An Introduction To The Theory of Local Zeta Functions: Studies in Advanced Mathematics
An Introduction To The Theory of Local Zeta Functions: Studies in Advanced Mathematics
An Introduction To The Theory of Local Zeta Functions: Studies in Advanced Mathematics
Studies in
Advanced
Mathematics
S.-T. Yau, Series Editor
An Introduction
to the Theory of
Local Zeta Functions
Jun-ichi Igusa
An Introduction
to the Theory
of Local Zeta Functions
Jun-ichi Igusa
2000 Mathematics Subject Classification. Primary 11Sxx, 11S40, 11Mxx, 11Gxx, 14Gxx.
Copying and reprinting. Individual readers of this publication, and nonprofit libraries
acting for them, are permitted to make fair use of the material, such as to copy a chapter for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Requests for such
permission should be addressed to the Acquisitions Department, American Mathematical Society,
201 Charles Street, Providence, Rhode Island 02904-2294, USA. Requests can also be made by
e-mail to [email protected].
c 2000 by the American Mathematical Society and International Press. All rights reserved.
Reprinted by the American Mathematical Society, 2007.
The American Mathematical Society and International Press retain all rights
except those granted to the United States Government.
Printed in the United States of America.
∞ The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.ams.org/
Visit the International Press home page at URL: http://www.intlpress.com/
10 9 8 7 6 5 4 3 2 1 12 11 10 09 08 07
Contents
1 Preliminaries 1
1.1 Review of some basic theorems . . . . . . . . . . . . . . . . . . . 1
1.2 Noetherian rings . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Hilbert’s theorems . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4 Bernstein’s theory 45
4.1 Bernstein’s polynomial bf (s) . . . . . . . . . . . . . . . . . . . . . 45
4.2 Some properties of bf (s) . . . . . . . . . . . . . . . . . . . . . . . 47
4.3 Reduction of the proof . . . . . . . . . . . . . . . . . . . . . . . . 49
4.4 A general theorem on D-modules . . . . . . . . . . . . . . . . . . 52
4.5 Completion of the proof . . . . . . . . . . . . . . . . . . . . . . . 55
vii
viii CONTENTS
Bibliography 227
Index 231
INTRODUCTION
Local zeta functions are relatively new mathematical objects. The first general
theorems were proved from 1968 to 1973. Since then, especially during the last
fifteen years, remarkable results have been obtained, allowing one to call the accu-
mulation a “theory.” Nevertheless, there remain several challenging problems whose
solution will make the theory much richer. The purpose of this book is to introduce
the readers to this theory. The book is written in such a way that it should be
appropriate for those who have mastered the “basic courses” taught in America for
first year graduate students. Assuming the reader has this background, nearly all
material will be explained with detailed definitions and proofs. There are, however,
two exceptions. We shall use Hironaka’s desingularization theorem and the func-
tional equations of Weil’s zeta functions over finite fields. We shall explain these
theorems by examples so that the readers can accept them with some understand-
ing. The references are given primarily to indicate our indebtedness to the authors
and not for the readers to consult.
Since local zeta functions are new, we shall define them briefly with details given
in the text. If k is a number field, any completion of k is called a local field. Every
local field K carries a Haar measure, and the rate of measure change under the
multiplication by a in K × = K\{0} defines its absolute value |a|K ; it is completed
by |0|K = 0. If now X = K n for some n ≥ 1, f is a K-valued non-constant
polynomial function on X, and Φ is in the Schwartz-Bruhat space S(x) of X, then
ZΦ (s) = |f (x)|sK Φ(x) dx, Re(s) > 0
X
is called a local zeta function. If Φ is the standard function on X, i.e., exp(−π t xx)
for K = R, exp(−2π t xx̄) for K = C, and the characteristic function of OK n
for a
p-adic field K with OK as its maximal compact subring, then we drop the subscript
Φ. Furthermore, we normalize the Haar measure dx on X so that Z(s) tends to
1 as s → 0. The set of ωs (·) = | · |sK for all s in C forms the identity component
ix
x INTRODUCTION
There is a condition n > 4 for the convergence of the series. Following Weil, we
call the RHS, the right-hand side, the Eisenstein-Siegel series, and the identity itself
with a modified LHS, the Siegel formula. Later, J. G. M. Mars proved the Siegel
formula for a certain cubic form. Toward the end of 60’s, we proved the Siegel
formula for the Pfaffian and determined all cases where the Siegel formula might
hold. However the proof of the convergence of the Eisenstein-Siegel series in general
became a serious difficulty. In order to overcome this obstacle, we introduced ZΦ (ω)
over K as above and showed that the general theorems on ZΦ (ω) can effectively be
used to examine the convergence problem. In fact, it was shown to be sufficient to
estimate Z(ω) for almost all p-adic completions K of k.
These were some of the developments up to the middle 70’s. Before we start an
explanation of later activities, we mention that the general theorems were proved
by using Hironaka’s theorem except for the second proof by Bernstein. In that
proof he used the following remarkable fact: If k◦ is any field of characteristic 0
and f (x) is in k◦ [x1 , . . . , xn ]\{0}, where x1 , . . . , xn are variables, then there exists
INTRODUCTION xi
becomes a holomorphic function on the whole s-plane. The proof via Hironaka’s
theorem shows that the poles of ZΦ (s) are negative rational numbers. On the other
hand, M. Sato developed his theory of prehomogeneous vector spaces in the middle
60’s. Suppose that G is a connected reductive algebraic subgroup of GLn (C) acting
transitively on the complement of an irreducible hypersurface f −1 (0) in Cn with
f (x) necessarily homogeneous of degree say d. Then without losing generality we
can normalize G and f (x) so that t G = Ḡ = G, f (x) is in R[x1 , . . . , xn ] and further
for a monic polynomial b(s) of degree d called Sato’s b-function. By definition bf (s)
is a factor
of b(s). Actually they are equal. It can be seen, e.g., as follows: If
b(s) = (s + λ), then Z(s) has the form
Γ(s + λ)
−ds
|f (x)|C exp(−2π xx̄) dx = (2π)
s t
·
Cn Γ(λ)
λ
for Re(s) > 0. This with the above results implies b(s) = bf (s). It also gives in the
prehomogeneous case another
proof to a general theorem of M. Kashiwara stating
that all λ’s in bf (s) = (s + λ) are positive rational numbers.
Now in the p-adic case, what we did after the middle 70’s was to compute Z(ω),
especially Z(s), for those f (x) which might give the Siegel formulas. In compiling
a list of Z(s), we gradually became interested in patterns appearing in the shape
of Z(s) as a rational function of t = ωs (π) = q −s , where q = card(OK /πOK ).
We therefore started a systematic computation of Z(s) for a larger class of f (x),
especially for those f (x) which appeared in Sato’s theory, hoping to find conjectures
on Z(s). It did not take too long to find the first conjecture stating that if f (x) is
a homogeneous polynomial in k[x1 , . . . , xn ]\k, where k is a number field as before,
then
degt (Z(s)) = −deg(f )
for almost all K. We might emphasize the fact that this conjecture was not sug-
gested by any existing theory, but it came from explicit computations. At any rate
the conjecture was investigated by D. Meuser and proved as stated by J. Denef in
the late 80’s. In a similar manner, the ever-increasing list of explicitly computed
Z(s) suggested a new type of functional equations satisfied by Z(s). This conjecture
also became a theorem by Meuser and Denef in the early 90’s. More precisely, the
xii INTRODUCTION
new functional equation was derived from the functional equations of Weil’s zeta
functions over finite fields proved by A. Grothendieck. We shall devote the last
chapter to a detailed explanation of their work.
We shall explain some problems on the denominator and the numerator of Z(s)
as a rational function of t. It is known for a general f (x) that except fora power of t
and the allowance of cancellation, the denominator of Z(s) is of the form (1−q −a tb )
for some positive integers a, b. Now in all known examples bf (−a/b) = 0, i.e., the
real parts of the poles of Z(s) are zeros of bf (s), and the order of each pole is at most
equal to the order of the corresponding zero. What it says is that bf (s) for some
hidden reason describes the poles of Z(s) also in the p-adic case. This is extremely
remarkable in view of the fact that bf (s) does not play any direct role in that case.
At any rate the problem is to convert the above experimental fact into a theorem.
In the two variable case, i.e., if n = 2, the problem was settled by F. Loeser. Also in
the prehomogeneous case, it was settled jointly by T. Kimura, F. Sato, and X.-W.
Zhu, except for the information on the orders of poles stated above. In the general
case, a solution seems to require a new theory.
Again, in the case of f (x) appearing in Sato’s theory, hence d = deg(f ) =
deg(bf ), if b > 1 in some factors 1 − q −a tb of the denominator of Z(s), then Denef’s
theorem suggests that Z(s) might have a nonconstant numerator. By going through
the list of Z(s), we notice that certain cubic polynomials in t of the same type appear
rather mysteriously in the numerators of Z(s) for those f (x) which do not have any
apparent similarity. No hint to solve this mystery can be found in the complex case
by the uniform simplicity of Z(s) mentioned above, and a similarly explicit and
general form of Z(s) is not known in the real case. At any rate, no conjecture of
any kind has been proposed on how to describe the numerator of Z(s). We might
finally make it clear that there are several important results, especially those by J.
Denef, which we did not mention in this book. The reader can find most of them
in Denef’s Bourbaki seminar talk [11] and our expository paper [31].
The author would like to thank Professor S.-T. Yau for kindly inviting him
to publish this book in the AMS-IP series. The author would also like to thank
Professor M. M. Robinson for her effort to bring the manuscript into this final form.
Finally, the author would like to gratefully acknowledge the invaluable assistance
by his wife, Yoshie, for providing ideal working conditions for the last fifty years.
https://doi.org/10.1090/amsip/014/01
Chapter 1
Preliminaries
1.1 Review of some basic theorems
We shall assume that the reader is familiar with definitions and basic theorems on
groups, rings, vector spaces, and modules in algebra. We shall review two theorems,
among others, which we shall use later.
We shall assume, unless otherwise stated, that all rings are associative. If A is
any ring with the unit element 1 = 0, then we shall denote by A× the group of units
of A. If Mn (A) is the ring of n × n matrices with entries in A, then Mn (A)× will
be denoted by GLn (A). Suppose that A is commutative. Then an element a of A
is called a zero divisor if ab = 0 for some b = 0 in A. If A has no zero divisor other
than 0, then it is called an integral domain; and an element a = 0 of A is called
irreducible if a is not in A× and if a = bc for b, c in A implies that either b or c is
in A× . An integral domain is called a unique factorization ring if every a = 0 in A
can be expressed uniquely, up to a permutation and elements of A× , as a product
of irreducible elements. The classical examples are the ring Z of integers and the
ring F [x] of polynomials in one variable x with coefficients in a field F . In general,
we have the following consequence of Gauss’ lemma:
If A is a unique factorization ring and x is a variable, then A[x] is also a unique
factorization ring.
We might give some explanation. If we denote by F the quotient field of A, then
every f (x) = 0 in F [x] can be written as f (x) = cf◦ (x) with c in F × and f◦ (x) = 0 in
A[x] such that its coefficients are relatively prime. We call such an f◦ (x) primitive.
According to the Gauss lemma, the product of primitive polynomials is primitive.
Therefore if f (x), g(x) are primitive and f (x) = g(x)h(x) with h(x) in F [x], then
necessarily h(x) is in A[x] and primitive. This implies the above statement. The
irreducible elements of A[x] are irreducible elements of A and primitive polynomials
which are irreducible in F [x]. At any rate, as a corollary we see that F [x1 , . . . , xn ],
where x1 , . . . , xn are variables, is a unique factorization ring.
We shall also use the following fact, which is sometimes called the “Principle of
the irrelevance of algebraic inequalities”:
Let F denote an infinite field and f (x), g1 (x), . . . , gt (x) elements of the polyno-
mial ring F [x] = F [x1 , . . . , xn ], in which g1 (x) = 0, . . . , gt (x) = 0. Suppose that
f (a) = 0 for every a = (a1 , . . . , an ) in F n satisfying g1 (a) = 0, . . . , gt (a) = 0. Then
f (x) = 0.
1
2 JUN-ICHI IGUSA
The proof is as follows. If we put h(x) = f (x)g1 (x) . . . gt (x), then by assumption,
h(a) = 0 for every a in F n . If we can show that this implies h(x) = 0, since F [x] is
an integral domain and g1 (x) = 0, . . . , gt (x) = 0, we will have f (x) = 0. Suppose
that h(x) = 0. Suppose further that n = 1. Then the number of zeros of h(x) in F
is at most equal to deg(h). Since F is an infinite field, we have a contradiction. We
shall therefore assume that n > 1 and apply an induction on n. If we write
with c0 (x ), c1 (x ), ... in F [x1 , . . . , xn−1 ], then ci (x ) = 0 for some i by h(x) = 0.
By induction, we can find a = (a1 , . . . , an−1 ) in F n−1 satisfying ci (a ) = 0. Then
h(a , xn ) = 0, hence h(a , an ) = 0 for some an in F . We thus have a contradiction.
We shall assume that the reader is familiar with topological spaces, their com-
pactness, continuous maps, etc., in general topology. We shall review some defini-
tions and theorems.
A topological space is a nonempty set X equipped with a family of its subsets,
called open, with the property that the family is closed under the taking of arbitrary
union and finite intersection. The family then contains X and the empty set ∅. If
Y is any nonempty subset of X, then Y will be considered as a topological space
by the induced topology. The family of open sets in Y consists of intersections of
Y and open sets in X. Complements of open sets are called closed. If A is any
subset of X, the intersection A of all closed sets containing A is called the closure
of A; it is the smallest closed set containing A. An open set containing a point is
called its neighborhood. A topological space is called a Hausdorff space if any two
distinct points have disjoint neighborhoods. A Hausdorff space X with the following
property is called compact: If X is the union of open sets in a family I, then X
is also the union of open sets in a suitable finite subfamily of I. Every nonempty
closed set in a compact space is compact. The Tychonoff theorem states that the
product space, i.e., the product set with the product topology, of any nonempty set
of compact spaces is compact. A Hausdorff space is called locally compact if every
point has a neighborhood with compact closure. The fields R and C of real and
complex numbers with the usual topology are locally compact. A Hausdorff space
is called normal if any two disjoint closed sets are contained in disjoint open sets.
Every compact space is normal. If ϕ is an R-valued function on any nonempty set
X satisfying ϕ(x) ≥ α for some α in R and for every x in X, we shall write ϕ ≥ α
and also α ≤ ϕ. If A, B are disjoint nonempty closed sets in a normal space X,
there exists an R-valued continuous function ϕ on X satisfying 0 ≤ ϕ ≤ 1 such that
ϕ = 0 on A, i.e., ϕ(x) = 0 for every x in A, and ϕ = 1 on B. This remarkable
theorem is called Urysohn’s lemma. We shall later use the following fact:
A Hausdorff space X is locally compact if and only if for every neighborhood U
of any point a of X there exist an R-valued continuous function ϕ on X satisfying
0 ≤ ϕ ≤ 1 and a compact subset C of U containing a such that ϕ = 0 on X\C and
ϕ(a) = 1.
We shall outline the proof. The if-part is clear. In fact, if Oϕ denotes the set of
all x in X where ϕ(x) > 0, then Oϕ is a neighborhood of a with its closure contained
in C, hence it is compact. Conversely, suppose that X is locally compact. Then
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 3
is a power series with center a and coefficients c0 , c1 , . . . all in C, then there exists
0 ≤ r ≤ ∞, called the radius of convergence, such that P (z) is convergent (resp.
divergent) for every z in C satisfying |z − a| < r (resp. |z − a| > r). Furthermore r
is given by
1
r= 1 ,
lim supn→∞ |cn | n
which is often called Cauchy-Hadamard’s formula. As a consequence, if we define
a power series Q(z) by a termwise differentiation of P (z) as
Q(z) = ncn (z − a)n−1 ,
n>0
then Q(z) has the same radius of convergence as P (z). Furthermore, the C-valued
function P on the disc |z − a| < r defined by P (z) is differentiable at every point of
the disc and its derivative P is given by the function Q defined by Q(z). If D is any
nonempty open subset of C and a is a point of D, we put ∂D = D̄\D and denote
by dis(a, ∂D) the distance from a to ∂D. We have 0 < dis(a, ∂D) ≤ ∞. Suppose
now that f is a C-valued function on D with the following property: At every point
a of D there exists a power series P (z) with a radius of convergence r > 0 such
that f (z) = P (z) for every z in D satisfying |z − a| < r. Then we say that f is
holomorphic on D. The above remark on P and Q shows that if f is holomorphic
on D, then its derivative f exists and is holomorphic on D. We call P (z) the
Taylor expansion of f (z) at a. A standard criterion for f to be holomorphic on D
is that f is differentiable at every point of D. In fact “Cauchy’s theorem” holds for
such an f , hence f (z) can be expressed by “Cauchy’s integral formula,” and hence
it can be expanded into a Taylor series. This line of argument implies that the
above r satisfies r ≥ dis(a, ∂D), and this fact supplies a basis for the holomorphic
continuation of f . At any rate, the above criterion implies the following criterion,
sometimes called Morera’s theorem:
Let f denote a C-valued continuous function on D such that its integral along
any closed curve of finite length in D is 0. Then f is holomorphic on D.
In the application D usually has the property that it is connected and simply
connected. Then Morera’s theorem becomes the converse of Cauchy’s theorem
mentioned above. At any rate the proof is straightforward. We may assume that D
4 JUN-ICHI IGUSA
is connected. Choose a and z from D and join them by a curve C of finite length
in D. Then, by assumption, the integral of f along C is independent of the choice
of C, hence we may denote it by F (z). In this way we get a well-defined function
F on D with f as its derivative. Therefore F , hence also f , is holomorphic on D.
If a is a point of D and the function of z defined by (z − a)k f (z) becomes
holomorphic on D for some k > 0, then a is called a pole of f . In such a case,
f (z) can be expanded into a Laurent series at a, which is similar to P (z) above but
with finitely many terms cn (z − a)n for n < 0. Furthermore, the above mentioned
Cauchy’s integral formula implies that
1 f (z)
cn = dz
|z−a|=r◦ (z − a)
2πi n+1
for every n, in which 0 < r◦ < dis(a, ∂D) and the integral is from θ = 0 to 2π
after writing z − a = r◦ exp(iθ). We shall use meromorphic functions, meromorphic
continuation, etc. later on.
We shall finally review two theorems in calculus and one theorem in analysis.
The first one is called Gauss’ theorem, and it is as follows:
Suppose that D is a bounded open subset of R3 such that ∂D = D̄\D is piecewise
smooth; let f1 (x), f2 (x), f3 (x) denote continuously differentiable functions on D̄.
Then
∂f1 ∂f2 ∂f3
(f1 (x)dx2 dx3 +f2 (x)dx3 dx1 +f3 (x)dx1 dx2 ) = ( + + ) dx1 dx2 dx3 .
∂D D ∂x 1 ∂x2 ∂x3
We have formulated the above theorems rather restrictively so that the integrals
become those familiar in calculus. Actually, the third theorem is proved in analysis
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 5
Lemma 1.2.1 Let a denote an ideal of a noetherian ring A which is different from
A and meet-irreducible in the sense that it cannot be expressed as an intersection
of two ideals of A both strictly containing a. Then a is a primary ideal.
Proof. We shall assume that a is not primary and derive a contradiction. Since
a = A is already assumed, we will have the situation that ab in a, a not in a,
and b not in r(a). Since the sequence a : b, a : b2 , . . . is increasing, we will have
a : bk = a : bk+1 for some positive integer k. Then both a1 = Abk +a and a2 = Aa+a
strictly contain a. If x is in a1 , then x = cbk + d for some c in A and d in a. If x is
also in a2 , since ab is in a, we see that bx is in a. Then c is in a : bk+1 = a : bk , hence
x is in a. Therefore a is the intersection of a1 and a2 , hence a is not meet-irreducible.
In a noetherian ring A every ideal a can be expressed as an intersection of a finite
number of meet-irreducible ideals. Otherwise, starting from a we can construct an
infinite sequence of strictly increasing ideals in A. On the other hand if q1 , . . . , qt are
primary ideals of A satisfying r(qi ) = p for all i and t > 0, then their intersection q is
6 JUN-ICHI IGUSA
a primary ideal and r(q) = p. Therefore, by Lemma 1 every ideal a can be expressed
as an intersection of primary ideals q1 , . . . , qt with distinct r(q1 ), . . . , r(qt ). If no
qi is redundant, then for the sake of simplicity we call such an expression a minimal
representation of a. We shall only use the following uniqueness theorem:
Proof. If t = 0, i.e., a = A, then the theorem holds trivially. Suppose that a has
another minimal representation as an intersection of primary ideals q1 , . . . , qs and
assume by induction that min(s, t) > 0. Since the situation is symmetric, we may
assume that p = r(qt ) is maximal among r(qi ), r(qj ) for all i, j. Then p is among
r(q1 ), . . . , r(qs ). Otherwise we will have
a : qt = qi = qj = a,
1≤i<t 1≤j≤s
We observe that they are minimal representations of the LHS. Therefore, by induc-
tion the two sets {r(q1 ), . . . , r(qt−1 )} and {r(q1 ), . . . , r(qs−1 )} are the same.
We shall use the operation S −1 by a multiplicative subset S of A, i.e., a multi-
plicatively closed subset containing 1, but only in the case where S is free from zero
divisors. In that case, S −1 A simply consists of a/s for all a in A and s in S with the
convention that a/s = a /s if and only if s a = sa . We convert S −1 A into a ring
as in the special case where A = Z and S is the set of positive integers. Then S −1 A
becomes a commutative ring with 1/1 as its unit element and A can be identified
with its subring under the correspondence a → a/1. If a is an ideal of A, then S −1 a
is defined as the set of a/s for all a in a and s in S. Then S −1 a becomes an ideal of
S −1 A. We observe that the operation a → S −1 a is monotone and commutes with
the taking of finite intersection. Furthermore, if a∗ is an ideal of S −1 A, then A ∩ a∗
is an ideal of A and a∗ = S −1 (A ∩ a∗ ) because a/s = (1/s)(a/1) for every a/s in a∗
with a = a/1 in A ∩ a∗ . Therefore if a∗1 , a∗2 , . . . form a strictly increasing sequence of
ideals of S −1 A, then A ∩ a∗1 , A ∩ a∗2 , . . . form a strictly increasing sequence of ideals
of A. This implies that S −1 A is noetherian if A is noetherian.
Lemma 1.2.2 Let q denote a primary ideal of A. If q and S are disjoint, then
S −1 q is a primary ideal of S −1 A, r(S −1 q) = S −1 r(q), and A ∩ S −1 q = q. If q and
S intersect, then S −1 q = S −1 A.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 7
Proof. Since the second part is clear, we shall only prove the first part. If q and S
are disjoint, then clearly S −1 q = S −1 A and A ∩ S −1 q = q because r(q) and S are
also disjoint. Furthermore, if a/s for a in A and s in S is in S −1 q, then a is in q.
In fact, if a/s = a /s with a in q and s in S, then s a = sa is in q and s is not in
r(q), hence a is in q. This implies r(S −1 q) = S −1 r(q). We shall show that S −1 q is
a primary ideal of S −1 A. If (a/s)(a /s ) is in S −1 q with a/s not in S −1 q, then aa
is in q with a not in q, hence a /s is in S −1 r(q) = r(S −1 q).
Proposition 1.2.1 In Theorem 1.2.1 suppose that qi and S are disjoint for i ≤ r
but not for i > r. Then
S −1 a = S −1 qi
1≤i≤r
−1
is a minimal representation of S a in S −1 A.
Proof. This follows immediately from Lemma 1.2.2 except for the fact that none
of S −1 q1 , . . . , S −1 qr is redundant. If we denote by a the intersection of q1 , . . . , qr ,
then by Lemma 1.2.2 we get A ∩ S −1 a = a . Therefore, if, e.g., S −1 qr is redundant,
we will have
a = A ∩ S −1 a = qi ,
1≤i<r
Corollary 1.2.1 Let A denote a local ring with m as its maximal ideal and M a
finitely generated A-module; for every a, x respectively in A, M denote by ā, x̄
their images in A/m, M/mM and convert M/mM into a vector space over A/m as
āx̄ = ax. Then M = Ax1 + . . . + Axn if and only if x̄1 , . . . , x̄n span M/mM over
A/m.
Proof. Since the other part is clear, we shall prove the if-part. We observe that
M/mM = (A/m)x̄1 + . . . + (A/m)x̄n can be rewritten as M = (Ax1 + . . . + Axn ) +
mM . If we put N = M/(Ax1 + . . . + Axn ), then N is a finitely generated A-module
and mN = N , hence N = 0.
Proof. We take any ideal A of A[x] and show that A has a finite ideal basis. We
shall exclude the trivial case where A = 0 and denote by a the set of coefficients of
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 9
the highest powers of x which occur in elements of A. If axm , bxn are the highest
terms of elements of A and m ≥ n, then xm−n · bxn = bxm is also the higest term of
an element of A. Therefore if a, b are in a, then a + b is in a and ca for any c in A
is clearly in a. This shows that a is an ideal of A. We choose g1 (x), . . . , gs (x) from
A such that the coefficients of the highest powers of x which occur in them form an
ideal basis for a. By the same adjustment as above, we may assume that xn for some
n > 0 is the highest power of x which occurs in all of them. We denote by A the
intersection of A and A+Ax+. . .+Axn−1 and put A = A[x]g1 (x)+. . .+A[x]gs (x).
We shall show that A = A + A . Since A contains A + A , we have only to show
that A is contained in A + A . Take any f (x) from A and put deg(f ) = m. If
m < n, then f (x) is in A . Therefore we shall assume that m ≥ n and apply an
induction on m. By the choice of g1 (x), . . . , gs (x) we can find a1 , . . . , as in A such
that for
f (x) = f (x) − ai xm−n gi (x)
1≤i≤s
we have deg(f ) < m. Then by induction f (x), hence also f (x), is in A + A . The
rest of the proof is as follows.
Since A is an A-submodule of the finitely generated A-module A + Ax + . . . +
Ax n−1
, by Lemma 1.3.1 it is finitely generated. Therefore A = Af1 (x)+. . .+Afr (x)
for some f1 (x), . . . , fr (x) in A. This implies
A= A[x]fi (x) + A[x]gj (x).
1≤i≤r 1≤j≤s
We shall next explain Hilbert’s Nullstellensatz. We shall start with its shallow
generalization for a better understanding. Since we shall not use this result, we
just outline its proof. Let A denote any commutative ring with the unit element.
Then a nilpotent element of A is clearly contained in every prime ideal of A. The
converse is true and the proof is quite simple. Consider the polynomial ring A[x] as
in Lemma 1.3.2 and take any a from the intersection of all prime ideals of A. Then
1 − ax is a unit of A[x]. Otherwise 1 − ax is contained in a maximal ideal, hence a
prime ideal, say, P of A[x]. This is clear if A, hence A[x], is a noetherian ring. In
the general case, we have only to use Zorn’s lemma. At any rate, if p denotes the
intersection of A and P, then p is a prime ideal of A, hence p contains a. Then P
contains (1 − ax) + ax = 1, a contradiction. Therefore (1 − ax)f (x) = 1 for some
f (x) in A[x], and this implies an+1 = 0 if deg(f ) = n. The above result implies
that for any ideal a of A, its root r(a) is the intersection of all prime ideals of A
10 JUN-ICHI IGUSA
Lemma 1.3.3 Let m denote any maximal ideal of the polynomial ring K[x] =
K[x1 , . . . , xn ]. Then the images of x1 , . . . , xn in K[x]/m are all algebraic over K.
Proof. We denote the image of xi in K[x]/m by xi , assume that they are not all
algebraic over K, and derive a contradiction. After a permutation we may assume
that yi = xi for i ≤ r are algebraically independent over K and xi for i > r are
algebraic over K(y), where y = (y1 , . . . , yr ). We choose d(y) = 0 from K[y] so
that if we denote d(y)xi for i > r by z1 , . . . , zs , then they become zeros of monic
polynomials with coefficients in K[y] of respective degrees say n1 , . . . , ns . Since
K[x ] = K[x1 , . . . , xn ] is a field, d(y)−1 is in K[x ], hence K[x ] = K[d(y)−1 , y, z],
where z = (z1 , . . . , zs ). Furthermore if we put N = n1 · · · ns and denote z1e1 · · · zses ,
where 0 ≤ ei < ni for all i, by w1 , . . . , wN , then we will have K[y, z] = K[y]w1 +
. . . + K[y]wN . On the other hand, for some y = (y1 , . . . , yr ) with yi algebraic over
K we have d(y ) = 0. We can take y from K r if K is infinite. If we denote by
p the kernel of the homomorphism K[y] → K[y ] defined by yi → yi , then p is a
prime ideal of K[y] not containing d(y) and p = 0 by r > 0, hence K[x ]p = K[x ].
We put S = K[y]\p, A = S −1 K[y], and M = S −1 K[y, z]. Then A is a local ring
with S −1 p as its maximal ideal and M = K[x ] = Aw1 + . . . + AwN . Furthermore,
(S −1 p)M = M . This implies the contradiction M = 0 by Theorem 1.2.2.
Now Hilbert’s Nullstellensatz is a consequence of Lemma 1.3.3 and the fact that
r(a) for any ideal a of K[x] is the intersection of all maximal ideals of K[x] which
contain a. The classical statement is as follows:
Theorem 1.3.2 Let f (x), f1 (x), . . . , fr (x) denote elements of the polynomial ring
K[x] = K[x1 , . . . , xn ] and Ω any algebraically closed extension of K such that
f (a) = 0 for every a = (a1 , . . . , an ) in Ωn satisfying fi (a) = 0 for all i. Then there
exists a positive integer e and a1 (x), . . . , ar (x) in K[x] such that
f (x)e = ai (x)fi (x).
1≤i≤r
Proof. We exclude the trivial case where f (x) = 0, introduce a new variable y, and
denote by a the ideal of K[x, y] generated by f1 (x), . . . , fr (x), 1 − f (x)y. Then
Lemma 1.3.3 and the assumption imply that a is not contained in any maximal
ideal of K[x, y], hence a = K[x, y], and hence
1= ai (x, y)fi (x) + a(x, y)(1 − f (x)y)
1≤i≤r
for some a1 (x, y), . . . , ar (x, y), a(x, y) in K[x, y]. If y e◦ is the highest power of y
which occurs in a1 (x, y), . . . , ar (x, y) and e = e◦ + 1, then by replacing y by 1/f (x),
we get the relation in the theorem with ai (x) = f (x)e ai (x, 1/f (x)) in K[x] for all i.
Finally, we shall explain Hilbert’s theorem on his characteristic functions.
Hilbert’s original proof depends on the theory of syzygies. Another proof by B. L.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 11
van der Waerden depends on ideal theory. We shall explain the proof by M. Nagata
[44] which seems to be the simplest known proof. We shall start with the following
well-known lemma, in which N denotes the set of nonnegative integers:
Lemma 1.3.4 Let χ(t) denote a polynomial of degree d in one variable t with
coefficients in a field of characteristic 0 such that χ(r) is in Z for all large r in N.
Then χ(t) is necessarily of the form
t
t
χ(t) = ai , = t(t − 1) . . . (t − i + 1)/i!
i i
0≤i≤d
with a0 , a1 , . . . , ad in Z.
Proof. Since the highest degree term of ti is ti /i!, we can uniquely write χ(t) as in
the lemma with a0 , a1 , . . . , ad in the field. Since they are clearly in Z for d = 0, we
shall assume that d > 0 and apply an induction on d. If we put σ(t) = χ(t+1)−χ(t),
then deg(σ) = d − 1 and σ(r) is in Z for all large r. Therefore if we write
t
σ(t) = bi ,
i
0≤i<d
for all r. The Hilbert characteristic function, abbreviated as Hf, is the polynomial
χ(M, t) of t in the following theorem:
12 JUN-ICHI IGUSA
Proof. If χ(M, t) exists, then we say that M has an Hf. To be proved is that every
finitely generated graded A-module M has an Hf. We observe that if M has an
Hf and M # is a new graded A-module defined as (M # )r = Mr+r◦ for some fixed
r◦ and for all r, then M # also has an Hf and χ(M, t), χ(M # , t) have the same
degree. Furthermore, if M is a graded A-submodule of M and if both M and
M/M have Hf’s, then M has χ(M , t) + χ(M/M , t) as its Hf. In particular, if
deg(χ(M , t)), deg(χ(M/M , t)) ≤ n, then deg(χ(M, t)) ≤ n. After these remarks
we write M = Af1 + . . . + Afm . We may assume that each fi is in Mri for some ri .
If m = 0, hence M = 0, then M has 0 as its Hf. Furthermore, if we can show that
every M with m = 1 has an Hf, then by induction M = Af1 + . . . + Afm−1 and
M = M/M will have Hf’s, hence M has an Hf. Therefore, we have only to show
that M = Af1 has an Hf and deg(χ(M, t)) ≤ n.
If we denote by a the kernel of the A-homomorphism from A to M defined by
a → af1 , then a is a homogeneous ideal of A and (A/a)r becomes K-isomorphic
to Mr+r1 , i.e., (A/a)r is mapped K-linearly and bijectively to Mr+r1 , for all r.
Therefore, if A/a has an Hf, so does M. Furthermore, since
n+r
dimK (Fr (A/a)) ≤ dimK (Fr (A)) = ,
n
we will have deg(χ(M, t)) ≤ n. Consider the set Σ of all homogeneous ideals a of
A, different from A, with the property that A/a does not have an Hf. We have
only to derive a contradiction assuming that Σ is not empty. We choose any a
which is maximal in Σ. Then a1 is strictly contained in A1 for otherwise a becomes
A1 + A2 + . . . and A/a will have 1 as its Hf. Therefore, we can choose f from
A1 \a1 . Then the homogeneous ideal Af + a strictly contains a, hence A/(Af + a)
has an Hf, and hence σ(t) = χ(A/(Af + a), t) is defined. We observe that a : f is a
homogeneous ideal of A containing a and that the correspondence a → af gives rise
to a K-isomorphism from Ar /(a : f )r to (Af + a)r+1 /ar+1 for every r. A simple
computation of dimensions then shows that
dimK (Fr+1 (A/a)) − dimK (Fr (A/(a : f ))) = dimK (Fr+1 (A/(Af + a)))
for every r, and the RHS is equal to σ(r + 1) for all large r. If a : f = a, then
we see, as in the proof of Lemma 1.3.4, that A/a has an Hf, which is not the case.
Therefore a : f strictly contains a, hence A/(a : f ) has an Hf, and then A/a also
has an Hf. This is a contradiction.
We remark that if a is any homogeneous ideal of A = K[x1 , . . . , xn ], different
from 0, then in
t
χ(a, t) = ai
i
0≤i≤n
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 13
Chapter 2
15
16 JUN-ICHI IGUSA
The set A[[x1 , . . . , xn ]] of all such formal power series forms a commutative ring. If
K is a complete field, basic properties of sequences and infinite series for K =
R, which one learns in calculus, remain valid for K. If a series in K[[x]] =
K[[x1 , . . . , xn ]] is convergent at every a in K n satisfying a < r for some r > 0,
then it is called a convergent power series. The set Kx
= Kx
1 , i. . . , xn
of all
such
Proof. We shall prove the first part, i.e., f = 0 implies f (x) = 0. This can be done
by using the principle of the irrelevance of algebraic inequalities in Chapter 1.1 or
directly as follows. Suppose first that n = 1 and
f (x) = c i xi , ck = 0
i≥k
in which fi (x ) are all in K[[x1 , . . . , xn−1 ]]. Then they are convergent at every a
in K n−1 satisfying a < r, where · is relative to K n−1 . Since fk (x ) = 0,
by induction we can find a such that fk (a ) = 0. Then f (a , xn ) = 0, hence
f (a , an ) = 0 for some an in K satisfying |an |K < r, and hence f = 0. This is a
contradiction.
We shall prove the second part. As in calculus, we may assume that n = 1.
Take a from K, h from K × satisfying |a|K + |h|K < r. Then we have
f (a + h) − f (a)
M r|h|K
− ici ai−1 ≤ .
h K (r − |a|K − |h|K )(r − |a|K )2
Theorem 2.1.1 (i) Let K denote an arbitrary field and assume for some m, n that
every Fi (x, y) in F (x, y) = (F1 (x, y), . . . , Fm (x, y)) is in
K[[x, y]] = K[[x1 , . . . , xn , y1 , . . . , ym ]] satisfying Fi (0, 0) = 0 and further
with aij , cijk in K, in which ci0k = 0 for |k| = 1, then aij = (∂Fi /∂yj )(0, 0) for all
i, j. The square matrix a with aij as its (i, j)-entry is in GLm (K) by assumption.
We observe that f (x) satisfying F (x, f (x)) = 0 is not affected by any invertible
K-linear transformation of entries of F (x, y). Therefore, after multiplying a−1 to
F (x, y), regarded as a column vector, we may assume that aij = δij , i.e., aii = 1
and aij = 0 for i = j. After this adjustment we write
fi (x) = dij xj
|j|>0
for all i, if we denote by fip (x) the homogeneous part of degree p in fi (x), then
F (x, f (x)) = 0 becomes equivalent to
fip (x) = cijk xj fα pαβ (x) ,
1≤α≤m 1≤β≤kα
Furthermore, if p > 1, then in (∗) we have pαβ < p for cijk = 0. In fact if pαβ ≥ p,
hence pαβ = p, for some α, β, then j = 0, kα = 1, and kα = 0 for α = α. This
implies |k| = 1 and cijk = ci0k = 0. We have thus shown that fi1 (x) is as above and
fip (x) for p > 1 is determined by fi p (x) for p < p. Therefore fi (x) is uniquely
determined by an induction on p.
Proof of (ii). We start with an additional observation still in the case where K is
an arbitrary field. The above proof shows that dij = cij0 for |j| = 1 and that dij
for |j| = p > 1 is a polynomial in ci j k and di j for |j | + |k | ≤ p and |j | < p
with coefficients in N. Therefore, again by an induction on p, we see that
then |cijk |K ≤ c◦ijk for all i, j, k and by (i) we have a unique fi◦ (x) =
(f1 (x), . . . , fn (x)) in R[[x]] satisfying fi◦ (0) = 0 and F ◦ (x, f ◦ (x)) = 0. Further-
◦ ◦
more
fi◦ (x) = d◦ij xj , d◦ij = Pij (c◦i j k )
|j|>0
with the same polynomial Pij as before. Since the coefficients of Pij are in N,
therefore, we get
for all i, j. If, for a moment, we accept the fact that fi◦ (x) is in Rx
, then by
definition fi (x) << fi◦ (x) for all i. Therefore the formal identity F (x, f (x)) = 0
implies F (a, f (a)) = 0 for all a near 0 in K n .
We shall show that if (a, b) is near (0, 0) in K n × K m and F (a, b) = 0, then
b = f (a). If y1 , . . . , ym
are other variables, then by the remark after Lemma 2.1.2
we can write
Gi (x, y) − Gi (x, y ) = Hij (x, y, y )(yj − yj )
1≤j≤m
hence
(δij − Hij (a, b, f (a)))(bj − fj (a)) = 0
1≤j≤m
for all i. Since Hij (a, b, b ) depends continuously on (a, b, b ) near (0, 0, 0) in K n ×
K m × K m by Lemma 2.1.3 and Hij (0, 0, 0) = 0, the coefficient-matrix is invertible
for (a, b) near (0, 0), hence b = f (a).
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 21
M
G◦i (x, y) = − M (1 + Y /r),
(1 − X/r)(1 − Y /r)
then there exists a unique f (x) = (f1 (x), . . . , fn (x)) with fi (x) in K[[x]] satisfying
fi (0) = 0 for all i and g(f (x)) = x. (ii) In the above situation, if K is a complete
field and gi (x) is in Kx
, then fi (x) is also in Kx
for all i. Furthermore if b
is near 0 in K n and a = g(b), then a is also near 0 in K n and b = f (a). Therefore
y = f (x) gives rise to a bicontinuous map from a small neighborhood of 0 in K n to
another.
This follows immediately from Theorem 2.1.1. We have only to take Fi (x, y) =
xi − gi (y) for 1 ≤ i ≤ m = n.
Corollary 2.1.2 If g(0) = 0 for g(x) in K[[x]], then 1/g(x) can be expressed
uniquely as an element of K[[x]]; if further g(x) is in Kx
, then 1/g(x) is also
in Kx
.
We may assume that g(0) = 1. Then we have only to apply Theorem 2.1.1 to
F (x, y) = g(x)(1 + y) − 1 for m = 1. We also mention that if char(K) does not
divide a positive integer m and if the m-th power map is surjective on K × , then
the m-th power map is also surjective on K[[x]]× . In fact if g(x) is any element
of K[[x]] with g(0) = 0, then am = g(0) for some a in K × . This time we apply
Theorem 2.1.1 to F (x, y) = (a + y)m − g(x).
Theorem 2.2.1 (i) If Fi (x, y) is in OK [[x, y]] and Fi (0, 0) = 0 for all i and further
then every fi (x) in the unique solution f (x) = (f1 (x), . . . , fm (x)) of F (x, f (x)) = 0
satisfying fi (0) = 0 is in OK [[x]]. (ii) If every Fi (x, y) is an SRP in
x1 , . . . , xn , y1 , . . . , ym , then every fi (x) is an SRP in x1 , . . . , xn . Furthermore if a
is in OK n
, then f (a) is in OK m
and F (a, f (a)) = 0, and if (a, b) in OK n
× OKm
satisfies
F (a, b) = 0, then b = f (a).
with aij , cijk in OK this time, in which ci0k = 0 for |k| = 1, then the square matrix
a with aij as its (i, j)-entry is in GLm (OK ) by assumption. Therefore, after the
normalization aij = δij by multiplying a−1 to F (x, y), the new cijk is still in OK
for all i, j, k. Since
fi (x) = dij xj , dij = Pij (ci j k ),
|j|>0
in which the coefficients of the polynomial Pij are in N, we see that fi (x) is in
OK [[x]] for all i.
Proof of (ii). We observe that the normalization aij = δij does not affect the
assumption that Fi (x, y) is an SRP in x1 , . . . , xn , y1 , . . . , ym for all i. Therefore
Fi# (x, y) = πFi (π −1 x, π −1 y) is in OK [[x, y]] and Fi# (0, 0) = 0 for all i, and further
∂(F1# , · · · , Fm
#
)/∂(y1 , · · · , ym )(0, 0) = 1.
Therefore if g(x) = (g1 (x), · · · , gm (x)), where gi (0) = 0, is the unique solution of
F # (x, g(x)) = 0, then every gi (x) is in OK [[x]] by what we have shown. On the
other hand, F (x, f (x)) = 0 implies F # (x, f # (x)) = 0, in which fi# (0) = 0 for all
i. Therefore by the uniqueness we get fi# (x) = gi (x), hence fi (x) is an SRP in
x1 , · · · , xn for all i. In particular if a is in OKn
, then f (a) is in OK m
and the formal
identity F (x, f (x)) = 0 implies F (a, f (a)) = 0.
We shall show that if (a, b) in OK n
× OK
m
satisfies F (a, b) = 0, then b = f (a). We
observe that the LHS of
Gi (x, y) − Gi (x, y ) = Hij (x, y, y )(yj − yj )
i≤j≤m
for all i with the coefficient-matrix in GLm (K), in fact in GLm (OK ) because its
determinant is in 1 + πOK , we see that b = f (a).
Theorem 2.2.1 has a corollary similar to that of Theorem 2.1.1. We shall use
the same notation as in that corollary.
Corollary 2.2.1 (i) If gi (x) is in OK [[x]] and gi (0) = 0 for all i and further
then every fi (x) in the unique solution of g(f (x)) = x satisfying fi (0) = 0 is also
in OK [[x]]. (ii) If every gi (x) is an SRP in x1 , . . . , xn , then every fi (x) is also an
n
SRP in the same variables, and y = f (x) gives rise to a bicontinuous map from OK
to itself
Theorem 2.3.1 If F (x, y) is in K[[x, y]] = K[[x1 , . . . , xn , y]] and has the property
that F (0, y) is different from 0 and has cy m for some m > 0 as its leading form,
then F (x, y) can be written uniquely as
in which E(x, y) is in K[[x, y]] with E(0, 0) = 0 and a1 (x), . . . , am (x) are in K[[x]],
hence necessarily ai (0) = 0 for all i and E(0, 0) = c. Furthermore if F (x, y) is
a convergent power series, then a1 (x), . . . , am (x) and E(x, y) are also convergent
power series.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 25
Proof. We shall first show that we indeed have ai (0) = 0 for all i and E(0, 0) = c.
Suppose that am−i (0) = 0 for all i < k and for some k < m. Then we will have
By comparing the coefficients of y k on both sides we get 0 = E(0, 0)am−k (0), hence
am−k (0) = 0. Therefore by induction we get ai (0) = 0 for all i. Then by comparing
the coefficients of y m on both sides we get c = E(0, 0).
After dividing F (x, y) and E(x, y) by c, we may assume that c = 1. Since E(x, y)
is a unit of K[[x, y]] by Corollary 2.1.2, we replace it by its inverse say H(x, y). Also
we replace F (x, y) by y m − G(x, y). Then the equation to be solved becomes
and denote the homogeneous parts of gi (x), hi (x) of degree j by gij (x), hij (x) for
all j. The condition that y m is the leading form of F (0, y) then becomes
for all l. In the above, δ0 (k) is the function taking the value 1 at k = 0 and 0
elsewhere. If we incorporate the fact that gi0 (x) = 0 for all i ≤ m, then we finally
get
(∗) hkl (x) = δ0 (k)δ0 (l) + gk+mi,0 (x)hil (x)
0≤i<k
+ gk+m−i,l−j (x)hij (x)
0≤i≤k+m 0≤j<l
induction on φ(k, l) for all k, l starting with h00 (x) = 1. Hence H(x, y) and also
a1 (x), . . . , am (x) are uniquely dertermined. We have thus shown that a solution by
formal power series exists and is unique.
We shall show that if G(x, y) is a convergent power series, then the unique
H(x, y) is also a convergent power series. As we have seen in section 2.1, if we
choose M , r > 0 suitably, then we will have
M
G(x, y) << G◦ (x, y) = (x1 + . . . + xn )p y q ,
r p+q
in which the summation is over N2 and indicates the restriction q > m for
p = 0. We shall show that the unique H ◦ (x, y) for G◦ (x, y) gives a dominant series
for H(x, y). We shall first make H ◦ (x, y) explicit. If we write
y i
G◦ (x, y) = ◦
gij (x) ,
r
i,j≥0
we will have j
◦ x1 + . . . + xn
gij (x) =M
r
◦
with the exception that gi0 (x) = 0 for all i ≤ m. Therefore by (∗) we get
y i
H(x, y) << H ◦ (x, y) = h◦ij (x) ,
r
i,j≥0
in which
The induction is thus complete. Therefore the series aij X j Y i is convergent for
−1 −1
|X| < β and |Y | < α .
The rest of the proof is straightforward. If F (x, y) is a convergent power series,
so is G(x, y) = y m − F (x, y), hence also H(x, y) by what we have shown. Since
H(0, 0) = 1, the inverse E(x, y) of H(x, y) is a convergent power series. Also
in which the coefficients a1 (x), . . . , am (x) are power series satisfying ai (0) = 0 for
all i, is called a Weierstrass polynomial.
If g is an
element of GLn (K) with gij as its (i, j)-entry, then the correspondence
xi → yi = gij xj gives rise to a K-automorphism of An . If f (x) is any element
of An \{0} with fm (x) as its leading form, since every complete field is infinite, we
will have fm (a) = 0 for some a in K n \{0}. We can then find g in GLn (K) with
a as its last column. In fact if the k-th entry of a is different from 0, we can take
e1 , . . . , ek−1 , ek+1 , . . . , en , where e1 = t (1, 0, . . . , 0), etc., as the first, ..., the (n−1)-
th columns of g. If in f (y) = f (gx) we put xi = 0 for all i < n, then we get cxm n +. . .
with c = fm (a) = 0. By applying the above observation to a(x)b(x)p(x) as f (x),
we may assume that a(0, xn )b(0, xn )p(0, xn ) = 0. Then by Theorem 2.3.1 we may
further assume that a(x), b(x), p(x) are Weierstrass polynomials in Bn = An−1 [xn ].
We know by induction and by a consequence of Gauss’ lemma that Bn is a unique
factorization ring. Furthermore a Weierstrass polynomial in Bn is a unit of An if
and only if it is 1; that a Weierstrass polynomial is irreducible in An if and only if
it is irreducible in Bn . Therefore p(x) divides either a(x) or b(x).
We remark that if K is any infinite field, the above proof is applicable without
any change to K[[x]] = K[[x1 , . . . , xn ]], hence it is also a unique factorization ring.
We might mention that both K[[x]] and Kx
are noetherian rings, and the idea
of the proof for K[[x]] is as follows. If A is any ideal of K[[x]], then the leading
forms of elements of A\{0} generate an ideal a of K[x]. If we choose a finite subset
I of A\{0} such that the set of leading forms of its elements forms an ideal basis
for a, then I forms an ideal basis for A.
ψV ◦ f ◦ φ−1
U : φU (U ∩ f
−1
(V )) → K dim(Y )
Therefore we have only to observe that the correspondence f → (∂f /∂xi )(a) defines
an element (∂/∂xi )a of Ta (X) and
∂ ∂
∂= ∂xi , [ , (dxj )a ] = δij
∂xi a ∂xi a
1≤i≤n
for all ∂ in Ta (X) and all i, j. We observe that if (df1 )a , . . . , (dfn )a for f1 , . . . , fn in
Oa are linearly independent, then by Corollary 2.1.1-(ii) there exists a neighborhood
V of a in U such that ψV = (f1 , . . . , fn ) gives a K-bianalytic map from V to an
open subset of K n . Therefore the addition of (V, ψV ) to the given atlas {(U, φU )}
on X will produce an equivalent atlas. We shall apply such a process whenever it
becomes necessary. We call (f1 , . . . , fn ) local coordinates of X around a.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 31
We shall use the Grassmann or the exterior algebra of a vector space. We shall
briefly recall its definition. In general, if A is a vector space over an arbitrary field K
equipped with a K-bilinear multiplication A × A → A, then A is called a K-algebra.
We shall consider, for the time being, only associative K-algebras each with the unit
element. If E is a vector space over K, assumed to be finite dimensional, then the
exterior algebra (E) of E is the K-algebra generated by E with v 2 = 0 for every
v in E as its “defining relation.” A more precise definition is as follows. If E, E
are vector spaces over K, their tensor product E ⊗ E is the vector space over K
with a K-bilinear map (v, v ) → v ⊗ v from E × E to E ⊗ E such that E ⊗ E
is spanned by the image and dimK (E ⊗ E ) = dimK (E) dimK (E ). If we choose
K-bases for E, E , the set of formal products of their members forms a K-basis for
E ⊗ E . This fact can be used as a non-intrinsic definition of E ⊗ E . At any rate,
if we denote by T (E) the direct sum of K, E, E ⊗ E, ... and define a product in
T (E) by ⊗, then T (E) becomes a K-algebra, and it is called the tensor algebra of
E. IfI(E) denotes the two-sided ideal of T (E) generated by v ⊗ v for all v in E,
then (E) is the factor ring T (E)/I(E). If v1 , . . . , vp are elements of E, the image
of v
1 ⊗ . . . ⊗ vp in (E) is denoted by v1 ∧ . . . ∧ vp and the K-span of such elements
p
by (E). We have
v ∧ v = 0, v ∧ v + v ∧ v = 0
p
for every v, v in E. As a vectorspace (E) is the direct sum of (E) for 0 ≤ p ≤ n
if dimK (E) = n, hence dimK ( (E)) = 2n .
If now we take Ωa (X) as E and if φU (x) = (x1 , . . . , xn ) with U containing a,
then we get
p
(Ωa (X)) = K (dxi1 )a ∧ . . . ∧ (dxip )a ;
i1 <...<ip
in which fU,i1 ...ip are K-valued functions on U . If they are all K-analytic functions
on U for every U , we say that α is a K-analytic differential form of degree p on
X. In particular, if f is any K-analytic function on X, then df is a K-analytic
differential form of degree 1 on X.
If f : X → Y is a K-analytic map of K-analytic manifolds and β is a K-
analytic differential form of degree p on Y , then we get a similar differential form
(δf )∗ (β) on X as follows. If a is any point of X and f (a) = b, then a K-linear map
δa f : Ta (X) → Tb (Y ) is defined as (δa f )(∂)(g) = ∂(g ◦ f ) for all g in OY,b and its
dual map (δa f )∗ : Ωb (Y ) → Ωa (X) is given by
Proof. Since F (x) is separably algebraic over F (x ), we can write F (x) = F (x , y)
with y satisfying f◦ (y) = 0 for a unique irreducible monic polynomial f◦ (t) in
F (x )[t], where t is a variable. Since y is a simple root of f◦ (t), we have (df◦ /dt)(y) =
0. By multiplying a common denominator of the coefficients, we can convert f◦ (t)
into a primitive polynomial f1 (x , t) in F [x ][t] = F [x , t]. Then by a consequence
of Gauss’ lemma the kernel of the F -algebra homomorphism F [x , t] → F [x , y]
defined by x → x , t → y is the principal ideal generated by f1 (x , t). Therefore
if we denote the coordinates on Ld+1 by v1 , . . . , vd+1 , then by our previous obser-
vation the correspondence ∂ → (∂x1 , . . . , ∂xd , ∂y) gives an L-linear bijection from
DerF (F [x , y], L) to the subspace of Ld+1 defined by
∂f1 ∂f1
(x , y) vi + (x , y) vd+1 = 0,
∂xi ∂t
1≤i≤d
Theorem 2.5.1 Let F denote a field with char(F ) = 0 and f (t) any element of the
polynomial ring F [t] = F [t1 , . . . , tn ]; define Cf as the set of all a in F n satisfying
(∂f /∂t1 )(a) = . . . = (∂f /∂tn )(a) = 0 and Vf as the set of f (a) for all a in Cf .
Then Vf is a finite subset of F .
If r = 0, hence a = F [t], we see that Cf and Vf are empty sets. Therefore we shall
assume that r > 0, choose q = qj and put p = r(q). If we denote by xi the image of
ti in F [t]/p, then F [x] = F [x1 , . . . , xn ] is an integral domain and (∂f /∂ti )(x) = 0
for 1 ≤ i ≤ n. We shall show that x0 = f (x) is in F .
If x0 is not in F , since F is algebraically closed, it is transcendental over F .
Therefore we may assume after a permutation that x0 , x1 , . . . , xd for some d ≥ 0
are algebraically independent over F and F (x) = F (x0 , x) is algebraic, necessarily
separable by char(F ) = 0, over F (x0 , x1 , . . . , xd ). Then by Lemma 2.5.1 there exists
an element ∂ of DerF (F [x], F (x)) so that ∂x0 takes any preassigned value in F (x).
On the other hand x0 = f (x) implies
∂f
∂x0 = (x) ∂xi = 0.
∂ti
1≤i≤n
Chapter 3
Hironaka’s desingularization
theorem
3.1 Monoidal transformations
Hironaka’s desingularization is achieved by successive monoidal transformations;
they have been known in algebraic geometry for a long time. We just mention
O. Zariski’s paper [62] which contains a rigorous definition of a general monoidal
transformation and its basic properties. We shall explain a monoidal transformation
with smooth center following A. Borel and J.-P. Serre [4]. We fix a complete field
K and start with a definition of the projective space Pn (K).
We regard two points of K n+1 \{0} to be equivalent if they differ by a scalar
factor in K × and denote the set of all equivalence classes by Pn (K). If t is a point
of Pn (K), therefore, it is represented by some (t1 , . . . , tn+1 ) in K n+1 \{0}, called
the homogeneous coordinates of t. The condition ti = 0 on t is independent of
the choice of its homogeneous coordinates and defines a subset Ui of Pn (K) for
1 ≤ i ≤ n + 1. If t is in Ui , then a map φ : Ui → K n is well defined as
t1 ti−1 ti+1 tn+1
φi (t) = ,··· , , ,··· , .
ti ti ti ti
We observe that φi is a bijection. We shall make the map
φj ◦ φ−1
i : φi (Ui ∩ Uj ) → φj (Uj ∩ Ui )
35
36 JUN-ICHI IGUSA
open if and only if U ∩ Ui is open in Ui for all i. Then Pn (K) becomes a Hausdorff
space. Furthermore, again by (∗), we see that {(Ui , fi )} gives an atlas on Pn (K),
hence Pn (K) becomes an n-dimensional K-analytic manifold. We keep in mind that
the map K n+1 \{0} → Pn (K) defined by (t1 , . . . , tn+1 ) → t is K-analytic, hence
continuous, and its restriction to the subset defined by max(|t1 |K , . . . , |tn+1 |K ) = 1
is surjective.
We now take an n-dimensional K-analytic manifold X and a closed submanifold
C with p = codimX (C) ≥ 2, and define the monoidal transformation f : X # → X
with center C. It will have the following properties: Firstly X # is also an n-
dimensional K-analytic manifold; secondly f is a K-analytic map which induces a
K-bianalytic map X # \f −1 (C) → X\C, where the preimage f −1 (C) of C under f
is a closed submanifold of X # of codimension 1, called the exceptional divisor of f ;
thirdly f −1 (a) for every a in C is a closed submanifold of X # which is K-bianalytic
to Pp−1 (K). In particular f is surjective. In the special case where C is a point f
is often called the quadratic transformation with center C.
We take an atlas {(U, φU )} on X with φU (x) = (x1 , . . . , xn ) such that if U ∩C =
∅, then it consists of all x in U satisfying x1 = . . . = xp = 0. We then define for each
U an n-dimensional K-analytic manifold U # equipped with a K-analyticsurjection
fU : U # → U and piece them together to get X # and f : X # → X as f U # = fU .
If U ∩ C = ∅, we simply take U # = U and fU = idU , the identity map of U . If
U ∩ C = ∅, then U # is the closed subset of U × Pp−1 (K) defined as follows: It
consists of all (x, t) satisfying xi tj − xj ti = 0 for 1 ≤ i < j ≤ p, in which (t1 , . . . , tp )
are the homogeneous coordinates of t. We put fU (x, t) = x. By definition if x
is not in C, then t has (x1 , . . . , xp ) as its homogeneous coordinates and fU−1 (x)
consists of the single point (x, t) for that t. On the other hand if x is in C, then
fU−1 (x) = x × Pp−1 (K). After this simple observation, we shall examine U # more
closely.
We take the open covering of Pp−1 (K) by Vi defined by ti = 0 and introduce
local coordinates of t in Vi as
t1 ti−1 ti+1 tp
(u1 , . . . , up−1 ) = ,... , , ,... ,
ti ti ti ti
Furthermore if (u1 , . . . , up−1 ) and (v1 , . . . , vp−1 ) are the local coordinates of t re-
spectively in V1 and V2 , then
# −1
φ#
2 ◦ (φ1 ) : φ# # # # # #
1 (U1 ∩ U2 ) → φ2 (U2 ∩ U1 )
is given by
1 u2 up−1
(x1 , xp+1 , . . . , xn , u1 , u2 , . . . , up−1 ) → x1 u1 , xp+1 , . . . , xn , , ,... , .
u1 u1 u1
Therefore it is clearly K-analytic. We have thus shown that {(Ui# , fi# )} gives an
atlas on U # , hence it becomes an n-dimensional K-analytic manifold. Also we
see by (∗∗) that the map fU : U # → U is K-analytic. Furthermore fU−1 (U ∩ C)
becomes a closed submanifold of U # of codimension 1 because we see by (∗∗) that it
is defined in each Ui# by xi = 0. Finally fU maps U # \fU−1 (U ∩ C) K-bianalytically
to U \(U ∩ C) because we see again by (∗∗) that on the open subset of U defined
by xi = 0 the inverse of fU is given by
x1 xi−1 xi+1 xp
x → xi , xp+1 , . . . , xn , , . . . , , ,... ,
xi xi xi xi
for 1 ≤ i ≤ p.
We shall construct X # and f : X # → X out of the set {(U # , fU )} for all U in
the atlas {(U, fU )} on X. We take U , U with U = U ∩ U = ∅. If U ∩ C = ∅,
then fU , fU respectively map fU−1 (U ), fU−1
(U ) K-bianalytically to U . We shall
#
identify the above open subsets of U , (U ) by this bijection. If U ∩ C = ∅, then
#
is based on the following fact: If φU (x) = (x1 , . . . , xn ), fU (x) = (x1 , . . . , xn ) and
(u1 , . . . , up−1 ) are the local coordinates of t in Vi , then
# −1 #
φ#
U ◦ (φU ) : φ# #
U (U ∩ (U × Vi )) → φU ((U ) ∩ (U × Vi ))
#
is given by
Since x1 , . . . , xn are K-analytic functions of (x1 , . . . , xn ), we see by (∗∗) that the
above map is K-analytic. Therefore if we define X # as the union of all U # and
topologize X # as in the case of Pn (K), then X # becomes a Hausdorff space and
the set of charts defined above gives an atlas on X # . In this way X # becomes an
n-dimensional K-analytic manifold. If finally we define f : X # → X as f U # = fU
for every U , then f becomes a K-analyic map with the properties stated in the
beginning.
We add the following remark for our later use. If K is any field with an absolute
value |·|K , then for any ε > 0 we define Iε as the set of all a in K satisfying |a|K ≤ ε.
By making ε smaller if necessary, we shall assume that ε = |a0 |K for some a0 in K × .
We observe that K is locally compact if and only if I1 is compact. This follows
from the fact that a → a−1 0 a maps Iε bicontinuously to I1 . If K is locally compact,
38 JUN-ICHI IGUSA
f −1 (0)sing = f −1 (0) ∩ Cf ,
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 39
in which Cf is the critical set of f . Then we have the following three possibilities:
Firstly f −1 (0) = ∅; secondly f −1 (0) = ∅ and f −1 (0)sing = ∅; thirdly f −1 (0)sing = ∅.
All these cases occur, e.g., for K = R. Simple respective examples are as follows:
f (x) = x2i + 1, x2i − 1, x2i − x2j (1 ≤ p ≤ n).
1≤i≤n 1≤i≤n 1≤i≤p p<j≤n
At any rate in the second case f −1 (0) becomes a closed submanifold of X of codi-
mension 1 with f as its local equation around every point of f −1 (0). In an over-
simplified manner we can say that Hironaka’s desingularization theorem or rather
its consequence gives a method to improve the third case. An exact statement,
including the trivial case where n = 1, is as follows:
Theorem 3.2.1 Let K denote a complete field with char(K) = 0 and f (x) any
element, not in K, of the polynomial ring K[x1 , . . . , xn ] for n ≥ 1; put X =
K n . Then there exist an n-dimensional K-analytic manifold Y , a finite set E =
{E} of closed submanifolds of Y of codimension 1 with a pair of positive integers
(NE , nE ) assigned to each E, and a K-analytic map h : Y → X satisfying the
following conditions: Firstly, h is the composite map of a finite number of monoidal
transformations each with a smooth center; secondly,
(f ◦ h)−1 (0) = E
E∈E
We observe that if
f −1 (0)smooth = f −1 (0)\Cf
is not empty and if we denote by E the union of those E not contained in
h−1 (f −1 (0)sing ), then h gives rise to a K-bianalytic map of E \h−1 (f −1 (0)sing )
to f −1 (0)smooth . We call E the strict transform of f −1 (0) under h. The nerve
complex N (E) of E with the function E → (NE , nE ), called the numerical data, on
the set of its vertices is an important combinatorial object associated with f (x) or
rather f −1 (0)sing .
40 JUN-ICHI IGUSA
in F [[x]][y] by a unit of F [[x, y]]. We shall assume, for the sake of simplicity, that
f (x, y) is irreducible in Ω[[x, y]] for an algebraically closed extension Ω of F . Then
Px (y) is irreducible in Ω((x))[y]. Therefore if η is a zero of Px (y), then L = Ω((x))(η)
is an extension of Ω((x)) of degree m. The fact is that if x1/m is any m-th root of
x, then L = Ω((x1/m )). One way to see this is as follows:
We observe that K = Ω((x)) is a complete non-archimedean field with OK =
Ω[[x]] and xOK as its maximal ideal. Therefore L is also a complete nonarchimedean
field and, since OK /xOK = Ω is algebraically closed, we will have OL = Ω[[πL ]]
and xOL = πL m
OL . This follows, e.g., from Proposition 11.6.1. Furthermore, as we
×
have remarked at the end of Chapter 2.1, the m-th power map from OL to itself is
surjective. Therefore x 1/m
is in OL and x 1/m
OL = πL OL , hence OL = Ω[[x1/m ]]
and L = Ω((x1/m )).
Since η is an element of OL , it becomes a power series in x1/m . We write this
“Puiseux series” as
η= a0i xi + a1i x(µ1 +i)/ν1 + a2i x(µ2 +i)/ν1 ν2 + . . .
0<i≤j0 0≤i≤j1 0≤i≤j2
(µg +i)/ν1 ···νg
+ agi x ,
i≥0
in which the exponents are strictly increasing, a10 a20 · · · ag0 = 0, µi , νi are relatively
prime positive integers with νi > 1 for 1 ≤ i ≤ g, and ν1 ν2 · · · νg = m. Furthermore
µ1 > ν1 . A basic fact is that the g pairs
depend only on the factor ring Ω[[x, y]]/Ω[[x, y]]f (x, y) and, to some extent, they
determine the ring. At any rate they are called the characteristic pairs of the factor
ring.
We shall now replace F by a complete field K and put X = K 2 . We shall
assume, for the sake of simplicity, that f −1 (0)sing = {0}. Then there exists a
unique shortest sequence of quadratic transformations such that their composition
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 41
in which a1 > a2 > . . . > at = kt > 1 for some t > 0 and k1 , . . . , kt−1 > 0. We shall
write a0 /a1 = [k0 , k1 , . . . , kt ]. In this notation we introduce kij as
NIi NIi−1 µi
= + − µi−1 νi2 ,
mi mi−1 νi
µi
n Ii = (nIi−1 + − µi−1 )νi
νi
for 1 ≤ i ≤ g with the understanding that (NI0 , nI0 ) = (0, 1), hence
n I1 µ1 + ν1
= .
NI1 µ1 ν1 · · · νg
Furthermore nI /NI > nI1 /NI1 if I < I1 and nI /NI > nIi /NIi if I > Ii for 1 ≤ i ≤ g,
hence nI1 /NI1 is smaller than any other nI /NI .
The above-outlined desingularization of f −1 (0) is due to F. Enriques and O.
Chisini, and it is entirely classical. The exact values of (NI , nI ) for all I and the fact
that nI1 /NI1 is smaller than any other nI /NI can be found in [24]. The structure of
42 JUN-ICHI IGUSA
Since the two curves in X1 with local equations y1 , 1−(x1 )3 y1 have normal crossings,
we apply a QT to X1 with 0 as its center. Then, by increasing the subscripts by 1,
in X2 , X2 we have
Since the two curves in X2 with local equations x2 , x2 y22 − 1 have normal crossings,
we apply a QT to X2 with 0 as its center. Then in X3 , X3 we have
4 12
f (x, y) = x12 6
3 y3 f3 (x3 , y3 ) = (x3 ) (y3 ) f3 (x3 , y3 ),
dx ∧ dy = x43 y32 dx3 ∧ dy3 = x3 (y3 )4 dx3 ∧ dy3 ,
f3 (x, y) = (1 − y)2 − x6 y 3 − 4x3 y 2 ,
f3 (x, y) = (x − 1)2 − x5 y 6 − 4x3 y 3 .
We apply a QT to X3 with (1, 0) as its center. Since (1, 0) in X3 and (0, 1) in X3
represent the same point, we could have applied a QT to X3 with (0, 1) as its center.
At any rate in X4 , X4 we have
f (x, y) = (1 + x4 )4 x14 12
4 y4 f4 (x4 , y4 )
= (1 + x4 y4 )4 (y4 )14 f4 (x4 , y4 ),
dx ∧ dy = (1 + x4 )x54 y44 dx4 ∧ dy4
= (1 + x4 y4 )(y4 )5 dx4 ∧ dy4 ,
f4 (x, y) = 1 − 4(1 + x)3 xy 3 − (1 + x)5 x4 y 6 ,
f4 (x, y) = x2 − 4(1 + xy)3 y − (1 + xy)5 y 4 .
44 JUN-ICHI IGUSA
If we denote by Y the union of X1 , X2 , X3 \ {(0, 1)}, X4 , X5 , X6 , X6 and define
h on each one of them as the composition of X1 → X, X2 → X1 → X, . . . ,
X6 → X5 → X4 → X3 → X2 → X1 → X, then h : Y → X gives a desingularization
of f −1 (0). In fact, a list of local equations for E1 , E2 , . . . , E7 is as follows:
in which fi , fi are fi (xi , yi ), fi (xi , yi ) for all i. We can easily verify that E =
{E1 , E2 , . . . , E7 } has normal crossings. Therefore
(Ni , ni ) = (4, 2), (6, 3), (12, 5), (14, 6), (15, 7), (30, 13), (1, 1)
Chapter 4
Bernstein’s theory
4.1 Bernstein’s polynomial bf (s)
We take a field K with char(K) = 0 and the polynomial ring K[x] =
K[x1 , . . . , xn ] for some n > 0. The formal differentiation ∂/∂xi in K[x] uniquely
extends to an element, also denoted by ∂/∂xi , of DerK (K(x), K(x)) for 1 ≤ i ≤ n.
We shall denote by Dn or simply by D the K-subalgebra of EndK (K(x)), the K-
algebra of all K-linear transformations in K(x), generated by the multiplication
by xi and ∂/∂xi for all i. Furthermore, by an abuse of notation, we shall write
D = K[x, ∂/∂x]. The 2n generators of D satisfy the following Heisenberg commu-
tation relation:
Lemma 4.1.1 Let E denote a 2n-dimensional vector space over K with a basis
ξ1 , . . . , ξn , η1 , . . . , ηn and T (E) the tensor algebra of E; let I(E) denote the two-
sided ideal of T (E) generated by
ξi ⊗ ξj − ξj ⊗ ξi , ηi ⊗ ηj − ηj ⊗ ηi , ξi ⊗ ηj − ηj ⊗ ξi + δij
45
46 JUN-ICHI IGUSA
At any rate, the set of all such polynomials b(s) and 0 forms an ideal in K0 [s],
which is a principal ideal ring. A monic polynomial b(s) of the smallest degree is
a generator of this ideal, and it is uniquely determined by f (x) and K0 . We shall
denote it by bf (s) and call bf (s) the Bernstein polynomial of f (x). In the special
case where
f (x) = x21 + . . . + x2n ,
hence bf (s) is a factor of (s + 1)(s + n/2). We shall see later that they are actually
equal. This case is classical. In the middle 60’s M. Sato proved a similar statement
for a large class of f (x) in his theory of prehomogeneous vector spaces. We shall
explain a part of this theory in Chapter 6.
P = P α wα , bf (s) = bα (s)wα ,
in which Pα is in K0 [s, x, ∂/∂x] and bα (s) is in K0 [s] for all α. Then we get
Pα · f (x) = bα (s) for all α with bα (s) = 0 for at least one α. Hence the Bernstein
48 JUN-ICHI IGUSA
polynomial bf (s) relative to K0 exists and clearly bf (s) = bf (s)c (s) for some c (s)
in K0 [s]. On the other hand, bα (s) = bf (s)cα (s) for some cα (s) in K0 [s] for all α.
By putting these together we get
cα (s)wα c (s) = 1.
Since c (s) = bf (s)/bf (s) is a monic polynomial, this implies c (s) = 1, hence
bf (s) = bf (s).
(ii) If the Bernstein polynomial of f (x) exists and if f (x ) is obtained from f (x)
by an invertible K0 -linear transformation x → x , then the Bernstein polynomial of
f (x ) also exists, and they are equal.
The proof is as follows: We denote by x, ∂/∂x, etc. the column vectors with xi ,
∂/∂xi , etc. as their i-th entries and put x = g −1 x for any g in GLn (K0 ). Then
∂/∂x = (t g −1 )∂/∂x . Therefore if we put f (x ) = f (x) = f (gx ) and assume that
P (s, x, ∂/∂x) · f (x) = bf (s), then we get
P (s, gx , (t g −1 )∂/∂x ) · f (x ) = bf (s).
Therefore bf (s) exists, and it divides bf (s). The situation is now symmetric and
bf (s) divides bf (s), hence they are equal.
(iii) If c is in K0× and if bf (s) exists, then bcf (s) also exists, and they are equal.
This follows immediately from the fact that each one of xi · ϕ, ∂/∂xi · ϕ for every
ϕ in K[x]f = K[x]cf relative to f (x) and cf (x) are equal.
(iv) If f (x) is in K0× , then bf (s) = 1 while if f (x) is in K0 [x]\K0 and bf (s)
exists, then it is divisible by s + 1.
Since the first part is clear, we shall prove the second part. Suppose that bf (s)
exists for some f (x) in K0 [x]\K0 and that P (s, x, ∂/∂x) · f (x) = bf (s). Then we
will have
P (s, x, ∂/∂x)f (x)s+1 = bf (s)f (x)s
for every s in Z, in particular for s = −1. We observe that for s = −1 the LHS is
in K0 [x] while the RHS is bf (−1)/f (x), and 1/f (x) is not in K0 [x]. This implies
bf (−1) = 0, hence bf (s) is divisible by s + 1.
(v) Suppose that f (x) is in K0 [x]\K0 and there is no point a in Ωn , where Ω is
any algebraically closed extension of K0 , satisfying
∂f ∂f
f (a) = (a) = . . . = (a) = 0.
∂x1 ∂xn
Then bf (s) = s + 1.
This can be proved as follows: By Hilbert’s Nullstellensatz there exist a0 (x),
a1 (x), . . . , an (x) in K0 [x] satisfying
∂f
a0 (x)f (x) + ai (x) = 1.
∂xi
1≤i≤n
Define P as
P = (s + 1)a0 (x) + ai (x) ∂/∂xi .
1≤i≤n
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 49
Then we have P · f (x) = s + 1. Since we have seen in (iv) that bf (s), if it exists, is
divisible by s + 1, we get bf (s) = s + 1.
If we take, e.g., C as K0 and f (x) from C[x]\C, then we see by (iv), (v) that
bf (s)/(s + 1) = 1 implies f −1 (0)sing = ∅. In a certain sense the size of bf (s)/(s + 1)
corresponds to the complexity of f −1 (0)sing . We might recall that to simplify
f −1 (0)sing or, more precisely, to replace f −1 (0) by the union of closed submanifolds
of codimension 1 with normal crossings is one of the main objectives of desingu-
larization. We shall determine bf (s) for an f (x) such that f −1 (0) itself has such a
simple structure. Namely, we shall prove the following statement:
If f (x) = xm1 . . . xn , where mi is in N for every i, then
1 mn
bf (s) = (s + j/mi )
1≤i≤n 1≤j≤mi
then we get
(d/mdx)m · f (x) = b(s),
hence bf (s) divides b(s). On the other hand, if P is in K0 [s, x, d/dx] and P · f (x) =
bf (s), then by expressing P as
P = cij (s)xi (d/mdx)j
i,j≥0
Lemma 4.3.1 If ϕ(s, x) is in K[x]f = K0 (s)[x, 1/f (x)] and P (s, x, ∂/∂x) is in
Dn , then for every r in Z we have
Proof. We first observe that the formula is valid in the special cases where
P (s, x, ∂/∂x) = xi , ∂/∂xi for 1 ≤ i ≤ n. Furthermore if the formula is valid for
P1 , P2 in D = Dn , then it is valid for c1 P1 + c2 P2 for any c1 , c2 in K = K0 (s).
Therefore we have only to show that the formula is valid also for P3 = P1 P2 , and
the proof is as follows: If we put
then
Proposition 4.3.1 Theorem 4.1.1 holds if and only if the D-module K[x]f is
finitely generated.
Proof. We shall, for the sake of simplicity, write f, ϕ(s), P (s) instead of f (x), ϕ(s, x),
P (s, x, ∂/∂x). We then have
D · f −r+1 = D · (f · f −r ) = Df · f −r ⊂ D · f −r ,
D · f −r ⊃ K[x] · f −r = K[x]f −r
for r = 0, 1, 2, . . . , hence
D · 1 ⊂ D · f −1 ⊂ D · f −2 . . . , D · f −r = K[x]f .
r≥0
P (s − 1) · 1 = f −1 , P (s − 2) · f −1 = f −2 , . . . ,
hence K[x]f = D · 1.
In general if K is an arbitrary field, A is a K-algebra, which is associative and
with the unit element, and M is an A-module, then the finite generation problem
of M as an A-module can sometimes be investigated by converting A and M into a
filtered K-algebra and a filtered A-module, and then passing to the corresponding
graded G(A)-module G(M ). In the present case, as we shall see, this method works
perfectly.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 51
We have already used graded algebras and modules to prove the existence of
Hilbert’s functions in Chapter 1.3. If a K-algebra A contains an increasing sequence
of subspaces F0 (A), F1 (A), F2 (A), . . . with A as their union satisfying
Proof. By assumption there exists a finite subset {ψ1 , . . . , ψk } of G(M ) such that
G(M ) = G(A)ψ1 + . . . + G(A)ψk . After expressing each ψi as a finite sum of
homogeneous elements, we may assume that ψi itself is homogeneous, i.e., in Gri (M )
for some ri in N. We shall show that if we choose ϕi from Fri (M ) with ψi as its
image in Gri (M ), then we will have
for all r in N. That will imply M = Aϕi . If we denote the RHS by Fi (M ), then
clearly Fr (M ) ⊃ Fr (M ) for all r. Therefore we have only to show that Fr (M ) ⊂
Fr (M ) also for all r. Since F−1 (0) = 0, this is clear for r = −1. Therefore we shall
apply an induction on r assuming that Fr−1 (M ) ⊂ Fr−1 (M ), hence Fr−1 (M ) =
Fr−1 (M ), for some r ≥ 0. We take ϕ arbitrarily from Fr (M ) and denote its image
in Gr (M ) by ψ. Then we have ψ = b1 ψ1 + . . . + bk ψk for some bi in Gr−ri (A)
for all
i. If we choose ai from Fr−ri (A) with bi as its image in Gr−ri (A), then ϕ − ai ϕi
is in Fr−1 (M ). Therefore ϕ is in
with cij in K for all i, j in Nn . We define Fr (D) by the condition that |i| + |j| ≤ r
for all r. The Heisenberg commutation relation implies that
xi (∂/∂x)j xi (∂/∂x)j = xi+i (∂/∂x)j+j + ci j xi (∂/∂x)j
Then the condition that {Fr (M )} forms an increasing sequence with M as its union
becomes α > 0. We observe that the condition Fi (D)Fj (M ) ⊂ Fi+j (M ) for all i,
j is equivalent to xi · Fr (M ), ∂/∂xi · Fr (M ) ⊂ Fr+1 (M ) for 1 ≤ i ≤ n and for all
r. Since xi · ϕ = xi ϕ, hence deg(xi · ϕ) = deg(ϕ) + 1, the first condition becomes
α ≥ 1. Since deg(∂/∂xi · ϕ) ≤ deg(ϕ) − 1 and ∂/∂xi · ϕ is in f −r−1 K[x] for every
ϕ in f −r K[x], there is no new condition from ∂/∂xi · Fr (M ) ⊂ Fr+1 (M ). We shall
therefore take α = 1. We have thus shown that if we define Fr (M ) as the subspace
of f −r K[x] consisting of all ϕ such that deg(ϕ) ≤ r, then M becomes a filtered
D-module.
We might remark that except when we regard K[x]f as a D-module, the field
K need not be K0 (s). It can be any field with char(K) = 0. Furthermore if D is
replaced by a general filtered K-algebra A, then char(K) need not be 0.
(∗) 0 → M → M → M → 0
0 → Fr (M ) → Fr (M ) → Fr (M ) → 0
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 53
for every r, which is exact as K-modules, i.e., vector spaces over K. In that case
we clearly have
for all r. In the case where A is a graded K-algebra and M , M , M are graded
A-modules, (∗) is called exact as graded A-modules if it gives rise to a sequence
0 → Mr → Mr → Mr → 0
for every r, which is exact as K-modules. The fact is that G is an “exact functor”
in the following sense: If (∗) is exact as filtered A-modules, then the associated
sequence
0 → G(M ) → G(M ) → G(M ) → 0
is exact as graded G(A)-modules. In fact, if we identify Fr (M ) with its image in
Fr (M ), then the kernel of the K-homomorphism from Gr (M ) = Fr (M )/Fr−1 (M )
to Gr (M ) = Fr (M )/Fr−1 (M ), which is clearly surjective, is
for all r. We observe that if A is a filtered K-algebra and M in the exact sequence
(∗) of A-modules is a filtered A-module, and if we put
Fr (M ) = Fr (M ) ∩ M , Fr (M ) = (Fr (M ) + M )/M
for all r, then M , M become filtered A-modules and (∗) is exact as filtered A-
modules.
Suppose now that M is a filtered A-module for a filtered K-algebra A and
for all r in N, hence F gives an (m, 1)-filtration. Also the filtration of the D-module
K[x]f defined in section 4.3 is an (n, (deg(f ) + 1)n )-filtration. At any rate the
following lemma is clear by definition:
Lemma 4.4.1 Let 0 → M → M → M → 0 denote an exact sequence of filtered
A-modules such that the filtrations of M , M are of types (d , e ), (d , e ) respec-
tively. Then the filtration of M is of type (d, e) where d = max(d , d ) and e = e ,
e + e , e according as d > d , d = d , d < d .
54 JUN-ICHI IGUSA
We also need the fact that the converse of Proposition 4.3.2 holds with addi-
tional information. Suppose that A is a filtered K-algebra and M is an A-module
considered as filtered A-modules under two filtrations F , F . We say that F , F are
equivalent if there exist r0 , s0 in N satisfying
Fr (M ) ⊂ Fr+s 0
(M ), Fr (M ) ⊂ Fr+r0 (M )
for all r. This is clearly an equivalence relation. It follows from the definition
that if F is of type (d, e) and F is equivalent to F , then F is also of type (d, e).
We say that a filtration F of M is standard if G(M ) becomes a finitely generated
G(A)-module.
Proposition 4.4.1 If A is a filtered K-algebra and M is a finitely generated A-
module, then M always has a standard filtration which is unique up to equivalence.
Proof. Suppose that M is generated as an A-module by its finite subset {ϕ1 , . . . , ϕk }.
If we regard Ak as an A-module by the prescription
a · (a1 , . . . , ak ) = (aa1 , . . . , aak ),
then the correspondence (a1 , . . . , ak ) → a1 ϕ1 + ... + ak ϕk gives rise to a surjective
A-homomorphism Ak → M with kernel, say N . If we put
Fr (Ak ) = Fr (A)k , Fr (M ) = Fr (A)ϕi , Fr (N ) = Fr (Ak ) ∩ N,
1≤i≤k
with aij , bji in A for 1 ≤ i ≤ k, 1 ≤ j ≤ l. We may assume that aij , bji are all
contained in Fr0 (A) for some r0 in N. If we put
s0 = max(r0 + r1 , . . . , r0 + rl ),
then we will have
Fr (M ) = Fr (A)ϕi ⊂ Fr+r0 (A)φj ⊂ Fr+s 0
(M ),
1≤i≤k 1≤j≤l
Fr (M ) = Fr−rj (A)φj ⊂ Fr+r0 (A)ϕi = Fr+r0 (M )
1≤j≤l 1≤i≤k
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 55
with deg(χ(G(M ), t)) ≤ m for all large r. Furthermore d, e are independent of the
choice of F by Proposition 4.4.1. Therefore we shall denote them by d(M ), e(M ).
The following statement is the main theorem in Bernstein’s theory:
Theorem 4.4.1 If M is any D-module with a (d, e)-filtration where D = Dn , then
necessarily d ≥ n. Furthermore if d = n, then the length of any strictly increasing
sequence of D-submodules of M is at most equal to e. In particular, the D-module
M is finitely generated.
Since the D-module K[x]f has an (n, (deg(f ) + 1)n )-filtration, Theorem 4.4.1
implies that it is a finitely generated D-module, and that implies Theorem 4.1.1 by
Proposition 4.3.1. As for the proof of Theorem 4.4.1 we shall show at this point
only the fact that the first part implies the second part. We, therefore, assume that
d = n for M and take any finite strictly increasing sequence of D-submodules of M :
0 = L0 ⊂ L1 ⊂ . . . ⊂ Lk .
0 = M0 ⊂ M1 ⊂ . . . ⊂ Mk .
We observe that Mi and Mi /Mi−1 are finitely generated D-modules different from
0 for 1 ≤ i ≤ k. Therefore, by the first part and Lemma 4.4.1 we get n ≤
d(Mi /Mi−1 ) ≤ d(Mi ) for 1 ≤ i ≤ k. Since Mi is an A-submodule of M which
has an (n, e)-filtration, we get d(Mi ) ≤ n, hence d(Mi ) = d(Mi /Mi−1 ) = n for
1 ≤ i ≤ n. Again by Lemma 4.4.1 we then get
for 1 < i ≤ k and e(M1 ) > 0 by definition, hence e(Mi ) ≥ i for 1 ≤ i ≤ k. Since
Mk is an A-submodule of M and d(Mk ) = n, we have e(Mk ) ≤ e, hence k ≤ e.
Therefore the first part of Theorem 4.4.1 indeed implies its second part.
algebraically closed. Then θ(p) = β · θ(t − α) is a unit of EndK (M ). Therefore
θ uniquely extends to a K-algebra homomorphism from K(t) to EndK (M ), which
we shall denote also by θ. We choose any ϕ = 0 from M , which is possible because
M = 0, and consider the K-linear map from K(t) to M defined by q → θ(q)ϕ. We
observe that the set {1/(t − α); α ∈ K} is linearly independent over K. Since K
is noncountable and dimK (M ) is at most countable, therefore, the above K-linear
map K(t) → M is not injective. Therefore θ(q)ϕ = 0 for some q = 0; but θ(q) is a
unit of EndK (M ), hence ϕ = 0, a contradiction.
We are ready to prove the first part of Theorem 4.4.1 stating that if M is a
D-module with a (d, e)-filtration, then d ≥ n. Since M = 0, it contains a finitely
generated D-module M = 0 and then d(M ) ≤ d. Therefore we have only to
show that if M = 0 is a fintely generated D-module, then d(M ) ≥ n. We keep in
mind that, since dimK (D) is countable and the D-module M is finitely generated,
dimK (M ) is at most countable. Since d(M ) ≥ 0, the above statement becomes
trivial for n = 0. Therefore we shall apply an induction on n assuming that n >
0. After tensorizing D and M over K by an extension of K, we may assume
that K is algebraically closed and noncountable. We shall denote the K-algebra
homomorphism D → EndK (M ) defining M as a D-module by θ. We shall assume
that d(M ) < n and derive a contradiction.
Put xn = t. Then by Lemma 4.5.2 there exists an element α of K such that
θ(t − α) is not a unit of EndK (M ). We may replace t − α by t and denote the kernel
and the cokernel of θ(t) respectively by M and M . If we put D = Dn−1 , then
θ(t) is a D -homomorphism, hence M and M are D -modules. Since θ(t) is not a
unit of EndK (M ), either M = 0, M = 0 or M = 0. We shall show that either
case will bring a contradiction.
Suppose first that M = 0, hence M = M/tM = 0. Since M is a finitely
generated D-module, it has a standard filtration F by Proposition 4.4.1, and it
gives rise to a standard filtration F of M as its image under M → M , i.e.,
as Fr (M ) = (Fr (M ) + tM )/tM for all r. We observe that Fr (M )/tFr−1 (M ) is
mapped surjectively to Fr (M ) under M → M and θ(t) is injective. Therefore we
get
dimK (Fr (M )) ≤ dimK (Fr (M )) − dimK (Fr−1 (M )) = O(r d(M )−1 ),
valid for all m in N. If now ϕ is in N , then tm ϕ = 0 for some m > 0. This implies
by (∗∗)
tm+1 ((∂/∂t)ϕ) = (∂/∂t)(tm+1 ϕ) − (m + 1)tm ϕ = 0.
Therefore (∂/∂t)ϕ is in N , hence N is stable under θ(∂/∂t). We have thus shown
that N is a D-submodule of M . By our remark in the beginning N is then a finitely
generated D-module, hence d(N ) is defined and d(N ) ≤ d(M ), and hence d(N ) < n
by assumption. Furthermore N = 0 because N contains Ker(θ(t)) = M = 0.
Therefore we can replace M by N , and we will have the situation that every element
ϕ of M satisfies tm ϕ = 0 for some m > 0 depending on ϕ. The rest of the proof is
as follows.
We shall show that Ker(θ(∂/∂t − α)) = 0 for all α in K. Suppose otherwise
and choose ϕ = 0 from the above kernel for some α in K; also choose the smallest
m > 0 satisfying tm ϕ = 0. Then (∂/∂t)ϕ = αϕ implies by (∗∗)
Chapter 5
Lemma 5.1.1 Let I = (−δ, δ) for any δ > 0 denote an open interval in R and θ a
continuous map from I to R with the property that θ(x + y) = θ(x) + θ(y) for all
x, y in I satisfying |x| + |y| < δ and θ(−x) = −θ(x) for all x in I. Then θ(x) = ax
for some a in R and for all x in I. In particular, θ uniquely extends to an element
of Hom(R, R).
59
60 JUN-ICHI IGUSA
We have thus shown that θ(x) = ax for all x in Q ∩ I, hence for all x in I by
continuity.
We shall determine Hom(R, C× × ×
1 ) and Hom(C1 , C1 ) by using Lemma 5.1.1 start-
ing with Hom(R, C1 ). If we restrict the homomorphism e : R → C×
×
1 defined by
x → e(x) = exp(2πix) to (−1/2, 1/2), then it has a unique inverse, say ψ, over
C× ×
1 \{−1}. If χ is an arbitrary element of Hom(R, C1 ), then I = (−δ, δ) for a small
δ > 0 and θ = ψ ◦ χ will have the property in Lemma 5.1.1, hence θ extends to
R and θ(x) = ax for a unique a in R. This implies χ(x) = e(ax) for all x in R.
The converse is obvious and Hom(R, C× 1 ) is isomorphic to R as χ → a. Similarly,
if χ is an element of Hom(C× × ×
1 , C1 ), then χ ◦ e becomes an element of Hom(R, C1 ),
hence χ(e(x)) = e(ax) for a unique a in R and for all x in R. Since χ ◦ e maps Z
to 1, we see that a is in Z, and χ(t) = ta for all t in C× 1 . The converse is obvious
and Hom(C× ×
1 , C1 ) is isomorphic to Z as χ → a. These are well-known examples in
Pontrjagin’s theory. We are ready to prove the following proposition.
Proposition 5.1.1 If K = R or C, then Ω(K × ) = Hom(K × , C× ) consists of all
ω such that
ω(a) = |a|sK (a/|a|)p ,
in which s is in C and p is in Z considered modulo 2 for K = R. Furthermore under
the correspondence ω → (s, p) the group Ω(K × ) is isomorphic to C × (Z/2Z) or
C × Z according as K = R or C.
Proof. If a is an arbitrary element of K × , then under its product expression a =
|a|(a/|a|) we have K × = R× × × ×
+ × K1 . Since Hom(K1 , R+ ) = 1, therefore, we see that
×
Ω(K ) is isomorphic to
Hom(R× × × × × ×
+ , R+ ) × Hom(R+ , C1 ) × Hom(K1 , C1 ).
where K = R, since K1× = {±1}, we take p mod 2. Then in both cases p becomes
unique. We have thus shown that if ω is any element of Ω(K × ), then we can write
with σ(ω) in R. If we express ω as in Proposition 5.1.1, then σ(ω) =Re(s), the real
part of s. If σ is arbitrary in R, then we denote by Ωσ (K × ) the open subset of
Ω(K × ) defined by σ(ω) > σ. We shall sometimes denote by Cσ the open subset of
C defined similarly by Re(s) > σ. In this notation Ω0 (K × ) becomes a union of the
right-half plane C0 .
0 ≤ ϕ < ∞, cϕ = |c| ϕ, ϕ + ϕ ≤ ϕ + ϕ
for all i, j in Nn . The set of such seminorms is countable. We further observe that
Φ∞ = Φ0,0 and Φ∞ = 0 implies Φ = 0. The following lemma is known in
general topology:
We shall outline the proof that E is a metric space, i.e., d(ϕ, ϕ ) = 0 if and only
if ϕ = ϕ , d(ϕ, ϕ ) = d(ϕ , ϕ), and d(ϕ, ϕ ) ≤ d(ϕ, ϕ ) + d(ϕ , ϕ ). The first two
properties are clear while the third property follows from the fact that the function
f (t) = t/(1 + t) for t ≥ 0 is monotone increasing, 0 ≤ f (t) < 1, and
In particular, S(X) is a metric space and the topology in S(X) is invariant under
any invertible R-linear transformation in X. The vector space S(X) with the so-
defined topology is called the Schwartz space of X. If {Φk } is a Cauchy sequence in
S(X), then {(∂/∂x)j Φk } forms a Cauchy sequence relative to the uniform norm for
every j in Nn . Therefore (∂/∂x)j Φk is uniformly convergent to Ψj , say, as k tends
to ∞ and if we put Ψ0 = Ψ, then (∂/∂x)j Ψ = Ψj for every j. This is well known
in calculus; it is proved by using the representation of a differentiable function as
an integral of its derivative and applying an elementary form of Lebesgue’s theorem
reviewed in Chapter 1.1. Furthermore, {Φk i,j } is a Cauchy sequence in R with
Ψi,j as its limit for every i, j. In particular, this shows that Ψ is an element
of S(X), hence S(X) is a complete metric space. We denote the topological dual
of S(X) by S(X) . In other words, S(X) is the subspace of the dual space of
S(X) consisting of its elements which are continuous functions on S(X), i.e., which
convert every null sequence in S(X) into a null sequence in C. An element of
S(X) is called a tempered distribution in X. We recall that a distribution in X was
introduced by L. Schwartz in [51]-1 by using the space D(X) of all C ∞ -functions
on X with compact support instead of S(X) and a tempered distribution in [51]-II
to discuss Fourier transformations.
We shall show, for our later use, the well-known fact that D(X) is dense in S(X).
We might start with A. Cauchy’s historical remark that the R-analytic function
exp(−1/x2 ) on R× completed by the value 0 at x = 0 becomes a C ∞ -function on
R with 0 as its Maclaurin expansion. We replace the above x by (1 − 4(t − a)2 )1/2
and put
θa (t) = exp(−1/(1 − 4(t − a)2 ))
for |t − a| < 1/2 and θa (t) = 0 for |t − a| ≥ 1/2. Then θa becomes a C ∞ -function
on R. Furthermore, if we put I = {0, ±1/2, ±1} and J = I ∪ {±3/2}, then
χ= θa / θb
a∈I b∈J
is a C ∞ -function on R such that χ(t) = 1 for |t| ≤ 1, χ(t) = 0 for |t| ≥ 2 and
χ(−t) = χ(t), 0 ≤ χ(t) ≤ 1 for all t in R. In particular, χ(t) becomes a C ∞ -
function of t2
. In the following lemma and also later r(x) denotes the distance from
0 to x, i.e., ( x2i )1/2 :
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 63
Lemma 5.2.2 Take any Φ from S(X) and put Φk (x) = χ(k−1 r(x))Φ(x) for k =
1, 2, . . . . Then {Φk } gives a sequence in D(X) which tends to Φ as k tends to ∞,
i.e., {Φk − Φ} is a null sequence in S(X).
tends to 0 as k tends to ∞ for all i, j. We observe that the RHS is at most equal to
(j!/j ! j !) (∂/∂x)j (χ(k−1 r(x)) − 1)Ψj (x)∞ ,
in which Ψj (x) = xi (∂/∂x)j Φ(x) and the summation is with respect to j in Nn
such that j = j − j is also in Nn . Since Ψj is in S(X), we have only to show that
This implies that (∗∗) holds also for β, hence (∗∗) holds for all α. Now if |α| = 0,
then
(∂/∂x)α (χ(k−1 r(x)) − 1)Φ(x)∞ ≤ 2rΦ∞ · k−1
and if |α| > 0, then
LHS ≤ φα,j ∞ Φ∞ · k−|α| .
|j|≤|α|
and the RHS tends to 0 as r(x) tends to ∞. Incidentally, the factor π in π · r(x)2
makes the integral of Φ over X to become 1.
We go back to the topological dual S(X) of S(X). In general, let E denote a
vector space over C and E a subspace of the dual space of E; let {Tk } denote any
sequence in E with the property that for every ϕ in E a finite limit
Proof. We take a sequence {Tk } in S(X) with the property that for every Φ in
S(X) a finite limit T (Φ) of {Tk (Φ)} as k → ∞ exists. To be proved is the fact that T
is continuous. We shall assume that T is not continuous and derive a contradiction
in three steps.
If T is not continuous, there exists a null sequence {Φp } in S(X) such that
{T (Φp )} is not a null sequence in C. By replacing {Φp } by a subsequence, we may
assume that |T (Φp )| ≥ c for all p and for some c > 0 independent of p. Since {Φp }
is a null sequence in S(X), we have Φp i,j → 0 as p → ∞ for every i, j in Nn .
Therefore for any given k, if we choose p = p(k) sufficiently large, we will have
for every k in N.
We shall next show that if for every k in N we suitably choose k , k ≥ k
depending on k and put ψk = ϕk , Sk = Tk , then we will have
(∗∗) |Sp (ψk )| ≤ 2p−k (0 ≤ p < k), |Sk (ψk )| > |Sk (ψp )| + k.
0≤p<k
is defined for every ϕ in E and ϕ → S(ϕ) defines an element S of the dual space
of E. We shall show that S is contained in E . We parametrize C by a continuous
function s(t) for 0 ≤ t ≤ 1 and for every k in N we put
S0 = 0, Sk = Tsi (si − si−1 ) (k > 0),
1≤i≤k
in which
1 Ts
ck = ds
2πi |s−s0 |=r (s − s0 )k+1
is in E for every k in Z. We call
Ts = ck (s − s0 )k
k∈Z
Σ = {s ∈ U1 ; Ts ∈ E },
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 67
Lemma 5.3.1 Let (X, dx) denote a measure space, e.g., a nonempty open subset
of Rn equipped with the usual measure, U a nonempty open subset of C, and f a
C-valued measurable function on X × U , e.g., a continuous function if X is an open
subset of Rn , with the following properties: (i) If C is any compact subset of U ,
there exists an integrable function φC ≥ 0 on X satisfying
Theorem 5.3.1 We take f (x) arbitrarily from R[x1 , . . . , xn ]\R, define an open
subset V of X = Rn as V = {x ∈ X; f (x) > 0}, and for any s in the right-half
plane C0 and any Φ in S(X) we put
s
f+ (Φ) = f (x)s Φ(x) dx
V
s
with the understanding that f+ (Φ) = 0 if V = ∅. Then f+
s
becomes an S(X) -valued
holomorphic function on C0 . Furthermore, if we put
bf (s) = (s + λ), γf (s) = Γ(s + λ),
λ λ
s
in which bf (s) is the Bernstein polynomial of f (x), then f+ /γf (s) has a holomorphic
continuation to the whole C.
for every x in X. Then for every (x, s) in V × C we have |f (x)s Φ(x)| ≤ φC (x) and
|f+
s
(Φ)| ≤ φC (x) dx ≤ φC (x) dx
V X
= M1 · Φ∞ + M2 · r dσ0 +n+1 Φ∞ ,
in which
M1 = M1σ0 Ωn /n, M2 = M2σ0 Ωn
s s
are both finite and independent of Φ. In particular, f+ (Φ) is defined and f+ converts
every null sequence in S(X) into a null sequence in C. Furthermore, by applying
Lemma 5.3.1 to this case with V , C0 , f (x)s Φ(x) respectively as X, U , f (x, s), we
see that f+ s
(Φ) is a holomorphic function on C0 . Therefore, f+ s
is an S(X) -valued
holomorphic function on C0 .
As for the second part, for the sake of simplicity, we shall write b(s), γ(s) instead
of bf (s), γf (s). By Bernstein’s theorem, i.e., Theorem 4.1.1, there exists an element
P (s) of R[s, x, ∂/∂x] satisfying
for every x in V . More precisely f (x)s is well defined for every x in V and P (s) is
applied to f (x)s+1 as a differential operator. If we take s from Cσ for a large σ > 0
depending on the order of P (s), then Lemma 5.3.2 becomes repeatedly applicable
and we get
s
b(s)f+ (Φ) = s+1
(P (s)f (x) )Φ(x) dx = f (x)s+1 (P (s)∗ Φ(x)) dx.
V V
therefore, we get
γ(s)−1 f+
s
(Φ) = γ(s + k)−1 · f (x)s+k Φk (x) dx
V
under the assumption that s is in Cσ for a large σ > 0. We observe that the LHS
is holomorphic on C0 while the RHS is holomorphic on C−k . Since k is arbitrary in
N, the above relation implies that f+
s
(Φ)/γ(s) has a holomorphic continuation to C.
s
Therefore we see by Proposition 5.2.1 that f+ /γ(s) has a holomorphic continuation
to C.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 71
The integral does not change even if we replace X by X\f −1 (0), which is the disjoint
union of {x ∈ X; f (x) > 0} and {x ∈ X; −f (x) > 0}. Therefore, if we introduce
s
f− (Φ) as (−f )s+ (Φ), we can write
s
ZΦ (ω) = ω(f )(Φ) = f+ (Φ) + (−1)p f−
s
(Φ).
Theorem 5.3.1 then shows that ZΦ (ω) has a meromorphic continuation to Ω(R× ).
Furthermore, since b−f (s) = bf (s), the poles of ZΦ (ω) on the s-plane are in the
union of −λ − N as −λ runs over the set of zeros of bf (s). The theorem also shows
that the order of a pole of ZΦ (ω) is at most equal to the order of the corresponding
zero of bf (s).
If now f (x) is in C[x1 , . . . , xn ]\C, then V = Cn \f −1 (0) is an analogue of {x ∈
R ; f (x) > 0}. However f (x)s is not well defined on V . With this situation in mind
n
Then
as1 ās2 = |a|sC (a/|a|)p , |a|C = |a|2 = aā
is well defined. We shall apply the above observation to a = f (x) for x in V .
We also make the following observation. If we put uα =Re(xα ), vα =Im(xα ) so
that xα = uα + ivα , x̄α = uα − ivα for 1 ≤ α ≤ n, then Cn becomes R2n under
the correspondence (x1 , . . . , xn ) → (u1 , v1 , . . . , un , vn ). Furthermore, if ϕ is any
differentiable function on a nonempty open subset of Cn , then
for every x in V , the RHS of the above relation is holomorphic on C−k/2 , and this
implies the second part.
Remark. We observe that if we put Q(s) = P̄ (s, x̄, ∂/∂ x̄) for P (s) = P (s, x, ∂/∂x),
then we have
Q(s2 )f (x)s1 f¯(x̄)s2 +1 = b̄f (s2 )f (x)s1 f¯(x̄)s2 ,
in which P̄ , b̄f are as f¯ the images of P , bf under the complex-conjugation applied
to their coefficients. In particular, if bf (s) is written as in Theorem 5.3.1, then
b̄f (s) = (s + λ̄),
λ
and this is the Bernstein polynomial of f¯(x). We define γ̄f similarly. More precisely,
we define γ̄f (s) as the product of Γ(s + λ̄). Then in the notation of the above proof
we have
γ̄f (s2 )−1 f s1 f¯s2 (Φ) = γ̄f (s2 + k)−1 · f (x)s1 f¯(x̄)s2 +k Ψk (x) dx,
V
so that the integral of ρε over X becomes 1. We are using the same notation as
in L. Schwartz [51], I, p. 22. If ϕ is any C-valued continuous function on X with
compact support, then the convolution of ϕ and ρε is defined as
(ϕ ∗ ρε )(x) = ϕ(y)ρε (x − y) dy.
X
∞
We see immediately that ϕ ∗ ρε is a C -function on X which vanishes outside the
ε-neighborhood of Supp(ϕ), i.e., the union of open balls of radius ε centered at all
points of Supp(ϕ). Furthermore we have
for every x in C.
Proof. In Chapter 1.1 we have recalled the following fact with proof: Let X denote
a locally compact space and U a neighborhood of any point a of X. Then there
exists a continuous function ϕa on X with compact support contained in U such
that 0 ≤ ϕa ≤ 1 and ϕa (a) = 1. If now {(U, φU )} is an atlas on X, by replacing
each U by many subsets with compact closure contained in U we may assume that
every Ū is compact and local coordinates are valid on Ū . Then by replacing Uα by
Uα ∩ U we may assume for every α that Ūα is compact and local coordinates are
valid on Ūα . We then construct ϕa for every a in C so that Supp(ϕa ) is contained
in some Uα . Since C is compact, we can find a finite set {ai ; i ∈ I} such that if
we put ϕi = ϕa for a = ai , then for every x in C we will have ϕi (x) > 0 for some i
in I. If we denote by ϕ the sum of all ϕi , then ϕ has a positive minimum η on C.
Furthermore if for every i we choose Uα containing Supp(ϕi ), denote it by Ui , and
replace Ūi by its bicontinuous image in Rn , then
then by using the mean-value theorem in calculus and the Schwarz inequality we
get |Φ(x) − Φ(y)| ≤ M · r(x − y) for every x, y in X. This implies
|(Φ ∗ χε − Φ)(x)| ≤ M · r(x)χε (x) dx = M Ωn ε · r n exp(−πr 2 ) dr,
X r≥0
which tends to 0 as ε → 0.
We shall next show that Φ ∗ χε for any fixed ε > 0 can be approximated by an
element of G(X). We subdivide X into small cubes with vertices in k−1 Zn for a
large k in N. We denote by {δi ; i ∈ Ik } the finite set of cubes which intersect C
and choose yi from C ∩ δi for every i in Ik . We observe that
Sk (x) = (kε)−n · Φ(yi ) exp(−πε−2 · r(x − yi )2 )
i∈Ik
is a Riemann sum for (Φ ∗ χε )(x) and that it is in G(X). We shall show that Sk
tends to Φ ∗ χε as k → ∞. We shall prove, more generally, that if ϕ is in S(X) and
Sk (x) = k−n · Φ(yi )ϕ(x − yi ),
i∈Ik
for every i, j in Nn . Since (∂/∂x) Sk (x) is a Riemann sum for (Φ ∗ (∂/∂x)j ϕ)(x), we
j
may assume as before that j = 0. We may also replace xi by r(x)m where m = |i|.
We shall show that for any given η > 0 we can make |r(x)m (Sk − Φ ∗ ϕ)(x)| < η
for all x in X. If r1 ≥ r0 and r(x) ≥ 2r1 , then by using r(x) ≤ 2 · r(x − y) and
r(x − y) ≥ r1 for Φ(y) = 0 we get
|r(x)m (Sk − Φ ∗ ϕ)(x)| ≤ |r(x)m Sk (x)| + |r(x)m (Φ ∗ ϕ)(x)|
≤ 2m Φ∞ r m+1 ϕ∞ (card(Ik )k−n + µ(C))r1−1 ,
in which µ(C) denotes the total measure of C. Since card(Ik )k−n tends to µ(C) as
k → ∞, the RHS becomes less than η for a large r1 and for all large k. We shall
fix such an r1 and take x from the remaining part, i.e., from the compact ball in
X defined by r(x) ≤ 2r1 . Since Supp(Φ) is also compact, clearly Φ(y)ϕ(x − y) is
equicontinuous in y, i.e., uniformly continuous in y with the uniformity independent
of x. Therefore we will have (2r1 )m |(Sk − Φ ∗ ϕ)(x)| < η for all large k.
Remark. If Φ(x) for every x in X is a holomorphic function of s in a fixed nonempty
open subset V of C, then in Sk (x) in the above proof s appears only in Φ(yi ) for i
in Ik . Therefore Φ can be approximated by
ci (s)(ϕi1 ⊗ · · · ⊗ ϕin )
in which (Nj , nj ) = (NE , nE ) with J bijective to the set of all E containing b and
ε, η are units of the local ring Ob of Y at b. We choose a small neighborhood Ub of
b over which the above local coordinates are valid and ε±1 , η ±1 are all K-analytic.
We apply Lemma 5.4.1 to Y , C, {Ub } instead
of X, C, {Uα }. In that way we get a
partition of unity {pi ; i ∈ I} such that pi (y) = 1 for every y in C. This implies
n −1
ω(f )(Φ) = ω(yj )Nj |yj |Kj (ω(ε)|η|K (Φ ◦ h)pi ).
i∈I j∈J
for yj in K × has a meromorphic continuation to the whole s-plane with its poles
defined by the condition that “Nj s + nj − 1 is in −1 − Nj |p|/2 − N,” i.e., s is in
−|p|/2 − (1/Nj )(nj + N), with the understanding that p = 0 for K = R. Finally,
for a similar reason as before, we can apply Lemma 5.4.2 and its remark to replace
the above Ψ by a tensor product of elements of S(K) without losing any pole. In
that way we get the description of the poles of ω(f ) as stated in the theorem.
5.5 An application
We shall give an application of Theorem 5.4.1 after recalling some basic facts on
Fourier transformations in S(X) and S(X) for X = Rn , n > 0. We reserve the
notation Φ for an arbitrary element of S(X), and we shall not repeat “for all Φ in
S(X)” all the time. We first observe that every C-valued continuous function ϕ on
X with polynomial growth gives rise to an element Tϕ of S(X) as
Tϕ (Φ) = ϕ(x)Φ(x) dx.
X
78 JUN-ICHI IGUSA
This implies that Tϕ converts every null sequence in S(X) into a null sequence in
C. The correspondence ϕ → Tϕ is clearly C-linear and Tϕ = 0 implies ϕ = 0. In
fact if ϕ = 0, then ϕ(a) = 0 for some a in X. Then Φ(x) = ρε (x − a), where ρε is
as in section 5.4, is in D(X) and Tϕ (Φ) = 0 if ε is small. We shall sometimes write
ϕ instead of Tϕ . We shall now define some operations on elements of S(X) in such
a way that they are compatible with the identification of ϕ and Tϕ .
If ϕ is any C ∞ -function on X such that all its derivatives have polynomial
growth, then Φ → ϕΦ gives a C-linear continuous map of S(X) to itself. Therefore
if T is in S(X) , then
(ϕT )(Φ) = T (ϕΦ)
defines an element of ϕT of S(X) . Similarly, if P is any element of C[x, ∂/∂x] and
P ∗ is its adjoint operator, then
(P T )(Φ) = T (P ∗ Φ)
defines an element P T of S(X)
. If now ϕ is any C-valued continuous integrable
function on X and [x, y] = xi yi for every x, y in X with xi , yi as their i-th
coordinates for 1 ≤ i ≤ n, then
(Fϕ)(x) = ϕ(y)e([x, y]) dy,
X
Φ1 ≤ Ωn (n−1 Φ∞ + r n+1 Φ∞ ).
hence
and this implies (ϕ∗ )∗ (x) = ϕ(−x) for every x in X. Since the C-span of the
above ϕ for all λ > 0 and a in X is dense in S(X) by Lemma 5.4.2 and since
F is continuous, we get (Φ∗ )∗ (x) = Φ(−x) for every x in X. In particular, F is
bicontinuous. Furthermore if ϕ is any continuous integrable function on X, then by
using Fubini’s theorem we get
∗
ϕ(x)Φ (x) dx = ϕ(x)Φ(y)e([x, y]) dxdy = ϕ∗ (x)Φ(x) dx.
X X×X X
T ∗ (Φ) = T (Φ∗ ),
then T ∗ is in S(X) and Tϕ∗ = (Tϕ )∗ . We shall formulate its immediate consequence
as a proposition for our later use.
Proposition 5.5.1 If ϕ is a continuous integrable function on X = Rn with inte-
grable Fourier transform ϕ∗ , then the Fourier inversion formula
hence (ϕ∗ )∗ = ψ.
We shall give two examples of T ∗ . Firstly
We might remark that if f (x) is any element of R[x1 , . . . , xn ]\R, then the hypersur-
face f −1 (0) in X is of measure 0. This can be proved, e.g., by using the Weierstrass
preparation theorem. We are ready to explain an elegant proof by M.F. Atiyah [1]
of the following theorem:
Theorem 5.5.1 If f (x) is any element of C[x1 , . . . , xn ]\C, there exists an element
T of S(X) for X = Rn satisfying f T = 1.
Proof. We may assume that f (x) is in R[x1 , . . . , xn ]\R and f ≥ 0. In fact, if
(f f¯)S = 1 for some S in S(X) , then T = f¯S is in S(X) and f T = 1. By
s s
assumption f− = 0, hence ωs (f ) = f+ , which we shall denote by f s . If we put
V = {x ∈ X; f (x) > 0}, then by definition
s
f (Φ) = f (x)s Φ(x) dx
V
fs = ck (s + 1)k
k∈Z
denote its Laurent expansion at −1 with ck in S(X) for all k. Since the poles
of f s are negative rational numbers by Theorem 5.4.1, we see that f s+1 = f f s is
holomorphic at s = −1. Therefore f ck = 0 for all k < 0 and
f s+1 = f c0 + (f ck )(s + 1)k .
k>0
Corollary 5.5.1 If P is any element of C[∂/∂x1 , . . . , ∂/∂xn ]\C, then there exists
an elementary solution for the differential operator P , i.e., an element S of S(X)
for X = Rn satisfying P S = δ0 .
P S = (f T )∗ = 1∗ = δ0 .
can be proved by using Gauss’ theorem. This shows that −1/(n − 2)Ωn r n−2 is
an elementary solution for ∆. The readers can learn the significance and further
examples of an elementary solution in Schwartz [51] and also in Gel’fand and Shilov
[16].
https://doi.org/10.1090/amsip/014/06
Chapter 6
Prehomogeneous vector
spaces
6.1 Sato’s b-function b(s)
We shall explain M. Sato’s theory of prehomogeneous vector spaces. More precisely,
we shall only explain regular prehomogeneous vector spaces up to their b-functions.
We start from the beginning: We say that a group G acts on a nonempty set X
if there exists a map G × X → X denoted by (g, x) → gx satisfying (gg )x =
g(g x), 1x = x for every g, g in G and x in X. In that case for every ξ in X
Gξ = {g ∈ G; gξ = ξ}, Gξ = {gξ; g ∈ G}
are called respectively the fixer of ξ in G and the G-orbit of ξ. We say that the
action of G on X is transitive and also X is a G-homogeneous space if X itself is a
G-orbit. In general, X becomes a disjoint union of G-orbits. We observe that Gξ
is a subgroup of G and the map G → X defined by g → gξ gives rise to a bijection
from the coset space G/Gξ to Gξ.
We shall now consider the case where X = Cn and G is a subgroup of GLn (C)
which is algebraic in the sense that it is the set of all common zeros of some poly-
nomials in the n2 entries of g with coefficients in C. We shall assume that the
closed topological subgroup G of GLn (C) is connected. If we regard elements of X
as column vectors, then G acts on X by matrix-multiplication. If X has a dense
G-orbit, then (G, X) is called a prehomogeneous vector space. If further there ex-
ists an irreducible polynomial f (x) in C[x1 , . . . , xn ] such that X\f −1 (0) becomes a
G-orbit, then the prehomogeneous vector space (G, X) is called regular.
Proof. Since X\f −1 (0) is a G-orbit, every g in G keeps f −1 (0) invariant, i.e.,
gf −1 (0) = f −1 (0). Therefore if, for a moment, we put fg (x) = f (gx) then
83
84 JUN-ICHI IGUSA
fg−1 (0) = f −1 (0). We observe that fg (x) is also irreducible. Therefore by Hilbert’s
Nullstellensatz, we see that fg (x) and f (x) divide each other, hence they differ by
a factor ν(g) in C× necessarily satisfying ν(gg ) = ν(g)ν(g ) for every g, g in G.
Furthermore ν(g) is a polynomial in the entries of g, hence ν is continuous. There-
fore ν is an element of Ω(G). If now F (x) is any relative G-invariant different from
0, then F −1 (0) is G-invariant. If F −1 (0) intersects the G-orbit X\f −1 (0), then
F −1 (0) will contain X\f −1 (0), which is dense in X, hence F (x) = 0. Since this is
not the case, F −1 (0) is contained in f −1 (0). Then by Hilbert’s Nullstellensatz F (x)
divides f (x)e for some positive integer e. Since f (x) is irreducible, this implies that
F (x) is a power of f (x) up to a factor in C× . Finally, write
f (x) = fi (x), fd (x) = 0
i≥d
with fi (x) homogeneous of degree i for all i ≥ d. Then f (gx) = ν(g)f (x) implies
fi (gx) = ν(g)fi (x) for every g in G and for all i. By what we have just shown, there
exists a positive integer e satisfying fd (x) = cf (x)e for some c in C× . By comparing
the degrees of both sides, we get c = e = 1 and fi (x) = 0 for all i > d.
We call f (x) in Proposition 6.1.1 a basic relative invariant of (G, X); it is unique
up to a factor in C× . We observe that if γ is an arbitrary element of GLn (C), then
(γGγ −1 , X) is also a regular prehomogeneous vector space with (γf )(x) = f (γ −1 x)
as its basic relative invariant. We say that (G, X) and (γGγ −1 , X) are equivalent.
This is clearly an equivalence relation. In the special case where γ, hence also γ −1 ,
is contained in the normalizer N (G) of G, i.e., if γGγ −1 = G, then f (γx) is another
basic relative invariant of (G, X), hence it differs from f (x) by a factor in C× .
Therefore, ν extends to N (G) as f (γx) = ν(γ)f (x) for every γ in N (G).
We shall assume from now on that G is reductive. This condition is known to
imply the existence of γ in GLn (C) such that γGγ −1 is invariant under g → t g −1 .
Therefore we shall simply assume that G satisfies this condition, i.e., t G = G. We
keep in mind that γGγ −1 also satisfies this condition if and only if t γγ is in N (G).
Corollary 6.1.1 If f (x) is a basic relative invariant of (G, X), then there exists
an element B(s) of C[s] satisfying
f (∂/∂x) · f (x) = B(s).
If B(s) = 0, then deg(B) ≤ deg (f ) and ν(t g) = ν(g) for every g in G. Furthermore,
B(s) depends, up to a factor in C× , only on the equivalence class of (G, X).
Proof. If deg(f ) = d, then by definition we can write
f (∂/∂x) · f (x) = φ0 (x) + φ1 (x)s + . . . + φd (x)sd
with φi (x) in C[x1 , . . . , xn , 1/f (x)] for 0 ≤ i ≤ d. We shall obtain some information
about the above expression assuming that it is different from 0. If we specialize s
to an element k of N, then we will have
f (∂/∂x)f (x)k+1 = φi (x)ki f (x)k .
0≤i≤d
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 85
in which
(∂/∂x)α xβ = βi (βi − 1) · · · (βi − αi + 1)xβi i −αi
1≤i≤n
Theorem 6.1.1 We have B(s) = b0 b(s), in which b0 > 0 and b(s) is a monic
polynomial of degree d = deg(f ) in R[s].
Proof. If xα is the highest monomial in f (x), then its coefficient is 1 by the above
normalization. Furthermore, xα(k+1) is the highest monomial in f (x)k+1 and
f (∂/∂x)k+1 f (x)k+1 = B(j)
0≤j≤k
hence
B(j) ≥ exp{dk log k(1 + o(1))},
0≤j≤k
and hence
(∗) B(j)/kδk → ∞
0≤j≤k
Γ(s + 1) = sΓ(s)
for Re(s) > 0. Since Γ(1) = 1 by definition, we conclude that Γ(s) has a mero-
morphic continuation to C with a pole of order 1 at every point of −N. These are
well-known elementary properties of Γ(s). Now by Weierstrass
s · (1 + s/n)e−s/n ,
with o(1) uniform in σ as |t| → ∞. Since the product-representation and the above
important asymptotic behavior of Γ(s) are seldom included in the basic course, we
shall give their proofs. We shall follow the presentation by H. Mellin [40] because
the proof by the Mellin transformation seems most appropriate.
Firstly if we express log n as the integral of 1/t from 1 to n and rewrite the n-th
term of the sequence defining C as
1/k − log n + 1/n = 1 − log n − 1/k ,
1≤k<n 1<k≤n
88 JUN-ICHI IGUSA
then we see that both {·} are positive and increase with n. Therefore, the sequence
is between 0 and 1, and monotone decreasing, hence the limit C exists. In the
following log(1 + x) is defined for |x| < 1 by its power series expansion taking
the value 0 at x = 0. We leave it as an exercise to show that if |x| ≤ 1/2, then
| log(1 + x) − x| ≤ |x|2 . Also we keep in mind that Re(log x) = log |x| is well defined
for every x in C× . We now start with
1/G(s) = eCs s · (1 + s/n)e−s/n .
n≥1
for all n ≥ 2R. Therefore, the above infinite product is absolutely and uniformly
convergent for |s| ≤ R, hence it defines an entire function of s with a zero of order
1 at every point of −N. Furthermore, if s is not in −N, then
G(s + 1)/G(s) = se−C · lim n/(s + n + 1) · exp 1/k − log n = s.
n→∞
1≤k≤n
This implies
G(s + 1) = sG(s)
for all s in C. Since 1/sG(s) takes the value 1 at s = 0, we get G(1) = 1. Further-
more,
π/G(s)G(1 − s) = −π/sG(s)G(−s) = πs · (1 − (s/n)2 ),
n≥1
G(s)G(1 − s) = π/ sin(πs).
Since we have
(1 + 1/n)s e−s/n = lim (n + 1)s exp − ( 1/k)s = e−Cs ,
n→∞
n≥1 1≤k≤n
We shall prove the asymptotic formula for G(s). Since G(s) becomes its complex
conjugate under the complex conjugation of s and
by replacing s by it we get
hence
|G(σ + it)/|t|σ G(it)| = |1 + 1/(n + it)|σ /|1 + σ/(n + it)|,
n≥0
and hence
log( |G(σ + it)/|t|σ G(it)| ) = Re σ log(1 + 1/(n + it)) − log(1 + σ/(n + it)) .
n≥0
in which
P (x) = σ · (−1)k (1 − σ k )/(k + 1) · xk−1
k≥1
in which
|RHS| ≤ M · 1/(n2 + t2 ) ≤ M t−2 (1 + π|t|/2),
n≥0
hence
then ψ(y) becomes a continuous integrable function on R with its Fourier transform
ψ ∗ (t) = φ(σ + it) also integrable on R. Therefore by Proposition 5.5.1, we get
(ψ ∗ )∗ (−y) = ψ(y). This implies the above proposition.
Now G(s) is a meromorphic function on C with a pole of order 1 at every point
of −N. Furthermore, G(s + 1) = sG(s), G(1) = 1 imply
for every x > 0 and k in N. The asymptotic formula for G(σ + it) as |t| → ∞
guarantees that G(σ + it) is an integrable function of t if σ is not in −N. We take
0 < σ < 1 and determine
1
ϕ(x) = G(s)x−s ds.
2πi σ+iR
We take n from N and R > 0, and consider the following path of integration:
σ − i ∞ → σ − iR → σ − n − iR → σ − n + iR → σ + iR → σ + i ∞.
The asymptotic formula again shows that the integrals along the two horizontal
paths tend to 0 as R → ∞. Therefore by Cauchy’s theorem we get
1
ϕ(x) = (−x)k /k! + G(s)x−s ds,
2πi σ−n+iR
0≤k<n
in which
G(s)x−s ds ≤ x−σ · |G(σ + it)| dt · xn (k − σ).
σ−n+iR R 1≤k≤n
We have chosen 0 < σ < 1. However, since G(s) is holomorphic on C0 , by the above
process of shifting the vertical line of integration we can replace σ by any positive
real number. Therefore, by Proposition 6.2.1 we get
G(s) = xs e−x d log x = Γ(s)
x>0
Proof. We recall that Z(s) is the LHS of the formula to be proved. If we write
|f (x)|sC = f (x)s f (x̄)s = b(s)−1 f (∂/∂x)f (x)s+1 f (x̄)s
and apply the same argument as in the proof of Theorem 5.3.2, then we get
b(s) Z(s) = (2π)d Z(s + 1)
for Re(s) > 0. Therefore, if we put
C(s) = (2π)ds Z(s) / Γ(s + λ),
λ
then C(s + 1) = C(s) for Re(s) > 0. Since Z(s) is a holomorphic function on C0 ,
we see that C(s) is a periodic holomorphic function on C with period 1. Therefore
C(s) becomes a meromorphic function of z = e(s) on C× and its Laurent expansion
at 0 can be written as
C(s) = ck z k = ck e(ks),
k∈Z k∈Z
in which
ck = exp(2πkt) · C(σ + it) e(−kσ) dσ.
R/Z
This implies
|ck | ≤ exp(2πkt) · (2π)dσ |Z(σ + it)| |Γ(σ + it + λ)| · dσ.
R/Z λ
92 JUN-ICHI IGUSA
then we have
Z(s) = (2π)−ds−n φ(s) Γ(ds + n).
Furthermore,
|φ(s)| ≤ 2 n−1
· |f (u)|σC du.
r(u)=1
for 1 ≤ σ ≤ 2.
If now we incorporate the asymptotic formula
for
Re(s) > 0. If we take the limit as s → 0, then Z(s) tends to 1, hence c0 =
Γ(λ)−1 . Since c0 = 0, we have Γ(λ)−1 = 0 for all λ.
Theorem 6.3.2 If b(s) is the Sato b-function of a regular prehomogeneous vector
space and bf (s) is the Bernstein polynomial of its basic relative invariant f (x), then
b(s) = bf (s). Furthermore, all its zeros are negative rational numbers, hence b(s)
is in Q[s].
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 93
Proof. Since bf (s) divides b(s), after changing the notation in Theorem 6.3.1 we
write
bf (s) = (s + λ), b(s) = bf (s) · (s + λ );
λ λ
also we put
γf (s) = Γ(s + λ), γ (s) = Γ(s + λ ).
λ λ
According to that theorem, Z(s) and γf (s)γ (s) differ by a holomorphic function on
C with no zeros. On the other hand we know by Theorem 5.3.2 that Z(s)/γf (s) is
a holomorphic function on C, hence γ (s) is also a holomorphic function on C. This
implies γ (s) = 1, hence b(s) = bf (s). We know furthermore, by Theorem 5.4.1,
that the poles of Z(s) are negative rational numbers. Therefore all zeros of b(s) are
negative rational numbers.
Actually the zeros of bf (s) are known to be negative rational numbers for an
arbitrary f (x) by M. Kashiwara [33]. We might mention that Theorem 6.3.1, in a
weaker form, is in [27]. We shall explain its counterpart in the real case:
Proof. This can be proved in the same way as Theorem 6.3.1. We have
for some M1 > 0 and σ1 , which are independent of t, as |t| → ∞. Therefore, we get
k = 0, hence C(s) = c0 . Since Z(s) tends to 1 as s → 0, we necessarily
ck = 0 for
have c0 = Γ(λ/2)−1 .
94 JUN-ICHI IGUSA
We shall discuss a classical example from various viewpoints. We start with the
following general remark. If we replace the normalized f (x) by a0 f (x) for any a0
in R× , then we will have
directly, in which X = Mn (R) and tr(y) for any square matrix y denotes the trace
of y. We denote the above Z(s) by Zn (s) and show by an induction on n that
Zn (s) = π −ns/2 · Γ((s + k)/2) Γ(k/2) .
1≤k≤n
If n > 1, we write x = (x1 x ) with x in Mn,n−1 (R) and put x1 = ru, where
r = r(x1 ), so that u is on the unit sphere and dx1 = r n−1 drdu. Then we get
Zn (s) = r s+n−1 exp(−πr 2 ) drdu | det(u x )|sR exp(−π tr(t x x )) dx .
Since the group of rotations in Rn acts transitively on the unit sphere, we can write
u = ge1 , where e1 = t (10 . . . 0), for some g in GLn (R) satisfying t gg = 1. In the
integral {·} above if we replace x by gx , then it becomes Zn−1 (s), hence
in view of the following obvious fact. Namely if I, I are finite subsets of C and
φ(s) is a holomorphic function on C with no zeros satisfying
Γ(s + λ) = φ(s) · Γ(s + λ ),
λ∈I λ ∈I
where X = Mn (C).
We might mention that (∗) follows from Capelli’s identity in invariant theory,
cf., e.g., H. Weyl [60]. It can also be proved, after R. Sasaki, as follows: If for
1 ≤ i1 < . . . < ik ≤ n, 1 ≤ j1 < . . . < jk ≤ n we put
and denote by Xi1 ...ik ,j1 ...jk the coefficient of xi1 j1 . . . xik jk in det(x), then we will
have
Pi1 ...ik ,j1 ...jk · det(x) = (s + i) Xi1 ...ik ,j1 ...jk .
1≤i≤k
Chapter 7
97
98 JUN-ICHI IGUSA
Lemma 7.1.2 Let G denote a locally compact group. Then G is totally discon-
nected if and only if the set of all compact open subgroups of G forms a base at its
unit element 1.
Proof. Since the if-part is clear, we shall again prove the only-if part. We take
any neighborhood U of 1. Then by Lemma 7.1.1 there exists a member A of T1
contained in U . We introduce a subset B of G as
B = {g ∈ G; gA ⊂ A}.
if and only if its intersection with DF is open in DF for all F . Let D(X) denote the
topological dual of D(X) with the so-defined topology. Then, since every C-linear
function on DF is continuous, we see that D(X) coincides with the dual space of
D(X). In particular, D(X) is complete. We call elements of D(X) distributions in
X.
We shall give another description of distributions in X. If T is in D(X) , for
every A in T (X) we put T (A) = T (χA ). Then we get a C-valued simply additive
function T on T (X) in the sense that if A is a necessarily finite disjoint union of
A1 , A2 , . . . in T (X), then T (A) = T (A1 ) + T (A2 ) + . . . . Conversely, suppose that
T is such a function. Express an arbitrary ϕ in D(X) as α1 χA1 + α2 χA2 + . . . with
α1 , α2 , . . . in C, in which A1 , A2 , . . . are disjoint members of T (X). Then we define
T (ϕ) as α1 T (A1 )+α2 T (A2 )+. . . . We shall show that T (ϕ) is well defined. Suppose
that β1 χB1 + β2 χB2 + . . . is a similar expression of ϕ, and put Cij = Ai ∩ Bj for
all i, j. Then we get
ϕ= αi χCij = βj χCij
i,j i,j
Proposition 7.1.1 Let X denote a locally compact totally disconnected space and
express it as a disjoint union of nonempty open and closed subsets Y and F , re-
spectively. Then we have an exact sequence of C-linear maps:
If T0 is any element of D(F ) , its image T0X in D(X) is defined as T0X (ϕ) = T0 (ϕ|F )
for every ϕ in D(X), and if T is any element of D(X) , its image T |Y in D(Y ) is
defined as (T |Y )(ψ) = T (ψ X ) for every ψ in D(Y ).
for every x in C. The proof is quite simple. At every x in C we can find by Lemma
7.1.1 a compact neighborhood Vx of x contained in some Uα . Since C is compact,
it can be covered by a finite number of Vx , say W1 , W2 , . . . . We have only to put
Ai = Wi \(W1 ∪ . . . ∪ Wi−1 )
for i = 1, 2, . . . . We shall prove the existence of O(T ). Define O(T ) as the union of
all open subsets Y of X such that T |Y = 0. We have only to show that T (ϕ) = 0
for every ϕ in D(X) with Supp(ϕ) contained in O(T ). If we apply the partition of
unity to C = Supp(ϕ) and Uα = Y above, then we get
ϕ= ϕχAi , T (ϕ) = T (ϕχAi ) = 0
i∈I i∈I
By definition we have
ϕ(x)µ(x)≤ |ϕ(x)|µ(x) ≤ µ(Supp(ϕ))ϕ∞
X X
because both sides are equal to µ(A)ν(B). If we accept the fact that D(X × Y )
is the C-span of the set of functions of the form χA ⊗ χB , then the above formula
holds for every ϕ in D(X × Y ). Now we have seen that D(X × Y ) is the C-span of
the set of χC for all C in T (X × Y ) and we observe that every such C is a finite
union of its subsets of the form A × B. Therefore we have only to observe further
that any intersection of sets of the form A×B is of that form and that if A1 , . . . , An
are subsets of any set, the characteristic function of their union is the sum of the
characteristic function of the intersection of Ai1 , . . . , Aip with the sign (−1)p−1 for
all i1 < . . . < ip .
Now G continuously acts on G by left multiplication, i.e., as g0 · g = g0 g. A
Haar measure µG or simply µ on G is a G-invariant measure different from 0. We
shall show that it exists and is unique up to a factor in R× + . We observe that if
we have a finite number of members A, B, . . . of T (G), they can be expressed as
disjoint unions of cosets gN by one compact open subgroup N of G. In fact, by
Lemma 7.1.2 and by the compactness of A, B, . . . they can be expressed as finite
unions of gi Ni , where every Ni is a compact open subgroup of G. We can then take
as N any compact open subgroup of G contained in all Ni . After this remark, we
fix a compact open subgroup N0 of G, choose N for A and B = N0 , and put
µ(A) = card(A/N ) card(N0 /N ),
for every A in T (G), i.e., T = T (N0 )µ, with T (N0 ) in C. If T is a Haar measure
on G, then T (N0 ) is in R×+.
If µ is a Haar measure on G and g is in G, then µ (A) = µ(Ag) also gives a Haar
measure on G, hence by the uniqueness we will have
Since
χA (gg0−1 ) = χAg0 (g), χA (g −1 ) = χA−1 (g),
the above properties of µ can also be expressed as
−1
ϕ(gg0 ) µ(g) = ∆G (g0 ) ϕ(g) µ(g),
G G
ϕ(g −1 )∆G (g −1 ) µ(g) = ϕ(g) µ(g)
G G
used gN instead of N g above. At any rate we have shown that G/H is a locally
compact totally disconnected space. We observe that G continuously acts on G/H
as g0 · p(g) = p(g0 g). The action is transitive and p is equivariant.
Since H is a closed subgroup of G, it is also a locally compact totally discon-
nected group. Therefore H has a Haar measure µH . If ϕ is in D(G), then ϕ(gh) as
a function of h is in D(H), hence
π(ϕ)(gH) = ϕ(gh) µH (h)
H
Proof. This can be proved in the same way as the statement in italics on p. 45
in A. Weil [56]. Suppose that T = 0 is in EG/H (ρ) and define S in D(G) as
S(ϕ) = T (π(ϕ)) for every ϕ in D(G). Then S is in EG (ρ) and S1 (ϕ) = S(ρ−1 ϕ)
defines an element S1 of EG (1). Since S1 = 0 by T = 0, we have S1 = cµ for
some c in C× , hence T (π(ϕ)) = cµ(ρϕ). In particular, dimC (EG/H (ρ)) = 1. If
we take h0 arbitrarily from H and replace ϕ(g) by ϕ1 (g) = ϕ(gh−1
0 ), then we get
π(ϕ1 ) = ∆H (h0 )π(ϕ), hence
We have only to show therefore that π(ϕ) = 0 implies µ(ρϕ) = 0. Now π(ϕ) = 0
implies
ϕ(gh−1 )∆H (h−1 )µH (h) = ϕ(gh)µH (h) = 0
H H
for every g in G. If we take any θ from D(G), multiply (ρθ)(g) to the LHS of the
above equation and integrate over G, then by using Fubini’s theorem we get
−1
∆H (h ) (ρθ)(g) ϕ(gh−1 ) µ(g) µH (h) = 0,
H G
in which
(ρθ)(g) ϕ(gh−1 ) µ(g) = ∆G (h) (ρθ)(gh) ϕ(g) µ(g).
G G
Therefore, by using ∆H (h) = ρ(h)∆G (h) and Fubini’s theorem again we get
(ρϕ)(g) θ(gh) µH (h) µ(g) = 0.
G H
If we choose A from T (G) which contains Supp(ϕ), e.g., A = Supp(ϕ), and specialize
θ to any element of D(G) satisfying π(θ) = χAH , then we finally get µ(ρϕ) = 0.
We shall prove another proposition for our later use. As we have remarked
in Chapter 5.1, if G is a compact group, then Hom(G, R× + ) = 1, hence Ω(G) =
Hom(G, C× 1 ). If further G is commutative, then elements of Ω(G) are called char-
acters of G. We keep in mind that if G is a finite abelian group, then G is isomorphic
to a product of cyclic groups, hence Ω(G) and G have the same order.
Proposition 7.2.2 Let G denote a compact totally disconnected abelian group and
µG its Haar measure normalized as µG (G) = 1; define an inner product (ϕ, ϕ ) of
every ϕ, ϕ in D(G) as
(ϕ, ϕ ) = ϕ(g)ψ(g)µG (g),
G
in which ψ(g) is the complex conjugate of ϕ (g). Then Ω(G) forms an orthonormal
basis for D(G), i.e., (χ, χ ) = 1 or 0 according as χ = χ or χ = χ for every χ, χ
in Ω(G) and every ϕ in D(G) can be expressed as a finite sum
ϕ= cχ χ
χ∈Ω(G)
Proof. We shall first prove the orthonormality of Ω(G). Since (χ, χ) = 1 is clear,
we shall show that (χ, χ ) = 0 for χ = χ . Since the complex conjugate of χ(g) is
χ(g)−1 , we have only to show that I = (χ, 1) represents 0 for χ = 1. Now χ = 1
means χ(g0 ) = 1 for some g0 in G. If we replace χ(g) in the integral I by χ(g0 g),
then the new integral is still I because µG is a Haar measure and also equal to χ(g0 )I
because χ(g0 g) = χ(g0 )χ(g), hence I = 0. We shall show that Ω(G) is complete,
106 JUN-ICHI IGUSA
i.e., it spans D(G). Take ϕ arbitrarily from D(G). Since ϕ is locally constant, by
Lemma 7.1.2 for every g in G there exists a compact open subgroup Hg of G such
that ϕ|gHg becomes a constant function on gHg . Since G is compact, it can be
covered by a finite number of gHg . Let H denote any compact open subgroup of G
contained in all such Hg . Then ϕ becomes a C-valued function on the finite abelian
group G/H. We observe that the dimension over C of the vector space of all such
ϕ is equal to the order of G/H, which is equal to the order of Ω(G/H) considered
as a subgroup of Ω(G). The orthonormality implies that elements of Ω(G/H) are
linearly independent over C, hence they form a basis for the above vector space.
Therefore, ϕ can be expressed as a linear combination of elements of Ω(G/H) with
coefficients as stated.
We might mention that in the notation of Proposition 7.2.2 we have the following
Plancherel formula:
|ϕ(g)|2 µG (g) = |cχ |2 .
G χ∈Ω(G)
then
(T0 Gξ)(π(ϕ)) = c · ρ(g)ϕ(g) µG (g)
G
for every ϕ in D(G) with c in C× independent of ϕ.
Proof. In view of Lemma 7.3.1, the above remark, and Proposition 7.2.1, we have
only to show that T0 |Gξ = 0. Suppose that T0 |Gξ = 0 and choose an open subset
U of X satisfying U ∩ F = Gξ. If we put V = U ∪ O(T ), then V is open in X.
Furthermore, if ϕ is in D(X) with Supp(ϕ) contained in V , then
Supp(ϕF ) ⊂ V ∩ F = U ∩ F = Gξ,
hence T (ϕ) = T0X (ϕ) = T0 (ϕ|F ) = 0 by the assumption that T0 |Gξ = 0. Therefore,
T |V = 0, hence V is contained in O(T ), and hence U is contained in O(T ). This
implies that Gξ is empty, which is a contradiction.
We might mention that we have been motivated by A. Weil [58], Chapter 3,
Lemma 16. An example of locally compact totally disconnected spaces which are
not countable at ∞ is R but with discrete topology. We keep in mind that every
locally compact space which is separable, i.e., has a countable base of open sets, is
countable at ∞. We also keep in mind that the condition ρ∆G |H = ∆H will never
be satisfied if ρ|H is not R×
+ -valued.
in which the entry matrices of a1 are 1, 0, 0, a∗ with a∗ = a22 − a21 a12 . We have
tacitly assumed that n > 1. If n = 1, then we simply have g1 = 1. At any rate, by
using elements of GLm (OK ), GLn (OK ) with entry matrices of the form 1, 0, 0, g ∗
with g ∗ respectively in GLm−1 (OK ), GLn−1 (OK ) we can simplify a∗ in the same
way as above, i.e., by an induction on n.
Lemma 7.4.2 We have
n
ρ(d) = µn (dOK ) = card(dOkn /π e dOK n
) card(OK n
/π e dOK n
)
= |π |K = | det(d)|K ,
ei
1≤i≤n
µn (f (a) + π e OK
n
) = µn (π e OK
n
) = µn (a + π e OK
n
),
hence y = f (x) is measure preserving. We know by Corollary 2.2.1 that the inverse
of y = f (x) has a similar form as y = f (x). Therefore, we have only to show
that f (a + π e b) is contained in f (a) + π e OK
n
for every b in OKn
. If we put g(x) =
−e
π (f (a + π x) − f (a)), then the entries of g(x) are all SRP’s in x1 , . . . , xn by
e
n n
(i), hence they are convergent at every b in OK . Therefore, g(b) is in OK , hence
e e n n
f (a + π b) is in f (a) + π OK for every b in OK .
We are ready to prove a change of variable formula in the p-adic case. The
formula is identical to the well-known formula in the archimedean case.
Proposition 7.4.1 If every fi (x) in f (x) = f1 (x), . . . , fn (x) is K-analytic
around some point a of K n and
for 1 ≤ i ≤ n. Therefore, in the general case we may assume that fi (x) is of the
above form with cij in K for all i, j. Since every fi (x) is a convergent power series
by assumption, there exists e0 in N such that cij π e0 |j| tends to 0 as |j| → ∞ for
1 ≤ i ≤ n. Then there exists e1 in N such that
we will have
cij = π kj · π |j|−1 cij ,
in which
kj = (e − e0 − 1)(|j| − 1) − e0 − e1 ≥ (e0 + e1 )(|j| − 2) ≥ 0
for all |j| ≥ 2. Therefore, as we have remarked, the formula is valid for x → g(x) =
(g1 (x), . . . , gn (x)). Since y = f (x) is the composition of x → π −e x, x → g(x), x →
π e x, the formula is valid for y = f (x).
After the above preparation, suppose that we have an n-dimensional p-adic
manifold X defined by an atlas {(U, φU )} and a K-analytic differential form α of
degree n on X; put φU (x) = (x1 , . . . , xn ) for every x in U . Then α|U has an
expression of the form
α(x) = fU (x) dx1 ∧ · · · ∧ dxn ,
in which fU is a K-analytic function on U . If A is any member of T (X) small
enough to be contained in U , then we define its measure µα (A) as
µα (A) = |fU (x)|K µn (φU (x))
A
= q −e · µn φU (fU−1 (π e OK
×
) ∩ A) .
e∈Z
We observe that the above series is convergent because fU (A) is a compact subset
of K, hence a subset of π −e0 OK for some large e0 in N, and then the summation
is restricted as e ≥ −e0 . The point is that if (U , φU ) is another chart and if A is
also contained in U , then we will have the same µα (A) relative to that chart. In
fact, if φU (x) = (x1 , . . . , xn ), then
fU (x) ∂(x1 , . . . , xn )/∂(x1 , . . . , xn ) = fU (x),
µn (φU (x)) = |∂(x1 , . . . , xn )/∂(x1 , . . . , xn )|K µn (φU (x))
for every x in U ∩ U , the first by definition and the second by Proposition 7.4.1.
Therefore, we indeed have
|fU (x)|K µn (φU (x)) = |fU (x)|K µn (φU (x)).
A A
Lemma 7.5.1 Let K denote a p-adic field and X any compact n-dimensional K-
analytic manifold for n > 0. Then X is K-bianalytic to the disjoint union r · OK
n
n
of r copies of OK for some 0 < r < q.
take Y = r · OK
n
and define β on each OK n
as dx1 ∧ · · · ∧ dxn so that µβ = µn in our
notation, then β is a gauge form on Y , hence α = f ∗ (β) is a gauge form on X, and
µα (X) = µβ (r · OK
n
) = r · µn (OK
n
) = r.
If now α, α are any two gauge forms on X, then we can find an atlas {(U, φU )} on
X with φU (x) = (x1 , . . . , xn ) such that
for 1 ≤ i ≤ m. We denote by Jx the jacobian matrix with ∂yi /∂xj = ∂(yi ◦ f )/∂xj
as its (i, j)-entry for 1 ≤ i ≤ m, 1 ≤ j ≤ n. Then f |U is submersive if and only
if the m rows of Jx are linearly independent, i.e., rank(Jx ) = m, for every x in
U . In that case, after a permutation of x1 , . . . , xn and by making U smaller if
necessary, we may assume that the first m columns of Jx are linearly independent.
Then by the implicit function theorem y1 ◦ f, . . . , ym ◦ f , xm+1 , . . . , xn form local
coordinates on X at every point of U . Therefore, we may replace xi by yi ◦ f for
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 115
1 ≤ i ≤ m possibly after making U still smaller. This implies that every submersive
map is open. Furthermore, it gives a simple description of the fiber Xb = f −1 (b)
of f over an arbitrary point b of Y . In fact, if b is not in f (X), then Xb = ∅.
If b = f (a) for some a in X, we may assume that a is in the above U . If we put
ψV (b) = (b1 , . . . , bm ), then Xb ∩U is defined by xi −bi = 0 for 1 ≤ i ≤ m. Therefore,
if we put p = n − m, then Xb becomes a closed p-dimensional submanifold of X.
Now let α denote a K-analytic differential form of degree n on X and β a gauge
form on Y . Then, in the above notation, we can write
α = f ∗ (β) ∧ γ
for b = f (a). We observe that the LHS does not depend on the local expressions of
α, β and the RHS is independent of γ. Therefore, the restriction of γ to Xb depends
only on α, β, hence it gives rise to a K-analytic differential form θb of degree p on
the whole Xb with the above local expression. We take a variable point y of Y and
define θy on Xy as above if Xy = ∅ and θy = 0 otherwise. We shall write
is in D(Y ) and
ϕ(x) µα (x) = Fϕ (y) µβ (y).
X Y
116 JUN-ICHI IGUSA
Chapter 8
Proof. We shall first prove the existence of ψ. If G is a finite abelian group, then
as we have already explained G, G∗ have the same order. Therefore, if H is a
subgroup of G, then the injective homomorphism G∗ /(G/H)∗ → H ∗ defined by
g ∗ → g ∗ |H is surjective because they have the same order. In particular, if we put
Ge = π −e OK /OK , the homomorphism G∗e+1 → G∗e dual to the inclusion Ge → Ge+1
is surjective for every e in N. Therefore, if χ1 = 1 is given in G∗1 , we can find χe
in G∗e satisfying χe+1 |Ge = χe for every e > 0 in N. If now a is arbitrary in K
and ā is its image in K/OK , then ā is contained in Ge for all large e. We observe
that ψ(a) = χe (ā) is independent of e and defines an element ψ of K ∗ satisfying
ψ|OK = 1 and ψ|π −1 OK = 1.
117
118 JUN-ICHI IGUSA
×
for every u in U = OK .
Since e(χ) > 0, therefore, the formula holds for e = 1. Suppose that e > 1. Then
we have
Then FΦ = Φ∗ is in S(X) for every Φ in S(X) and further (Φ∗ )∗ (x) = Φ(−x) for
every x in X. In particular, the Fourier transformation F gives a C-linear bijection
from S(X) to itself.
Proof. Since S(X) is the C-span of χA for all A in T (X), we may assume that
Φ = χA . We may further assume that A = a + π e OK
n
for some a in X and e in Z.
We then have
are called respectively the order and the angular component of a, abbreviated as
×
ord(a) and ac(a). Therefore K × is bicontinuously isomorphic to Z × OK under a →
× × × ∗ ×
(ord(a), ac(a)), hence Ω(K ) is isomorphic to C × (OK ) as ω → (ω(π), ω|OK ). In
× × × ∗
particular, Ω(K ) becomes the disjoint union of countable copies of C with (OK )
× ∗ ×
as its index set. We observe that (OK ) is the union of finite groups (OK /(1 +
π e OK ))∗ for e = 1, 2, . . . .
We define an element ωs of Ω(K × ) for every s in C as
If, for every ω in Ω(K × ), we choose s from C satisfying ω(π) = q −s , then we can
write
ω(a) = ωs (a)χ(ac(a)),
×
in which χ = ω|OK . We keep in mind that the above s is not unique. In fact, the
correspondence s → t = q −s gives a bicontinuous isomorphism
C (2πi/ log q)Z → C× .
in which U = 1 + π e a−1 OK .
×
then Φk is in S(K). Furthermore, if we put U = OK as before, then
I = lim Φk (x) dx = lim ω(x)N |x|n−1
K dx
k→∞ K k→∞ πj U
e≤j<k
= lim (q −n tN )j · χ(u)N du.
k→∞ U
e≤j<k
If χN = 1, since |q −n tN | < 1 in view of |t| < 1, we get the first expression for I. On
the other hand, if χN = 1, by the orthogonality of characters of U we get I = 0.
Suppose next that a is not in π e OK so that U above becomes a subgroup of U .
Then we simply have
I = ω(a)N |a|nK · χ(u)N du,
U
with u(0) in K × and ci in K for all i = 0 in Nn such that the series is convergent on
π k OK
n
for some k in N. Secondly, by making k larger if necessary, we may assume
×
that u(0)−1 u(y) − 1 is contained in π e(χ) OK for all y in π k OK
n
, where χ = ω|OK .
Then we will have
Theorem 8.2.1 Assume that char(K) = 0 and let f (x) denote an arbitrary el-
ement of K[x1 , . . . , xn ]\K; take ω, Φ respectively from Ω0 (K × ), S(X), where
X = K n . Then
ω(f )(Φ) = ω(f (x))Φ(x) dx
X\f −1 (0)
× ∗
becomes holomorphic on the punctured t-plane C× for all χ in (OK ) .
Proof. In the above definition of ω(f )(Φ), since every ω in Ω0 (K × ) has a continuous
extension to K as ω(0) = 0, we could have used X as the domain of integration.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 123
We shall fix χ and Φ, and use Lemma 5.3.1 to prove the first part. Choose any
compact subset C of Ω0 (K × ), put
and define φC as
φC = Φ∞ M σ0 χA ,
in which χA is the characteristic function of A. Then
for every x in X\f −1 (0) and ω in C, and the integral of φC over X\f −1 (0) is finite.
Since ω → ω(f (x)) for every x in X\f −1 (0) is a holomorphic function on Ω(K × ),
hence on Ω0 (K × ), by that lemma ω(f )(Φ) is holomorphic on Ω0 (K × ), and this
implies the first part.
As for the main part, at every point b of Y we can choose a chart (U, φU ) such
that U contains b, φU (y) = (y1 , . . . , yn ) and
N n −1
f ◦h=ε· yj j , h∗ dxk = η · yj j · dyk ,
j∈J 1≤k≤n j∈J 1≤k≤n
in which (Nj , nj ) = (NE , nE ) with J bijective to the set of all E containing b and
ε, η are units of the local ring Ob of Y at b with ε(y), η(y) having expansions similar
to the expansion of u(y) discussed before. Since h is proper and A = Supp(Φ) is
compact open, we see that B = h−1 (A) is in T (Y ). Therefore, we can express B
as a necessarily finite disjoint union of members Bα of T (Y ) such that each Bα is
contained in some U above. Since Φ is locally constant, after subdividing Bα we
may assume that (Φ ◦ h)|Bα = Φ(h(b)) and further that φU (Bα ) = c + π e OK n
for
−1 −1
some c = (c1 , . . . , cn ) in K and e in N. Since h : Y \(f ◦ h) (0) → X\f (0) is
n
instead of ZΦ (ω) and, changing the notation slightly, we shall write Z(s) instead of
Z(ωs ). If σ(ω), Re(s) > 0, therefore, we have
Z(ω) = ω(f (x)) dx, Z(s) = |f (x)|sK dx.
n
OK n
OK
for all i in N. Therefore, if t = ωs (π) = q −s for Re(s) > 0, i.e., |t| < 1, then we get
Z(s) = |f (x)|sK dx = |f (x)|sK dx
n \f −1 (0) ×
OK i≥0 f −1 (π i OK )
= (ci q −ni − ci+1 q −n(i+1) )ti = P (t) − t−1 (P (t) − 1).
i≥0
Then
P (t) = (1 − tZ(s))/(1 − t)
with s and t related as t = q −s for Re(s) > 0. In particular, P (t) is a rational
function of t.
We might mention that the Poincaré series of f (x) was introduced and its ra-
tionality was conjectured in the joint book by S. I. Borewicz and I. R. Šafarevič [5],
which is based on a course given by the second author at the Moscow University.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 125
The function FΦ (g) on G is then defined as the above integral of Φ(x) over the fiber
f −1 (g) by µg (x), and it is continuous.
We shall now make the following specialization: X = K n for a p-adic field K,
G = K, and f : X → K is the K-analytic map defined by an arbitrary f (x) in
K[x1 , . . . , xn ]\K, and we shall examine FΦ , FΦ∗ with all details. As we have shown
in Proposition 8.1.1, the bicharacter ψ(ab) of K × K puts K into a duality with
itself. Therefore FΦ∗ becomes the function on K defined by
FΦ∗ (i∗ ) = Φ(x)ψ(i∗ f (x)) dx
X
in which ψ((i∗ − i∗0 )f (x)) = 1 for every x in Supp(Φ), hence the integral is 0, i.e.,
FΦ∗ (i∗ ) = FΦ∗ (i∗0 ).
n
In the special case where Φ is the characteristic function of OK , we shall write
∗ ∗
F instead of FΦ . If further the coefficients of f (x) are in OK , then we will have
F ∗ |OK = 1. We shall examine F ∗ (i∗ ) for i∗ in K\OK . If we write i∗ = π −e u with
×
e = −ord(i∗ ) > 0 and u in OK , then
F ∗ (i∗ ) = ψ(i∗ f (x)) dx
n
ξ+π e OK
n,
ξ∈OK mod π e
= q −ne · ψ(i∗ f (ξ)).
n,
ξ∈OK mod πe
lim µn (f −1 (π e OK ) ∩ A) = 0.
e→∞
Proof. This lemma can be proved elementarily by using the Weierstrass preparation
theorem. In the case where char(K) = 0, the case in which we shall be interested,
it can also be proved as follows. If we put
×
ce = µn (f −1 (π e OK ) ∩ A), t = q −s ,
with ce = 0 only for a finitely many e < 0. By Theorem 8.2.1 we know that
ZΦ (ωs ) is a rational function of t with poles possibly at 0 and outside the unit disc.
Therefore by Cauchy-Hadamard’s formula for the radius of convergence of a power
series we get
lim sup |ce |1/e < 1,
e→∞
hence 0 ≤ ce ≤ r for some 0 < r < 1 and for all large e. We have only to observe,
e
finally, that
µn (f −1 (π e OK ) ∩ A) = ci ≤ r i = r e /(1 − r)
i≥e i≥e
then these unions become disjoint for all large e. Furthermore, since A is in T (X),
by applying Lemma 8.3.1 to f − i0 instead of f for every i0 in S, we see that
µn (De ) → 0 as e → ∞. Therefore, we get
Φ(x) dx = lim Φ(x) dx = lim FΦ (i) di = FΦ (i) di.
X e→∞ X\De e→∞ K\Ee K\S
We have used Theorem 7.6.1 to see that the two integrals under the limit signs are
equal and the fact that the second limit exists because the first limit exists.
respectively for all i∗ in K and all i in K\S. Furthermore, the integral over π −e OK
becomes independent of e for all large e.
= lim card{ξ ∈ OK
n
, mod π e ; f (ξ) ≡ i mod π e }/q (n−1)e
e→∞
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 129
with the expression under the limit sign independent of e for all large e. Such an
F (i) is called a local singular series.
We shall determine F in the case where f (x) = x1 xm+1 + . . . + xm x2m . We
have seen that F ∗ (i∗ ) = max(1, |i∗ |K )−m for every i∗ in K. Since Cf = {0}, hence
S = {0}, we take i from OK \{0}. Then by using Theorem 8.3.1 we can easily see
that
F (i) = (1 − q −m ) q −(m−1)k .
0≤k≤ord(i)
Lemma 8.4.1 Let m denote a positive integer not divisible by char(K). Then for
any e > ord(m) in N the m-th power map gives a surjection from 1 + π e OK to
1 + mπ e OK .
e(Φ) = m0 + e
will then have the required property, i.e., we shall derive a contradiction assuming
that ZΦ (ω) = 0 for e(χ) > e(Φ). At any rate ZΦ (ω) = 0 implies
n −1
I= ω(ε(y)) ω(yj )Nj |yj |Kj · dy = 0
n
c+π e OK j∈J
for some c.
Suppose first that J = ∅. Then ε(y) = ε(b)(1 + y1 ), hence
I = q −(n−1)e ω(ε(b)) · χ(1 + y1 ) dy1 = 0.
c1 +π e OK
×
Since 1 + c1 is in 1 + π m0 +1 OK , hence in OK , this implies
χ(y1 ) dy1 = 0,
1+π e OK
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 131
hence χ(1 + π e OK ) = 1, i.e., e(χ) ≤ e contradicting e(χ) > e(Φ) ≥ e.
Suppose next that J = ∅. Then we express c + π e OK n
as a disjoint union of
−1
c +π OK observing that e(χ) > e(Φ) ≥ e. Since ε(b) ε(y) is in OK [[y1 , . . . , yn ]],
e(χ) n
we then have
ω(ε(y))(c + π e(χ) OKn
) = ω(ε(b))χ(ε(b)−1 ε(c ))
for every c . Since I = 0, therefore, we get
n −1
ω(yj )Nj |yj |Kj dyj = 0
cj +π e(χ) OK
for some c and j. If cj is in π e(χ) OK , then χNj = 1 by Lemma 8.2.1, hence χNj = 1
on 1 + π m0 +1 OK . Since ord(Nj ) ≤ m0 , we see by Lemma 8.4.1 that χ = 1 on
1 + π 2m0 +1 OK . This implies e(χ) ≤ 2m0 + 1 contradicting e(χ) > e(Φ) ≥ 2m0 + 1.
If cj is not contained in π e(χ) OK , then m0 + 1 ≤ ord(cj ) < e(χ) and χNj = 1 on
1 + π e(χ) (cj )−1 OK by Lemma 8.2.1, hence χNj = 1 on 1 + π e(χ)−m0 −1 OK . Since
e(χ)−m0 −1 > m0 ≥ ord(Nj ), we see by Lemma 8.4.1 that χ = 1 on 1+π e(χ)−1 OK
contradicting the definition of e(χ). We have thus proved the first part.
As for the second part, since the expression of ZΦ (ω) in terms of FΦ (i) has
already been proved, we shall prove the converse. Since FΦ is a locally constant
function on K × , for every e in Z, not the above e, the function u → FΦ (π e u) is in
×
D(OK ). Therefore, by Proposition 7.2.2 we can write
FΦ (π e u) = ce,χ χ(u)
× ∗
χ∈(OK )
Therefore by the first part, ce,χ = 0 for e(χ) > e(Φ) and
(1 − q −1 )−1 q e · Rest=0 (ZΦ (ω)t−e−1 ) χ(u)−1 = ce,χ−1 χ(u)−1 ,
e(χ)≤e(Φ) e(χ)≤e(Φ)
In the following, we shall only use the property of ZΦ (ω) being meromorphic on
Ω(K × ) and examine Condition (A) for a fixed Φ. In doing so, we shall use the
following lemma:
Lemma 8.4.2 Let φ(t) denote a meromorphic function on |t| ≤ r, i.e., on an open
set containing the closed disc, for some r > 0. Then a finite limit
a = lim r e · Rest=0 (φ(t)t−e−1 )
e→∞
exists if and only if φ(t) − b/(1 − r−1 t) is holomorphic on 0 < |t| ≤ r for some b in
C, and in that case a = b.
Proof. We consider the vector space over C of all meromorphic functions φ(t) on
|t| ≤ r and denote the Laurent expansion of each φ(t) at 0 by
φ(t) = c e te , ce = Rest=0 (φ(t)t−e−1 ).
e∈Z
We introduce its subspaces En and Ec respectively with poles at most at 0 and with
r e ce having a finite limit as e → ∞, and put
E = C · 1/(1 − r −1 t) + En .
Then the first part of the lemma can be restated as Ec = E. If φ(t) is in En , then it
is convergent at r, hence {r e ce } is a null sequence. Therefore, if φ(t) − b/(1 − r −1 t)
is in En , then r e ce tends to b as e → ∞. Hence Ec = E also implies the second
part of the lemma. Since we have shown that E is contained in Ec , we have only
to show that Ec is contained in E.
We observe that En , Ec , E are all stable under the multiplication by t, hence
that they are C[t]-modules. If φ(t) is in Ec , then the sequence {r e ce } is bounded,
hence φ(t) is holomorphic on 0 < |t| < r, and hence
φ(t) ≡ cα,m /(1 − α−1 t)m mod En
|α|=r 1≤m≤mα
for some cα,m in C with cα,mα = 0. We shall show that mα ≤ 1 for all α. Suppose
that mα > 1 for some α. Then Ec contains
(1 − α−1 t)mα −2 · (1 − β −1 t)mβ · φ(t) ≡ ce te mod En ,
β=α e∈Z
in which
ce = (cα,mα (e + 1) + cα,mα −1 )α−e
for all e in N. This contradicts the definition of Ec because the sequence {r e ce } is
unbounded. We shall next show that mα = 0 for α = r. Suppose that mα = 1 for
some α = r. Then Ec contains
(1 − β −1 t) · φ(t) ≡ ce te mod En ,
β=α e∈Z
in which ce = cα,1 α−e for all e in N. This again contradicts the definition of Ec
because r e ce has no limit as e → ∞. Therefore φ(t) is in E.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 133
then
|FΦ∗ (i∗ )| ≤ γ(Φ) · max(1, |i∗ |K )−σ+ε
for all i∗ in K, in which γ(Φ) > 0 is independent of i∗ .
Proof. We have seen in section 8.1 that the Fourier transform of an integrable
function is continuous. Therefore (2) implies (1) by Theorem 8.3.1. In the notation
of the proof of the second part of Theorem 8.4.1 we see that (1) implies
−1 −1
lim ce,χ = (1 − q ) · lim FΦ (π e u) χ(u)−1 du = δ1 (χ) FΦ (0),
e→∞ e→∞ ×
OK
term by any r satisfying 0 < r < q σ , then we get a bounded sequence, i.e., we have
ce,χ−1 = δ1 (χ)b + O((qr −1 )e )
as e → ∞. On the other hand, by using Theorem 8.3.1 and Corollary 8.1.1 we get
FΦ∗ (π −e u) = (1 − q −1 ) ck,1 q −k + q −e · ce−e(χ),χ−1 q e(χ) g(χ) χ(u)
k≥e χ
×
for every e in Z and u in OK , in which the summation in χ is finite. Since e(1) = 1
−1
and g(1) = −q , the contribution from the above series and the term for χ = 1
is O(r −e ) while the contribution from each term for χ = 1 is also O(r −e ) both as
e → ∞. If ε > 0, then we can take r = q σ−ε . Therefore, we get
∗
FΦ∗ (i∗ ) = O(r ord(i ) ) = O(|i∗ |−σ+ε
K )
as |i∗ |K → ∞.
134 JUN-ICHI IGUSA
is its Laurent expansion at α = (π), then all ck are in S(X) with ck = 0 for
k < −n by Theorem 8.2.1 and with Supp(ck ) contained in f −1 (0) for k < 0 by
what we have just shown. We shall denote the order of the pole α by m = mα and
put T = c−m . Then 0 < m ≤ n, T = 0 and Supp(T ) is contained in f −1 (0). We
shall obtain more precise information about T in the case where f (x) is a relative
invariant.
We recall that GLn (K) is a locally compact totally disconnected group. In fact,
the compact open subgroups 1n + π e Mn (OK ) of GLn (K) for all e > 0 in N form a
base at its unit element 1n . Furthermore, if we denote the n × n diagonal matrix
with d1 , . . . , dn as its diagonal entries by diag{d1 , . . . , dn }, then we see by Lemma
7.4.1 that
GLn (K) = ∪ GLn (OK ) diag{π e1 , . . . , π en }GLn (OK ),
in which the union is taken for all increasing sequences (e1 , . . . , en ) in Z. Since all
double cosets above are compact subsets of GLn (K), it is countable at ∞. Conse-
quently, if G is any closed subgroup of GLn (K), then it is also a locally compact
totally disconnected group which is countable at ∞. We observe that any subgroup
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 135
× ×
for ω close to , i.e., ω|OK = |OK and t close to α. Since
g · T = ρ (g)−1 T
In the above theorem it is, of course, enough to assume that the number of
G-orbits in f −1 (0) is countable. The theorem can be applied to the case where f (x)
is a basic relative invariant of a regular prehomogeneous vector space. We might
mention that the theorem was proved in that case under the assumption that the
number of G-orbits in f −1 (0) is finite. In fact, our proof in [25] is basically the same
as the one we have just given. We might remark that the countability assumption
on the number of G-orbits is very strong. However, it has been reported by A.
Gyoja [18] that he succeeded in removing that restriction.
https://doi.org/10.1090/amsip/014/09
Chapter 9
Some homogeneous
polynomials
9.1 Quadratic forms and Witt’s theorem
We shall make ourselves familiar with those homogeneous polynomials for which we
shall later compute Z(s). We shall start with quadratic forms. We take an arbitrary
field K and consider vector spaces over K all assumed to be finite dimensional. If
X is such a vector space, as before we shall denote its dual space by X ∗ and put
[x, x∗ ] = x∗ (x) for every (x, x∗ ) in X × X ∗ . A quadratic form Q on X is a K-valued
function on X satisfying the following conditions:
(Q1) Q(λx) = λ2 Q(x) for every λ in K and x in X.
(Q2) Q(x, y) = Q(x + y) − Q(x) − Q(y) is K-bilinear on X × X.
It follows from the definition that Q(x, y) = Q(y, x), Q(x, x) = 2Q(x), and Q|Y for
any subspace Y of X is a quadratic form on Y . The value of Q on any K-linear
combination λ1 x1 + . . . + λn xn of x1 , . . . , xn in X can be determined by the formula
Q λ i xi = Q(xi )λ2i + Q(xi , xj )λi λj ,
1≤i≤n 1≤i≤n 1≤i<j≤n
137
138 JUN-ICHI IGUSA
such that the first p elements form a K-basis for Y and identify
X with K n as x = λi wi → λ = t (λ1 · · · λn ). We denote by h the p × n matrix
with hij = Q(wi , wj ) as its (i, j)-entry for 1 ≤ i ≤ p, 1 ≤ j ≤ n and by h0 the
p × p submatrix of h composed of its first p columns. Then Q|Y is nondegenerate
if and only if det(h0 ) = 0 and x is in Y ⊥ if and only if hλ = 0. Since rank(h) =
p = dimK (Y ) and n = dimK (X), therefore we get (∗). We observe that, in the case
where Q|Y is nondegenerate, if Q is nondegenerate (resp. reduced), then Q|Y ⊥ is
nondegenerate (resp. reduced). We also remark that if Q|Y and Q|Y ⊥ are both
nondegenerate, then (Y ⊥ )⊥ = Y and the relation of Y and Y ⊥ is symmetric.
We shall prove Witt’s theorem in [61], which is fundamental in the theory of
quadratic forms. We shall start with three lemmas.
Lemma 9.1.1 Let Y , Y denote subspaces of a vector space X both different from
X. Then their union is also different from X, i.e., strictly smaller than X.
Proof. Suppose that X is the union of Y and Y . Then, by replacing Y , Y by
larger subspaces if necessary, we may assume that they are both of codimension 1
in X. We can then write X = Y + Ka for any a in X\Y . Also we can express Y
as the set of all x satisfying [x, a∗ ] = 0 for some a∗ = 0 in X ∗ . Since X is the union
of Y and Y , we see that y + λa for every y in Y and λ = 0 in K is in Y , hence
[y + λa, a∗ ] = [y, a∗ ] + [a, a∗ ]λ = 0.
We can take y = 0, and we get [a, a∗ ] = 0. Then [y + λa, a∗ ] = 0 for every y in Y
and λ in K, hence X is contained in Y . This is a contradiction.
Lemma 9.1.2 Suppose that Q is a reduced quadratic form on X and Q(a) = 0 for
some a = 0 in X. Then Q(a, b) = 1 and Q(b) = 0 for some b in X. Furthermore,
if we put X = a, b
⊥ , then Q|a, b
is nondegenerate, hence Q|X is reduced and
X becomes the orthogonal direct sum
X = a, b
⊕ X .
Proof. Since Q is reduced and Q(a) = 0, a = 0, we see that a is not in X ⊥ . There-
fore, the K-linear map x → Q(a, x) from X to K is surjective, hence Q(a, b0 ) = 1
for some b0 in X. If we put b = b0 − Q(b0 )a, then Q(a, b) = 1 and Q(b) = 0.
Futhermore, if λ, µ, λ , µ are in K, then
Q(λa + µb, λ a + µ b) = λµ + µλ ,
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 139
Proof. We shall only assume, until the last moment, that Q is reduced. Since b is in
Y ⊥ and gY = Y , we see that gb is in Y ⊥ . Therefore, if we put c = gb − Q(a, gb)b,
then c is in Y ⊥ and Q(a, c) = 0, hence c is in X ⊥ . Since Q(c) = Q(gb) = 0 and Q
is reduced, we get c = 0, hence gb = Q(a, gb)b with τ0 = Q(a, gb) = 0. If x is in X ,
we can write gx = (lx)b + σx with K-linear maps l : X → K and σ : X → X . If
σc = 0 for some c in X , then gx = 0 for x = −τ0−1 (lc )b + c , hence x = 0, and
hence c = 0. Therefore, σ is an injection, hence a bijection, and the condition of g
being an isometry simply becomes σ in O(Q|X ).
We now take λ0 , µ0 from K and x0 from X , put g ∗ a = λ0 a + µ0 b + x0 , and
require that g ∗ acts on Y as g. Then the condition on the K-linear map g ∗ : X → X
so defined to give an element of O(Q) becomes λ0 = 0 and
for every λ in K and z in Z. This is equivalent to Q(a) = Q(b) and a−b in Z ⊥ . If now
H = (a−b)⊥ , then H is contained in Y ∪Y . Otherwise H contains c which is not in
Y ∪Y . Then we can easily verify that the prescription g1 : λa+µc+z → λb+µc+z
for every λ, µ in K and z in Z defines an isometry g1 from Y + Kc to Y + Kc
with the property that g1 |Y = g0 and g1 |Z1 = idZ1 for Z1 = Z + Kc, hence
Z1 = Z becomes a member of Σ. This contradicts the assumption. Therefore, H is
contained in Y ∪ Y , hence H becomes the union of H ∩ Y and H ∩ Y . This implies
by Lemma 9.1.1 that H is contained either in Y or in Y . Since the codimension of
H in X is at most 1 for any Q, this implies that either H = Y or H = Y , hence
either Q(a, a − b) = 0 or Q(b, a − b) = 0. Since Q(a) = Q(b), we then get both
Q(a, a − b) = 0 and Q(b, a − b) = 0, hence H = Y = Y . If we put a1 = a − b, then
a1 = 0 and Q(a1 ) = 0. Therefore by Lemma 9.1.2, we will have X = a1 , b1
⊕ X
for some b1 satisfying Q(a1 , b1 ) = 1 and Q(b1 ) = 0 with X necessarily contained
in Y . Then g0 becomes an isometry from Y = Ka1 + X to itself. Since Q is
nondegenerate by assumption, we see by Lemma 9.1.3 that g0 uniquely extends to
an element of O(Q).
Theorem 9.1.2 If Q is a reduced quadratic form on a vector space X over an
arbitrary field K, then X contains 2p elements a1 , . . . , ap , b1 , . . . , bp satisfying
Q(ai ) = Q(bi ) = 0, Q(ai , bi ) = 1 for 1 ≤ i ≤ p such that it becomes an orthogonal
direct sum
X = a1 , b1
⊕ · · · ⊕ ap , bp
⊕ X0
with Q|X0 anisotropic. Furthermore, if Q is nondegenerate, then up to an isometry
the anisotropic kernel X0 does not depend on the choice of such a Witt decomposition
of X.
Proof. If Q is reduced but not anisotropic, then by Lemma 9.1.2 we will have
X = a, b
⊕ X , in which Q(a) = Q(b) = 0, Q(a, b) = 1 and Q|X is reduced. If
Q|X is not anisotropic, we can apply Lemma 9.1.2 to Q|X . In that way we get a
decomposition of X as stated. If
some computation to prepare for section 9.3. As a side result, we shall show that
the anisotropic kernel X0 in Theorem 9.1.2 is unique up to an isometry. Actually,
although we have not included its proof in this book, the above uniqueness holds
for any K. At any rate, we shall start with some general observations without
assuming that K is finite.
In the case where char(K) = 2, there is a classical diagonalization of an arbi-
trary quadratic form Q on X stating that X has a K-basis {w1 , . . . , wn }, hence
dimK (X) = n, such that
Q xi w i = Q(wi )x2i
1≤i≤n 1≤i≤n
but if q is even, there is only one case and card(Q−1 (i)) = q 2m for every i.
Proof. If dimK (X) = 2m, then by Lemmas 9.2.1, 9.2.2 the anisotropic kernel has the
two possibilities stated in the theorem. If we can prove the formula for card(Q−1 (0)),
then it will show that the two cases are independent of the decomposition of X in
Theorem 9.1.2. Furthermore, if L = K(ξ) and ξ 2 − aξ + b = 0 as before, then with
respect to the K-basis {1, ξ} for L we have d(N ) = a2 − 4b, which is (ξ − ξ q )2 .
Therefore if q is odd, then d(N ) is not in (K × )2 , hence χ(d(N )) = −1, and hence
χ(Q) = χ(d(Q)). On the other hand, if dimK (X) = 2m + 1, then by Lemma 9.2.2
the anisotropic kernel is of the form Kw for some w in X and Q(w)(K × )2 clearly
determines its isometry class. Since χ(2d(Q)) = χ(Q(w)) if q is odd, what remains
to be proved is only the formula for card(Q−1 (i)).
First of all, in the notation of Theorem 9.1.2 we have
Q (xi ai + xi bi ) + x0 = t x x + Q(x0 )
1≤i≤p
for every x = t (x1 . . . xp ), x = t (x1 . . . xp ) in K p and x0 in X0 . If now Q(x) =
t
x x for x , x in K m and i is in K × , then Q(x) = i implies x = 0. The number
of such x is q m − 1 and for each x = 0 the number of x satisfying t x x − i = 0
is q m−1 , hence
If Q(x) = t x x + N (ξ) for x , x in K m−1 and ξ in L, then card(Q−1 (0)) is the
sum of the numbers of solutions of N (ξ) = −t x x = 0 and = i in K × . Since
card(N −1 (0)) = 1 and card(N −1 (i)) = q + 1 for i in K × , we therefore get
As for card(Q−1 (i)) for i in K × , if N (α) = i, then Q−1 (1) is mapped bijectively to
Q−1 (i) under (x , x , ξ) → (ix , x , αξ). Therefore, card(Q−1 (i)) is independent of
i, hence (q − 1)card(Q−1 (i)) + card(Q−1 (0)) = q 2m , and hence
In the above argument we have tacitly assumed that m > 1, but the formulas are
valid also for m = 1.
If Q(x) = t x x + ax20 for x , x in K m , x0 in K, and a is in K × , we add the
numbers of solutions in x , x of t x x = −ax20 for all x0 in K. In that way we get
As for card(Q−1 (i)) for i in K × , we shall first assume that q is odd. We introduce
a quadratic form Q1 (x, x1 ) = t x x + ax20 − ix21 , in which x1 is a new variable, and
compute card(Q−1 1 (0)) directly and also as the sum of the numbers of solutions in
x of Q1 (x, x1 ) = 0 for all x1 in K. Since χ(d(Q1 )) = χ(ai), we then get
card(Q−1
1 (0)) = q
2m+1
+ χ(ai)(q m+1 − q m ) = q 2m + (q − 1)card(Q−1 (i)).
This implies
card(Q−1 (i)) = q 2m + χ(ai)q m ,
in which χ(ai) = χ(2d(Q)i). We shall next assume that q is even. Then the square
map gives an automorphism of K. Therefore, if we write a = b2 , i = j 2 with b in
K × , j in K, then −ax20 +i = (bx0 +j)2 . Since x0 → bx0 +j gives a bijection from K
to itself, if we put Q0 (x) = t x x +x20 , then card(Q−1 (i)) =card(Q−1
0 (0)). Therefore,
we get card(Q−1 (i)) = q 2m by the above result or from q·card(Q−1 (i)) = q 2m+1 for
all i in K.
[i] = 1 − q −i , [i]+ = 1 + q −i
for every i in N.
We observe that G = GLn (K) acts transitively on S = K n \{0} by matrix-
multiplication. If we express any element g of GLn (K) by its 1 × 1, 1 × (n − 1),
(n − 1) × 1, (n − 1) × (n − 1) entry matrices g11 , g12 , g21 , g22 , then the fixer H
of e1 = t (1 0 . . . 0) in G is defined by g11 = 1, g21 = 0 necessarily with g22 in
GLn−1 (K). Therefore, if K = Fq , since card(S) = q n [n], we get
We have tacitly assumed that n > 1, but if n = 1, then GL1 (K) = K × , hence
card(GL1 (K)) = q[1]. Therefore, by an induction on n we get
2
card(GLn (K)) = q n · [i].
1≤i≤n
146 JUN-ICHI IGUSA
We next take a reduced quadratic form Q on a vector space X over K and define
the orthogonal group O(Q) of Q as in section 9.1. We observe that if g : X → X
is any K-linear transformation satisfying Q(gx) = Q(x) for every x in X, then g
is necessarily an injection, hence a bijection. In fact, if gx = 0 for some x in X,
then Q(x, y) = Q(gx, gy) = 0 for every y in X and Q(x) = Q(gx) = 0, hence x is
in X ⊥ ∩ Q−1 (0), and hence x = 0 because Q is reduced. Therefore O(Q) consists
of all such g.
After this remark, we shall consider the special case where X = K 2m and
Q(x) = x2i−1 x2i = x1 x2 + x3 x4 + . . . ,
1≤i≤m
hence
Still in the case where X = K 2m with K arbitrary for a moment, if there exists
a separable quadratic extension L of K generated by a zero ξ of t2 − at + b for a, b
in K, we put
Q(x) = x2i−1 x2i + (x22m−1 + ax2m−1 x2m + bx22m )
1≤i<m
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 147
and write O2m (K) instead of O(Q). If we take g = (w1 w2 . . . w2m ) from M2m (K),
then the condition for g to be in G = O2m (K) becomes Q(wi ) = 0 except for
Q(w2m−1 ) = 1, Q(w2m ) = b, Q(w2i−1 , w2i ) = 1 except for Q(w2m−1 , w2m ) = a,
and Q(wi , wj ) = 0 for all other i < j. If K = Fq , hence L = Fq2 , and if m > 1,
then exactly in the same way as in the previous case but by using the formula in
Theorem 9.2.1 for card(Q−1 (0)) for the above Q this time we get
In the case where m = 1, if we put w1 = t (α, γ), w2 = t (β, δ), and N = NL/K , then
the above condition becomes N (α+γξ) = 1, N (β +δξ) = b, and N (α+γξ, β +δξ) =
a. Therefore, η = (β+δξ)(α+γξ)−1 is also a zero of t2 −at+b, hence η = ξ or η = ξ q .
The subgroup of O2 (K) defined by η = ξ is isomorphic to Ker(N ) as g → α + γξ
and O2 (K) has a coset by that subgroup represented by g with α = 1, β = a,
γ = 0, δ = −1. In particular, card(O2 (K)) = 2q[1]+ . Therefore, we get
card(O2m (K)) = 2q m(2m−1) [m]+ · [2i].
1≤i<m
for some a in K × . Since O(Q) does not change even if we multiply any element of
K × to Q, we may further assume that a = 1. We shall write O2m+1 (K) instead of
O(Q). Then if K = Fq , by the same argument as before, we get
The above computation shows that if q is even in K = Fq , then Sp2m (K) and
O2m+1 (K) have the same order, hence there exists a bijection from Sp2m (K) to
O2m+1 (K). Actually, they are isomorphic for any perfect field K with char(K) = 2.
The proof of this remarkable fact is simple and it is as follows. We change our
notation and put
Q(x) = x20 + t x x
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 149
hence n(1) = 1 for otherwise n(1) = 0. But then n(a) = n(a1) = n(a)n(1) = 0,
hence n(a, b) = 0 for all a, b in C, and hence C = C ⊥ ∩ n−1 (0) = 0, a contradiction.
We polarize (1) in b, i.e., we replace b by b1 + b2 , then by b1 , b2 and subtract the
second from the first. In that way we get
then 1 = 1 and n(a ) = n(a), hence by polarization n(a , b ) = n(a, b). Furthermore,
by replacing a1 , a2 , b1 , b2 respectively by a, 1, b, c and also by a, c, b, 1 we get
We rewrite the first polarized form of (1) as n(ab, ac) = n(a)n(b, c). Since n(ab, ac) =
n(a (ab), c) by (2), if we put d = a (ab)−n(a)b, then n(d, c) = 0 for all c in C. By us-
ing (1), (2), and n(a ) = n(a) we also have n(d) = 0. Hence d is in C ⊥ ∩ n−1 (0) = 0,
i.e.,
we have used (2), n(a , b ) = n(a, b), and (2) again. We similarly have
t(ab) = t(ba).
(ab)a2 = a(ba2 )
a ◦ b = (1/2)(ab + ba),
152 JUN-ICHI IGUSA
A = Ke1 + Ke2 + X,
(α1 e1 + α2 e2 + x)(β1 e1 + β2 e2 + y) = γ1 e1 + γ2 e2 + z,
If now C is any composition algebra and α1 , α2 , x are the (1, 1), (2, 2), (1, 2)
entries of a in H2 (C) then a → α1 e1 + α2 e2 + x gives an isomorphism from H2 (C)
to A = Ke1 + Ke2 + C where Q(x) = n(x). Therefore, H2 (C) is a Jordan algebra.
We shall show that H3 (C) is also a Jordan algebra.
We take a, b, c from H3 (C) and keep in mind that aij = aji , etc.; we put a◦a◦a =
a ◦ (a ◦ a), e = 13 . We introduce the following quadratic and cubic forms on A:
Since t(aij ajk aki ) is invariant under even permutations of aij , ajk , aki and is equal
to t((aij ajk aki ) ) = t(aik akj aji ), we see that t(aij ajk aki ) = t(a23 a31 a12 ) for all
distinct i, j, k. Since
Q(a, b) = aii bii + n(aij , bij ),
i i<j
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 153
hence
We shall use this fact very often. If we replace b above by a and a ◦ a, then we get
(a ◦ a)ii = a2ii + n(aij ) + n(aik ), (a ◦ a)ij = (aii + ajj )aij + aik akj ;
(a ◦ a ◦ a)ii = a3ii + (2aii + ajj )n(aij ) + (2aii + akk )n(aik ) + t(a23 a31 a12 ),
(a ◦ a ◦ a)ij = (a2ii + aii ajj + a2jj + n(a23 ) + n(a31 ) + n(a12 ))aij + Q(a, e)aik akj .
By using these we can easily verify that a satisfies the following basic cubic equation
in which
Q1 (a) = (1/2)Q(a, e)2 − Q(a).
If we take a variable t and define the first polar det1 (a, b) of det(a) as
Similarly, by taking the first polar of the above cubic equation for a we get
in which Q1 (a, b) = Q(a, e)Q(b, e) − Q(a, b). If we denote the LHS of the above
equation by L(a, b), then we get a ◦ L(a, b) − L(a, a ◦ b) = 0 because both terms are
0. By using the cubic equation for a and the above expression for Q1 (a, b), we can
rewrite this equation as
If we replace det1 (a, b) by Q(a# , b) with a# as above and use the fact that a ◦ a# =
det(a)e, then we get
We have thus shown that H3 (C) is a Jordan algebra for any composition algebra
C. If we assume that char(K) = 3, in addition to char(K) = 2, then Q|Ke is
nondegenerate, hence A = Ke ⊕ (Ke)⊥ . The proof can be slightly shortened
by using this fact. At any rate, the fact is that if K is algebraically closed and
char(K) = 2, then every simple Jordan algebra over K is isomorphic to Hn (C) for
some n ≥ 1 for the C in (C1), (C2), (C3), Ke1 + Ke2 + X with a nondegenerate
quadratic form Q on X, or H3 (C) for the C in (C4). We again refer the details to
Jacobson [32]. We might mention, for a comparison, that if A is a simple associative
algebra over any algebraically closed field K, then by a special case of Wedderburn’s
theorem A is isomorphic to Mn (K) for some n ≥ 1.
then by Cayley’s theorem we will have φ(a) = 0. We might recall its proof: If we
put (te − a)# = φ(t)(te − a)−1 and write
φ(t) = αi ti , (te − a)# = bi ti
0≤i≤n 0≤i≤n
then Hn (C) consists of a = (x, t x) for all x in Mn (K). Furthermore, we see that a
satisfies the same equation in Hn (C) as x in Mn (K). Therefore, we can take the
determinant of x as det(a). If C is as in (C3), hence C = M2 (K), then Hn (C)
consists of all a in M2n (K) satisfying Jn t aJn−1 = a, i.e., such that x = aJn is in
Alt2n (K), in which Jn and Alt2n (K) are as in section 9.3. If x is in Alt2n (K), i.e.
an alternating matrix in M2n (K), we define its Pfaffian Pf(x) as
Pf(x) = ε(σ)xσ(1),σ(2) · · · xσ(2n−1),σ(2n) ,
σ
det(a) = a11 a22 a33 + t(a23 a31 a12 ) − a11 n(a23 ) − a22 n(a31 ) − a33 n(a12 ).
We shall obtain a classical expression for this cubic form. If we write an arbitrary
element a of M3 (C) as a = (a1 , a2 ) with a1 , a2 in M6 (K), then a is in H3 (C) if and
only if a1 is in H3 (M2 (K)) and the entries of a2 satisfy (a2 )ii = 0 for 1 ≤ i ≤ 3,
(a2 )ij = −(a2 )ji for 1 ≤ i < j ≤ 3. Therefore, by what we have shown above
y = a1 J3 is in Alt6 (K). We shall denote the (i, j)-entry of y by yij ; also we put
z11 z12 . . . z16
(a2 )23 (a2 )31 (a2 )12 =
z21 z22 . . . z26
and wij = z1i z2j − z1j z2i for 1 ≤ i, j ≤ 6. Then in det(a) those terms which are
free from zij will give Pf(y). Furthermore, we can easily verify that
det(a) − Pf(y) = − yij wij = −t z1 yz2 ,
i<j
for dimK (C) = 1, 2, 4, 8. In the notation of section 9.4 we define the Freudenthal
quartic on X after H. Freudenthal [14] as
in which y, z are in Alt8 (K). Actually, the above cubic form Pf(y) − t z1 yz2 and
J(y, z) are in E. Cartan’s thesis of 1894 but with some incorrectness about J(y, z),
which was corrected as above by Freudenthal. Furthermore, he and T. A. Springer
developed the theory of Jordan algebras in connection with exceptional simple
groups. At any rate, if C = M2 (K) + M2 (K), then there exists a K-linear map
X → Alt8 (K)2 under which f (x) becomes −4J(y, z). We might give one such map.
We write a = (a1 , a2 ), b = (b1 , b2 ) as before and define y , z in Alt6 (K) with their
(i, j)-entries yij , zij for 1 ≤ i, j ≤ 6 as y = a1 J3 , z = J3 b1 . We next define y, z in
Alt8 (K) with their additional (i, j)-entries yij , zij for i, j = 7, 8 as
y17 y27 . . . y67
= J1 (b2 )23 (b2 )31 (b2 )12 J3 , y78 = b0 ,
y18 y28 . . . y68
z17 z27 ... z67
= (a2 )23 (a2 )31 (a2 )12 , z78 = a0 .
z18 z28 ... z68
Then, after some lengthy computation, we can see that f (x) = −4J(y, z). Since we
shall not use this fact, the details will not be given.
In order to satisfactorily examine Freudenthal quartics in the four cases, we need
to prove a large number of formulas in the theory of Jordan algebras. Instead, we
shall explain only one case where such a preparation will not be necessary. In fact,
we shall use the method we have used in [22], pp. 1021-1023 which depends only
on the usual matrix computation. p n
We know that the action of GLn (K) on K n extends to K for all p. In
particular, for p = 3 we have
(g · x)ijk = gii gjj gkk xi j k
for every g in GLn (K) with gii as its (i, i )-entry and for every i, j, k, in which the
summation is for all i , j , k . We observe that if x = 0, hence xijk = 0 for some i <
j < k and if g is the permutation-matrix representing (1i)(2j)(3k), then (g · x)123 =
3
xijk = 0. We shall assume, from now on, that n = 6, hence dimK ( K n ) = 20,
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 157
and denote the 3 × 3 entry matrices of g at (1, 1), (1, 2), (2, 1), (2, 2) respectively
by α, β, γ, δ. We observe that if x123 = 0 in x, then there exists a unique g with
α = δ = 13 , β = 0, such that (g · x)ijk = 0 for all i < j ≤ 3 < k and the unique γ is
given by
−x123 γ = (x23i x31i x12i )i=4,5,6 ;
we shall denote the RHS by a. In other words, a is the element of M3 (K) with
(x23i x31i x12i ) as its (i − 3)-th row for i = 4, 5, 6. Similarly, if x456 = 0, there
exists a unique g with α = δ = 13 , γ = 0 such that (g ·x)ijk = 0 for all i ≤ 3 < j < k
and the unique β is given by
then by using the above expressions for a0 , b0 , b and the general formulas
3
Proposition 9.5.1 Let K denote any field with char(K) = 2, identify K 6 with
K 2 + M3 (K)2 under the K-linear bijection x → (a0 , b0 ; a, b), in which
f (g · x) = det(g)2 f (x);
if f (x) = 0 and θ = f (x)1/2 , then there exists an element g of SL6 (K(θ)) satisfying
g · x = (1, θ; 0, 0); and if θ is in L× for any field L, the fixer of (1, θ; 0, 0) in SL6 (L)
consists of all g with its entry matrices α, β, γ, δ satisfying det(α) = det(δ) = 1,
β = γ = 0.
Proof. In order to prove the first part, i.e., f (g · x) = det(g)2 f (x), we can replace K
by its algebraic closure, and we shall use the principle of the irrelevance of algebraic
inequalities in Chapter 1.1. We take t = (g, x) regarding the 56 entries of g, x as
variables and put
i.e., F (t ) = 0. Therefore, by the above principle F (t) = 0, hence F (t ) = 0 for all
t in K 56 .
We shall now go back to the original notation. Suppose that f (x) = 0 and
θ = f (x)1/2 . Then for some permutation matrix g0 we will have (g0 · x)123 = 0.
Since f (g0 · x) = det(g0 )2 f (x) = f (x), the second part follows from what we have
shown earlier. As for the third part, the proof of Lemma 9.5.2 shows that if g is
in the fixer of (1, θ; 0, 0) in GL6 (L) and if det(α) = 0, then det(α) = det(δ) = 1
and β = γ = 0 while if det(α) = 0, then det(g) = −1. Therefore, if g is in SL6 (L),
then we only have the first possibility. Furthermore, every such g is in the fixer of
(1, θ; 0, 0).
160 JUN-ICHI IGUSA
with the usual understanding that Fm,0 (a, t) = 1; we extend the definition by
Fm,n (a, t) = 0 for n < 0. If we put
hence Fm,n (a) = Fn,m (a), and Fi,j (a)Fi+j,k (a) = Fi,j+k (a)Fj,k (a). We also observe
that
and further Fm,n (a, t) = Fm−k,n (a, ak t) for all m, n, k in Z. In this notation the
Gauss identity can be written as
(G0) Fi,j (a)ai(i−1)/2 ti = (1 + ai−1 t)
i+j=n 1≤i≤n
for every n ≥ 0. We might also mention that (∗) for t = 1 is tacitly used in Gauss’
proof, in fact as follows: The coefficient of ti in (1 + an t)Θ(n) is
By (G0) we have
(1 − ai+k t) = Fi,j (a) ai(i−1)/2 (−ak+1 t)i .
1≤i≤n−k i+j=n−k
Since the other product is Fk,n−k (a), the LHS of (G1) becomes
(−1)i Fk,n−k (a) Fi,j (a) ai(i+1)/2+(i+k)k ti+k .
i+j+k=n
If we put l = i + k, hence j = n − l, by using Fk,n−k (a) Fi,j (a) = Fk,i (a) Fi+k,j (a),
the above expression can be rewritten as
Fk,i (a) ak(k−1)/2 (−1)k · (−1)l Fl,n−l (a) al(l+1)/2 tl .
0≤l≤n k+i=l
If we put Dm,n (a, t) = RHS - LHS, then Dm,n (a, t) = 0 for n ≤ 0. Therefore, we
shall assume that n > 0 and prove Dm,n (a, t) = 0 by an induction on n. If we apply
(∗) to Fm,n (a, t) and Fk,n−k (a, t), then we get
2
Dm,n (a, t) = Fm,n−1 (a, t) − Fm−k,k (a) Fk,n−k−1 (a, t) ak tk
0≤k≤n
Fm−k,k (a) Fk−1,n−k (a, t) ak (a−1 t)k .
2
+ an Fm−1,n (a, t) −
0≤k≤n
We observe that in the first {·} the term for k = n is 0, hence {·} is equal
to Dm,n−1 (a, t), which is 0 by induction. On the other hand, if we replace
Fm−1,n (a, t) and Fk−1,n−k (a, t) in the second {·} respectively by Fm,n (a, a−1 t) and
Fk,n−k (a, a−1 t), then it becomes Dm,n (a, a−1 t). Therefore, we get Dm,n (a, t) =
an Dm,n (a, a−1 t). Since a, t are variables and Dm,n (a, t) is a polynomial in t, this
implies
Dm,n (a, t) = Am,n (a)tn
162 JUN-ICHI IGUSA
for some Am,n (a) which is independent of t. If we can show that Am,n (a) = 0, then
we will have Dm,n (a, t) = 0. Since Am,n (a) is the coefficient of tn in Dm,n (a, t), by
going back to its definition we get
amn − (−1)n (1 − ai ) · a−n(n+1)/2 Am,n (a)
1≤i≤n
= (−1)k Fm−k,k (a) · (1 − an−k+i ) · ak(k−1)/2 .
0≤k≤n 1≤i≤k
Therefore, we have only to show that the above RHS is equal to amn . By definition
we have
Fm−k,k (a) (1 − an−k+i ) = Fn−k,k (a) (1 − am−k+i )
1≤i≤k 1≤i≤k
Since Fl,n−l (a) Fi,j (a) = Fi,n−i (a) Fj,l (a), it can be rewritten as
Fi,n−i (a) ami Fj,l (a) aj(j−1)/2 (−1)j .
0≤i≤n j+l=n−i
As in the proof of (G1), we see by (G0) that {·} above represents 1 or 0 according
as n − i = 0 or n − i > 0. Therefore, the above expression is indeed equal to amn .
Finally, we shall show that
Fm−k,k (a) Fk,n−k (a, t) ak −k tk = Fm−1,n (a, t) + tFm,n−1 (a, t).
2
(G3)
0≤k≤n
0≤k≤n
2
+ Fm−k−1,k (a) Fk,n−k (a, t) ak tk .
0≤k≤n
By (G2) the second sum is equal to Fm−1,n (a, t). We observe that in the first sum
the term for k = 0 is 0. If we replace k by k + 1, the sum becomes
2
t· Fm−k−1,k (a) Fk+1,n−k−1 (a, t) ak (at)k ,
0≤k<n
in which Fk+1,n−k−1 (a, t) = Fk,n−k−1 (a, at). Therefore, again by (G2), it is equal
to t · Fm−1,n−1 (a, at) = t · Fm,n−1 (a, t).
https://doi.org/10.1090/amsip/014/10
Chapter 10
Computation of Z(s)
10.1 Z(ω) in some simple cases
We recall that
Z(ω) = ω(f (x)) dx,
X◦
in which X ◦ = OK n
, dx = µn is the Haar measure on X = K n normalized as
µn (X ) = 1, and f (x) is in K[x1 , . . . , xn ]\K. As for ω, it is in Ω0 (K × ), i.e.,
◦
t = ω(π) satisfies 0 < |t| < 1. We have denoted Z(ωs ) by Z(s). This is a p-adic
analogue of
|f (x)|s exp(−π t xx) dx.
Rn
There is a remarkable difference between them about the explicit computability for
a given f (x). In the real case Z(s) is hardly computable; while in the p-adic case it
is a rational function of t and has been computed for a large number of f (x). The
significance of an ever-increasing list of explicitly computed Z(s) is that it allows us
to formulate conjectures and proceed to their proofs. At any rate, in this chapter
we shall explain some Z(s) in the above-mentioned list. We shall start with the
simplest cases after the following general remarks.
×
Suppose that OK acts on X as (u, x) → u · x keeping µn and X ◦ invariant;
suppose further that f (u · x) = um f (x) for some m > 0 in N. Then by replacing x
by u · x in the integral defining Z(ω) we get
Z(ω) = χ(u)m Z(ω)
× ×
for every u in OK , in which χ = ω|OK . Therefore, Z(ω) = 0 unless χm = 1.
◦
We remark that X \{0} is the disjoint union of π e Un for all e in N, where Un =
×
X ◦ \πX ◦ , hence U1 = OK and µn (Un ) = [n] in the notation of Chapter 9.3. We put
[i, j] = 1 − q −i tj , [i, j]+ = 1 + q i tj
so that [i] = [i, 0], [i]+ = [i, 0]+ for every i, j in N.
Suppose now that f (x) = det(x) for x in X = Mn (K), hence dimK (X) = n2 .
If we define u · x as the multiplication by u to the first column x1 of x, then we
will have det(u · x) = u det(x), hence Z(ω) = 0 unless χ = 1. If we denote Z(s) by
Zn (s), then by the same method as in Chapter 6.3, we can show that
Zn (s) = [k]/[k, 1].
1≤k≤n
163
164 JUN-ICHI IGUSA
In fact, as a special case of Lemma 8.2.1, we get Z1 (s) = [1]/[1, 1]. If n > 1, we
write x = (x1 x ) and apply the second remark above to the domain of integration
by dx1 . In that way, we get
Zn (s) = (q −n t)i dx1 | det(x1 x )|sK dx ,
i≥0 Un (X )◦
In fact, since Pf(x) = x12 for n = 1, we get Z1 (s) = [1]/[1, 1]. In the case where
n > 1, the first row of x in X ◦ is of the form π i (0 t x1 ) for some i in N and x1 in
U2n−1 , hence x1 = g e1 for some g in SL2n−1 (OK ). If we denote by g the element of
SL2n (OK ) with 1, 0, 0, g as its entry matrices and if x in X = Alt2n−2 (K) is the
right-bottom entry matrix of g −1 x t g −1 , then x is in (X )◦ and Pf(x) = π i Pf(x ).
Since d(π i x1 ) = q −(2n−1)i dx1 , therefore, we get
Zn (s) = (q −(2n−1) t)i dx1 |Pf(x )|sK dx
i≥0 U2n−1 (X )◦
If in the above {·} we write x = ge1 for some g in GLm (OK ) and replace x by
t −1
g x , then it becomes [1]/[1, 1], hence
f (x) = t x x + π e x0
for any e in N so that X = K 2m+1 . If we denote Z(s) by ϕ(e) in this case, then
clearly ϕ(0) = [1]/[1, 1]. If e > 0, then we write
ϕ(e) = dx |t x x + π e x0 |sK dx dx0 ,
m m+1
OK OK
m m
and split OK above into Um and πOK . Then, by the repeatedly used argument, we
get
ϕ(e) = [m] · [1]/[1, 1] + q −m t · ϕ(e − 1).
This is a recursion formula by which we can express ϕ(e) in terms of ϕ(0). In that
way we get
|t x x + π e x0 |sK dx = [1] [1, 1][m, 1] · [m] + [0, 1]q −m (q −m t)e
X◦
for every e in N. We might remark that Z(s) in the third case becomes the limit of
the above expression as e → ∞.
As the fourth simple case we take x from X = M2m,2n (K) where m ≥ n and
in which Jm is as in Chapter 9.3 the element of Alt2m (Z) with 1 as its (2i − 1, 2i)-
entry for 1 ≤ i ≤ m and 0 as its (i, j)-entry for all other i < j. If we define u · x
as the multiplication by u to the first column of x, then f (u · x) = uf (x), hence
Z(ω) = 0 unless χ = 1. If we denote Z(s) by Zm,n (s), then we have
Zm,n (s) = [2(m − k + 1)][2k − 1] [2(m − k + 1), 1][2k − 1, 1].
1≤k≤n
The proof is by now a familiar argument. We write the first column of x as π i x1 for
some i in N and x1 in U2m . If we denote by Sp2m (OK ) the intersection of Sp2m (K)
and GL2m (OK ), then it consists of all g = (w1 w2 . . . w2m ) in M2n (OK ) satisfying
t
w2i−1 Jm w2i = 1 for 1 ≤ i ≤ m and t wi Jm wj = 0 for all other i < j. Therefore by
166 JUN-ICHI IGUSA
the same argument as in Chapter 9.3 we see that every x1 in U2m can be written
as x1 = ge1 for some g in Sp2m (OK ). If we put X = M2m,2n−1 (K), then similarly
as before we get
−2m i
Zm,n (s) = (q t) dx1 |f (x1 x )|sK dx
i≥0 U2m (X )◦
= [2m]/[2m, 1] · |f (e1 x )|sK dx .
(X )◦
If in the above {·} we write y2 = g ∗ e∗1 for some g ∗ in GL2n−1 (OK ), where e∗1 is
the e1 in K 2n−1 , and replace y by g ∗ y, then f (e1 t (y1 π i y2 y )) will be replaced by
det(g ∗ )f (e1 t (y1 π i e∗1 y )). Therefore, we get
I = [2n − 1] · q −(2n−1)i |f (e1 t (y1 π i e∗1 y ))|sK dy.
i≥0 Y◦
We shall denote the above integral over Y ◦ by Ii . The point is that if we put
X = M2m−2,2n−2 (K), x = (e1 t (y1 π i e∗1 y )), and denote by x in (X )◦ the
right-bottom entry matrix of x, then we have
with the understanding that Zm−1,n−1 (s) = 1 for n = 1. This implies the above-
stated expression for Zm,n (s). We shall see later in this chapter that if we replace
Jm above by an element h of Sym2m (Z) with 1 as its (2i − 1, 2i)-entry for 1 ≤ i ≤ m
and 0 as its (i, j)-entry for all other i < j, and accordingly Pf(t xJm x) by det(t xhx),
then the corresponding Zm,n (s) will have an entirely different expression.
Finally, as the fifth simple case we take x = (y, (z1 z2 )), where y is in Y =
Alt6 (K) and (z1 z2 ) is in M6,2 (K) hence dimK (X) = 27, and
We shall use capital letters for the spaces of variables expressed by small letters
such as X, x and Y , y above. We write Z(s) as an integral with respect to dy dz1
followed by an integration by dz2 . If we split (Z2 )◦ \{0} into π i U6 for all i in N,
write z2 in U6 as z2 = ge1 for some g in SL6 (OK ) and replace y, z1 respectively by
t −1
g yg −1 , gz1 , then we get
Z(s) = [6] · q −6i |Pf(y) + π i t y1 w|sK dydw,
i≥0 Y ◦ ×W ◦
If f¯(ā) = 0, then f (a+πx) ≡ f (a) ≡ 0 mod π for all x in X ◦ , hence the contribution
of all such ā in Ē is q −n (card(Ē) − N ). If f¯(ā) = 0 and (∂ f¯/∂xi )(ā) = 0 for some
i, then we use Lemma 7.4.3. If we put
We clearly have ϕ(0) = [1]/[1, 1]. In the case where i > 0, if we apply SPF to ϕ(i),
we get −n i
q t · ϕ(0) 0<i≤d
ϕ(i) = [n, d]Z(s) +
−n d
q t · ϕ(i − d) i ≥ d,
in which Z(s) is for f (x). If we write i = dk + i0 , where 0 ≤ i0 < d, then by an
induction on j we get
ϕ(i) = [nj, dj]Z(s) + (q −n td )j · ϕ(d(k − j) + i0 )
for 0 ≤ j ≤ k, hence
[n(k + 1), d(k + 1)]Z(s) + (q −n td )k · q −n ti0 · [1]/[1, 1] i0 > 0
ϕ(i) =
[nk, dk]Z(s) + (q −n td )k · [1]/[1, 1] i0 = 0.
In particular, if we replace f (x) by a quadratic form Q(x), then we will have
[n(i + 1)/2, i + 1]Z(s) + (q −n/2 t)i−1 · q −n t · [1]/[1, 1] i odd
ϕ(i) =
[ni/2, i]Z(s) + (q −n/2 t)i · [1]/[1, 1] i even.
170 JUN-ICHI IGUSA
In other words if Q̄ is hyperbolic and only in that case, the above two expressions
become equal, and
|Q(x) + π e y|sK dxdy = [1] [1, 1][n/2, 1] · [n/2] + [0, 1]q −n/2 (q −n/2 t)e
X ◦ ×OK
X ◦ = (OK a + OK b) ⊕ L.
we may assume that they are in OK but not all in πOK . If we denote by J the set
of all i for which c̄i = 0 so that
f¯(x) = c̄i xdi i = 0
i∈J
and by NJ the number of zeros of f¯(x) in Ē = Fnq , then NJ has been expressed in
terms of Jacobi sums by A. Weil [57]. Furthermore, S̄ is defined by xi = 0 for all i
in J. Therefore, by SPF we get
Z(s) = |f (x)|sK dx = RJ + q −card(J) te · |f1 (x)|sK dx,
X◦ X◦
in which
RJ = (1 − q −n NJ ) + (q −n NJ − q −card(J) )[1]t/[1, 1].
As for e and f1 (x), we replace xi in f (x) by πxi for all i in J and write the new
f (x) as π e f1 (x) with f1 (x) satisfying the same condition as f (x). We can then
apply the same argument to f1 (x). By repeating this process, we get a sequence
f0 (x) = f (x), f1 (x), f2 (x), . . . . If we write
fj (x) = cji xdi i ,
1≤i≤n
for all j in N. The point is that, in view of ord(cji ) ≤ max(ord(ci ), di ) for all i, j,
the sequence Z0 (s) = Z(s), Z1 (s), Z2 (s), . . . becomes periodic, i.e. Zj (s) = Zj (s)
for some j < j . Then [α, β]Zj (s) for some α, β > 0 in N becomes known, hence
all Zk (s), in particular Z(s), will be known. In the following, we take c1 = c2 =
. . . = cn = 1, {d1 , d2 , . . . , dn } = {2, 3}, {2, 3, 5} and make the corresponding Z(s)
explicit.
If f (x) = x21 + x32 , then NJ = q not only for J = {1}, {2} but also for J = {1, 2}
because the plane curve f −1 {0} is parametrized as x1 = u3 , x2 = −u2 . Therefore,
we get RJ = [1][j, 1]/[1, 1] for j = card(J). If we denote RJ by Rj for j = card(J),
then we get the following sequence:
This implies
Z(s) = |x21 + x32 |sK dx
X◦
= [1]/[1, 1][5, 6] · (1 − q −2 t(1 − t) − q −5 t5 ).
172 JUN-ICHI IGUSA
If f (x) = x21 + x32 + x53 , then NJ = q 2 for all J. This is clear for card(J) = 1 and also
for J = {1, 2} by what we have shown. If x21 + x53 = 0, x3 = 0, and x1 = ux3 , then
u2 + x33 = 0, hence NJ = q 2 for J = {1, 3}. If x32 + x53 = 0, x3 = 0, and x2 = ux3 ,
then u3 + x23 = 0, hence NJ = q 2 for J = {2, 3}. Finally, if x21 + x32 + x53 = 0, x3 = 0,
and x1 = u1 x33 , x2 = u2 x23 , then (u21 + u32 )x3 + 1 = 0, hence
NJ = q(from x3 = 0) + (q 2 − q)(from x3 = 0) = q 2
also for J = {1, 2, 3}. Therefore, we get RJ = [1][j, 1]/[1, 1] for j = card(J). If we
denote RJ by Rj for j = card(J), then we get the following sequence:
(0, 0, 0) = R3 + q −3 t2 (0, 1, 3), (0, 1, 3) = R1 + q −1 t(1, 0, 2),
ga = ḡā.
×
Proof. We first observe that ν(g) for every g in G(OK ) is in OK . In fact, if g is in
G(OK ), hence in Mn (OK ), since f (x) is in OK [x1 , . . . , xn ], we see that f (gx) is also
in OK [x1 , . . . , xn ]. Therefore, if ci xi is any term of f (x), then f (gx) = ν(g)f (x)
×
implies that ν(g)ci is in OK for every i. Since ci is in OK for at least one i, we see
that ν(g) itself is in OK . Since ν(g)ν(g ) = 1 and g is in G(OK ), hence ν(g −1 )
−1 −1
×
is in OK , we see that ν(g) is in OK .
◦
Secondly, if a, b are in X , g is in G(OK ), and b̄ = ḡā, then the integrals of
|f (x)|sK over a + πX ◦ and b + πX ◦ are equal. In fact, since ḡā = ga and G(OK )
keeps X ◦ invariant, we have b + πX ◦ = ga + πX ◦ = g(a + πX ◦ ). Furthermore since
the action of G(OK ) on X ◦ is measure preserving, we get
|f (x)|sK dx = |f (x)|sK dx = |f (x)|sK dx.
b+πX ◦ g(a+πX ◦ ) a+πX ◦
We have tacitly used the fact that |f (gx)|K = |ν(g)|K |f (x)|K = |f (x)|K .
We have called the above theorem a key lemma in [28]. Actually, in a slightly
different form it appeared in our paper of 1977 and in the above form with several
applications in [25]. We might mention that it would look more natural to use a
G(OK )-orbital decomposition of X ◦ . However, that approach did not work except
for some very simple cases. The usefulness of the key lemma comes from the facts
that firstly the G(Fq )-orbital decomposition of Fnq is not too difficult to obtain and
secondly in the partial integral of |f (x)|sK over ξ + πX ◦ we can replace f (x) by a
simpler polynomial in a smaller number of variables. This second process has been
formalized as a supplement to the key lemma in [28].
At any rate, we remark that if f¯(ξ)¯ = 0, then the partial integral over ξ + πX ◦
−n
is clearly q . On the other hand, if f (x) is homogeneous of degree d, then the
174 JUN-ICHI IGUSA
[n, d]Z(s) = card(f¯−1 (F×
q ))q −n
+ card(G(F q ) ¯
ξ) · |f (x)|sK dx.
ξ∈R ξ+πX ◦
Furthermore, if f¯(ξ)
¯ = 0 and (∂ f¯/∂xi )(ξ)
¯ = 0 for some i, then the proof of Theorem
10.2.1 shows that
|f (x)|sK dx = [1]q −n t/[1, 1].
ξ+πX ◦
for every g in G̃(OK ), then we can use the decomposition of Fnq into G̃(Fq )-orbits
instead of G(Fq )-orbits. We can also use the obvious fact that if G̃(Fq )ξ̄ denotes
the fixer of ξ¯ in G̃(Fq ) defined by ρ̄(ḡ)ξ¯ = ξ,
¯ then
¯ = card(G̃(Fq )) card(G̃(Fq ) ).
card(ρ̄(G̃(Fq ))ξ) ξ̄
We shall give an application of the key lemma which will illustrate how it works.
We start with some preparations. We take an arbitrary field F and we let GLn (F )
act on Symn (F ) as (g, x) → g ·x = gxt g. We keep in mind that rank(g ·x) = rank(x)
for every g in GLn (F ) and x in Symn (F ). We write n = p + k, where 1 ≤ p ≤ n,
and denote by g1 , g12 , g21 , g2 the p × p, p × k, k × p, k × k entry matrices of g; if
p = n, hence k = 0, it is understood that g = g1 . Also if x1 , x2 are respectively
in Mp (F ), Mk (F ), we denote by x1 ⊕ x2 the element of Mn (F ) with x1 , 0, 0, x2
as its entry matrices. Suppose now that x0 , x0 are in Symp (F ) and det(x0 x0 ) = 0.
Then we have
g · (x0 ⊕ 0) = x0 ⊕ 0
We have only to use these facts repeatedly to verify the above statement.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 175
ck = q −n(n+1)/2 · card(Σp ),
Proof. Suppose first that char(F ) = 2. If for every x in Symn (F ) we put Qx (u) =
t
uxu, where u is in F n , then the correspondence x → Qx gives a bijection from
Symn (F ) to the set of all quadratic forms on F n . Furthermore, if we let GLn (F )
act on this set as (g, Q) → (g · Q)(u) = Q(t gu), then the above bijection becomes
equivariant. Therefore by the classical diagonalization of a quadratic form and
by Theorem 9.2.1, we get the lemma except for the expression for ck . If ξ is a
representative of one of the two GLn (F )-orbits in Σp for F = Fq expressed as in
the lemma, then
g · (1 ⊕ J1 ) = 13 .
Actually, there are six elements such as g and one of them has 0 only at (2, 3) and
(3, 2), i.e., all other entries are 1. Therefore if p is even, then the GLp (F )-orbit of x0
contains either Jp/2 or ξ0 = 12 ⊕Jp/2−1 while if p is odd, it contains ξ1 = 1⊕J(p−1)/2 .
176 JUN-ICHI IGUSA
We observe that, in the case where p is even, the GLn (F )-orbits of ξ with Jp/2 and
ξ0 as x0 are different because Qξ = 0 for one and Qξ = 0 for another. Therefore,
the number of GLn (F )-orbits in Σp is 2 if p is even and 1 if p is odd.
We shall determine the fixers in GLn (F ) of the above representatives. We have
only to determine the fixers in GLp (F ) of ξ0 , ξ1 . We for a moment change the
notation and replace p by n and write ξ instead of ξ0 , ξ1 . Also we denote by H the
fixer of ξ in GLn (F ). Suppose first that n = 2m, put J = Jm−1 , and denote by a,
b, c, d the entry matrices of g, in which a is in M2 (F ), etc. Then g is in H, i.e.,
gξ t g = ξ for ξ = 12 ⊕ J, if and only if
at a + bJ t b = 12 , at c + bJ t d = 0, ct c + dJ t d = J.
We shall examine the above {·}. If we denote by g the element of GLn (K) such
that t g has 1p , −π(ξ0 + πx1 )−1 x12 , 0, 1k as its entry matrices, then g is in GLn (OK )
and
g · (ξ + πx) = (ξ0 + πx1 ) ⊕ π(x2 − π t x12 (ξ0 + πx1 )−1 x12 ).
Therefore, if we denote the original Z(s) by Zn (s), then {·} becomes tk Zk (s), hence
| det(x)|sK dx = q −n(n+1)/2 tk Zk (s).
ξ+πX ◦
This is a recursion formula by which we can determine Z(s) = Zn (s) starting with
Z0 (s) = 1. The result is as follows:
0≤k≤m 1≤i≤m−k
in which a# = Adj(a) = det(a)a−1 , etc. We shall first determine the GL6 (Fq )-
3 6
orbital decomposition of X(Fq ) = Fq . For a moment, we take any field F with
char(F ) = 2. We recall that f (g · x) = det(g)2 f (x) for every g in GL6 (F ) and x
in X. We express g by its 3 × 3 entry matrices α, β, γ, δ as before. We can easily
verify that if β = γ = 0 in g, then
where t is a variable, then we get f1 (ξ1 , x) = 4x456 . Since for every g in GL6 (F )
we have fp (g · ξ, g · x) = det(g)2 fp (ξ, x), if g · ξ1 = ξ1 and g · x = x , then x456 =
det(g)2 x456 , i.e.,
g4i g5j g6k xijk = det(g)2 x456
1≤i,j,k≤6
for every x in X(F ). We shall for a moment restrict our attention to the submatrix
(γ δ) of g. If we denote by πijk the determinant of the 3 × 3 matrix made up of its
i-th, j-th, k-th columns, then we can rewrite the above condition as
πijk xijk = det(g)2 x456
1≤i<j<k≤6
for every x, hence π456 = det(g)2 = 0. In particular, the 4-th, 5-th, 6-th columns
are linearly independent. Therefore, π145 = π146 = π156 = 0 implies that the
first column is 0. Similarly, we see that the second and the third columns are 0,
hence γ = 0 in g. On the other hand, if v = v1 e1 + v2 e2 + v3 e3 for vi in F ,
then ξ1 v = (v1 e4 + v2 e5 + v3 e6 )e123 . By applying the above g to this we see that
α = det(α)δ, i.e., δα# = 13 . Conversely, if for any α in GL3 (F ) we define δ as
above and take β = γ = 0 in g, then g is in the fixer of ξ1 in GL6 (F ). Furthermore,
if α = δ = 13 and γ = 0 in any g in GL6 (F ), then we can easily verify that
g · ξ1 = ξ1 + tr(β)e123 .
Therefore, the kernel of the surjective homomorphism from the fixer of ξ1 in GL6 (F )
to GL3 (F ) defined by g → α consists of all g with α = δ = 13 , γ = 0, and tr(β) = 0.
If F = Fq , this implies
in which g11 is in GL4 (F ) satisfying g11 J2 t g11 = λJ2 , g22 , λ are in F × , and other
entry matrices are free. Conversely, if g is of this form, then g is in GL6 (F ) and
g · ξ2 = ξ2 . We further observe that g → λ defines a surjective homomorphism from
the fixer of ξ2 in GL6 (F ) to F × because the diagonal matrix with λ, 1, λ, 1, 1, λ−1
as its diagonal entries is in the fixer for every λ in F × . Therefore, if F = Fq , then
Finally, if ξ3 = e123 , then Lemma 9.5.1 shows that the fixer of ξ3 in GL6 (F ) consists
of all g with α in SL3 (F ), δ in GL3 (F ), γ = 0, and β free in M3 (F ). If F = Fq ,
therefore, we get
In the above {·} we have ξ + πx = (a0 , πb0 ; πa, u + πb), in which a0 = 1 + πa0 .
Therefore, if we define an element g of SL6 (OK ) by α = δ = 13 , β = 0, and
γ = (a0 )−1 πa, then, as we have seen in Chapter 9.5, we have
The above integral can be reduced to an integral in Corollary 10.2.1 by the method
we have repeatedly used in section 10.1. We write b = (b1 b ) with a 3 × 2 submatrix
b and express the integral as an integral by db followed by an integration by db0 db1 .
Since the first column of u + πb can be completed to an element of SL3 (OK ), if b
denotes the bottom 2 × 2 submatrix of b , then the integral by db becomes an
integral by db . In that way we get
|f (ξ + πx)|sK dx = t2 · |b20 + det(b )|sK db0 db
X◦ OK ×M2 (OK )
Suppose next that i = 3 and put ξ = −ξ3 = (1, 0; 0, 0). Then by the first part
of the above argument we get
|f (ξ + πx)|K dx =
s
|(πb0 )2 + det(πb)|sK db0 db.
X◦ OK ×M3 (OK )
then
|y02 + π det(y)|sK dy0 dy = P (t)/[1, 1][5, 2][7, 2].
OK ×M3 (OK )
182 JUN-ICHI IGUSA
in which C(t) is
[5]+ − (1 + q −1 + q −2 − q −6 )q −5 t + (1 − q −4 − q −5 − q −6 )q −6 t2 + [5]+ q −12 t3 .
We recall that if C is any composition algebra over K and X = K 2 + H3 (C)2 ,
then the corresponding Freudenthal quartic f (x) on X is defined in Chapter 9.5.
We have just computed Z(s) in one case. Actually all Z(s) have been computed in
[25]. If for any m, n in N satisfying m ≥ n and for variables a, t, we put
Cm,n (a, t) = (1+am ) − (1 + am−n + a2m−2n − a2m−n )an+1 t
+ (1 − an − am − a2m−n )am+1 t2 + (1 + am )a2m+2 t3 ,
then for C = (C1), (C2), (C3), (C4) in Chaper 9.4, we have
|f (x)|sK dx = [1][3k/2 + 2]C3k/2+2,k+2 (q −1 , t)/
X◦
[1, 1][k + 3, 2][2k + 3, 2][3k + 4, 2],
in which k = dimK (C) = 1, i.e., k = 2, 4, 8; if k = 1, then it becomes
[1][4][7, 1] [1, 1][4, 2][7, 2].
The computation has been done uniformly. Instead of (∗) we have
[6k + 8, 4]Z(s) = [1][3k + 4] + [1][k + 2][3k/2 + 2][2k + 2]q −1 t/[1, 1]
+ c2 I2 + c3 I3 ,
in which
c2 = q −(k+3) [3k/2][3k/2 + 2][2k + 2]/[k/2],
c3 = q −(3k+4) [k/2 + 1]+ [k + 1]+ [3k/2 + 2],
I2 = [1][k + 3, 1]t2 /[1, 1][k + 3, 2],
I3 = ([1]t2 /[1, 1][k + 3, 2][2k + 3, 2]){1 − (1 + q −k/2 − q −k−1 )q −k/2−2 t
+ (1 + q −k/2 − q −k/2−1 − q −k−1 − q −3k/2−1 )q −k/2−2 t2 + q −3k−6 t3 }
for k = 2, 4, 8. This implies the above result. Actually Z(s) has been computed
also in some “twisted cases.” Furthermore, in the case where k = 8 not only Z(s)
but also Z(ω) was computed earlier via Weil’s function F ∗ (i∗ ). At any rate, by T.
Kimura [34] we have
bf (s) = (s + 1)(s + (k + 3)/2)(s + k + 3/2)(s + 3k/2 + 2)
for k = 1, 2, 4, 8.
We further mention that Z(s) for the norm form has been computed in all twisted
cases. If C is the unique unramified quadratic extension of K, i.e., the field K2 in
the notation of Chapter 11.6, the result is stated in [25]. Furthermore the norm
form of any associative simple K-algebra can be handled in the same way as det(x)
while every octonian K-algebra is isomorphic to the one in (C4). That leaves only
the case where C is either a ramified quadratic extension of K or a quaternion K-
algebra not isomorphic to M2 (K). In those rather difficult cases the computation
has been carried out by M. M. Robinson [46].
184 JUN-ICHI IGUSA
g13 = −αh−1t
0 (δ
−1
γ), g23 = −(1/2)γh−1t
0 (δ
−1
γ),
then we get a surjective homomorphism Kx → O(Q1 ) as g → g44 such that its kernel
is isomorphic as g → (b, a) to Mm1 ,k (F )×Altk (F ) with the following multiplication:
We have seen such a group in the proof of Lemma 10.3.1. At any rate Gx has now
been determined. We keep in mind that
up to a factor in (F × )2 .
We now take Fq as F , write X (F ) instead of X , and choose a complete set of
representatives of G(F )-orbits in X (F ). In doing so, we shall assume that d(Q) is
in (F × )2 if m is even so that, by Theorem 9.2.1, the anisotropic kernel of Q becomes
0. Then with respect to a suitable F -basis for V we can identify V with F m so that
we can write Q(v) = (1/2)t vhv, in which
0 1n 0
h = 1n 0 0
0 0 h
for some h . We then have X = Mm,n (F ) and iX (x) = t xhx for all x in X. We
define Σp as in Lemma 10.3.1 and choose representatives ηp of GLn (F )-orbits in
Σp of the form ηp = y0 ⊕ 0 with y0 in Symp (F ), hence det(y0 ) = 0. We keep in
mind that the number of ηp is 2 for p > 0 and 1 for p = 0. If we define an element
ξp of Mm,n (F ) as ξp = t (ηp 1n 0), then ξp is in X (F ) and (1/2)iX (ξp ) = ηp .
Therefore, {ξp } for all ηp and for 0 ≤ p ≤ n forms a complete set of representatives
of G(F )-orbits in X (F ).
We shall compute card(G(F ) · ξp ) for each ξp . We have card(G(F ) · ξp ) =
card(G)/card(Gx ) for x = ξp , in which card(G) = card(O(Q)) card(GLn (F )) and
Gx has been made explicit. Therefore, if we put
γ = q −m(m−1)/2 · card(O(Q))
and define γ0 , γ1 similarly for O(Q0 ), O(Q1 ), then we can easily verify that
q −mn · card(G(F ) · ξp ) = q −k(k+1)/2 · [k + i] · γ/γ0 γ1 .
1≤i≤p
pm1
Furthermore, since d(Q0 )d(Q1 ) = (−1) d(Q), where d(Q) = 1 if m is even, up to
a factor in (F × )2 , by using the formulas in Chapter 9.3 we get
γ =2· [2i]
1≤2i<m
Therefore, the reduction of Z(s) to Z0 (s) can be done by the following lemma:
Lemma 10.5.1 We only assume that m ≥ n in X = Mm,n (K) and that f (x) is
any homogeneous polynomial of degree d in πi (x) with coefficients in K. Then
|f (x)|sK dx = 1 [m − k + 1, d] · |f (x)|sK dx.
X◦ 1≤k≤n (X )◦
We shall postpone the proof of this lemma to section 10.6 and quickly finish the
computation of Z(s), i.e., that of Z0 (s). We shall use ηp = y0 ⊕ 0, ξp = t (ηp 1n 0)
for the liftings of the previous ηp , ξp to Symn (OK ), (X )◦ so that, e.g., det(y0 ) ≡ 0
mod π instead of det(y0 ) = 0; also we shall normalize h as before with entries in
OK this time. We observe that if we denote by x22 the k × k submatrix of x with
its (i, j)-entries for p < i, j ≤ n as the entries of x22 , then we will have
y + πy πy12
f (ξp + πx) = 2n · det 0 t 11 ,
π y12 πy22
in which the entries of y11 , y12 , y22 are SRP’s in the entries of x. We further observe
that y22 ≡ (1/2)(x22 + t x22 ) mod π and the above determinant is equal to
If we put
ck = q −mn card(G(F ) · ξ¯p )
ηp
for F = Fq and ξ¯p = ξp mod π, then by using the key lemma, the above result on
ck and Proposition 10.3.1 we get the following preliminary result:
188 JUN-ICHI IGUSA
in which
ck = q −k(k+1)/2 · [k + i] · [2i] [2i] · [2i]
1≤i≤p 1≤2i<m 1≤2i≤p 1≤2i≤m1
& '
[m/2 − k]+ p even
[m/2] m even
1 p odd
·
1 m odd,
&
[k + 1, 1] k even
Ik = 1 [1, 1] · [2i − 1]/[2i + 1, 2] ·
[k] k odd
1≤2i≤k
for 0 ≤ k ≤ n.
yi = πi (xx−1
1 )yi◦ = (−1)
n−j
cj yi◦
×
with yi◦ in OK and yi in OK , hence cj is in OK . Therefore, xx−1
1 with 1n as its top
n × n submatrix is in Mm,n (OK ), hence
1n ×
V =f OK .
Mr,n (OK )
×
Since OK , OK are compact and f is continuous, we see that V is compact.
Once we know that X, Y are locally compact, then by Lemma 7.3.1 we can
identify X, Y respectively with G/Gξ , G/Gη under g → x = gξ, g → y = g · η. We
shall denote by X ◦ the compact open subset of X consisting of all x in Mm,n (OK )
×
such that rank(x mod π) = n, i.e., with πi (x) in OK for some i. The condition
190 JUN-ICHI IGUSA
dg = | det(g)|−p
K µp2 (g)
gives a Haar measure on Gξ and d(gg0 ) = | det(g0 )|nK dg for every g0 in Gξ , hence
∆Gξ (g) = | det(g)|nK . If g is an element of Gη , hence g11 is in H, g21 = 0, and if for
a moment µH denotes any Haar measure on H, then
gives a Haar measure on Gη and d(gg0 ) = | det(g0 )|nK dg for every g0 in Gη , hence
∆Gη (g) = | det(g)|nK .
The above information about the modules on G, Gξ , Gη implies, in view of
Proposition 7.2.1, that X, Y have measures µX , µY = 0 satisfying
for every g in G. Actually, the restriction of µmn on Mm,n (K) to its open subset X
is such a measure µX and we simply take µX = µmn |X. Furthermore, we normalize
µH as
µH (H(OK )) = q −(n −1) card(H(F )) =
2
[k],
1<k≤n
for every ϕ in D(G) provided that the Haar measures dg, dk, dk respectively on
G, Gξ , Gη are suitably normalized. If we put
Φ(x) = ϕ(gk) dk,
Gξ
where x = gξ, then we know, cf. loc. cit., that the correspondence ϕ → Φ gives a
C-linear surjection from D(G) to D(X). Furthermore, if we put h = (k )11 , then
we have gk kξ = xh. Therefore, by Proposition 7.2.1 we have
ϕ(gk ) dk = ϕ(gk k) dk µH (h) = Φ(xh)µH (h).
Gη H Gξ H
Before we proceed further, we make the following remark. The relatively invariant
measure µY on Y remains relatively invariant under the normalizer of ρ(G) in
GLN (K). In particular, if c is in K × , then µY (cy) differs from µY (y) by a factor in
R×+ independent of y. By Lemma 7.2.1 this factor gives a continuous homomorphism
from K × to R× + . Therefore, it is of the form |c|K for some σ in R independent of c.
σ
On the other hand, if we put g = π1m , then ρ(g)y = π n y and | det(g)|nK = q −mn ,
hence µY (π n y) = q −mn µY (y). By putting these together we see that σ = m, hence
µY (cy) = |c|m
K µY (y)
for every c in K × .
We now express Y as the disjoint union of π k Y ◦ for all k in Z. Then by using
the above remark we can rewrite the integration formula as
(∗) ϕ(f (x))Φ(x)µX (x) = q −mk · ϕ(π k y) Φ(xh)µH (h) µY (y),
X k∈Z Y◦ H
in which f (x) = π k y. The above {·} can be made explicit, e.g., as follows: We
observe that it is invariant under x → gx for any g in G(OK ), hence it is independent
of y. We take any k from N and in (∗) we replace ϕ by the constant 1 and Φ by
the characteristic function of the subset of X0 defined by the condition that f (x)
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 193
is in π k Y ◦ . We observe that this subset is the disjoint union of E(e) for all e in Nn
satisfying |e| = e1 + . . . + en = k. Therefore we get
µX (E(e)) = q −mk · · · µY (Y ◦ ).
|e|=k
We put all these together and use the fact that Mm,n (OK )\X0 is of measure 0 by
µX (X0 ) = 1. In that way we get the following proposition:
Proposition 10.6.1 Let ϕ denote any C-valued continuous function on the image
of Mm,n (K) under x → (πi (x)) and Um,n the compact open subset of Mm,n (OK )
defined by rank(x mod π) = n. Then we have
ϕ(f (x)) dx = q −(m−k+1)ek ϕ(π |e| f (x)) dx,
Mm,n (OK ) e∈Nn 1≤k≤n Um,n
Now Lemma 10.5.1 follows from Proposition 10.6.1. We have only to denote
f (x) in Proposition 10.6.1 say by p(x), express f (x) in Lemma 10.5.1 by h(p(x))
with h(y) homogeneous of degree d, and take ϕ(y) = |h(y)|sK .
in which
'
[m/2] n even
m even
[m/2][m/2, 1]+ n odd '
A=
[n + 1, 1] n even
m odd
[m − n + 1, 1] n odd
while Ak = 1 except for the case where m, n are both even, and in that case
2
−k
(G3) Fm−k,k (a)Fk,n−k (a, t)ak tk = Fm−1,n (a, t) + tFm,n−1 (a, t),
0≤k≤n
in which
Fm,n (a, t) = (1 − am+i t)/(1 − ai ), Fm,n (a) = Fm,n (a, 1)
1≤i≤n
for m, n in Z with the understanding that Fm,n (a, t) = 0 for n < 0. In all cases, we
replace q −2 by a and q −1 t2 by t. More precisely, we shall be replacing the variables
a, t in (G2) and (G3) by q −2 , q −1 t2 .
If m, n are not both even, hence Ak = 1, we replace them respectively by
2(m + n + 1), 2n + 1 m even, n odd
2(m + n) + 1, 2n m odd, n even
2(m + n) + 1, 2n + 1 m odd, n odd.
and the sum over 0 ≤ k ≤ n is equal to Fm−1,n (a, t) + tFm,n−1 (a, t) by (G3).
Therefore, if in the original notation we put
We can easily verify the fact that Cm,n (a, t) is the only element of Q(a)[t] with
these properties. Furthermore, we can show that Cm,n (a, t) is irreducible in C(a)[t]
for n > 0 while clearly
We also mention that bf (s) for f (x) = det(t xhx) was computed as an example
of their general theory in a joint paper [50] by M. Sato, M. Kashiwara, T. Kimura,
and T. Oshima. The result is
bf (s) = (s + (k + 1)/2)(s + (m − k + 1)/2).
1≤k≤n
We observe that, in all examples which we have computed or mentioned, the real
parts of the poles of Z(s) for f (x) are the zeros of bf (s) with the order of each pole
at most equal to the order of the corresponding zero. As we have emphasized in the
Introduction, to convert this experimental fact into a theorem is an open problem.
As we have also mentioned in the Introduction, without the information about the
orders this has been proved by T. Kimura, F. Sato, and X.-W. Zhu [35] in the case
where f (x) is the basic relative invariant of an irreducible regular prehomogeneous
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 197
vector space. In their proof Theorem 8.5.1 and a theorem of M. Sato play key roles.
Also in the case where the number of variables in f (x) is 2 and in cases of some
generality, it has been proved by F. Loeser [37], [38].
Now, in spite of the complexity of the expression of Z(s) for the Gramian f (x) =
det(t xhx), it has remarkably simple properties. Firstly, if we denote by degt (Z(s))
the degree of Z(s) as a rational function of t, then we have
If we examine other examples of Z(s), then we will find that the above property is
shared by all Z(s) provided that f (x) is homogeneous and, e.g., non-zero coefficients
×
of f (x) are all units of OK . Secondly, the coefficients of the expressions of Z(s) as
a rational function of a = q −1 and t for f (x) = det(t xhx) are numerical constants
independent of K. Therefore, the process of replacing a, t by a−1 , t−1 , i.e., q by
q −1 , makes sense, and we have
Z(s)|q
→q−1 = t2n Z(s) = tdeg(f ) Z(s).
Z(s)|q
→q−1 = t4 Z(s) = tdeg(f ) Z(s).
Chapter 11
199
200 JUN-ICHI IGUSA
for all large r. After P. Samuel [47] we call d = deg(χ(G(A), t)) the dimension of A
and write dim(A) = d. We observe that
t+n
χ(G(A), t) = χ(K[x], t) − χ(a, t) = − χ(a, t)
n
and that, by the remark after Theorem 1.3.3, we have
t
deg χ(a, t) − <n
n
if a = 0. Therefore, we get d ≤ n with the equality if and only if a = 0. We shall
summarize the above observations as follows:
Proposition 11.1.1 Let A denote a local ring with m as its maximal ideal such
that A contains a field K satisfying A = K + m and G(A) the graded K-algebra
with Gr (A) = mr /mr+1 for all r. Then there exists a polynomial χ(t) satisfying
dimK (A/mr+1 ) = dimK (Gi (A)) = χ(r)
i≤r
p = Aa1 + . . . + Aap
for 0 ≤ p ≤ n. Then A/p is also a regular local ring, hence p is a prime ideal of A,
and dim(A/p) = n − p.
Proof. Since the local ring A is regular, every a in A gives rise to a unique sequence
f0 (x), f1 (x), f2 (x), . . . with fi (x) in K[x]i satisfying
a≡ fi (a) mod mr+1 ,
i≤r
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 201
Then A and A/p are regular local rings with dim(A) = n and dim(A/p) = n − p.
Proof. We apply the K-automorphism of K[x] defined by xi → yi = xi − ai for
1 ≤ i ≤ n and reduce the general case to the case where a = 0. Since we can
write A = K[x] + mr for all r, we can identify G(A) with K[x]. Therefore, A
is a regular local ring with dim(A) = n. Furthermore, by assumption there exist
1 ≤ j1 < . . . < jp ≤ n such that if g(x) denotes the determinant of the of p × p
submatrix of J(x) obtained by crossing out its j-th columns for all j = j1 , . . . , jp ,
then g(0) = 0. We observe that every f (x) in K[x] satisfying f (0) = 0 and its first
polar
(∂f /∂x1 )(0)x1 + . . . + (∂f /∂xn )(0)xn
have the same image in m/m2 . Therefore, if we denote xj for j = j1 , . . . , jp by
t1 , . . . , td , then the images of f1 (x), . . . , fp (x), t1 , . . . , td in m/m2 form its K-basis,
hence
m= Afi (x) + Atj .
1≤i≤p 1≤j≤d
It then follows from Proposition 11.1.2 that p is a prime ideal of A and A/p is a
regular local ring with dim(A/p) = d = n − p.
We remark that if we put P = K[x] ∩ p, then P is a prime ideal of K[x]
satisfying S −1 P = p and a minimal representation of the ideal of K[x] generated
by f1 (x), . . . , fp (x) is of the form
K[x]fi (x) = P ∩ Qi ,
1≤i≤p 1<i≤t
202 JUN-ICHI IGUSA
in which the primary ideal Qi of K[x] intersects S for 1 < i ≤ t. This follows from
Lemma 1.2.2 and Proposition 1.2.1.
then fij,i j (c) = 0 for all fij,i j (t) in I. Therefore, c is in Y . Conversely, suppose
that c is any point of Y with homogeneous coordinates (c11 , c12 , . . . ) so that ci0 j0 = 0
for some i0 , j0 . Denote by a, b the points of Projm , Projn with (c1,j0 , . . . , cm+1,j0 ),
(ci0 ,1 , . . . , ci0 ,n+1 ) as their respective homogeneous coordinates. Then we see that
c is the image of (a, b). We have thus shown that Y is the image of X. We shall
show that the surjective map X → Y is injective, hence bijective. Suppose that
(a , b ) in X has the same image as (a, b) and that (a1 , . . . , am+1 ), (b1 , . . . , bn+1 )
are the homogeneous coordinates of a , b . Then we will have ai bj = λai bj for all
i, j with some λ in Ω× . If ai0 bj0 = 0, then ai0 bj0 = 0, hence we may assume that
ai0 = bj0 = ai0 = bj0 = 1. This implies λ = 1, ai = ai bj0 = ai bj0 = ai , and
similarly bj = bj for all i, j, hence (a, b) = (a , b ). We now define OX so that the
bijection X → Y becomes an isomorphism. We might mention that OX can be
defined directly by using doubly homogeneous polynomials. We also remark that
if for any 1 ≤ i0 ≤ m + 1 we define an open subset Ui0 of Projm as before and
put J = {ti0 ,j ; 1 ≤ j ≤ n + 1}, then the above isomorphism X → Y gives rise
to an isomorphism from the product Ui0 × Projm to the difference of Y and the
closed subset of ProjN defined by J. Therefore, Aff m × Projn can be considered as
a quasi-projective variety.
As a rather special case of the above, we regard any closed subset of Aff n as a
quasi-projective variety, and call it an affine variety. If X is such a variety, then
the set I(X) of all f (t) in Ω[t] = Ω[t1 , . . . , tn ] which vanish at every point of X
forms an ideal of Ω[t]. We shall show that if g(t) is any element of Ω[t] different
from 0, then the open subset Xg of X consistsing of all a in X for which g(a) = 0
can be considered as an affine variety. In fact, if s is a variable, then the set Y of
all common zeros of elements of I(X) and h(t, s) = g(t)s − 1 in Ω[t, s] is an affine
variety in Aff n+1 . Furthermore, the correspondence a → (a, b), where b = 1/g(a),
gives a bijection from Xg to Y . If an element of OY,(a,b) is represented by a function
defined by f ∗ (t, s)/g ∗ (t, s) for f ∗ (t, s), g ∗ (t, s) in Ω[t, s] and g ∗ (a, b) = 0 and if e is
at least equal to the degrees in s of f ∗ (t, s), g ∗ (t, s), then
in which f (t) = g(t)e f ∗ (t, 1/g(t)), g (t) = g(t)e g ∗ (t, 1/g(t)) are in Ω[t] and g (a) =
0. This implies that X → Y is a morphism. Since the projection Y → X is a
morphism, the bijection X → Y is an isomorphism.
We shall make some remarks on the local ring OX,a for any quasi-projective
variety X. Firstly, OX,a will not change even if we replace X by Xg for any
homogeneous polynomial g(t) satisfying g(a) = 0. We observe that Xg for a suitable
g(t) becomes an affine variety. In fact, by definition X can be expressed as X =
X1 \X2 for some closed subsets X1 , X2 of Projn and a is not in X2 . Therefore,
we can find a homogeneous polynomial g0 (t) in I(X2 ) satisfying g0 (a) = 0. If
further a is in Ui , then we can take g(t) = g0 (t)ti . Therefore, after replacing X by
Xg , we may assume that X is an affine variety in Aff n . We change the notation
accordingly and write Ω[t] = Ω[t1 , . . . , tn ]. We shall show that if S denotes the set
of all g(t) in Ω[t] satisfying g(a) = 0, where a = (a1 , . . . , an ), then Oa = OX,a can
be identified with S −1 Ω[t]/S −1 I(X). In particular, Oa is a noetherian ring. We
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 205
We say, after O. Zariski [63], that a is a simple point of X if the equality holds, i.e.,
if Oa is a regular local ring. We sometimes denote by Xsmooth the set of all simple
points of X. If X = Xsmooth , then we call X nonsingular or smooth. Furthermore
if d = dima (X) is independent of a in X, then we say that X is d-dimensional
or, more precisely, everywhere d-dimensional. We observe that Aff n and Projn are
both smooth and everywhere n-dimensional.
Finally, suppose that a point a of Projn with homogeneous coordinates
(a1 , . . . , an+1 ) is contained in Ui and Uj , i.e., ai aj = 0. Then we clearly have
F (φi (a)) = F (φj (a)); we denote this field by F (a). If F (a) = F we say that a is an
F -rational point. If X is a quasi-projective F -variety in Projn , we denote by X(F )
the set of all F -rational points of X.
If we use (y1 , yn+1 , . . . , y2n−1 ) as the coordinates on Y1 , then the jacobian of h|Y1
becomes
∂(x1 , . . . , xn )/∂(y1 , yn+1 , . . . , y2n−1 ) = y1n−1 .
Also (y1 , yn+1 , . . . , y2n−1 ) = (x1 , x2 /x1 , . . . , xn /x1 ), hence h|Y1 gives rise to an
isomorphism from Y1 \y1−1 (0) to X\x−1 1 (0). At any rate Y is a closed smooth n-
dimensional subvariety of X × Projn−1 .
We now take f (x) from F [x1 , . . . , xn ]\F , denote by f −1 (0) the set of all a in
X satisfying f (a) = 0, and examine the effect of h : Y → X on the hypersurface
f −1 (0) in X. We shall assume that if f0 (x) is the leading form of f (x), i.e., the
homogeneous part of f (x) of the smallest degree, then N = deg(f0 ) > 0 and not all
partial derivatives of f0 (x) vanish at any point of f0−1 (0) other than 0. We observe
that the preimage (f ◦ h)−1 (0) of f −1 (0) in Y under h is the union of its intersection
with Yα for 1 ≤ α ≤ n. Since they are all similar, we take α = 1 and use the same
notation as above. We then have
in which
f1 (y) = f0 (1, yn+1 , . . . , y2n−1 ) + y1 f (y)
with f (y) in F [y1 , yn+1 , . . . , y2n−1 ]. In general, if f1 (t), . . . , fp (t) in Ω[t1 , . . . , tn ]
satisfy the condition in Theorem 11.1.1 for K = Ω, then we say that the p hyper-
surfaces f1−1 (0), . . . , fp−1 (0) in Aff n are transversal at a. In that case the theorem
implies that a is a simple point of their intersection. We shall show that f1−1 (0)
and y1−1 (0) are transversal at every point of their intersection.
If b = (b1 , bn+1 , . . . , b2n−1 ) is any point of their intersection, then b1 = 0 and
f1 (b) = f0 (c) = 0, in which c = (1, bn+1 , . . . , b2n−1 ). If f1−1 (0) and y1−1 (0) are not
transversal at b, then
for 1 ≤ i < n. If we write down Euler’s identity for f0 (x) and evaluate both sides at
c, then we also have (∂f0 /∂x1 )(c) = 0. Since f0 (c) = 0 and c = 0, this contradicts
the assumption that f0−1 (0)\{0} is smooth.
If we further assume that f (x) is homogeneous, hence f (x) = f0 (x), then
f −1 (0)\{0} is smooth. In that case since f1 (y) = f (1, yn+1 , . . . , y2n−1 ), the above
argument shows that f1−1 (0) is smooth. Therefore, if we denote by E1 , E2 the
closed subsets of Y such that their intersections with Y1 are respectively y1−1 (0),
f1−1 (0) and their intersections with Yα are similar, then (f ◦ h)−1 (0) becomes the
union of E1 , E2 and they are both smooth, (n − 1)-dimensional, and have normal
crossings in the sense that they are transversal at every point of their intersection.
Furthermore, the bijection from Y \h−1 (0) to X\{0} is an isomorphism. We also
observe that Y and E1 , E2 are F -varieties.
We shall now introduce the concept of Hironaka’s desingularization of a hyper-
surface by using some terminology which will be explained in a slightly different
language.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 207
Yαβ = h−1 −1
1 (0) ∩ . . . ∩ hm (0) ∩ Uβ .
If we write all E’s which intersect Yαβ by E1 , . . . , Ep , then the second condition
is that there exist f1 (t), . . . , fn (t) in k[t] such that Ei ∩ Yαβ = fi−1 (0) ∩ Yαβ for
1 ≤ i ≤ p and
d(t) = ∂(f1 , . . . , fn , h1 , . . . , hm ) ∂(t1 , . . . , tm+n )
h∗ (dx1 ∧ . . . ∧ dxn )Yαβ = η · fi (y)ni −1 · df1 (y) ∧ . . . ∧ dfn (y),
1≤i≤p
in which ε, η are units of k[Yαβ ]. These conditions are satisfied for a suitable choice
of {gβ (t)}.
We shall express the above conditions by identities in k[t]; we shall use e0 , e,
etc. to denote nonnegative integers.
First of all, by Hilbert’s Nullstellensatz there
exist gβ (t) in k[t] satisfying gβ (t)gβ (t) = 1. If we put
then k[Uβ ] = k[t, 1/gβ (t)] implies k[Yαβ ] = k[Uβ ]/k[Uβ ]a = k[y, 1/gβ (y)]. Therefore,
every element of k[Yαβ ] is of the form P (y)/gβ (y)e0 for some P (t) in k[t] and e0 ; it
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 209
is a unit of k[Yαβ ] if and only if P (t)P (t) ≡ gβ (t)e mod a for some P (t) in k[t] and
e. In particular, ε = ε0 (y)/gβ (y)eε , η = η0 (y)/gβ (y)eη and
ε0 (t)ε0 (t) ≡ η0 (t)η0 (t) ≡ gβ (t)e mod a
for some ε0 (t), ε0 (t), η0 (t), η0 (t) in k[t] and eε , eη , e. Furthermore,
d (t)d(t) = gβ (t)e
for some d (t) in k[t] and e, and
gβ (t)e f (t) ≡ ε0 (t)gβ (t)e0 · fi (t)Ni mod a
1≤i≤p
We might mention the proof in the special case where n = 4. If we use the notation
in section 11.2, then Proj1 × Proj1 is isomorphic to the surface in Proj3 defined by
t11 t22 − t12 t21 , hence M is C-bianalytic to P1 (C) × P1 (C), and hence
We might also mention that if Q(x) is of the above form, then the open subset of
X defined by x1 = 0 is isomorphic to Aff n−2 .
Before we explain the second example, we shall recall a system of equations
defining the Grassmann variety. We shall follow the presentation in G. B. Gurevich
[17]. Wetake an m-dimensional vector space V over an arbitrary field F and
n
examine
0 < n ≤ m. We shall write products of elements
(V ) for a fixed
of
n (V ) without using .n We first observe that if W is a subspace of V , then
n (W ) is a subspace
of (V
). Furthermore, if W1 , W2 are subspaces
of V , then
(W1 ∩ W2 ) = n (W1 ) ∩ n (W2 ). Therefore, for any xin n (V ) there exists
n
the smallest subspace Wx of V such that x is contained in (Wx ). We shall later
give an explicit description of Wx in terms of x. We observe that dimF (Wx ) ≥ n
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 211
with xi1 ...in in F . We define xi1 ...in for all 1 ≤ i1 , . . . , in ≤ m so that we get an
alternating tensor, which we call the representative tensor of x. We shall find the
necessary and sufficient condition in terms of xi1 ...in for x to be decomposable. We
observe that if x is decomposable, i.e., if x = v1 . . . vn for some v1 , . . . , vn in V ,
hence vj = y1j e1 + . . . + ymj em with yij in F , then the representative tensor of x
becomes
xi1 ...in = πi1 ...in (y),
in which y is the m × n matrix with yij as its (i, j)-entry and πi1 ...in (y) for i1 <
. . . < in is the determinant of the n × n submatrix of y obtained by crossing out its
k-th rows for k = i1 , . . . , in , and they are not all 0.
n
We take x from (V ) and ϕ from the dual space V ∗ of V . If xi1 ...in is the
representative tensor of x, then we define ∂ϕ x in n−1 (V ) as
(∂ϕ x)i1 ...in−1 = xi1 ...in−1 i ϕ(ei )
1≤i≤m
in which
zi1 ...in+1 ,j2 ...jn = (−1)k xi1 ...ik−1 ik+1 ...in+1 xik j2 ...jn ,
1≤k≤n+1
for all i > s and for all ϕ2 , . . . , ϕn in V ∗ . This implies that xii2 ...in = 0 for all
i2 , . . . , in if i > s, hence xi1 ...in = 0 only for i1 , . . . , in ≤ s. Therefore x is in
n
(W ), hence Wx is contained in W , hence Wx = W .
We are ready to show that x = 0 in n (V ) is decomposable if and only if its
representative tensor satisfies the following quadratic equations:
(∗∗) (−1)k xi1 ...ik−1 ik+1 ...in+1 xik j2 ...jn = 0
1≤k≤n+1
for all i1 < . . . < in+1 and j2 , . . . , jn . First, suppose that x is decomposable. Then
xv = 0 for all v in Wx , hence for all v = ∂2 . . . ∂n x. Then by (∗) we get (∗∗). Next,
suppose that (∗∗) is satisfied. Then by (∗) we get xWx = 0. If we choose an F -basis
v1 , . . . , vr for Wx , then x can be expressed uniquely as an F -linear combination of
vi1 . . . vin for 1 ≤ i1 < . . . < in ≤ r. Since xvi = 0 for 1 ≤ i ≤ r, we see that r = n
and x is in F v1 . . . vn .
We shall now explain the second example. We shall take Fq , Ω and later C
as F . We denote by Y the open subset of Mm,n consisting of all y with linearly
independent columns and consider ProjN for N + 1 = m n with homogeneous
coordinates xi1 ...in for 1 ≤ i1 < . . . < in ≤ m. Then the correspondence y →
x where xi1 ...in = πi1 ...in (y) gives a morphism from Y to the projective variety
X in ProjN defined by (∗∗), which is called the Grassmann variety. We have
seen in Chapter 10.6, though in different language, that the open subset of X
defined by xi1 ...in = 0 is isomorphic to Mm−n,n for every i1 , . . . , in . Therefore,
X is an irreducible smooth d-dimensional projective F -variety for d = (m − n)n.
Furthermore, if F = Fq , then the results there imply that
card(X(Fq )) = (1 − q m−k+1 )/(1 − q k ).
1≤k≤n
We recall that in the notation of Chapter 9.6 the RHS is Fm−n,n (q). We further
recall the Gauss identity:
Fm−n,n (q) q n(n−1)/2 tn = (1 + q i t).
0≤n≤m 0≤i≤m−1
This fact was known in classical algebraic geometry dealing with Schubert varieties.
We refer to C. Ehresmann [13] for a topological proof.
After these rather special examples, we shall state the main theorem on Weil’s
zeta functions. We recall that an element α of an extension of Q is called an
algebraic integer if it is a zero of a monic polynomial with coefficients in Z.
Theorem 11.4.1 Let X denote a smooth everywhere d-dimensional projective Fq -
variety. Then there exist nonnegative integers Bi for 0 ≤ i ≤ 2d satisfying Bi =
B2d−i and for each i algebraic integers αij = 0, in fact, all their conjugates over Q
in C having the absolute value q i/2 , for 1 ≤ i ≤ Bi with the following properties:
Firstly,
card(X(Fqe )) = (−1)i αij
e
0≤i≤2d 1≤j≤Bi
for 0 ≤ i ≤ 2d.
As we have stated in the Introduction, this is one of the two theorems which
we shall use without proof. Actually, the deeper part of the theorem stating that
αij are algebraic integers of absolute value q i/2 will not be used. At any rate, the
theorem stated above was conjectured by A. Weil and proved, in the above general
form by A. Grothendieck and P. Deligne. We refer to Deligne [8] for the history up
to his decisive contribution after the work of B. Dwork, Grothendieck and others
all with references. We might at least recall that Weil’s zeta function Z(t) of X is
defined for t near 0 in C as
dlogZ(t)/dt = card(X(Fqe ))te−1 , Z(0) = 1.
e≥1
The first and the second parts of the theorem then imply its rationality
Z(t) = (1 − αij t)εi , εi = (−1)i+1
0≤i≤2d 1≤j≤Bi
such a θ exists for every q and Fq3 = Fq (θ). We denote by σ the Fq -automorphism
of Fq3 defined by σθ = θ q and by a, b the points of Proj3 (Fq3 ) respectively with
(θ, σθ, σ 2 θ, 1), (θ, σ 2 θ, σθ, 0) as their homogeneous coordinates. We further denote
by L, L , L the lines in Proj3 respectively through {a, b}, {σa, σb}, {σ 2 a, σ 2 b}.
Then by the choice of θ we see that L, L , L are mutually disjoint, σ gives a
permutation L → L , L → L , L → L, and their union X is a smooth everywhere
1-dimensional projective Fq -variety. Since card(Proj1 (Fq )) = 1 + q for every q, we
get card(X(Fqe )) = 3(1 + q e ) or 0 according as e ≥ 1 is in 3Z or not. If ζ is any
element of C satisfying 1 + ζ + ζ 2 = 0, then we can write
card(X(Fqe )) = (1 + ζ e + ζ 2e )(1 + q e )
and define Y ◦ , (Yα )◦ as their respective intersections with Y (K) for all α. We
introduce the coordinates (t1 , . . . , tm+n ) as in section 11.3 and identify (X × Zα )◦
with OK m+n
. If we denote by (Yαβ )◦ the open subset of (Yα )◦ consisting of all b such
m+n
that gβ (b) is a unit of OK , i.e., the set of all b in OK satisfying h1 (b) = . . . =
×
hm (b) = 0 and gβ (b) in OK , then (Yα ) becomes the union of (Yαβ )◦ for all β. We
◦
We observe that Za (s) depends on the image ā of a in X̄(Fq ) rather than a itself.
If we take any b from h−1 (a + πX ◦ ), then b is in (Yαβ )◦ for some α, β. We observe
that if b is any point of Y ◦ with the same image in Ȳ (Fq ) as b, then b is also in
(Yαβ )◦ because b is in (Yα )◦ and gβ (b ) is a unit of OK . Furthermore,
i∈I i∈I
µY (b + πy) = |η · fi (b + πy)ni −1 |K · q −n dz = q −n · |πzi |nKi −1 · dz.
i∈I i∈I
216 JUN-ICHI IGUSA
This implies
Ni s+ni −1
Za (s) = q −n · |πzi |K dz,
n
OK
b̄ i∈I
−1
in which the summation is over the set h̄ (ā)(Fq ) and I for each b̄ can be identified
with the set of all Ē’s which contain b̄. Since the integral of |z|s−1K over OK is
(1 − q −1 )/(1 − q −s ), the above formula implies the following theorem of Denef:
Theorem 11.5.1 Take an arbitrary subset I of E and a point a of X ◦ with its
image ā in X̄(Fq ); denote by cI = cI (ā) the number of all b̄ in h̄−1 (ā)(Fq ) such that
b̄ is in Ē(Fq ) if and only if E is in I. Then
Za (s) = |f (x)|sK dx = q −n · cI · ((q − 1)/(q NE s+nE − 1)).
a+πX ◦ I E∈I
Consequently, the above limit is finite and different from 0. Since Z(s) is a rational
function of t, this implies that degt (Z(s)) = −d.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 217
then we have
Z(s) = q −n · card(ĒI0 (Fq )) · (q − 1)/(q NE s+nE − 1) .
I E∈I
h̄0 (t) are relatively prime and ḡ0 (t) = 0, deg(ḡ0 ) = deg(g0 ). Also, we fix an integer
d ≥ deg(g0 h0 ) and take any c(t) from OK [t] with deg(c) ≤ d. Then there exist a(t),
b(t) in OK [t] satisfying
a(t)h0 (t) + b(t)g0 (t) = c(t), deg(a) < deg(ḡ0 ), deg(b) ≤ d − deg(ḡ0 ).
In fact, since ḡ0 (t), h̄0 (t) are relatively prime, we have
for some ā(t), b̄(t) in F [t]. By replacing ā(t) by its residue mod ḡ0 (t), we may
assume that deg(ā) < deg(ḡ0 ). We then have deg(b̄) ≤ d − deg(ḡ0 ). Therefore, we
have only to apply the above remark with m = n = d + 1 to the coefficients of a(t),
b(t) as x and the coefficients of c(t) as y. We are ready to prove Hensel’s lemma,
which is as follows:
Lemma 11.6.1 Let f (t) denote any element of OK [t]\OK such that its image f¯(t)
in F [t], where F = OK /πOK , splits as f¯(t) = ḡ0 (t)h̄0 (t), in which ḡ0 (t), h̄0 (t) are
relatively prime and ḡ0 (t) = 0. Then there exist g(t), h(t) in OK [t] with deg(g) =
deg(ḡ0 ) satisfying
Proof. We shall exclude the trivially simple case where f¯(t) = 0, hence h̄0 (t) = 0.
We put d = deg(f ) and choose g0 (t), h0 (t) from OK [t] with ḡ0 (t), h̄0 (t) as their
images in F [t] and satisfying deg(g0 ) = deg(ḡ0 ), deg(h0 ) = deg(h̄0 ). That is clearly
possible and d ≥ deg(f¯) = deg(ḡ0 h̄0 ) = deg(g0 h0 ). We introduce unknown elements
gi (t), hi (t) of OK [t] satisfying
We have only to show that the equation f (t) = g(t)h(t) is solvable. If we put
c1 (t) = π −1 (f (t) − g0 (t)h0 (t)) and ck (t) = 0 for k > 1, then ck (t) is in OK [t] and
deg(ck ) ≤ d for every k. Furthermore, the equation g(t)h(t) = f (t) can be replaced
by the following sequence of equations:
gk (t)h0 (t) + hk (t)g0 (t) = ck (t) − gi (t)hk−i (t)
0<i<k
for k = 1, 2, 3, . . . . The first equation is g1 (t)h0 (t) + h1 (t)g0 (t) = c1 (t), which is
solvable in g1 (t), h1 (t) by the previous observation. Therefore, we shall apply an
induction on k assuming that k > 1. Then in the k-th equation, the RHS is an
already known polynomial in OK [t] of degree at most d, hence for the same reason
it is also solvable in gk (t), hk (t).
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 219
then [L : K] = re.
Proof. We shall first show that L has a nonarchimedean absolute value |·|L satisfying
|a|L = |a|nK for every a in K, where n = [L : K]. We start with a review of the
norm N from L to K for any field K. We choose a K-basis u1 , . . . , un for L. Then
we get a K-algebra homomorphism ρ : L → Mn (K) as
Since ρ(α) is unique up to ρ(α) → gρ(α)g −1 for some g in GLn (K) coming from
a change of the K-basis for L, N α = det(ρ(α)) is well defined. Furthermore,
N (αβ) = (N α)(N β) for every α, β in L. We need another property of N α. If f0 (t)
is the irreducible monic polynomial in K[t] of degree d satisfying f0 (α) = 0, then
1, α, . . . , αd−1 form a K-basis for K(α). If we compute the norm N0 α of α from
K(α) to K by using this basis, we get N0 α = (−1)d f0 (0). If we choose a K(α)-basis
v1 , . . . , ve for L, then
form a K-basis for L so that n = de. If we compute N α by using this basis, we get
with aij in K for all i, j then, since K is complete, we have only to show that
{ai1 }, . . . , {ain } are all Cauchy sequences in K. If k denotes the number of nonzero
sequences among them, the statement clearly holds for k ≤ 1. We shall therefore
assume that k > 1 and apply an induction on k. If we can derive a contradiction
from the assumption that one of the k sequences is not a Cauchy sequence, then
the induction will be complete. By converting the double sequence {αi − αj } into
a sequence and replacing it by a subsequence, we may assume that {αi } is a null
sequence in L while the absolute values of all terms of one of the k sequences in
K are at least > 0. After a permutation of u1 , . . . , un , we may assume that
{ai1 }, . . . , {aik } are the nonzero sequences and |aik |K ≥ for all i. Then the
sequence in L with
αi /aik − uk = (aij /aik )uj
1≤j<k
as its i-th term is convergent, hence a Cauchy sequence, and hence the coefficients
of u1 , . . . , uk−1 form Cauchy sequences in K by induction. If we denote their limits
by b1 , . . . , bk−1 , we get b1 u1 + . . . + bk−1 uk−1 + uk = 0. This contradicts the fact
that u1 , . . . , un form a K-basis for L.
Finally, since the image of L× under |·|L is a subgroup of the image of K × under
| · |K , it is discrete. Therefore, the ideal of nonunits of OL can be written as πL OL ,
r
and πOL = πL OL for some positive integer r. On the other hand, if w1 , . . . , we
are the elements of OL such that their images w̄1 , . . . , w̄e in OL /πL OL are linearly
independent over F = OK /πOK , then w1 , . . . , we are linearly independent over K,
hence e ≤ [L : K]. We shall assume that w̄1 , . . . , w̄e form an F -basis for OL /πL OL .
Since
r−1 r−1
1, πL , . . . , πL , π, ππL , . . . , ππL ,...
are elements of OL whose orders are respectively 0, 1, 2, . . . , every element α of
OL can be written as
i
α= aij wj πL
0≤i<r 1≤j≤e
with aij in OK for all i, j. We shall show that the expression is unique. We have
only to show that if the RHS is 0, then aij = 0 for all i, j. Suppose that the RHS is
0 for some aij in OK not all 0. Then, after cancelling a power of π, we may assume
that they are not all in πOK , hence their images āij in OK /πOK are not all 0. If
we take the image of the RHS in OL /πL OL , then we get ā01 w̄1 + . . . + ā0e w̄e = 0,
hence ā0j = 0, and hence a0j = πb0j with b0j in OK for all j. We can repeat the
same argument after cancelling πL , and we get a1j = πb1j with b1j in OK for all
j. By continuing this process, we will see that all aij are in πOK , a contradiction.
If we allow aij to have a power of π as a denominator, then every α in L can be
expressed uniquely as above with aij in K. We have thus shown that [L : K] = re.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 221
. . . . It is very likely that the two conditions are sufficient for the functional equa-
tion Z(u−1 , v −1 ) = v deg(f ) · Z(u, v).” The conjecture was investigated by D. Meuser
by using Denef’s formula in Theorem 11.5.1 and, under a certain condition on the
numerical data, was settled by her discovery of the relation between the conjectural
functional equation and the functional equations of Weil’s zeta functions over fi-
nite fields. Later J. Denef succeeded in removing this condition and his proof was
simplified by J. Oesterlé. We shall explain Denef-Meuser’s joint paper [10] on this
subject. As we shall see, their results go much further than the conjecture. We
refer to their paper for more detailed history and further important results.
Lemma 11.7.1 Let Ai denote a subset of a finite set S for all i in an index set I0
and I, J subsets of I0 . Then for any given I, we have
card Ai \ Ai = (−1)card(J\I) card Ai .
i∈I i∈I J⊃I i∈J
Proof. We take a arbitrarily from S and show that a contributes the same number
to both sides of the identity. Now a is contained in Ai1 , . . . , Aip for some i1 , . . . , ip
but not in Ai for any other i. First, suppose that I is not contained in {i1 , . . . , ip },
i.e., a is not contained in Ai for some i in I. Then, clearly, a contributes 0 to both
sides. Next, suppose that I is contained in {i1 , . . . , ip }. We put n = p − card(I).
If n = 0, i.e., I = {i1 , . . . , ip }, then a contributes 1 to both sides. If n > 0, then
a contributes 0 to the LHS. On the other hand, a is contained in Ai for all i in J
if and only if J is a subset of {i1 , . . . , ip }. Therefore, the contribution of a to the
RHS is
n
(−1)j = (1 − 1)n = 0.
j
0≤j≤n
Lemma 11.7.2 Let I0 denote a finite index set, i an element of I0 , and I, J subsets
of I0 ; further, let xi , yJ denote variables. Then we have
(−1)card(J\I) yJ · xi = yI · (xi − 1).
I J⊃I i∈I I i∈I
in X defined by f (x) is invariant. The fact, which we accept, is that the above
h : Y → X can be constructed in such a way that it is preserved under the action
of Ω× . We shall give some details. We recall that X ∗ = Projn−1 is the factor
space of X\{0} by Ω× . If Y ∗ denotes the factor space of Y \h−1 (0) by Ω× , then Y ∗
becomes a smooth (n − 1)-dimensional k-subvariety of X ∗ × Projm and h gives rise
to a Hironaka’s desingularization h∗ : Y ∗ → X ∗ of the hypersurface in X ∗ defined
by f (x). Furthermore, if E denotes the set of all E in E not contained in h−1 (0),
then for every E in E the factor space E ∗ of E\h−1 (0) by Ω× becomes a smooth
(n − 2)-dimensional k-subvariety of Y ∗ . More generally, for every subset I of E
EI∗ = E∗
E∈I
we will have
|f (x)|sK dx = q −n card (ĒI0 \h̄−1 (0))(Fq ) · xE .
X ◦ \πX ◦ I⊂E E∈I
We shall now apply Lemmas 11.7.1, 11.7.2 to the above summation in I with I0 = E ,
AE = (Ē\h̄−1 (0))(Fq ), and
yI = card AE .
E∈I
In doing so, we use the obvious fact that if S, S0 are subsets of a set, then the op-
eration S → S\S0 commutes with the taking of union, intersection, and difference.
We see that yI = (q − 1)card(ĒI∗ (Fq )) and that the above summation in I is equal
to
(q − 1) · card(ĒI∗ (Fq )) · (xE − 1).
I⊂E E∈I
We keep in mind that we can replace the above K by any finite algebraic ex-
tension L, e.g., by Ke in Theorem 11.6.1. We shall now explain an observation
by J. Oesterlé about the extendability of certain functions on N\{0} to Z\{0}. It
depends on the following lemma:
We observe that the rational function of t so defined depends only on ϕ and that
it has αi−1 as a pole or a zero of order |mi | according to whether mi is positive or
negative for 1 ≤ i ≤ r.
If now ϕ in A is expressed as in Lemma 11.7.3 possibly with mi = 0 for some
i and αi = αj for some i = j, then the function ϕ# on Z defined by the same
expression for all n in Z depends only on ϕ. In order to prove this fact, we have
only to show that ϕ = 0 implies ϕ# = 0. Let {β} denote the set of distinct
α1 , . . . , αr and for each β define mβ as the sum of all mi for αi = β. Then ϕ = 0
implies mβ = 0 for all β by Lemma 11.7.3, hence
ϕ# (n) = mi αin = mβ β n = 0
1≤i≤r β
in which ϕi is in A and ϕ#
i is in A# for all i in N. We observe that B, B # are
commutative
# rings with 1 and the subsets S, S # of B, B # defined by the condition
that ϕi (n)z in = 0 for every n in Z\{0} and for a variable z in C× are both
multiplicative and free from zero divisors. Therefore, C = S −1 B, C # = (S # )−1 B #
are defined, and C consists of elements of the form
Φ(z, n) = ϕi (n)z in / ψi (n)z in ,
i i
in which
for all e in N\{0}. We recall that ĒI is either empty or a smooth projective Fq -
variety of dimension rI = n − card(I) − 1. Therefore, Theorem 11.4.1 is applicable
to ĒI . In that way, we get
card(ĒI (Fqe )) = (−1)i αIij
e
0≤i≤2rI 1≤j≤BIi
226 JUN-ICHI IGUSA
[4] A. Borel and J.-P. Serre, Le théorème de Riemann-Roch, Bull. Soc. math.
France, 86 (1958), 97-136.
[7] C. Chevalley, The algebraic theory of spinors, Columbia Univ. Press (1954).
[8] P. Deligne, La conjecture de Weil, Inst. Hautes Études Sci. Publ. Math.,
43 (1974), 273-307.
[9] J. Denef, On the degree of Igusa’s local zeta function, Amer. J. Math., 109
(1987), 991-1008.
[10] J. Denef and D. Meuser, A functional equation of Igusa’s local zeta func-
tions, Amer. J. Math., 113 (1991), 1135-1152.
[11] J. Denef, Report on Igusa’s local zeta functions, Sém. Bourbaki 741 (1991),
1-25; Asterisque No. 201-203, 359-386.
[12] L. E. Dickson, Linear groups with an exposition of the Galois field theory,
B. G. Teubner (1901).
227
228 JUN-ICHI IGUSA
[18] A. Gyoja, Functional equation for Igusa local zeta functions, JAMI lecture
(1993).
[19] D. Hilbert, Über die vollen Invariantensysteme, Math. Ann. 42 (1893), 313-
373.
[21] J. Igusa, Analytic groups over complete fields, Proc. Nat. Acad. Sci., 42
(1956), 540-541.
[25] J. Igusa, Some results on p-adic complex powers, Amer. J. Math., 106
(1984), 1013-1032.
[29] J. Igusa, Universal p-adic zeta functions and their functional equations,
Amer. J. Math., 111 (1989), 671-716.
[30] J. Igusa, A stationary phase formula for p-adic integrals and its applica-
tions, Algebraic Geometry and its Applications, Springer-Verlag (1994),
175-194.
INTRODUCTION TO THE THEORY OF LOCAL ZETA FUNCTIONS 229
[31] J. Igusa, On local zeta functions, Amer. Math. Soc. Transl. (2) 172, (1996),
1-20.
[35] T. Kimura, F. Sato, and X.-W. Zhu, On the poles of p-adic complex powers
and b-functions of prehomogeneous vector spaces, Amer. J. Math., 112
(1990), 423-437.
[40] H. Mellin, Abriss einer einheitlichen Theorie der Gamma- und der hyper-
geometrischen Funktionen, Math. Ann., 68 (1910), 305-337.
[41] D. Meuser, On the poles of a local zeta function for curves, Invent. math.,
73 (1983), 445-465.
[42] D. Meuser, On the degree of a local zeta function, Comp. Math., 62 (1987),
17-29.
[44] M. Nagata, Local rings, Interscience Tracts in Pure and Appl. Math., 13
(1962).
[46] M. M. Robinson, The Igusa local zeta function associated with the singular
cases of the determinant and the Pfaffian, J. Number Theory, 57 (1996),
385-408.
[51] L. Schwartz, Théorie des distributions. I, II, Hermann & Cie (1950,1951).
[54] L. Strauss, Poles of a two variable p-adic complex power, Trans. Amer.
Math. Soc., 278 (1983), 481-493.
[55] J. Tate, Fourier analysis in number fields and Hecke’s zeta-functions, The-
sis, Princeton (1950).
[58] A, Weil, Sur la formule de Siegel dans la théorie des groupes classiques,
Acta Math., 113 (1965), 1-87.
231
232 JUN-ICHI IGUSA
International
Press
www.intlpress.com
AMSIP/14.S