with v > 0. The function |m + w|−2s is periodic in u
m∈Z
and one may easily derive the Fourier expansion
(2s − 1)
|m + w|−2s = 2π |2v|1−2s
m∈Z
(s)2
2π s 1 −s s− 1
+ |v| 2 |m| 2 K s− 12 (2π |m|v)e2πimu .
(s) m=0
It now follows from the definition of the Eisenstein series and the above Fourier
expansion (after making the substitution d = mcq 2 + r ) that
L(2s, χ)E(z, s, χ )
∞
ys
= y s L(2s, χ ) + χ (d) ·
c=1 d∈Z
|cq 2 z + d|2s
s
∞ −2s
2
y
cq
= y s L(2s, χ ) + |c|−2s χ (r ) m + z + r
q4 cq 2
c=1 r =1 m∈Z
s
∞
cq 2
y
= y s L(2s, χ ) + |c|−2s χ (r )
q4 c=1 r =1
0
(2s − 1) 2π s y 12 −s
× 2π |2y| 1−2s
+
(s)2 (s) m=0 m
1
2πim x+ r 2
× K s− 12 (2π|m|y)e cq
√
2π s y 1
= y s L(2s, χ ) + |m|s− 2 K s− 12 (2π |m|y)e2πimx
q 4s (s) m=0
∞
2
cq
2πimr
−2s
× |c| χ (r )e cq 2 . (7.3.8)
c=1 r =1
In fact, this follows easily because every r in the above sum is of the form
r = r1 + tq with 1 ≤ r1 ≤ q and 0 ≤ t < cq. Hence, the sum takes the form
q qc−1 2πir1 m 2πimt
χ (r1 )e cq 2 e cq
r1 =1 t=0
220 The Gelbart–Jacquet lift
and the inner sum over t above is zero unless m =
cq for some
∈ Z. Com-
bining (7.3.8) and (7.3.9), we obtain
L(2s,χ )E(z, s,χ )
√
2π s y σ2s−1 (m, χ̄ )
= y L(2s, χ ) + τ (χ )
s
K s− 12 (2π |m|qy)e2πimqx ,
3s− 12 1
q (s) m=0 |m|s− 2
(7.3.10)
where
m s
σs (m, χ ) = χ (m/
) · .
|m
≥1
Next, we compute
the Fourier
expansion around z = 0. We make use of the
0 −1
fact that wq = is a normalizer for 0 (q 2 ). This means that
q2 0
wq 0 (q 2 )wq−1 = 0 (q 2 ),
which can be easily seen from the calculation
0 −1 a b 0 q −2 d −c
= ∈ 0 (q 2 ).
q2 0 cq 2 d −1 0 −bq 2 a
Thus, for every γ ∈ ∞ \0 (q 2 ), there exists a unique γ ∈ ∞ \0 (q 2 ) such
that γ wq = wq γ . It follows, as in the computation for (7.3.8), that
−1 1
L(2s,χ)E 2 , s, χ = L(2s, χ) χ (γ ) (γ wq z)s
q z 2 γ ∈ \ (q 2 ) ∞ 0
1
= L(2s, χ) χ̄ (γ ) (wq γ z)s
2 γ ∈∞ \0 (q 2 )
s
L(2s, χ ) −1
= χ̄ (d) 2
2 q · (az + b)/(cz + d)
a b
∈ ∞ \0 (q 2 )
c d
a
ys ∞
χ (a) ys ∞
χ (a) r −2s
= = z + m +
q 2s a=1 b∈Z |az + b|2s q 2s a=1 a 2s m∈Z r =1 a
2π 21−2s y 1−s (2s − 1)
= L(2s − 1, χ )
q 2s (s)2
√
2π s y 1
+ 2s σ2s−1 (m, χ )K s− 12 (2π|m|y)e2πimx .
q (s) m=0 |m|s− 12
The functional equation follows after comparing the above with (7.3.10).
7.3 Statement and proof of the Gelbart–Jacquet lift 221
f, f χ̄ · E(∗, s, χ 2 ),
Proof of the Gelbart–Jacquet lift (EBV condition), Step III In Step II we have
shown how to obtain the meromorphic continuation and functional equation for
the twisted L-function
∞
L f × f (s, χ)
L χ (s) = A(n, 1)χ (n)n −s = .
n=1
L(s, χ )
It does not follow from these methods that L χ (s) is entire, however. The problem
is that we do not know that L(s, χ ) divides L f × f (s, χ), and, for all we know,
L χ (s) could have poles at all the zeros of L(s, χ ).
With a brilliant idea, Shimura (1975) was able to show that L(s, χ ) always
divides L f × f (s, χ), so that L χ (s) is entire for all Dirichlet characters χ . We
now explain Shimura’s idea.
It follows from the methods used to prove Lemma 7.3.5 that
∞
∞
L(s, χ) χ (n)a(n 2 )n −s = χ (n)a(n)2 n −s , (7.3.11)
n=1 n=1
where a(n) is the nth Fourier coefficient of f as in (7.2.1). Here, notice the
difference from a(n 2 ) to a(n)2 . Shimura showed that
∞
χ (n)a(n 2 )n −s
n=1
√
∞ ∞
a(n 2 )n δ χ (n) 1 dy
= 2π 1 K ν f − 12 (y) y s− 2 .
n=1 (2πn 2 )s− 2 y
0
has at most a simple pole at s = 34 − 2δ . Actually, such a pole can only occur if
χ is the trivial character, and in this case, we know from the Rankin–Selberg
convolution in Step II that it cannot occur. The residue of the half-integral
1
weight Eisenstein series is, up to a constant factor, y 2 θ(z), where
∞
2
θ(z) = e2πin z
n=−∞
denotes the classical theta function. As a consequence, one obtains the inter-
esting result that
1 d xd y
f (z) |θ(z)|2 y 2 = 0.
y2
0 (4q 2 )\h2
Finally, from the growth properties of Ẽ(z, s, χ ), one may obtain the EBV
(entire and bounded in vertical strips of fixed width) conditions needed for the
completion of the proof of the Gelbart–Jacquet lift.
Proof One easily verifies that the map is well defined and injective. It is
surjective because every relatively prime triple of integers can be completed to
a matrix in S L(3, Z).
⎛ ⎞
∗ ∗ ∗
Let γ = ⎝ ∗ ∗ ∗ ⎠ be a representative for the coset \S
ˆ L(3, Z). Then
a b c
Det(z)
Det(γ z) = 32 ,
y12 |az 2 + b|2 + (ax3 + bx1 + c)2
where z 2 = x2 + i y2 , and Det : h3 → R+ denotes the determinant of a matrix,
where the matrix is in Iwasawa canonical form as in Proposition 1.2.6.
Proof Brute force computation.
Lemmas 7.4.1 and 7.4.2 then imply that the series above converges absolutely
for (s) > 1, and has the explicit representation
2 s
1 y1 y2
E(z, s) = . (7.4.3)
2 a,b,c ∈ Z y 2 az 2 + b2 + (ax3 + bx1 + c)2 3s/2
1
(a,b,c)=1
The most important properties of E(z, s) are given in the next proposition.
Proposition 7.4.4 The Eisenstein series E(z, s) has a meromorphic continu-
ation to all s ∈ C and satisfies the functional equation
−3s/2 3s −(3−3s)/2 3 − 3s
π ζ (3s)E(z, s) = π ζ (3 − 3s) Ẽ(z, 1 − s),
2 2
where Ẽ(z, 1 − s) = E(w t z −1 w, s) is the dual Eisenstein series, and w is the
long element of the Weyl group, as in Proposition 6.3.1. Furthermore,
∗ −3s/2 3s
E (z, s) := π ζ (3s)E(z, s),
2
is holomorphic except for simple poles at s = 0, 1, with residues −2/3, 2/3,
respectively.
Proof The results will follow from the Fourier–Whittaker expansion in exactly
the same way in which they were obtained for Eisenstein series on S L(2, Z)
in Theorems 3.1.8 and 3.1.10. We break the sum for E ∗ (z, s), given in (7.4.3),
into two pieces corresponding to a = 0 and a = 0. It follows that E ∗ (z, s)
= E 1∗ (z, s) + E 2∗ (z, s) where
2 s
3s y1 y2
∗ −3s/2
E 1 (z, s) = π 2 ,
2 b,c ∈ Z2 −(0,0) y b2 + (bx1 + c)2 3s/2
1
7.4 Rankin–Selberg convolution for G L(3) 225
and
∞
∞ ∞ ∞
s
E 2∗ (z, s) =2 y12 y2
a=1 j,
=−∞
−∞ −∞
−3s/2
(3s/2) · e−2πi( ju+
v)
π
× 2 3s/2 dudv. (7.4.5)
y 2 az 2 + u + (ax3 + ux1 + v)2
1
1 s
But E 1∗ (z, s) is y1 y2 times an S L(2, Z) Eisenstein series with Fourier expan-
2
× e2πi (
a(x3 +ξ3 +ξ2 x1 )+( j−
(x1 +ξ1 ))a(x2 +ξ2 )−m 1 ξ1 −m 2 ξ2 )
∞ ∞ 2πi (−
v+ay2 u(
(x1 +ξ1 )− j))
e
× ay2 2 2 3s/2 dudv dξ1 dξ2 dξ3 .
−∞ −∞
y1 y2 a 2 (u 2 + 1) + v 2
226 The Gelbart–Jacquet lift
when m 2 = 0, and to
−3s/23s 2 s
δm 1 ,0 · 2π y1 y2
2
∞ ∞ ∞
2 2 2 2 −3s/2
× ay2 y1 y2 a (u + 1) + v 2 dudv,
a=1
−∞ −∞
⎪
⎪ if m 1 = 0, m 2 = 0,
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎨ 1−s 1−( 2s )
|m 2 | 2 −1 σ2−3s (|m 2 |)K 3s2 −1 (2π|m 2 |y2 )e2πim 2 x2 ,
3s
4y1 y2
⎪
⎪ if m 1 = 0, m 2 = 0,
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
⎪
−3s 1−3s
2y12s y2s π 2 3s2 ζ (3s) + 2y11−s y2s π 2 3s−1 ζ (3s − 1)
⎪
⎪
⎪
⎩ 1−s 2−2s 1− 3s2
3s 2
+ 2y1 y2 π 2 − 1 ζ (3s − 2), if m 1 = m 2 = 0.
The meromorphic continuation and functional equation of E(z, s) are an
immediate consequence of the above explicit computation of the Fourier
coefficients given in (7.4.7).
γ ∈U (Z)\S L(2,Z) m 1 =1 m 2 =0
|m 1 m 2 | 1 2
2
(7.4.8)
∞ B(m 1 , m 2 )
γ
g(z) = WJacquet M z, ν , ψ1, |mm2 | ,
γ ∈U2 (Z)\S L(2,Z) m 1 =1 m 2 =0
|m 1 m 2 | 1 2
7.4 Rankin–Selberg convolution for G L(3) 227
Now
⎛ ⎞
∗ ∗ ∗
⎝
= ∗
ˆ ∗ ∗ ⎠ ⊂ S L(3, Z)
0 0 1
228 The Gelbart–Jacquet lift
where
⎛ ⎞
1 ∗ ∗
U3 (Z) = ⎝ 1 ∗ ⎠ ⊂ S L(3, Z).
1
This result was also obtained in Lemma 5.3.13. Since the Fourier expansion
of f (7.4.10) is given by a sum over S L(2, Z), we may unfold further using
(7.4.11) to obtain
⎛ ⎛ ⎞ ⎞
y1 y2
|cz 2 +d|
⎜ ⎜ ⎟ ⎟
× W̄ Jacquet ⎝ M ⎝ y1 · |cz 2 + d| ⎠, ν , ψ1,1 ⎠
1
s
× y12 y2 d ∗ z,
We will end this section with an explicit computation of the Euler product
for L f ×g assuming that f, g are both eigenfunctions of the Hecke operators and
L f (s), L g (s) have Euler products as given in Definition 6.5.2.
Proposition 7.4.12 (Euler product) Let f, g be Maass forms for S L(3, Z)
with Fourier coefficients A(m 1 , m 2 ), B(m 1 , m 2 ), respectively, as in (7.4.8). Let
∞ ∞
A(m 1 , m 2 ) B(m 1 , m 2 )
L f ×g (s) := ζ (3s)
m 1 =1 m 2 =1 m 2s
1 m2
s
then
−1
3
3
αi, p β j, p
L f ×g (s) = 1 − .
p i=1 j=1 ps
Proof The proof is based on a special case of Cauchy’s identity (Weyl, 1939),
(McDonald, 1979), (Bump, to appear), which takes the form:
3
3
∞
∞
(1 − αi β j x)−1 = Sk1 ,k2 (α1 , α2 , α3 )Sk1 ,k2 (β1 , β2 , β3 )x k1 +2k2
i=1 j=1 k1 =0 k2 =0
× (1 − α1 α2 α3 β1 β2 β3 x 3 )−1 , (7.4.13)
where
k1 +k2 +2
x1
k +1 x2k1 +k2 +2 x3k1 +k2 +2
x 1 x2k1 +1 x3k1 +1
1
1 1 1
Sk1 ,k2 (x1 , x2 , x3 ) := 2
x1 x22 x32
x1 x2 x3
1 1 1
that
A p k1 , p k2 = Sk1 ,k2 (α1, p , α2, p , α3, p ), B p k1 , p k2 = Sk1 ,k2 (β1, p , β2, p , β3, p ),
(7.4.14)
This is proved by noting that the ratio of determinants in the definition of the
Schur polynomial can be computed explicitly to yield
If we then multiply both sides of the above formula by p −(k1 +k2 )s and sum over
k1 , k2 , then the identity (7.4.15) follows after some algebraic manipulations.
Now, by Section 6.5, we have
It follows that
∞
∞
A p k1 , p k2 p −(k1 +k2 )s
k1 =0 k2 =0
∞
∞
= (1 − p −2s ) A p k1 −k , p k2 −k p −(k1 +k2 )s
k1 =0 k2 =0 k≤min(k1 ,k2 )
∞
∞
= (1 − p −2s ) A p k1 , 1 A 1, p k2 p −(k1 +k2 )s
k1 =0 k2 =0
−2s
−1
= (1 − p ) 1 − A( p, 1) p −s + A(1, p) p −2s − p −3s
−1
× 1 − A(1, p) p −s + A( p, 1) p −2s − p −3s
7.4 Rankin–Selberg convolution for G L(3) 231
−1 −1 −1
= (1 − p −2s )(1 − α1 p −s ) (1 − α2 p −s ) (1 − α3 p −s )
−1 −1 −1
× (1 − α2 α3 p −s ) (1 − α3 α1 p −s ) (1 − α1 α2 p −s )
∞
∞
= Sk1 ,k2 (α1 , α2 , α3 ) p −(k1 +k2 )s .
k1 =0 k2 =0
as a factor. By comparing degrees one easily sees that the determinant of the Van-
dermonde matrix is the product (7.4.16) up to a constant factor. To show that the
constant factor is 1, one may consider the main diagonal term x1n−1 x2n−2 · · · xn1 xn0 ,
which is exactly what one gets by taking the first positive terms in each factor
of the product (7.4.16). This proves that
n−1
x1 x2n−1 · · · xnn−1
n−2
x1 x2n−2 · · · xnn−2
.. .. .. = (x − x ).
. . ··· . i j (7.4.17)
x ··· 1≤i< j≤n
1 x2 xn
1 1 ··· 1
Next, we consider Cauchy’s determinant which is defined to be the deter-
minant of the n × n matrix whose i, jth entry is 1/(1 − xi y j ) where xi , y j
(1 ≤ i, j ≤ n) are variables and n ≥ 2. We write this matrix as
1
.
1 − xi y j 1≤i, j≤n
In fact, the numerator of (7.4.19) has degree n(n − 1) while the denominator
has degree n 2 which is also the case for the determinant because all of its terms
are products of n factors, each having one term in the denominator and none
in the numerator with a net increase of n in the denominator. To show that the
constant multiple must be 1, let xi = −yi−1 for i = 1, 2, . . . , n. In this case,
(7.4.19) becomes
)
(yi − y j )2
1≤i< j≤n
2−n · ) .
(yi + y j )2
1≤i< j≤n
where the denominator is the Vandermonde determinant which has the value
given by (7.4.17).
)
n )
n
where α = αi and β = βi .
i=1 i=1
where
a
((x1 , . . . , xn )) = (σ )σ x1
1 · · · xn
n ,
σ ∈Sn
in which Sn is the group of permutations of {1, 2, . . . , n}, and (σ ) is the sign
of the permutation σ .
It is clear that the polynomial a
(x) := a
((x1 , . . . , xn )) satisfies
σ (a
(x)) = (σ )a
(x)
for any σ ∈ Sn , and, therefore, vanishes unless
1 , . . . ,
n are all distinct. So,
we may assume
1 >
2 > · · · >
n ≥ 0. In view of the skew symmetry, we
may write
(
1 , . . . ,
n ) = (λ1 + n − 1, λ2 + n − 2, λ3 + n − 3, . . . , λn ).
It follows that
λ j +n− j
a
(x) = aλ+δ = Det xi 1≤i, j≤n
,
is the L-function associated to a Maass form on G L(n), then one may consider
−1
L(s, ∨k ) := 1 − α p,i1 α p,i2 · · · α p,ik p −s ,
p 1≤i 1 ≤i 2 ≤···≤i k ≤n
235
236 Bounds for L-functions and Siegel zeros
Axiom 8.1.4 (Euler product) We may express log L(s) by the Dirichlet series
∞
bm
log L(s) =
m=2
ms
a p a p
= O(1).
p≤x p
See (Kowalski, 2003), (Ramakrishnan and Wang, 2003) for further discus-
sions of the following very interesting conjecture.
It follows from (Conrey and Ghosh, 1993) that Conjectures 8.1.5 and 8.1.6
imply Conjecture 8.1.9.
In order to classify the Dirichlet series L(s) in the Selberg class, it is conve-
nient to introduce the degree d L of L ∈ S as
n
dL = 2 λj.
j=1
m
s
(s m ) = · · |s|m < − · |s|m
|s| 2
it follows that for T > T0 , sufficiently large, that
+|s|α
|,m (s)| ≤ e− 2 |s|
m
· |φ(s)| ≤ |φ(s)| ≤ B,
Therefore,
|φ(s)| ≤ e|s| B
m
(8.2.2)
on LT . A similar argument can be given for the case t < 0 from which it follows
that (8.2.2) also holds on any such L−T . By the maximum principle, we can
now assert that (8.2.2) holds for any > 0 on the entire strip with vertical
sides consisting of the lines (s) = σ1 , (s) = σ2 . Letting → 0 in (8.2.2)
establishes that
|φ(s)| ≤ B
everywhere inside this strip which proves the theorem in the particular case
considered.
In the general case, let
α
u(s) = e M(s) log((−is) ) ,
where the logarithm has its principal value. Then the function u(s) is holomor-
phic for σ1 ≤ (s) ≤ σ2 , and (s) = t ≥ 1. If we write M(s) = M(σ ) + ibt,
then
Hence |u(s)| = |t| M(σ ) eO(1) , and, therefore, the function (s) = φ(s)/u(s)
satisfies the same conditions as φ(s) did in the first part. Thus, (s) is bounded
in the strip and the theorem follows.
We have thus established that L(s) is bounded for (s) = 1 + and is bounded
by
λ(1+2)
n
A 1+2
(1 + |λt + κ j |)
j=1
+
∞
Theorem 8.3.3 (Approximate functional equation) Let L(s) = am /m s ∈ S,
m=1
be entire and satisfy the functional equation (8.3.1). Define qw as in (8.3.2).
Then there exists a smooth function F : (0, ∞) → C such that for every w ∈ C
with 0 ≤ (w) ≤ 1, we have
∞ ∞
am m am m
L(w) = F + λw F̄ ,
m=1
mw qw m=1
m 1−w q1−w
and which is bounded in the vertical strip −2 < (s) < 2. For every w ∈ C,
and x > 0, we define
2+i∞
1 G(s + w) ds
Hw (x) := h(s)x −s , (8.3.5)
2πi G(w) s
2−i∞
where
n
G(s) = (λs + µ j ),
j=1
1++i∞
1 As+w G(s + w)L(s + w) ds
I L (w) := h(s) .
2πi Aw G(w) s
1+−i∞
If we shift the line of integration to the left we pick up a residue of the pole of
the integrand at s = 0. It follows, after applying the functional equation (8.3.1),
and then transforming s → −s, that
−1−+i∞
1 A1−s−w G(1 − s − w)L(1 − s̄ − w̄) ds
I L (w) = L(w) + w
h(s)
2πi A G(w) s
−1−−i∞
A1−w G(1 − w)
= L(w) −
2πi Aw G(w)
1++i∞
As+1−w G(s + 1 − w) L̃(s + 1 − w) ds
× h(s) ,
A1−w G(1 − w) s
1+−i∞
or equivalently
A1−w G(1 − w)
L(w) = I L (w) + · I L̃ (1 − w), (8.3.6)
Aw G(w)
+
∞
where L̃(s) = am /m s denotes the dual L-function. In (8.3.6) we may substi-
m=1
tute the Dirichlet series for L(s) and L̃(s) and integrate term by term. It follows
that
∞
am m ∞
am m
L(w) = Hw + λ w H1−w . (8.3.7)
m=1
mw A m=1
m 1−w A
Consequently
)
n
|(λ(w + s) + µ j )|
G(w + s)
j=1 π
qσw e 2 nλ|s| .
G(w) = )
n (8.3.8)
|(λw + µ j )|
j=1
In (8.3.9), under the assumption that h(s) has sufficient decay, we may shift
the line of integration either to the left (picking up the residue 1 at the pole at
s = 0) or we may shift to the right. After shifting to an arbitrary line R(s) = σ
and differentiating k times with respect to x, we obtain from (8.3.9) that
k
d (−1)k
F(x) = δσ,k +
dx 2πi
σ+i∞
G(s + w) −s−k ds
× h(s)s(s + 1) · · · (s + k − 1)q−s
w x ,
G(w) s
σ −i∞
(8.3.10)
1 σ < 0, k = 0,
where δσ,k =
0 otherwise.
It now follows from (8.3.8) and (8.3.10) that
⎛ σ +i∞ ⎞
k
d π
F(x) = δσ,k + Oσ,k ⎝ e 2 n|s| (1 + |s|)k · |h(s)| · x −σ |ds|⎠ .
dx
σ −i∞
Definition 8.4.1 (Siegel zero) Fix a constant c > 0 and an integer n ≥ 1. Let
L(s) ∈ Sn satisfy the functional equation given in Axiom 8.1.3. Assume that
L(β) = 0 for some real β satisfying
c
1− ≤ β ≤ 1,
log(λA + 1)
where λ = max (λi + |µi |). Then β is termed a Siegel zero for L(s) relative
1≤i≤n
to c.
Interlude on the history of Siegel zeros Let D < √0 denote the fundamental
discriminant of an imaginary quadratic field k = Q( D). Then D ≡ 1 (mod 4)
and square-free, or of the form D = 4m with m ≡ 2 or 3 (mod 4) and square-
free. Define
group of non-zero fractional ideals ba
h(D) = #
group of principal ideals (α), α ∈ k ×
to be the cardinality of the ideal class group of k. Gauss (1801) showed (using the
language of binary quadratic forms) that h(D) is always finite. He conjectured
that
h(D) → ∞ as D → −∞,
Gauss made the remarkable conjecture that his tables were complete. In modern
parlance, we may rewrite Gauss’ tables in the form (see also (Goldfeld, 2004)):
h(D) 1 2 3 4 5
# of fields 9 18 16 54 25 . (8.4.2)
largest |D| 163 427 907 1555 2683
for some M ≥ 1, B ≥ 0, and all t ∈ R. If L(s) has no real zeros in the range
1 − (1/log M) < s < 1, then there exists an effective constant c(B) > 0 such
that
R −1 ≤ c(B) log M.
Proof Let r > B be a fixed integer. We shall make use of the well-known
integral transform
2+i∞ r
1 xs 1
1 − x1 , x > 1,
ds = r !
2πi 2−i∞ s(s + 1) · · · (s + r ) 0, 0 < x ≤ 1,
which is proved by either shifting the line of integration to the left (if x > 1)
or to the right (if 0 < x ≤ 1), and then computing the sum of the residues with
Cauchy’s theorem. Now,
∞
a(m)
L(s) =
m=1
ms
Let us now use Lemma 8.4.3 to relate the Siegel zero with the Gauss class
number problem for imaginary quadratic fields. Let D <√0 be the fundamen-
tal discriminant of an imaginary quadratic field k = Q( D). The relation is
through the zeta function
∞
a(m)
ζk (s) = ζ (s)L χ (s) = ,
m=1
ms
Actually, much stronger bounds are known, but we do not need them here. It
follows from (8.4.7), (8.4.8), (8.4.9) that ζk (s) satisfies the conditions required
in Lemma 8.4.3. It then follows from Lemma 8.4.3 that ζk (s) has a Siegel zero
if the class number h(D) is so small that
1
|D| 2
h(D) . (8.4.10)
log |D|
The above result was first published by Landau (1918), but Landau attributes
this result to a lecture given by Hecke.
8.5 Siegel’s theorem 249
When combined with the Landau–Hecke result (8.4.10) this gave an uncondi-
tional proof of Gauss’ conjecture that the class number of an imaginary quadratic
field goes to infinity with the discriminant. The surprising aspect of this chain of
theorems is that first one assumes the Riemann hypothesis to establish a result
and then one assumes that the Riemann hypothesis is false to obtain the exact
same result! This is now called the Deuring–Heilbronn phenomenon, but has
the defect of being totally ineffective. Siegel (1935) practically squeezed the
last drop out of the Deuring–Heilbronn phenomenon. He proved the following
theorem, which is the main subject matter of this section.
√
Theorem 8.5.1 (Siegel’s theorem) Let Q( D) be an imaginary quadratic
field with fundamental discriminant D < 0 and class number h(D). Then
for every > 0, there exists a constant c > 0 (which cannot be effectively
computed) such that
Remarks Landau (1935) proved Theorem 8.5.1 with = 18 (also not effec-
tive). Siegel’s theorem, with any > 0, appeared in the same volume of Acta
Arithmetica, but did not reference Landau’s result at all!
Note that since h(D) is a positive rational integer, it follows from (8.4.8) that
L(1, χ ) |D|− 2 .
1
and the fact that χ , χ can only take values among {−1, 0, +1}, one readily
establishes that Z (s) is a Dirichlet series whose first coefficient is 1 and whose
other coefficients are non-negative.
Lemma 8.5.2 For every > 0, there exists χ (mod D ) and β ∈ R satisfying
1 − < β < 1 such that Z (β) ≤ 0 independent of what χ (mod D) may be.
Proof If there are no zeros in [1 − , 1] for any L(s, χ ), then Z (β) < 0 if
1 − < β < 1 since ζ (β) < 0 and all L-functions L(s, χ) will be positive in
the interval. Here we use the fact that L(s, χ ) is positive for (s) > 1 and can
only change sign in the interval [1 − , 1] if the L-function vanishes in the
interval.
On the other hand, if such real zeros do exist, let β be such a zero, with χ
the corresponding character. Then Z (β) = 0 independent of χ .
Tatuzawa (1951) went a step beyond Siegel and proved that Siegel’s
Theorem 8.5.1 must hold with an effectively computable constant c > 0 for
all D < 0, except for at most one exceptional discriminant D. This shows that
the family of real Dirichlet L-functions can have at most one L-function with a
Siegel zero. It can be shown (see (Davenport, 1967)) that complex Dirichlet
L-functions cannot have Siegel zeros. Thus, the entire family of G L(1)
L-functions can have at most one exceptional L-function with a Siegel zero.
The exceptional L-function must correspond to a real primitive Dirichlet char-
acter associated to a quadratic field.
so that
1
b=− .
ρ ρ
It follows that
m m G (s) L (s) 1
+ + log A + + = .
s s−1 G(s) L(s) ρ s −ρ
Lemma 8.4.3 when we choose ζ (s)L(s, χ) as our Dirichlet series, which has
all the required properties of Lemma 8.4.3 and has a simple pole at s = 1 with
residue L(1, χ). A real breakthrough was achieved in (Hoffstein and Lockhart,
1994) when they took the study of Siegel zeros outside the classical domain of
Dirichlet L-functions and considered, for the first time, the question of whether
such zeros could exist for L-functions associated to automorphic forms on
G L(3) occurring as symmetric square lifts (Gelbart–Jacquet lifts) from G L(2).
It now became possible to obtain lower bounds of special values of L-functions
in the same way as the classical methods (using the Deuring–Heilbronn phe-
nomenon) gave lower bounds for L(1, χ) with χ a real primitive Dirichlet
character. This established a powerful new tool in modern analytic number
theory.
In (Hoffstein and Lockhart, 1994), it was shown that if f is a Maass form
for S L(2, Z) which is an eigenfunction of the Hecke operators, and F is its
symmetric square lift (Gelbart–Jacquet lift) to S L(3, Z), then the lifted L-
function,
L f × f (s)
L F (s) = , (8.7.1)
ζ (s)
given in Theorem 7.3.2 cannot have a Siegel zero. Actually, they proved a
more general result valid for congruence subgroups of S L(2, Z) and also con-
sidered Gelbart–Jacquet lifts of both holomorphic modular forms and non-
holomorphic Maass forms. Their proof was based on a generalization of Siegel’s
Theorem 8.5.1 and, thereby, was not effective. In the appendix of their paper
Goldfeld, Hoffstein and Lieman (1994) obtained an effective version of their
theorem which was based on the Siegel-zero Lemma 8.6.1. This effective proof
is the subject matter of this section.
The key idea in the proof of the non-existence of Siegel zeros for L-functions
of type (8.7.1) is the construction of an auxiliary L-function which has non-
negative Dirichlet coefficients and a multiple pole at s = 1, and satisfies the
requirements of Lemma 8.6.1. Accordingly, we introduce
+
∞ +
∞
Here, if L F (s) = c(m)m −s , then L F×F (s) = ζ (3s) |c(m)|2 m −s as we
m=1 m=1
recall from Proposition 7.4.12. In terms of Euler products, if
α p −1
α p −1
L f (s) = 1− s 1− s ,
p p p
254 Bounds for L-functions and Siegel zeros
with α p · α p = 1, and
−1 −1
−1
α 2p 1 α p 2
L F (s) = 1− 1− s 1− ,
p p ps p ps
then
−1
−1 −1
−1
α 4p α 2p 1 α 2p
L F×F (s) = 1− 1− 1− s 1−
p ps ps p ps
−1
−1 −1
1 α p 2 1
× 1− s 1− 1− s
p ps p
−1
−1
α p 2 α p 4
× 1− 1− . (8.7.3)
ps ps
It follows that
Finally, we obtain
+
∞
Lemma 8.7.5 Let Z (s) = a(m)m −s be given by (8.7.2). Then a(1) = 1
m=1
and a(m) ≥ 0 for m = 2, 3, 4, . . .
8.7 Non-existence of Siegel zeros for Gelbart–Jacquet lifts 255
Proof The fact that a(1) = 1 is an immediate consequence of the Euler prod-
uct for Z (s). Now, the Euler product for Z (s) takes the form
−1
α p1 α p 2
1−
p ps
where the product is taken over all sixteen possible pairs (1 , 2 ), and where
1 , 2 , independently run through the values 2, 0, 0, 2. If one takes logarithms,
it follows that the pth term in the expansion of log Z (s) is
∞
(α 2
+ α −2
+ 2)(α 2
+ α −2
+ 2)
.
=1
p
s
Since α 2 , α 2 , are either non-negative real numbers or lie on the unit circle, it
follows that the above series has non-negative terms. Consequently, so does the
series for Z (s).
Proof We make use of (8.7.4). It follows from the Euler product that for s ∈ R,
s > 1 that Z (s)/Z (s) < 0. In (Bump and Ginzburg, 1992) it is shown that when
f is not a lift from G L(1), then L F (s, ∨2 ) has a simple pole at s = 1, and any
zero of L F (s) will be a zero of Z (s) with order at least 3. Consequently, if
L(F, s) has a Siegel zero, then Z (s) = ζ (s)L F (s)3 L F (s, ∨2 ) will have 3 Siegel
zeros. Since Z (s) has a pole of order 2, we obtain a contradiction from the
Siegel zero Lemma 8.6.1.
In the classical case, the non-existence of Siegel zeros implies a lower bound
for the class number of an imaginary quadratic field. One may ask what takes
the place of class numbers in the G L(3) setting. The answer to this question
is given in Lemma 8.4.3 which says that we will obtain a lower bound for the
residue (at s = 1) of the relevant L-function. We state an important and useful
corollary to Theorem 8.7.6.
Corollary 8.7.7 Let f be a Maass form for S L(2, Z) of type ν. Then the
Petersson inner product of f with itself satisfies
1
f, f .
log(1 + |ν|)
256 Bounds for L-functions and Siegel zeros
L f (s) has a Siegel zero. Let L f (s, ∨2 ) denote the symmetric square lift as in
Theorem 7.3.2, which is associated to a cusp form F on G L(3).
Now, construct
where f˜ denotes the Maass form dual to f . Let D(s) denote the Rankin–Selberg
convolution of Z (s) with itself. Then D(s) will have a pole of order 3, but it
will have two copies of L f (s) and an additional two copies of L f˜ (s) as factors.
So if L f (s) has a Siegel zero then D(s) will have four Siegel zeros and a pole
of order 3 at s = 1 which again contradicts the Siegel zero Lemma 8.6.1. This
establishes that L-functions associated to non-self dual Maass forms on G L(n),
for n ≥ 3, cannot have Siegel zeros. Note that this situation is analogous to the
way one proves that complex Dirichlet L-functions cannot have Siegel zeros.
Finally, we consider the case of self dual Maass forms on G L(n) with n ≥ 3.
We again follow Hoffstein and Ramakrishnan (1995) who proved that Siegel
zeros cannot exist if one assumes Langlands’ conjectures. Let f denote a self
dual Maass form on G L(n). Assume there exists some g = f where g is not
an Eisenstein series such that
and take the Rankin–Selberg convolution of Z with itself. Then this Rankin–
Selberg convolution will have non-negative coefficients, a pole of order 3 at
s = 1, and it will have L f (s)4 as a factor assuming everything else is analytic. If
L f (s) had a Siegel zero, then this would contradict the Siegel zero Lemma 8.6.1.
For the case of G L(3), Hoffstein and Ramakrishnan (1995) proved there are
no Siegel zeros subject to a certain analyticity hypothesis. This hypothesis was
subsequently proved in (Banks, 1997) which establishes that there are no Siegel
zeros on G L(3) except for the obvious cases.
9
The Godement–Jacquet L-function
259
260 The Godement–Jacquet L-function
∞
∞
γ
φ(z) = ··· φ̃ (m 1 ,...,m n−1 ) z ,
γ ∈Un−1 (Z)\S L(n−1,Z) m 1 =1 m n−2 =1 m n−1 =0
1
Note the change of notation in the Fourier expansion above to conform with
Remark 9.1.1.
Now, we have shown that φ̃ (m 1 ,...,m n−1 ) (z) is a Whittaker function. Further,
φ̃ (m 1 ,...,m n−1 ) (z) will inherit the growth properties of the Maass form φ and will
satisfy the conditions of the multiplicity one theorem of Shalika (1974) which
states that only the Jacquet Whittaker function (5.5.1) can occur in the Fourier
expansion of a Maass form for S L(n, Z) and that φ̃ (m 1 ,...,m n−1 ) must be a constant
multiple of the Jacquet Whittaker function. It follows from Theorem 5.3.2 and
Proposition 5.5.2 that if φ is a Maass form of type ν = (ν1 , . . . , νn−1 ) ∈ Cn−1
for S L(n, Z) then
∞
∞ A(m 1 , . . . , m n−1 )
φ(z) = ···
γ ∈Un−1 (Z)\S L(n−1,Z) m 1 =1 m n−2 =1 m n−1 =0
)
n−1
|m k |k(n−k)/2
k=1
γ
× WJacquet M · z, ν, ψ1,...,1, m n−1 , (9.1.2)
1 |m n−1 |
where
⎛ ⎞
m 1 · · · m n−2 · |m n−1 |
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
M =⎜ m1m2 ⎟, A(m 1 , . . . , m n−1 ) ∈ C,
⎜ ⎟
⎝ m1 ⎠
1
9.2 The dual and symmetric Maass forms 261
and
⎛⎛ ⎞⎞
1 u n−1
⎜⎜ 1 u n−2 ∗ ⎟⎟
⎜⎜ ⎟⎟
⎜⎜ .. .. ⎟⎟ 2πi u 1 +···+u n−2 +u n−1
ψ1,...,1, ⎜⎜ . . ⎟⎟ = e .
⎜⎜ ⎟⎟
⎝⎝ 1 u1 ⎠⎠
1
A(m 1 ,... ,m n−1 )
The particular normalization )
n−1 is chosen so that later formulae are
|m k |k(n−k)/2
k=1
as simple as possible.
Lemma 9.1.3 (Fourier coefficients are bounded) Let φ be a Maass form for
S L(n, Z) as in (9.1.2). Then for all non-zero integers m 1 , . . . , m n−1 ,
A(m 1 , . . . , m n−1 )
= O(1).
)
n−1
|m k |k(n−k)/2
k=1
where
d ∗ x = d x1 · · · d xn−1 d xi, j .
1≤i< j+1≤n
Proposition 9.2.1 Let φ(z) be a Maass form of type (ν1 , . . . , νn−1 ) ∈ Cn−1 as
in (9.1.2). Then (for !x" the largest integer ≤ x)
⎛ ⎞
(−1)[n/2]
⎜ 1 ⎟
⎜ ⎟
φ̃(z) := φ(w · t (z −1 ) · w), w = ⎜ . ⎟
⎝ .. ⎠
1
is a Maass form of type (νn−1 , . . . , ν1 ) for S L(n, Z). The Maass form φ̃ is
called the dual Maass form. If A(m 1 , . . . , m n−1 ) is the (m 1 , . . . , m n−1 )-Fourier
coefficient of φ then A(m n−1 , . . . , m 1 ) is the corresponding Fourier coefficient
of φ̃.
then
⎛ ⎞⎛ ⎞
1 δx1 y1 y2 · · · yn−1
⎜ 1 −x2 ∗ ⎟ ⎜ y2 y3 · · · yn−2 ⎟
⎜ ⎟⎜ ⎟
−1 −1 ⎜ . . ⎟ ⎜ . ⎟
w · (z ) · w =⎜
t .. .. ⎟⎜ .. ⎟
⎜ ⎟⎜ ⎟
⎝ 1 −xn−1 ⎠ ⎝ yn−1 ⎠
1 1
(9.2.2)
!n/2"+1
where δ = (−1) . One may then show that
φ̃(z) du = 0,
(S L(n,Z)∩U )\U
9.2 The dual and symmetric Maass forms 263
to pick off the (m 1 , . . . , m n−1 )-Fourier coefficient, then because x j and xn− j
are interchanged (for j = 1, . . . , n − 1) we will actually get A(m n−1 , . . . , m 1 ).
In the S L(2, Z) theory, the notions of even and odd Maass forms (see
Section 3.9) played an important role. If a(n) is the nth Fourier coefficient
of an S L(2, Z) Maass form then a(n) = ±a(−n) depending on whether the
Maass form is even or odd. We shall see that a similar phenomenon holds in the
case of S L(n, Z) when n is even. On the other hand, if n is odd then all Maass
forms are actually even (see Section 6.3 for the example of n = 3).
Consider a diagonal matrix δ of the form
⎛ ⎞
δ1 · · · δn−1
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
δ := ⎜ δ1 δ2 ⎟
⎜ ⎟
⎝ δ1 ⎠
1
)
n
with j ∈ {+1, −1}, for j = 1, 2, . . . , n and j = 1. It is also invariant by
j=1
the central element
⎛ ⎞
−1
⎜ −1 ⎟
⎜ ⎟
⎜ .. ⎟
⎝ . ⎠
−1
which has determinant −1. Since these elements generate all possible Tδ , this
proves Tδ φ = φ for all Tδ .
Next, let
⎛ ⎞
−1
⎜ 1 ⎟
⎜ ⎟
Tδ0 = ⎜ .. ⎟.
⎝ . ⎠
1
Then since δ0 zδ0 transforms xn−1 → −xn−1 and fixes x j with 1 ≤ j ≤ n − 2,
it follows that the integral
1 1 3 4
··· Tδ0 φ(z) e−2πi m 1 x1 +···+m n−1 xn−1
d∗x
0 0
picks off the A(m 1 , . . . , m n−2 , −m n−1 ) coefficient of φ, and this must be the
same as A(m 1 , . . . , m n−2 , m n−1 ) because Tδ0 φ = φ.
We next show that the theory of symmetric Maass forms for S L(n, Z) (with
n even) is very similar to the S L(2, Z) theory. Basically, there are only two
types of such Maass forms, even and odd Maass forms.
Proposition 9.2.6 Assume n ≥ 2 is an even integer. Let φ be a symmetric
Maass form for S L(n, Z) with Fourier coefficients A(m 1 , . . . , m n−1 ) as in the
expansion (9.1.2). Fix
⎛ ⎞
−1
⎜ 1 ⎟
⎜ ⎟
Tδ0 = ⎜ .. ⎟.
⎝ . ⎠
1
Then for all m j ≥ 1 (with j = 1, 2, . . . , n − 1), we have
A(m 1 , . . . , m n−2 , m n−1 ) = ±A(m 1 , . . . , m n−2 , −m n−1 ),
according as Tδ0 φ = ±φ. The Maass form φ is said to be even or odd
accordingly.
266 The Godement–Jacquet L-function
Proof The diagonal elements with ±1 entries are generated by such elements
with determinant 1 and the additional special element Tδ0 . Since φ is invariant
under left multiplication by S L(n, Z) it follows that φ is symmetric if and only
if Tδ φ = ±φ.
Now δ0 zδ0 transforms xn−1 → −xn−1 and fixes x j with 1 ≤ j ≤ n − 2, it
follows that the integral
1 1 3 4
··· Tδ0 φ(z) e−2πi m 1 x1 +···+m n−1 xn−1
d∗x
0 0
picks off the A(m 1 , . . . , m n−2 , −m n−1 ) coefficient of φ, and this must be the
same as ±A(m 1 , . . . , m n−2 , m n−1 ) because Tδ0 φ = ±φ.
where f ∈ L (\X ), x ∈ X, and αi are given by (9.3.1). The Hecke ring consists
2
g → t g, g ∈ ,
where tg denotes the transpose of the matrix g. It is again clear that the conditions
of Theorem 3.10.10 are satisfied so that the Hecke ring is commutative.
The following lemma is analogous to Lemma 3.12.1, which came up in the
S L(2, Z) situation.
m0
Proof First of all we claim the decomposition is disjoint. If not, there exists
⎛ ⎞
γ1,1 γ1,2 ··· γ1,n
⎜ γ2,1 γ2,2 ··· γ2,n ⎟
⎜ ⎟
⎜ . .. .. ⎟ ∈
⎝ .. . . ⎠
γn,1 γn,2 ··· γn,n
268 The Godement–Jacquet L-function
such that
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
γ1,1 γ1,2 · · · γ1,n c1 c1,2 · · · c1,n c1 c1,2 · · · c1,n
⎜ γ2,1 γ2,2 · · · γ2,n ⎟ ⎜ c2 · · · c2,n ⎟ ⎜ ⎟
c2 · · · c2,n
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ . .. .. ⎟ · ⎜ .. .. ⎟ = ⎜ .. .. ⎟ .
⎝ .. . . ⎠ ⎝ . . ⎠ ⎝ . . ⎠
γn,1 γn,2 · · · γn,n cn cn
(9.3.4)
Note that the above shows that c
= c
(1 ≤
≤ n). Therefore, (9.3.4) takes
the form
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 γ1,2 · · · γ1,n c1 c1,2 · · · c1,n c1 c1,2 · · · c1,n
⎜ ⎟
1 · · · γ2,n ⎟ ⎜ ⎜ ⎟
c2 · · · c2,n ⎟ ⎜ ⎜ c2 ⎟
· · · c2,n
⎜ ⎟
⎜ . .. ⎟ · ⎜ . .. ⎟ = ⎜ .. .. ⎟ .
⎝ . . . ⎠ ⎝ . . . ⎠ ⎝ . . ⎠
1 cn cn
Since 0 ≤ ci,
, ci,
< c
for 1 ≤ i <
≤ n, one concludes that γi,
= 0 for
1 ≤ i <
≤ n, and the decomposition is disjoint as claimed.
Now, by Theorem 3.11.2, every element on the right-hand side of (9.3.3) can
be put into Smith normal form, so must occur as an element on the left-hand
side of (9.3.3). Similarly, by Theorem 3.11.1, every element on the left-hand
side of (9.3.3) can be put into Hermite normal form, so must occur as an element
on the right-hand side of (9.3.3). This proves the equality of the two sides of
(9.3.3).
By analogy with the S L(2, Z) situation (see (3.12.3)), it follows that for
every integer N ≥ 1, we have a Hecke operator TN acting on the space of
square integrable automorphic forms f (z) with z ∈ hn . The action is given by
the formula
⎛⎛ ⎞ ⎞
c1 c1,2 · · · c1,n
⎜⎜ c2 · · · c2,n ⎟ ⎟
1 ⎜⎜ ⎟ ⎟
TN f (z) = n−1/2 f ⎜⎜ . . ⎟ · z ⎟ (9.3.5)
N )
n ⎝⎝ .. .. ⎠ ⎠
c
=N
=1 cn
0≤ci,
<c
(1≤i<
≤n)
9.3 Hecke operators for SL(n, Z) 269
Note the normalizing factor of 1/N (n−1)/2 which was chosen to simplify later
formulae. Clearly, T1 is just the identity operator.
The C-vector space L2 (\hn ) has a natural inner product, denoted , , and
defined by
f, g = f (z)g(z) d ∗ z,
\hn
for all f, g ∈ L2 (\hn ), where d ∗ z denotes the left invariant measure given in
Proposition 1.5.3.
In the case of S L(2, Z), we showed in Theorem 3.12.4 that the Hecke oper-
ators are self-adjoint with respect to the Petersson inner product. For S L(n, Z)
with n ≥ 3, it is no longer true that the Hecke operators are self-adjoint. What
happens is that the adjoint operator is again a Hecke operator and, therefore,
the Hecke operator commutes with its adjoint which means that it is a normal
operator.
Theorem 9.3.6 (Hecke operators are normal operators) Consider the Hecke
operators TN , (N = 1, 2, . . . ) defined in (9.3.5). Let TN∗ be the adjoint operator
which satisfies
TN f, g = f, TN∗ g
for all f, g ∈ L2 (\hn ). Then TN∗ is another Hecke operator which commutes
with TN so that TN is a normal operator. Explicitly, TN∗ is associated to the
following union of double cosets:
⎛ ⎞
N · m −1
0
( ⎜ N · (m 0 m 1 )−1 ⎟
⎜ ⎟
⎜ . ⎟ .
⎝ . . ⎠
1 ···m n−1 =N
m n0 m n−1
−1
N · (m 0 · · · m n−1 )
(9.3.7)
Proof It follows from (9.3.3), and also from the fact that transposition is an
antiautomorphism (as in the proof of Theorem 3.10.10), that
⎛ ⎞
m 0 · · · m n−1
( ⎜ .. ⎟ ( (
⎜ . ⎟
⎜ ⎟= α = α. (9.3.8)
⎝ m0m1 ⎠ α∈S α∈S
n n−1
m0 m1 ···m n−1 =N N N
m0
270 The Godement–Jacquet L-function
Since the action of the Hecke operator is independent of the choice of right
coset decomposition, we obtain
1
TN f, g = (n−1)/2 f (αz) g(z) d ∗ z
N α∈S N
\hn
1
= (n−1)/2 f (z) g(α −1 z) d ∗ z (9.3.9)
N α∈S N
\hn
⎛⎛ ⎞ ⎞
N
⎜⎜ .. ⎟ ⎟
1 ⎜⎜ . ⎟ −1 ⎟ ∗
= f (z) g ⎜⎜ ⎟α · z ⎟ d z,
N (n−1)/2 α∈S N
⎝⎝ N ⎠ ⎠
\hn
N
Finally,
⎛ if ⎞
we multiply both sides of (9.3.10) by the diagonal matrix
N
⎜ .. ⎟
⎠, with N = m n0 m 1 · · · m n−1 , it follows that the adjoint Hecke
n−1
⎝ .
N
9.3 Hecke operators for SL(n, Z) 271
operator defined by (9.3.9) is, in fact, associated to the union of double cosets
given in (9.3.7). This completes the proof.
The Hecke operators commute with the G L(n, R)-invariant differential oper-
ators, and they also commute with the operators Tδ given in (9.2.3). It follows
by standard methods in functional analysis, that we may simultaneously diago-
nalize the space L2 (S L(n, Z)\hn ) by all these operators. We shall be interested
in studying Maass forms which are eigenfunctions of the full Hecke ring of all
such operators. The following theorem is analogous to Theorem 3.12.8 which
came up in the S L(2, Z) situation.
Theorem 9.3.11 (Multiplicativity of the Fourier coefficients) Consider
∞
∞ A(m 1 , . . . , m n−1 )
φ(z) = ···
γ ∈Un−1 (Z)\S L(n−1,Z) m 1 =1 m n−2 =1 m n−1 =0
)
n−1
|m k |k(n−k)/2
k=1
⎛⎛ ⎞ ⎞
m 1 · · · |m n−1 |
⎜⎜ .. ⎟ ⎟
⎜⎜ . ⎟ γ ⎟
× WJacquet ⎜⎜ ⎟· z, ν, ψ1,...,1, mn−1 ⎟,
⎝⎝ m1 ⎠ 1 |m n−1 |⎠
1
a Maass form for S L(n, Z), as in (9.1.2). Assume that φ is an eigenfunction
of the full Hecke ring. If A(1, . . . , 1) = 0, then φ vanishes identically. Assume
φ = 0 and it is normalized so that A(1, . . . , 1) = 1. Then
Tm φ = A(m, 1, . . . , 1) · φ, ∀ m = 1, 2, . . .
Furthermore, we have the following multiplicativity relations
A m 1 m 1 , . . . , m n−1 m n−1 = A(m 1 , . . . , m n−1 ) · A m 1 , . . . , m n−1 ,
if (m 1 · · · m n−1 , m 1 · · · m n−1 ) = 1, and
m 1 cn m 2 c1 m n−1 cn−2
A(m, 1, . . . , 1)A(m 1 , . . . , m n−1 ) = A , ,...,
)
n c1 c2 cn−1
c
=m
=1
c1 |m 1 ,c2 |m 2 ,...,cn−1 |m n−1
⎛ ⎞ ⎛ ⎞
1 x1,2 x1,3 ··· x1,n y1 y2 · · · yn−1
⎜ 1 x2,3 ··· x2,n ⎟ ⎜ y1 y2 · · · yn−2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ .. .. .. ⎟ ⎜ .. ⎟
x =⎜ . . . ⎟, y=⎜ . ⎟.
⎜ ⎟ ⎜ ⎟
⎝ 1 xn−1,n ⎠ ⎝ y1 ⎠
1 1
In view of Theorem 5.3.2, Remark 9.1.1, and formula (9.1.2), we may write
for, m 1 , . . . , m n−1 ≥ 1,
1 1
··· φ(z) e−2πi (m 1 xn−1,n +m 2 xn−2,n−1 +···+m n−1 x1,2 ) d ∗ x
0 0
⎛⎛ ⎞ ⎞
m 1 y1 · · · m n−1 yn−1
A(m 1 , . . . , m n−1 ) ⎜⎜ .. ⎟ ⎟
⎜⎜ . ⎟ ⎟
= n−1 · WJacquet ⎜⎜ ⎟, ν, ψ1,...,1⎟,
) k(n−k)/2 ⎝⎝ m 1 y1 ⎠ ⎠
mk
k=1 1
(9.3.12)
)
where d ∗ x = d xi, j . While the notation adopted in Remark 9.1.1 is
1≤i< j≤n−1
useful in most cases, it makes the Hecke operator computations extremely
gruesome, so we temporarily return to our earlier notation for x.
If φ is an eigenfunction of the Hecke operator Tm defined by (9.3.5), then we
have Tm f (z) = λm f (z) for some eigenvalue λm . We can compute λm directly
using a variation of (9.3.12). We begin by considering for m = 1, 2, . . . ,
⎛⎛ ⎞ ⎞
m 1 y1 · · · m n−1 yn−1
A(m 1 , . . . , m n−1 ) ⎜⎜ .. ⎟ ⎟
⎜⎜ . ⎟ ⎟
λm n−1 · WJacquet ⎜⎜ ⎟, ν, ψ1,...,1 ⎟
) k(n−k)/2 ⎝⎝ m 1 y1 ⎠ ⎠
mk
k=1 1
m m
1
= ··· Tm φ(z) e−2πi (m 1 xn−1,n +m 2 xn−2,n−1 +···+m n−1 x1,2 ) d ∗ x
m n(n−1)/2
0 0
9.3 Hecke operators for SL(n, Z) 273
m m
1
= ···
m (n+1)(n−1)/2 )
n
c
=m 0 0
=1
0≤ci,
<c
(1≤i<
≤n)
⎛ ⎛ ⎞ ⎞
⎛ ⎞ 1 x1,2 x1,3 ··· x1,n
⎜ c1 c1,2 · · · c1,n ⎜ ⎟ ⎟
⎜⎜ 1 x2,3 ··· x2,n
⎜⎜ c2 · · · c2,n ⎟
⎟⎜
⎜
..
⎟ ⎟
⎟ ⎟
× φ ⎜⎜ ⎟ ⎜ .. .. ⎟ · y⎟
⎜⎝ .. .
.. ⎠ ⎜ . . .⎟ ⎟
⎝ . ⎝ 1 xn−1,n ⎠ ⎠
cn
1
× e−2πi (m 1 xn−1,n +m 2 xn−2,n−1 +···+m n−1 x1,2 ) d ∗ x. (9.3.13)
Next, if we let
⎛ ⎞
⎛ ⎞ 1 x1,2 x1,3 ··· x1,n
c1 c1,2 · · · c1,n ⎜ ⎟
⎜ 1 x2,3 ··· x2,n
⎜ c2 · · · c2,n ⎟
⎟⎜
⎜
..
⎟
⎟
⎜ .. ..
.. .. ⎟ ⎜ . . . ⎟
⎝ . . ⎠⎜⎝
⎟
1 xn−1,n ⎠
cn
1
⎛ ⎞
1 x1,2 x1,3 ··· x1,n ⎛ ⎞
⎜ 1
x2,3 ···
x2,n ⎟ c1
⎜ ⎟⎜ c2 ⎟
⎜ .. .. .. ⎟⎜ ⎟
=⎜ . . . ⎟⎜ .. ⎟,
⎜ ⎟⎝ . ⎠
⎝ 1
xn−1,n ⎠
cn
1
1
j
xi, j = ci,k xk, j , (9.3.14)
c j k=i
with the understanding that ci,i = ci and xi,i = 1 for i = 1, 2, . . . , n. Note the
special case:
ci xi,i+1 + ci,i+1
xi,i+1 = (i = 1, 2, . . . , n − 1), (9.3.15)
ci+1
of (9.3.14).
274 The Godement–Jacquet L-function
With the computations (9.3.14), (9.3.15), the right-hand side of the identity
(9.3.13) becomes
ci m +
j ci,k xk, j
cj + cj
k=i+1
1
m (n+1)(n−1)/2 )
n 1≤i< j≤n
c
=m +
j ci,k xk, j
=1 cj
0≤ci,
<c
(1≤i<
≤n) k=i+1
⎛⎛ ⎞ ⎞
1 x1,2 x1,3 ··· x1,n ⎛ ⎞
⎜⎜ ⎟ c1 ⎟
⎜⎜ 1 x2,3 ··· x2,n ⎟⎜ ⎟ ⎟
⎜⎜ .. ⎟⎜ c2 ⎟ ⎟
× φ ⎜⎜ .. .. ⎟⎜ ⎟ · y⎟
⎜⎜ . . . ⎟⎝ .. ⎠ ⎟
⎝⎝ ⎠ . ⎠
1 xn−1,n
cn
1
+ cr,r +1 m n−r
n−1 +
n−1
cr +1 m n−r
2πi −2πi xr,r +1 cj
×e ·e · d xi, j .
cr cr
r =1 r =1
ci
In view of the periodicity of the above integrand, we may deduce that it takes
the form:
ci m
c j
1
m (n+1)(n−1)/2 )
n 1≤i< j≤n
c
=m 0
=1
0≤ci,
<c
(1≤i<
≤n)
⎛⎛ ⎞ ⎞
1 x1,2 x1,3 ··· x1,n ⎛ ⎞
⎜⎜ ⎟ c1 ⎟
⎜⎜ 1 x2,3 ··· x2,n ⎟⎜ ⎟ ⎟
⎜⎜ .. ⎟⎜ c2 ⎟ ⎟
× φ ⎜⎜ .. .. ⎟⎜ ⎟ · y⎟
⎜⎜ . . . ⎟⎝ .. ⎠ ⎟
⎝⎝ ⎠ . ⎠
1 xn−1,n
cn
1
+ cr,r +1 m n−r
n−1 +
n−1
cr +1 m n−r
2πi −2πi xr,r +1 cj
×e ·e · d xi, j .
cr cr
r =1 r =1
ci
⎧)
n
2πi
+ cr,r +1 m n−r
n−1
⎨ ct−1 if cr | m n−r (1 ≤ r < n),
t
=
cr
e r =1 t=2
0≤ci,
<c
(1≤i<
≤n)
⎩
0 otherwise.
9.3 Hecke operators for SL(n, Z) 275
Consequently, it follows that our integral (9.3.13) may be written in the form:
ci m
n c j
1
ctt−1
m (n+1)(n−1)/2 )
n t=2 1≤i< j≤n
c
=m 0
=1
⎛⎛ ⎞ ⎞
1 x1,2 x1,3 ··· x1,n ⎛ ⎞
⎜⎜ ⎟ c1 ⎟
⎜⎜ 1 x2,3 ··· x2,n ⎟⎜ ⎟ ⎟
⎜⎜ .. ⎟⎜ c2 ⎟ ⎟
× φ ⎜⎜ .. .. ⎟⎜ ⎟ · y⎟
⎜⎜ . . . ⎟⎝ .. ⎠ ⎟
⎝⎝ ⎠ . ⎠
1 xn−1,n
cn
1
+
n−1
cr +1 m n−r
−2πi xr,r +1 cj
×e · d xi, j .
cr
r =1
ci
Further, if F : R → C is a periodic
*M function satisfying
*1 F(x + 1) = F(x) and
M is a positive integer, then 0 F(x) d x = M · 0 F(x) d x. Consequently, the
integral above is equal to
n 1
1
ctt−1
m (n−1)/2 )
n t=2 1≤i< j≤n
c
=m 0
=1
⎛⎛ ⎞ ⎞
1 x1,2 x1,3 ··· x1,n ⎛ ⎞
⎜⎜ ⎟ c1 ⎟
⎜⎜ 1 x2,3 ··· x2,n ⎟⎜ ⎟ ⎟
⎜⎜ .. ⎟⎜ c2 ⎟ ⎟
× φ ⎜⎜ .. .. ⎟⎜ ⎟ · y⎟
⎜⎜ . . . ⎟⎝ .. ⎠ ⎟
⎝⎝ ⎠ . ⎠
1 xn−1,n
cn
1
+
n−1
cr +1 m n−r
−2πi xr,r +1
×e d xi, j .
cr
r =1 (9.3.16)
) *1
Finally, the multiple integral, · · · , above can be evaluated with
1≤i< j≤n 0
(9.3.12) and has the value
A mcn−1
1 cn
, mc2n−2
cn−1
, . . . , m n−1
c1
c2
)
n−1 k(n−k)/2
cn+1−k m k
cn−k
k=1
⎛⎛ ⎞ ⎞
m 1 y1 · · · m n−1 yn−1 cn
⎜⎜ .. ⎟ ⎟
⎜⎜ . ⎟ ⎟
× WJacquet ⎜⎜ ⎟, ν, ψ1,...,1 ⎟ ,
⎝⎝ m 1 y1 cn ⎠ ⎠
cn
276 The Godement–Jacquet L-function
= ctt−1 ·
m (n−1)/2 )
n t=2
)
n−1
cn+1−k m k
k(n−k)/2
c
=m cn−k
=1 k=1
⎛⎛ ⎞ ⎞
m 1 y1 · · · m n−1 yn−1 cn
⎜⎜ .. ⎟ ⎟
⎜⎜ . ⎟ ⎟
× WJacquet ⎜⎜ ⎟, ν, ψ1,...,1 ⎟ .
⎝⎝ m 1 y1 cn ⎠ ⎠
cn
Note that we may cancel the Whittaker functions on both sides of the above
identity because the Whittaker functions are invariant under multiplication by
scalar matrices. Further, one easily checks the identity
)
n
ctt−1
t=2
k(n−k)/2 = (c1 · c2 · · · cn ) = m (n−1)/2 .
(n−1)/2
)
n−1
cn+1−k
cn−k
k=1
for all non-negative integers i 1 , . . . , i n−1 . One may then proceed to products
of two primes, products of three primes, etc. to eventually obtain that if
A(1, . . . , 1) = 0 then all coefficients A(m 1 , . . . , m n−1 ) must vanish.
If f = 0 then we may assume it is normalized so that A(1, . . . , 1) = 1.
If we now choose m 1 = m 2 = · · · = m n−1 = 1, it immediately follows from
(9.3.17) that λm = A(m, 1, . . . , 1). Substituting this into (9.3.17), and changing
indices (on the c j s), proves the identity
1 n−2
· (−1)r A( p k−r , 1, . . . , 1)A( 1, . . . , p , . . . , 1)
p ks r =0 " #$ %
position r +1
A( p k+1
, 1, . . . , 1) + (−1) n−2
A( p k−n+1
, 1, . . . , 1)
= . (9.4.1)
p ks
9.4 The Godement–Jacquet L-function 279
If we define
∞
A( p k , 1, . . . , 1)
φ p (s) := ,
k=0
p ks
Definition 9.4.3 Let s ∈ C with (s) > (n + 1)/2, and let f (z) be a Maass
form for S L(n, Z), with n ≥ 2, which is an eigenfunction of all the Hecke
operators as in Theorem 9.3.11. We define the Godement–Jacquet L-function
L f (s) (termed the L-function associated to f ) by the absolutely convergent
series
∞
L f (s) = A(m, 1, . . . , 1)m −s = φ p (s),
m=1 p
Remark It is clear that the L-function associated to the dual Maass form f˜
takes the form
∞
L f˜ (s) = A(1, . . . , 1, m)m −s .
m=1
Set
⎛ ⎞
m 1 · · · m n−2 · |m n−1 |
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
M =⎜ m1m2 ⎟,
⎜ ⎟
⎝ m1 ⎠
1
and z = x y with
⎛ ⎞
1 xn−1 x1,3 ··· x1,n
⎜ 1 xn−2 ··· x2,n ⎟
⎜ ⎟
⎜ .. .. .. ⎟ ,
x =⎜ . . . ⎟
⎜ ⎟
⎝ 1 x1 ⎠
1
⎛ ⎞
y1 y2 · · · yn−1
⎜ y1 y2 · · · yn−2 ⎟
⎜ ⎟
⎜ .. ⎟
y=⎜ . ⎟.
⎜ ⎟
⎝ y1 ⎠
1
A simple computation gives
⎛ ⎞
1 |m n−1 |xn−1
⎜ ∗ ⎟
⎜ ⎟
⎜ ⎟
⎜ .. .. ⎟
M·x =⎜ ⎜
. . ⎟ · M.
⎟
⎜ 1 m 2 x2 ⎟
⎜ ⎟
⎝ 1 m 1 x1 ⎠
1
It follows from Definition 5.4.1 (2), that for any integers 1 , . . . , n−1 , the
Jacquet Whittaker function satisfies
WJacquet (M z, ν, ψ1 ,...,n−1 ) = e2πi [m 1 1 x1 +···+m n−2 n−2 xn−2 +|m n−1 |n−1 xn−1 ]
× WJacquet (M · y, ν, ψ1 ,...,n−1 ).
γ
we may put the G L(n, R) matrix, · z, into Iwasawa form:
1
⎛ γ γ γ ⎞
1 xn−1 x1,3 · · · x1,n
⎜ γ γ ⎟
⎜ 1 xn−2 · · · x2,n ⎟
γ ⎜ ⎟
·z ≡⎜⎜
..
.
..
.
.. ⎟
⎟
1 ⎜ .
⎝ γ ⎟
1 x1 ⎠
1
⎛ γ γ γ ⎞
y1 y2 · · · yn−1
⎜ γ γ γ ⎟
⎜ y1 y2 · · · yn−2 ⎟
⎜ ⎟
×⎜⎜ . ⎟ (mod Z n O(n, R)),
.. ⎟
⎜ ⎟
⎝ y
γ ⎠
1
1
(9.4.5)
γ
where x1 = an−1,1 x1,n + an−1,2 x2,n + · · · + an−1,n−1 x1 .
It immediately follows from Proposition 5.5.2, (9.4.4) and (9.4.5) that we
may write
γ
WJacquet M · · z, ν, ψ1 ,...,n−1
1
3 γ γ
4
= e2πi m 1 1 (an−1,1 x1,n +an−1,2 x2,n +···+an−1,n−1 x1 )+m 2 2 x2 +···+|m n−1 |n−1 xn−1
⎛ ⎛ γ γ ⎞ ⎞
y1 · · · yn−1
⎜ ⎜ .. ⎟ ⎟
⎜ ⎜ . ⎟ ⎟
× WJacquet ⎜ M · ⎜ ⎟ , ν, ψ 1,...,1 ⎟ . (9.4.6)
⎝ ⎝ y
γ ⎠ ⎠
1
1
Finally, we obtain the following theorem which is the basis for the construc-
tion of the L-function L f (s) (given in Definition 9.4.3) as a Mellin transform.
Theorem 9.4.7 Let f (z) be a Maass form of type ν for S L(n, Z) as in (9.1.2).
Then we have the representation,
∞
∞ A(m 1 , . . . , m n−1 )
f (z) = ···
γ ∈Un−1 (Z)\S L(n−1,Z) m 1 =1 m n−2 =1 m n−1 =0
)
n−1
|m k |k(n−k)/2
k=1
γ γ
2πim 1 (an−1,1 x1,n +···+an−1,n−1 x1 ) 2πi (m 2 x2 +···+m n−1 xn−1 )
×e e
× WJacquet (M · y γ , ν, ψ1,...,1 ),
282 The Godement–Jacquet L-function
γ
where γ is given by (9.4.4) and, x γ , y γ , are defined by: z ≡ x γ · yγ ,
1
as in (9.4.5).
Proof The proof follows from the Fourier expansion (9.1.2) and the identity
(9.4.6).
Corollary 9.4.8 Let f (z) be a Maass form of type ν for S L(n, Z) as in (9.1.2).
Then
⎛⎛ ⎞ ⎞
1 0 · · · 0 u 1,n
1 1 ⎜ ⎜⎜
⎜ ..
.
.. .
. ..
.. ⎟ ⎟
. ⎟ ⎟
n−2
⎜⎜ ⎟ ⎟ −2πiu 1
··· f ⎜⎜ 1 0 u ⎟ · z ⎟ e du 1 du j,n
⎜⎜ n−2,n ⎟ ⎟
0 0 ⎝ ⎝ 1 u1 ⎠ ⎠ j=1
1
∞
∞
= ···
γ ∈Un−2 (Z)\S L(n−2,Z) m 2 =1 m n−2 =1 m n−1 =0
A(1, m 2 , . . . , m n−1 ) 2πi (x1 +m 2 x2γ +···+m n−1 xn−1
γ
)
× n−1 e
)
|m k |k(n−k)/2
k=2
⎛⎛ γ γ ⎞ ⎞
m 2 · · · |m n−1 |y1 y2 · · · yn−1
⎜⎜ .. ⎟ ⎟
⎜⎜ . ⎟ ⎟
⎜⎜ ⎟ ⎟
× WJacquet ⎜⎜ m
γ
2 y1 y2
⎟, ν, ψ1,...,1 ⎟ .
⎜⎜ ⎟ ⎟
⎝⎝ y1 ⎠ ⎠
1
1
9.4 The Godement–Jacquet L-function 283
1 1
n−2
2πim 1 an−1,1 u 1,n +···+an−1,n−2 u n−2,n +an−1,n−1 u 1
··· e e−2πiu 1 du 1 du j,n
j=1
0 0
The proof of Corollary 9.4.8 follows from this after noting that
⎛ ⎞
a1,1 ··· a1,n−2 a1,n−1
⎜ a2,1 ··· a2,n−2 a2,n−1 ⎟
⎜ ⎟
⎜ .. .. .. .. ⎟
γ =⎜ . . . . ⎟ ∈ Un−1 (Z)\S L(n − 1, Z)
⎜ ⎟
⎝ an−2,1 an−2,n−2 an−2,n−1 ⎠
0 ··· 0 1
⎛ ⎞
a1,1 ··· a1,n−2 0
⎜ a2,1 ··· a2,n−2 0⎟
⎜ ⎟
⎜ .. .. .. .. ⎟ ∈ U (Z)\S L(n − 2, Z).
γ =⎜ . . . .⎟ n−2
⎜ ⎟
⎝ an−2,1 an−2,n−2 0 ⎠
0 ··· 0 1
Theorem 9.4.9 Let f (z) be a Maass form of type ν for S L(n, Z) as in (9.1.2).
Then for
⎛ ⎞
1 0 u 1,3 ··· u 1,n
⎜ 1 u n−2 ··· u 2,n ⎟
⎜ ⎟
n−2
⎜ .. .. .. ⎟ ,
û = ⎜ . . . ⎟ d ∗ û = uj du i, j ,
⎜ ⎟
⎝ 1 u1 ⎠
j=1 2≤i< j−1≤n−1
1
284 The Godement–Jacquet L-function
we have
1 1
··· f (ûz)e−2πi (u 1 +···+u n−2 ) d ∗ û
0 0
A(1, . . . , 1, m)
= e2πimxn−1 e2πi(x1 +···+xn−2 )
m=0
|m| n−1/2
⎛⎛ ⎞ ⎞
|m|
⎜⎜ 1 ⎟ ⎟
⎜⎜ ⎟ ⎟
× WJacquet ⎜⎜ .. ⎟ · y; ν, ψ1,...,1 ⎟ .
⎝⎝ . ⎠ ⎠
1
The modern theory of Eisenstein series began with Maass (1949), who formally
defined and studied the series (see Section 3.1)
1 ys
E(z, s) = ,
2 (c,d)=1 |cz + d|2s
285
286 Langlands Eisenstein series
Example 10.1.2 (Parabolic subgroups of G L(3, R)) There are three stan-
dard parabolic subgroups of G L(3, R) corresponding to the three partitions:
3 = 1 + 1 + 1, 3 = 1 + 2, 3 = 2 + 1.
Explicitly, we have
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎨ ∗ ∗ ∗ ⎬ ⎨ ∗ ∗ ∗ ⎬
P1,1,1 = ⎝ 0 ∗ ∗ ⎠ , P1,2 = ⎝0 ∗ ∗⎠ ,
⎩ ⎭ ⎩ ⎭
0 0 ∗ 0 ∗ ∗
⎧⎛ ⎞⎫
⎨ ∗ ∗ ∗ ⎬
P2,1 = ⎝ ∗ ∗ ∗ ⎠ .
⎩ ⎭
0 0 ∗
Example 10.1.3 (Parabolic subgroups of G L(4, R)) There are seven stan-
dard parabolic subgroups of G L(4, R) corresponding to the seven partitions:
4 = 1 + 1 + 1 + 1, 4 = 1 + 1 + 2, 4 = 1 + 2 + 1,
4 = 2 + 1 + 1, 4 = 1 + 3, 4 = 2 + 2, 4 = 3 + 1.
Explicitly, we have
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎪
⎪ ∗ ∗ ∗ ∗ ⎪⎪ ⎪
⎪ ∗ ∗ ∗
∗ ⎪⎪
⎨⎜ ⎬ ⎨⎜ ⎬
0 ∗ ∗ ∗⎟ 0 ∗ ∗⎟
∗
P1,1,1,1 = ⎜ ⎝
⎟ , P1,1,2 = ⎜ ⎟ ,
⎪
⎪ 0 0 ∗ ∗ ⎠⎪
⎪ ⎪
⎪
⎝ 0 0 ∗ ⎠⎪
∗ ⎪
⎩ ⎭ ⎩ ⎭
0 0 0 ∗ 0 0 ∗∗
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎪
⎪ ∗ ∗ ∗ ∗ ⎪⎪ ⎪
⎪ ∗ ∗ ∗ ∗ ⎪⎪
⎨⎜ ⎬ ⎨⎜ ⎬
0 ∗ ∗ ∗⎟ ∗ ∗ ∗ ∗⎟
P1,2,1 = ⎜ ⎝
⎟ , P2,1,1 = ⎜ ⎟ ,
⎪
⎪ 0 ∗ ∗ ∗ ⎠⎪
⎪ ⎪
⎪
⎝ 0 0 ∗ ∗ ⎠⎪
⎪
⎩ ⎭ ⎩ ⎭
0 0 0 ∗ 0 0 0 ∗
288 Langlands Eisenstein series
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎪
⎪ ∗ ∗ ∗ ∗ ⎪⎪ ⎪
⎪ ∗ ∗ ∗ ∗ ⎪ ⎪
⎨⎜ ⎬ ⎨⎜ ⎬
0 ∗ ∗ ∗⎟ ∗ ∗ ∗ ∗⎟
P1,3 = ⎜⎝0 ∗
⎟ , P3,1 = ⎝⎜ ⎟ ,
⎪
⎪ ∗ ∗ ⎠⎪
⎪ ⎪
⎪ ∗ ∗ ∗ ∗ ⎠⎪⎪
⎩ ⎭ ⎩ ⎭
0 ∗ ∗ ∗ 0 0 0 ∗
⎧⎛ ⎞⎫
⎪
⎪ ∗ ∗ ∗ ∗ ⎪⎪
⎨⎜ ⎬
∗ ∗ ∗ ∗⎟
P2,2 = ⎜⎝0 0
⎟ .
⎪
⎪ ∗ ∗ ⎠⎪
⎪
⎩ ⎭
0 0 ∗ ∗
Definition 10.1.4 (Associate parabolics) Fix integers n ≥ 2, 1 < r ≤ n.
Two standard parabolic subgroups Pn 1 ,...,nr , Pn 1 ,...,Pn r of G L(n, R), corre-
sponding to the partitions,
n = n 1 + · · · + n r = n 1 + · · · + n r ,
n = n 1 + · · · + n r = n 1 + · · · + n r .
where
⎧⎛ ⎞⎫
⎪ In 1 ∗ ··· ∗ ⎪
⎪
⎪ ⎪
⎪⎜ 0
⎪ ··· ∗ ⎟⎪⎪
⎨⎜ In 2 ⎟⎪⎬
⎜ ··· ∗ ⎟
Nn 1 ,...,nr = ⎜ 0 0 ⎟ (Ik = k × k Identity matrix),
⎪⎜ .
⎪ ⎟⎪
⎪
⎪ ⎝ .. .. ..
.
.. ⎠⎪⎪
⎪
⎪
⎩
. . ⎪
⎭
0 0 ··· In r
is the unipotent radical and
⎧⎛ ⎞⎫
⎪ mn 1 0 ··· 0 ⎪
⎪
⎪ ⎪
⎪⎜ 0 mn 2
⎪ ··· ⎟⎪
⎪
⎨⎜ 0 ⎟⎪
⎬
⎜ ··· ⎟
Mn 1 ,...,nr = ⎜ 0 0 0 ⎟ (mk ∈ G L(k, R)),
⎪
⎪ ⎜ . ⎟⎪
⎪
⎪ ⎝ .. .. ..
.
.. ⎠⎪
⎪
⎪
⎪
⎩
. . ⎪
⎭
0 0 · · · mnr
is the so-called Levi component. The Levi component further decomposes into
the direct product
of G L(3, R).
⎧⎛ ⎞⎫
⎨ ∗ ∗ ∗ ⎬
P1,1,1 = ⎝ 0 ∗ ∗⎠
⎩ ⎭
0 0 ∗
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎫
⎨ 1 ∗ ∗ t1 0 0 ±1 0 0 ⎬
= ⎝0 1 ∗⎠ · ⎝ 0 t2 0⎠·⎝ 0 ±1 0⎠ ,
⎩ ⎭
0 0 1 0 0 t3 0 0 ±1
(t1 , t2 , t3 > 0)
⎧⎛ ⎞⎫
⎨ ∗ ∗ ∗ ⎬
P1,2 = ⎝0 ∗ ∗⎠
⎩ ⎭
0 ∗ ∗
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎫
⎨ 1 ∗ ∗ t1 0 0 ±1 0 0 ⎬
= ⎝0 1 0⎠ · ⎝ 0 t2 0⎠·⎝ 0 a b⎠ ,
⎩ ⎭
0 0 1 0 0 t2 0 c d
(t1 , t2 > 0, ad − bc = ±1)
⎧⎛ ⎞⎫
⎨ ∗ ∗ ∗ ⎬
P2,1 = ⎝∗ ∗ ∗⎠
⎩ ⎭
0 0 ∗
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎫
⎨ 1 0 ∗ t1 0 0 a b 0 ⎬
= ⎝ ⎠
0 1 ∗ · 0⎝ t1 0⎠·⎝c d 0⎠ ,
⎩ ⎭
0 0 1 0 0 t2 0 0 ±1
(t1 , t2 > 0, ad − bc = ±1).
Example 10.2.6 (Langlands decomposition for P1,3,2 ) Finally, we shall give
an example of the Langlands decomposition for the parabolic subgroup P1,3,2
of G L(6, R). We have
⎧⎛ ⎞⎫
⎪
⎪ ∗ ∗ ∗ ∗ ∗ ∗ ⎪ ⎪
⎪
⎪⎜ ⎟⎪⎪
⎪
⎪ ⎜ ∗ ∗ ∗ ∗ ∗ ⎟ ⎪
⎪
⎪
⎨⎜ ⎟⎪⎬
⎜ ∗ ∗ ∗ ∗ ∗⎟
P1,3,2 = ⎜ ⎟
⎪
⎪ ⎜ ∗ ∗ ∗ ∗ ∗ ⎟⎪
⎪⎜
⎪ ⎟⎪⎪
⎪
⎪ ⎝ ∗ ∗ ⎠⎪⎪
⎪
⎪
⎩ ⎪
⎭
∗ ∗
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎪
⎪ 1 ∗ ∗ ∗ ∗ ∗ ⎪ ⎪ ⎪
⎪ ∗ ⎪
⎪
⎪
⎪ ⎜ ⎪
⎟⎪ ⎪
⎪ ⎜ ⎟⎪⎪
⎪
⎪ ⎜ 1 0 0 ∗ ∗ ⎪
⎟⎪ ⎪
⎪ ⎜ ∗ ∗ ∗ ⎟⎪⎪
⎪
⎨⎜ ⎪
⎟⎬ ⎨⎜⎪ ⎟⎪⎬
⎜ 0 1 0 ∗ ∗⎟ ⎜ ∗ ∗ ∗ ⎟
= ⎜ ⎟ · ⎜ ⎟ ,
⎪
⎪ ⎜ 0 0 1 ∗ ∗ ⎟⎪ ⎪ ⎜ ∗ ∗ ∗ ⎟⎪
⎪⎜
⎪ ⎟⎪
⎪
⎪
⎪
⎪
⎪ ⎜ ⎟⎪⎪
⎪
⎪
⎪ ⎝ 1 0 ⎪
⎠⎪ ⎪⎝⎪ ∗ ∗ ⎠⎪⎪
⎪
⎩ ⎪
⎭ ⎩ ⎪ ⎪
⎭
0 1 ∗ ∗
unipotent radical Levi component
10.2 Langlands decomposition of parabolic subgroups 291
where t1 , t2 , t3 > 0 and the block matrices in M1,3,2 have determinant ±1.
The parabolic subgroups of G L(n, R) can be characterized as stabilizers of
flags on Rn . A flag of Rn is a sequence of subspaces:
φ ⊂ V1 ⊂ V2 · · · ⊂ Vr = Rn
where ⊂ denotes a proper subset, and φ is the empty set. The action of G L(n, R)
on a flag (V1 , . . . , Vr ) is defined in the canonical way. That is if g ∈ G L(n, R)
then the action is given by g(V1 , . . . , Vr ) = (gV1 , . . . , gVr ), where the action
of g on an element (a1 , . . . , an ) ∈ Rn is given by matrix multiplication
g · t(a1 , . . . , an ). The standard complete flag is φ ⊂ V1 ⊂ · · · ⊂ Vn = Rn
where Vi = Re1 ⊕ · · · ⊕ Rei and ei is the vector (of length n) with a 1 in
the ith position and zeros elsewhere. The stabilizer of a subflag (Vd1 , . . . , Vdr )
(with 0 < d1 < · · · < dr = n) of the standard flag has the form
⎛ ⎞
m11 m12 m13 · · ·
⎜ 0 m22 m23 · · · ⎟
⎜ ⎟
⎜ 0 0 m33 · · · ⎟ , (10.2.7)
⎝ ⎠
.. .. .. ..
. . . .
V0 to be the empty set. In the case that P takes the form (10.2.7), we have
gi = mii . The unipotent radical N P of P is the subgroup of all g ∈ P so that
gi is the identity on Vi /Vi−1 for every i. For P of the form (10.2.7), mii must
be the identity matrix for every i. In a similar manner we may define the Levi
component using flags. For each 1 ≤ i ≤ r choose a complementary subspace
X i ⊂ Rn so that Vi = Vi−1 ⊕ X i . The Levi component M P of P is defined to
be the subgroup of P consisting of all g ∈ P which stablize each X i . In the
case that P is of the form (10.2.7), the Levi component requires that mi j = 0
for 1 ≤ i < j ≤ n.
Example 10.3.1 (Weyl group for G L(3, R)) The Weyl group W3 is the group
of six elements:
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 1 0 0 0 1 0
⎝0 1 0⎠, ⎝0 0 1⎠, ⎝1 0 0⎠,
0 0 1 0 1 0 0 0 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 0 0 1 0 0 1
⎝0 0 1⎠, ⎝1 0 0⎠, ⎝0 1 0⎠.
1 0 0 0 1 0 1 0 0
Recall that the standard Borel subgroup Bn of G L(n, R) is the group of invertible
upper triangular matrices.
G L(n, R) = Bn Wn Bn .
Proof Let
⎛ ⎞
g11 g12 ··· g1n
⎜ g21 g22 ··· g2n ⎟
⎜ ⎟
g=⎜ . .. .. ⎟ ∈ G L(n, R).
⎝ .. . ··· . ⎠
gn1 gn2 ··· gnn
10.3 Bruhat decomposition 293
Let gn
denote the first non-zero entry in the bottom row of g. Then by right
multiplication by some b1 ∈ Bn , we can change this entry to 1 and make the
rest of the bottom row 0. The matrix gb1 now takes the form
⎛ ⎞
g11 ··· g1
··· g1n
⎜ g21
···
g2
···
g2n ⎟
⎜ ⎟
⎜ .. . . ⎟
gb1 = ⎜ . ··· .
. ··· . ⎟
. .
⎜ ⎟
⎝ g
· · · gn−1,
· · · gn−1,n ⎠
n−1,1
0 ··· 1 ··· 0
If we now multiply gb1 on the left by some matrix
⎛ ⎞
b11 b12 b13 · · · b1n
⎜ 0 b22
b23 ⎟
· · · b2n
⎜ ⎟
⎜ ⎟
· · · b3n
b1 = ⎜ 0 0 b33 ⎟ ∈ Bn ,
⎜ . .. .. .. .. ⎟
⎝ .. . . . . ⎠
0 0 0 ··· 1
it is easy to see that we may choose the bi, j (1 ≤ i < j ≤ n) so that b1 gb1 takes
the form
⎛ ⎞
g11 ··· 0 ··· g1n
⎜ g21
··· 0 ···
g2n ⎟
⎜ ⎟
⎜ .. . . ⎟
b1 gb1 = ⎜ . ··· . ···. . ⎟
. .
⎜ ⎟
⎝ g ··· 0 ··· g ⎠
n−1,1 n−1,n
0 ··· 1 ··· 0
We call this clearing the (n)th row and (
)th column.
We next consider the first non-zero entry in the (n − 1)st row of b1 gb1 .
Suppose it is gn−1,
1
.We may again multiply b1 gb1 on the right by some element
b2 ∈ Bn so that we change this entry to 1 and make all other entries in the
(n − 1)st row 0. By left multiplication by some b2 we can make all the other
entries in the (
1 )st column 0 which results in clearing the (n − 1)st row and
(
1 )st column.
Continuing in this manner, we obtain a set of n matrices
b1 gb1 , b2 b1 gb1 b2 , b3 b2 b1 gb1 b2 b3 , ... , bn · · · b3 b2 b1 gb1 b2 b3 · · · bn
g ∈ G L(n, R), define Mλ (g) to be the k × k minor of g formed from the bottom
k rows and the columns
1 , . . . ,
k , indexed by the elements of λ. We may
express Mλ (g) using wedge products of e1 , e2 , . . . , en , where ei denotes the
column vector of length n with a 1 at position i and zeros elsewhere. We
have
Mλ (g)e1 ∧ · · · ∧ en = e1 ∧ · · · ∧ en−k ∧ g · e
1 ∧ · · · ∧ g · e
k . (10.3.3)
Proof It easily follows from Proposition 10.3.2 that every g ∈ G L(n, R) has a
Bruhat decompostion g = u 1 cwu 2 with u 1 , u 2 ∈ Un , w ∈ Wn , and c a diagonal
matrix. To determine c, we utilize (10.3.3) and write
We now write u 2 = (µi, j ) (where µi, j denotes the i, j entry of the matrix
u 2 ). Note that for any 1 ≤
≤ n,
u 2 e
= u r,
er .
r =1
with
ij if i + j ≤ n,
bi, j =
(n − i)(n − j) if i + j ≥ n,
z → p · z
Definition 10.4.2 The series (10.4.1) is called the minimal parabolic Eisen-
stein series for .
Proof The fact that E Pmin (z, s) is invariant under is easy to prove because
the function is formed as a sum over a coset of the group . For the absolute
convergence, we follow Godement (see (Borel, 1966)). It is enough to show
that for every point z 0 ∈ \hn and some (non-zero volume) compact subset C z0
of \hn (with z 0 ∈ C z0 ), that the integral
E P (z, s)d ∗ z
min
C z0
10.4 Minimal, maximal, and general parabolic Eisenstein series 297
converges. Now, it follows from Proposition 1.3.2 that there will be only finitely
many γ ∈ Pmin \ such that γ z 0 ∈ √3 , 1 . By a continuity argument, one may
2 2
deduce, for sufficiently small C z0 , that there are only finitely many γ ∈ Pmin \
such that γ z ∈ √3 , 1 for all z ∈ C z0 . We immediately deduce that there exists
√ 2 2
n−1
× yk−k(n−k)−1 dyk ,
k=1
Pn 1 ,...,nr /(K · R× ) → C,
by the formula
r
Is (g, Pn 1 ,...,nr ) = Det mn (g) si ,
i
i=1
for all
⎛ ⎞ ⎛ ⎞
In 1 ∗ ··· ∗ mn 1 (g) 0 ··· 0
⎜ 0 In 2 ··· ∗ ⎟ ⎜ 0 mn 2 (g) · · · 0 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ··· ∗ ⎟ ⎜ 0 ··· 0 ⎟
g=⎜ 0 0 ⎟·⎜ 0 ⎟ ∈ Pn 1 ,...,nr .
⎜ . .. .. .. ⎟ ⎜ . .. .. .. ⎟
⎝ .. . . . ⎠ ⎝ .. . . . ⎠
0 0 ··· In r 0 0 · · · mnr (g)
Remarks Here mni is the whole group G L(n i , R) while mni (g) is a particular
+
r
element in this group. The condition n i si = 0 insures that
i=1
for any matrix δ of the form δ = t · In with t ∈ R× , so that Is (∗, Pn 1 ,...,nr ) is well
defined on Pn 1 ,...,nr /R× . It is also clear that Is (gk, Pn 1 ,...,nr ) = Is (g, Pn 1 ,...,nr ) for
k ∈ K since the determinant of an orthogonal matrix has absolute value 1.
10.4 Minimal, maximal, and general parabolic Eisenstein series 299
then z ∈ Pn 1 ,...,nr , since the standard Borel subgroup lies in every parabolic
subgroup Pn 1 ,...,nr . It follows that
s1
s2
n 1
n−n
Is (z, Pn 1 ,...,nr ) = Y j1 · Y j2
j1 =n−n 1 +1 j2 =n−n 1 −n 2 +1
s3
sr
n−n 1 −n 2
nr
× Y j3 ··· Y jr .
j3 =n−n 1 −n 2 −n 3 +1 jr =1
One easily checks that Is (z, Pn 1 ,...,nr ) is precisely the standard function Is (z)
for suitable choice of s depending on s.
we have
1 s1
Is (z, P1,2 ) = (y1 y2 )s1 · y1s2 = y12 y2 .
m P : P → M,
by the formulae
⎛ ⎞
mn 1 (g) 0 ··· 0
⎜ 0 mn 2 (g) · · · 0 ⎟
⎜ ⎟
⎜ ··· 0 ⎟
m P (g) = ⎜ 0 0 ⎟,
⎜ . .. .. .. ⎟
⎝ .. . . . ⎠
0 0 · · · mnr (g)
for all
⎛ ⎞ ⎛ ⎞
In 1 ∗ ··· ∗ mn 1 (g) 0 ··· 0
⎜ 0 In 2 ··· ∗ ⎟ ⎜ 0 mn 2 (g) ··· 0 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ··· ∗ ⎟ ⎜ 0 ··· ⎟
g =⎜ 0 0 ⎟·⎜ 0 0 ⎟ ∈ P.
⎜ . .. .. .. ⎟ ⎜ . .. .. .. ⎟
⎝ .. . . . ⎠ ⎝ .. . . . ⎠
0 0 ··· In r 0 0 · · · mnr (g)
(10.5.1)
We may now define the Langlands Eisenstein series twisted by Maass forms of
lower rank.
and put = S L(n, Z). Then for φ a cusp form on M and z ∈ hn , we define the
Eisenstein series E P (z, s, φ) by the infinite series
E P (z, s, φ) = φ (m P (γ z)) · Is (γ z, P),
γ ∈P∩\
where
ψm (u) = e2πi(m 1 u 1,2 +···+m n−1 u n−1,n ) ,
and U denotes the group of upper triangular matrices with 1s on the diagonal.
The computation of this Fourier coefficient is based on the explicit Bruhat
decomposition given in Proposition 10.3.6:
(
G L(n, R) = Gw,
w∈W
where W denotes the Weyl group and
G w = U DwU = U w DU,
with U the group of upper triangular matrices with 1s on the diagonal (as
above), and where D denotes the multiplicative group of diagonal matrices
with non-zero determinant. Consider Pmin (Z) = Pmin ∩ S L(n, Z).
The minimal parabolic Eisenstein series E Pmin (z, s) is constructed as a sum
over the left quotient space Pmin (Z)\S L(n, Z). By the Bruhat decomposition,
we may realize this left quotient space as a union:
( <
Pmin (Z)\S L(n, Z) = Pmin (Z) (S L(n, Z) ∩ G w ).
w∈W
Lemma 10.6.3 The group w = w−1 · t Pmin (Z) · w ∩ Pmin (Z) acts properly
on the right on the left coset space Pmin (Z)\(S L(n, Z) ∩ G w ).
Proof For each w ∈ W , we introduce two additional spaces.
Uw = (w −1 · U · w) ∩ U
(10.6.4)
Ūw = (w −1 · t U · w) ∩ U.
To get a feel for these spaces, consider the example of G L(3, R) where we have
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 ∗ ∗ 1 1
Uw1 = ⎝ 1 ∗ ⎠ ∩ U, Ūw1 = ⎝ ∗ 1 ⎠ ∩ U, w1 = ⎝ 1 ⎠,
1 ∗ ∗ 1 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 ∗ ∗ 1 1
Uw2 = ⎝ 1 ⎠ ∩ U, Ūw2 = ⎝ ∗ 1 ∗ ⎠ ∩ U, w2 = ⎝ 1⎠,
∗ 1 ∗ 1 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 ∗ 1 ∗ 1
Uw3 = ⎝ ∗ 1 ∗ ⎠ ∩ U, Ūw3 = ⎝ 1 ⎠ ∩ U, w3 = ⎝ 1 ⎠,
1 ∗ ∗ 1 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 1 ∗ ∗ 1
Uw4 = ⎝ ∗ 1 ∗ ⎠ ∩ U, Ūw4 = ⎝ 1 ⎠ ∩ U, w4 = ⎝ 1⎠,
∗ 1 ∗ 1 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 ∗ 1 ∗ 1
Uw5 = ⎝ 1 ⎠ ∩ U, Ūw5 = ⎝ ∗ 1 ∗ ⎠ ∩ U, w5 = ⎝ 1 ⎠,
∗ ∗ 1 1 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 1 ∗ ∗ 1
Uw6 = ⎝ ∗ 1 ⎠ ∩ U, Ūw6 = ⎝ 1 ∗ ⎠ ∩ U, w6 = ⎝ 1 ⎠.
∗ ∗ 1 1 1
⎛ ⎞ ⎛ ⎞
1 ∗ 1 ∗
For example, Uw5 = ⎝ 1 ⎠ and Ūw5 = ⎝ 1 ∗ ⎠ . We may think of
1 1
Ū w as the space opposite to Uw in U . It is clear that for each w ∈ W , we have
U = Uw · Ū w = Ū w · Uw . (10.6.5)
Returning to the proof of our lemma, it is plain that for every w ∈ W , the group
w acts on Pmin (Z)\(S L(n, Z) ∩ G w ) in the sense that right multiplication by
w maps left cosets to left cosets. We only need to show that this action is
proper, i.e., only the identity element acts trivially. To show this, suppose that
γ ∈ w = w−1 · t Pmin (Z) · w ∩ Pmin (Z) = Ū w (Z)
10.6 Fourier expansion of minimal parabolic Eisenstein series 305
fixes the left coset Pmin (Z) · b1 cwb2 where c ∈ D, and without loss of generality,
b1 ∈ Uw , b2 ∈ Ū w .
Since
Pmin (Z) · b1 cwb2 · γ = Pmin (Z) · b1 cwb2 ,
it follows that
b2 γ b2−1 ∈ Uw ∩ Ū w = {1}.
This proves that γ must be the identity matrix and the action is proper.
Note that the last integral on the right-hand side of (10.6.6) will be a degenerate
Whittaker function, as in Definition 5.10.1, if w is not the long element of the
Weyl group.
It is instructive to illustrate the double integral (10.6.6) with an example. We
shall consider G L(4, R). In this case
⎛ ⎞
1 ∗ ∗ ∗
⎜ 1 ∗ ∗⎟
U (R) = ⎜ ⎝
⎟.
1 ∗⎠
1
For our example, we will let
⎛ ⎞
1
⎜ 1 ⎟
w=⎜
⎝
⎟.
1⎠
1
In this case,
⎛ ⎞ ⎛ ⎞
1 1 ∗ ∗ ∗
⎜ 1 ∗⎟ ⎜ 1 ∗ ⎟
Uw (R) = ⎜
⎝
⎟, Ū w (R) = ⎜ ⎟,
1 ∗⎠ ⎝ 1 ⎠
1 1
and the double integral (10.6.6) takes the form:
1 ∞ ∞ ∞ ∞
e−2πim 3 u 3,4 du 3,4 · Is (wu 2 z)ψm (u 2 ) d ∗ u 2 ,
0 −∞ −∞ −∞ −∞
where
⎛ ⎞
1 u 1,2 u 1,3 u 1,4
⎜ 1 u 2,3 ⎟
u2 = ⎜
⎝
⎟,
⎠ d ∗ u 2 = du 1,2 du 1,3 du 1,4 du 2,3 .
1
1
10.7 Meromorphic equation of Eisenstein series 307
and where Det is the determinant function on hn . While it is not yet clear that
(10.7.1) converges absolutely for (s) > 2/n, this will follow directly from
(10.7.4). Note that in (10.7.1), we must put γ z in canonical Iwasawa form (as
in Proposition 1.2.6) before actually taking the determinant. The meromorphic
continuation and functional equation of E P (z, s) can be obtained from the
Poisson summation formula
1
f (m · z) = fˆ(m · (t z)−1 ), (10.7.2)
m∈Z n |Det(z)| m∈Zn
b1 = a1 y1 · · · yn−1
b2 = (a1 x1,2 + a2 )y1 · · · yn−2
..
.
bn = (a1 x1,n + a2 x2,n + · · · + an−1 xn−1,n + an ). (10.7.3)
10.7 Meromorphic equation of Eisenstein series 309
γ z = τ · k · r In ,
b12 + · · · + bn2 = r 2 .
Consequently
Then we have the Fourier transform fˆu (x) = (1/u n/2 ) f u (x/u) . We shall make
use of the Poisson summation formula (10.7.2) with this choice of function f u .
It follows from (10.7.3), (10.7.4), and the integral representation of the Gamma
310 Langlands Eisenstein series
function that
∞ =
>
E ∗P (x, s) = Det(z)
s
f u ((a1 , . . . , an ) · z) − f ((0, . . . , 0))
(a1 ,...,an )∈Zn
0
du
. × u ns/2
u
The proposition follows by breaking the integral into two parts: [0, 1] and
[1, ∞], and then applying the Poisson summation formula (10.7.2) just as we
did in the proof of the functional equation of the Riemann zeta function given
on page 1.
1
as in Section 10.6. Here C(z, s) denotes the degenerate terms in the Fourier
expansion associated to (m 1 , m 2 , . . . , m n−1 ) with m i = 0 for some 1 ≤ i
≤ n − 1. We shall assume E(z, s) is normalized (multiplied by a suitable
10.8 The L-function associated to parabolic Eisenstein series 311
for some fixed w in the Weyl group of G L(n, R) and all diagonal matrices
y as in (5.9.1). This is due to the fact that the action of the Weyl group on
the group of diagonal matrices just permutes the diagonal elements so that the
sum over 1 ≤ c1 , c2 , . . . , cn ∈ Z does not change. We shall show that E(z, s)
is actually an eigenfunction of the Hecke operators. The Hecke relations in
Section 9.3 imply that if A(( p k , 1, 1, . . . , 1), s) satisfies the functional equation
s → s for s, s given by (10.8.2) and all primes p and all k = 1, 2, . . . , then
A((m 1 , m 2 , . . . , m n−1 ), s) must also satisfy the functional equations (10.8.2)
for all m 1 , m 2 , . . . , m n−1 ∈ Zn−1 . It immediately follows from Theorem 5.9.8
+
j−1
that for s j,k = (nsn−k+i − 1)/2, we have
i=0
n−1
∗ − 12 −s j,k 1
E (z, s) = π + s j,k ζ (1 + 2s j,k )(E(z, s) − C(z, s))
j=1 j≤k≤n−1
2
= E ∗ (z, s ) (10.8.3)
for all s, s satisfying (10.8.2) It may also be shown that C(z, s) satisfies these
same functional equations.
n−1
n−1 b sj
Is (z) = yi i, j
i=1 j=1
with
ij if i + j ≤ n,
bi, j =
(n − i)(n − j) if i + j ≥ n.
312 Langlands Eisenstein series
Proof It follows from the Hermite normal form, Theorem 3.11.1, that for any
α ∈ Sm 1 , γ ∈ S L(n, Z), there exists a unique α ∈ Sm 1 and γ ∈ S L(n, Z) such
that γ α = αγ . The result follows easily from this.
10.8 The L-function associated to parabolic Eisenstein series 313
∞ ⎜⎜ m
⎟⎟
⎜⎜ .. ⎟⎟
= m −s
Iv− n1 ⎜⎜ . ⎟⎟
⎜⎜ c1 ···cn−2 cn−1
2 ⎟⎟
m=1 1≤c1 ,...,cn−1 ∈Z ⎝⎝ ⎠⎠
)
n−1 m
c
m 1
=1
⎛⎛ c1 ⎞⎞
m
∞
∞
∞ ⎜⎜ .. ⎟⎟
⎜⎜ . ⎟⎟
= ··· (mc1 · · · cn−1 )−s Iv− n1 ⎜⎜ ⎟⎟ .
c1 =1 cn−1 =1 m=1
⎝⎝ cn−1 ⎠⎠
m
1
(10.8.5)
The above computation immediately implies the following theorem.
314 Langlands Eisenstein series
where
n
s−λi (v) s − λi (v) n
s − λi (v)
π− = π− 2
ns
G Ev (s) = 2 ,
i=1
2 i=1
2
⎛⎛ ⎞⎞
c1
⎜⎜ .. ⎟⎟
⎜⎜ . ⎟⎟
Is − n1 ⎜⎜ ⎟⎟ .
1≤c1 ,c2 ,...,cn ∈Z
⎝⎝ cn−1 ⎠⎠
)
n
c
=m 1 cn
=1
)
r
for all z ∈ hn , and λ ∈ Cn−1 is such that φi (mni (z)) is of type λ.
i=1
Proof Since the invariant differential operators commute with the action of
G L(n, R), it is enough to check that
r
s
φi mni (z) · Det mni (z) i , (10.9.2)
i=1
r
s
Wλi mni (z) Det mni (z) i
i=1
r
s
= Iλi mni (wuz) ψi (u) · Det mni (z) i d ∗ u
i=1
Uni (R)
r
s
= Iλi mni (wuz) ψi (u) · Det mni (wuz) i d ∗ u,
i=1
Uni (R)
since Det(wu) = 1. It immediately follows that the above, and hence, (10.9.2) is
) a
n−1
an eigenfunction because it is obtained from y
for suitable a1 , a2 , . . . , an−1
=1
by integrating a set of left translates. To show that E P (z, s, φ) is an eigenfunction
of type λ + s one uses the fact that Iλ (z) · Is (z) = Iλ+s (z).
318 Langlands Eisenstein series
where
⎛ ⎞
c1 c1,2 · · · c1,n
⎜ c2 · · · c2,n ⎟
⎜ ⎟
c=⎜ .. .. ⎟ .
⎝ . . ⎠
cn
Note that
mni (cz) = mni (c) · mni (z).
Define ηi = n 1 + n 2 + · · · + n i−1 with the convention that η1 = 0. Then for
each i = 1, 2, . . . , r , we have
⎛ ⎞
cηi +1 cηi +1,ηi +2 · · · cηi +1,ηi+1
⎜ cηi +2 · · · cηi +2,ηi+1 ⎟
⎜ ⎟
mni (cy) = ⎜ .. .. ⎟
⎝ . . ⎠
cηi+1
⎛ ⎞
y1 y2 · · · yn−ηi −1
⎜ y1 y2 · · · yn−ηi −2 ⎟
⎜ ⎟
×⎜ .. ⎟.
⎝ . ⎠
y1 y2 · · · yn−ηi+1
10.10 The constant term 319
r
s +η s
m −(n−1)/2 TCi φi mni (z) · Ci i i · Det mni (y) i
C1 C2 ···Cr =m i=1
r
s +η s
= m −(n−1)/2 Ai (Ci ) φi mni (z) · Ci i i · Det mni (y) i ,
C1 C2 ···Cr =m i=1
Propositions 10.9.1 and 10.9.3 allow one to show that the non-degenerate
terms in the Whittaker expansion of general Langlands Eisenstein series for
S L(n, Z) with n ≥ 2 have a meromorphic continuation and satisfy the same
functional equation as the Whittaker functions that occur in the Fourier–
Whittaker expansion.
r
s
E P (s, z, φ) = φi mni (γ z) · Det mni (γ z) i (10.10.1)
γ ∈(P∩)\ i=1
because
if and only if u 1 u −1
2 ∈ N (Z) ∩ (γ
−1
P(Z)γ ).
Now, by the Bruhat decomposition (see Propositions 10.3.2, 10.3.6), each
γ ∈ P(Z)\/N (Z)
Since φ is a cusp form, the inner integral vanishes unless 0 N is the identity and
this will only happen if P and P are associate with w ∈ (P, P ).
⎞⎛⎛ ⎞
x3 1
y1 y2 x2
where = S L(3, Z), z = ⎝ x1 ⎠⎝ 1
y1 ⎠, with z 2 = x2 + i y2 ,
1 1
and the slash operator |γ denotes the action of γ on z.
322 Langlands Eisenstein series
Proposition 10.11.2 The constant terms of the Eisenstein series E P2,1 (z, s, φ)
given in (10.11.1) along the various parabolic subgroups take the form:
⎛⎛ ⎞ ⎞
1 1 1 1 u2 u3
E P2,1 ⎝⎝ 1 u 1 ⎠ z, s, φ ⎠ du 1 du 2 du 3 = 0,
0 0 0 1
⎛⎛ ⎞ ⎞
1 1 1 u3 s
E P2,1 ⎝⎝ 1 u 1 ⎠ z, s, φ ⎠ du 1 du 3 = 2 y12 y2 φ(z 2 ),
0 0 1
⎛⎛ ⎞ ⎞
1 1 1 u2 u3
E P2,1 ⎝⎝ 1 ⎠ z, s, φ⎠ du 2 du 3 = 2y11−s y22−2s φ (λ − 1) φ(z 1 ),
φ (λ)
0 0 1
where
s + + ir s + − ir
φ (s) = π −s L φ (s),
2 2
(with = 0 or 1 according as φ is even or odd), is the completed L-function of
the Maass form φ (of type 12 + ir ) as in Proposition 3.13.5, and where
3s − 12 if φ is even
λ =
3s + 12 if φ is odd.
Proof Omitted.
Recently, Sarnak (2004), Gelbart, Lapid and Sarnak (2004) obtained explicit
zero–free regions for general automorphic L-functions by use of the Jacquet–
Shalika method. We shall now present a short exposition of this method by
10.12 An application of the theory of Eisenstein series 323
considering the classical case of G L(2) and then the not so classical case of
G L(3).
where
√ s − 12 ζ (2s − 1)
φ(s) = π , σs (n) = ds,
(s) ζ (2s) d|n
d>0
and
∞
1 du
e− 2 y(u+(1/u)) u s
1
K s (y) = .
2 0 u
If ζ (1 + it0 ) = 0 for some t0 ∈ R, then one easily sees that if
then E ∗ (z, (1 + t0 )/2) must be a non-constant Maass form of type 14 + t02 . This
is because the constant term of the Eisenstein series E ∗ (z, s) will vanish when
s = (1 + t0 )/2. It is easy to show that E ∗ (z, (1 + t0 )/2) is non-constant, in
particular non-zero, because the sum:
1
σ1−2s (n)|n|s− 2 K s− 12 (2π|n|y)e2πinx
n=0
will be non-zero high in the cusp. The key point is that an Eisenstein series can
never be a cusp form because Eisenstein series are orthogonal to cusp forms,
so the inner product
? @
1 + t0 1 + t0
E ∗, , E ∗,
2 2
would have to be zero. This contradicts the fact that E (∗, (1 + t0 )/2) is not iden-
tically zero. Since we have obtained a contradiction, our original assumption:
that ζ (s) vanished on the line (s) = 1, must be false!
(s) = 1, then there exists a special value of s such that E ∗P2,1 (z, s, φ) will be a
Maass form for S L(3, Z), i.e., its constant term will vanish. One again obtains
a contradiction by taking the inner product of E ∗ with itself.
φ ∈ L2 (S L(3, Z)\h3 ).
1
φ00 (z) = φ, E(∗, s1 , s2 ) E 00 (z, s1 , s2 ) ds1 ds2 . (10.13.2)
(4πi)2
(si )= 13
The idea of the proof is to embed h2 in h3 and show that φ21 is invariant under
S L(2, Z) with this embedding. Then, we use a G L(2) spectral decomposition
of φ21 .
Following (Garrett, 2002), for z ∈ h3 = S L(3, R)/S O(3, R), we can write:
⎛ ⎞⎛ ⎞⎛ √ ⎞
1 0 x3 0 1/ l 0√ 0
⎜ ⎟
z = ⎝ 0 1 x1 ⎠ ⎝ z 2 0⎠ ⎝ 0 1/ l 0 ⎠ ,
0 0 1 0 0 1 0 0 l
The proof of Lemma 10.13.5 is fairly straightforward and we will omit it. We
will, however, give the calculation of φ, E j (∗, s). It is part of our assumption
326 Langlands Eisenstein series
Note that if we apply Mellin inversion to the results of Lemma 10.13.6 and
by our assumptions move the line of integration to ( 12 ), we have the following.
1
a j (l) = φ21 (∗, l), u j = φ, E j (∗, s) l 3−3s ds. (10.13.7)
πi ( 12 )
Now we can put this together with the spectral decomposition of φ21 (z 2 , l)
to get
∞
1
φ21 (z 2 , l) = φ, E j (∗, s) l 3−3s u j (z 2 ) ds
πi j=0 ( 12 )
1
+ φ21 (∗, l), E(∗, v) E(z 2 , v) dv.
4πi ( 12 )
Make the change of variable s → 1 − s in the first integral. Then the equation
above becomes:
∞
1
φ21 (z 2 , l) = φ, E j (∗, s) l 3s u j (z 2 ) ds
πi j=0 ( 12 )
1
+ φ21 (∗, l), E(∗, v) E(z 2 , v) dv. (10.13.8)
4πi ( 12 )
Let us now assume the following proposition, whose proof will be deferred
to later.
This proves (10.13.3). This is almost the end because (10.13.4) is proved
using (10.13.3) and a few properties of Eisenstein series. On G L(3), there is an
involution which preserves the Iwasawa form. The involution ι is defined by
⎛ ⎞
1
ι t −1
z = w z w, where w=⎝ 1 ⎠ . (10.13.11)
1
For any automorphic form φ(z), we can define another automorphic form,
also on S L(3, Z), denoted φ̃(z) given by
It is easy to see that φ̃, ψ̃ = φ, ψ. It follows that φ̃, E j (∗, s) =
φ, Ẽ j (∗, s). Also, we have the following equalities:
where
j (λ − 1)
θ (s) = .
j (λ)
The coefficients in L j (v) are real since they come from the Hecke form u j , so
we know that θ (s) = θ(s̄). Applying these observations to (10.13.13) gives
∞
1
φ12 (z) = θ(s̄)φ, E j (∗, s) Ẽ j,12 (z, 1 − s) ds
2πi j=0
( 12 )
1
+ φ, E(∗, s2 , s1 ) E 12 (z, s2 , s1 ) ds1 ds2 ,
(4πi)2
( 13 ) ( 13 )
∞
1
φ12 (z) = φ, E j (∗, s) E j,12 (z, s) ds
2πi j=0
( 12 )
1
+ φ, E(∗, s1 , s2 ) E 12 (z, s1 , s2 ) ds1 ds2 .
(4πi)2
( 13 ) ( 13 )
It was shown in Chapter 6 that the function det z = y12 y2 , is invariant under
β ∈ P21 . So, det βz = det z.
Previously we had dropped the dependence of n in the notation, but to avoid
confusion we need to introduce n in the notation. Thus (n) = S L(n, Z) and
∞ (n) is the set of upper-triangular matrices in S L(n, Z) with 1s on the diagonal.
s1 +(s2 /2) 3s2 /2
E(z, s1 , s2 ) = 2
y1 y2 y2 ,
α∈P21 \(3) β∈∞ (3)\P21 β α
s1 +(s2 /2) 3s2
= 2 y12 y2 E z2, ,
2
α∈P21 \(3) α
s2 3s2
= E∗ z, s1 + , .
2 2
In the above we have used ∞ (3)\ P21 ≡ ∞ (2)\ S L(2, Z), a fact which is easy
to prove and also follows as a special case of Lemma 5.3.13.
Now, we will drop the dependence of the notation on n. To prove the proposi-
tion, we want to calculate the following inner product, in the region of absolute
convergence of E ∗ (z, v1 , v2 ). In a manner similar to the calculation of Lemma
10.13.6, we have
φ, E ∗ (∗, v1 , v2 ) = φ(z) E ∗ (z, v1 , v2 ) d ∗ z
\h3
= φ(z) E ∗ (z, v1 , v2 ) d ∗ z,
\h3
10.13 Langlands spectral decomposition for S L(3, Z)\h3 331
∞ 1 1
=3 φ(z)y12v1 y2v1 l −3 E(z 2 , v2 )
0 0 0
S L(2,Z)\h2
dl
× d ∗ z 2 d x3 d x1 ,
l
∞
dl
=3 φ21 (z 2 , l) l 3v1 −3 E(z 2 , v2 ) d ∗ z 2 ,
0 l
S L(2,Z)\h2
∞
dl
=3 φ21 (∗, l), E(∗, v2 ) l 3v1 −3 .
0 l
Let a(l) = φ21 (∗, l), E(∗, v2 ), then we have just shown that
In the last step we have used the assumption that φ is orthogonal to all the
residues of the Eisenstein series to move the line to (v1 ) = 12 .
We have just shown that
1
φ21 (∗, l), E(∗, v2 ) = φ, E ∗ (∗, v1 , v2 ) l 3−3v1 dv1 . (10.13.14)
2πi
( 12 )
Now let v1 = s1 + (s2 /2) and v2 = 3s2 /2. From the definition of E ∗ (z, v1 , v2 ),
we get the following.
3 2 s1 +(s2 /2) 3s2
LHS = φ, E(∗, s1 , s2 ) y1 y2 E z2, ds1 ds2 .
4(2πi)2 ( 13 ) ( 13 ) 2
(10.13.15)
Using the fact that 21 (z 2 , l) is invariant under S L(2, Z) and so has a Fourier
expansion, it is easy to show that for any form (z),
21 (z) = 0,n (z),
n∈Z
where
⎛⎛ ⎞ ⎞
1 1 1 1 u2 u3
n 1 ,n 2 (z) = ⎝⎝0 1 u 1 ⎠ z⎠ e(−n 1 u 1 − n 2 u 2 ) du 1 du 2 du 3 .
0 0 0
0 0 1
We will need some results from (Bump, 1984) and will use the notation from
there. In (Bump, 1984), it is shown that the expansion of E 0,n is the sum of
three Whittaker functions indexed by w ∈ V, where V is a certain subset of the
Weyl group W . (In the notation of (Bump, 1984), V = {w1 , w2 , w4 }.) So,
1 (s1 ,s2 )
E 0,n (z, s1 , s2 ) = aw (n, s1 , s2 ) W0,n (z, w). (10.13.16)
2 w∈V
Using the expansion of the G L(2) Eisenstein series, it is easy to show that
the factor (y12 y2 )s1 +(s2 /2) E (z 2 , 3s2 /2) which appears in (10.13.15) is exactly:
2 s1 +(s2 /2) 3s2 1 (s1 ,s2 )
y1 y2 E z2, = aw (n, s1 , s2 ) W0,n (z, w2 ).
2 2 n∈Z 2
10.13 Langlands spectral decomposition for S L(3, Z)\h3 333
Now in (10.13.15), we will divide the integral into three parts and in each
part let (s1 , s2 ) → (u(s1 ), u(s2 )) for u ∈ U. We choose the subset U of W in
such a way as to get the sum over u to give us E 0,n (z, s1 , s2 ). We can do this
because the Whittaker functions have certain transformation properties. We
(s1 ,s2 )
will be interested in the transformation of W0,n (z, w2 ) under the action of the
Weyl group on (s1 , s2 ); specifically for w ∈ V, there exists a u ∈ U depending
on w, such that
(u(s1 ),u(s2 )) (s1 ,s2 )
W0,n (z, w2 ) = Bwu W0,n (z, w). (10.13.17)
Recall that w ∈ V are those that appear in the expansion of E 21 .
LHS
3 (s1 ,s2 )
= 2
φ, E(∗, s1 , s2 ) aw2 (n, s1 , s2 )W0,n (z, w2 ) ds1 ds2 ,
n∈Z
8(2πi) ( 1
) ( 1
)
3 3
1
= φ, E(∗, u(s1 ), u(s2 )) aw2 (n, u(s1 ), u(s2 ))
n∈Z u∈U
8(2πi)2
( 13 ) ( 13 )
(u(s1 ),u(s2 ))
× W0,n (z, w2 ) ds1 ds2 ,
1
= φ, E(∗, u(s1 ), u(s2 )) aw2 (n, u(s1 ), u(s2 ))
n∈Z w∈V
8(2πi)2
( 13 ) ( 13 )
(s1 ,s2 )
× Bwu W0,n (z, w)ds1 ds2 .
Also note that at (si ) = 13 , we have (u(si )) = 13 .
From Bump again, we can show that for any w ∈ W,
E(z, w(s1 ), w(s2 )) = ϕw (s1 , s2 )E(z, s1 , s2 )
for some ϕw independent of z. This is done as follows. First, let us define, as in
(Bump, 1984), the function G(z, s1 , s2 ) in the following way.
G(z, s1 , s2 ) = B(s1 , s2 ) E(z, s1 , s2 ),
where
ζ (3s1 ) ζ (3s2 ) ζ (3s1 + 3s2 − 1) 3s1 3s2
B(s1 , s2 ) =
4π 3s1 +3s2 −1/2 2 2
3s1 + 3s2 − 1
× E(z, s1 , s2 ).
2
Now from (Bump, 1984), we know that G(z, s1 , s2 ) is invariant under the
action of the Weyl group on the pair (s1 , s2 ). In particular, it follows,
B(w(s1 ), w(s2 )) E(z, w(s1 ), w(s2 )) = B(s1 , s2 )E(z, s1 , s2 ). (10.13.18)
334 Langlands Eisenstein series
B(s1 , s2 )
So, ϕw (s1 , s2 ) = . From the definition B(s1 , s2 ) = B(s1 , s2 ).
B(w(s1 ), w(s2 ))
At (si ) = 13 , we have that si = 23 − si and (w(si )) = 13 . It follows that at
(si ) = 13 ,
2
B − s1 , 23 − s2
ϕw (s1 , s2 ) = 2 3
.
B 3
− w(s1 ), 23 − w(s2 )
Consequently,
2
1 B − s1 , 23 − s2
LHS = φ, E(∗, s1 , s2 ) · 2 3
2(4πi)2 n∈Z w∈V B 3
− u(s1 ), 23 − u(s2 )
si = 13
(s1 ,s2 )
× aw2 (n, u(s1 ), u(s2 ))Bwu W0,n (z, w) ds1 ds2 .
Now if we consider this last equality along with the transformation property
(u(s1 ),u(s2 )) (s1 ,s2 )
W0,n (z, w2 ) = Bwu W0,n (z, w)
and use the fact that the Whittaker functions are linearly independent, we get
the equality:
B(s1 , s2 ) (s1 ,s2 )
aw (n, s1 , s2 )W0,n (z, w)
B(u(s1 ), u(s2 ))
(s1 ,s2 )
= aw2 (n, u(s1 ), u(s2 ))Bwu W0,n (z, w).
So,
B(s1 , s2 )
aw (n, s1 , s2 ) = aw2 (n, u(s1 ), u(s2 ))Bwu . (10.13.22)
B(u(s1 ), u(s2 ))
Now (10.13.21) reduces to showing that
B (u(s1 ), u(s2 )) B 23 − s1 , 23 − s2
= 2 . (10.13.23)
B(s1 , s2 ) B 3 − u(s1 ), 23 − u(s2 )
To prove this and to finish the proof of the lemma, we have to show that the
product
2 2
B(s1 , s2 ) B − s1 , − s2
3 3
is invariant under the action of the Weyl group on (s1 , s2 ). Since
2 2
B(s1 , s2 ) B − s1 , − s2
3 3
is a product of six gamma functions and six zeta functions and since the quan-
tities
are permuted by the Weyl group, this completes the proof of Lemma 10.13.19
and Proposition 10.13.10.
336 Langlands Eisenstein series
Poincaré series and Kloosterman sums associated to the group S L(3, Z) were
introduced and studied in (Bump, Friedberg and Goldfeld, 1988) following the
point of view of Selberg (1965). A very nice exposition of the G L(2) theory
is given in (Cogdell and Piatetski-Shapiro, 1990). The method was first gen-
eralized to G L(n) in (Friedberg, 1987), (Stevens, 1987). In (Bump, Friedberg
and Goldfeld, 1988) it is shown that the S L(3, Z) Kloosterman sums are hyper
Kloosterman sums associated to suitable algebraic varieties. Non-trivial bounds
were obtained by using Hensel’s lemma and Deligne’s estimates for hyper-
Kloosterman sums (Deligne, 1974) in (Larsen, 1988), and later (Dabrowski
and Fisher, 1997) improved these bounds by also using methods from algebraic
geometry following (Deligne, 1974). Sharp bounds for special types of Kloost-
erman sums were also obtained in (Friedberg, 1987a,c). In (Dabrowski, 1993),
the theory of Kloosterman sums over Chevalley groups is developed. Impor-
tant applications of the theory of G L(n) Kloosterman sums were obtained in
(Jacquet, 2004b) (see also (Ye, 1998)).
Another fundamental direction for research in the theory of Poincaré series
and Kloosterman sums was motivated by the G L(2) Kuznetsov trace formula,
(see (Kuznecov, 1980) and also (Bruggeman, 1978)). Generalizations of the
Kuznetsov trace formula to G L(n), with n ≥ 3 were obtained in (Friedberg,
1987), (Goldfeld, 1987), (Ye, 2000), but they have not yet proved useful for
analytic number theory. The chapter concludes with a new version of the G L(n)
Kuznetsov trace formula derived by Xiaoqing Li.
337
338 Poincaré series and Kloosterman sums
1s on the diagonal. Note that for fixed integers m 1 , m 2 , . . . , m n−1 , the character
ψ may be defined by
ψ(u) = ψ (u) = e2πi (m 1 u 1 +m 2 u 2 +···+m n−1 u n−1 )
m 1 ,...,m n−1 (11.1.1)
where
⎛ ⎞
1 u n−1
⎜ 1 u n−2 ∗ ⎟
⎜ ⎟
⎜ .. .. ⎟
u=⎜ . . ⎟ ∈ Un (R).
⎜ ⎟
⎝ 1 u1 ⎠
1
The Poincaré series is constructed from the following functions:
r The I (z)–function given in (5.1.1).
s
r A bounded function: eψ : hn → C which is characterized by the property:
φ, P(∗, s, eψ )
= φ(z) P(z, s, eψ ) d ∗ z
\hn
n−1
dyi
= φ(x y)eψ (y) · ψ̄m 1 ,...,m n−1 (x) d x j,k i(n−i)+1
,
1≤ j<k≤n i=1 yi
Un (Z)\hn
where we have used the explicit measure d ∗ z given in Theorem 1.6.1. The inner
d x-integral picks off the (m 1 , . . . , m n−1 )th Fourier coefficient, and the result
follows.
and Dn denotes the subgroup of diagonal matrices in G L(n, R). We let Un (Z)
denote the matrices in Un with integer coefficients. Let c1 , . . . , cn−1 be non-zero
integers, and set
⎛ ⎞
1/cn−1
⎜ cn−1 /cn−2 ⎟
⎜ ⎟
⎜ .. ⎟
c=⎜ . ⎟ (11.2.1)
⎜ ⎟
⎝ c2 /c1 ⎠
c1
provided this sum is well defined, (i.e., it is independent of the choice of Bruhat
decomposition for γ ). Otherwise, the Kloosterman sum is defined to be zero.
The Kloosterman
sum is
based on the Bruhat decomposition. It is convenient
1 b1 /c1 1 b2 /c1
to set b1 = , b2 = . In this case, the Bruhat decom-
1 1
±1 ∗
position says that every γ ∈ S L(2, Z), γ = can be written in the
0 ±1
form
−1
1 b1 /c1 c1 −1 1 b2 /c1
γ =
1 c1 1 1
b1 (b1 b2 − 1)/c1
=
c1 b2
where
then the Kloosterman sum in (11.2.4) can be written in the more traditional form
2πi Mβ+N β̄
Sw (ψ, ψ , c) := S(M, N , c1 ) = e c1
. (11.2.5)
β (mod c1 )
β β̄ ≡ 1 (mod c1 )
One may use this to systematically solve for the coefficients of b1 and b2 . In
this manner, one obtains
⎛ ⎞
1 c1 a1,2 − a1,1ca2 3,2 ac1,11
⎜ ⎟
b1 = ⎜
⎝ 1 a2,1 ⎟
c1 ⎠
,
1
⎛ a3,2 a3,3 ⎞
1 c1 c1
⎜ ⎟
b2 = ⎝ 1 c1 a2,3 − a2,1 a3,3 ⎠. (11.2.9)
c2
1
342 Poincaré series and Kloosterman sums
Sw (ψ M , ψ N , c)
6M a 7 6 N (c a −a a ) N a 7
M (c a −a a )
2πi 1c 2,1 + 2 1 1,2c 1,1 3,2 2πi 1 1 2,3c 2,1 3,3 + 2c 3,2
= e 1 2 ·e 2 1 ,
a2,1 ,a3,2 (mod c1 )
c1 a2,3 −a2,1 a3,3 (mod c2 )
c1 a1,2 −a1,1 a3,2 (mod c2 )
subject to the constraint that there exist integers a1,2 , a1,3 , a2,2 such that
c1 (a1,2 a2,3 −a1,3 a2,2 )+a3,2 a1,3 a2,1 − a1,1 a2,3 + a3,3 a1,1 a2,2 − a1,2 a2,1 =1,
ψ(cwuw −1 ) = ψ (u)
for all u ∈ (w−1 Un (R)w) ∩ Un (R). Hence the Kloosterman sum is non-zero
only in this case.
Remarks The Kloosterman zeta function was first introduced, for the case of
G L(2), in (Selberg, 1965) who showed the meromorphic continuation in s and
the existence of infinitely many simple poles on the line (s) = 12 occurring at
the eigenvalues of the Laplacian. Selberg used Weil’s bound for G L(2) Kloost-
erman sums to deduce that the Kloosterman zeta function converged absolutely
in the region (s) > 34 . He was able to immediately deduce from this the bound
≥ 163
for the lowest eigenvalue of the Laplacian (see (Selberg, 1965)).
For (s) > 12 + , it was shown in (Goldfeld and Sarnak, 1983) that the
Kloosterman zeta function is
1
|s| 2
O .
(s) − 12
Such results have not been obtained for higher-rank Kloosterman zeta functions.
This would constitute an important research problem.
As an example, consider
⎛ ⎞
g1,1 g1,2 g1,3
⎝
g = g2,1 g2,2 g2,3 ⎠ ∈ G L(3, R).
g3,1 g3,2 g3,3
Then the Plücker coordinates are
The key theorem needed for the evaluation of Kloosterman sums is the
following (see (Friedberg, 1987)).
Proof It is convenient to have a notation for the minors that occur in the
Plücker coordinates. Fix n ≥ 2. For 1 ≤ k ≤ n − 1, let Lk denote the set of all
k-element subsets of {1, 2, . . . , n} with the lexicographical ordering <, where
{
1 , . . . ,
k } <
1 , . . . ,
k
if
1 <
1 . If the first elements are equal, but only in that case, you look at the
second and compare them, whichever is higher, and that is the “greater” set. One
continues naturally in this manner obtaining an ordering which is analogous to
alphabetical order. Then the set Lk can be used to index the minors that occur
in the Plücker coordinate ρk . For example, if {
1 , . . . ,
k } ∈ Lk , then we may
consider the minor formed with the columns
1 , . . . ,
k .
For each k with 1 ≤ k < n − 1, we may list the elements of Lk+1 (in some
order) as follows:
- .
Lk+1 = λk,1 , . . . , λk, n
k+1
with
ν1 , . . . , νn , ν1,2 , ν1,3 , . . . , νn−1,n , . . . , ν2,3,...,n ⊆ R2 −2
n
(11.3.3)
satisfying
⎛ ⎞
νλk,1
⎜ νλk,2 ⎟
⎜ ⎟
⎜ .. ⎟ ∈ Image(T ), (11.3.4)
⎜ . ⎟
⎝ ⎠
νλ
n
k, k+1
n
for all 1 ≤ k < n − 1, where T = (ti, j ) (with 1 ≤ i ≤ k+1 , 1 ≤ j ≤ n) is
the linear transformation
(−1)r −1 νλk,i −{ j} if j =
k,i,r ∈ λk,i ,
ti, j = (11.3.5)
0 otherwise.
Here λk,i is a set of k integers and λk,i − { j} means to delete the element j
from the set λk,i . Note also that the linear transformation T maps Rn to Rn
by simple left matrix multiplication of T on an n-dimensional column vector
in Rn . Thus, given coordinates νλ (with λ ∈ Lk ) we see that (11.3.4), (11.3.5)
gives a condition which allows us to see which νλ (with λ ∈ Lk+1 ) occur.
11.3 Plücker coordinates and the evaluation of Kloosterman sums 345
V1 := v ∈ V vλ = 0 for some λ ∈ Ln−1 .
To see this, note that the map M factors through Un (R)\G L(n, R). Furthermore,
condition (11.3.4) holds for M(g) since Mλ (g) can be computed for λ ∈ Lk in
terms of Mλ (g) with λ ∈ Lk−1 for all 1 < k < n, by expanding the determinant
along the top row. Since Det(g) = 0 it follows that M maps into V1 . Now
given v ∈ V1 , the condition (11.3.4) guarantees that there exists an (n − 1) × n
matrix with appropriate minors. Since vλ = 0 for some λ ∈ Ln−1 , this can be
completed to a matrix g ∈ G L(n, R) with image v. Consequently, the map
M is onto V1 . It remains to show that the map M : Un (R)\G L(n, R) → V1 is
injective.
To see this, note first that by Proposition 10.3.6, if
(1)
1 if λ ∈ λ(1)
1 , . . . , λn−1
Mλ (g) =
0 otherwise,
g −1 (e1 ) ∧ · · · ∧ g −1 (en−k ) ∧ e
1 ∧ · · · ∧ e
k
= h −1 (e1 ) ∧ · · · ∧ h −1 (en−k ) ∧ e
1 ∧ · · · ∧ e
k .
346 Poincaré series and Kloosterman sums
Hence
It follows that
1 if λ = λ(1)
k ,
Mλ (gh −1 ) =
0 if λ = λ(1)
k ,
and
The Plücker relations are defined by a set of quadratic forms in the bottom row
based minor determinants of any n × n matrix in G L(n, R). These forms have
coefficients ±1. They must vanish if the values assigned to symbols representing
the minor determinants come from any square matrix. The number of Plücker
relations grows rapidly with n, because each j × j sub-matrix, with elements
chosen from the bottom j rows and any j columns, also gives rise to a set of
relationships of the given type. In dimension 2 there are no relationships and in
dimension 3 just one, the Cramer’s rule relationship ν1 ν23 − ν2 ν13 + ν3 ν12 = 0.
(Here νλ , where λ is an ordered subset of {1, 2, . . . , n}, is used to represent the
minor determinant based on the bottom |λ| rows and the columns indexed by
the elements of λ.)
Example 11.3.11 (G L(4) Plücker relations) I would like to thank Kevin
Broughan for this example in which ten relations for dimension n = 4 are
derived.
(i) The simplest relation is obtained by expanding the matrix using the bottom
row-based minors of size n − 1 along the bottom row. By Cramer’s rule we
obtain
(−1)1+1 ν1 ν234 + (−1)1+2 ν2 ν134 + (−1)1+3 ν3 ν124 + (−1)1+4 ν4 ν123 = 0.
348 Poincaré series and Kloosterman sums
(ii) Now let the rows of a fixed but arbitrary 4 × 4 matrix be represented by
the vectors a1 , a2 , a3 , a4 . Then because for all vectors v, v ∧ v = 0:
0 = (a3 ∧ a4 ) ∧ (a3 ∧ a4 ) = α e1 ∧ e2 ∧ e3 ∧ e4 ,
where the ei are the standard unit vectors and α is a real constant with value
ν12 ν34 − ν24 ν13 + ν14 ν23 , which is therefore 0.
(iii) The relation ν24 ν123 + ν12 ν234 − ν23 ν124 = 0 will now be derived. First
expand the wedge product of the bottom two rows:
a3 ∧ a4 = ν12 e1 ∧ e2 + ν13 e1 ∧ e3 + ν14 e1 ∧ e4
+ ν23 e2 ∧ e3 + ν24 e2 ∧ e4 + ν34 e3 ∧ e4 .
Then
a3 ∧ a4 = e2 ∧ (−ν12 e1 + ν23 e3 + ν24 e4 )
+ ν13 e1 ∧ e3 + ν14 e1 ∧ e4 + ν34 e3 ∧ e4
= e2 ∧ ω + η,
say, where ω is a 1-form and η a 2-form with e2 not appearing. Then
a2 ∧ a3 ∧ a4 = ν123 e1 ∧ e2 ∧ e3 + ν234 e2 ∧ e3 ∧ e4 + ν124 e1 ∧ e2 ∧ e4
so
a2 ∧ a3 ∧ a4 ∧ ω = (ν24 ν123 + ν12 ν234 − ν23 ν124 )e1 ∧ e2 ∧ e3 ∧ e4 + 0
= λ e1 ∧ e2 ∧ e3 ∧ e4 .
Since each term of η has two of {e1 , e3 , e4 } and each term of ω one of this
set we can write
η ∧ ω = (ν12 ν34 − ν24 ν13 + ν14 ν23 )e1 ∧ e3 ∧ e4 = 0
by the relation derived in (ii) above.
But then
a2 ∧ a3 ∧ a4 ∧ ω = a2 ∧ (a3 ∧ a4 ) ∧ ω
= a2 ∧ (e2 ∧ ω + η) ∧ ω
= a2 ∧ (η ∧ ω) = 0,
so, therefore, λ = 0, and we get ν24 ν123 + ν12 ν234 − ν23 ν124 = 0.
Three similar relationships are derived by factoring out in turn each of the
unit vectors e1 , e3 and e4 .
(iv) The four remaining relations are obtained by applying the dimension 3
relations to each of the four subsets of {1, 2, 3, 4} of column numbers of size 3.
11.3 Plücker coordinates and the evaluation of Kloosterman sums 349
Example 11.3.12 (The w = In−1 ±1 Kloosterman sum)
This example was
first worked out in (Friedberg, 1987). For w = In−1 ±1 , we have
⎛ ⎞
1 ∗ 0
⎜ .. .. ⎟
⎜ . .⎟
Uw := (w −1 Un (R)w) ∩ Un (R) = ⎜ ⎟.
⎝ 0⎠
1
It follows from Proposition 11.2.10 (brute force computation) that the Kloost-
erman sum is non-zero only when c1 |c2 , c2 |c3 , . . . , cn−2 |cn−1 .
Next, we show that the map M : G L(n, R) → Rn −2 arising in the proof of
2
Theorem 11.3.2, composed with projection, gives a bijection between the sets
γ ∈ Un (Z)\S L(n, Z) ∩ G w / w diag(γ ) = c
and
-
S = (νn , νn−1,n , . . . , ν2,3,...,n ) ∈ Zn−1 νn (mod c1 ), νn−1,n (mod c2 ), . . . ,
.
ν2,3,...,n (mod cn−1 ), (νn− j+1,...,n , c j /c j−1 ) = 1 for all 1 ≤ j ≤ n − 1 ,
with c0 = 1.
To see the bijection, note that the image of M composed with projection
is contained in S, since Det(γ ) = 1 implies that the relative primality condi-
tions must hold. Further, the map is one-to-one because the information in S
determines all minors of the form
Mn−k,...,
n−i,...,n
(γ ), i <k (mod ck )
by induction on k for
ck
Mn−k,...,
n−i,...,n
(γ ) = · Mn−k+1,...,
n−i,...,n
(γ ).
ck−1
where
Sw (ψ M , ψ N , cc ) = Sw (ψ M , ψ N , c) · Sw (ψ M , ψ N , c ).
Proof This was first proved for n = 3 in (Bump, Friedberg and Goldfeld,
1988). See (Friedberg, 1987), (Stevens, 1987) for the proof.
Let
P(z, s, eψ ) := Is (γ z)eψ (γ z)
γ ∈Un (Z)\
be a Poincaré series for S L(n, Z) as in Definition 11.1.2. The main goal of this
section is an explicit computation of the N = (N1 , . . . , Nn−1 ) ∈ Zn−1 Fourier
coefficient
P(uz, s, eψ )ψ N (u) d ∗ u, (11.5.1)
Un (Z)\Un (R)
with
⎛ ⎞
1 u n−1
⎜ 1 u n−2 u i, j ⎟
⎜ ⎟
⎜ .. .. ⎟
u=⎜ . . ⎟ ∈ Un (R),
⎜ ⎟
⎝ 1 u1 ⎠
1
)
d ∗u = du i, j , and ψ N (u) = e2πi(u 1 N1 +···+u n−1 Nn−1 ) .
1≤i< j≤n
The computation of (11.5.1) requires some additional notation.
11.5 Fourier expansion of Poincaré series 353
We are now ready to state and prove the main theorem of this section which
was first proved in (Bump, Friedberg and Goldfeld, 1988) for n = 3, and then
more generally in (Friedberg, 1987).
Proof It follows from (11.5.1) and the definition of the Poincaré series that
P(uz, s, eψ M )ψ N (u)d ∗ u = Is (γ uz)eψ M (γ uz)ψ N (u)d ∗ u.
γ ∈Un (Z)\
Un (Z)\Un (R) Un (Z)\Un (R)
The sum over γ ∈ Un (Z)\ on the right above may be rewritten using the
Bruhat decomposition (see Section 11.2) after taking an additional quotient by
354 Poincaré series and Kloosterman sums
γ = b1 cwb2 , b1 , b2 ∈ Un (Q),
Warning Note that this does not conform to the definition of hn as given in
Definition 1.2.3, although it is a very close approximation. We shall only adopt
this notation temporarily in this section because of the particular applications
in mind, and the fact that the Selberg transform may take a very simple form in
some cases.
For z, z ∈ hn and every k ∈ Cc∞ (S O(n, R)\S L(n, R)/S O(n, R)), a space
which contains infinitely differentiable compactly supported S O(n, R) bi-
invariant functions on S L(n, R), we define an automorphic kernel
K (z, z ) = k(z −1 γ z ). (11.6.1)
γ ∈S L(n,Z)
Clearly
Here
⎛ ⎞
1 x1,2 x1,3 ... x1,n
⎜ 1 x2,3 ... x2,n ⎟
⎜ ⎟
⎜ .. .. ⎟
x =⎜ . . ⎟, xi := xn−i,n−i+1 (i = 1, . . . , n − 1),
⎜ ⎟
⎝ 1 xn−1,n ⎠
1
356 Poincaré series and Kloosterman sums
⎛ ⎞
y1 y2 . . . yn−1
⎜ y1 y2 . . . yn−2 ⎟
⎜ ⎟ −1/n
⎜ .. ⎟
y=⎜ . ⎟ · y1n−1 y2n−2 . . . yn−2
2
yn−1 ,
⎜ ⎟
⎝ y1 ⎠
1
(11.6.4)
Define
eψ M (y, z ) := k(z −1 z )ψ M (x) d ∗ x. (11.6.8)
Un (Z)\Un (R)
= ψ M (u)eψ M (y, z ),
so that eψ M (y, z ) is an e-function. It is also clear that PM (y, z ) ∈
L2 (S L(n, Z)\hn ) as a function of z . Hence, PM (y, z ) is a Poincaré series
as in Definition 11.1.2 with s = (0, 0, . . . , 0).
11.6 Kuznetsov’s trace formula for SL(n, Z) 357
Ūw
= k(z −1 · wcuy ) ψ M (x) ψ Nv (u) d ∗ x d ∗ u,
Ūw Un (Z)\Un (R)
(11.6.10)
and Sw (ψ M , ψ Nv ; c) is the Kloosterman sum as in Definition 11.2.2.
The other way of computing the Fourier coefficient (11.6.9) is to make use
of the Langlands spectral decomposition (Langlands, 1966, 1976) (see also
Theorem 10.13.1). Langlands spectral decomposition states that
L2 (S L(n, Z)\hn ) = L2discrete (S L(n, Z)\hn ) ⊕ L2cont (S L(n, Z)\hn )
where
L2discrete (S L(n, Z)\hn ) = L2cusp (S L(n, Z)\hn ) ⊕ L2residue (S L(n, Z)\hn ).
Here L2cusp denotes that the space of Maass forms, L2residue consists of iterated
residues of Eisenstein series twisted by Maass forms, and L2cont is the space
spanned by integrals of Langlands Eisenstein series.
For each k ∈ Cc∞ (S O(n, R)\S L(n, R)/S O(n, R)), we can define an integral
operator L k which acts on f ∈ L2 (S L(n, Z)\hn ), by
(L k f )(z) := f (w)k(z −1 w) d ∗ w
hn
= f (γ w)k(z −1 γ w) d ∗ w
γ ∈
S L(n,Z)\hn
= f (w)K (z, w) d ∗ w.
S L(n,Z)\hn
(L k φ)(z) = k̂(ν)φ(z),
is called the Harish transform of k. One can think of the Selberg transform
as the Mellin transform of its Harish transform. The Selberg transform has an
inversion formula due to Harish-Chandra, Bhanu-Murthy, Gangolli, etc., see
(Terras, 1988) and the references there.
For simplicity, we introduce the parameters ak for 1 ≤ k ≤ n − 1, which are
defined by
j(n − j) ak
n− j n−1
+ = b ji νi
2 k=1
2 i=1
with
n−1
|z i | 2 (ān−i /2)−(ān−i+1 /2)−1) ,
1
pa (z) =
i=1
11.6 Kuznetsov’s trace formula for SL(n, Z) 359
with
n
ai = 0.
i=1
Proof See page 88 in (Terras, 1988). Our model hn can be identified with her
model SP n of n × n positive definite matrices of determinant 1 by considering
the map: hn → SP n : z → z t z.
where for each φ j in the cuspidal spectrum, every Eisenstein series E in the
continuous spectrum, and all N = (N1 , . . . , Nn−1 ) ∈ Zn−1 with N1 · · · Nn−1 =
0, we have
⎛ ⎞
N1 · · · |Nn−1 |
⎜ .. ⎟
⎜ . ⎟
∗ ⎜ ⎟
N := ⎜ N1 N2 ⎟,
⎜ ⎟
⎝ N1 ⎠
1
A j (N )
WJacquet (N ∗ y , ν j ) = φ j (x y )ψ N (x) d ∗ x,
)
n−1
|Ni |i(n−i)/2 Un (Z)\Un (R)
i=1
and, correspondingly,
ε(N , s) ∗
WJacquet (N y , s) = E(x y , s)ψ N (x) d ∗ x.
)
n−1
|Ni |i(n−i)/2 Un (Z)\Un (R)
i=1
The contribution from the residual spectrum is 0 due to the lack of non-
degenerate Whittaker models for the residual spectrum.
× WJacquet (N ∗ y , s)WJacquet (M ∗ y, s) ds
∞
∞
= ··· Sw ψ M , ψ Nv ; c Jw y, y ; ψ M , ψ Nv , c .
w∈Wn v∈Vn c1 =1 cn−1 =1
j(n − j)
n− j n−1
+ αk = b ji νi (11.6.15)
2 k=1 i=1
n−1
αn = − αi . (11.6.16)
i=1
Remark Stade proves the above formula for (s) large, but one may obtain
|α j | < 12 by Proposition 12.1.9, and similarly for βk , so the right-hand side of
the formula is analytic for s 1. On the other hand Jacquet (2004a) shows
the integral on the left is absolutely convergent for (s) 1, so by analytic
continuation Stade’s formula holds for s 1.
We are now ready to state and prove the Kuznetsov trace formula.
n−1
(n− j)(1− j) dt j
× 2 π (n− j) t j
j=1
tj
11.6 Kuznetsov’s trace formula for SL(n, Z) 363
⎛ ⎞
t1 · · · tn−1
⎜ .. ⎟
⎜ . ⎟
t := ⎜ ⎟.
⎝ t1 ⎠
1
ti = Mi yi = Ni yi , (i = 1, 2, . . . , n − 1).
and then integrate over (R+ )n−1 . Theorem 11.6.19 immediately follows from
Proposition 11.6.17.
Concluding Remarks
(i) The test function k can be extended to a larger space, say, Harish-
Chandra’s Schwartz space, as long as the convergence of both sides is not a
problem.
(ii) The nice feature of the above Kuznetsov type formula is that the residual
spectrum does not appear, while its appearance is inevitable in the Selberg–
Arthur trace formula. In this sense, the Kuznetsov formula is more handy
to treat the cuspidal spectrum.
(iii) In the case of G L(2), Kuznetsov also derived an inversion formula so that
one can put any good test function on the Kloosterman sum side. Such an
inversion formula on G L(n) does not exist yet, although it is certainly a
very interesting research problem.
364 Poincaré series and Kloosterman sums
365
366 Rankin–Selberg convolutions
theorem of Shintani (1976) which relates the Fourier coefficients of Maass forms
with Schur polynomials as was discussed in Section 7.4, and ties everything
together with an old identity of Cauchy. Such formulae were originally con-
jectured by Langlands (1970) and generalizations were found by Kato (1978)
and Casselman and Shalika (1980). We shall follow this approach as in (Bump,
1984, 1987, 2004). Alternatively, a basic reference for an adelic treatment of the
Rankin–Selberg theory is Cogdell’s, analytic theory of L-functions for G L n , in
(Bernstein and Gelbart, 2003).
The Rankin–Selberg convolution is one of the most important constructions
in the theory of L-functions. Naturally it has had inumerable generalizations.
The excellent survey paper of Bump (to appear) gives a panoramic overview
of the entire subject. We give applications of the Rankin–Selberg convolution
method towards the generalized Ramanujan and Selberg conjectures. In par-
ticular, the chapter concludes with the theorem of Luo, Rudnick and Sarnak
(1995, 1999) which has been used to obtain the current best bounds for the
Ramanujan and Selberg conjectures.
∞
∞ A(m 1 , . . . , m n−1 )
f (z) = ···
γ ∈Un−1 (Z)\S L(n−1,Z) m 1 =1 m n−2 =1 m n−1 =0
)
n−1
|m k |k(n−k)/2
k=1
γ
× WJacquet M · z, ν f , ψ1,...,1,m n−1 /|m n−1 | ,
1
∞
∞ B(m 1 , . . . , m n−1 )
g(z) = ···
γ ∈Un−1 (Z)\S L(n−1,Z) m 1 =1 m n−2 =1 m n−1 =0
)
n−1
|m k |k(n−k)/2
k=1
γ
× WJacquet M · z, νg , ψ1,...,1, m n−1 /|m n−1 | ,
1
(12.1.1)
as in (9.1.2).
12.1 The GL(n) × GL(n) convolution 367
Proof The proof follows as in Section 7.4. The key point is that the identity
(7.4.14) generalizes. In this case, we have
A p k1 , p k2 , . . . , p kn−1 = Sk1 ,...,kn−1 (α p,1 , α p,2 , . . . , α p,n ),
and Theorem 12.1.3 is a consequence of Cauchy’s identity, Proposition 7.4.20.
Our main result is the following generalization of Theorems 7.2.4 and 7.4.9.
Theorem 12.1.4 (Functional equation) For n ≥ 2, let f, g be two Maass
forms of types ν f , νg for S L(n, Z), as in (12.1.1), whose associated L-functions
L f , L g , satisfy the functional equations:
n −s+λi (ν f ) s − λi (ν f )
f (s) := π 2 L f (s) = f˜ (1 − s)
i=1
2
n −s+λ j (νg ) s − λ j (νg )
g (s) := π 2 L g (s) = g̃ (1 − s),
j=1
2
as in Theorem 10.8.6 and Remark 10.8.7 where f˜, g̃, are the dual Maass forms
(see Section 9.2). Then the Rankin–Selberg L-function, L f ×g (s) (see Defini-
tion 12.1.2) has a meromorphic continuation to all s ∈ C with at most a simple
368 Rankin–Selberg convolutions
Remark It follows from (10.8.5) and Remark 10.8.7 that the powers of π
occurring in the above functional equations take the much simpler form:
n
−s+λi (ν)
n
n −s+λi (ν f )+λ j (νg )
= π −ns/2 , = π −n
2
π 2 π 2 s/2
.
i=1 i=1 j=1
Now
⎛ ⎞
∗ ··· ∗ ∗
⎜ .. .. .. ⎟
⎜ ··· .⎟
P = ⎜. . ⎟ ⊂ ,
⎝∗ ··· ∗ ∗⎠
0 ··· 0 1
is generated by matrices of type
⎛ ⎞ ⎛ ⎞
0 0 ··· 0 r1
⎜ .. ⎟ ⎜ .. .. .. ⎟
⎜ .⎟ ⎜.
⎜ mn−1 ⎟, ⎜ ··· . . ⎟
⎟,
⎜ ⎟ ⎝0
⎝ 0⎠ ··· 0 rn−1 ⎠
0 ··· 0 1 0 ··· 0 1
12.1 The GL(n) × GL(n) convolution 369
It follows that
say.
The meromorphic continuation and functional equation of L f ×g (s) now
follows from the meromorphic continuation and functional equation of
E P (z, s), (given in Proposition 10.7.5) provided we know the meromorphic
continuation and functional equation of G ν f ,νg (s), the Mellin transform of the
product of Whittaker functions. This latter functional equation was obtained in
a very explicit form by Stade (2001) which allows one to prove the functional
equation stated in Theorem 12.1.4.
370 Rankin–Selberg convolutions
The correct form of the functional equation can immediately be seen. Of course,
one needs to deal with the difficult convergence problems when taking Rankin–
Selberg convolutions which are not of rapid decay, but the techniques for dealing
with this are now well known (Zagier, 1982), (Liemann, 1993).
A(m, 1, . . . , 1) m (n−1)/2 ,
which is easily proved by shifting the line of integration to the left and apply-
ing Cauchy’s residue theorem. In view of the fact that the Rankin–Selberg
L-function, L f × f (s), specified in Definition 12.1.2 has positive coefficients
and |A(1, . . . , 1, m)|2 occurs as a coefficient of m −s , one obtains for fixed σ,
sufficiently large and all |m| sufficiently large that
σ+i∞
1 (2m)s
|A(1, . . . , 1, m)|
2
L f × f (s) ds. (12.1.7)
2πi s
σ −i∞
Remark 12.1.8 Note that the method for proving Proposition 12.1.6 also
shows that A(m 1 , m 2 , . . . , m n−1 ) behaves like a constant on average. If we
write
∞
b(m)
L f × f (s) =
m=1
ms
1
The power Det(z)s− 2 , above, was chosen to make the final formulae as simple
as possible. In view of the Fourier expansion of g given in (12.2.1), we may
apply the unfolding trick to g and obtain
1
f · ḡ, |Det(∗)|s̄− 2
∞ ∞ B(m 1 , . . . , m n )
= ··· f (z)
)
n
m 1 =1 m n−1 =1 m n =0 |m k |k(n+1−k)/2 Un (Z)\hn
k=1
⎛⎛ ⎞ ⎞
m 1 · · · |m n |
⎜⎜ .. ⎟ ⎟ s− 1
⎜⎜ ⎟ z ⎟
× W̄Jacquet⎜⎜ . ⎟ , νg , ψ1,...,1, |mmnn | ⎟ Det(z) 2 d ∗ z.
⎝⎝ m1 ⎠ 1 ⎠
1
374 Rankin–Selberg convolutions
where In is the n × n identity matrix. Since f (z) and d ∗ z are invariant under
scalar multiplication, it follows that
1
f · ḡ, |Det(∗)|s̄− 2
∞ ∞ B(m 1 , . . . , m n )
= ··· f (z)
)
n
m 1 =1 m n−1 =1 m n =0 |m k |k(n+1−k)/2 Un (Z)\hn
k=1
⎛⎛ ⎞ ⎞
m 2 · · · m n y1 · · · yn−1
⎜⎜ m · · · m y · · · y ⎟ ⎟
⎜⎜ 2 n−1 1 n−2 ⎟ ⎟
⎜⎜ ⎟ ⎟
⎜⎜ .. ⎟ ⎟
× WJacquet ⎜ ⎜⎜
⎜ . ⎟, νg , ψ1,...,1 ⎟
⎟ ⎟
⎜⎜ m 2 y1 ⎟ ⎟
⎜⎜ ⎟ ⎟
⎝⎝ 1 ⎠ ⎠
1
−n (s− 12 )
× e−2πi (m 2 x1 +m 3 x2 +···+m n xn−1 ) m 1 Det(y)s− 12 d ∗ z
∞ ∞ A(m 2 , . . . , m n ) B(m 1 , . . . , m n )
= ··· n n−1 s
m 1 =1 m n−1 =1 m n =0 m 1 m 2 · · · |m n |
∞ ∞
y
, νg , ψ1,...,1 Det(y)s− 2 d ∗ y,
1
× · · · WJacquet y, ν f , ψ1,...,1 W̄Jacquet
1
0 0
(12.2.3)
because the above integral in d ∗ x picks off the (m 2 , m 3 , . . . , m n )th Fourier
) −k(n−k)
n−1
coefficient (Theorem 5.3.2) and d ∗ y = yk dyk /yk as in (1.5.4).
k=1
Define
1
f ×g (s) := f · ḡ, |Det(∗)|s̄− 2 . (12.2.4)
Theorem 12.2.5 (Functional equation) Let f, g be Maass forms asso-
ciated to S L(n, Z), S L(n + 1, Z), respectively, with Fourier expansions
given by (12.2.1). Then the Rankin–Selberg L-function, L f ×g (s), defined in
Definition 12.2.2 has holomorphic continuation to all s ∈ C and satisfies the
functional equation
f ×g (s) = f˜×g̃ (1 − s),
where f ×g (s) is defined in (12.2.4), (12.2.3), and f˜, g̃ denote the dual Maass
forms as in Section 9.2.
12.2 The GL(n) × GL(n + 1) convolution 375
z −→ w · t (z −1 ) · w −1 := ι z,
where w is the long element as in Proposition 9.2.1. It follows from (9.2.2) that
⎛ ⎞⎛ ⎞
1 δx1 y1 y2 · · · yn−1
⎜ 1 −x2 ∗ ⎟⎜ y2 y3 · · · yn−2 ⎟
⎜ ⎟⎜ ⎟
ι ⎜ . . ⎟ ⎜ . ⎟
z=⎜ .. .. ⎟⎜ .. ⎟,
⎜ ⎟⎜ ⎟
⎝ 1 −xn−1 ⎠ ⎝ yn−1 ⎠
1 1
!n/2"+1
where δ = (−1) . It follows from Proposition 9.2.1, and the fact that f, g
are automorphic that
ι
˜f (ι z) = f (z), z z
g̃ =g .
1 1
Consequently,
1
f ×g (s) = f · ḡ, |Det(∗)|s̄− 2
z
|Det(z)|s− 2 d ∗ z
1
= f (z) · g
1
S L(n,Z)\hn
ι
z
f˜(ι z) · g̃ |Det(z)|s− 2 d ∗ z.
1
= (12.2.6)
1
S L(n,Z)\hn
It follows from the above remarks, after applying the substitution (12.2.7) to
(12.2.6), that formally
wz
|Det(z)| 2 −s d ∗ z
1
f ×g (s) = f (wz) · g̃
˜
1
S L(n,Z)\hn
z
|Det(z)| 2 −s d ∗ z
1
= f˜(z) · g̃
1
S L(n,Z)\hn
= f˜×g̃ (1 − s).
The above formal proof can be made rigorous by breaking the integrals from 0 to
∞ into two pieces [0, 1], [1, ∞] and applying Riemann’s method for obtaining
376 Rankin–Selberg convolutions
the functional equation of the Riemann zeta function. We will not pursue this
here.
∞
∞ A(m 1 , . . . , m n−1 )
f (z) = ···
γ ∈Un−1 (Z)\S L(n−1,Z) m 1 =1 m n−2 =1 m n−1 =0
)
n−1
|m k |k(n−k)/2
k=1
γ
× WJacquet M · z, ν f , ψ1,...,1, mn−1 ,
1 |m n−1 |
(12.3.1)
∞
∞ B(m 1 , . . . , m n −1 )
g(z) = ··· n) −1
γ ∈Un −1 (Z)\S L(n −1,Z) m 1 =1 m n −2 =1 m n −1 =0 |m k |k(n −k)/2
k=1
γ
× WJacquet M · z, νg , ψ1,...,1, m n −1 .
1 |mn −1 |
where
G L(n, R) ∗
Pn,1 (R) = ⊂ G L(n + 1, R).
0 1
12.3 The G L(n) × G L(n ) convolution with n < n 377
Let g be a Maass form for S L(n , Z) as in (12.3.1). For z ∈ Pn,1 (R) we define
⎛⎛ ⎞⎞
u 1,n+2 ··· u 1,n
⎜⎜ .. .. ⎟⎟
⎜⎜ z . ··· . ⎟ ⎟
⎜⎜ ⎟⎟
⎜⎜⎜⎜ .. ⎟⎟⎟
1 1
⎟
n −(n −n−1)/2 ⎜⎜ u · · · . ⎟⎟
Pn (g)(z) := |Det(z)| · · · g ⎜⎜ n+1,n+2
⎟⎟
⎜⎜ .. .. ⎟⎟
⎜⎜ 1 . . ⎟ ⎟
0 0 ⎜⎜ ⎟⎟
⎜⎜ . ⎟⎟
⎝⎝ .. u
n −1,n
⎠⎠
1
× e−2πi (u n+1,n+2 +u n+2,n+3 +···+u n −1,n ) du i, j .
n+2≤ j≤n
1≤i< j
It is clear that if p ∈ Pn,1 (Z), then Pnn (g)( pz) = Pnn (g)(z).
Lemma 12.3.3 Fix 2 ≤ n < n − 1, and let g be given by (12.3.1). Then for
z ∈ Pn,1 (R), we have
Pnn (g)(z)
∞
∞
B(1, . . . , 1, m n −n , . . . , m n −1 )
= |Det(z)|−(n −n−1)/2 ··· n)
−1
γ ∈Un (Z)\S L(n,Z) m n −n =1 m n −1 =1
|m k |k(n −k)/2
k=n −n
⎛⎛ ⎞ ⎞
m n −n · · · m n −1
⎜⎜ m n −n · · · m n −2 ⎟ ⎟
⎜⎜ ⎟ ⎟
⎜⎜ .. ⎟ γz ⎟
× WJacquet ⎜⎜ . ⎟· , νg , ψ1,...,1 ⎟,
⎜⎜ ⎟ In −n ⎟
⎝⎝ m n −n ⎠ ⎠
In −n
where Ir denotes the r × r identity matrix.
Proof Let z ∈ Pn,1 (R). Since Pnn (g(z)) is invariant under left multiplication
by Pn,1 (Z) it follows from Section 5.3 that it has a Fourier expansion as in
Theorem 5.3.2. After some computation, the proof follows as in the proof of
Theorem 9.4.7.
With these preliminaries out of the way, we may proceed to describe the
Rankin–Selberg convolution. Let f, g be Maass forms for S L(n, Z), S L(n , Z),
respectively as in (12.3.1). The requisite convolution is given by the
1
inner product, f · Pnn (g), |Det(∗)|s̄− 2 , taken over the fundamental domain
378 Rankin–Selberg convolutions
S L(n, Z)\G L(n, R). Lemma 12.3.3 allows us to unravel Pnn (g) in this inner
product. It follows that
1
f · Pnn (g), |Det(∗)|s̄− 2
∞ ∞
B(1, . . . , 1, m n −n , . . . , m n −1 )
= f (z) ··· n)
−1
m n −n =1 m n −1 =1
Un (Z)\G L(n,R) |m k |k(n −k)/2
k=n −n
⎛⎛ ⎞ ⎞
m n −n · · · m n −1
⎜⎜ m n −n · · · m n −2 ⎟ ⎟
⎜⎜ ⎟ ⎟
⎜⎜ .. ⎟ z ⎟
× W Jacquet ⎜⎜ . ⎟· , νg , ψ1,...,1 ⎟
⎜⎜ ⎟ In −n ⎟
⎝⎝ m n −n ⎠ ⎠
In −n
n −n
× |Det(z)|s− 2 d ∗ z.
∞
B(m, 1, . . . , 1) n
L g (s) = s
= (1 − β p,i p −s )−1 ,
m=1
m p i=1
Proof The proof follows the ideas in Section 7.4, but requires a modification
of Cauchy’s identity (Proposition 7.4.20). For details, see (Bump, 1984, 1987,
to appear).
as in Remark 10.8.7 where f˜, g̃ are the dual Maass forms (see Section 9.2). Then
the Rankin–Selberg L-function, L f ×g (s) (see Definition 12.3.4) has a holomor-
phic continuation to all s ∈ C. Furthermore, L f ×g (s) satisfies the functional
equation
n n −s+λi (ν f )+λ j (νg ) s − λi (ν f ) − λ j (νg )
f ×g (s) := π 2 L f ×g (s)
i=1 j=1
2
= f˜×g̃ (1 − s).
380 Rankin–Selberg convolutions
Remark Note that as in the remark after Theorem 12.1.4, the power of π in
the above theorem simplifies to π −nn s/2 .
z −→ w · t (z −1 ) · w−1 := ι z,
f˜ (ι z) = f (z),
and
ι
z z
g̃ =g .
In −n In −n
Consequently,
1
f ×g (s) = f · Pnn (g), |Det(∗)|s̄− 2
z
|Det(z)|s− 2 d ∗ z
1
= f (z) · Pnn (g)
I n −n
S L(n,Z)\G L(n,R)
ι
z
f˜(ι z) · Pnn (g̃) |Det(z)|s− 2 d ∗ z.
1
=
In −n
S L(n,Z)\G L(n,R)
(12.3.7)
It follows from the above remarks, after applying the substitution (12.3.8) to
(12.3.7), that formally
z
|Det(z)| 2 −s d ∗ z
1
f ×g (s) = f˜(wz) · ι ◦ Pnn ◦ ι (g̃)
1
S L(n,Z)\G L(n,R)
z
|Det(z)| 2 −s d ∗ z
1
= f˜(z) · ι ◦ Pnn ◦ ι (g̃)
1
S L(n,Z)\G L(n,R)
=
˜ f˜×g̃ (1 − s),
where
1
˜ f˜×g̃ (s) = f˜ · ι ◦ Pnn ◦ ι (g̃),
|Det(∗)|s̄− 2 .
The above formal proof can be made rigorous by breaking the integrals from 0 to
∞ into two pieces [0, 1], [1, ∞] and applying Riemann’s method for obtaining
the functional equation of the Riemann zeta function.
Once the form of the functional equation is obtained, the precise Gamma
factors in the functional equation can be deduced by using the functional equa-
tion of the minimal parabolic Eisenstein series as a template as we did at the
end of Section 12.1. We leave these computations to the reader.
where
⎛ ⎞
m 1 · · · m n−2 · |m n−1 |
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
M =⎜ m1m2 ⎟, A(m 1 , . . . , m n−1 ) ∈ C,
⎜ ⎟
⎝ m1 ⎠
1
382 Rankin–Selberg convolutions
and
⎛⎛ ⎞⎞
1 u n−1
⎜⎜ 1 u n−2 ∗⎟ ⎟
⎜⎜ ⎟⎟
⎜⎜ ⎟⎟
ψ1,...,1, ⎜⎜ ..
.
..
. ⎟⎟ = e2πi (u 1 +···+u n−2 +u n−1 ) .
⎜⎜ ⎟⎟
⎝⎝ 1 u 1 ⎠⎠
1
We further assume that f is normalized so that A(1, . . . , 1) = 1 and that f is an
eigenfunction of the Hecke operators given in (9.3.5). It was shown in (9.4.2)
that the Godement–Jacquet L-function
∞
A(n, 1, . . . , 1)
L f (s) =
n=1
ns
has an Euler product given by
L f (s) = 1 − A( p, . . . , 1) p −s + A(1, p, . . . , 1) p −2s −
p
−1
· · · + (−1)n−1 A(1, . . . , p) p (1−n)s + (−1)n p −ns
n
= (1 − α p,i p −s )−1 , (12.4.2)
p i=1
This conjecture has been proved for holomorphic modular forms on G L(2)
in (Deligne, 1974) (see Section 3.6) but is still a major unsolved problem for
Maass forms. Deligne shows that every holomorphic modular form (Hecke
eigenform) over Q, say, is associated to an algebraic variety and that the pth
Fourier coefficient can be interpreted in terms of the number of points on this
variety over F p , the finite field of p elements. From this point of view, the
Ramanujan conjecture is equivalent to the Riemann hypothesis for varieties
over finite fields first conjectured in (Weil, 1949) and, in a stunning tour de
12.4 Generalized Ramanujan conjecture 383
force, finally proved in (Deligne, 1974). The problem with trying to generalize
this approach to non-holomorphic automorphic forms is that there seem to be no
visible connections between the theory of Maass forms and algebraic geometry.
There is yet another fundamental conjecture which was originally formu-
lated by Selberg (1965) (see also Section 3.7). Classically this is known as the
Selberg eigenvalue conjecture. For the cognoscenti, it is clear from the adelic
point of view that the Selberg eigenvalue conjecture is really the generalized
Ramanujan conjecture at the infinite prime. In the next section, we will give a
more elementary explanation of why these two conjectures can be placed on an
equal footing. Here is the generalized Selberg eigenvalue conjecture.
Conjecture 12.4.4 (Selberg eigenvalue conjecture) For n ≥ 2, let f (z) be a
Maass form of type (ν1 , ν2 , . . . , νn−1 ) ∈ Cn−1 for S L(n, Z) as in (12.4.1). Then
1
(νi ) = , (λi (ν)) = 0,
n
for i = 1, 2, . . . , n − 1, with λi as defined in Theorem 10.8.6 and
Remark 10.8.7.
Remark The first to observe that the classical Ramanujan conjecture con-
cerning the Fourier coefficients of (z) can be reformulated to a very general
conjecture on G L(n) appears to have been Satake (1966). The generalized
Ramanujan conjecture has been established for automorphic cusp forms of
G L(n, F), of algebraic type satisfying a Galois conjugacy condition, where F
is a complex multiplication field (see (Harris and Taylor, 2001)). The proof is
a remarkable tour-de-force combining the Arthur–Selberg trace formula and
the theory of Shimura varieties. The generalized Ramanujan conjecture is still
unproven for Maass forms for S L(n, Z) of the type considered in this book.
While the Selberg eigenvalue Conjecture 12.4.4 is not hard to prove for
S L(2, Z) (see Theorem 3.7.2) it is still an unsolved problem for congruence
subgroups (see Conjecture 3.7.1); and, of course, Conjecture 12.4.4 can be easily
generalized to higher level congruence subgroups of S L(n, Z) with n ≥ 2.
In (Kim and Sarnak, 2003), one may find the current world record for both the
Ramanujan conjecture and Selberg eigenvalue conjecture for G L(2) (this
includes the case of congruence subgroups of higher level). The precise bounds
obtained are
7
|α p,1 |, |α p,2 | ≤ p 64 ,
and correspondingly,
7
(ν) ≤ .
64
384 Rankin–Selberg convolutions
In (Selberg, 1965) the bound ≤ 1/4 instead of ≤ 7/64 as in (Kim and Sarnak,
2003) was attained. Selberg’s result was slightly improved to < 1/4 in (Gelbart
and Jacquet, 1978). Further improvements were given in (Serre, 1981) (see also
1 1
(Serre, 1977)), who obtained |α p,i | ≤ p 5 , Shahidi (1988), |α p,i | < p 5 , Duke
1
and Iwaniec (1989), found a different proof of |α p,i | ≤ 2 p 5 , Bump, Duke,
5
Hoffstein and Iwaniec (1992), |α p,i | ≤ p 28 . In the case of G L(n) with n ≥ 2,
Jacquet and Shalika (1981) obtained the bound < 1/2. This bound was proved
for S L(n, Z) in Propositions 12.1.6, 12.1.9. For the finite primes p, Serre (1981)
suggested that one could do better by a clever use of the Rankin–Selberg L-
function. This led to the bound
1 1
≤ − 2 (12.4.5)
2 dn + 1
for Maass forms on G L(n) over a number field of degree d. In (Luo, Rudnick and
Sarnak, 1995, 1999) the dependence on d was removed. We give an exposition
of the Luo, Rudnick and Sarnak method (over Q) in the next section.
Remark In (Luo, Rudnick and Sarnak, 1999) the above bound was obtained
for Maass forms on G L(n) over an arbitrary number field.
Proof Recall the definition of L f (s) given in (12.4.2).
n
L f (s) = (1 − α p,i p −s )−1 .
i=1 p
12.5 The Luo–Rudnick–Sarnak bound for Ramanujan conjecture 385
Further, as in the proof of Theorem 12.1.4, Remark 10.8.7, and the comments
after it, the function L ( f ⊗χ )× f (s) satisfies the functional equation
( f ⊗χ)× f (s)
n n s+aχ −λi (ν)−λ j (ν)
q 2 s − λi (ν) − λ j (ν) + aχ
= L ( f ⊗χ )× f (s)
i=1 j=1
π 2
= χ · ( f˜⊗χ̄ )× f˜ (1 − s), (12.5.2)
where
√ 2
i aχ q n 0 if χ (−1) = 1,
χ = , aχ =
τ (χ ) 1 if χ (−1) = −1.
It follows from the methods used to prove Theorem 12.1.4 that ( f ⊗χ )× f (s)
has a holomorphic continuation to all s ∈ C except for simple poles at s = 1
and s = 0, the latter simple poles can only occur if χ ≡ 1 is the trivial character.
Note that at finite level there can be finitely many χ for which this has a pole
(of course, this impacts nothing). This result was also proved in much greater
generality in the combination of papers: (Shahidi, 1981, 1985), (Jacquet and
Shalika, 1981, 1990), (Jacquet, Piatetski-Shapiro and Shalika, 1983), (Moeglin
and Waldspurger, 1989).
We also require, for any fixed prime p0 , the modified Rankin–Selberg L-
function
n
n
1 − α p0 ,i α p0 , j χ ( p0 ) p0−s , (12.5.3)
p
L ( f0 ⊗χ)× f (s) = L ( f ⊗χ )× f (s) ·
i=1 j=1
which is the same as L ( f ⊗χ )× f (s) except that the Euler factor at p0 has been
removed.
The key idea in the proof of Theorem 12.5.1 is the following lemma.
386 Rankin–Selberg convolutions
We defer the proof of this lemma until later and continue with the proof of
Theorem 12.5.1.
It follows from (12.5.3) that
n
n
−1
1 − α p0 ,i α p0 , j χ ( p0 ) p0−s
p
L ( f ⊗χ)× f (s) = L ( f0 ⊗χ )× f (s) · .
i=1 j=1
or
n
n
s − λi (ν) − λ j (ν)
, (12.5.5)
i=1 j=1
2
p
must be a zero of L ( f0 ⊗χ )× f (s).
Assume that χ ( p0 ) = 1 and for some 1 ≤ i ≤ n we have |α p0 ,i |2 = p β with
β > 1 − (2/(n 2 + 1)). Then (1 − |α p0 ,i |2 p0−s )−1 has a pole at s = β. Sim-
ilarly, assume that (λi (ν)) = β for some i = 1, . . . , n − 1, and for some
β > 1 − n 22+1 . Then the Gamma factor
n
s − 2(λi (ν))
i=1
2
Remark The above proof yields a slightly better bound for the case of
G L(2) Maass forms when combined with the Gelbart–Jacquet lift. For example,
12.5 The Luo–Rudnick–Sarnak bound for Ramanujan conjecture 387
With the notation of (12.5.3), for a primitive character χ (mod q), the func-
tional equation (12.5.2) may be rewritten in the form.
q (n 2 /2)−n 2 s
p p
L ( f0 ⊗χ )× f (s) = χ G p0 (s) L ( f0¯⊗χ̄)× f¯ (1 − s), (12.5.6)
π
where
) 1−s−λi (ν)−λ j (ν) −1
n )n
2
1 − α p0 ,i α p0 , j p0−(1−s)
i=1 j=1
G p0 (s) = −1 .
)n ) n
α p0 ,i α p0 , j p0−s
s−λi (ν)−λ j (ν)
2
1−
i=1 j=1
It follows easily from the Euler product (Theorem 12.1.3) and the coefficient
bound in Proposition 12.1.6 that for β ≥ 1, we have L ( f ⊗χ )× f (β) = 0. This was
proved in much greater generality in (Shahidi, 1981). Consequently, we may
assume that 1 − (2/(n 2 + 1)) < β < 1. The key idea of the proof is to show
that for > 0, and Q sufficiently large that
∗ p
L ( f0 ⊗χ )× f (β) Q 2− , (12.5.7)
q∼Q χ (mod q)
+∗
where means that the sum ranges over primitive characters χ (mod q)
+
satisfying χ ( p0 ) = χ (−1) = +1, and means we sum over primes
q∼Q
Q ≤ q ≤ 2Q. The lower bound (12.5.7) immediately implies Lemma 12.5.4.
To prove (12.5.7), we require an auxilliary compactly supported smooth
function h : [A, B] → R, where 0 < A < B, and h(y) ≥ 0, and in addition
∞
h(y)
dy = 1.
0 y
388 Rankin–Selberg convolutions
Define
∞
dy
h̃(s) := h(y)y s ,
y
0
while
h 1 (y) = 0, if y ≥ B.
where
and, we may recall that the λi (ν) occur in the Gamma factors of G p0 (s) as in
(12.5.6). To prove (12.5.8) shift the line of integration to the right to the line
(s) = M. The result follows since G p0 (s) has at most polynomial growth in
s in fixed vertical strips while h̃(−s) has rapid decay. To prove (12.5.9), shift
the line of integration to the left. We may assume β0 + β − 1 < 0, otherwise,
(12.5.9) is obvious. After shifting to the left, we will pick up the first simple
pole at (s) = β0 + β − 1. The bound (12.5.9) follows immediately.
The next step in the proof is the derivation of an approximate functional
equation for L ( f ⊗χ )× f (s) of the type previously derived in Section 8.3. The
12.5 The Luo–Rudnick–Sarnak bound for Ramanujan conjecture 389
On the other hand, we may shift the line of integration to the left, apply the
functional equation (12.5.6), and then let s → −s, to obtain
1 ds
I(β, Y ) := L ( f ⊗χ )× f (β) + h̃(s) L ( f ⊗χ )× f (s + β) Y s ,
2πi s
(s)=−1
(12.5.11)
which is equal to
n 2
q n22 (1−2β) ∞
b̄(m)χ̄ (m) π
χ h 2 mY · .
π m=1
m 1−β q
Combining (12.5.10), (12.5.11), and the above calculation, we obtain, for any
Y > 1, the approximate functional equation
∞
b(m) m
L ( f ⊗χ )× f (β) = χ (m) · h 1
m=1
mβ Y
q n22 (1−2β) ∞
b̄(m) mY π n
2
− χ χ̄ (m) · h 2 .
π m=1
m 1−β q n2
390 Rankin–Selberg convolutions
− n2
χ̄ (m) · h 2 . (12.5.12)
π 2 (1−β)
m=1
m 1−β q n2
The final step in the proof of Lemma 12.5.4 uses the approximate func-
tional equation (12.5.12) to deduce (12.5.7). We shall restrict ourselves to
prime q so that all non-trivial characters are automatically primitive. We
require:
⎧
⎪
⎪ m ≡ 0 (mod q)
⎨0,
χ (m) = q−1
− 1, m ≡ ±1 (mod q) (12.5.13)
⎪
⎪
2
χ (mod q) ⎩−1, otherwise.
χ=χ0 , χ(−1) = +1
The contribution of the first sum on the right-hand side of (12.5.12) to the
average (12.5.14) is
∞
b(m) m
χ (m) · h 1
q∼Q χ (mod q) m=1
mβ Y
χ=χ0 , χ (−1) = +1
q −1 b(m) m b(m) m
= h1 − h1 .
q∼Q
2 m≡±1 (mod q) m β Y q∼Q (m,q)=1
mβ Y
(12.5.15)
We will show that the main contribution to (12.5.15) comes from the term
m = 1. All other terms will give a much smaller contribution. In fact, the
contribution of the term m = 1 is
q −1 1 q −1
h1 = .
q∼Q
2 Y q∼Q
2
where we have used the fact that for m = 1, the number of different
representations m = 1 + dq = 1 + d q is O(m ). By the properties of the
Rankin–Selberg convolution L f × f (s), (see Section 12.1, and in particular,
Remark 12.1.8) we see that b(m) ≥ 0 of all m = 1, 2, . . . , and
b(m) ∼ c f x, (x → ∞)
m≤x
+ +
∞
b(m)m
for some c f > 0. It follows that m (β)
|h 1 (m/Y )| QY 1−β+ and,
q∼Q m=1
therefore,
q −1 b(m) m
h1 QY 1−β+ .
q∼Q
2 m≡±1 (mod q)
mβ Y
m=1
By the same type of computation, one also obtains a bound for the last sum on
the right-hand side of (12.5.15):
b(m) m
h1 QY 1−β+ .
q∼Q (m,q)=1
mβ Y
χ̄(m) · h 2
q∼Q χ (mod q)
π (n 2 /2)(1−β) m=1 m 1−β q n2
χ=χ0 ,χ (−1)=+1
q n 2+1 −n 2 β |b(m)| mY π n
2
∞ 2
· h 2
n2
q∼Q π 2
(1−β)
m=1
m 1−β q n2
q n 2+1 −n 2 β
∞
r Y π n β dr
2 2
f2 r
n2
q∼Q π 2
(1−β) q n2 r
1
q
n 2 +1
2
n2
q∼Q π 2 (1+β)
· Yβ
2
1+ n 2+1
Q · Y −β .
Collecting together all the previous computations with the approximate func-
tional equation (12.5.12) yields the asymptotic formula
q −1 n 2 +1
L ( f ⊗χ)× f (β) = + O QY 1−β+ + Q 1+ 2 · Y −β .
q∼Q χ (mod q) q∼Q
2
χ=χ0 ,χ(−1)=+1
The case when p0 is a prime is more difficult to deal with. In this situtation
we replace (12.5.13) with
∗ = Nq if m = 1,
χ (m) = (12.5.18)
β|q χ (mod β) ≥0 otherwise,
χ ( p0 )=χ(−1)=1
where
∗
Nq = 1,
β|q χ (mod β)
χ ( p0 )=χ(−1)=1
The key point is that one needs to know that Nq is large for many values of
q. More precisely, one requires that for every > 0 and all Q sufficiently large
Nq Q 2− . (12.5.19)
Q<q<2Q
Theorem 12.6.1 (Strong multiplicty one) Let f, g be two Maass forms for
S L(n, Z) as in (12.1.1) with Fourier coefficients
A( p, 1, . . . , 1) = B( p, 1, . . . , 1)
not have a pole then by standard techniques in analytic number theory the sum
+
A( p, 1, . . . , 1) · B( p, 1, . . . , 1) would be small as x → ∞.
p≤x
About 25 years ago I was discussing analytic number theory with Jean-Pierre
Serre. I distinctly recall how he went to the blackboard and wrote down the
Euler product
n
(1 − α p,i p −s )−1
p i=1
corresponding to an automorphic form on G L(n), and then pointed out that one
of the most important problems in the theory of L-functions was to obtain the
analytic properties of the higher kth symmetric power L-functions given by
−1
1 − α p,i1 α p,i2 · · · α p,ik p −s .
p 1≤i 1 ≤i 2 ≤···≤i k ≤n
http://www.sunsite.ubc.ca/DigitalMathArchive/Langlands/functoriality.html
395
396 Langlands conjectures
−1
1 − α p,i1 α p,i2 · · · α p,ik p −s .
p i 1 <i 2 <···<i k
The local Langlands conjectures (Carayol, 1992, 2000) have seen much
greater advances. In (Laumon, Rapoport and Stuhler, 1993), the local Lang-
lands conjecture was proved for local fields of prime characteristic. This was
followed by Harris and Taylor (2001) who gave a proof for characteristic zero
and then Henniart (2000) gave a simplified proof. In (Drinfeld, 1989) a proof
of Langlands functoriality conjecture was obtained for G L(2) over a function
field. Finally Lafforgue (2002) established Langlands conjectures for G L(n)
(with n ≥ 2) in the function field case.
Other references for Langlands conjectures include (Borel, 1979), (Bump,
1997), (Gelbart, 1984), (Arthur, 2003), (Bernstein and Gelbart, 2003), (Moreno,
2005).
ρ : G → G L(V )
Now let
1 1 1
α = 23 , β = e2πi/3 · 2 3 , γ = e4πi/3 · 2 3 .
pO K = P1e · · · Pre
where
= [G : DP ] then the distinct conjugate divisors to P are just the divisors
P gi , (1 ≤ i ≤
). It follows that
= r and the order of DP is e f. For unramified
primes this is just f .
Consider the residue fields FP := O K /P and Fp := Ok /p, respectively.
Then FP , Fp are finite fields.It is not hard to see that the elements of the
decomposition group DP are automorphisms of FP /Fp. Indeed, if g ∈ DP and
α g ≡ β g (mod P).
large. The Artin L-function (denoted L(s, ρ, K /k)) attached to this data is given
by the Euler product
−1 −1
Det I − ρ (FrP ) N (p)−s Det I − ρ (FrP ) V IP N (p)−s ,
p unramified p ramified
where I is the identity matrix and P is any prime above p. Furthermore, for p
ramified, the quotient DP /IP acts on the subspace V IP of V on which IP acts
trivially. The notation ρ (FrP ) |V IP means that the action of ρ (FrP ) is restricted
to V IP .
Artin’s method is to show that in spite of the differences in the definitions the function
L(s, ρ, K /F) attached to a one-dimensional ρ is equal to a Hecke L-function L(s, χ)
where χ = χ (ρ) is a character of F ∗ \I F . He employed all the available resources
of class field theory, and went beyond them, for the equality of L(s, ρ).
It is not hard to show (see Heilbronn’s article in (Cassels and Fröhlich, 1986))
that Artin’s L-function L(s, ρ, K /k) can be expressed as a product of rational
powers of abelian L-functions of Hecke’s type, where the abelian L-functions
are associated to intermediate fields k ⊆ ⊆ K with K / abelian. A major
advance on Artin’s conjecture in the case when Gal(K /k) is not an abelian
402 Langlands conjectures
group was made by Brauer (1947) who proved that all irreducible represen-
tations ρ of the Galois group G can be expressed as Z-linear combinations
of induced representations of one-dimensional representations on subgroups
of G. As a consequence, he showed that L(s, ρ, K /k) can be expressed as a
product of integer powers of abelian L-functions of Hecke’s type. This proved
that Artin’s L-functions extended to meromorphic functions in the entire com-
plex s plane and satisfied a functional equation. The problem was that Brauer
could not exclude negative integral powers so that Artin’s conjecture was still
unproven.
Further advances on Artin’s Conjecture 13.1.19 did not come until Langlands
changed the entire landscape of research around this problem by making the
striking conjecture that Artin’s L-functions should be L-functions associated to
Maass forms on G L(n).
When n = 2 and the image of ρ in P G L(2, C) is a solvable group, Artin’s
conjecture was solved in (Langlands, 1980). The ideas in this paper played a
crucial role in Taylor and Wiles (1995) proof of Fermat’s Last Theorem. In
(Langlands, 1980) the conjecture is proved for tetrahedral and some octahedral
representations and in (Tunnell, 1981) the results are extended to all octahedral
representations. When n = 2 and the projective image is not solvable, the only
possibility is that the projective image is isomorphic to the alternating group A5 .
These representations are called icosahedral because A5 is the symmetric group
of the icosahedron. Joe Buhler’s Harvard Ph.D. thesis (see (Buhler, 1978)) gave
the first example where Artin’s conjecture was proved for an icosahedral repre-
sentation. The book (Frey, 1994) proves Artin’s conjecture for seven icosahedral
representations (none of which are twists of each other). In (Buzzard and Stein,
2002), the conjecture is proved for eight more examples. A further advance
was made in (Buzzard, Dickinson, Shepherd-Barron and Taylor, 2001) who
proved Artin’s conjecture for an infinite class of icosahedral Galois represen-
tations which were disjoint from the previous examples.Very little is known
for n > 2.
we shall attempt to motivate the definition and give a feeling for the important
program that Langlands has created.
The key new idea introduced by Langlands in his 1967 letter to Weil, and also
introduced independently in (Gelfand, Graev and Pyatetskii-Shapiro, 1990) is
the notion of an automorphic representation. We have intensively studied
Maass forms for S L(n, Z). These are examples of automorphic forms. The
leap to automorphic representation is a major advance in the subject with
profound implications.It was first intensively researched, for the case of G L(2),
in (Jacquet and Langlands, 1970).
In the interests of notational simplicity and an attempt by the writer to
explain in as simple a manner as possible the ideas behind the functorial-
ity conjectures of Langlands, we shall restrict our discussion, for n ≥ 2, to
the group S L(n, Z) which acts on G L(n, R) by left matrix multiplication. An
automorphic form is then a Maass form for S L(n, Z) as in Definition 5.1.3, a
Langlands Eisenstein series for S L(n, Z) as in Chapter 10, or the residue of such
an Eisenstein series. By Langlands spectral theorem (see (Langlands, 1966) and
also Theorem 10.13.1 for the case of G L(3)) these automorphic forms generate
the C-vector space
<
Vn := L2 S L(n, Z) G L(n, R) O(n, R) · R× .
ρ : G L(n, R) → End(Vn ),
where
(ρ(g)F)(z) := F(z · g)
n
−1
L(s, ρ , K /k) = 1 − λp,i N (p)−s ,
p j=1
Other examples of this type can be given by considering, for example, the
exterior product of two vector spaces (see Section 5.6), and then forming the
exterior product of two Galois representations or by taking higher symmetric
or exterior powers.
Yet another type of interesting operation that can be done with representa-
tions is to consider induced representations. This corresponds to induction from
a Galois subgroup (see (Bump, 1997). Langlands derived from this process his
theory of base change (Langlands, 1980).
An even deeper theorem in Galois representations is Artin’s reciprocity law
(Artin, 1927) (see also the introductory article (Lenstra-Stevenhagen, 2000))
which generalized Gauss’ law of quadratic reciprocity, and included all known
reciprocity laws up to that time. Langlands formulated an even more general
version of Artin reciprocity in the framework of automorphic representations.
Of course, Langlands did not stop here. This was the starting point. For a gen-
eral connected reductive algebraic group G he introduced, in (Langlands, 1970),
the dual group or what is now known as the L-group, L G.He then formulated
his now famous principle of functoriality which states that given any two con-
nected reductive algebraic groups H, G and an L-homomorphism L H → L G
then this should determine a transfer or lifting of automorphic representations
of H to automorphic representations of G.
In fact, the modern theory of automorphic forms allows us to parameterize
each automorphic representation of a reductive group such as H by a set of
semisimple conjugacy classes {cv } in LH , a complex Lie group, where v runs
over almost all the places of the defining global field. The functoriality prin-
ciple then roughly states that for each “L–homomorphism” f LH → LG, the
collection { f (cv )} defines, in fact, an automorphic representation of G, or more
precisely a “packet” of automorphic representations of G.
List of symbols
, d∗z gl(n, R)
, ⊗
(Rn ) G L(2, R)
,
(Rn ) 1 G L(3, R)
!, " 2 G L(n, R)
⊗
(Rn ) (z) G L(n, Z)
∧ Det G L(V )
∨ eψ Gw
⊗ E(z, s) Gw
An 1 ,...,nr E ∗ (z, s) γa
A(m 1 , . . . , m n−1 ) E(z, s) ∞
α p,i E(z, s, χ) (N )
αp E ∗ (z, s, χ) 0 (N )
Bn E P (z, s) w
bi, j E ∗P (z, s) h(d)
B(s1 , s2 ) E Pmin (z, s) h
C G () E P (z, s, φ) hn
C+∞ Em (z, s) ♥
χ (n) Ei j In
D d IP
Dn fˆ Ir
Dn f˜ Is
Dα Fq Iν (z)
Di j F(z, h, q) Iν (z, Pn 1 ,...,nr )
DP FrP Ks
d(n) G ν (s) K ν (y)
d(z, z ) G̃ ν (s) k(u)
d ∗u gcd k(z)
407
408 List of symbols
A.1 Introduction
This appendix is the manual for a set of functions written to assist the reader to
understand and apply the theorems on GL(n, R) set out in the main part of the
book. The software for the package is provided over the world wide web at
http://www.math.waikato.ac.nz/∼kab
A.1.1 Installation
First connect to the website given in the paragraph above and click on the link
for GL(n)pack listed under “Research” to get to the GL(n)pack home page.
Instructions on downloading the files for the package will be given on the home
page. If you have an earlier version of GL(n)pack first delete the file gln.m,
the documentation gln.pdf and the validation program glnval.nb. The name of
the file containing the package is gln.m. To install, if you have access to the file
system for programs on your computer, place a copy of the file in the standard
repository for Mathematica packages – this directory is called “Applications”
on some systems. You can then type gln.m and then press the Shift and Enter
keys to load the package. You may need administrator or super-user status to
complete this installation. Alternatively place the package file gln.m anywhere
in your own file system where it is safe and accessible.
409
410 Appendix The GL(n)pack Manual
SetDirectory[”c:\your\directory\path”];
<<gln.m;ResetDirectory[]<Shift/Enter>
or
Get[”gln.m”,Path->{”c:\your\directory\path”}]
<Shift/Enter>
Instructions for Unix/Linux systems These are the same as for Windows,
but the path-name syntax style should be like /usr/home/your/subdirectory.
Instructions for Macintosh systems These are the same as for Windows,
but the path-name syntax style should be like HD:Users:ham:Documents:.
All systems The package should load printing a message. The functions of
GL(n)pack are then available to any Mathematica notebook you subsequently
open.
There are many issues to do with computer algebra and mathematical soft-
ware that will arise in any serious evaluation or use of Mathematica and
GL(n)pack. A comment on one aspect: GL(n)pack function arguments are first
evaluated and then checked for correct data type. If the user calls a function
with an incorrect number of arguments or an argument of incorrect type, rather
than issue a warning and proceeding to compute (default Mathematica style),
GL(n)pack prints an error message, aborts the evaluation and returns the user
to the top-level, no matter how deeply nested the function which makes the
erroneous call happens to be placed. This is a tool for assisting users to debug
programs which include calls to this package.
A.1.7 Acknowledgements
Kevin Broughan had assistance from Columbia University and the University
of Waikato and a number of very helpful individuals while writing the package.
These included Ross Barnett, Mike Eastwood, Sol Friedberg, David Jabon (for
the Smith Normal Form code), Jeff Mozzochi and Eric Stade, in addition to
Dorian Goldfeld, who’s detailed text and explicit approach made it possible.
Their contribution is gratefully acknowledged.
A.2 Functions for GL(n)pack 413
m is a positive integer with value 2 or more being the order of the operator,
expr is a Mathematica expression, normally in the Iwasawa matrix or variables
and other parameters, which can be symbolically differentiated,
iwa is a numeric or symbolic matrix in Iwasawa form,
value is an expression or number being the result of applying the Casimir
operator in the Iwasawa variables to the expression.
Example
BlockMatrix (blm)
This function returns a specified sub-block of a matrix. The entries of the sub-
block must be contiguous.
Block matrices are used in a number of places, but most especially in
Chapter 10 on Langlands Eisenstein series.
a is a matrix of CREs,
rows is a list of two valid row indices for a, being the first and last sub-block
rows,
columns is a list of two valid column indices for a, being the first and last
sub-block columns,
b is the sub-block of a with the specified first and last row and column sub-block
indices.
Example
BruhatCVector (bcv)
In the explicit Bruhat decomposition of a non-singular matrix a, the diagonal
matrix c has a special form, each element being the ratio of absolute values of
minor determinants (ci ) of the original matrix a with the element in the (i, i)th
418 Appendix The GL(n)pack Manual
position being cn−i+1 /cn−i for 2 ≤ i ≤ n − 1 with the (n, n)th element being
c1 and the 1, 1th, det(w)det(a)/cn−1 . This function returns those ci .
See Section 10.3 and Proposition 10.3.6.
BruhatCVector[a] −→ c
Example
BruhatForm (bru)
This function finds the factors of a non-singular matrix, which may have entries
which are polynomial, rational or algebraic expressions, so that the matrix can
be expressed as the product of an upper triangular matrix with 1s on the leading
diagonal (unipotent), a diagonal matrix, a permutation matrix (with a single 1
in each row and column), and a second unipotent matrix. When an additional
constraint, namely, that the transpose of the permutation matrix times the second
upper triangular matrix is lower triangular, then the factors are unique. This is
the so-called Bruhat decomposition.
See Section 10.3 Proposition 10.3.6. The decomposition is used in
Section 10.6 to derive the Fourier expansion of a minimal parabolic Eisenstein
series.
A.3 Function descriptions and examples 419
Example
CartanForm (car)
This function gives a form of the Cartan decomposition of a numeric real
non-singular square matrix, namely the factorization a = k · ex p(x) where k
is orthogonal and x symmetric. It follows from this that the transpose of an
invertible matrix satisfies an equation
t
a = k.a.k
CartanForm[a] −→ {k, s}
Example
ConstantMatrix (com)
This function constructs a constant matrix with specified element value.
This function can be used together with other functions to construct matrices.
ConstantMatrix[c, m, n] −→ a
CreQ (crq)
This function checks to see if its argument evaluates to a so-called Canonical
Rational Expression (CRE), i.e. a number (real or complex, exact or floating
A.3 Function descriptions and examples 421
CreQ[e] −→ P
e is a Mathematica expression,
P is True if e is a CRE and False otherwise.
Example
DiagonalToMatrix (d2m)
This function takes a list and constructs a matrix with the list elements as the
diagonal entries.
Diagonal matrices appear in many places, including in the Iwasawa and
Bruhat decompositions.
DiagonalToMatrix[di] −→ a
EisensteinFourierCoefficient (efc)
This function returns the nth term of the Fourier expansion of an Eisenstein
series for GL(2), with an explicit integer specified for n.
See Section 3.1, especially Theorem 3.1.8.
EisensteinFourierCoefficient[z, s, n] −→ v
z is a CRE, s is a CRE,
n is an integer being the index of the nth coefficient,
v is a complex number or symbolic expression representing the nth Fourier
term of the Eisenstein Fourier expansion for GL(2) with parameters
z and s.
Example
EisensteinSeriesTerm (est)
This function returns the term of the Eisenstein series E(z, s) for GL(2), namely
the summand of:
1 ys
E(z, s) =
2 a,b∈Z,(a,b)=1 |az + b|2s
EisensteinSeriesTerm[z, s, ab] −→ v
z is a CRE, s is a CRE,
ab is a list of two integers {a, b}, at least one of which must be non-zero,
v is a complex number or symbolic expression representing the term of the
Eisenstein series for GL(2) with parameters z, s, a, b.
A.3 Function descriptions and examples 423
Example
ElementaryMatrix (elm)
This function returns a square matrix having 1s along the leading diagonal and
with a given element in a specified off-diagonal position.
ElementaryMatrix[n, i, j, c] −→ e
Example
FunctionalEquation (feq)
This function, for each index i, returns a list of affine combinations of its
variables representing the ith functional equation for the Jacquet–Whittaker
function of order n ≥ 2.
See Section 5.9, especially equations (5.9.5), (5.9.6) and Example 5.9.7.
FunctionalEquation[v, i] −→ vp
Example
GetCasimirOperator (gco)
This function computes the Casimir operator acting on an arbitrary noun func-
tion and with respect to the Iwasawa variables. Note that this function makes
an explicit brute-force evaluation of the operator, so is not fast, especially for
n ≥ 3.
See Proposition 2.3.3, Example 2.3.4 and Section 6.1.
GetCasimirOperator[m,n,“x”,“y”,“f”] −→ Operator
m is positive integer with value 2 or more being the order of the operator,
n is a positive integer with value 2 or more being the dimension of the Iwasawa
form,
“x” is a string, being the name of the symbol such that the variables in the upper
triangle of the matrix given by the Iwasawa decomposition are x[i, j],
“y” is a string, being the name of the symbol such that the terms in the first
n − 1 positions of the leading diagonal of the Iwasawa decomposition
are
y[1] · · · y[n − 1], y[1] · · · y[n − 2], . . . , y[1],
A.3 Function descriptions and examples 425
“f” is a string being the name of a function of noun form (i.e. it should not be
defined as an explicit Mathematica function or correspond to the name of
an existing function) which will appear as partially differentiated by the
computed Casimir operator, Operator is an expression in the variables
(x[i, j], 1 ≤ i < j ≤ n), (y[i], 1 ≤ i ≤ n − 1)
and the partial derivatives of the function with name “f” with respect to
argument slots of f arranged in the order (x1,1 , x1,2 , . . . , y1 , . . . , yn−1 ).
Example
GetMatrixElement (gme)
The specified element of a matrix is returned.
GetMatrixElement[a, i, j] −→ e
a is a matrix of CREs,
i is the row index of the element,
j is the column index of the element,
e is the (i, j)th element of a.
See also: MatrixColumn, MatrixRow, MatrixBlock.
426 Appendix The GL(n)pack Manual
GlnVersion (glv)
This function prints out the date of the version of GL(n)pack which is being
used, followed by the version of Mathematica. It has no argument, but the
brackets must be given.
GlnVersion[] −→ True
HeckeCoefficientSum (hcs)
This function takes a natural number m, a list of natural numbers
{m 1 , . . . , m n−1 } and a string for a function name and finds the terms in the
sum right-hand side
λm A(m 1 , . . . , m n−1 ) = A(c0 m 1 /c1 , c1 m 2 /c2 , . . . , cn−2 m n−1 /cn−1 )
)n−1
where the summation is over all (ci ) such that i=0 ci = m and ci |m i , 1 ≤ i ≤
n − 1.
See Theorem 9.3.11 and equation (9.3.17).
HeckeEigenvalue (hev)
This function returns the value of the mth eigenvalue of the ring of HeckeOpera-
tors acting on square integrable automorphic forms f (z) for hn . Note that when
the Euler product of a Maass form is known, the Fourier coefficients which
appear in the expressions for the eigenvalues (the A in λm = A(m, 1, . . . , 1))
A.3 Function descriptions and examples 427
HeckeEigenvalue[m, n, a] −→ λm
Example
HeckeMultiplicativeSplit (hms)
This function takes a list of natural numbers {m 1 , . . . , m n−1 }, finds the primes
and their powers that divide any of the m i , and returns a list of lists of those
primes and their powers. The purpose of this function is the evaluation of
the Hecke Fourier coeffients of a Maass form in terms of Schur polynomials
when the Euler product coeffients of the form are known. If p1 , . . . , pr are
the primes and ki, j is the maximum power of pi dividing m j , then the Fourier
coefficient
r
k k
A(m 1 , . . . , m n−1 ) = A pi i,1 , . . . , pi i,n−1 .
i=1
HeckeMultiplicativeSplit[m] −→ list
Example
HeckeOperator (hop)
This function computes the nth order Hecke operator which acts on square
integrable forms on hn .
See Section 9.3, especially formula (9.3.5).
HeckeOperator[n, z, f] −→ Tn (f(z))
n is a natural number being the order of the operator,
z is a square matrix of CREs of size n,
f is a string being the name of a function of a square matrix of size n,
Tn (f(z)) is an expression representing the action of the nth Hecke operator on
the matrix function f(z).
Example
HeckePowerSum (hps)
This function takes a natural number m, a list of natural numbers
{m 1 , . . . , m n−1 } and a string for a function name and finds the powers of any
fixed prime in the sum right-hand side
λm A(m 1 , . . . , m n−1 ) = A(c0 m 1 /c1 , c1 m 2 /c2 , . . . , cn−2 m n−1 /cn−1 )
)n−1
where the summation is over all (ci ) such that i=0 ci = m and ci |m i , 1 ≤ i ≤
n − 1 in case m and each of the m i is a power of a fixed prime. The powers
that appear in the expansion are the same for any prime. The purpose of this
function is to simplify the study of the multiplicative properties of the Fourier
coefficients.
See Theorem 9.3.11 and equation (9.3.17).
a is a non-negative integer, being the power a of any prime p such that pa is the
index of the eigenvalue λpa ,
as is a list of non-negative integers representing the powers of a fixed prime
which appear in the multi-index of a Fourier coefficient,
list consists of a sum of terms B[b1,i , . . . , bn−1,i ] such that the corresponding
term in the Hecke sum would have a value A(pb1,i , . . .).
Example
HermiteFormLower (hfl)
This function computes the lower left Hermite form h of a non-singular integer
matrix a, and a unimodular matrix l such that a = lh. This Hermite form is a
lower triangular integer matrix with strictly positive elements on the diagonal
of increasing size, and such that each element in the column below a diagonal
entry is non-negative and less than the diagonal entry.
See Theorem 3.11.1.
430 Appendix The GL(n)pack Manual
HermiteFormLower[a] −→ {l, h}
Example
HermiteFormUpper (hfu)
This function computes the upper Hermite form h of a non-singular integer
matrix a, and a unimodular matrix l such that a = lh. This Hermite form is an
upper triangular integer matrix with strictly positive elements on the diagonal
of increasing size, and such that each element in the column above a diagonal
entry is non-negative and less than the diagonal entry.
See Theorem 3.11.1.
HermiteFormUpper[a] −→ {l, h}
Example
IFun (ifn)
This function returns the value
n−1
n−1 b νj
Iν (z) = yi i, j
i=1 j=1
Example
InsertMatrixElement (ime)
An element is inserted into a matrix returning a new matrix and leaving the
original unchanged.
InsertMatrixElement[e, i, j, a] −→ b
e is a CRE being the element to be inserted,
i is the row index of the position where the element is to be inserted,
j is the column index of the position where the element is to be inserted,
a is the original matrix of CREs,
b is a new matrix, being equal to a but with e in the (i, j)th position.
See also: DiagonalToMatrix, MatrixJoinHorizontal, MatrixJoinVertical.
IwasawaForm (iwf)
This function computes the Iwasawa form of a non-singular real matrix a.
This consists of the product of an upper triangular unipotent matrix x and a
diagonal matrix y with strictly positive diagonal entries such that, for some
non-singular integer matrix u, real orthogonal matrix o and constant diagonal
matrix δ, a = u.x.y.o.δ. This function returns a single matrix z = x.y.
See Section 1.2.
IwasawaForm[a] −→ z
a is a non-singular square matrix of CREs,
z is an upper-triangular matrix with positive diagonal entries, being the Iwasawa
form of a.
A.3 Function descriptions and examples 433
Example
The Iwasawa decomposition of the matrix
a b
c d
is found.
IwasawaXMatrix (ixm)
This function returns the unipotent matrix x corresponding to the decomposition
z = x.y of a matrix z in Iwasawa form.
See Proposition 1.2.6 and Example 1.2.4
IwasawaXMatrix[w] −→ x
Example
In this example the x-matrix, x-variables, y-matrix and y-variables are extracted
from a generic matrix in Iwasawa form.
434 Appendix The GL(n)pack Manual
IwasawaXVariables (ixv)
This function returns the x-variables from a matrix z = x.y in Iwasawa form.
These are the elements in the strict upper triangle of the matrix x in row order.
See Definition 1.2.3 and Proposition 1.2.6.
IwasawaXVariables[w] −→ l
w is a square non-singular matrix of CREs which must be in Iwasawa form,
l is a list of the form {x1,2 , . . . , x1,n , x2,3 , . . . , xn−1,n }.
See also: IwasawaForm, IwasawaXMatrix, IwasawaYVariables, IwasawaY-
Matrix.
IwasawaYMatrix (iym)
This function returns the y-matrix from the decomposition z = x.y of a matrix
z in Iwasawa form.
A.3 Function descriptions and examples 435
IwasawaYVariables (iyv)
This function returns a list of the y-variables from the Iwasawa decomposition
of a matrix z = x.y.
See Definition 1.2.3 and Proposition 1.2.6.
IwasawaYVariables[z] −→ L
z is a square non-singular matrix of CREs which must be in Iwasawa form,
L a list {y1 , . . . , yn−1 } of the y-variables of the Iwasawa form.
See also: IwasawaForm, IwasawaYMatrix, IwasawaXVariables, IwasawaX-
Matrix.
IwasawaQ (iwq)
This function tests a Mathematica form or expression to see whether it is a
non-singular square matrix in Iwasawa form.
See Section 1.2.
IwasawaQ[z] −→ value
z is a Mathematica form,
value is True if z is a matrix of CREs in Iwasawa form, False otherwise.
See also: IwasawaForm, MakeZMatrix.
KloostermanBruhatCell (kbc)
This function takes an explicit permutation matrix w with all other arguments
being symbolic. It returns rules which solve for x and y in the square matrix
Bruhat decomposition equation a = x.c.w.y assuming c is in “Friedberg form”,
x and y are unipotent and y satisfies t w.t y.w is upper triangular. These rules
are not unique.
See Chapter 11, especially Section 11.2. Also Lemma 10.6.3.
436 Appendix The GL(n)pack Manual
KloostermanBruhatCell[a,x,c,w,y] −→ rules
a is a symbol which will be used as the name of an n × n matrix,
x is a symbol which will be used as the name of a unipotent matrix,
c is a symbol which will be used as the name of an array c[i] representng a list
of n − 1 non-zero integers specifying the diagonal of a matrix. (Note that
the 1st element of the diagonal represents the term det(w)/c[n − 1], the
second c[n − 1]/c[n − 2] and so on down to the last c[1] as in the notation
of (11.2.1).),
w is an n × n matrix which is zero except for a single 1 in each row and column,
being an explicit element of the Weyl Group Wn ,
y is a symbol which will be used as the name of a unipotent matrix which satisfies
t t
w. y.w is upper triangular making the decomposition unique, given a,
rules is a list of rules of the form x[i, j] → eij or y[i, j] → eij where the eij are
expressions in the a[i, j] and c[i].
Example
KloostermanCompatibility (klc)
This function takes an explicit permutation matrix w, with remaining arguments
symbolic, and returns a list of values, each element being a different type
of constraint applicable to any valid Kloosterman sum based on w. The first
element is a list of forms restricting the characters. The second is a set of
divisibility relations restricting the values of the diagonal matrix c. And the third
is the set of minor relations. A typical approach to forming Kloosterman sums
would be to first run this function, determine a valid set or sets of parameters
from the symbolic output, and then run KloostermanSum using explicit integer
values of valid parameters.
A.3 Function descriptions and examples 437
KloostermanSum (kls)
This function computes the generalized Kloosterman sum for SL(n, Z) for
n ≥ 2 as given by Definition 11.2.2. When n = 2 this coincides with the clas-
sical Kloosterman sum. More generally the sum is
S(θ1 , θ2 , c, w) := θ1 (b1 )θ2 (b2 )
γ =b1 cwb2
438 Appendix The GL(n)pack Manual
where
γ ∈ Un (Z)\S L(n, Z) ∩ G w / w
and w = t w.t U n (Z).w ∩ Un (Z) and G w is the Bruhat cell associated to the
permutation matrix w. Since these sums are only well defined for some par-
ticular compatible values of the arguments the user is advised to first run
KloostermanCompatibility with an explicit w to determine those values.
) 2
Note that the complexity of the algorithm is O( 1≤i≤n−1 |ci |n ) = O(cn ) where
c = max|ci |.
See Chapter 11.
Example
This n = 4 example shows how KloostermanCompatibility should be run after
selecting a permutation matrix. Then KloostermanSum is called with compat-
ible arguments.
A.3 Function descriptions and examples 439
Example
This first illustrates commutativity of the LongElement sums (c/f (Friedberg
(1987), its Proposition 2.5)), then Proposition 11.4.1 and finally is given an
example of a classical sum showing it is real.
LanglandsForm (llf)
This function returns a list of the three matrices of the Langlands decomposition
of a square matrix in a parabolic subgroup specified by a partition of the matrix
dimension.
See Section 10.2.
LanglandsForm[p, d] −→ {u, c, m}
Example
LanglandsIFun (lif)
This function computes the summand for Langlands’ Eisenstein series with
respect to a specified parabolic subgroup.
See Chapter 10, Definition 10.4.5.
LanglandsIFun[g, p, s] −→ Is (g.z)
Example
LeadingMatrixBlock (lmb)
This function extracts a leading matrix block of specified dimensions.
LeadingMatrixBlock[a, i, j] −→ b
LongElement (lel)
This function constructs the so-called long element of the group GL(n,Z), a
matrix with 1s along the reverse leading diagonal and 0s elsewhere.
See Chapter 5.
LongElement[n] −→ w
Example
LowerTriangleToMatrix (ltm)
This function takes a list of lists of increasing length and forms a matrix with
zeros in the upper triangle and the given lists constituting the rows of the lower
triangle.
LowerTriangleToMatrix[l] −→ a
Example
A.3 Function descriptions and examples 443
MakeBlockMatrix (mbm)
This function takes a list of lists of matrices and creates a single matrix wherein
the jth matrix element of the ith sublist constitutes the (i, j)th sub-block of this
matrix. In order that this construction succeed, the original matrices must have
compatible numbers of rows and columns, i.e. the matrices in each sublist must
have the same number of rows for that sublist and for each j the jth matrix in
each sublist must have the same number of columns. In spite of this restriction,
the function is a tool for building matrices rapidly when they have a natural
block structure.
MakeBlockMatrix[A] −→ B
A is a list of lists of equal length of matrix elements, each matrix having CRE
elements,
B is a single matrix with sub-blocks being the individual matrices in A.
Examples
MakeMatrix (mkm)
This function returns a symbolic matrix of given dimensions.
444 Appendix The GL(n)pack Manual
MakeMatrix[“a”, m, n] −→ A
“a” is a string being the name of the generic symbolic matrix element variable
a[i, j],
m is a strictly positive integer representing the number of rows of A,
n is a strictly positive integer representing the number of columns of A,
A is a symbolic matrix with (i, j)th entry a[i, j].
Example
MakeXMatrix (mxm)
This function returns a symbolic upper triangular square matrix of given dimen-
sion with 1s on the leading diagonal, i.e. a unipotent matrix.
See Definition 1.2.3.
MakeXMatrix[n, “x”] −→ u
Example
MakeXVariables (mxv)
This function returns a list of the x-variables which appear in the symbolic
generic Iwasawa form of a square matrix of given dimension.
See Definition 1.2.3.
MakeXVariables[n, “x”] −→ l
n is a strictly positive integer representing the size of the matrix,
“x” is a string being the name of the generic list element variable x[i, j],
l is a list of the x-variables in order of increasing row index.
See also: MakeXMatrix, MakeYMatrix, MakeYVariables, MakeZMatrix.
446 Appendix The GL(n)pack Manual
MakeYMatrix (mym)
This function returns a symbolic diagonal matrix of given dimension with val-
ues on the leading diagonal being the product of the y-variables of a matrix
expressed in Iwasawa form.
See Definition 1.2.3 and the manual entry for MakeXMatrix.
MakeYMatrix[n, “y”] −→ d
n is a strictly positive integer representing the size of the matrix,
“y” is a string being the name of the generic symbolic matrix element variable
y[i] such that the jth diagonal element is the product y[1]y[2] · · · y[n − j],
d is a diagonal matrix.
See also: MakeXMatrix, MakeXVariables, MakeYVariables, MakeZMatrix.
MakeYVariables (myv)
This function returns a symbolic list of the n − 1 y-variables which would occur
in the Iwasawa form of a matrix of size n × n.
See Definition 1.2.3 and the manual entry for MakeXMatrix.
MakeYVariables[n, “y”] −→ l
n is a strictly positive integer representing the size of the matrix,
“y” is a string being the name of the generic variable y[i],
l is a list of the form {y[1], . . . , y[n − 1]}.
See also: MakeXMatrix, MakeXVariables, MakeYMatrix, MakeZMatrix,
MakeZVariables.
MakeZMatrix (mzm)
This function returns a symbolic upper triangular square matrix of given dimen-
sion being in generic Iwasawa form.
See Example 1.2.4 and the manual entry for MakeXMatrix.
MakeZMatrix[n, “x”,“y”] −→ u
n is a strictly positive integer representing the size of the matrix,
“x” is a string being the name of the generic symbolic Iwasawa x-variable
x[i, j],
“y” is a string being the name of the generic symbolic Iwasawa y-variable y[i],
u is an upper-triangular symbolic matrix with (i, j)th element having the form
x[i, j]y[1] · · · y[n − j].
A.3 Function descriptions and examples 447
MakeZVariables (mzv)
This function returns a list of the variables which occur in the Iwasawa form
for a matrix with generic symbolic entries and of given size.
See the manual entry for MakeXMatrix.
MakeZVariables[n, “x”, “y”] −→ l
n is a strictly positive integer representing the size of the matrix,
“x” is a string being the name of the generic symbolic matrix element x[i, j]
with i > j,
“y” is a string being the name of the generic symbolic matrix element y[i],
l is a list of the Iwasawa variables with the x-variables first in order of increasing
row index followed by the y-variables:
{x[1, 2], . . . , x[1, n], x[2, 3], . . . , x[n − 1, n], y[1], . . . , y[n − 1]}.
See also: MakeXMatrix, MakeXVariables, MakeYMatrix, MakeYVariables,
MakeZMatrix.
MatrixColumn (mcl)
This function returns a given column of a matrix.
MatrixColumn[m, j] −→ c
m is a matrix of CREs,
j is a valid column index for m,
c is the jth column of m returned as a list.
See also: MatrixRow.
MatrixDiagonal (mdl)
This function extracts the diagonal of a matrix.
MatrixDiagonal[a] −→ d
a is a square matrix of CREs,
d is a list, being the diagonal entries of a in the same order.
See also: DiagonalToMatrix.
448 Appendix The GL(n)pack Manual
MatrixJoinHorizontal (mjh)
This function assembles a new matrix by placing one matrix to the right of
another compatible matrix.
MatrixJoinHorizontal[a, b] −→ c
a is a matrix of CREs,
b is a matrix with the same number of rows as a,
c is a matrix with block decomposition c = [a|b].
Example
MatrixJoinVertical (mjv)
This function assembles a new matrix by placing one matrix above another
compatible matrix.
MatrixJoinVertical[a, b] −→ c
a is a matrix of CREs,
b is a matrix with the same number of columns as a,
c is a matrix with block decomposition having a above b.
See also: MatrixJoinHorizontal.
MatrixLowerTriangle (mlt)
This function extracts the elements in the lower triangle of a square matrix,
including the diagonal, and returns them as a list of lists.
MatrixLowerTriangle[a] −→ t
a is a square matrix of CREs,
t is a list of lists where the ith element of the jth list represents the (j, i)th
element of a.
See also: MatrixUpperTriangle, LowerTriangleToMatrix.
A.3 Function descriptions and examples 449
MatrixRow (mro)
This function returns a given row of a matrix.
MatrixRow[m, i] −→ r
m is a matrix of CREs,
i is a row index of m,
r is a list representing the ith row of m.
MatrixUpperTriangle (mut)
This function extracts the elements in the upper triangle, including the diagonal,
of a square matrix and returns a list of lists of the elements from each row.
MatrixUpperTriangle[a] −→ t
Example
ModularGenerators (mog)
This function returns a list of two matrix generators for the subgroup of the
group of integer matrixes with determinant 1, i.e. generators of SL(n, Z).
See Chapter 5, especially Section 5.9.
450 Appendix The GL(n)pack Manual
ModularGenerators[n] −→ g
Example
MPEisensteinGamma (eig)
This function computes the gamma factors for the minimal parabolic Eisenstein
series
n
s − λi (v)
G Ev (s) = π −ns/2 .
i=1
2
MPEisensteinGamma[s,v] −→ G
Example
MPEisensteinLambdas (eil)
This function computes the functions λi (v) : Cn−1 → C such that the L-
function associated with the minimal parabolic Eisenstein series L Ev (s) is a
product of shifted zeta values
n
L Ev (s) = ζ (s − λi (v)).
i=1
MPEisensteinLambdas[v] −→ L
Example
MPEisensteinSeries (eis)
This function computes the L-function associated with the minimal parabolic
Eisenstein series E v (z) as a product of shifted zeta values
n
L Ev (s) = ζ (s − λi (v)).
i=1
MPEisensteinSeries[s,v] −→ Z
Example
MPExteriorPowerGamma (epg)
This function returns the gamma factors for the kth symmetric L-function asso-
ciated with a minimal parabolic Eisenstein series.
See the introduction to Chapter 13.
MPExteriorPowerGamma[s,v,k] −→ G
Example
MPExteriorPowerLFun (epl)
This function returns the kth exterior power of the L-function of a minimal
parabolic Eisenstein series as a product of zeta values.
See the introduction to Chapter 13.
This function can be used to show that exterior power L-functions satisfy a
functional equation.
A.3 Function descriptions and examples 453
MPExteriorPowerLFun[s,v,k] −→ Z
Example
MPSymmetricPowerLFun (spf)
This function returns the kth symmetric power of the L-function of a minimal
parabolic Eisenstein series as a product of zeta values.
See the introduction to Chapter 13.
This can be used to show that symmetric power L-functions satisfy a func-
tional equation.
MPSymmetricPowerLFun[s,v,k] −→ Z
Example
454 Appendix The GL(n)pack Manual
MPSymmetricPowerGamma (spg)
This function returns the gamma factors for the kth symmetric L-function asso-
ciated with a mimimal parabolic Eisenstein series.
See the introduction to Chapter 13.
MPSymmetricPowerGamma[s,v,k] −→ G
Example
NColumns (ncl)
This function gives the number of columns of a matrix.
NColumns[a] −→ n
a is a matrix of CREs,
n is the number of columns of a.
NRows (nro)
This function gives the number of rows of a matrix.
NRows[a] −→ m
a is a matrix of CREs,
m is the number of rows of a.
ParabolicQ (paq)
This function tests a square matrix to see whether it is in a given parabolic
subgroup as specified by a non-trivial partition of the matrix dimension.
See Chapter 10, especially Section 10.1.
ParabolicQ[a, d] −→ ans
Example
PluckerCoordinates (plc)
This function takes an n × n square matrix and returns a list of lists of the so-
called Plücker coordinates, namely the values of all of the bottom j × j minors
with 1 ≤ j ≤ n − 1.
See Chapter 11, Section 11.3, Theorem 11.3.1.
PluckerCoordinates[a] −→ value
a is an n × n matrix of CREs,
value is a list of lists being the values of all of the j × j minor determinants with
1 ≤ j ≤ n − 1 based on the bottom row and taking elements from the bot-
tom j rows. The jth sublist has the j × j minors in lexical order of the column
indices.
Example
PluckerInverse (pli)
This function takes a list of lists of integers, which could be the Plücker
coordinates arising from a square matrix, and returns such a matrix having
determinant 1. The matrix is not unique but PluckerInverse followed by
A.3 Function descriptions and examples 457
PluckerInverse[Ms] −→ a
Example
PluckerRelations (plr)
This function computes all the known quadratic relationships between the
minors of a generic square n × n matrix known as the Plücker coordinates.
See Chapter 11.
458 Appendix The GL(n)pack Manual
In the case that n = 2 there are none and for n = 3 one. For n > 3 the
number grows dramatically. No claim is made that this function returns, for
any given n, a complete set of independent relationships. By “complete”
is meant sufficient to guarantee the coordinates arise from a member of
S L(n, Z).
PluckerRelations[n, v] −→ relations
Example
RamanujanSum (rsm)
This function computes the Ramanujan sum s(n, c) for explicit natural number
values of n, c, namely
c
s(n, c) = e2πi(r/c) .
r =1,(r,c)=1
RemoveMatrixColumn (rmc)
A given row is removed from a matrix, creating a new matrix and leaving the
original unchanged.
RemoveMatrixColumn[a, j] −→ b
a is a matrix of CREs,
j is a valid column index of a,
b is a matrix with all columns identical to a except the jth which is missing.
See also: RemoveMatrixRow.
RemoveMatrixRow (rmr)
A given row is removed from a matrix, leaving the original unchanged.
RemoveMatrixRow[a, i] −→ b
a is a matrix of CREs,
i is a valid row index of a,
b is a matrix with all rows identical to a except the ith which is missing.
See also: RemoveMatrixColumn.
460 Appendix The GL(n)pack Manual
SchurPolynomial (spl)
This function computes the Schur polynomial in n variables x1 , . . . , xn with
n − 1 exponents k1 , . . . , kn−1 , that is to say the ratio of the determinant of a
k +···+ki−1 +n−i
matrix with (i, j)th element 1 for i = n and x j 1 for 1 ≤ i ≤ n − 1,
to the determinant of the matrix which is 1 for i = n and x j for 1 ≤ i ≤ n − 1.
n−i
SmithElementaryDivisors (sed)
This function computes the elementary divisors of a non-singular n × n integer
matrix a, i.e. for each j with 1 ≤ j ≤ n, the gcd d j (a) of all of the j × j
minor determinants. If s j is the jth diagonal entry of the Smith form then
s j = d j (a)/d j−1 (a).
SmithElementaryDivisors[a] −→ l
a is a non-singular n × n integer matrix,
l is a list of the n Smith form elementary divisors of a in the order
{d1 (a), . . . , dn (a)}.
See also: SmithForm, SmithInvariantFactors, HermiteFormUpper, Hermite-
FormLower.
A.3 Function descriptions and examples 461
SmithForm (smf)
This function returns the Smith form diagonal matrix d of a square non-singular
matrix a with integer entries. This matrix d satisfies 0 < di,i and di,i | di+1,i+1
for all i ≤ n − 1. It also returns unimodular matrixes l, r such that a = l.d.r .
See Theorem 3.11.2.
SmithForm[a] −→ {l, d, r}
a is non-singular integer matrix,
l is a unimodular matrix,
d is a diagonal matrix, being the Smith Form of a, r is a unimodular matrix.
Example
The Smith form of a 4 by 4 matrix is computed and the result checked.
SmithInvariantFactors (sif)
This function computes the invariant factors of the Smith form of a non-singular
integer matrix a. These are all of the prime powers which appear in the diagonal
entries of the Smith form of a.
SmithInvariantFactors[a] −→ l
a is an n × n non-singular integer matrix,
l is a list of prime powers.
See also: SmithForm, SmithElementaryDivisors, HermiteFormUpper,
HermiteFormLower.
462 Appendix The GL(n)pack Manual
SpecialWeylGroup (swg)
This function, for each natural number n, returns the group of n × n matrices
with each entry being 0 or ±1, and having determinant 1. There are 2n−1 n! such
matrices.
See Sections 6.3 and 6.5.
SpecialWeylGroup[n] −→ g
Example
SubscriptedForm (suf)
This function takes a Mathematica expression and prints it out in such a way that
subexpressions of the form x[n1 , n2 , . . . , nj ], where the ni are explicit integers,
are printed in the style
xn1 ,n2 ,...,nj .
The value of this function is for improving the look of expressions for inspec-
tion and should not be used otherwise. Compare the Mathematica function
MatrixForm. Not all expressions can be subscripted using this function.
SubscriptedForm[e] −→ f
e is a Mathematica expression,
f is a subscripted rendition of the same expression.
Example
SwapMatrixColumns (smc)
Two columns of a matrix are exchanged creating a new matrix and leaving the
original unchanged.
464 Appendix The GL(n)pack Manual
SwapMatrixColumns[a, i, j] −→ b
a is a matrix of CREs,
i is a valid column index for a,
j is a valid column index for a,
b is a matrix equal to a except the ith and jth columns have been exchanged.
See also: SwapMatrixRows, ElementaryMatrix.
SwapMatrixRows (smr)
Two rows of a matrix are exchanged creating a new matrix and leaving the
original unchanged.
SwapMatrixRows[a, i, j] −→ b
a is a matrix of CREs,
i is a valid row index for a,
j is a valid row index for a,
b is a matrix equal to a except the ith and jth rows have been exchanged.
See also: SwapMatrixColumns, ElementaryMatrix.
TailingMatrixBlock (tmb)
This function returns a tailing matrix block of specified dimensions leaving the
original matrix unchanged.
TailingMatrixBlock[a, i, j] −→ b
a is a matrix of CREs,
i is a positive integer less than the number of rows of a,
j is a positive integer less than the number of columns of a,
b is the tailing block of a with i rows and j columns.
See also: LeadingMatrixBlock, BlockMatrix.
UpperTriangleToMatrix (utm)
This function takes a list of lists of strictly decreasing length and forms a matrix
with zeros in the lower triangle and with the given lists constituting the rows of
the upper triangle. The length of the matrix is the length of the first sublist. The
last sublist has length 1 and each successive sublist has length one less than the
preceding sublist.
UpperTriangleToMatrix[u] −→ a
u is a list of lists of CREs of decreasing length representing the elements of an
upper triangular submatrix including the diagonal,
A.3 Function descriptions and examples 465
VolumeBall (vbl)
This function computes the volume of an n-dimensional ball with given radius.
VolumeBall[r, n] −→ Vol
Example
VolumeFormDiagonal (vfd)
This function computes the differential volume form for the set of diagonal
matrices
A
n
dai ,
i=1
VolumeFormDiagonal[“a”, n] −→ Form
VolumeFormGln (vfg)
This function computes the differential volume form for the matrix group
G L(n, R) using the wedge product.
See Sections 1.4 and 1.5 and Proposition 1.4.3.
VolumeFormGln[“g”, n] −→ Form
“g” is a string which will be the name of a two-dimensional array symbol,
n is a positive integer representing the dimension of the matrices,
Form is the diagonal volume form based on the variables g[i, j].
Example
VolumeFormHn (vfh)
This function computes the differential volume form for the generalized upper
half-plane.
See Definition 1.2.3 and Proposition 1.5.3.
VolumeFormHn[“x”, “y”, n] −→ Form
“x” is a string which will be the name of a two-dimensional array symbol,
“y” is a string which will be used as the name of a one-dimensional array
symbol,
n is a positive integer representing the dimension of the matrices which appear
in the Iwasawa decomposition,
Form is the volume form based on the variables x[i, j], y[j].
Example
VolumeFormUnimodular (vfu)
This function computes the differential volume form for the group of unimod-
ular matrices, i.e. real upper-triangular with 1s along the leading diagonal.
See Sections 1.4 and 1.5.
A.3 Function descriptions and examples 467
VolumeFormHn[“x”, n] −→ Form
“x” is a string being the name of an array symbol,
n is a positive integer representing the dimension of the matrices,
Form is the volume form based on the variables x[i, j].
Example
VolumeSphere (vsp)
This function computes the n-dimensional volume of the sphere S n in Rn+1 .
VolumeSphere[r, n] −→ Vol
r is a CRE being the radius of the sphere,
n is the dimension of the sphere,
Vol is the volume of the sphere computed using n-dimensional Lebesgue
measure.
See also: VolumeBall, VolumeHn.
Wedge,d
This function computes the Wedge product of any finite number of functions or
differential forms in an arbitrary number of dimensions. It works with the dif-
ferential form operator d. Note that these functions have a different construction
from others in GL(n)pack , and have limited error control. An alternative to the
function Wedge is the infix operator which may be entered into Mathematica
468 Appendix The GL(n)pack Manual
by typing a backslash, and open square bracket, the word “Wedge” and then a
closing square bracket. It prints like circumflex, but is not the same. Note that
wedge products of vectors are not currently supported.
See Sections 1.4, 1.5, 5.6, 5.7 and 5.8.
Wedge[f1 , f2 , . . . , fn ] −→ value
fi is an expression or a form,
value is the wedge product of the functions or forms fi .
Example
In this example the function Wedge is used in conjunction with the differential
form generator function d. Note that symbols, such as a, can be declared to be
constant explicitly by setting, d[a] = 0.
WeylGenerator (wge)
This function returns a set of matrix generators for the Weyl subgroup of the
group of integer matrices with determiant ±1, which consists of all matrices
with exactly one ±1 in each row and column. A single call returns a single
generator.
A.3 Function descriptions and examples 469
See Chapter 6.
Also see the manual entry for SpecialWeylGroup.
WeylGenerator[n,i,j] −→ g
Example
WeylGroup (wgr)
This function returns, for each whole number n, a list of all of the Weyl group
of n by n permutation matrices.
See the proof of Proposition 1.5.3.
WeylGroup[n] −→ {m1 , m2 , . . . , mk }
Example
470 Appendix The GL(n)pack Manual
Whittaker (wit)
This function computes a symbolic interated integral representatin of the gen-
eralized Jacquet Whittaker function WJacquet (also written W J ) of order n,
for n ≥ 2, as defined by Equation (5.5.1). See Proposition 3.4.6, Section 3.4,
Equation (5.5.1) and Equation (5.5.5). The algorithm uses the recursive rep-
resentation of the Whittaker function defined by Stade (1990, Theorem 2.1)
related to that used in the book as follows. Let W S and W S∗ be Stade’s Whittaker
and Whittaker starred functions respectively and let represent the gamma
factors for either form. Then
Q · W S∗ = · W J = W J∗ = W S
where
n−1
−µ j
Q = Iν (y) yj ,
j=1
n− j
µj = r j,k ,
k=1
k+ j−1
nνi j
r j,k = − .
i=k
2 2
Example
WhittakerGamma (wig)
This function returns the gamma factors for the generalized Jacquet Whittaker
function. See Definition 5.9.2. Note that although this definition differs from
that in (Stade, 1990), the gamma factor that it represents is the same.
WhittakerGamma[v] −→ value
v is a list of n − 1 CREs which, if any are numerical, satisfy vi > 1/n,
value is an expression, being the product of the gamma factors for the Whittaker
function of order n.
See also: Whittaker.
472 Appendix The GL(n)pack Manual
WMatrix (wmx)
This function returns the so-called w-matrix, with (−1)!n/2" in the (1, n)th
position and 1 in every other reversed diagonal position, a member of SL(n, Z).
See Section 5.5.
WMatrix[n] −→ w
n is a strictly positive integer with n ≥ 2,
w is an n by n matrix with each element 0, except the (1, n)th which is (−1)!n/2"
and every (i, n − i + 1)th which is 1 for 2 ≤ i ≤ n.
Example
ZeroMatrix (zmx)
This function returns a zero matrix of given dimensions.
ZeroMatrix[m,n] −→ Z
m is a strictly positive integer representing the number of matrix rows,
n is a strictly positive integer representing the number of matrix columns,
Z is a zero matrix with m rows and n columns.
See also: ConstantMatrix.
References
Ahlfors L. (1966), Complex Analysis, Int. Series in Pure and Applied Math. New York,
McGraw–Hill Book Company, pp. 133–7.
Arthur J. (1979), Eisenstein series and the trace formula, in Automorphic forms, rep-
resentations and L-functions (Proc. Sympos. Pure Math., Oregon State University,
Corvallis, OR, 1977), Proc. Sympos. Pure Math., Vol. XXXIII, Providence, RI,
Amer. Math. Soc., Part 1, pp. 253–74.
(1989), Unipotent automorphic representations: conjectures, Orbites unipotentes et
représentations, II, Astérisque No. 171–2, 13–71.
(2002), A note on the automorphic Langlands group, Dedicated to Robert V. Moody,
Canad. Math. Bull. 45 (4), 466–82.
(2003), The principle of functoriality. Mathematical challenges of the 21st century,
(Los Angeles, CA, 2000). Bull. Amer. Math. Soc. (N.S.) 40 (1), 39–53.
Artin E. (1923), Über eine neue Art von L-Reihen, Hamb. Math. Abh. 3, 89–108.
(Collected papers, Edited by Serge Lang and John T. Tate. Reprint of the 1965
original. New York–Berlin, Springer-Verlag, (1982), pp. 105–24).
(1927), Beweis des allgemeinen Reziprozitätsgesetzes, Hamb. Math. Abh., 353–63.
(Collected papers, Edited by Serge Lang and John T. Tate. Reprint of the 1965
original. New York–Berlin, Springer-Verlag, (1982), pp. 131–41).
(1930), Zur Theorie der L-Reihen mit allgemeinen Gruppencharakteren, Hamb. Math.
Abh. 8, 292–306. (Collected papers, Edited by Serge Lang and John T. Tate. Reprint
of the 1965 original. New York–Berlin, Springer-Verlag, (1982) pp. 165–79).
Arveson W. (2002), A short course on spectral theory, Graduate Texts in Mathematics
209, New York, Springer-Verlag.
Atkin A. O. and Li W. C. (1978), Twists of newforms and pseudo-eigenvalues of W -
operators, Invent. Math. 48 (3), 221–43.
Aupetit B. (1991), A primer on spectral theory, Universitext. New York, Springer-Verlag.
Baker Alan. (1971), Imaginary quadratic fields with class number “2”, Ann. Math. (2)
94, 139–52.
Baker Andrew. (2002) Matrix groups, An introduction to Lie group theory, Springer
Undergraduate Mathematics Series, London, Springer-Verlag.
Banks W. (1997), Twisted symmetric-square L-functions and the nonexistence of Siegel
zeros on GL(3), Duke Math. J. 87 (2), 343–53.
473
474 References
Conrey J. B. and Ghosh A. (1993), On the Selberg class of Dirichlet series: small degrees,
Duke Math. J. 72, 673–93.
Conway J. B. (1973), Functions of One Complex Variable, Graduate Texts in Mathe-
matics, Springer-Verlag.
Curtis M. L. (1984), Matrix groups, Second edition, Universitext, New York, Springer-
Verlag.
Dabrowski R. (1993), Kloosterman sums for Chevalley groups, Trans. Amer. Math. Soc.
337 (2), 757–69.
Dabrowski R. and Fisher B. (1997), A stationary phase formula for exponential sums
over Z/ p m Z and applications to G L(3)-Kloosterman sums, Acta Arith. 80 (1),
1–48.
Davenport H. (1974), Multiplicative Number Theory, Markham Publishing (1967), Sec-
ond Edition (Revised by H. Montgomery) Graduate Texts in Mathematics, Springer-
Verlag.
Deligne P. (1974), La conjecture de Weil. I, (French) Inst. Hautes Études Sci. Publ. Math.
43, 273–307.
(1977), Application de la formule des traces aux sommes trigonométriques, in SGA4 12 ,
Springer Lecture Notes 569, 168–232.
Deuring M. (1933), Imaginäre quadratische Zahlkörper mit der Klassenzahl (1), Math.
Zeit. 37, 405–15.
Drinfeld V. G. (1989), Cohomology of compactified moduli varieties of F-sheaves of
rank 2, (Russian) Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI)
162 (1987), Avtomorfn. Funkts. i Teor. Chisel. III, 107–158, 189; translation in J.
Soviet Math. 46 (2), 1789–821.
Duke W. and Iwaniec H. (1989), Estimates for coefficients of L-functions, I, Automorphic
Forms and Analytic Number Theory (Montreal, PQ, 1989), CRM, pp. 43–47.
Edelen D. G. B. (1985), Applied exterior calculus, A Wiley-Interscience Publication,
New York, John Wiley.
Frey G. (editor) (1994), On Artin’s Conjecture for Odd 2-dimensional Representations,
Berlin, Springer-Verlag.
Friedberg S. (1987a), Poincaré series, Kloosterman sums, trace formulas, and automor-
phic forms for G L(n), Séminaire de Théorie des Nombres, Paris 1985–86, Progr.
Math. 71, Boston, MA, Birkhäuser, 53–66.
(1987b), A global approach to the Rankin–Selberg convolution for G L(3,Z), Trans.
Amer. Math. Soc. 300, 159–78.
(1987c), Poincaré series for G L(n): Fourier expansion, Kloosterman sums, and
algebreo-geometric estimates, Math. Zeit. 196 (2), 165–88.
Friedberg S. and Goldfeld D. (1993), Mellin transforms of Whittaker functions, Bull.
Soc. Math. France, 121, 91–107.
Garrett P. (2002), Volume of S L(n,Z)\S L(n,R) and Spn (Z)\Spn (R), preprint.
Gauss C. F. (1801), Disquisitiones Arithmeticae, Göttingen; English translation by
A. Clarke, revised by W. Waterhouse, New Haven, Yale University Press, 1966;
reprinted by Springer-Verlag, 1986.
Gelbart S. (1984), An elementary introduction to the Langlands program, Bull. Amer.
Math. Soc. (N.S.) 10 (2), 177–219.
Gelbart S. and Jacquet H. (1978), A relation between automorphic representations of
G L(2) and G L(3), Ann. Sci. École Norm. Sup. 4e série 11, 471–552.
References 477
Gelbart S., Lapid E. and Sarnak P. (2004), A new method for lower bounds of L-functions,
(English. English, French summary) C. R. Math. Acad. Sci. Paris 339 (2), 91–4.
Gelbart S. and Shahidi F. (1988), Analytic properties of automorphic L-functions, Per-
spectives in Mathematics 6, Boston, MA, Academic Press.
Gelfand I. M., Graev M. I. and Pyatetskii-Shapiro I. I. (1990), Representation theory
and automorphic functions, Translated from the Russian by K. A. Hirsch. Reprint
of the 1969 edition. Generalized Functions 6, Boston, MA, Academic Press.
Godement R. (1966), The spectral decomposition of cusp-forms, Algebraic Groups
and Discontinuous Subgroups (Proc. Sympos. Pure Math., Boulder, CO, 1965),
pp. 225–34 Amer. Math. Soc., Providence, RI.
Godement R. and Jacquet H. (1972), Zeta functions of simple algebras, Lecture Notes
in Mathematics, Vol. 260. Berlin–New York, Springer-Verlag.
Goldfeld D. (1974), A simple proof of Siegel’s theorem, Proc. Nat. Acad. Sci. USA. 71,
1055.
(1976), The class number of quadratic fields and the conjecture of Birch and
Swinnerton-Dyer, Ann. Scuola Norm. Sup. Pisa (4) 3, 624–63.
(1985), Gauss’ class number problem for imaginary quadratic fields, Bull. Amer.
Math. Soc. 13, 23–37.
(1987), Kloosterman zeta functions for G L(n,Z ), Proceedings of the International
Congress of Mathematicians, Vol. 1, 2 (Berkeley, CA, 1986), Providence, RI, Amer.
Math. Soc., pp. 417–24.
Goldfeld D., Hoffstein J., Lieman D. (1994), An effective zero free region, in the
Appendix of Coefficients of Maass forms and the Siegel zero, Ann. of Math. 140,
177–81.
(2004), The Gauss class number problem for imaginary quadratic fields, in Heegner
Points and Rankin L-series, MSRI Publications 49, 25–36.
Goldfeld D. and Satnak P. (1983), Sums of Kloosterman sums, Invent. Math. 71 (2),
243–50.
Gross B. and Zagier D. B. (1986), Heegner points and derivatives of L-series, Invent.
Math. 84, 225–320.
Halmos P. (1974), Measure Theory, Graduate Texts in Mathematics 18, Springer-Verlag,
250–62.
Hamburger H. (1921), Über die Riemannsche Funktionalgleichung der ζ –Funktion,
Math. Zeit. 10, 240–54.
Harcos G. (2002), Uniform approximate functional equations for principal L-functions,
IMRN 18, 923–32.
Harish-Chandra (1959), Automorphic forms on a semi-simple Lie group, Proc. Nat.
Acad. Sci. U.S.A. 45, 570–3.
(1966), Discrete series for semisimple Lie groups, II. Explicit determination of the
characters, Acta Math. 116, 1–111.
(1968), Automorphic forms on semisimple Lie groups, Notes by J. G. M. Mars. Lecture
Notes in Mathematics, No. 62, Berlin–New York, Springer-Verlag.
Harris M. and Taylor R. (2001), The geometry and cohomology of some simple Shimura
varieties, With an appendix by Vladimir G. Berkovich, Ann. Math. 151, Princeton,
NJ, Princeton University Press.
Hecke E. (1920), Eine neue Art von Zetafunktionen und ihre Beziehungen zur Verteilung
der Primzahlen, Zweite Mitteilung, Math. Zeit. 6, 11–51.
478 References
Langlands R. P. (1966), Eisenstein series, and Volume of the fundamental domain for
some arithmetical subgroups of Chevalley groups, in Proc. Symp. Pure Math.,
Vol. IX, Algebraic Groups and Discontinuous Subgroups, Amer. Math. Soc.,
pp. 235–57.
(1970), Problems in the theory of automorphic forms, in Lecture Notes in Mathematics
170, Springer-Verlag.
(1971), Euler Products, Yale University Press.
(1976), On the functional equations satisfied by Eisenstein series, Lecture Notes in
Mathematics 544, Springer-Verlag.
(1980), Base Change for G L(2), Princeton, NJ, Princeton University Press.
Larsen M. (1988), Estimation of S L(3,Z) Kloosterman sums, in the Appendix
to Poincaré series and Kloosterman sums for S L(3,Z), Acta Arith. 50,
86–9.
Laumon G., Rapoport M. and Stuhler U. (1993), D-elliptic sheaves and the Langlands
correspondence, Invent. Math. 113 (2), 217–338.
Lavrik A. F. (1966), Funktional equations of Dirichlet L-functions, Soviet Math. Dokl.
7, 1471–3.
Lax P. D. and Phillips R. S. (1976), Scattering Theory for Automorphic Functions, Ann.
Math., No. 87. Princeton, NJ, Princeton University Press, pp. 7–11.
Lenstra H. W. and Stevenhagen P. (2000), Artin reciprocity and Mersenne primes, Nieuw
Arch. Wisk. (5) 1, 44–54.
Li W. C. (1975), Newforms and functional equations, Ann. Math. 212, 285–315.
(1979), L-series of Rankin type and their functional equations, Ann. Math. 244 (2),
135–66.
Lieman D. B. (1993), The GL(3) Rankin–Selberg convolution for functions not of rapid
decay, Duke Math. J. 69 (1), 219–42.
Lindenstrauss E. and Venkatesh A. (to appear), Existence and Weyl’s law for spherical
cusp forms.
Luo W., Rudnick Z. and Sarnak P. (1995), On Selberg’s eigenvalue conjecture, Geom.
Funct. Anal. 5 (2), 387–401.
(1999), On the generalized Ramanujan conjecture for G L(n), Automorphic forms,
automorphic representations, and arithmetic (Fort Worth, TX, 1996), Proc. Sympos.
Pure Math. 66, Part 2, Amer. Math. Soc., Providence, RI, 301–10.
Maass H. (1949), Über eine neue Art von nichtanalytischen automorphen Funktionen
und die Bestimmung Dirichletscher Reihen durch Funktionalgleichungen, Ann.
Math. 122, 141–83.
(1964), Lectures on Modular Functions of One Complex Variable, Tata Inst. Fund.
Research, Bombay.
Macdonald I. G. (1979), Symmetric Functions and Hall Polynomials, Oxford, Clarendon
Press, Ch. 1, Sect. 4, pp. 32–4.
Manin J. I. and Pančiškin A. A. (1977), Convolutions of Hecke series, and their values
at lattice points, (Russian) Mat. Sb. (N.S.) 104(146) (4), 617–51.
Miller S. D. (2001), On the existence and temperedness of cusp forms for S L 3 (Z),
J. Reine Angew. Math. 533, 127–69.
Miller S. D. and Schmid W. (2004), Summation formulas, from Poisson and Voronoi
to the present, in Noncommutative Harmonic Analysis, Progr. Math. 220, Boston,
MA, Birkhäuser, 419–40.
References 481
Roquette P. (2000), On the history of Artin’s L-functions and conductors. Seven letters
from Artin to Hasse in the year 1930, Mitteilungen der Mathematischen Gesellschaft
in Hamburg, Band 19, pp. 5–50.
Sarnak P. (1990), Some Applications of Modular Forms, Cambridge University Press
99.
(2004), Nonvanishing of L-functions on (s) = 1, Contributions to Automorphic
forms, Geometry, and Number Theory, Baltimore, MD, Johns Hopkins University
Press, pp. 719–32.
(to appear), The generalized Ramanujan conjectures.
Satake I., Spherical functions and Ramanujan conjecture, (1966) Algebraic Groups and
Discontinuous Subgroups (Proc. Sympos. Pure Math., Boulder, CO, 1965), Amer.
Math. Soc., Providence, RI, 258–64.
Selberg A. (1940), Bemerkungen über eine Dirichlesche reihe, die mit der theorie der
modulformer nahe verbunden ist, Arch. Math. Naturvid. 43, 47–50.
(1956), Harmonic analysis and discontinuous groups in weakly symmetric Rieman-
nian spaces with applications to Dirichlet series, J. Indian Math. Soc. 20, 47–87.
(1960), On discontinuous groups in higher dimensional symmetric spaces, in Contri-
butions to Function Theory, Tata Institute, Bombay, 147–64.
(1963), Discontinuous groups and harmonic analysis, Proc. Int. Congr. Mathemati-
cians (Stockholm, 1962), pp. 177–89.
(1965), On the estimation of Fourier coefficients of modular forms, Proc. Sympos.
Pure Math., Vol. VIII, Amer. Math. Soc., Providence, RI, 1–15.
(1991), Old and new conjectures and results about a class of Dirichlet series, Collected
Papers (Vol. II), Springer-Verlag, 47–63.
Serre J. P. (1977), Modular forms of weight one and Galois representations, Algebraic
number fields: L-functions and Galois properties (Proc. Symp. Durham University,
1975) Academic Press, 193–268.
(1981), Letter to J. M. Deshouillers.
Shahidi F. (1981), On certain L-functions, Amer. J. Math. 103, 297–355.
(1985), Local coefficients as Artin factors for real groups, Duke Math. J. 52 (4),
973–1007.
(1988), On the Ramanujan conjecture and finiteness of poles for certain L-functions,
Ann. Math. 127, 547–84.
(1989), Third symmetric power L-functions for G L(2), Comp. Math. 70, 245–73.
(1990a), On multiplicativity of local factors, in Festschrift in honor of I. I. Piatetski-
Shapiro, Part II, Israel Math. Conf. Proc. 3, Jerusalem, Weizmann, 279–89.
(1990b), A proof of Langlands conjecture on Plancherel measures; complementary
series for p-adic groups, Ann. Math. 132, 273–330.
(1992), Twisted endoscopy and reducibility of induced representations for p-adic
groups, Duke Math. J. (1) 66, 1–41.
Shalika J. A. (1973), On the multiplicity of the spectrum of the space of cusp forms of
G L n , Bull. Amer. Math. Soc. 79, 454–61.
(1974), The multiplicity one theorem for GL(n), Ann. Math. 100, 171–93.
Shimura G. (1971), Arithmetic Theory of Automorphic Functions, Princeton University
Press.
(1973), On modular forms of half-integral weight, Ann. Math. 97, 440–81.
References 483
(1975), On the holomorphy of certain Dirichlet series, Proc. London Math. Soc. 31
(3), 79–98.
Shintani T. (1976), On an explicit formula for class-1 “Whittaker Functions” on G L(n)
over p-adic fields, Proc. Japan Acad. 52.
Siegel C. L. (1935), Über die Classenzahl quadratischer Zahlkörper, Acta Arith. I, 83–6.
(1936), The volume of the fundamental domain for some infinite groups, Trans. Amer.
Math. Soc. 30, 209–18.
(1939), Einheiten quadratischer Formen, Hamb. Abh. Math. 13, 209–39.
(1980), Advanced Analytic Number Theory, Tata Institute Fund. Research, Bombay.
Soundararajan K. (2004), Strong multiplicity one for the Selberg class, Canad. Math.
Bull. 47 (3), 468–74.
Stade E. (1990), On explicit integral formulas for G L(n,R)-Whittaker functions, Duke
Math. J. 60 (2), 313–362.
(2001), Mellin transforms of GL(n,R) Whittaker functions, Amer. J. Math. 123 (1),
121–61.
(2002), Archimedean L-factors on G L(n) × G L(n) and generalized Barnes integrals,
Israel J. Math. 127, 201–19.
Stark H. (1967), A complete determination of the complex quadratic fields of class
number one, Mich. Math. J. 14, 1–27.
(1972), A transcendence theorem for class number problems I, II, Ann. Math. (2) 94
(1971), 153–73 and 96, 174–209.
Stevens G. (1987), Poincaré series on G L(r ) and Kloostermann sums, Ann. Math. 277
(1), 25–51.
Tate J. (1968), Fourier analysis in number fields and Hecke’s zeta function, Thesis,
Princeton (1950), also appears in Algebraic Number Theory, edited by J. W. Cassels
and A, Frohlich, New York, Academic Press, pp. 305–47.
Tatuzawa T. (1951), On a theorem of Siegel, Japan J. Math. 21, 163–78.
Taylor R. and Wiles A. (1995), Ring-theoretic properties of certain Hecke algebras,
Ann. Math. (2) 141 (3), 553–72.
Terras A. (1985), Harmonic Analysis on Symmetric Spaces and Applications I, Berlin,
Springer–Verlag.
(1988), Harmonic Analysis on Symmetric Spaces and Applications II, Berlin,
Springer–Verlag.
Titchmarsh E. C. (1986), The Theory of the Riemann Zeta Function, Second Edition
revised by D. R. Heath-Brown, Oxford Science Publications, Oxford University
Press.
Tunnell J. (1981), Artin’s conjecture for representations of octahedral type, Bull. Amer.
Math. Soc. (N.S.) 5 (2), 173–75.
Venkov A. B. (1983), Spectral theory of automorphic functions, A translation of Trudy
Mat. Inst. Steklov. 153 (1981), Proc. Steklov Inst. Math. 1982, no. 4(153).
Vinogradov I. and Takhtadzhyan L. (1982), Theory of Eisenstein series for the group
S L(3,R) and its application to a binary problem, J. Soviet Math. 18, 293–324.
Wallach N. (1988), Real Reductive Groups, I, Academic Press.
Watkins M. (2004), Class numbers of imaginary quadratic fields, Math. Comp. 73 (246),
907–38.
Weil A. (1946), Sur quelques résultats de Siegel, Brasiliensis Math. 1, 21–39.
484 References
(1949), Numbers of solutions of equations in finite fields, Bull. Amer. Math. Soc. 55,
497–508.
(1967), Über die Bestimmung Dirichletscher Reihen durch Funktionalgleichungen,
Ann. Math. 168, 149–56.
Weyl H. (1939), Classical Groups, their Invariants and their Representations, Princeton
University Press, Ch. 7, sect. 6, 202–3.
Whittaker E. T. and Watson G. N. (1935), A Course in Modern Analysis, Cambridge
University Press.
Ye Y. (2000), A Kuznetsov formula for Kloosterman sums on G L n , Ramanujan J. 4 (4),
385–95.
(1998), Exponential sums for G L(n) and their applications to base change, J. Number
Theory 68 (1), 112–30.
Zagier D. (1982), The Rankin–Selberg method for automorphic functions which are not
of rapid decay, J. Fac. Sci. University of Tokyo Sect. IA Math. 28, 415–37.
(unpublished), Unpublished notes on Kuznetsov’s formula on S L(2,Z).
Index
abelian L-function, 401, 402 Bessel function, 65, 67, 72, 134, 411, 470
Abel transform, 101 bi-invariant function, 355
acts continuously, 3, 6, 74, 163, 266 BlockMatrix, 417
additive character, 63, 64, 66, 117 Bohr–Mollerup theorem, 155
additive twist, 206 Bombieri’s theorem, 354, 393
adjoint, 166, 168, 269, 270, 277 bracket, 40, 45, 47
algebra, 39, 40, 42, 43, 44, 47, 50, 114, 139 BruhatCVector, 418
algebraic integer, 60 Bruhat decomposition, 292–294, 303, 305,
algebraic variety, 351, 382 320, 339–341, 350, 353–355, 417, 421,
antiautomorphism, 76, 77, 80, 164, 167, 266 435
ApplyCasimirOperator, 416 BruhatForm, 418
approximate functional equation, 241–245, Bump’s double Dirichlet series, 186–193
388–392
arithmetic group, 235, 285 CartanForm, 419
Arthur–Selberg trace formula, 383 Casimir operator, 47–50, 416
Artin’s conjecture, 236, 401, 402 Cauchy’s determinant, 231, 232, 234
Artin L-function, 396–401, 404, 405 Cauchy’s identity, 229, 231, 232, 234, 367, 379
Artin reciprocity law 406 Cauchy–Schwartz type inequality, 137
associate parabolic center, 9, 11, 25, 44, 46, 289
associative algebra, 39, 40, 43, 45 center of the universal enveloping algebra,
automorphic, 70, 88, 96, 119, 120, 161, 181, 46–50, 114
188, 195, 196, 201, 216, 217, 218, 222, central character, 162
319, 375, 380, 405 character, 115, 116, 128–130, 144, 145, 152,
automorphic condition, 161, 216, 262 155, 195, 207, 216, 217, 218, 221, 250,
automorphic form, 54, 76, 81, 89, 114, 161, 307, 317, 337–342, 350–353, 360, 401,
166, 216, 218, 222, 223, 253, 257, 263, 437, 438, 470
268, 328, 365, 370, 383, 395, 396, character sum, 207
403–406, 426 characteristic zero, 397
automorphic kernel, 355 Chevalley group, 27, 114, 337
automorphic relation, 222 class field theory, 401
automorphic cuspidal representation, 403 class number, 245, 248, 249, 255
automorphic representation, 365, 396, class number one, 246
403–406 commensurator, 74, 164, 266, 267
commutativity of the Hecke ring, 76
base change, 406 compatibility condition, 342, 350
basis, 16, 38–41, 135–137, 195, 324, 359, 406 composition of differential operators, 52
485
486 Index
constant term, 286, 321, 322, 323, 334 flag, 291, 292
convergence, 224, 300 Fourier transform, 1, 2, 30, 59, 102, 104,
for S L(2, Z), 112, 225, 313, 323 106–107, 113, 309
for S L(3, Z), 112, 321 Fourier expansion, 54, 56, 58–70, 95, 97, 114,
for S L(n, Z), 302, 318, 319, 396, 403 118, 119, 158, 173, 200, 216, 219, 225,
functional equation, 97, 229, 307, 310, 329, 228, 278, 339, 352, 369, 373
370 of an Eisenstein series, 54, 58–62, 95, 225,
Fourier expansion, 56, 58, 59, 60, 62, 225, 228, 286, 303, 310, 319, 323, 327, 332,
286, 303, 339, 422 418, 422
general parabolic, 295 of a Maass form, 63, 64, 69, 70, 114, 128,
Langlands, 285, 286, 301, 316, 318–320, 129, 159, 214, 260, 261, 282, 302, 315,
357, 403, 417, 441 366, 372, 374, 376, 381
L-function associated to, 89, 90, 310, 314 Fourier transform, 1, 2, 30, 59, 102, 104, 106,
meromorphic continuation, 224, 285, 307, 107, 113, 309
310 Fourier Whittaker expansion, 66–73, 89,
maximal parabolic, 223, 295, 297, 300, 307, 163, 174, 179, 180, 224, 225, 279, 286,
308, 309, 321, 327, 368 319
minimal parabolic, 295, 296, 297, 303, 310, Frobenius automorphism, 398–400
313, 314, 315, 316, 324, 325, 327, 330, function field, 397
339, 370, 381, 418, 450, 451–454 functional analysis, 83, 168, 271
residue, 324, 331, 357, 403 FunctionalEquation, 424
template, 370 functional equation, 1–3, 38, 54
twisted by Maass forms, 286, 301, 302, 315, of Eisenstein series, 97, 212, 229, 307, 310,
316, 321, 357 329, 370
ElementaryMatrix, 423 of K-Bessel function, 134
elementary row and column operations, 79 of L-functions, 54, 62, 86, 91, 178, 182, 187,
endomorphism ♥, 105, 106, 108 188, 194, 195, 201, 202, 206, 208–211,
endomorphism J , 100, 105 213–217, 221, 227, 240–245, 251, 252,
Epstein zeta function, 308, 309 278, 314, 315, 365, 367–371, 374, 379,
equivalence class, 4 385, 387–392, 396, 402, 452, 453
equivalent, 4, 6, 7, 216 of the Riemann zeta function, 2, 3, 54, 90,
Euclidean topology, 20 214, 310, 376, 381
Euler product, 1, 54, 76, 85, 90, 95, 212, 215, for the Selberg class, 245
229, 235–237, 250, 253, 255–257, 365, template, 314
367, 373, 379, 382, 387, 395, 396, 398, of Whittaker functions, 114, 134, 145, 319,
401, 404, 405, 426, 427 424
even Maass form, 73, 74, 86, 87, 90, 91, 99, functoriality, 396, 397, 402–406
100, 106 fundamental discriminant, 249, 250
even primitive character, 208 fundamental domain, 6–8, 15, 27, 31, 34, 36,
explicit Bruhat decomposition, 303 55, 70, 71, 117, 377
exponential sum, 56 fundamental unit, 60
exterior algebra, 139
exterior power, 134–137, 396, 406, Galois conjugacy, 383
452–454 Galois conjugate, 339, 400
exterior square, 397 Galois extension, 397–401, 404
exterior square lift, 397 Galois group 397–402, 404
exterior power L-function, 396, 452 Galois representation, 402, 404–406, 474, 475,
482
finite field, 69, 351, 382, 398–400 Galois subgroup, 404
finite order, 236 Gamma factor, 211, 213, 231, 245, 372, 383,
finite prime, 382, 384 386, 388, 414, 415, 450, 452, 454, 470,
finite propagation speed, 111, 112 471, 475
488 Index
irreducible representation, 402 Laplacian, 39, 56, 82, 93, 100, 110–112, 216,
Iwasawa decomposition, 8–16, 24, 28, 31, 299, 285, 343
309, 424, 431, 433, 435, 466 Laurent expansion, 158
IwasawaForm, 432 laws of composition, 420
IwasawaQ, 435 LeadingMatrixBlock, 441
IwasawaXMatrix, 431 left equivalent, 78, 79
IwasawaXVariables, 434 left Haar measure, 20, 21
IwasawaYMatrix, 434 left invariant measure, 24, 27, 74, 81, 117,
IwasawaYVariables, 435 166, 269
Iwasawa form, 146, 150, 175, 185, 281, 299, level, 70, 216, 217, 383, 385
308, 328, 413, 416, 424, 432–435, Levi component, 289–292
445–447, 467, 470 L-function, 54, 76, 114, 172–178, 187, 188,
210, 239–243, 250, 255–257, 277, 278,
Jacobi identity, 40 281, 282, 322, 323, 365, 395–396
Jacobi theta function, 2, 221–223 associated to an Eisenstein series, 89, 90,
Jacobian, 124 286, 313–315, 451–454
Jacquet Whittaker function, 139, 142, 154, associated to a Maass form, 91, 174, 177,
159, 160, 175, 260, 280, 307, 360, 470, 194, 213–215, 235, 257, 279, 322, 323,
471 367, 372, 403
Jacquet’s Whittaker function, 129, 130, 136, Artin, 396–401, 404, 405
141, 144, 150, 152, 361 Dirichlet, 236, 237, 244, 249, 251, 252, 253,
256, 315
K-Bessel function, 65, 134 Hecke, 62, 401, 402
KloostermanBruhatCell, 435, 436 L-group, 406
KloostermanCompatibility, 436, 437, 438 Lie algebra, 19, 39–42, 44, 45, 50
KloostermanSum, 436–438 Lie algebra of G L(n, R), 19, 42, 50
Kloosterman sum, 337–343, 349–353, 357, Lie bracket, 41–43
362, 363, 391, 436–438 Lie group, 39, 42, 270, 396, 404
Kloosterman zeta function, 342, 343 lifting, 397, 404
Kronecker symbol, 222, 248 local Langlands conjecture, 397
Kuznetsov trace formula, 337, 354, 355, locally compact, 19–23
360–363 locally compact Hausdorff topological group,
19, 20, 21, 23
Landau’s lemma, 372, 395 LongElement, 439, 442
Langlands, 27, 195, 236, 285, 286, 366, 395, long element, 150, 224, 306, 307, 317, 341,
397, 401, 402, 403, 406 350, 351, 375, 380, 442
LanglandsIFun, 441 long element Kloosterman sum, 351, 342
LanglandsForm, 440 LowerTriangleToMatrix, 442, 448
Langlands conjecture, 195, 395–397, 402 Luo–Rudnick–Sarnak theorem, 366, 384, 387,
Langlands decomposition, 288–290, 298–301, 393
320, 440
Langlands Eisenstein series, 285, 286, 301, Maass form, 54, 62, 83, 86, 112, 216, 235,
316, 318, 319, 320, 357, 403, 417, 256, 257, 286, 310, 313, 382, 383, 403,
441 426, 427
Langlands functoriality, 396, 397, 402, 405 even and odd, 73, 74, 86, 87, 88, 90, 100,
Langlands global functoriality conjectures, 105, 106, 162, 265
396, 397 for S L(2, Z), 62, 63, 67, 69–74, 86, 91, 92,
Langlands–Shahidi method, 286 95, 210, 221, 253, 255, 256, 321–323, 387
Langlands spectral decomposition, 286, 319, for S L(3, Z), 99, 159–163, 172, 174, 176,
324, 357 182, 187, 188, 194, 196, 197, 201, 213,
Laplace operator, 38, 39, 70, 73, 95 214, 215, 227, 229, 255, 302, 324, 387
490 Index
Selberg spectral decomposition, 54, 94, 95, symmetric group, 25, 135, 292, 397, 402
285 symmetric kth, power L-function, 235, 395,
Selberg transform, 102, 355, 358 453, 454
self-adjoint, 81, 103, 106, 108, 110–112, 156, symmetric Maass form, 161, 261, 265, 315
166, 269 symmetric operator, 62
self dual, 213 symmetric power L-function, 256, 286, 395,
semigroup, 76, 80, 164, 266, 267 396, 453
Siegel set, 15, 17, 18, 108, 113, 117, 129, 131 symmetric product, 405
Siegel zero, 235, 245–249, 251–258 symmetric square, 241
Siegel zero lemma, 251, 255–258 symmetric square lift, 253–257, 397, 398
Siegel’s theorem, 249, 251, 253 symmetric square representation, 406
simple reflection, 145, 148
simultaneous eigenfunctions, 83, 277 TailingMatrixBlock, 464
skew symmetry, 40, 234 Taylor series, 43
SmithElementaryDivisors, 460 template, 314, 315, 370, 381, 396
SmithForm, 461 tensor product, 41, 134, 137, 397, 404, 405
SmithInvariantFactors, 461 tensor product lift, 397
Smith normal form, 54, 77–79, 81, 165, 268 tetrahedral representation, 402
smooth function, 32, 38, 39, 44, 46, 50, 55, 62, theta function, 2, 221–223
94, 102, 106, 110, 115, 117, 128, 242, time reversal symmetry, 111–112
308, 316, 387 topological group, 19–21, 23
solvable group, 402 topological space, 3–6, 14, 19, 20, 74–76, 163,
SpecialWeylGroup, 462 266
special orthogonal group S O(n, R), 16, 17, 27, trace formula, 99–101, 337, 354, 355, 359,
28, 33, 35, 139, 325, 355, 357–362 360–363, 383
spectral decomposition, 54, 94, 95, 285, 286, transfer, 406
319, 324–327, 355, 357 transpose-inverse, 30, 37
sphere (n, dimensional), 27, 32, 467 transpose of a matrix, 9, 80, 164, 267, 418, 419
spherical coordinates, 32 transposition, 26, 80, 81, 83, 135, 167, 269
split classical group, 397 trivial character, 217, 223, 385
split semisimple group, 99 trivial representation, 401
splits completely, 400 twist, 91, 194, 195, 206, 208, 214, 237, 402
square integrable, 55, 70, 81, 94, 106, 268, twisted L-function, 214, 215, 217, 221
426, 428 type, 9, 62, 74, 83, 84, 86, 90–92, 99, 100,
stability group, 216 114–117, 128, 130, 134, 137, 145,
stabilizer, 108, 121, 291 160–163, 176, 177, 179, 180, 182, 188,
Stade’s formula, 361, 362 193, 197, 210, 211, 213, 214, 217, 225,
standard L-function, 173 227, 228, 236, 241, 253, 255, 260,
standard parabolic subgroup, 286–288, 295, 262–265, 281, 282, 283, 315–317, 322,
297, 301, 316, 319 323, 352, 358, 366–368, 371, 372, 376,
Stirling, 157, 240, 243 381, 383, 384, 387, 401, 402, 404
Stirling’s formula, 240
strong multiplicity one, 393 unfolding, 222, 227, 339, 368, 373
subconvexity bounds, 354 unimodular, 20, 21, 429, 430, 461, 466
SubscriptedForm, 463 unipotent matrix, 151, 259, 418, 419, 432,
super diagonal, 129, 159, 175, 259, 409 433, 436, 444
SwapMatrixColumns, 463, 464 unipotent radical, 289–291
SwapMatrixRows, 464 unique factorization, 246, 399
symmetric, 6, 62, 162, 264, 266, 419 unit, 60
symmetric cube, 235, 286, 297 universal enveloping algebra, 9, 40–47, 50,
symmetric fourth power, 235, 297 114
Index 493