Zeta Functions of Reductive Groups and Their Zeros (PDFDrive)
Zeta Functions of Reductive Groups and Their Zeros (PDFDrive)
REDUCTIVE GROUPS
AND THEIR ZEROS
Lin Weng
Kyushu University, Japan
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TAIPEI • CHENNAI • TOKYO
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.
Printed in Singapore
v
b2530 International Strategic Relations and China’s National Security: World at the Crossroads
Introduction
Here, MtotF,n denotes the moduli space of (OF -)lattices Λ of rank n (with OF the
integer ring of F), Θ(Λ) and vol(Λ) denote the theta series and the co-volume of
Λ, respectively. If it were convergent when n ≥ 2, (0.1) would be viewed as a rank
n non-abelian zeta function of F, since this integration recovers the Dedekind zeta
function of F if n = 1, and the rank n lattices are associated to the group GLn , for
which only GL1 is commutative. Obviously, using the Mellin transform, (0.1) can
be rewritten as
Z ∞Z
dT
Θ(Λ) vol(Λ) s dµ(Λ)
0 Mtot
F,n [T ]
T
Z Z ∞ !
1 dT
= Θ(T 2n · Λ) T s dµ(Λ) (0.2)
tot
MF,n [1] 0 T
Z
= E(Λ,
b s) dµ(Λ),
Mtot
F,n [1]
where, for T > 0, Mtot F,n [T ] denotes the moduli space of rank n lattices of co-
volumes T > 0, and E(Λ, b s) denotes the (complete) Epstein zeta functions of Λ.
Since Epstein zeta functions are special kinds of Eisenstein series and hence are
well-known to be of slow growth, and Mtot F,n [T ] are not compact, e.g. in the case
F = Q, Mtot F,n [1] is isomorphic to SL n (Z)\SL n (R)/SOn , all the integrations above
(over moduli spaces) diverge.
The first task of this book is to remedy the above constructions to obtain con-
vergent integrations. Motivated by Mumford’s stability in algebraic geometry, it
is only natural to truncate the moduli spaces Mtot F,n [T ] using the stability condition
so as to obtain the compact moduli spaces MF,n [T ] and hence also the moduli
space MF,n of semi-stable lattices of rank n.
vii
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page viii
viii Introduction
Theorem 0.1. Non-abelian zeta functions satisfy all basic zeta properties.
Namely,
(0) Up to a constant factor depending on n and the local and global units of F,
ζF,1 (s) = b
b ζF (s).
(1) The integration (0.3) defines a homomorphic function in s and admits a unique
meromorphic continuation b ζF,n (s) to the whole complex s-plane.
ζF,n (s) satisfies the standard functional equation
(2) Zeta function b
ζF,n (1 − s) = b
b ζF,n (s).
(3) Zeta function bζF,n (s) has two singularities only, i.e. two simple poles at s =
0, 1, and the residues admit the following geometric interpretations
ζF,n (s) = vol MF,n [1] .
Res s=1b
One of the central themes of this book is to expose algebraic, analytic and
geometric structures of these zeta functions. In particular, a weak Riemann hy-
pothesis for the rank n non-abelian zeta function will be established, ensuring that
all but finitely many zeros of b ζQ,n (s) lie on the central line <(s) = 1/2 when n ≥ 2.
To explain this, as in (0.2) above, the Mellin transform transforms (0.3) into
Z
bζF,n (s) = E(Λ,
b s) dµ(Λ). (0.4)
MF,n [1]
Since the Epstein zeta function E(Λ, s) coincides with the Eisenstein series
E SLn /Pn−1,1 (1, g; s), induced from the constant function one on the Levi subgroup
of the maximal subgroup Pn−1,1 of SLn corresponding to the ordered partition
n = (n − 1) + 1, the rank n non-abelian zeta function may be viewed as a special
geometric Eisenstein period, defined as the integration of Eisestein series over
some compact moduli spaces.
Naturally associated to zeta functions, Eisenstein series play a central role in
number theory as well. From their theories, ranging from the classical Siegel to the
modern Langlands, and the trace formula techniques, of the primitive Selberg and
the powerful Arthur, for an L2 automorphic function ϕ on a Levi subgroup M of a
parabolic subgroup P of a reductive group G, with respect to a sufficiently regular
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page ix
Introduction ix
Definition 0.2. The T -period of a split reductive group G over a number field F
is defined by
X ehwλ−ρ,T i Y ζF (hλ, αi)
ωG;T .
b
(λ) := (0.6)
F
ρ, αi
Q
α>0, wα<0 ζF (hλ, αi + 1)
w∈W α∈∆ hwλ − b
Here W, ∆, ρ, > and < 0 denote the Weyl group, the set of simple roots, the Weyl
vector and the positive and negative roots in the induced root system, respectively.
x Introduction
Up to this point, despite all the main constructions have emerged, fine struc-
tures for non-abelian zeta functions have not been exposed because of the lack of
any connection between geometric Eisenstein periods and analytic Eisenstein pe-
riods. Our next main theme is to show that these two types of Eisenstein periods
coincide, by disclosing an intrinsic relation between Arthur’s analytic truncation
over SLn (F)\SLn (A)1 /K and the geometric truncation of stability on Mtot F,n . This
is a number theoretic analogue of a result of Lafforgue and leads to the following:
Theorem 0.2. Under the identification SLn (F)\SLn (A)1 /K ' Mtot
F,n [1],
(1) (Analytic Aspect of Stability and Parabolic Reduction) For g ∈ SLn (A)
and a normalized convex polygon p on [0, n],
1(pg ≤ p) = (−1)|P|−1 1(pδg
X X
P >P p). (0.9)
P δ∈P(F)\G(F)
(2) (Equivalence of Analytic and Geometric Eisenstein Periods) For the char-
acteristic function 1MF,n [1] of MF,n [1] ⊂ Mtot
F,n [1],
Therefore, the rank n non-abelian zeta function b ζF,n (s) can be calculated
through a multiple residue of the SLn -period ωSL
F
n
(λ) in (0.8), and hence becomes a
linear combination of (products of) complete Dedekind zeta functions with ratio-
nal function coefficients. This result, apparently surprising, confirms once again
a hidden general principle that non-abelian invariants are built up from abelian
invariants via non-abelian symmetric structures.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page xi
Introduction xi
There are totally 2n−3 (n + 2) special elements, as exposed jointly with Kim.
Moreover, when working on the special uniformity of zeta functions for function
fields, jointly with Zagier, we are able to characterize all these (n − 1)-special
elements, or better, the special permutations, after identifying the Weyl group W
of SLn with the symmetric group Sn on n symbols. The up-shot is the following
Theorem 0.4. The SLn -zeta function b ζFSLn (s) and hence the rank n non-abelian
zeta function bζF,n (s) of F are given explicitly by
n
X
bζF,n (s) = b ζFSLn (ns − n) = F,n (ns − n) · ζF (ns − n + a).
R(a) b (0.13)
a=1
Our first application of this theorem is a closed formula for the volume of the
moduli space MF,n [1], since, by Theorems 0.1(3) and 0.4,
n
X (a)
vol MF,n [1] = Res s=1 RF,n (ns − n) · ζF (ns − n + a) .
b (0.15)
a=1
Corollary 0.1 (Semi-Stable Volume). The volume of the moduli space MQ,n [1]
of semi-stable OF -lattices of rank n and volume one is given by
n
X X vF,n1 · · ·b vF,nk
vol MF,n [1] = (−1)k−1 .
b
(0.16)
k=1 n ,...,n >0
(n1 + n2 ) · · · (nk−1 + nk )
1 k
n1 +...+nk =n
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page xii
xii Introduction
Theorem 0.5 (Weak Riemann Hypothesis). Let n ≥ 2. All but finitely many
1
ζQ,n (s) lie on the central line <(s) = .
zeros of the zeta function b
2
When n = 2, 3, 4 and 5, Suzuki-Lagaruas, Suzuki and Ki establish the full
Riemann hypothesis for b ζQSLn (s), respectively. Our first proof of this theorem was
obtained using the techniques developed in these works and a joint work on the
functional equation of bζFSLn (s) with Kim. In this book, as claimed above, we give
a simplified proof using Theorem 0.4. The key is to use a meromorphic decom-
position of the rank n zeta function induced by the refined symmetric structures
hidden in (0.13). Namely,
≤ n+1
≥ n+1 ≤ n+1
≥ n+1
ζQ 2 (s)
b
ζQ 2 (s) + b
ζQ,n(s) = b ζQ 2 (s) = b
ζQ 2 (s) · 1 + (0.17)
b
≥ n+1
ζ 2 (s)
b
Q
where
≥ n+1 X0
ζQ 2 (s) :=
b
F,n (ns − n) · ζF (ns − n + a),
R(a) b
a≥ n+1
2
n+1 X0 (0.18)
≤
ζQ 2 (s) :=
b
F,n (ns − n) · ζF (ns − n + a)
R(a) b
a≤ n+1
2
P
X0
a≥ n+1
2
n even X0
with n+1
:= and a similar . Surely,
a≤ n+1
a≥ 2
a≥ n+1 − 1 a= n+1
P P
n odd 2
2 2 2
(0.17) is compatible with the functional equation since
n+1 n+1
≥ ≤
ζQ 2 (1 − s) = b
b ζQ 2 (s). (0.19)
Hence, to locate the zeros of rank n zeta function b ζQ,n (s), it suffices to expose
≥ n+1 ≤ n+1 . ≥ n+1
the structures of the factor ζ
b 2
Q (s) and the quotient ζ 2 (s) b
b
Qζ 2 (s). Being the
Q
combinations of Riemann zeta functions with coefficients of rational functions,
≥ n+1 ≤ n+1
ζQ 2 (s) and hence also b
b ζQ 2 (s) admit Hadamard products. Say,
n
≥ n+1 Y ≥ n+1
ξQ (s) := (ns − n + k) · b
2
ζQ 2 (s) = κn ecn s Pn (s) Hn (s) (0.20)
k=1
where
Introduction xiii
n+1
≥
ζQ 2 (s) in the
Theorem 0.6. Let n ≥ 2. There are at most finitely many zeros of b
region of <(s) ≥ 1/2 and hence bζQ,n (s) satisfies the weak Riemann hypothesis,
provided that ∆Q,n > 0.
xiv Introduction
Definition 0.7. The secondary big Delta pair correlation sequence for the zeros
ζQ,n (s) are defined by
of b
n
∆n,k := δn,k − 1 · log γn,k .
(0.28)
2π
1 Here and in the sequel, when n = 2, stronger results hold. For details, please see Part 6.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page xv
Introduction xv
Theorem 0.9. Let ϕ(x) be a function in C 1 (R)!such that ϕ0 (x) 1 for |x| < 1,
d ζ0 n
ϕ0 (x) x−2 for |x| ≥ 1, and ϕ < + int is bounded on R. Then for n ≥ 2,
dt ζ 2
Z ∞
1 X 1 n
ϕ ∆n,k = ϕ(u) du.
lim m u, (0.30)
T →∞ Nn (T ) 2π −∞ 2
0<γk ≤T
Z ∞
Here m(u, σ) := M(u + iv, σ) dv.
−∞
With all these said, reemerged is the role played by the stability condition for
non-abelian zeta functions b ζF,n (s). Note that, instead of T = 0 above, correspond-
ing to the stability condition, every positive T ≥ 0 can be used to introduce a
ζF,n
T -version of b T
(s) satisfying bζF,n
0
(s) = b
ζF,n (s). Even many tautological zeta prop-
erties can be similarly established for ζF,n bT
ζQ,3
(s), numerical calculations with b T
(s)
indicate that, only when T = 0, the Riemann hypothesis holds. So, stability does
play a key role for non-abelian zeta functions to satisfy the Riemann hypothesis.
We will not explore the deep reasons behind. Instead, in this book, we de-
velop a new theory for zeta functions of general split reductive groups G, or
better, of split reductive groups G and their maximal parabolic subgroups P, over
number fields. The searching process for these zeta functions b ζFG/P (s) is faithfully
recorded in Part 4, with a theoretic preparation in Part 3, in which genuine mul-
tiple L-functions for reductive groups are introduced using Langlands’ theory of
Eisenstein series and analytic and geometric truncations. The up-shot is an ana-
logue of Definitions 0.2 and 0.3. The functional equation of bζFG/P (s) is established
essentially following Komori in Part 4 as well. In Parts 5 and 6, we prove a weak
Riemann Hypothesis for the zeta functions b ζFG/P (s). It consists of two aspects.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page xvi
xvi Introduction
This book ends with five appendices on our joint works with K. Sugahara on
new arithmetic cohomologies for arithmetic varieties, ranging from the construc-
tions of arithmetic adelic cohomology groups in higher dimensions, to a theory of
ind-pro topologies in arithmetic dimension two, and the reciprocity laws for arith-
metic surfaces. They may be viewed as natural continuations of the cohomology
theory in arithmetic dimension one of Part 1.
Acknowledgement. I would like to thank Hida for his consistent support. The
first suggestion to the author on writing a book about the zetas for reductive groups
came from him. I also would like to thank Lafforgue for his help. Not only the
quantitive relation between stability and parabolic reduction is motivated by his
work, but the title of the book was proposed by him. The zeta project, as a part of
the Geometric Arithmetic program, was initiated in 1999. Working alone mainly,
the author has benefit from the visits to IHES, MPIM, and UCLA, where many
components of this project were done. I would like to thank them for providing
excellent research environments. Without the works of Suzuki, Komori and all our
co-authors, we cannot present a complete story on the weak Riemann hypothesis
of our zeta functions. I would like to thank them for their contributions. Special
thanks are due to Sugahara for agreeing to publish our joint works as the appen-
dices to this book. At last, but not the least, I would like to thank my family for
their understanding, trust and support given to me in all these years.
The zeta project has been partially supported by JSPS.
December 21, 2017 14:18 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page xvii
Contents
Introduction vii
1. Semi-Stable Lattice 3
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Projective Modules . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.1 Invertible Modules . . . . . . . . . . . . . . . . . . . . 7
1.2.2 General Projective Modules . . . . . . . . . . . . . . . 8
1.3 Stability of Lattices . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.1 Lattices . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.2 Covolumes of Lattices . . . . . . . . . . . . . . . . . . 10
1.3.3 Stability of Lattices . . . . . . . . . . . . . . . . . . . 11
1.3.4 Canonical Filtration . . . . . . . . . . . . . . . . . . . 13
1.4 Volume of Lattice: Special Linear Group . . . . . . . . . . . . . 17
1.4.1 Metrics of Lattices . . . . . . . . . . . . . . . . . . . . 17
1.4.2 Special Metrics of Lattices . . . . . . . . . . . . . . . . 18
1.5 Automorphisms of Lattices . . . . . . . . . . . . . . . . . . . . 20
1.5.1 General Automorphisms . . . . . . . . . . . . . . . . . 20
1.5.2 Special Automorphisms . . . . . . . . . . . . . . . . . 21
1.5.3 Unit Automorphisms . . . . . . . . . . . . . . . . . . . 23
1.6 Compact Moduli Spaces of Semi-Stable Lattices . . . . . . . . . 25
xvii
December 21, 2017 14:18 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page xviii
xviii Contents
2. Geometry of Numbers 27
2.1 Global Cohomology . . . . . . . . . . . . . . . . . . . . . . . . 27
2.1.1 Adelic Ring . . . . . . . . . . . . . . . . . . . . . . . 27
2.1.2 Adelic Cohomology Groups . . . . . . . . . . . . . . . 28
2.1.3 Topological Duality: Local and Global Pairings . . . . 29
2.2 Duality and Riemann-Roch Theorem in Arithmetic . . . . . . . 30
2.2.1 Nine-Diagram . . . . . . . . . . . . . . . . . . . . . . 30
2.2.2 Arithmetic Riemann-Roch Theorem . . . . . . . . . . . 32
2.3 Arithmetic Vanishing Theorem . . . . . . . . . . . . . . . . . . 36
2.3.1 Elementary Properties of 0-th Cohomology . . . . . . . 36
2.3.2 Ampleness and Vanishing Theorem . . . . . . . . . . . 39
Contents xix
xx Contents
Contents xxi
xxii Contents
Contents xxiii
12. Algebraic and Analytic Structures and Weak Riemann Hypothesis 299
12.1 Criterion for Weak Riemann Hypothesis . . . . . . . . . . . . . 299
12.1.1 Discriminant of (G, P) . . . . . . . . . . . . . . . . . . 299
12.1.2 Criterion of Weak Riemann Hypothesis . . . . . . . . . 301
12.2 Refined Symmetries of Zeta Functions . . . . . . . . . . . . . . 301
12.2.1 Positively Oriented . . . . . . . . . . . . . . . . . . . . 301
12.2.2 Refined Symmetries . . . . . . . . . . . . . . . . . . . 302
12.3 Dominate Term . . . . . . . . . . . . . . . . . . . . . . . . . . 304
12.3.1 Entire Function Oriented . . . . . . . . . . . . . . . . . 304
12.3.2 Terms with Maximal Discrepancy . . . . . . . . . . . . 306
12.3.3 Relation between W p and W p . . . . . . . . . . . . . . 309
12.3.4 Leading Polynomials . . . . . . . . . . . . . . . . . . 312
12.4 Zero Free Regions . . . . . . . . . . . . . . . . . . . . . . . . . 313
12.4.1 Lie Theoretic Structures Involved . . . . . . . . . . . . 313
12.4.2 Estimations on Right Half Plane . . . . . . . . . . . . . 318
12.4.3 Normalization Factor . . . . . . . . . . . . . . . . . . 319
12.4.4 Zero Free Region on Right Half Plane . . . . . . . . . 320
12.4.5 Zero Free Region on Left Half Plane . . . . . . . . . . 322
12.5 Hadamard Product . . . . . . . . . . . . . . . . . . . . . . . . 325
12.5.1 Distributions for Zeros of A p (s) . . . . . . . . . . . . . 325
12.5.2 Hadamard Product . . . . . . . . . . . . . . . . . . . . 328
12.6 Proof of Theorem . . . . . . . . . . . . . . . . . . . . . . . . . 332
12.7 Discriminant and Residue of Period . . . . . . . . . . . . . . . 333
12.7.1 Residue of Period . . . . . . . . . . . . . . . . . . . . 333
12.7.2 Period of Pseudo Root System . . . . . . . . . . . . . . 335
12.7.3 Residue and Leading Coefficient . . . . . . . . . . . . 337
12.A Decomposition of Σ p (1) . . . . . . . . . . . . . . . . . . . . . . 339
12.B Geometric Interpretation of c p . . . . . . . . . . . . . . . . . . 348
13. Weak Riemann Hypothesis for Zeta Functions of Exceptional Groups 349
13.1 Weyl Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
13.1.1 Root Systems . . . . . . . . . . . . . . . . . . . . . . 349
13.1.2 Weyl Group for D5 . . . . . . . . . . . . . . . . . . . . 350
December 21, 2017 14:18 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page xxiv
xxiv Contents
15. Special Uniformity of Zeta Functions and Weak Riemann Hypothesis 381
15.1 Special Uniformity of Zeta Functions . . . . . . . . . . . . . . . 381
15.1.1 Basic Properties of Truncations . . . . . . . . . . . . . 381
15.1.2 Analytic and Geometric Periods . . . . . . . . . . . . . 383
15.2 Weak Riemann Hypothesis for Non-Abelian Zeta Functions . . . 384
15.2.1 Parabolic Reduction, Stability and Volumes . . . . . . . 384
15.2.2 Existence of Stable OF -Lattices . . . . . . . . . . . . . 386
15.2.3 Weak Riemann Hypothesis . . . . . . . . . . . . . . . 386
Contents xxv
17. Weak Riemann Hypothesis for Zeta Functions of Reductive Groups 427
17.1 Volumes of Semi-Stable Moduli Spaces . . . . . . . . . . . . . 427
17.1.1 Parabolic Reduction, Stability and the Volumes . . . . . 427
17.2 Riemann Hypothesis for Zeta Functions of Reductive Groups . . 431
17.2.1 Geometric Interpretation of Discriminant . . . . . . . . 431
17.2.2 Riemann Hypothesis for Weng Zeta Functions . . . . . 432
xxvi Contents
Contents xxvii
Bibliography 519
Index 525
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 1
PART 1
Non-abelian zeta functions for number fields are constructed as the integrations
over the moduli spaces of semi-stable lattices together with the integrants involv-
ing a newly developed arithmetic cohomology theory. Like their classical ana-
logues, these new yet genuine zeta functions are well-defined, satisfy the standard
zeta properties, including a weak Riemann Hypothesis. In this part, we first re-
call the basic theory for the lattices over the ring of integers of number fields,
introduce the stability conditions and construct the moduli spaces for these lat-
tices. This makes the integral domains easily accessible. Then, we develop an
arithmetic cohomology theory for arithmetic curves by introducing two types of
cohomology groups. We show that not only the arithmetic counts for these groups
satisfy a duality and an arithmetic Riemann-Roch theorem, but its refined arith-
metic vanishing theorem under the stability condition implies naturally that the
integrations above are well-defined. In addition, all the zeta properties can be
established tautologically in this language. For example, the functional equation
comes from the numerical duality and the arithmetic Riemann-Roch theorem, and
the geometric interpretation of the residues at one of the zeta functions as the
volumes of the moduli spaces of semi-stable lattices with fixed volumes comes
from the construction and the vanishing theorem. To end this part, we express all
the non-abelian zeta functions as the integrations of the (complete) Epstein zeta
functions over some compact moduli spaces of semi-stable lattices.
1
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 3
Chapter 1
Semi-Stable Lattice
1.1 Motivation
3
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 4
K ∗ := a = (α) ∈ I0K : av = α ∈ K ∗ ∀v ∈ S ,
Set U K,fin := U K ∩ IK,fin . In addition, let Pic(OK ) be the (Arakelov) Picard group of
K, defined as the abelian group of isometric classes invertible OK -lattices and let
Pic0 (OK ) be the subgroups consisting of degree zero classes. Then, with respect
to the natural locally compact (ind-pro) topology on IK , we have
such that N(a) = N(I(a))−1 . In particular, the class group CLK is finite.
(4) Pic(OK ) of K, resp. Pic0 (K), is isomorphic to U K,fin \IK /K ∗ , resp.
U K,fin \I0K /K ∗ . In particular, the degree and the norm map factor through
Pic(OK ) with Pic0 (K) taking values one and zero, respectively.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 5
Semi-Stable Lattice 5
We may write an idele a ∈ IK as a = (afin .a∞ ) with afin ∈ IK,fin and a∞ ∈ IK,∞ .
In particular, if dµ(a) denotes the normalized Haar measure on IK , induced by
vol(Ov ) = 1 for all v ∈ S fin , and the canonical Euclidean, resp. Hermitian, metric
on R, resp. on C, (see e.g. [118]), we obtain a measure dµ(a) on IK and hence on
Pic(OK ) via dµ(a) = dµ(afin ) · dµ(a∞ ).
Definition 1.1. Let K be a number field. The (complete) Dedekind zeta function
of K is defined by
s
s
s r1 r2 X
ζK (s) := ∆K2 2π− 2 Γ
b (2π)−s Γ(s) N(I)−s <(s) > 1. (1.2)
2 0,I E O K
Semi-Stable Lattice 7
(1) The rank of the OK -module M, denoted by rankOK (M), or simply rank(M), is
defined to be the dimension of the K-linear space MK := M ⊗OK K.
(2) The OK -module M is called projective if there exists an OK -module N such
that M ⊕ N is a free OK -module.
Proof. Without loss of generality, we may assume that both a and b are integral.
From the short exact sequences of OK -modules
0 → a ∩ b → a ⊕ b → a + b → 0, (1.7)
we have a ⊕ b ' (a + b) ⊕ (a ∩ b). If a and b are co-prime in OK , then
a + b ' OK , and a ⊕ b ' OK ⊕ a b. We are done. For general a and b,
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 9
Semi-Stable Lattice 9
from the unique factorization property for integral ideals of the Dedekind do-
main OK , there exist α, β ∈ K such that α a and β b are co-prime. Therefore,
a ⊕ b ' α a ⊕ β b ' OK ⊕ (α a) (β b) ' OK ⊕ a b.
1.3.1 Lattices
For later use, we denote by rankOK (Λ) the rank of the OK -module P and denote
by rankZ (Λ) the rank of the Z-module P. Since an OK -ideal is of rank n as a Z-
module, where n = r1 + 2r2 = [K : Q] is the extension degree of K over Q,
rankZ (Λ) = rankOK (Λ) · rankZ (OK ) = rank(Λ) · n. (1.9)
Accordingly, V(P) := P ⊗Z R = σ∈S∞ V(P)σ is an n-dimensional real space.
Q
Moreover, ρ induces a natural metric on V(P), which we also denote by ρ by an
abuse of notations.
X X
hu, wi∞ := huσ , wσ iσ + <(huσ , wσ iσ ) ∀u, w ∈ V(P). (1.10)
σ:R σ:C
We denote by ρZ the induced metric on V(P).
Semi-Stable Lattice 11
Definition 1.8. An OK -lattice Λ is called semi-stable (resp. stable), if, for all sub
OK -lattices Λ1 of Λ,
vol(Λ1 )rankOK (Λ) ≥ (resp. >) vol(Λ)rankOK (Λ1 ) . (1.14)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 12
Remark 1.1. (1) This definition of stability for lattices is motivated by Mum-
ford stability for vector bundles over integral, regular, projective curves defined
over algebraic closed fields. Indeed, by the Arakelov-Riemann-Roch theorem, i.e.
Proposition 1.5, the inequality (1.14) is equivalent to
degar (Λ1 ) degar (Λ)
≤ or equivalently µ(Λ1 ) ≤ µ(Λ) ∀Λ1 ≤ Λ (1.15)
rankOK(Λ1 ) rankOK(Λ)
where, by definition, for an OK -lattice Λ, its µ-invariant is defined by
degar (Λ)
µ(Λ) := . (1.16)
rankOK(Λ)
In this form, (11.2) is the same as the µ-semi-stability condition of Mumford for
vector bundles over curves.
(2) The main reason for Mumford to introduce the stability condition is to con-
struct a projective moduli spaces for geometric objects. In the case of vector bun-
dles over a proper regular integral curve over algebraic closed field, the space of
isomorphic classes of vector bundles of rank r and degree d is not even Hausdorff.
However, if we consider only stable bundles, their moduli space becomes quasi-
projective. In addition, by adding the so-called S-classes of semi-stable bundles,
we obtain a projective moduli space.
(3) Stability condition for lattices was introduced when we constructed the non-
abelian zeta function for number fields in 1999. See, however, [104, 105].
(1) Λ is semi-stable.
(2) For all proper sub OK -lattices Λ1 of Λ,
volLeb (Λ1 )rankOK (Λ) ≥ volLeb (Λ)rankOK (Λ1 ) .
(3) For all quotient OK -lattices Λ2 of Λ,
volLeb (Λ)rankOK (Λ2 ) ≥ volLeb (Λ2 )rankOK (Λ) .
Semi-Stable Lattice 13
Claim 1.1. Let Λ = (P, ρ) be an OK -lattice. The following two are equivalent.
(1) Λ is semi-stable.
(4) For all projective sub OK -modules P1 of P,
vol(Λ ∩ P1 )rankOK (Λ) ≥ vol(Λ)rankOK (Λ∩P1 ) .
Proof. (4) implies (1) is clear. The opposite direction is slightly complicated.
The point is that while Λ ∩ P1 is an OK -lattice, it may not be a sub OK -lattice
of Λ. However, it is well known that, for P1 ⊆ P, there exists a projective sub
OK -module P1env , the envelop of P1 in P, such that P1 ⊆ P1env is a finite index sub
OK -module and the quotient P/P1env is projective. In addition, by Definition 1.5,
we certainly have vol(Λ ∩ P1 ) = vol(Λ ∩ P1env ) [P1env : P1 ]. Therefore
env
vol(Λ ∩ P1env )rankOK (Λ) ≥ vol(Λ)rankOK (Λ∩P1 ) ,
implies (4). This proves the claim and hence the lemma.
(a) one, denoted by L(0, Λ1 ), of slope µ(Λ1 )− 21 log ∆K , connecting Q(0) to Q(Λ1 ),
(b) the other, denoted by L(Λ1 , Λ), of slope µ(Λ/Λ1 )− 21 log ∆K , connecting Q(Λ1 )
to Q(Λ).
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 14
This, together with the line segment L(0, Λ), forms a triangle ∆(0, Λ1 , Λ) in the
xy-plane. In this language, the stability condition is naturally translated.
Corollary 1.2. For the OK -lattice, the following conditions are equivalent.
Lemma 1.3. Let Λ be a fixed OK -lattice. For any constant c ∈ R∗+ , there exist
only finitely many sub OK -lattices Λ1 of Λ such that vol(Λ1 ) ≤ c. In particular,
the subset vol(Λ1 ) : Λ1 ⊆ Λ of R≥0 is discrete and bounded from below.
Proof. By the restriction of scalars, we may reduce the proof to Z-lattices. Fur-
thermore, as we did in the proof of Lemma 1.1, passing to the enveloping sub
Z-modules, it suffices to treat the cases Λ1 ≤ Λ. For this, we use an induction on
r. Let r1 be the rank of Λ1 = Λ ∩ P1 . If r1 = 1, the Λ1 ’s are lattices vectors in Λ
and hence form a discrete subset in the metrized space ΛR := Λ ⊗Z R. Obviously,
this discrete subset admits only finitely many intersections with the compact ball
of radius c, centered at the origin. Hence, when r1 = 1 or r = 1, we are done. As-
sume now the same holds for all Z-lattices of rank ≤ r − 1. To treat general r1 , we
use the map Λ 7→ ∧r1 Λ, which sends Λ1 to ∧r1 Λ1 . By the induction hypothesis,
we may assume that ∧r1 Λ1 is not degenerate. That is, det Λ1 := L is a Z-lattice of
rank one. By the case r1 = 1 proved above, there are only finitely many L’s such
that vol(L) ≤ c. This then completes the proof, since vol(det Λ1 ) = vol(Λ1 ) and,
for a fixed L, there are only finitely many Λ1 such that det Λ1 = L.
Hence, within Λ, there exists a unique sub OK -lattice Λ1 such that, not only
(a) the slope of L(0, Λ1 ) attains the maximum at a sub OK -lattices Λ1 of Λ, but
(b) the rank of Λ1 is also maximum at a sub OK -lattices Λ01 satisfying that the
slope of L(0, Λ01 ) is equal to the slope of L(0, Λ1 ).
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 15
Semi-Stable Lattice 15
(1) The i-th successive graded OK -lattice gri (Λ) := Λi+1 Λi is semi-stable.
(2) Not only the slope of L 0, gri (Λ) attains the maximum among all sub OK -
lattices of Λ Λi , but the rank of Λi+1 attains the maximum among all sub
OK -lattices Λi+1 satisfying the same conditions.
(3) µ gri (Λ) > µ gri+1 (Λ) . Or equivalently,
Proof. By the restriction of scalars, we may simply work over Z. Let P1 and P2
be the projective Z-modules associated to Λ1 and Λ2 , respectively. Since Λ1 and
Λ2 are sub OK -lattices of Λ, Λ1 ∩Λ2 = Λ1 ∩ P2 = Λ2 ∩ P1 and Λ1 = (Λ1 +Λ2 )∩ P1
are sub OK -lattices of Λ2 and Λ1 + Λ2 , respectively. Here, as usual, Λ1 + Λ2 is a
sub OK -lattice of Λ generated by Λ1 and Λ2 . Consequently, it suffices to prove
vol(Λ1 + Λ2 /Λ1 ) ≤ vol(Λ2 /Λ1 ∩ Λ2 ). (1.20)
Obviously, there exists a full filtration P1 ∩ P2 = P12,0 ⊂ P12,1 ⊂ · · · ⊂
P12,k = P2 , such that, as Z-modules, ((P12,i + P1 )/P1 ) ((P12,i−1 + P1 )/P1 ) '
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 16
(P12,i /P1 ∩ P2 ) (P12,i−1 /P1 ∩ P2 ) are of relative rank one. In terms of lattices,
since (P12,i + (P12,i−1 + P1 )) (P12,i−1 + P1 ) = (P12,i + P1 ) (P12,i−1 + P1 ) and
P12,i P12,i−1 = P12,i (P12,i ∩ (P12,i−1 + P1 ), it suffices to verify that
Here ∗ denotes the lattice induced by ∗. These may be viewed as special cases of
(1.20) if we set Λ1 = P12,i−1 + P1 and Λ2 = P12,i . Hence it suffices to prove (1.20)
in the case of relative rank one. For this, replacing Λ1 and Λ2 by Λ1 Λ1 ∩ Λ2 and
Λ2 Λ1 ∩Λ2 , respectively, we may further assume that Λ1 ∩Λ2 = 0. Accordingly,
let s be a generator of Λ2 and π1 be the orthogonal projection to Λ1 ⊗Z Q. Then
vol(Λ2 ) = |s| ≥ |π1 (s)| = vol Λ2 Λ1 ∩ Λ2 , as desired.
Back to the proof of the proposition. Let Λ1 and Λ01 be two canonical sub OK -
lattices of Λ. By Lemma 1.4, Λ1 + Λ01 is also canonical. Thus, by the maximality
of the rank, Λ1 +Λ01 = Λ1 = Λ01 . This implies that the first canonical sub OK -lattice
of Λ, and hence all levels canonical sub OK -lattices of Λ, are unique.
Semi-Stable Lattice 17
Ll
j=0 Λ /Λ
j j+1
Moreover, up to isometry, the graded lattice Gr(Λ) := is uniquely
determined by Λ.
Even the Jordan-Hölder filtrations play a key role in introducing S-classes for
vector bundles, in this book, it will not be used. We leave its proof to the reader.
Lemma 1.5.
(1) For g ∈ GL(r, R), resp. g ∈ GL(r, C), there is an associated Euclidean metric
ρ(g) on Rr , resp. Hermitian metric ρ(g) on Cr , defined by
hx, yiρ(g) := hg x, g yi ∀x, y ∈ Rr , resp. ∀x, y ∈ Cr .
Here on the right hand side, h·, ·i denotes the canonical pairing.
(2) Each Euclidean metric on Rr , resp. Hermitian metric on Cr , is determined by
h·, ·iρ(g) for some g ∈ GL(r, R), resp. g ∈ GL(r, C).
(3) Let g, g0 ∈ GL(r, R), resp. ∈ GL(r, C). The following two are equivalent.
(a) On Rr , resp. on Cr , h·, ·iρ(g) ≡ h·, ·iρ(g0 ) . n o
(b) There is an orthogonal matrix O ∈ O(n) = O ∈ GL(r, R) OOt = Ir , resp. a
n o
unitary matrix U ∈ U(n) = U ∈ GL(r, C) UU t = Ir , such that g0 = g · O,
resp. g0 = g · U.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 18
Let Λ = (P, (ρσ )) be an OK -lattice of rank r. Since, for t > 0 vol(tΛ) = trn vol(Λ),
where n := [K : Q], we conclude that
(1) The parametrizing space of metric classes over OK -lattices of rank r natu-
rally decomposes into a disjoint union of the subspaces indexed by T ∈ R∗+ ,
parametrizing the metric classes over OK -lattices of rank r and volume T .
(2) For fixed T , T 0 ∈ R∗+ , the above scaling induces a natural bijection between
the sub collection associated to T and the sub collection associated to T 0 .
For this reason, in the sequel, we concentrate on OK -lattices with a fixed volume.
Fix a OK -ideal a = ai , let Λ0 = (P0 , ρ0 ) be the OK -lattice of rank r such that
(i) P0 = O⊕(r−1)X ⊕ a, and (ii) ρ0 is the metric on V(P0 ) corresponding to the image
of (Ir , . . . , Ir ) in GL(r, R) O(n) r1 × GL(r, C) U(n) r2 . Then, by Example 1.1,
r−1 1 r
vol(Λ0 ) = ∆K2 · N(a) · ∆K2 = N(a) · ∆K2 . As usual, set then
be the special orthogonal group and the special unitary group, respectively. Then,
by Proposition 1.8, what we have just said proves the following:
r
vol O⊕(r−1) ⊕ a, ρ = N(a) ∆K2 .
K
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 19
Semi-Stable Lattice 19
There is a natural procedure to reduce parameter spaces from the general linear
group to the special linear group. To explain this, we give a decomposition of
GL(r, C)/U(n) with respect to a fixed branch of r-th root of unity on C defined by
' '
GL(r, C) −→ SL(r, C) × C∗ −→ SL(r, C) × S1 × R∗+
1 1 det g (1.22)
, | det g|
g 7→ pr g, det g 7→ p g,
det g r
det g | det g|
and
!
' 1
U(n) −→ SU(r) × S , 1
U 7→ √r U, det U , (1.23)
det U
where S1 := z ∈ C : |z| = 1 denotes the unit circle in C∗ . As a direct consequence,
GL+ (r, R) := g ∈ GL(r, R) : det g > 0 and O+ (r) := O ∈ O(n) : det g > 0 .
n o n o
Obviously,
In the previous section, a parameter space for the metrics of OK -lattices of rank r is
constructed, using classifications of projective OK -modules and metrics on Rr and
Cr . For arithmetic applications, in this section, we construct a moduli space for
the OK -lattices of rank r, by examining the automorphism groups of the associated
projective OK -modules.
For this purpose, we first study elements of AutOX (O⊕(r−1) K ⊕ a), viewed as a sub-
group of GL(r, K).
Let A ∈ AutOK O⊕(r−1) ⊕ a . Then (i) det A ∈ U K , (ii) A O⊕(r−1) ⊕a ⊆ O⊕(r−1)
K K K ⊕
−1 ⊕(r−1) ⊕(r−1)
a and (iii)
A O K ⊕ a ⊆ O K ⊕ a. As an element in GL(r, K), we write
A = ai j with ai j ∈ K. In this way, by examining how the entires ai j ’s of A act on
lattice vectors x1 , . . . , xn t of O⊕(r−1)
K ⊕ a, we conclude that
arr , ai j ∈ OK
−1
⊕(r−1) ⊕(r−1) air ∈ a, ar j ∈ a
⊕ a = GL OK .
AutOK OK ⊕ a := (ai j ) ∈ GL(r, K) :
i, j = 1, · · · , r − 1
det(a ) ∈ U
ij K
Semi-Stable Lattice 21
determined by
a
.
.
.
\
O
⊕(r−1)
⊕a = .
AutOK OK A ∈ GL(r, K) K : det A ∈ U K
a
a . . . a−1 OK
−1
This is still not enough, since we cannot take the r-th root of units in K. To over-
come this, observe that, if ε ∈ U K is a unit, diag(ε, . . . , ε) ∈ AutOK O⊕(r−1) K ⊕a ,
and its determinant εr belongs to
U Kr := εr : ε ∈ U K .
Lemma 1.8. The quotient groups U K /U K+ and U K+ U Kr,+ are all finite.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 22
Proof. By Dirichlet’s unit theorem, the image log kU K k of U K (under the natural
logarithm-norm map) is a Z-lattice of rank r1 + r2 − 1 in Rr1 +r2 . Moreover, kU Kr k =
rkU K k, consisting of elements which can be written as the r-powers of elements
in kU K k. Therefore, rankZ kU Kr k = rankZ kU K k. Consequently, up to a finite torsion
subgroup consisting of the roots of unity in K, U Kr is a finite index subgroup of
U K . But U Kr,+ ⊂ U Kr ⊂ U K . This also shows that U Kr,+ is a finite index subgroup of
U K since U K2r ⊆ U Kr,+ .
Aut+OK O⊕(r−1) ⊕ a n U K+ .
K ⊕ a ' SL O⊕(r−1)
K (1.28)
Semi-Stable Lattice 23
Now, we define kU Kr,+ k := { kεk ∈ R∗+ : ε ∈ U Kr,+ } to be the norm group of U Kr,+ .
Note that this eliminates possible ambiguity caused by the roots of unity contained
in K. As a direct consequence from the above discussion, we have (proved) the
following:
This map transforms the multiplicative space (R∗+ )r1 +r2 into the additive space
Rr1 +r2 isometrically. By Dirichlet’s unit theorem, the image log kU K k of the unit
group U K of K is a full rank Z-lattice of the metrized hyper-plane
X
H := (xσ )σ∈S∞ ∈ Rr1 +r2 : [Kσ : R] xσ = 0 .
σ∈S ∞
Consequently, by Lemma 1.9, the image log kU Kr,+ k of the subgroup U Kr,+ is a full
rank sub Z-lattice in H as well. We denote by Dr,+ K a fundamental parallelepiped
of the lattice log kU Kr,+ k in H.
Similarly, let S be the norm-one hypersurface in (R∗+ )r1 +r2 defined by
Y
S := (xσ )σ∈S∞ ∈ (R∗+ )r1 +r2 : xσ[Kσ :R] = 1 . (1.30)
s∈S ∞
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 24
This decomposition is very important. Indeed, for a fixed T ∈ R∗+ , the isometry
classes of OK -lattices of rank r and volume T form a natural subspace of that for
all lattices. Denote by Mtot
K,r [T ] the moduli space of isometry classes of OK -lattices
of rank r and volume T .This subspace can be read from the above decomposition
easily, since the final factor R∗+ represents this volume parameter. That is to say,
we have the following:
Here R+K is the narrow regulator of K (to be defined at the end of the proof below).
Proof. Let F+K be the fundamental domain of the unit sub-group U K+ in S. The log-
d(x)
arithm log k k : (R∗+ )r1 +r2 → Rr1 +r2 = σ∈S∞ R transforms
Q
into the Lebesgue
(x)
measure d(x) on Rr1 +r2 . Hence,
Semi-Stable Lattice 25
where I := [0, 1] denotes the unit interval. By Dirichlet’s unit theorem, there exist
ε1 , . . . , εr1 +r2 −1 such that U K+ , modulo roots of unity, is generated by them. Thus,
the full rank sub lattice log |F+K × I| is generated by (r1 + r2 − 1) vectors
t
e j := log |σ1 (εi )|σ1 , . . . , log |σr1 (ε j )|σr1,log |σr1 +1 (ε j )|2σr1 +1 , . . . log |σr1 +r2 (ε j )|2σr1 +r2
Since the ε j ’s are units, adding the first r1 +r2 −1 lines to the last one, all entries of
r1 +r2
1 X
the new line becomes 0 except the last one, which is ep = 1. By definition,
n i=1 i
the absolute value of the determinant of this latest matrix is called the narrow
regulator of K and denoted by R+K . Therefore, we have vol(F+K ) = R+K . This
completes the proof since |U K+ /U Kr,+ | = µ+ (r, K).
(1) pΛ (0) = 0.
(2) pΛ is affine over the closed interval rankOK (Λi ), rankOK (Λi+1 ) , and
(3) pΛ (rankOK Λi ) = degar (Λi ) − rankOK (Λi ) · µ(Λ).
Obviously, the canonical polygon is well-defined on the space Mtot K,r and its sub-
∗
spaces Mtot
K,r [T ] for a fixed T ∈ R+ .
Recall also that, for a positive real number t, the scaling assignment Λ 7→ tΛ
induces a natural isomorphism between the moduli spaces
Mtot tot r[K:Q]
K,r [T ] ' MK,r [t T] ∀T ∈ R∗+ .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 26
For this reason, when necessary, we fix the volume, say, to be one, for the lattices
to be considered.
On the other hand, by Theorem 1.2, the moduli spaces of OK -lattices of rank
r are not compact, because of the existence of parabolic cusps. To overcome the
problem, we have introduced the stability condition.
Theorem
n 1.3.h Letip : [0, r] →o R be a fixed convex polygon. Then,
h r/2the
i subspace
[Λ] ∈ MtotK,r ∆ r/2
K : p Λ ≤ p is a compact subspace of M tot
K,r ∆ K . Inh partic-
i
ular, the moduli spaces of semi-stable OK -lattices of rank r and volume ∆r/2 K is
compact.
For later use, for T > 0, denote by MK,n [T ] the moduli space of OK -lattices
of rank r and volume T .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 27
Chapter 2
Geometry of Numbers
Let K be an algebraic number field. Denote by OK its ring of integers. Let S fin ,
resp. S ∞ , be the set of inequivalent normalized non-Archimedean places, resp.
Archimedean places, of K, and set S = S fin S ∞ . For each v ∈ S , write Kv the
F
v-completion of K, and, for each v ∈ S fin , denote by Ov the ring of integers in Kv .
Finally, let r1 , resp. r2 , be the number of real, resp. complex, places of K in S ∞ .
The adelic ring A = AK :=
Q0
v∈S Kv of K is defined to be the restricted
Q
product of the Kv ’s with respect to the Ov ’s, i.e. the collection of (xv )v∈S ∈ v∈S Kv
satisfying xv ∈ Ov for almost all v ∈ S fin . Set Afin := 0v∈S fin Kv and call it
Q
the ring of finite adeles of K and set A∞ := σ∈S ∞ Kv = Rr1 × Cr2 . There is a
Q
natural topology on A, and hence on Afin , generated by the basis of open subsets
Q Q
v∈I Uv × v<S Ov , where I ⊂ S are finite and contains S ∞ and Uv ⊂ Kv are open.
27
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 28
As usual, denote by GLn (A) the adelic general linear group, consisting of ele-
Q
ments (gv )v∈S of v∈S GLn (Kv ) satisfying gv ∈ GLn (Ov ) for almost all v ∈ S fin .
For convenience, we also write elements (gv )v∈S of GLn (A) as (gp .gσ )p∈S fin ,σ∈S ∞ or
simply (gp .gσ ) if there is no confusion.
Geometry of Numbers 29
Proof. Denote by φ : An∞ → An /(An (g) + K n ) the morphism induced from the
embedding A∞ ,→ A. It is sufficient to show that
(i) φ is surjective,
and
(ii) Ker φ = i Hfin0
(K, g) .
(ii) trivially comes from the definition of H 0 (K, g). To prove (i), we show the
equivalent condition that Anfin ⊆ An (g) + K n . But this is a direct consequence
of Proposiiton 2.1(4). Indeed, for any g = (gp .gσ ) ∈ An , there exists a finite set
S 0 ⊆ S fin such that gq ∈ GLn (Oq ) for q < S 0 . But, by the strong approximation
theorem, there exists a ∈ K n such that gp (ap ) − a ∈ Onp , for all p ∈ S 0 , and that
a ∈ Onq for all q < S 0 . This proves the lemma and hence also the proposition.
There is a canonical duality between H 0 and H 1 . To see this, we use the pairing
h·, ·i : A × A → R/Z ' S1 ⊂ C∗
(2.2)
(x, y) 7→ v∈S hxv , yv iv .
P
n
Y √
Here hxv , yv iv := e2π −1·χv (xv,i ·yv,i )
with χv := λv ◦ TrFv /Qv , and
i=1
Qv Qv /Zv ,→ Q/Z ,→ R/Z, v = p ∈ S fin
λv :=
(2.3)
R R/Z,
v ∈ F∞ .
bn ' An (as
By Tate [113], this global pairing induces a natural isomorphisms A
n ⊥ n
locally compact groups) and (K ) ' K (as discrete subgroups). On the other
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 30
hand, for the local pairings h·, ·ip of finite places p, by a direct local calculation
from the definition (see e.g. [113]), we have a⊥p ' a−1 p · ∂p , where ∂p denotes
the local differential module at p (dual to Op ). Put this globally, denote by κF
−ord (∂ )
the canonical idele πp p p .1 , where πp denotes a uniformizer of Op , Then the
above proof shows that the dual lattice Λ(g)∨ of Λ(g) is given by Λ(g)∨ ' Λ(κF ·
g−1 ). In addition, since H 0 is insensitive when local or global units are multiplied,
the cohomology group H 0 (K, κF · g−1 ) is well-defined since it does not depend on
the πp ’s used. Thus by (2.1), there is an isomorphism
Y Y .
H 1 (K, g) ' Rn × Cn , ρ(g∞ ) Λ(g).
σ:R τ:C
Consequently, if we denote by X
b the Pontryagin dual of a locally compact space
X, what we have said proves the following:
2.2.1 Nine-Diagram
H 0 (Kp , g) := {x ∈ Kpn : g · x ∈ Op }.
0 0
[g1 ] ≤ [g2 ] if and only if Hfin (F, g1 ) ⊆ Hfin (F, g2 ).
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 31
Geometry of Numbers 31
For g ∈ GLn (A), there is a 9-diagram with exact rows and columns given by
0 0 0
↓ ↓ ↓
0→ H 0 (K, g) → An (g) → An (g)/An (g) ∩ K n → 0
↓ ↓ ↓
0→ K n
→ An
→ A /K n
n
→0
↓ ↓ ↓
0 → K n /An (g) ∩ K n → An /An (g) → H 1 (K, g) →0
↓ ↓ ↓
0 0 0.
Consequently, for any two elements g = (gp .gσ ) and g0 = (g0p .g0v ) satisfy the
condition that gv = g0v ∀v ∈ S ∞ and [(gp )] ≤ [(g0p )], since the middle row remains
unchanged in the corresponding 9 diagrams, there exists a long exact sequence
With g∞ fixed, the quotient An (g0 ) An (g) measures only differences at finite
places. Namely degfin (g0 ) − degfin (g0 ) = deg(g0 ) − deg(g0 ), where
X
degfin (g) := degfin (det g) = ordp (detgp ) · Np log p.
p
This then suggests that, to obtain an arithmetic Riemann-Roch theorem, the arith-
metic counts hi (i=0,1) of H i should satisfy the following axioms.
In particular, by examining the diagonal elements of GLn (A), which are also
elements of n-copies of GL1 (A), it is natural to introduce
∆(K, n) = n · ∆(K).
Here 1Onv denotes the characteristic function of Onv in Kvn , and | · |can,σ denotes the
norm induced by the canonical metric for each of the Archimedean place σ of K.
1Onv (gv · xv )
X Y Y 2
X Y 2
= e−πNσ |gσ ·xσ |can,σ = e−πNσ |gσ ·xσ |can,σ .
x∈H 0 (K,g) v∈S fin σ∈S ∞ x∈H 0 (K,g) σ∈S ∞
Geometry of Numbers 33
X √ P
−1 v∈S χv (ξv ηv ) b
f (η) := e−2π f (ξ), ∀η ∈ H 1 (K, g).
c
b
ξ∈H[
1 (K,g)
f |H 0 (K,g) ∈ L2 H 0 (K, g) , 2
(K, g) ,
H 1[ f ∈ L2 H 1 (K, g) .
1 (K,g) ∈ L
f H[
b c
b
Here, as usual, for a measurable space X, L2 (X) denotes the associated space
of L2 -functions. Accordingly, we introduce the arithmetic counts for the groups
H 0 (K, g) and H 1 (K, g) with respect to the test function f by setting
Z
2 2
X
0
f (x) dx = f (x) ,
#Ar, f H (K, g) :=
H 0 (K,g) x∈H 0 (K,g)
Z
2 2
X
#Ar, f H 1[ f (y) dy = f (y) ,
(K, g) := b b
1 (K,g)
H[ 1 (K,g)
y∈H[
Z
2
H 1 (K, g) :=
#Ar, f f (z) dz.
c
b
H 1 (K,g)
[
1
[
That is to say, we count H 1 (K, g) by the numeration function e(∗) := eA2 (g∗) ,
2
1
since 1Onp = 1O2 n .
p
This, together with arithmetical duality H 1 (K, g) ' H 0 (K,[κF · g−1 ), implies that
ecA (g−1 a) = #Ar H 0 (K, κF · g−1 ),
X
(K, g) =
#Ar H 1[
1 (K,g)
a∈H[
To verify Axiom 0, we use the Tate Riemann-Roch theorem ([113]), obtained from
the Poisson summation formula for adelic spaces K n ,→ An . Namely,
1
eA (g(a)) = eA (g−1 (a)).
X X
· b
a∈K n kdetgk a∈K n
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 35
Geometry of Numbers 35
Similarly, we define
X
Nσ |x|2ρσ
#Ar H 0 (K, Λ) := .
P
e−π σ∈S ∞
x∈H 0 (K,Λ)
This coincides with k0 (F, Λ) in [32] obtained using arithmetic effectivity. Set
h0 (K, Λ) := log k0 (F, Λ). (2.4)
1
To define h , we use the counting function e(∗) on H (K, Λ) obtained using the
1
P 2
Fourier transformZfrom the function e−π σ∈S ∞ Nσ |∗|ρσ . That is
#Ar H 1 (K, Λ) := e(x)dµ(x) and h1 (K, Λ) := log #Ar H 1 (K, Λ).
x∈H 1 (K,Λ)
As a direct consequence of the Plancherel theorem, the Poisson summation for-
mula, and the Arakelov Riemann-Roch theorem, we have the following:
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 36
We recall some elementary results concerning h0 for lattices following [32, 35].
Specify now f (x) = e−πhx,xi , then bf = f . Easily, the sum on the right hand side of
(2.5) is maximal if e2πihv,yi = 1 for all y. This implies that v ∈ Λ.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 37
Geometry of Numbers 37
Definition 2.6. Let Λ be a full rank Z-lattice in an Euclidean space V = (V, h·, ·i).
(1) The first Minkowski successive minimum λ1 (Λ) of Λ is defined to be the min-
imal length among the lengths of all non-zero vectors in Λ. Namely,
λ1 = λ1 (Λ) := min kxk : x ∈ Λ, x , 0 .
Proof. It suffices to prove the first two since (3) then comes automatically.
(1) For a positive real number t and an element x0 in V, set
Corollary 2.2.
√
(1) Let Λ be a Z-lattice of rank n such that λ1 (Λ) ≥ n. Then
3n π 2
#H 0 (Q, Λ) ≤ 1 + e−πλ1 (Λ) .
π − log 3
(2) Let K be a number field of degree n over Q and let Λ be a rank one OK -lattice
satisfying degar (Λ) ≤ 0. Then
3n π − 2 degar (Λ)
h0 (K, Λ) ≤ 1 + e−πn e n .
π − log 3
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 39
Geometry of Numbers 39
√
Proof. (1) Write λ for λ(Λ). If t ≥ 1, 2 t + 1 ≤ 3t . Hence, by Lemma 2.3(3),
Z ∞
2
#H 0 (Q, Λ) − 1 ≤ πλ21 3nt e−πλ1 t dt
1
∞ 3n πλ21 3n π
Z
e(−πλ1 +n log 3)t dt =
2 2 2
= πλ21 e−πλ1 ≤ e−πλ1 .
1 πλ21 − n log 3 π − log 3
kxkσ
Here Nσ appearing in comes from the difference between canonical metric
Nσ
and Lebesgue metric. Hence, by the geometric-arithmetic mean inequality,
!2 1
X X kxk σ
Y kxkσ !2Nσ n
kxk2 := kxk2σ ≥ Nσ ≥ n
σ∈S ∞ σ∈S ∞
Nσ σ∈S ∞
Nσ (2.7)
2
2
= n #(Λ OK x)e− degar (Λ) n ≥ n e− n degar (Λ) .
This gives (2.6), and hence (2) since h0 (K, Λ) ≤ #H 0 Q, ResK/Q (Λ) − 1.
(2) a is called ample if for each g = (gp .gσ ) ∈ GLn (A) and for a sufficiently
positive m, the unit ball B0 (1) centered at 0 of the metrized space Rr1 ×
Cr2 n , ρ(am 0 m
∞ · g∞ ) contains a basis of the lattice H (K, a · g) whose elements
are positive at finite places.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 40
(1) a is ample.
(2) a is positive.
(3) lim h1 (K, am · g) = 0 for all g ∈ GLn (A).
m→∞
Geometry of Numbers 41
Now, fix a minimal vector e(−mk ) ∈ H 0 (K, a−mk · g) such that ke(−mk )k =
λ1 (−mk ), and denote by Λ1 (−mk ) the rank one OK -sublattices of H 0 (K, a−mk · g)
generated by e(−mk ). Then, by the biggest slope and highest rank property of
H 0 (K, a−m ) · Λ1 in the canonical filtration of h0 (K, a−mk · g), we conclude that
deg H 0 (K, a−mk ) · Λ1
deg Λ1 (−mk ) ≤ .
rank H 0 (K, a−mk ) · Λ1
Consequently,
lim deg Λ1 (−mk ) = −∞.
k→∞
Hence, by applying Proposition 7.1 of [35], or the same, estimation (2.6), to the
rank one Λ1 (−mk ), we see that
λ1 (−mk ) = ke(−mk )k ≥ [K : Q] · e− [K:Q ] ·deg Λ1 (−mk ) .
2
Chapter 3
3.1.1 Stability
Let K be a number field with OK the integer ring and ∆K the absolute value of the
discriminant. Denote by MK,n (d) the moduli space of semi-stable OK -lattices of
rank n and degree d. By the Arakelov-Riemann-Roch theorem,
n
− log vol(Λ) = degar (Λ) − log ∆K ,
2
h n/2 i
MK,n (d) is simply the moduli space MK,n ∆K e introduced in the first chapter,
−d
Here a1 , . . . , ah denote by integral OK -ideals such that the class group CLK of K is
given by CLK = [a1 ], [a2 ], . . . , [ah ] , Fr,+
K is the fundamental domain of the unit
group U Kr,+ := U Kr ∩ U K+ , and the lower index ss stands for the collections of the
modular points corresponding to the semi-stable OK -lattice of rank r (and volume
one). In addition, by Theorem 1.3, MK,n (d) is compact.
More generally, By Theorem 1.1, the moduli space Mtot K,n of isometric classes
of OK -lattices of rank n admits a natural modular representation
G
Fn,+ ⊕(n−1)
⊕ a) SL(n, R) SO(n) r1 × SL(n, C) SU(n) r2 ss × R∗+ .
K × SL(OK
a∈{a1 ,...,ah }
Example 3.1. If n = 1, the stability condition is automatic, and the moduli spaces
d
Mtot
K,n and MK,n (d) coincide with the Arakelov-Picard groups Pic(K) and Pic (K)
(see e.g [66]). It is well known that Picd (K) is compact.
43
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 44
Similarly, there are subspaces MK,n (d)<T , MK,n (d)≥T , MK,n (d)>T and MK,n (d)=T .
By the compactness of MK (n, d), we have the following:
Proposition 3.1. For (n, d) ∈ Z>0 × Z, there exist MK (n, d), mK (n, d) ∈ R such
that MK,n (d)<m = MK,n (d)>M = ∅ unless mK (n, d) ≤ n and M ≤ MK (n, d).
= max h0 (K, Λ∨ ⊗ ωK ) : Λ ∈ MK n, d
n n o
= max h0 (K, Λ) + log ∆K − d : Λ ∈ MK n, d
2
n
=MK (n, d) + log ∆K − d ,
2
as desired.
Proof. By Theorem 2.2(1), it suffice to establish (1). For any non-zero vector x
of Λ, denote by L(x) = OK · x, the OK -lattice generated by x in Λ. Then, by
2
estimation (2.6), kxk2 ≥ [K : Q] · e− [K:Q] ·degar (L(x)) . On the other hand, since Λ is
semi-stable,
degar (Λ) log n
degar (L(x)) ≤ ≤ −[K : Q] · .
n 2
Consequently, kxk ≥ λ1 ≥ [K : Q] · n = rankZ Λ. Since the OK -lattice Λ, viewed
2
Proof. The first and the second statements are obvious. Recall that, from the
definition and the Arakelov Riemann-Roch theorem, MK (n.d) − d − n2 log ∆K =
max h1 (K, Λ) : Λ ∈ MK,n (d) . Hence the third inequality is a direct consequence
of Theorem 3.1(2). Finally, to prove limd→∞ ∆K (n.d) = 0, it suffices to notice that,
for a fixed ε, the constant dK (n, ε) chosen for the three inequalities is effective.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 46
3.2.1 Definition
Let K be an algebraic number field. Denote by OK its ring of integers and denote
by ∆K the absolute value of its discriminant. For n ∈ Z>0 , denote by MK,n :=
F
d∈R MK,n (d) the moduli space of semi-stable OK -lattices of rank n. Denote by
dµ the natural measure on MK,n .
Definition 3.1. The rank n non-abelian zeta function bζK,n (s) of K is defined by
n s Z 0 deg(Λ)
ζK,n (s) := ∆K2 ·
b eh (K,Λ) − 1 · e−s dµ(Λ) <(s) > 1. (3.2)
Λ∈Mtot
K,n
Like classical zeta functions, such as the Dedekind zeta functions, these rank n
ζK,n (s) satisfy the standard zeta properties.
non-abelian zeta functions b
Proof. There are two proofs presented in this book. Namely, one proof given here
uses the arithmetic cohomology theory, and the other proof, to be given in the next
part, uses the theory of Eisenstein series.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 47
The assertion (0) is proved in Proposition 1.2. To prove the rest, denote by
MK,n [≥ T ], resp. MK,n [≤ T ], the moduli spaces of rank n semi-stable OK -lattices
of volumes ≥ T , resp. ≤ T . Obviously,
[ G
MK,n = MK,n [≤ T ] MK,n [≥ T ] = MK,n [T ].
T >0
Moreover, if we denote by dµT the natural induced volume form on MK,n [T ], then
dT
dµ = · dµT . Accordingly,
T Z 0
ζK,n (s) =
b eh (K,Λ) − 1 · vol(Λ) s · dµ(Λ)
Λ∈Mtot
K,n
Note that there is a natural isomorphism between MK,n [1] → MK,n [T ], say by
log T
sending Λ 7→ Λ ⊗ Λ0 for a fixed rank one lattice Λ0 of degree . Hence
n
Z 1 1
dT
α(s) = Ts · vol MK,n [1] = vol MK,n [1] · .
0 T s
Similarly,
Z 1
β(s) = vol(Λ)1−s · dµ(Λ) = −vol MK,n [1] · .
Λ∈MK,n [≥1] 1−s
From all these, we have
ζK,r (s) = I(s) + I(1 − s) − α(s) + β(s)
b
!
1 1
= I(s) + I(1 − s) + vol MK,n [1] · − .
s−1 s
This completes the proof since I(s) is a holomorphic function in s.
Let a1 = OK , a2 , n. . . , ah be integral
o OK -ideals such that the ideal class group CLK
of K is given by [a1 ], . . . , [ah ] . From the classification of projective OK -modules
in Corollary 1.1, for a rank n projective OK -module P, there exists one and only
one a ∈ a1 , . . . , ah such that P is isomorphic to Pn.a = O⊕(n−1)
K ⊕ a.
Let MK,n be the total moduli space of OK -lattices of rank n, and let Mtot
tot
K,a,n be
the moduli space of OK -lattices of rank n with underlying projective OK -module
Pn.a . From above,
h
G
K,n =
Mtot K,ai ,n .
Mtot
i=1
Similar to non-abelian zeta functions, (with almost the same proof we con-
clude that) these partial zeta functions satisfy the standard zeta properties with
a modified functional equation. Indeed, for integral OK -ideals a, b, if [a] = [b]
as ideal classes, then b ζK,n.a (s) = b
ζK,n.b (s). Hence, we may generalize the above
ζK,n.a (s) for an integral OK -ideal a to the case when a is a fractional
definition of b
OK -ideal. Under this modification, it is then rather easy to check that
ζK,n.a (1 − s) = b
b ζK,n.dnK a−1 (s),
since dual lattices of the OK -lattices with underlying projective module O⊕(n−1)
K ⊕a
have d(n−1)
K ⊕ d K a−1
as their underlying projective modules. Here d K denotes the
differential ideal of K. All this then gives a proof of the following:
Lemma 3.2. Let K be a number field and let a be OK -ideal of K. In addition, let
a1 , . . . , ah be integral OK -ideals such that the ideal class group CLK of K is given
by CLK = [a1 ], . . . , [ah ] . Then,
Xh
(0) b ζK,n (s) = bζK,n.ai (s) and b
ζK,a,n (s) = b
ζK,b,n , where b ∈ [a] is an OK -ideal.
i=1
(1) (Meromorphic Continuation) ζK,a,n (s) is a well-defined holomorphic function
b
in s when <(s) > 1, and admits a meromorphic continuation, denoted also by
ζK,a,n (s), to the whole complex s-plane.
b
(2) (Functional Equation) b ζK,a,n (1 − s) = b
ζK,dnK a−1 ,n (s).
(3) (Singularities & Residues) As a meromorphic function over the complex s-
plane, b ζK,a,n (s) admits only two singularities, namely, two simple poles at
s = 0, 1 with the residues ±vol MK,a,n [1] .
By definition,
X X X
h0 (K,Λ)
−1= kxσ kρσ .
2 2
e exp −π kzσ kρσ − 2π
z∈Λr{0} σ:R σ:C
a point z in Λ maps to the corresponding point (zσ ) in (Rn )r1 ×(Cn )r2 , whose norms
are given by kzσ kρσ = kgσ zσ k. Here, we have assumed that the metric ρσ is defined
by gσ · gtσ for a certain gσ ∈ GL(n, R) when σ is real, and by gσ · ḡtσ for a certain
gσ ∈ GL(n, C) when σ is complex.
Recall that kgσzσ k is O(n), resp. U(n) invariant for real, resp. complex, σ.
0
Hence vol(Λ) and eh (K,Λ) − 1 · (e−s )− log vol(Λ) are well-defined over the space
r1 r2
GL(n, R)/O(n) × GL(n, C)/U(n) .
Moreover, there is a scalar operation Λ 7→ tΛ for t > 0, which keeps the un-
derlying projective module unchanged, but rescale the metrics by a constant t.
Accordingly,
X X X
h0 (K,tΛ)
e −1= exp − π 2
kxσ kρ − 2π kxσ kρ · t ,
σ
2 2
σ
x∈Λr{0} σ:R σ:C
the volume vol(tΛ), resp. the volume form dµ(Λ), is changed to tn[K:Q] vol(Λ),
dt
resp. · dµ1 (Λ1 ). Here dµ1 (Λ1 ), or simply dµ(Λ1 ) if there is no confusion,
t
denotes the corresponding volume form on the moduli space of semi-stable OK -
lattices of rank n corresponding to the points in MK,a,n 1 . This later space is
simply isomorphic to
/ r1 r2
(n−1)
Fn,+
K × SL(O K ⊕ a) SL(n, R)/SO(n) × SL(n, C)/SU(n) × {1} . (3.3)
ss
h n i
In the sequel, we identify the moduli space MK,a,n N(a) · ∆K with this realization
2
Z ∞
dtB 1 s
s
Therefore, by applying Mellin transform e−At t s= · A− B · Γ (when-
0 t B B
ever the two sides make sense), we obtain that, for <(s) > 1,
ζK,n.a (s)
b Z
n s
= Rn,+
K dµ1 (Λ)
N(a)∆K2 Λ∈MK,a,n 1
X Y − ns2 ns Y − ns2 !
× πkxσ kρσ Γ · 2πkxσ kρσ Γ(ns)
x∈Λr{0} σ:R
2 σ:C
ns ns r r
Z X 1
1 −ns 2
= Rn,+K π− 2 Γ 2π Γ(ns)
ns
dµ(Λ).
2 Λ∈MK,a,n 1 x∈Λr{0} kxkΛ
Here Rn,+
K denotes the (co)volume of the unit lattice U Kn,+ := U Kn ∩ U K+ .
Theorem 3.3 (Weak Riemann Hypothesis). For n ≥ 2, all but finitely many
1
ζK,n (s) lie on the central line <(s) = .
zeros of b
2
For details, please refer to Parts 4, 5 and 6.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 53
PART 2
53
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 55
Chapter 4
In this chapter, for a number field K, with r1 real embeddings and r2 complex
embeddings, and a fractional ideal a of K, we first define a fractional action of
the modular group SL(a ⊕ OK ) on the hyperbolic space H r1 × Hr2 , where H,
resp. H, denotes the complex upper half plane, resp. the hyperbolic 3 space.
Then we establish a natural correspondence between the associated cusps and
ideal classes, and construct a fundamental domain for the stabilizer subgroup of
a cusp in SL(a ⊕ OK ). Finally, we introduce a notion of distances to cusps, and
construct a fundamental domain for the action SL(a ⊕ OK ) on H r1 × Hr2 .
To make the constructions and structures here more transparent, most of the
sections are divided into three layers. To be more precise, the first reviews the
classical theory for the action of Fuchsian groups of first kind on the upper half
plane, the second recalls Siegel’s famous theory on the actions of the modular
group SL(a ⊕ OK ) associated to totally real field K on H r1 , and the final develops
a new theory for the action of the modular group SL(a⊕OK ) associated to a general
number field K on the hyperbolic space H r1 × Hr2 . We claims neither credits nor
the materials for the theories and constructions appeared in the first two layers,
which mainly follows Kubota ([62]) and Siegel ([101]), respectively.
In this section, we first recall some basic facts on the upper half plane and the
hyperbolic 3-space and their associated modular actions. Then we construct a
hyperbolic parameter space for all isometric classes of rank two OK -lattices.
55
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 56
" #
0
Back to H itself. If M is a stabilizer of i = ∈ H in SL(2, R), then
1
ai + b
!
ab
= i. This implies that a = d, b = −c, or the same, ∈ SO(2). Hence,
ci + d cd
the stabilizer group of i in SL(2, R) is equal to SO(2). Moreover, for (x, y) ∈ H,
! √ !
1 x y 0 1x
if M := · ∈ SL(2, R), then M(i) = (i y) = x + iy. Hence, the
01 0 √1y 01
action of SL(2, R) on H is transitive. Therefore, there is a natural identification
between the quotient SL(2, R)/SO(2) and H:
'
SL(2, R)/SO(2) −→ H
(4.7)
[g] 7→ g · (i).
Similar to the upper half plane model in plane hyperbolic geometry, the follow-
ing upper half space H in Euclidean 3-space R3 is a convenient model for 3-
dimensional hyperbolic space
Similarly to the upper half plane H, the upper half 3-space H also admits a
quotient space interpretation. Indeed, for the action of SL(2, C) on H above, the
stabilizer group of j = (0, 0, 1) ∈ H in SL(2, C) is equal to SU(2), and this action
of SL(2, C) on H is transitive. Therefore, there is a natural identification between
the quotient SL(2, C)/SU(2) and H:
'
SL(2, C)/SU(2) −→ H
(4.17)
[M] 7→ M · ( j).
M · SU(2). Hence, we can parametrize the image of SL(2, C) SU(2) in P+C by the
.
|z|2 +r2
z
t
matrix Y = M · M = rz̄ 1r . Therefore, the above identification can be more
r r
explicitly written as
! √ !
1 z r 0 1z
M 7→ P M := ( j) = (r j) = z + r j ∈ H. (4.18)
0 √1r
01 01
In other words,
ad − bc = 1,
!
ab
a, d ∈ OK , b ∈ a,
∈ SL(OK ⊕ a) if and only if (4.20)
cd
c ∈ a−1 .
By Theorem 1.1, all isometry classes of rank two OK -lattices are parametrized
by
G
F+K × SL(OK ⊕ a) SL(2, R)/SO(2) r1 × SL(2, C)/SU(2) r2 × R∗+ .
a∈{a1 ,...,ah }
Theorem 4.1. All isometry classes of rank two OK -lattices are parametrized by
G
F+K × SL(OK ⊕ a) H r1 × Hr2 × R∗+ .
(4.22)
a∈{a1 ,...,ah }
Put this in a more concrete term, for rank two OK -lattices Λ = (OK ⊕ a, ρ),
when the volumes of Λ’s are fixed, the associated metrics ρ = (ρσ ) are repre-
sented by the matrices g = (gσ )σ∈S∞ with gσ ∈ SL(2, Kσ ). Moreover, by (4.7) and
(4.17), the rank two Ok -lattice Λ associated to a fixed g corresponds uniquely to
the equivalence class of the point g(ImJ) = (gσ τσ )σ∈S∞ ∈ H r1 × Hr2 , where
iσ := (0, 1)
σ real
ImJ := ( i, . . . i, j, . . . , j ) τσ =
and (4.23)
|{z} | {z } jσ := (0, 0, 1) σ complex.
r1 r2
For this reason, in the rest of this chapter, we study only the quotient space
SL(OK ⊕ a) H r1 × Hr2 .
A proof of this classical result can be easily found in literatures, and hence
will be omitted.
As a direct consequence, there is the following elegant result on minimal gen-
erators of a fractional ideals of an algebraic number field, even many standard
textbooks on algebraic number theory omit it.
Lemma 4.1. Let a, b be two fractional ideals of K. There exist two elements
α, β ∈ K such that
OK · α + a · β = b. (4.24)
Proof. We first reduce the proof to the case b is an integral OK -ideal. Indeed,
by definition, there exist b ∈ K and an integral OK -ideal b0 such that b = b · b0 .
Hence, if there exist two elements α0 , β0 ∈ K such that OK · α0 + a · β0 = b0 , then
b = b · b0 = OK · (bα0 ) + a · (bβ0 ). Therefore, α = bα0 and β = bβ0 would do the job
for the original a, b.
Similarly, we can reduce the proof to the case a is an integral OK -ideal. Indeed,
by definition, there exists an a ∈ K ∗ such that a · a =: a0 is integral. Thus, if there
exist two elements α0 , β0 ∈ K such that OK ·α0 +a0 ·β0 = b, then b = OK ·α0 +a0 ·β0 =
OK · α0 + a · (a−1 β0 ). This implies that α = α0 and β = a−1 β0 would do the job for
the original a, b.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 62
Hence, from now on, we may and will assume that both a, b are integral OK -
ideal. Choose then β ∈ a−1 b\{0} so that a · β ⊆ a · a−1 b = b. By the unique factor-
ization theorem of integral OK -ideals into product of prime ideals for Dedekind
domain OK (see e.g. [66]), there exist distinct prime ideals pi ’s and integers ni ’s
and mi ’s satisfying 0 ≤ ni ≤ mi (1 ≤ i ≤ l) such that
Yl Yl
b= pni ⊇ a · β = pmi .
i=1 i i=1 i
Theorem just cited, there exists an element α ∈ OK such that α ≡ bi mod pni i +1
(1 ≤ i ≤ l). Since, in terms of local orders at each pi , νpi (α) = νpi (b), we conclude
that α ∈ b. Consequently, OK · α + a · β ⊆ b.
On the other hand, if p is a prime ideal of OK which does not lie in the collec-
tion p1 , . . . , pl , then 0 = νp (β) = νp (b). This means that, for all primes p of OK ,
νp (OK · α + a · β) = inf νp (OK · α), νp (a · β) = νp (b). Therefore, OK · α + a · β = b.
α α∗
!
(1) ∈ SL(2, K), and
β β∗
(2) OK β∗ + a−1 α∗ = b−1 .
α α∗ α
! " #! " #
1
1
In particular, when acting on ∞ ∈ P (K), we have = .
β β∗ 0 β
that
(OK β∗ + a−1 α∗ ) · (OK α + aβ) = OK · β∗ α + a−1 · α∗ α + aa−1 · α∗ β + a · ββ∗ .
= OK .
This then proves (2).
Before treating general number fields, we here give an example treated by Siegel
in [101]. Let K be a number field of degree n over Q. K is called to-
tally real if all Archimedean places of K are real. In this case, we denote by
K = K, K (2) , . . . , K (n) the collection of all conjugates of K. Thus, if we set
(1)
Hb := H t P 1 (R), there is a natural embedding
P1 (K) ,→ P1 (R) n ⊂ H bn , a 7→ (a(1) , · · · , a(n) ) =: a.
(4.25)
" #
a
In the sequel, for a point µ ∈ P 1 (K), we choose a, b ∈ OK such that µ = . In
b
a 1
this case, for simplicity, we also write µ = . For example, ∞ = .
b 0
αβ
!
Let G = SL(2, K), Z = {±I} and M = ∈ G. For z = (z1 , . . . , zn ) ∈ H bn ,
γδ
α( j) z j + β( j)
we set z M := (z∗1 , . . . , z∗n ), where z∗j := ( j) for all j = 1, . . . n. This induces
γ z j + δ( j)
a natural action of G, and hence an action of G/Z, on H bn defined by
G× H bn −→ H b n
(4.26)
(M, z) 7→ z M .
For simplicity, if z j = x j + iy j and z∗j = x∗j + iy∗j with x j , y j , x∗j , y∗j ∈ R, we set
x := (x1 , . . . , xn ), y := (y1 , . . . , yn ), x M := (x1∗ , . . . , xn∗ ), y M := (y∗1 , . . . , y∗n ).
Hence, if z = x + iy ∈ H bn , then z M = x M + iy M ∈ H bn . In addition, we
αz + β
" #
1
symbolically denote z M by . For example, in the case ∞ = , we have
γz + δ 0
α(1) α(n)
!
∞M = , . . . , . Accordingly, by the calculations in §4.1.1, we have
γ(1) γ(n)
z M M2 = (z M ) M ∀z ∈ H bn , ∀M1 , M2 ∈ G. (4.27)
1 2 1
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 64
αβ
!
Let now Γ := SL(2, OK ) be the subgroup of G consisting of elements
γδ
where α, β, γ, δ ∈ OK . The quotient group ΓK := Γ/Z is traditionally called narrow
Hilbert modular group.
By definition, two elements!µ, ν in P1 (K) are called Γ-equivalent, denoted by
αβ αµ + β
µ ∼ ν, if there exists M = ∈ Γ such that ν = µ M = . By (4.27), this
γδ γµ + δ
∼ is an equivalence relation. Denote, as usual, by [µ] the equivalence class of
µ ∈ P 1 (K).
a
For µ = ∈ P1 (K) with a, β ∈ OK , we denote by a := OK a + OK b its
b
associated integral ideal.
αµ + β αβ
!
Proof. Indeed, if ν = µ M = ∼ µ for ∈ SL(2, OK ), then the ideal
γµ + δ γδ
αβ
!
OK (αp + β) + OK (γp + δ) corresponding to ν coincides with a, since is
γδ
uni-modular. Thus, all elements in the equivalence class [µ] associated to µ in
P1 (K) correspond to the same ideal class [a]. On the other hand, if a1 = (c)a is an
integral ideal in the ideal class of [a], then c ∈ K and a1 = OK (ca) + OK (cb)
ca
with ca, cb ∈ OK . Hence, we may take µ = ∈ P1 (K). Conversely, if
cb
a1 is an integral ideal associated to µ ∈ P 1 (K), there exists a1 , b1 ∈ OK such
a a1
that a1 = OK a1 + OK b1 and µ = = . This implies ab1 = a1 b and hence
b b1
OK a OK b1 = OK a1 OK b. Therefore, up to factor of rational ideals, we may and
hence will assume that OK a and OK b as also OK a1 and OK b1 are mutually co-
OK a OK b1 OK a1 OK b
prime. Since = , this implies that, up to factor of rational
a a1 a1 a
O K a O K a1 OK b OK b1
ideals, = and similarly = . This means that a1 = (c)a for
a a1 a a1
a certain c ∈ K. Consequently, for this c ∈ K, we have a1 = ac and b1 = bc.
Proposition 4.1. Let h be the class number of K. Then there exist h elements
µ1 , . . . , µh ∈ P 1 (K) such that
n o
P1 (K) ∼ = [µ1 ], . . . , [µh ] .
(4.28)
ai
OK ai + OK bi . In this way, we obtain an element µi := ∈ P 1 (K) such that
bi
ϕ([µi ]) = [ai ] for all 1 = 1, . . . , h. This proves the surjectivity of ϕ.
a a1
Conversely, for a fixed i, assume that the ϕ-images of both µ = , µ1 = ∈
b b1
P (K) with a, b, a1 , b1 ∈ OK are the same ideal class [ai ]. By Lemma 4.3, we may
1
We are ready to deal with general number fields. Let K be an algebraic number
field with OK the ring of integers. Denote by r1 , resp. r2 , the number of real, resp.
complex, places of K. Let a be an integral OK -ideal. Induced from the Minkowski
embedding K ,→ Rr1 × Cr2 , we obtain two embeddings
The product space H r1 × Hr2 admits a natural boundary Rr1 × Cr2 , or better,
1
P (R)r1 × P 1 (C)r2 . By extending the above fractional action to this boundary,
the natural action of SL(OK ⊕ a) on P 1 (K) then induces a SL(OK ⊕ a)-equivalent
embedding P 1 (K) ,→ P 1 (R)r1 × P 1 (C)r2 . Accordingly, we introduce the following
Definition 4.1. We call SL(OK ⊕ a)-orbits in P1 (K) the cusps of the quotient space
SL(OK ⊕ a) H r1 × Hr2 , or the cusps of the arithmetic group SL(OK ⊕ a). Quite
often, we also call representatives of these orbits as cusps of SL(OK ⊕ a).
The cusps of SL(OK ⊕ a) play an important role in the sequel. For later use,
we here give an elegant classification of cusps, rooted back to Maaβ.
α
" #
Proof. To start with, let P = be a point in P1 (K) so that α, β ∈ K, and define
β
π(P) to be the ideal class [OK ·α+a·β] associated to the fractional ideal OK ·α+a·β.
α α
" # " #
Proof. (1) Indeed, if P = 1 = 2 , we have, as ideal classes,
β1 β2
h i h β2 i h α1 i h i
OK · α1 + a · β1 = (OK · α1 + a · β1 ) = OK · β2 · + a · β2 = OK · α2 + a · β2 .
β1 β1
!
ab
(2) For γ = ∈ SL(OK ⊕ a), we have that a, d ∈ OK , b ∈ a, c ∈ a−1 . Note
cd
aα + bβ
" #
h i
that γ(P) = . Hence π γ(P) = OK · (aα + bβ) + a · (cα + dβ) . Now, by
cα + dβ
definition,
OK · (aα + bβ) + a · (cα + dβ) = (aα) · OK + (bβ) · OK + (cα) · a + (dβ) · a
⊆ α · OK + β · a + α · (a−1 · a) + β · a = α · OK + β · a.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 67
This implies also that OK · (aα + bβ) + a · (cα + dβ) ⊇ α · OK + β · a, since the
determinant of γ is one. Therefore, we have
h i h i
π γ · P = OK · (aα + bβ) + a · (cα + dβ) = OK · α + a · β .
This verifies (2).
Back to the proof of Theorem 4.3. Let h = hK denote the class number of K.
Choose now integral OK -ideals a1 , . . . , ah such that CLK = [a1 ], . . . , [ah ] . We
want to show that the elements of P1 (K) are divided into h equivalence classes by
the action of SL(OK ⊕ a). From Claim 4.1, there is a well-defined map
α
" #
Π : SL(OK ⊕ a) P1 (K) → CLK , 7→ [OK · α + a · β].
(4.32)
β
It suffices to show that Π is a bijection.
Surjectivity of Π can be easily verified. Indeed, by Lemma 4.1, for all ideals
" # αi , βi ∈ OK such that ai = OK · αi + a · βi . This
ai (1 ≤ i ≤ h), there exist elements
α
implies that the Π-image of i is exactly ai .
βi
To prove that Π is an injection, let ζ1 , ζ2 be two elements of P1 (K) such " # that
α
Π(ζ1 ) = Π(ζ2 ) = [b] for a fixed OK -ideal b ∈ a1 , . . . , ah . Write ζ1 := 1 and
β1
α2
" #
ζ2 := . We have ( f1 )b = OK α1 + aβ1 and ( f2 )b = OK α2 + aβ2 for suitable
β2
f1 , f2 ∈ K. We want to show that there exists an element γ ∈ SL(OK ⊕ a) such that
γζ2 = ζ1 .
For this, we first assume that ( f1 )b = ( f2 )b = b. By definition, this is equivalent
to the conditions that
OK α1 + aβ1 = b and OK α2 + aβ2 = b. (4.33)
α1 α∗1 α α ∗
! !
By Lemma 4.68, there exist M1 := and M2 := 2 ∗2 ∈ SL(2, K) such
β1 β∗1 β2 β2
that
OK β∗1 + a−1 α∗1 = b−1 and OK β∗2 + a−1 α∗2 = b−1 . (4.34)
Clearly, M1 (∞) = ζ1 and M2 (∞) = ζ2 . Hence, if we set γ := M1 · M2−1 , then
α2 α
" #! " #
γ ∈ SL(2, K), and γ = 1 . In other words, γ(ζ2 ) = ζ1 . Thus, to complete
β2 β1
our proof in this special situation, it suffices to verify the following:
a+b·0
! " # " #
ab 1
Moreover, from definition, γ = ∈ Γ∞ if and only if = . That
cd c+d·0 0
!
a b
means that γ = . Thus, if we let
0 a−1
( ! ) ( ! )
a b 1b
B(R) := : 0 , a ∈ R, b ∈ R and N(R) := : b ∈ R (4.39)
0 a−1 01
be the Borel subgroup B(R) of SL(2, R) and the associated unipotent radical N(R),
respectively, then Γ∞ = Γ ∩ B(R). Note that N(R) is isomorphic to the additive
group R+ , and B(R) is isomorphic to the semi-direct product R∗ by R+ . Hence the
group Γ∞ has a relatively simple structure.
Let dZ be the rank one Z-lattice associated to the cusp ζ. A standard funda-
mental domain for the action of Γζ on H is set to be
( ) ( )
d d d d
Aζ · z = x+iy ∈ H : − < x ≤ = Aζ · z = x+iy ∈ C : y > 0, − < x ≤ .
2 2 2 2
In the sequel, to simplify our discussions, we assume that Γ is reduced at all of its
cusps, unless otherwise is stated. This implies, in particular, that, for a cusp ζ of
Γ, Γζ /Z(Γ) ' Z, where Z(Γ) := Γ ∩ {±I} denotes the center of Γ.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 70
If Λ is the rank two Z-lattice naturally associated to the cusp ζ, there is then a
standard fundamental domain for the action of Γζ on H given by:
n o n o
Aζ · (x+iy)+r j ∈ H : (x, y) ∈ Λ = Aζ · x+iy+r j : x, y, r ∈ R, r > 0, (x, y) ∈ Λ .
We write CΓ for the set of cusps of Γ. Clearly, the group Γ leaves CΓ invariant,
and breaks CΓ into Γ-equivalent classes. It is known that if Γ ⊂ SL(2, C) is a dis-
crete group of finite co-volume, then Γ has only finitely many Γ-classes of cusps.
In this latter case, we denote these finite cusps by η1 , . . . , ηh , and, choose, once and
hence for all, elements A1 , . . . , Ah ∈ SL(2, C) such that η1 = A1 ·∞, . . . , ηh = Ah ·∞.
Similarly, as in the case of upper half plane, cusps are closely related with
the concept of parabolic elements. By definition, an element γ ∈ SL(2, C) r {±I}
is called parabolic if | Tr(γ) | = 2. It is not difficult to check that an !element
11
γ ∈ SL(2, C) is parabolic if and only if it is conjugate in SL(2, C) to . Hence
01
ζ ∈ P 1 (C) is a cusp of Γ if and only if ζ is the fixed point of a certain parabolic
element in Γ (on the boundary P 1 (C) of H).
In this subsection, to motivate our later discussions for general number fields,
we recall the corresponding constructions for totally real fields, mainly following
Siegel ([101]).
Let K be a totally real field of degree n over Q with OK its ring of integers. With
respect to the Minkowski embedding K ,→ K ⊗Q R = Rn , for a ∈ K, denote by
(a(i) ) the image of a in Rn .
Let Γ = SL(2, OK ) be the Hilbert modular group and Hn := H n be the product
of n copies of the upper half-plane. Then Γ may be viewed as a group of analytic
automorphisms of Hn via the assignment
αz + β α( j) z j + β( j) αβ
! !
z = (z1 , . . . , zn ) 7→ z M := := ( j) ∀M = ∈ Γ. (4.44)
γz + δ γ z j + δ( j) γδ
In the sequel, we shall identify M ∈ Γ with the modular substitution z 7→ z M if
there is no confusion.
As usual, for a point z of Hn , let Γz := M ∈ Γ : z M = z be the isotopy group
of z in Γ. More generally, for two subsets D and D0 of Hn , we define ΓD,D0 to
be the collection of elements M ∈ Γ satisfying the condition that D M ∩ D0 , ∅.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 72
αβ
!
Proof. Let M = ∈ ΓD,D0 . Assume that w = (w1 , . . . , wn ) ∈ D and write
γδ
w M = (w01 , . . . , w0n ) ∈ D0 . For each j = 1, . . . , n, write w j = u j + iv j and
w0j = u0j + iv0j with u j , u0j ∈ R, v j , , v0j ∈ R>0 . By a direct calculation, we have
vj
v0j = ( j) . Since D and D0 are compact, there exists a constant c > 0,
|γ w j + δ( j) |2
depending only on D and D0 but not on the j’s, such that |γ( j) w j + δ( j) | < c for all
j = 1, 2, . . . , n. This implies that, for all M ∈ ΓD,D0 , the associated γ’s and δ’s are
with bounded norms and hence should all belong to a certain finite subset of OK .
δ −γ
! !
0 1
To go further, let A := . Obviously, A · M · A = M =
t −t
.
−1 0 −β α
Accordingly, introduce the modular transformation ι : z 7→ zA on Hn . Note that
D M ∩ D0 , ∅ if and only if ι(D) M−t ∩ ι(D0 ) , ∅. Even ι(D) and/or ι(D0 ) may not be
compact, since D and D0 are compact, by a direct calculation as above but tracing
back to D and D0 , we conclude that, for all M ∈ ΓD,D0 , the associated α’s and β’s
are with bounded norms and hence should all belong to a certain finite subset in
OK . Therefore, ΓD,D0 itself is finite.
In particular, if we choose D = D0 = {z} ⊂ Hn , from above, we conclude that
Γz = Γ{z},{z} is finite.
Proof. Let z ∈ Hn and let U 0 be an open neighborhood of z such that the closure
U 0 of U 0 in Hn is compact. By applying Lemma 4.4 to the case D = D0 = U 0 ,
we conclude that there exist at most finitely many M ∈ Γ, denoted by M1 , . . . , Mr ,
such that D M ∩ D , ∅. By reordering if necessary, we may and hence will assume
that M1 , . . . , M s are the only elements which do not belong in Γz . Accordingly,
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 73
Proposition 4.2. The Hilbert modular group SL(2, OK ) acts properly and discon-
tinuously on Hn .
ρi ρ∗i ρi bζi∗ ε 0
! ! !
as in Lemma 4.2, then Ai · ∞ = λi , and MAi = , where both
σi σ∗i σi η∗i 0 ε
elements b ζi∗ = (αρ∗i + βσ∗i )ε and η∗i = (γρ∗i + δσ∗i )ε lie in a−1 i . By definition,
ρi σi − σi ρ∗i = 1 = ρi η∗i − σib
∗
ζi∗ . We obtain ρi (η∗i − σ∗i ) = σi (ξ∗ − ρ∗ ). This implies
that
(ρi ) (η∗i − σ∗i ) (σi ) (ξ∗ − ρ∗ )
· = · . (4.45)
ai a−1
i
ai a−1
i
(ρi ) (σi ) (ρi )
Again, since and are prime to each other, we conclude that divides
ai ai ai
(ξ∗ − ρ∗ )
. Hence, there exists an OK -ideal b such that (ξ∗ − ρ∗ ) = a−2
i b (ρi ), or
a−1
i
i such that ζi = ρi + ρi ζ. Consequently, by (4.45), we
the same, there exists ζ ∈ a−2 b∗ ∗
have ηi = σi + σi ζ. Therefore,
∗ ∗
ε ζε−1
!
MAi = Ai .
0 ε−1
This then completes our proof of the following:
α
" #
Proposition 4.3. (Siegel) Let K be a totally real field. Let λ = ∈ P 1 (K) such
β
α α∗
!
that α, β ∈ OK , and that, as OK ideals, (α) is prime to (β). Denote by A = ∈
β β∗
SL(2, K) a special transformation in Lemma 4.2. Then, the stabilizer group Γµ of
λ in Γ = SL(2, OK ) is given by;
ε ζ
( ! )
Γλ = A : ζ ∈ ai , ε ∈ U K A−1 .
−2
0 ε−1
for its trace. If λ ∈ K, N(λ) and Tr(λ) coincide with the usual norm and trace. In
addition, we denote zA−1 = (z∗1 , . . . , z∗n ), and write z j = x j + iy j and z∗j = x∗j + iy∗j
with x j , x∗j ∈ R, y j , y∗j ∈ R>0 . For later use, set y∗ := yA−1 := (y∗1 , . . . , y∗n ).
1
(a) The first coordinate p of z is preserved. Indeed, by definition, we have
N(y∗ )
N(yA−1 M ) = N(yT A−1 ) = N(ε2 )N(yA−1 ) = N(yA−1 ) = N(y∗ ).
(b) The next (n − 1) coordinates Y j ’s are shifted by k j ’s, i.e.
Y1 , Y2 , . . . , Yn−1 7→ Y1 + k1 , Y2 + k2 , . . . , Yn−1 + kn−1 .
the n-rowed square matrix (α(i j) ) and V denotes the diagonal matrix
(ε(1) )2 , . . . , (ε(n) )2 .
ε 0
!
Proof. Indeed, for ε = ±εk11 · · · εkn−1
n−1
and Mε = A A−1 , the changes on
0 ε−1
Y1 , Y2 , . . . , Yn−1 under the substitution z 7→ z Mε are given by:
Y1 , Y2 , . . . , Yn−1 7→ Y1 + k1 , Y2 + k2 , . . . , Yn−1 + kn−1 .
Hence, by choosing k1 , . . . , kn−1 properly, we may assume that the coordinates
Y1 , . . . , Yn−1 of z Mε satisfy (4.47). Moreover, since, for ζ = m1 α1 + . . . + mn αn and
1 ζ −1
!
Mζ = A A , the changes on X1 , X2 , . . . , Xn under the substitution z 7→ z Mζ
01
are given by:
X1 , X2 , . . . , Xn 7→ X1 + m1 , X2 + m2 , . . . , Xn + mn .
Hence, we can choose suitable m1 , . . . , mn such that coordinates X1 , . . . , Xn of
z Mζ Mε satisfy (4.47). Therefore, z Mζ Mε is reduced with respect to Γµ .
Let K be an algebraic number field with r1 , resp. r2 , real, resp. complex, embed-
dings of K. Denote by OK the ring of integers of K. Let a be an OK -ideal of K.
There is a natural action of Γ := SL(OK ⊕ a) on H r1 × Hr2 where, as above, H,
resp. H, is the complex upper half plane, resp. the upper half space. Our main aim
in this subsection is to construct a canonical fundamental domain for the quotient
space SL(OK ⊕ a) H r1 × Hr2 , which parametrize rank two OK -lattices of volume
one.
α
" #
Fix a point η = ∈ P 1 (K). Here, we assume, as we may, that α, β ∈ OK . Denote
β
by Γη its associated stabilizer group. Namely,
Γη := γ ∈ SL(OK ⊕ a) : γη = η .
(4.48)
Our main purpose here is to construct a fundamental domain for this action. For
this, we first determine the structure of Γη .
Since ab−2 is a OK -lattice, we call η a cusp of Γ and ab−2 the lattice associated to
the cusp η.
Proof. Since all elements in A−1 Γη A fix ∞, so they are given by upper triangle
! !
−1 a b u z
matrices. Accordingly, we write an element A A of A Γη A as
−1
for
cd 0 u−1
suitable u, z ∈ K. This implies that
α α∗ u z a b α α∗ aα + bβ ∗
! ! ! ! ! !
uα ∗
= and hence = .
β β∗ 0 u−1 c d β β∗ uβ ∗ cα + dβ ∗
In particular, as OK -modules, (uα) = (aα + bβ) and (uβ) = (cα + dβ). This implies
that, as fractional OK -ideals,
(u)(α, β) = (uα, uβ) = (aα + bβ, cα + dβ). (4.51)
Recall that, from (4.30), a, d ∈ OK , b ∈ a and c ∈ a . From d(aα + bβ) − b(cα +
−1
Proof. Since
β∗ −α∗ a b α α∗
! ! ! !
u z
· · = , (4.54)
−β α cd β β∗ 0 u−1
by a direct calculation, we have that
z = (a − d)α∗ β∗ − c(α∗ )2 + b(β∗ )2 . (4.55)
For entries involved, as mentioned above, a, d ∈ OK , b ∈ a, c ∈ a . Moreover, −1
as required.
ε z ε
! " #! " # " #
1 1
Finally, if z ∈ ab−2 , and ε ∈ U K , we have = = . Conse-
0 ε−1 0 0 0
ε z
!
quently, ∈ A−1 Γη A.
0 ε−1
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 80
Now we are ready to construct a natural fundamental domain for the action of
Γη ⊂ SL(OK ⊕ a) on H r1 × Hr2 .
α α∗
!
Let A = ∈ SL(2, K) be as in the previous subsection. Obviously,
β β∗
α
" # !
u 0
A(∞) = η = and Γη = Γ0η × A A−1
: u ∈ U K . (4.56)
β 0 u−1
! !
1 z −1
u 0 uz
Here Γη := A
0
A : z ∈ ab . Since
−2
(z) = −1 = u2 z, the isotropy
01 0 u−1 u
group of η, or the same, the stabilizer group of η, in A−1 SL(OK ⊕ a)A is generated
by:
Next we analysis how these two types affect our construction of a fundamental
domain for Γη . First, to deal with translations, we introduce the following:
Definition 4.4.
1 +r2 1 +r2
n o
on the norm-one hyper-surface S := y ∈ Rr>0 : N(y) = 1 in Rr>0 . By taking
logarithms, S is transformed bijectively into a trace-zero hyperplane H, which is
isomorphic to Rr1 +r2 −1
log Xr1 +r2
S ←→ H := (a1 , . . . , ar1 +r2 ) ∈ Rr1 +r2 : ai = 0 ' Rr1 +r2 −1 ,
n o
i=1
log y1 , . . . , log yr1 +r2 .
y 7→
Accordingly, the actions of unite in U K2 on S are transformed into the actions on
Rr1 +r2 −1 by translations: ai 7→ ai + log ε(i) . By Dirichlet’s Unit Theorem, the
logarithm of U K2 is a full rank lattice in Rr1 +r2 −1 . Consequently, the exponential
map transforms a fundamental parallelepiped of this lattice back to a fundamental
domain, denoted by SUK2 , for the action of the unit group U K2 on the hyper-surface
1 +r2
S. Hence, the cone based on SUK2 , that is the subset R>0 · SUK2 ⊂ Rr>0 , is a
2 r1 +r2
fundamental domain for the action of U K on R>0 .
Therefore, if we denote by T a fundamental domain for the action of the
translations by elements of ab−2 on Rr1 × Cr2 , then what we have just said proves
the following:
Back to the case of a single upper half plane. Let Γ be a Fuchsian group of first
kind. By definition, a connected domain F of H is called a fundamental domain
for the action of Γ on H if the following conditions are satisfied:
Proposition 4.5. (See e.g. [62]) The set F is a fundamental domain for the action
of Γ on H. Moreover, we have
(iv) For Γ, a Fuchsian group of the first kind, every vertex of F on P1 (R) is a cusp
of Γ.
(v) Every cusp of Γ is equivalent to a vertex of F on P 1 (R).
To end this discussion, we point out that Γ\H admits a natural Riemann sur-
face structure. To explain this, let CΓ ⊆ P 1 (R) be the set of all cusps of Γ and
H ∗ := HΓ∗ := H ∪ CΓ . And, for a fixed l > 0, let Ul := z ∈ H : =(z) > l
and set Ul∗ := Ul ∪ {∞}. Then, there exists a natural topology structure on H ∗
characterized by
α α
! " #
ab
Naturally, H is Hausdorff. Indeed, if σ =
∗
∈ Γ and x = = ,
cd β β
( )
=(z)
we have σ(Ul ) = z ∈ H : > l . In particular, σ(Ul ) is contained in a
|cz + d|2
circle of radius (2l2 )−1 which is tangent to the x-axis. As usual, if x ∈ CΓ , we call
σ(Ul ) a neighborhood of x in H as well. For example, if ∞ is a cusp of Γ, then a
fundamental neighborhood system of Γ\H ∗ at ∞ is given by F ∩ Ul∗ l>0 . In fact,
it is easy to check that F ∩ Ul∗ = z ∈ H ∗ : =(z) > l hz 7→ z + mi is a rectangle
of width m, starting from y = l and towards infinity.
Tautologically, the action of Γ on H extends naturally to H ∗ . And, with re-
spect to the induced quotient topology Γ H ∗ is Hausdorff as well. Moreover, we
have the following:
(i) Two points of Hλ are Γ-equivalent if and only! if they are Γ∞ -equivalent.
1 λi −1
(ii) If A1 , A2 ∈ SL(2, C) satisfying that Ai A ∈ Γ for a certain λi ∈ C r {0},
0 1 i
and that A1 (∞), A2 (∞) ∈ P 1 (C) are Γ-equivalent, then A1 ·Hλ1 ∩ A2 ·Hλ2 = ∅
for all γ ∈ Γ.
(iii) For ζ ∈ P 1 (C), if Γζ contains a parabolic element, then ζ is a cusp of Γ.
Namely, Γµ is associated with a rank two Z-lattice.
Indeed, over H, the topology induces by these hornbills is equivalent to the stan-
dard topology. To extend this topology to H∗ , it suffices to add the horoballs
touching P 1 (C) at the λ as systems of open neighborhoods for λ ∈ P 1 (C). Tauto-
logically, we have
Let K be a totally real field of degree n with OK its ring of integers. Let Γ :=
SL(2, OK ) be the Hilbert modular group. Denote by Hn := H n be the product of
n copies of the upper half plane H. With respect to the natural action of Γ on Hn ,
the quotient space Γ\Hn is not compact. To compatify it, we add the cusps of Γ.
To motivate our later discussion for general number fields, in this subsection, we
follow Siegel ( [101]) closely to first introduce a Γ-invariant function measuring
the difference between a cusp of Γ and a modular point in Γ\Hn , and then to use
this function to construct a fundamental domain for the action of Γ on Hn .
To facilitate our ensuing discussion, we first introduce some notations. For a point
z = (zk ) = (xk + iyk ) ∈ Hn with xk ∈ R, yk ∈ R>0 , we denote x = xz := ! (xk ) ∈ R ,
n
ab
y = yz := (yk ) ∈ Rn>0 , and let N(y) := nk=1 yk . Moreover, if M =
Q
∈ SL(2, K),
cd
a(k) z + b(k)
!
we write z M for M(z) = (k) . Here, for an element α ∈ K, α(k) denotes its
c z + d(k)
image of the k-th real embedding K → R. Accordingly, we write y-part of M(z) !
yk
as y M(z) , or simply y M . An easy calculation implies that y M = (k) . In
! |c zk + d(k) |2
yk y
the sequel, we denote the n-tuple (k) simply by .
|c zk + d(k) |2 |cz + d|2
ρ ρ
" #
Let µ = = ∈ P 1 (K) be a cusp of Γ. As before, we assume that ρ, σ ∈ OK
σ σ
ρ ξ
!
and set b := OK ρ + OK σ. Accordingly, fix a special transformation A := ∈
ση
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 86
1
For example, if µ = ∞, then ∆(z, ∞) = p . Hence, the larger is the N(y),
N(y)
the smaller is the ∆(z, ∞).
For later use, we next give some basic properties for the function ∆(z, µ).
(2) ∆(z, µ) is well-defined. That is, it does not depend on the elements ρ, σ and
the matrix A used.
Proof. These are very easy to verify. Indeed, for (1), by definition,
− 1 − 1 1
∆(z M , µ M ) = N y(zM )A−1 M−1 2 = N yz(A−1 M−1 M) 2 = N yA−1 − 2 = ∆(z, µ).
ρ1 ρ ζ
!
For (2), let µ = and A1 = 1 1 ∈ SL(2, K) with ρ1 , σ1 ∈ OK and A1 satisfy-
σ1 σ1 η1
ing the conditions described before the definition. Then, A−1 A1 ∈ Γ!∞ . Therefore,
ε ζ
by Theorem 4.4, or better, Proposition 4.3, we have A−1 A1 = for a certain
0 ε−1
unit ε in K. This implies that N(yA−1 1
) = N(yA−1 )N(ε−2 ) = N(yA−1 ).
As above, write OK ρ+OK σ = a. Assume first that a attaches the minimal norm
among OK -ideals in its associated ideal class [a]. By Dirichlet’s unit theorem, for
every α = (α1 , . . . , αn ) ∈ (K ∗ )n , there exists a unit ε in K such that, for a certain
constant c1 > 0 depending only on K,
pn
αi · ε(i) ≤ c1 |N(α)| (i = 1, . . . , n). (4.61)
−(σx + ρ)2 + σ2 y2
!
Since N keeps unchanged when (ρ, σ) is changed to an as-
y
sociated pair (ερ, εσ) for an ε ∈ U K , there exists a constant c2 > 0, depending
only on c1 , such that, up to associations, for all (ρ, σ)’s,
2
(−σ(i) xi + ρ(i) )2 + σ(i) y2i
≤ c2 , i = 1, 2, . . . , n. (4.62)
yi
This implies that (−σ(i) xi + ρ(i) ) and σ(i) and, hence, ρ(i) and σ(i) are all bounded.
Therefore, up to associations, there exist only finitely many pairs of (ρ, σ) sat-
isfying (4.62). Among these finite pairs, obviously, there ! exists at least a pair,
−(σx + ρ)2 + σ2 y2
say (ρ1 , σ1 ), at which the values of N attaches its minimum.
y
Obviously, (ρ1 , σ1 ) satisfies the inequality (4.60).
Next, to treat general case, let a ∈ K be an element satisfying a =
(a) OK ρ1 + OK σ1 . In the inequality (4.60), if we put ρ = ρ0 := aρ1 and
σ = σ0 := aσ1 , by the minimum property of (ρ1 , σ1 ), we have N(a) ≥ 1.
On the other hand, since a is of minimum norm among OK -ideals of its class,
N(a) ≤ N(a−1 a). Hence N(a) ≤ 1 and N(a) = 1. This then essentially com-
ρ0
pletes the proof. Indeed, by definition, a = OK ρ0 + OK σ0 , and, for λ0 := , we
σ0
1 1
Lemma 4.8. ([101]) There exists a constant d > 0, depending only on K, such
that, if ∆(z, µ1 ) < d and ∆(z, µ2 ) < d for all z = x + iy ∈ Hn , then µ1 = µ2 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 88
ρ1 ρ2
Proof. Write µ1 = and µ2 = and assume that ρ1 , σ1 , ρ2 , σ2 ∈ OK . Then
σ1 σ2
the conditions ∆(z, µ1 ) < d and ∆(z, µ2 ) < d are equivalent to that
1 1
N − (σ1 x + ρ1 )2 y−1 + σ21 y 2 < d and N − (σ2 x + ρ2 )2 y−1 + σ22 y 2 < d.
As in the proof of Lemma 4.7, by multiplying ρ1 and σ1 with a suitable unit ε ∈
U K , if necessary, we may assume that there exists a constant c1 > 0 such that
(i) 2
1 xi − ρ1 ) yi + σ1 yi < c1 d
(σ(i) (i) 2 −1 2/n
for all i = 1, 2, . . . , n. This implies that
1
(i) − 2 √ 1 √ 1/n
− σ(i)
1 xi + ρ1 yi < c1 d1/n and σ(i)
1 yi <
2
c1 d .
Similarly, there exists a constant c2 > 0 such that
1
(i) − 2 √ 1 √
− σ(i)
2 xi + ρ2 yi < c2 d1/n and σ(i)
2 yi <
2
c2 d1/n .
Therefore,
N(ρ1 σ2 − ρ2 σ1 ) < 2c1 c2 d2/n n ,
1 1 1 1
(i) − 2 (i) − 2
1 σ2 −ρ2 σ1 = −σ1 xi +ρ1 yi ·σ2 yi − −σ2 xi +ρ2 yi ·σ1 yi . Hence,
since ρ(i) (i) (i) (i) (i) (i) 2 (i) (i) 2
Lemma 4.9. ([101]) There exists a constant c > 0, depending only on K, such
that, for each z = x + iy ∈ Hn , there exists a cusp µ such that ∆(z, µ) < c.
p
This is certainly satisfied when ci = d j = ∆K (i, j = 1, 2, . . . , n). In
2n
Let now a be an OK -ideal such that its norm is minimal among all OK -ideals
of in the (ideal) class [OK ρ + OK σ]. Then, there exist an element a of K such that
aρ ρ
a OK ρ + OK σ = a and N(a) ≤ 1. Therefore, for the cusp µ := = ,
aσ σ
1 1
∆(z, µ) = N(a) · N −(σx + ρ) y + σ y ≤ 2 ∆K 2 .
2 −1 2 2 n
We next use the properties established for ∆(z, µ) above to construct a fundamental
domain for the action of Γ = SL(2, OK ) in Hn following Siegel ([101]). To begin
with, for a cusp µ ∈ P 1 (K) and r > 0, let
Uµ,r := z ∈ Hn : ∆(z, µ) < r
(4.64)
be an influence sphere of µ.
(1) Let Dλ is the fundamental domain of Γµ as in §4.3.3.3. Then the subset Uµ,r is
Γµ -invariant. In particular, Dµ ∩ Uµ,r is a fundamental domain of Γµ in Uµ,r .
(2) Let d be the constant in Lemma 4.8. Then the Uµ,d ’s are mutually
n o disjoint.
(3) Let c be the constant in Lemma 4.9. Then the collection Uµ,c is an
µ∈P1 (K)
open covering of Hn .
Proof. The assertion (2), resp. (3), is a direct consequence of Lemma 4.8, resp.
Lemma 4.9. As for (1), note that, if M ∈ Γµ , then µ M = µ. Hence, by
Lemma 4.6(1), we have
∆(z M , µ) = ∆(z M , µ M ) = ∆(z, µ) < r ∀z ∈ Uµ,r
This implies that z M ∈ Uµ,r . Therefore, Uµ,r is Γµ -invariant.
(1) The ∆-invariant ∆(z) of z ∈ Hn is defined by ∆(z) := inf λ∈P 1 (K) ∆(z, µ).
(2) If ∆(z, µ) = ∆(z), z is defined to be semi-reduced with respect to the cusp µ.
(3) The subset Fµ to be the collection of all elements in Hn which are semi-
reduced with respect to µ.
By Lemma 4.7, ∆(z) = ∆(z, µ) holds for at least one cusp µ of Γ. So the assertion
(2) makes sense.
(1) If ∆(z) = ∆(z, µ), then ∆(z) is Γµ -invariant. That is, for z ∈ Hn and M ∈ Γµ ,
∆(z M ) = ∆(z). In particular, Fµ is Γµ -invariant.
(2) For the constant c in Lemma 4.9, Fµ ⊂ Uµ,c .
Proof. Since ∆(z, µ) is Γ-invariant,
∆(z M ) = inf ∆(z M , µ) = inf ∆(z, µ M−1 ) = inf ∆(z, µ) = ∆(z).
µ µ µ
This proves (1). (2) comes directly from the constructions of Fµ and Uµ,c .
Definition 4.7.
(ii) If at least one equality holds in the inequalities (4.65), z is called a bound-
ary point of Di . We denote by ∂Di the set of boundary points of Di .
(iii) A point z ∈ F is called an inner point of F , if z is an inner point of some
Di . Similarly, we may define a boundary point of F and hence obtain the
boundary ∂F of F (consisting of all boundary points of F ).
(1) The set Di coincides with the topological closure D0i of D0i .
(2) The set of inner points of F is open in Hn .
in (σ(i) , ρ(i) )’s, there exists a constant c = c(z) which depends continuously on
1
z, but not on µ, such that ∆(z, µ) ≥ c N(σ2 + ρ2 ) 2 . Consequently, if z0 ∈ D0i ,
there is a sufficiently small neighborhood U of z0 such that, for all z ∈ U,
c0 1
∆(z, µ) ≥ N(σ2 + ρ2 ) 2 , where c0 = c(z0 ). Moreover, by striking U if neces-
2
sary, we may and hence will assume that ∆(z, µi ) ≤ 2∆(z0 , µi ) for all z ∈ U. This
implies that
c0 1 ρ
∆(z, µ) − ∆(z, µi ) ≥ N(σ2 + ρ2 ) 2 − 2∆(z0 , µi ) ∀z ∈ U, ∀µ = ∈ P 1 (K).
2 σ
Hence, following the same arguments in the proof of Lemma 4.7, there
are only finitely many non-associated pairs of integers (ρ, σ) such that
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 92
c0 1
N(σ2 + ρ2 ) 2 ≤ 2∆(z0 , µi ). Hence ∆(z, µ) > ∆(z, µi ) for all µ , µi , except for
2
finitely many points λ1 , . . . , λk ∈ P 1 (K). Therefore, by the continuity of ∆(z, λ j )’s
in z, there exists a neighborhood V of z0 such that, for j = 1, . . . , k, we have
∆(z, λ j ) > ∆(z, µi ), for all z ∈ U and λ j , µi . Consequently, if µ , µi , we have
∆(z, µ) > ∆(z, µi ) for all z ∈ U ∩ V. Furthermore, it is easy to see that, if necessary,
by shrinking V, the inequalities (4.47) hold for all z ∈ V, in addition. This implies
that the open neighborhood U ∩ V of z0 is contained entirely in D0i .
Consequently, for z ∈ Di , the curve defined by z(t) = (x∗ + ity∗ )Ai , t ≥ 1 lies
entirely in Di . Hence, for sufficiently large t, the z(t) ’s are all contained in D0i .
This can be used to show that Di and D0i are connected. For details, please refer
to [101].
Now we are ready to stated the following main result of Chapter III in
[101].Recall that for F ⊆ Hn and M ∈ Γ, FM is defined to be the collection of
al the z M ’s for all z ∈ F .
Proof. (1) By the comments after Definition 4.6, for every z ∈ Hn , there exists a
cusp µ such that ∆(z) = ∆(z, µ). Since there exist a certain 1 ≤ i ≤ h and an M ∈ Γ
satisfying µ = (µi ) M , we have ∆(z M−1 ) = ∆(z) = ∆(z, µ) = ∆(z M−1 , µi ). Hence
z M−1 ∈ Fµi . Furthermore, by definition, there is an Mi ∈ Γµi such that (z M−1 ) Mi is
reduced with respect to Γµi . Therefore, z Mi M−1 ∈ F .
(2, 3) Let z1 , z2 ∈ F . Assume that z1 ∈ Di , z2 ∈ D j for some 1 ≤ i, j ≤ h, and
that z1 = (z2 ) M for a certain M in Γ r {±I2 }. Since z1 ∈ Fµi , by the Γ-invariance,
we have ∆(z1 , µi ) ≤ ∆ z1 , (µi ) M−1 = ∆(z2 , µ j ). Similarly, we have ∆(z2 , µ j ) ≤
∆ z2 , (µi ) M = ∆(z1 , µi ). Therefore, ∆(z1 , µi ) = ∆(z2 , µi ) = ∆ z1 , (µ j ) M . This
implies that, with the constants c and d as in Lemmas 4.8 and 4.9, either one
of the following two statements holds: (i) ∆(z1 , µi ) = ∆(z2 , µ j ) < d, or (ii)
d ≤ ∆(z1 , µi ), ∆(z2 , µ j ) ≤ c.
We first treat the case ∆(z1 , µi ) = ∆(z2 , µ j ) < d. By Lemma 4.8, µi = (µ j ) M ,
since ∆(z1 , µi ) < d and ∆(z1 , (µ j ) M ) < d. By definition, for different i and j, µi
and µ j are not Γ-equivalent. Hence, i = j and M ∈ Γµi . Furthermore, since both
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 93
z1 and z2 are contained in Dµi and z1 = (z2 ) M for a certain M ∈ Γµi , both z1 and z2
are contained in ∂F . In addition, there are only finitely many such M’s in the set
i=1 Γµi .
Sh
Next we deal with the case d ≤ ∆(z1 , µi ), ∆(z2 , µ j ) ≤ c. By Lemmas 4.8 and
4.9, µi and µ j are not Γ-equivalent. To go further, for each 1 ≤ k ≤ h, denote by
Bk the set of z ∈ Hn satisfying d ≤ ∆(z1 , µk ) ≤ c and z ∈ Dµk . By Lemma 4.1, Bi
and hence B = hi=1 Bi are compact. Hence, both z2 and z1 = (z2 ) M belong to a
S
compact set. Consequently, by Lemma 4.4, there are only finitely many such M in
Γ, depending only on B, and hence only on K. Since ∆(z1 , µi ) = ∆ z1 , (µ j ) M and
(µ j ) M , µi , z1 and similarly z2 are contained in the boundary ∂F . This implies that
there are points z, arbitrarily close to z1 and z2 , such that ∆(z1 , µi ) , ∆ z, (µ j ) M .
Therefore, for each of the two cases, there are no two distinct inner points
of F which are Γ-equivalent. Moreover, F intersects only finitely many of its
neighborhoods, namely the F M ’s in both cases, at points on the boundary.
Remark 4.1. The manifolds defined by the ∆ equation above are seen to be gen-
ρi
eralizations of the isometric circles for a Fuchsian group. Indeed, for µi = and
σi
ρ
(µ j ) M = , the condition becomes N |σi z − ρi | = N |σz − ρ| . Hence, when n = 1
σ
and µi = ∞, the condition becomes |σz − ρ| = 1. This is simply the ‘isometric
ηz − ξ
circle’ corresponding to the transformation z 7→ on H.
−σz + ρ
Let K be a number field of degree n over Q. Denote by OK its ring of integers and
by ∆K the absolute value of its discriminant. Denote by r1 , resp. r2 , the number of
real embeddings, resp. complex embeddings, of K. Then n = r1 + 2r2 . Moreover,
for an element a ∈ K, denote (a(i) ) its image under the Minkowski embedding
K ,→ Rr1 × Cr2 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 94
Recall that, for a point τ = (z1 , . . . , zr1 ; P1 , · · · , Pr2 ) ∈ H r1 × Hr2 , we have let
ImJ(τ) := =(z1 ), . . . , =(zr1 ), J(P1 ), . . . , J(Pr2 ) ∈ Rr1 +r2 , where =(z) = y for z =
x + iy ∈ H with x ∈ R, y ∈ R>0 and J(P) = v for P = z + v j ∈ H with z ∈ C, v ∈
R>0 . Set now
r1 r2
Y Y
N(τ) := N ImJ(τ) = J(P j )2 = y1 · . . . · yr1 · v1 · . . . · vr2 2 . (4.67)
=(zi ) ·
i=1 j=1
η and τ by
=(z1 ) · · · =(zr1 ) · J(P1 )2 · · · J(Pr2 )2
µ(η, τ) := N a−1 · (OK α + aβ)2 · Qr1
i=1 |(−β zi + α )| j=1 k(−β P j + α )k
(i) (i) 2
Qr2 ( j) ( j) 2
1 N(ImJ(τ))
= · .
N(ab ) kN(−βτ + α)k2
−2
Lemma 4.13. Let τ ∈ H r1 × Hr2 and η ∈ P 1 (K). For the reciprocal distance
µ(η, τ) between η and τ, we have
(1) It is well defined. Namely, it does not depend on the choices of α and β.
(2) It is SL(OK ⊕ a)-invariant. That is to say,
µ(γη, γτ) = µ(η, τ), ∀γ ∈ SL(OK ⊕ a).
α α
" # " 0#
Proof. (1) If η = = in P1 (K), there exists λ ∈ K ∗ such that
β β0
α
" 0#
α0 = λ · α and β0 = λ · β. Therefore, µ(η, τ) in terms of 0 is given by
β
1 N(ImJ(τ))
· 0 τ + α0 )k2
where b = OK α + aβ = (λ) · b. Hence, µ(η, τ),
0 0 0
N(ab0 −2 ) kN(−β
α
" 0#
in terms of 0 , is equal to
β
N(λ)2 N(ImJ(τ)) 1 N(ImJ(τ))
· = · .
N(ab ) N(λ) · kN(−βτ + α)k
−2 2 2 N(ab ) kN(−βτ + α)k2
−2
α
" #
This is nothing but µ(η, τ) in terms of .
β
! (2) for a fixed pair α, β of the cusp η. Assume
(2) By (1), it suffices to verify
α α∗
then α, β ∈ OK . Let Aη = be one of the special transformations introduced
β β∗
β∗ −α∗
!
just before Definition 4.8. Since A−1 η = , the second row of A−1
η is to-
−β α
tally determined by α and β, even A may not be unique. With this, by a direct
calculation, we have
1
µ(η, τ) = · N ImJ A−1 η (τ) .
(4.70)
N(ab−2 )
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 96
Claim 4.5. The denominator and the numerator of the right hand of (4.70) are
Γ-invariant. In particular, µ(η, τ) is Γ-invariant.
!
ab
Proof. Let γ = ∈ Γ. By definition, a, d ∈ OK and b ∈ a and c ∈ a−1 . Since
cd
aα + bβ α
" # " #
γ(η) = =: 1 with α1 = aα + bβ and β1 = cα + dβ. Then
cα + dβ β1
b1 := OK α1 +aβ1 = OK aα+OK bβ+acα+adβ ⊆ OK α+aβ+OK α+aβ = OK α+aβ = b.
Since γ, being uni-modular, belongs to Γ, a similar argument as above implies that
b ⊆ b1 and hence b1 = b. Therefore, N(ab−2 1 ) = N(ab ). Hence, the denominator
−2
To show that the numerator is also Γ-invariant, we first verify the following
Claim 4.6. Let η ∈ P 1 (K) and let γ ∈ Γ be one of the special permutations associ-
ated to the cusp η in Lemma 4.2. Then γAη is a special permutation associated to
the cusp γ(η). Namely
Aγη = γAη . (4.71)
α α α∗
" # ! !
a b
Proof. Let η = with α, β ∈ OK , γ = ∈ Γ and Aη = ∈ SL(2, K).
β c d β β∗
Then
aα + bβ aα∗ + bβ∗ aα + bβ
! !
γAη = ∈ SL(2, K) and γ(η) = ∈ P 1 (K).
cα + dβ cα∗ + dβ∗ cα + dβ
From the proof of the previous claim, OK (aα + bβ)a(cα + dβ) = OK α + aβ = b.
Moreover,
OK (cα∗ + dβ∗ ) + a−1 (aα∗ + bβ∗ ) ⊆ b−1 + b−1 + b−1 + b−1 = b−1
since OK β∗ + a−1 α∗ = b−1 and a, d ∈ OK , b ∈ a, c ∈ a−1 . Hence, similar to the
proof of the previous lemma, this implies that OK (cα∗ +dβ∗ )+a−1 (aα∗ +bβ∗ ) = b−1 ,
since γ ∈ Γ is unimodular.
−1
γη (γτ) = γAη
A−1 η γ (γτ) = Aη γ γτ = Aη (τ).
(γτ) = A−1 −1 −1 −1 −1
Claim 4.7. There exists a positive constant c, depending only on K, such that, if
∆(η, τ) < c and ∆(η0 , τ) < c for τ ∈ H r1 × Hr2 and η, η0 ∈ P 1 (K), then η = η0 .
Proof. Similar to the proof of Lemma 4.7, for every (r1 + r2 )-tuple (t1 , . . . , tr1 +r2 )
of non-zero real numbers, by Dirichlet’s unit theorem, there exists a unit ε ∈ K
such that
1
ti ε(i) ≤ c · N(t) r1 +r2 ∀i = 1, . . . , r1 + r2
Qr1 Qr1 +r2 2
where N(t) := i=1 ti · j=r1 +1 t j and c is a constant depending only on K. Hence,
for a suitable constant T > 0 to be determined later, by multiplying α and β with
a suitable uint in K if necessary, we may and hence will assume that
2
max =(zi )−1 − β(i) zi + α(i) , J(P j )−2 − β( j) P j + α( j)
1 2 1 2
−r −r
≤ c · ∆(η, τ) 1 +r2 · C r1 +r2 ≤ c · T 1 +r2 · C r1 +r2 .
This implies
n o
max − β(i) <(zi )+α(i) · =(zi )−1/2 , − β( j) Z(P j )+α( j) · J(P j )−1
− 2(r 1+r 1
≤ c1/2 · T 1 2) · C r1 +r2 , (4.73)
− 2(r 1+r ) 1
n o
max β(i) · =(zi )1/2 , β( j) · J(P j ) ≤ c1/2 · T 1 2 ·C r1 +r2
.
In parallel, for α0 and β0 , we obtain similar inequalities. On the other hand, for
real places
α(i) (β0 )(i) − β(i) (α0 )(i) = − β(i) <(zi ) + α(i) =(zi )−1/2 · (β0 )(i) =(zi )1/2
Lemma 4.15. There exists a positive real number T := T (K), depending only on
K, such that, for τ ∈ H r1 × Hr2 , there exists a cusp η such that µ(η, τ) > T .
Proof. By Theorem 4.3, there are finitely many inequivalent cusps. Hence, it
suffices to show that, there exist α, β in OK , both of which are not zero at the same
time, satisfying the inequality
2
N(−βτ + α) · N ImJ(τ) −1 ≤ T −1 .
Definition 4.9.
(1) For a modular point τ ∈ H r1 × Hr2 and a Γ-cusp η ∈ P 1 (K), we define the
distance d(η, τ) of the modular point τ from the cusp η by
1
d(η, τ) := .
µ(η, τ)1/2
α
" #
(2) For a Γ-cusp η = ∈ P 1 (K), we define the influence sphere Sη of η by
β
n o
Sη := τ ∈ H r1 × Hr2 : µ(η, τ) ≥ µ(η0 , τ), ∀η0 ∈ P1 (K) .
−1/2
For example, if η = ∞, the distance d(η, ∞) is simply N(τ) · N(a) , since,
N(a(OK · 1 + a · 0)2 )N(τ)
by definition, µ(∞, τ) = = N(a) · N(τ).
|N(−0τ + 1)|2
d(γ−1 η, τ) ≥ d(η, τ)
k k
d(η, γτ) ≤ d(γη, γτ)
and that the inequalities are strict if γη , η.
iη : Γη Sη ,→ SL(OK ⊕ a) H r1 × Hr2 .
(4.75)
Similar to the totally real case, these h pieces of Sη ’s glue together along the
so-called ‘pants’ of their boundaries. Note that the action of Γη on H r1 × Hr2 is
free. Therefore, all fixed points of SL(OK ⊕ a) on H r1 × Hr2 lie on the boundaries
of Sη , characterized by the conditions that µ(η, τ) = µ(η0 , τ) for some η0 ∈ P1 (K).
Denote, as before, by η1 , . . . , ηh all the inequivalent cusps for the action of
SL(OK ⊕ a) on H r1 × Hr2 , and choose Aηi ∈ SL(2, K) to satisfy conditions listed
shortly before Definition 4.8, so that, in particular, Aηi (∞) = ηi , i = 1, 2, . . . , h.
Moreover,
1 +r2
n o
(i) Let S be the norm-one hyper-surface S := y ∈ Rr>0 : N(y) = 1 , and denote
by SUK2 a fundamental domain of the action of U K2 on S.
(ii) Let T be a fundamental domain
n for the translation actions of elements in ab−2o
on R × C , and let E := τ ∈ H r1 × Hr2 : ReZ (τ) ∈ T , ImJ (τ) ∈ R>0 · SUK2
r1 r2
To end this discussion, finally, we give an equivalent description for the funda-
mental domain constructed above. For this, we first introduce a natural geometric
truncation within this fundamental domain. Let T > 0 be a constant.
n o
(iii) We let S T := SL(OK ⊕ a) τ ∈ H r1 × Hr2 : µ(η, τ) ≤ T ∀η ∈ CSL(OK ⊕a) . And,
n o
(iv) We set W(η, T ) := τ ∈ H r1 × Hr2 : µ(η, τ) ≤ T .
Assume that T is sufficiently large satisfying the conditions that, for all Γ-cusps
η ∈ P1 (K), W(η, T ) ⊆ Sη and that W(η, T ) W(η0 , T ) = 0
T
∅ if [η] , [η] . Ac-
cordingly, the boundary ∂S T consists of h components iη Γη ∂W(η, T ) , each of
dimension 2r1 + 3r2 − 1. Moreover,
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 101
Yr1 Yr2
1 +r2
(v) We let Σ := (t1 , . . . , tr1 : s1 , . . . , sr2 ) ∈ Rr>0 : ti s2j = 1 , and in-
i=1 i=1
troduce a semi-product E := Rr1 × Cr2 n Σ.
n o
(vi) We set Fei (Yi ) := τ ∈ H r1 × Hr2 : ReZ(τ) ∈ Σ, ImJ(τ) ∈ R>T · SUK2 , and let
Fi (Yi ) = Ai · Fei (Yi ).
Clearly, the Fi (Yi )’s are disjoint from each other if the Yi ’s are sufficiently large.
All in all, what we have just discussed proves the following main theorem of
this chapter.
Theorem 4.9. For the action of the modular group SL(OK ⊕ a) on the space
H r1 × Hr2 , we have
We end this chapter by pointing out that our constructions here are compatible
with more general theories of Baily-Borel [8, 9], Borel-Serre [15], Satake [94–96]
and Mumford et al. [1, 54]. We leave the details to the reader.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 103
Chapter 5
Let K be an algebraic number field of degree n (over Q). Denote by OK its ring
of integer, and by ∆K the norm of its discriminant. Let CLK be the ideal class
group of K and let a1 , . . . , ah be the OK -ideals such that CLK = [a1 ], . . . , [ah ] .
Denote by r1 , resp. r2 , the number of real embeddings and complex embeddings
of K. Let S ∞ be the set of Archimedean places of K, and for each σ ∈ S ∞ , set
Nσ = 1, resp. 2, if σ is real, resp. complex. For an element a of K, write (aσ ) =
(aσ )σ∈S∞ , or simply (a(i) ), its image under the canonical Minkowski embedding
K ,→ K ⊗Q R = Rr1 × Cr2 .
Theorem 5.1. All isometric classes of rank two OK -lattices are parametrized by
the modular points in the space
G
F+K × SL(OK ⊕ a) H r1 × Hr2 × R∗+ .
(5.1)
a∈{a1 ,...,ah }
103
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 104
This is simply Theorem 1.1 of Part 1. More precisely, the factors in (5.1)
characterize the following inner structures of rank two OK -lattices.
(1) The factor R∗+ parameterizes the volumes of rank two OK -lattices.
(2) The factor F+K takes care of the actions of lattice automorphisms in the form
εI2 with ε ∈ U K and I2 the identity matrix of size 2.
(3) The factor SL(OK ⊕ a) H r1 × Hr2 parameterizes all isometric classes of rank
two OK -lattices with volume N(a)∆K .
Note that, by scaling a lattice with a constant factor t, the associated volume
for the lattice is changed by a factor tn . Hence, we may equally view the factor
SL(OK ⊕ a) H r1 × Hr2 as a parameter space for all isometric classes of rank two
OK -lattices with volume one.
Proposition 5.1. The isometric classes of all rank two semi-stable OK -lattices
are parametrized by
G
F+K × SL(OK ⊕ a) H r1 × Hr2 ss × R∗+ .
(5.3)
a∈{a1 ,...,ah }
There is a natural metric on the space (5.3). Indeed, the invariant Haar measure
dt
gives a canonical metric on the volume part R∗+ . Moreover, starting from the
t
classical Euclidean metric on Rr1 +r2 −1 , through Dirichlet’s unit theorem, there is
an induced flat metric on F+K . Finally, the product of the hyperbolic metrics on
H r1 × Hr2 induces a natural quotient metric on SL(OK ⊕ a) H r1 × Hr2 , since
canonical hyperbolic metric on H and on H are SL(OK ⊕ a)-invariant.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 105
Lemma 5.1. ( [38]) Let Λ = (OK ⊕ a, (ρσ )) be a rank two OK -lattice of volume
N(a)∆K . We have
(1) For! every rank one sub OK -lattice Λ1 of Λ, there ! exists a non- zero vector
x x
∈ K 2 such that Λ1 is contained in Λ b−1 , where b := OK x + a−1 y.
T
y y
!
x
(2) Λ is semi-stable if and only if, for all nonzero elements ∈ K2,
y
Y x ! 2
σ
≥ N ab2 .
σ∈S
yσ ρσ
∞
Proof. (1) By Lemma 1.3 in Part 1, for a rank one ! sub OK -module Λ1 of OK ⊕ a, !
x T x
there exist a fractional ideal c and an element ∈ K such that Λ1 = Λ c .
2
y y
(See the discussion before Definition 1.7 in Part 1 for the meaning of this notation.)
Accordingly, set b = OK x + a−1 y. By definition, c · x ⊆ OK and c y ⊆ a. Hence
b · c ⊆ OK x + a−1 y · c = c · x + a−1 (c · y) ⊆ OK + a−1 · a = OK .
aσ bσ
∈ SL(2, Kσ ), let ρ = ρg = (ρgσ )σ∈S∞ be the associated metrics on the
cσ dσ
space (R2 )r1 × (C2 )r2 . We have
! ! !−1
2
Y
2 Nσ ∗∗ ab
(x, y) ρ = |aσ xσ + cσ yσ | + |bσ xσ + dσ yσ |
2
=N (IJ) .
Λ
σ∈S ∞
xy cd
(5.6)
1
= .
(bσ xσ + dσ yσ ) + (aσ xσ + cσ yσ )2
2
This latest expression is simply equal to the factor associated to the real place
σ on the right hand side of (5.7).
(b) When σ is a complex place, for a point P = Z + V j ∈ H with Z ∈ C and
V ∈ R, if we set M(Z + V j) =: Z ∗ + V ∗ i with Z ∗ ∈ C and V ∗ ∈ R, then
V V
V ∗ := = . Thus, for P = (IJ)σ = j, we have
|cσ Z + dσ | + |σ | V
2 2 2 kcσ P + dσ k2
Z = 0, V = 1, and the corresponding (V ∗ )2 is simply given by:
!2
1
|(aτ xτ + cτ yτ ) · 0 + (bτ xτ + dτ yτ )|2 + |aτ xτ + cτ yτ |2 · 12
!2
1
= .
|bτ xτ + dτ yτ |2 + |aτ xτ + cτ yτ |2
This latest expression is obviously equal to the factor associated to the com-
plex place σ on the right hand side of (5.7).
To complete the proof, it suffices to combine both real and complex places
discussed.
We are now ready to state the following one of the main results in this Part.
As an example, we here reproduce the classical reduction theory with our lan-
guage. Clearly, the moduli space of isometric classes of rank two Z-lattices can
be proved to be identified with the double coset space SL(2, Z) SL(2, R) SO(2),
or better, with the quotient space SL(2, Z) H for the natural fractional actions. A
standard fundamental domain D ∈ H is given by:
( )
1 1
D := z ∈ H : |z| ≥ 1, − ≤ x < r {y = 1} ∪ {|z| = 1 : x ≥ 0} . (5.9)
2 2
Obviously, for a rank two Z-lattice Λ of volume 1 in the standard Euclidean space
R2 , there exists an element x ∈ Λ\{0} such that its length kxk is equal to the first
Minkowski successive minimum λ1 of Λ, defined by λ1 := minv∈Λ { kλk}. Thus,
with a rotation if necessary, we may and hence will assume that x = (λ1 , 0).
Accordingly, by the classical reduction theory, there exists a vector ω := x0 + iy0
1
in D such that Λ is a lattice of volume λ−2 1 =: y0 generated by (1, 0) and ω.
λ1
To see when these lattices are semi-stable, we introduce, for a fixed real number
T ≥ 1, the subset DT of D by
DT := z ∈ D : y = =(z) ≤ T .
(5.10)
This, in terms of distance to cusps can be simply characterized by
( )
1
DT = z ∈ D : d(∞, z) ≥ . (5.11)
T
Indeed, (since K = Q,) there is one and only one cusp class, namely, ∞, for the
modular SL(2, Z). Consequently, for the lattices corresponding to the points in
DT , since λ−2
1 = y0 ≤ T , we have λ1 ≤ T . In other words,
−2
1
vol(Λ1 )2 ≥ ∀ rank one Λ1 ⊂ Λ. (5.12)
T
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 109
Q,2 1 ' DT .
M≥T
In other words, the moduli space of rank two semi-stable lattices of volume
one is given by the space D1 , the part of D under the line y = 1.
Proposition 5.2. Let T ≥ 1 and η1 , η2 ∈ P 1 (K) be two cusps of SL(OK ⊕ a). Then
the following two statements are equivalent:
Proof. (2) implies (1) is clear. Hence, it suffices to prove the converse that, if
there exists an element τ ∈ H r1 × Hr2 satisfying both d(τ, η1 ) < 1 and d(τ, η2 ) < 1,
then η1 = η2 .
α α
" # " #
Write η1 = 1 , η2 = 2 with α1 , β1 , α2 , β2 ∈ OK . Set b1 := OK α1 + aβ1
β1 β2
and b2 = OK α2 + aβ2 . Then, the conditions that ηi ∈ Xηi (T ) for i = 1, 2 simply
mean that
! ! ! !
∗ ∗ ∗ ∗
N τ · N(a b1 ) > 1,
−1 2
N τ · N(a−1 b22 ) > 1.
−β1 α1 −β2 α2
Moreover, by definition, we have
α2 α∗2 β∗2 −α∗2
! ! ! ! !! !
∗ ∗ ∗ ∗
N (τ) · N(a b1 ) = N
−1 2
· · (τ) · N(a−1 b21 )
−β1 α1 −β1 α1 β2 β∗2 −β2 α2
! !
∗ ∗
N (τ)
−β2 α2
! ! !
∗∗ ∗ ∗
=N · (τ) · N(a b1 ) =
−1 2
· N(a−1 b21 ).
c d −β2 α2 2
!
∗ ∗
c· (τ) + d
−β2 α2
Here c := α1 β2 − β1 α2 and d := −β1 α∗2 + α1 β∗2 .
!
∗ ∗
Hence, if we set τ0 := (τ), it suffices to prove the following
−β2 α2
Lemma 5.4. With the same notation as above, assume that the following three
conditions are satisfied:
Proof. Indeed, much stronger results, namely, the corresponding inequality for
each Archimedean place, can be verified by direct calculations. We only check
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 111
local inequalities for real places, since complex cases are similar. Thus, let σ be a
real place. Then, by definition,
2
cσ zσ + dσ = (cσ xσ + dσ )2 + c2σ y2σ ≥ c2σ y2σ ,
as desired.
Hence, by condition (ii) (of the lemma), N(τ0 ) · N(a−1 b21 ) > N(c)2 · N(τ0 )2 .
This means that N(a−1 b21 ) > N(c)2 · N(τ0 ). Therefore, by (i), we further have
N(a−1 b21 ) · N(a−1 b22 ) > N(c)2 , or better, N(a−1 b1 b2 ) > N(c). This contradicts with
(5.13). This completes our proof of the lemma and the proposition.
Then, together with Theorem 5.2, what we have just proved in Proposition 5.2
implies the following:
Theorem 5.3. There is a natural identification between the following two spaces:
(1) The moduli space of rank two semi-stable OK -lattices of volume N(a)∆K with
OK ⊕ a as the underlying projective module.
(2) The truncated compact domain D1 consisting of modular points in the funda-
mental domain D whose distances to all cusps are not less than 1.
5.2.1 Definition
Let MF,2 be the moduli space of semi-stable OK -lattices of rank 2, and, for a fixed
T > 0, let MK,2 [T ] denote the moduli space of rank two semi-stable OK -lattices
of volumes T . By Theorem 5.1 (and Theorem 5.3), we have
G
MK,2 ' F2K × Da,1 × R∗+ (5.14)
a=a1 ,...,ah
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 112
where F2K is the torus U K2 \Rr1 +r2 −1 , and Da,1 ⊆ H r1 × Hr2 is the truncated funda-
mental domain associated to SL(OK ⊕ a). Moreover, as mentioned before, induces
from the standard flat metric on Rr1 +r2 −1 , resp. hyperbolic metrics dhyp on H and
H, there is a natural metric dx on F2K , resp. dµss on D1 . These two metrics, to-
dt
gether with the canonical Haar measure on R∗+ , induce a natural metric dµ on
t
MK,2 . Symbolically, this means
dt
dµ := dx × dµss × . (5.15)
t
By an abuse of notations, sometimes, we also write dµss simply as dµ, if there is
no confusion.
For each rank two semi-stable OK -lattice Λ = (OK ⊕ a, (ρσ )) ∈ MK,2 , by Defi-
nition 2.2(2), the associated 0-th geo-arithmetical cohomology h0 (K, Λ) is defined
by
X X X
h (K, Λ) := log exp − π
0 2
kxσ kρσ − 2π kxσ kρσ ,
2
(5.16)
x∈Λ σ:R σ:C
where x ∈ OK ⊕a and (xσ )σ∈S∞ denotes its image under the Minkowski embedding
OK ⊕ a ,→ K 2 ,→ (R2 )r1 × (C2 )r2 .
Definition 5.3. Let K be a number field. The rank two non-abelian zeta function
ζK,2 (s) of K is defined by:
b
Z 0
ζK,2 (s) :=
b eh (K,Λ) − 1 · vol(Λ) s dµ(Λ), <(s) > 1. (5.17)
Λ∈MK,2
By §3.2.2, the integrations above are all convergent. Hence, b ζK,2 (s) is well-
defined. More generally, we have the following results on basic zeta properties
ζK,2 (s).
of b
(1) (Meromorphic Continuation) Rank two non-abelian zeta function b ζK,2 (s) is a
well-defined holomorphic function when <(s) > 1, and admits a meromor-
phic continuation, denoted also by b ζK,2 (s), to the whole complex s-plane.
ζK,2 (s) satisfies the standard functional equation
(2) (Functional Equation) b
ζK,2 (1 − s) = b
b ζK,2 (s).
(3) (Singularities and Residues) b ζK,2 (s) admits only two singularities. More pre-
cisely, it admits two simple poles, namely, at s = 0, 1, with the residue
ζK,2 (s) = vol MK,2 ∆K .
Res s=1b
1n 1n
Γ∞ = : n ∈ Z , a typical element of Γ∞ acts on z ∈ H by a shift
01 01
z 7→ z + n. Hence, if f (z) is an automorphic function, f (Az) is a periodic function
in z of period 1, i.e. it satisfies the condition that f (A(z + 1)) = f (Az). Therefore,
if we define the m-th Fourier coefficient of f by
Z 1 √
am (y) := f (Az) e(−mx) dx, z = x + −1y (5.20)
0
√
where e(x) = exp(2π −1x), we obtain a Fourier expansion of f at the cusp κ
X
f (Az) = am (y) e(mx), (5.21)
m∈Z
provided that f satisfies the following growth conditions:
where
Z 1 Z ∞ X
am (y, s) = E(z, s) e(−mx) dx = y(γz) s e(−mx) dx. (5.24)
0 0 γ∈Γ∞ \Γ
!
GG ab
Γ∞ Γ Γ∞ = Γ∞ Γ∞ γΓ∞ , c > 0, d (mod c), γ = ∈ Γ.
c,d
cd
!
ab
Proof. Write Γ∞ Γ Γ∞ = γ Γ∞ γΓ∞ with γ = ∈ Γ. Since Γ does not
F
cd
!
t b
contain any hyperbolic element , t > 0, the double coset Γ∞ γΓ∞ is equal
0 t−1
!
1b
to Γ∞ Γ∞ if c = 0. On the other hand, if c , 0, we may assume c > 0,
01
a0 b0
!
by replacing γ with −γ if necessary. Moreover, for γ0 = 0 0 with c0 > 0,
c d
Γ∞ γΓ∞ = Γ∞ γ0 Γ∞ if and only if c = c0 and d ≡ d0 (mod c). This gives the
decomposition in the lemma. To prove (1), it sufficient to note that, for a given
c > 0, there exist only finitely many d’s which are incongruent mod c, such that
(c, d) is the second row of some γ ∈ Γ. To prove (2), we use the fact that Γ∞ γγ0 and
Γ∞ γγ00 are different cosets in Γ∞ \Γ, if γ and γ0 are different elements of Γ∞ .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 116
Consequently, we have
Z ∞ X
am (y, s) = δ0m y s +
a
y(γz) s e(−mx) dx
−∞ b
γ= ∈Γ∞ \Γ/Γ∞ , c,0
c d
∞
XZ ys
= δ0m y s + e(−mx) dx
−∞ |cz + d|
2s
c,d
X 1 X md ! Z ∞ ys
= δ0m y +
s
e · e(−mx) dx
|c|2s d c −∞ |x2 + y2 |2s
c
X 1 X md ! Z ∞ 1
= δ0m y + y
s 1−s
· e · e(−mt)dt,
c
|c|2s
d
c −∞ (1 + t2 ) s
!
∗∗
where ∈ Γ and c > 0, d (mod c). For later use, we set
cd
1 Γ(s − 2 )
X 1 X md ! 1
φm (s) = · e and φ(s) := π 2 φ0 (s). (5.25)
c
|c|2s d c Γ(s)
To state the Fourier expansion for Eisenstein series, as usual, we define the
modified Bessel K-function K s (z) by
∞
π I−s (z) − I s (z) X ( 21 z) s+2m
K s (z) := with I s (z) := . (5.26)
2 sin sπ m=0
m!Γ(s + m + 2)
Proposition 5.4. (see e.g. [62]) The m-th coefficient of the Fourier expansion of
E(z, s) is given by
s
2π 1 1
φm (s) · |m| s− 2 y 2 K s− 12 (2π |m|y) m , 0
Γ(s)
am (y, s) =
y s + φ(s)y1−s
m = 0.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 117
1 Γ(s − 2 )
X 1 X md ! 1
φi j,m (s) = · e , and φ ij (s) := π 2 φi j,0 (s),
c
|c|2s d c Γ(s)
with a similar calculation as above, we obtain the following:
Proposition 5.5. (see e.g. [62]) The m-th coefficient of the Fourier expansion of
Ei (z, s) with respect to the j-th cusp κ j is given by
s
2π 1 1
φi j,m (s) · |m| s− 2 y 2 K s− 12 (2π|m|y) m , 0
Γ(s)
ai j,m (y, s) =
δi j y s + φi j (s)y1−s
m = 0.
For later use, we set Φ(s) := (φi j (s)). This matrix is symmetric, since it is
invariant under the involution map x 7→ x−1 .
To end this discussion, we mention that there is an alternative approach to
obtain the Fourier expansion above by using the fact that the Eisenstein series E
is a solution of the partial differential equation ∆E = λE with λ = s(s − 1). This
implies that the coefficients ai j,m (y, s) satisfy the condition
d2 ai j,m λ
2
− 4π2 m2 + 2 ai j,m = 0. (5.28)
dy y
Here ∆ denotes the hyperbolic Laplacian in §4.1.1. Using the solutions of this
well-known second order ordinary differential equation, based on the growth be-
haviors of the Eisenstein series near the cusps, we would also arrive at the conclu-
sions that the constant terms in the Fourier expansion are some combinations of
y s and y1−s and that the non-constant terms are essentially the Bessel K-functions.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 118
We start with the following analogue of Proposition 5.4 for functions over H. Let
Λ be a full rank lattice in C and denote Λ∨ the dual lattice of Λ defined by Λ∨ :=
µ ∈ C : hµ, zi ∈ Z, ∀z ∈ Λ , where h , i denotes the usual Hermitian product on
C. Let f : H → C be a Λ-invariant C 2 -function, satisfies the differential equation
Here ∆ denotes the hyperbolic Laplacian on H. (See e.g. §4.1.2 for the precise
definition.) Assume that f (z + r j) is of slow growth as r → ∞, that is, as r → ∞,
f (z + r j) = O(rk ) for a constant k uniformly with respect to z ∈ C.
Write the hyperbolic Laplacian ∆ in terms of the coordinates z, r. Then the con-
dition (5.29) implies that the qµ (r)’s satisfy the following second order ordinary
differential equations with λ = 2s(2 − 2s)
2
!
2 d d
r 2 − r + λ − 4π |µ| r · qµ (r) = 0.
2 2 2
(5.31)
dr dr
It is well known that, see e.g. [25], the series on the right hand converges for
<(s) > 1, is of slow growth and satisfies the partial differential equation
∆E = −λE. (5.37)
Moreover, if η = B (∞) ∈ P (C) is a distinct cusp of Γ,
−1 1
1ω
( ! )
(BΓB )∞ = BΓη B =
−1 0 0 −1
:ω∈Λ , (5.38)
01
for a certain rank two lattice Λ ⊂ C. Since the function P 7→ E A (B−1 P, s) is
Λ-invariant, the Eisenstein series E A (B−1 P, s) admits a Fourier expansion:
X
E A (B−1 P, s) = a0 r2s + b0 r2−2s + aµ r · K2s−1 (2π|µ|r2 ) · e hµ, zi . (5.39)
0,µ∈Λ∨
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 120
Theorem 5.5. ([25]) For <(s) > 1, the Fourier expansion of the Eisenstein series
E A (B−1 P, s) is given by
π X
E A (B−1 P, S ) = δη,ζ [Γζ : Γ0ζ ]|d0 |−4s r2s + |c|−4s r2−2s
vol(P) s ∗ ∗∈R
cd
2s d
1 2π X X e(−hµ, i)
+ 2s−1 2π|µ|r · e hµ, zi .
c
|µ|2s−1 · 2
· rK
vol(P) Γ(2s) 0,µ∈Λ∨ ∗ ∗
∈R |c|
4s
c d
1 X Z
aµ (r, s) = r AMB−1 (z + r j) 2s · e − hµ, zi dx dy.
Proof. Set
vol(P) M∈Γ0 \Γ P
ζ
Then
X
E A (B−1 P, S ) = aµ (r, s) e hµ, zi , ∀P = z + r j ∈ H.
(5.42)
µ∈Λ∨
Z !2s
X r
=
· e hµ, zi dx dy
∗ ∗ |cz + d| + |c| r
2 2 2
∈R C
c d
!2s
e(hµ, dc i)
Z
X r
=
2 + r2
· e |µ| · x dx dy,
∗ ∗ |c| 4s
C |z|
∈R
c d
Remark 5.1. The factor [Γζ : Γ0ζ ] takes care of elliptic points, which are factored
out in our discussion. In the case of SL(OK ⊕ a), these elliptic points are induced
from units, since the fields involved now are imaginary quadratic fields (hence by
Dirichlet’s unit theorem, there are only finitely many of them.) We thank Elstrodt
for explaining this to us.
We next apply the results above to obtain Fourier expansion of Eisenstein series
associated to Γ := SL(2, OK ), where K is an imaginary quadratic field, following
closely [25].
!2s
X r
ba (P, s) :=N(a)2s
E <(s) > 1.
c,d∈ar{(0,0)}
kcP + dk2
(2) We define the partial zeta functions ζa,b (s) associated to a, b, and ζ[a] (s) asso-
ciated to [a] ∈ CLK by:
X X
ζa,b (s) := N(ab−1 ) s N(λ)−s , ζ[a] (s) := N(b)−s <(s) > 1.
λ∈ab−1 r{0} b∈[a],b⊆OK
(ii) The function ζa,b (s) depends only on the ideal classes [a], [b] ∈ CLK .
(iii) We have
ζa,b (s) = ζa·b−1 ,OK (s) and ζa,OK (s) = |U K | · ζ[a−1 ] (s).
In particular, ζa,b (s) = |U K | · ζ[mn−1 ] (s).
ba (P, s), Eb (P, s), ζa,b (s) and ζ[a] (s) are
As to be expected, the functions E
closely related.
This is well-defined and gives a surjective map. Easily, every (c, d) admits pre-
cisely |U K | different inverse images.
Consequently, to obtain the Fourier expansions for the Eisenstein series above,
it suffices to treat Eb (P, s). For this, we next point out that the function Eb (P, s)
αβ
!
is essentially the Eisenstein series E A (P, S ). Indeed, for A = ∈ SL(2, K),
γδ
let aA = OK γ + OK δ, bA = OK α + OK β. In this way, we obtain two maps from
SL(2, K) to the group of fractional ideals of K, namely, the assignments A 7→ aA
and A 7→ bA , respectively. By Lemma 4.1, these maps are surjective.
δ αβ
" # !
Lemma 5.8. (see e.g. [25]) Let ζ = ∈ P 1 (K) be a cusp of Γ and A = ∈
−γ γδ
SL(2, K) be the corresponding matrix as in Lemma 4.2. Then
1
E A (P, s) = (NaA )−2−2s EaA (P, s) <(s) > 1.
2
2
" #(c, d) ∈ K which generate aA as an OK -module.
Proof. Let L be the set of pairs
δ
!
d
For each (c, d) ∈ L, since A = ∞, there exists an M ∈ Γ such that M =
−γ −c
δ
!
. Consequently, there exists an natural assignment
−γ
φ : L → Γ0ζ \Γ, (c, d) 7→ Γ0ζ M.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 124
With all these, by Theorem 5.5, we can obtain the Fourier expansion of Ea (P, s)
and hence that of E ba (P, s). In the sequel, till the end of this discussion for imag-
inary quadratic field, we let η ∈ P 1 (K) be a cusp of Γ, let B ∈ SL(2, K) be a
matrix satisfying !B(η) = ∞,)and let a be a fractional ideal of K. For later use, set
ω
(
1
BΓ0η B−1 = : ω ∈ Λ , where Λ ⊂ C is a rank two OK -lattice. Let R0 be a
01
maximal system of representatives of (c, d) of a ⊕ a B−1 /BΓ0η B−1 with c , 0.
Corollary 5.1. ( [25]) The Fourier expansion of the Eisenstein series E ba (B−1 P, s)
is given by
Decomposing the right hand side into two parts, depending on whether c = 0 or
not, we get
N(a)2s
Z
X r 2s
aω0 (r, s) = · e hω0 , xi dx dy
|Λ| kdk2
(0,d)∈(a⊕a)B−1 ,d,0 P
(5.45)
N(a)2s
Z
X r 2s
+ 0
,
· e − hω zi dx dy.
P kcP + dk
|Λ| 2
(c,d)∈(a⊕a)B ,c,0
−1
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 125
Observe that the first term on the right hand side of this equation vanishes term-
wise when ω0 , 0, and moreover, when ω0 = 0, it is equal to
X
N(a)2s |d|−4s r2s . (5.46)
(0,d)∈(a⊕a)B−1 ,d,0
of the right hand side in the corollary. As for other terms, it suffices to apply the
same method of the proof for Theorem 5.5 to the second sum in (5.45). we leave
details to the reader.
ba (B−1 P, s) is given by
Theorem 5.6. ([25]) The Fourier expansion of E
(5.47)
21+2s π2s N(aB ) X 2s−1 4π|ω|r2
! !
2ω
+ |ω| σ a,aB (ω, 1−2s)rK 2s−1 √ e h √ , zi .
∆Ks Γ(2s) ∆K ∆K
0,ω∈a2
Let K be a number field of degree n with OK its ring of integer and ∆K the norm of
its discriminant. Denote by S the set of Archimedean places of K, and denote by
r1 , resp. r2 , the number of real embeddings, resp. complex embeddings, of K. We
have n = r1 + 2r2 . Moreover, for an element a ∈ K, denote by aσ its image in the
σ-completion Kσ of K. Similarly, we write its image of Minkowski embedding
as (a(i) ). Denote by U K the unit group of K, and by U K+ the collection of ε ∈ U K
such that εσ > 0 for all real σ ∈ S .
n o
where U K2 := ε2 : ε ∈ U K .
r −times r −times
z1 }| { z2 }| {
where ImJ := (i, . . . , i, j, . . . , j) ∈ H r1 × Hr2 , and k · k is the same as in Defini-
tion 4.8. Since, for ε ∈ U K , kεk = 1, we have
s
s X N(ImJ(τΛ ))
EΛ (s) := ΓR (2s) ΓC (2s) N(a)∆K
b r1 r2
2 . (5.49)
(x,y)∈OK ⊕a/U K r{(0,0)} x · τΛ + y
That is to say,
b2,a (τ, s) = ΓR (2s)r1 ΓC (2s)r2 N(a)∆K s E2,a (τ, s).
E (5.50)
Proof. (1) is rather standard whose proof can be found in any textbook on classi-
cal Eisenstein series. (2) is simply (5.49).
α
" #
To simplify our notation, write Γ = SL(OK ⊕ a). For a cusp η = ∈ P 1 (K), by
β
Lemma 4.2,! we may and hence will choose a (normalized) special transformation
α α∗
A= ∈ SL(2, K) such that, if we set b := OK α + aβ, then OK β∗ + aα∗ = b−1 .
β β∗
Consequently, A(∞) = η, and, by Theorem 4.4, the stabilizer group of η in Γ is
simply
1ω
( ! )
A−1 Γ0η A = : ω ∈ ab−2 . (5.51)
01
Since E2,a (τ, s) is Γ-invariant, it is Γ0η -invariant as well. Hence, E2,a (A(τ), s) is
invariant under the parallel transforms by the elements of ab−2 . This, together
with the slow growth property mentioned above, implies that there is a natural
Fourier expansion for E2,a (Aτ, s) as follows:
X
aω0 ImJ(τ), s · e2πihω ,ReZ(τ)i ,
0
E2,a (Aτ, s) =
(5.52)
ω0 ∈(ab−2 )∨
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 128
where (ab−2 )∨ denotes the dual lattice of ab−2 . Thus, if we denote by P a funda-
mental parallelogram of ab−2 in Rr1 × Cr2 , then, by definition,
1
aω0 (ImJ(τ), s) :=
vol(ab−2 )
!s
(5.53)
Z
X N(ImJ(τ)) −2πihω0 ,ReZ(τ)i Y Y
× e dx σ dxτ dyτ .
(c,d)∈(O ⊕a)A/U P kcτ + dk2
σ:R τ:C
K K
(c,d),(0,0)
1
(1) When c = 0, up to the factor , the corresponding contribution is
vol(ab−2 )
given by
Z !s
X N(ImJ(τ)) Y Y
e−2πihω ,ReZ(τ)i
0
dxσ dxτ dyτ
(0,d)∈(OK ⊕a)A/U K P kdk2 σ:R τ:C
d,0
! sZ (5.54)
X N(ImJ(τ)) Y Y
e−2πihω ,ReZ(τ)i
0
= dxσ dxτ dyτ .
(0,d)∈(OK ⊕a)A/U K
kdk2 P σ:R τ:C
d,0
So depending on whether ω0 = 0 or not, this may be further divided into two
subcases.
(1.a) When ω0 , 0, Z
we have
Y Y
e−2πihω ,ReZ(τ)i
0
dxσ · dxτ dyτ = 0. (5.55)
P σ:R τ:C
That is to say, the corresponding Fourier coefficient aω0 is zero.
(1.b) When ω0 = 0, sinceZ Y Y
dxσ · dxτ dyτ = vol(ab−2 ),
P σ:R τ:C
we have !s
X N(ImJ(τ))
a0 ImJ(τ), s =
(0,d)∈(OK ⊕a)A/U K ,d,0
kdk2
X (5.56)
= N(ImJ(τ)) s N(d)−2s .
(0,d)∈(OK ⊕a)A/U K ,d,0
X
To understand this latest summation N(d)−2s , we first note
(0,d)∈(OK ⊕a)A/U K ,d,0
αα α α∗ ∗
! !
that, by definition, (OK ⊕ a)A = (OK ⊕ a) , where is a matrix in
β β∗ β β∗
SL(2, F) such that, if b = OK α + aβ, then OK β∗ + aα∗ = b−1 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 129
By combining (5.55), (5.56) and Claims 5.2, 5.3, we obtain the following:
Claim 5.4. Let (c, d) ∈ (OK ⊕ a)A and ω ∈ ab−2 . If c , 0, (c, cω + d) ∈ (OK ⊕ a)A.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 130
α α∗
!
Proof. It suffices to deal the component cω + d. By Lemma 4.2, for A = ∈
β β∗
SL(2, F), we have α ∈ b, β ∈ a−1 b, α∗ ∈ ab−1 , β∗ ∈ b−1 . Hence,
X Z !s
1 N(ImJ(τ)) Y Y
· e−2πihω ,ReZ(τ)i
0
= dxσ · dxτ dyτ
Rr1 ×Cr2 kcτ + dk
−2
vol(ab ) 2
∗ ∗ σ:R τ:C
∈R
c d
Therefore,
aω0 ImJ(τ), s
1 X e2πihω0 , dc iZ !s
N(ImJ(τ)) −2πihω0 ,ReZ(τ)iY Y
= −2 2s 2
e dxσ dxτ dyτ .
vol(ab ) N(c) Rr1 ×Cr2 kτk σ:R τ:C
∗ ∗
∈R
c d
(2.a) When ω0 = 0,
Z !s
1 X 1 N(ImJ(τ)) Y Y
a0 ImJ(τ), s = dxτ dyτ .
k dxσ k
vol(ab−2 ) N(c)2s Rr1 ×Cr2 kτk2 σ:R τ:C
∗ ∗
∈R
c d
Z !s
N(ImJ(τ)) Y Y
· e−2πihω ,ReZ(τ)i
0
× 2
dxσ · dxτ dyτ .
Rr1 ×Cr2 kτk σ:R τ:C
(2.b.ii) Over each Archimedean place R, we have, by using the calculation over R
in (5.61),
Z !2s Z !2s
r −2πi|ω0 |·x r 0
2 + r2
e dx dy = 2 + y2 + r 2
e−2πi|ω |·x dx dy
C |z| R2 x
Z !2s
1 0 x x y
= y 2 r−2s e−2πi|ω |· r ·r d d · r2
R2 ( r ) + ( r ) + 1
x 2 r r
0
e−2πi|ω |·x·r
Z
=r 2−2s
dx dy
2 (x + y + 1)
2 2 2s
ZR Z
dy 0
= r2−2s 2 + x2 + 1)2s
· e−2πi|ω |rx dy dx
R R (y
Z Z √
1 y x + 1 −2πi|(ω0 )|rx
2
=r 2−2s
d√ e dx (5.62)
x2 + 1 (x + 1)
2 2 2s
R R ( √ y + 1) 2s
x2 +1
0
e−2πi|ω |rx
Z Z !
dt
=r 2−2s
· dx
R (1 + t )
2 2s 1
R (x2 + 1)2s− 2
1 Γ(2s − 21 )
Z 0
e−2πi|ω |rx
=r 2−2s
· π · 2
· dx
Γ(2s)
1
R (x2 + 1)2s− 2
2 + r2
e−2πi|ω |·x dx dy
C |z|
Z 2s
1 de−2πi|ω0 |· rx ·r x · rd y · r
=r −2s
C ( ) +1
|z|
2 r r
Z r
1 0 (5.63)
= r2−2s e−2πi|ω |·x·r · dx dy
C (1 + |z| )
2 2s
Theorem 5.7. The Fourier expansions for the Eisenstein series E2,a (Aτ, s) around
the cusp A(∞) is given by:
e2πihω , c i
0 d
1 X 1 1
+ · N(ImJ(τ)) 2 · N(ω0 ) s− 2
vol(ab−2 ) ∗ ∗∈R N(c)2s
c d
!r !r
2π s 1 Y 0 2π2s |ω0 |2s−1 2 Y
× K 1 (2π|ω |σ yσ ) K2s−1 (2π|ω0 |τ rτ ).
Γ(s) σ:R s− 2 Γ(2s) τ:C
Using technics in [25], we can write down all the Fourier coefficients explic-
itly, and give them arithmetic interpretations. However, for our application to the
rank two zeta functions in this book, this theorem on the constant term is already
sufficient, we omit the details to the reader.
Let Γ be a congruence subgroup of SL(2, Z). Induced from the canonical action
of SL(2, R), there is a natural action of Γ on H. Denote by κ1 = ∞, κ2 , . . . , κh
the inequivalent cusps of Γ. For each i = 1, . . . , h, let Γi := Γκi be the isotropy
group of κi , and fix an Ai ∈ GL2 (Q) satisfying A1 = I2 and Ai (∞) = κi . Obviously,
i Γi Ai = Γ∞ . In addition, we assume that
A−1
!
Γ is reduced at infinity, i.e. the
11
stabilizer group Γ∞ of ∞ is generated by .
01
Recall that the Eisenstein series Ei (z, s) of Γ at κi is defined by
X
Ei (z, s) := i γz),
y s (A−1 (5.64)
γ∈Γi \Γ
where y(x + iy) := y. For example, E1 (z, s) = E(z, s) is the Eisenstein series at ∞.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 134
Set
E1 (z, s)
E (z, s)
E(z, s) := 2 . (5.65)
· · ·
Eh (z, s)
By Proposition 5.5, there is a matrix Φ(s) of functions such that
E(z, s) = Φ(s)E(z, 1 − s), Φ(s)Φ(1 − s) = Ih (5.66)
where
X 1 X
Φ(s) = (φi j )h×h and φi j,0 (s) = 2s
∗ ∗
1. (5.67)
|c|
∈Ai ΓA j , d (modc)
−1
c>0
cd
Let F(z) be a Γ-modular function. Recall that F(z) is of slow growth at infinity,
i.e. it satisfies the condition that
F(Ai z) = ψi (y) + O(y−N ), as y = =(z) → ∞, ∀N. (5.68)
In the sequel, till the end of this subsection, we assume, in addition, that, as z → 0,
the ψi (y)’s admit the following asymptotic expansions
l
X ci j αi j
ψi (y) = y logni j y, where ci j , αi j ∈ C, ni j ∈ Z≥0 . (5.69)
j=1
ni j !
(1) The function vector R(F, s) is well-defined for <(s) sufficiently large.
(2) (Functional Equation) R(F, s) = Φ(s)R(F, 1 − s).
1
(3) Each component of R(F, s) = h(s) + Φ(s)h(1 − s) + is entire.
ζ(2s)
s(s − 1)b
Proof. Let D be the standard fundamental domain for the action of SL(2, Z) on
H, and let DΓ be a fundamental domain of Γ with |x| ≤ 21 . Set
To compute the summation on the right hand side, we divide it into two parts.
!
ab
(1) When a/c ∼ ∞, let γ0 = ∈ Γ. Then γ0−1 S a/c = z ∈ H : =(z) > T .
cd
Hence we have
" Z ∞Z ∞
F(z)y s dµ = F(z)=(γ0 z) s dµ
S a/c T −∞
Z ∞ Z 1/2 ∞
X " X
= F(z) =(γ0 (z + n)) s dµ = F(z) =(γz) s dµ,
T −1/2 S ∞ (T )
a ∗
n=−∞ γ= ∈Γ
c ∗
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 136
! !
a∗ 1n
where the sum runs over all γ = ∈ Γ, in the form γ0 for some n ∈ Z.
c∗ 01
Therefore, by an obvious relation
! G
G G G a ∗
Γ∞ \Γ/{±1} = I2 ,
a ∗
∗
c>0 a modz, a/c∼∞ ∈Γ c
c∗
we have
∞
X X " "
F(z)y dµ =
s
F(z) · E(z, s) − y s dµ.
c=1 a(modc),(a,c)=1,a/c∼∞ S a/c S ∞ (T )
Therefore,
Z Z T h "
X h i
F(z)E(z, s)dµ = a∞
0 (y)y
s−2
dy − F(Ai z) E(Ai z, s) − δi∞ y s dµ,
DT 0 i=1 S ∞ (T )
Z T Z 1 Z T
since we have easily F(z)y s dµ = a∞
0 (y)y
s−2
dy.
0 0 0
Z 1
To go further, let e j∞ := e j∞ (y, s) := E(A j z, s)dµ be the constant term in
0
the Fourier expansion of E(z, s) at κ j . By Proposition 5.4, ei∞ = δi∞ y s + φi∞ y1−s .
Thus, for <(s) sufficiently large, we have
Z Z T Xh " h i
F(z)E(z, s)dµ = a∞0 (y)y s−2
dy − F(Ai z) E(Ai z, s) − ei∞ dµ
DT 0 i=1 S ∞ (T )
h "
X
− F(Ai z)φi∞ y1−s dµ
i=1 S ∞ (T )
Z T h "
X h i
= a∞
0 (y)y
s−2
dy − F(Ai z) E(Ai z, s) − ei∞ dµ
0 i=1 S ∞ (T )
h
X Z ∞
− φi∞ ai0 (y)y−1−s dy.
i=1 T
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 137
h i
The difference E(Ai z, s) − ei∞ is an entire function of s and is of rapid decay with
respect to y. Hence we have
Z h "
X h i
F(z)E(z, s)dµ + F(Ai z) E(Ai z, s) − ei∞ dµ
DT i=1 S ∞ (T )
(5.74)
Z T h
X Z ∞
= a∞
0 (y)y
s−2
dy − φi∞ ai0 (y)y−1−s dy.
0 i=1 T
l
ci j ∂ni j
Z T
T s+αi j −1
X !
Set hiT (s) := ψi (y)y dy =
s−2
and hT (s) = (hiT (s)).
0 j=1
n i j ! ∂sni j s + α − 1
i j
We have
Z T Z T Z T
ai0 (y)y s−2 dy = ai0 (y) − ψi (y) y s−2 dy + ψi (y)y s−2 dy
0 0 0
Z ∞
= Ri (F, s) − ai0 (y) − ψi (y) y s−2 dy + hiT (s).
T
Z ∞ Z T
On the other hand, ψi (y)y dy = − −s−1
ψi (y)y−s−1 dy = −hiT (1 − s),
Z ∞ T 0
Moreover, since
" X h Z ∞ X Z ∞
F(Ai z)ei∞ dµ = a∞
0 (y)y
s−2
dy + φi∞ ai0 (y)y−s−1 dy,
S ∞ (T ) i=1 T i T
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 138
we have
h
X
R∞ (F, s) + h∞
T (s) + φi∞ hiT (1 − s)
i=1
Z h "
X h i
= F(z)E(z, s)dµ + F(Ai z) E(Ai z, s) − ei∞ dµ
DT i=1 S ∞ (T )
Z ∞ X h Z ∞
+ a∞ − ψ∞ (y) y s−2 dy + φi∞ ai0 (y) − ψi (y) y−1−s dy
0 (y)
T i=1 T
Z h "
X
= F(z)E(z, s)dµ + F(Ai z)E(Ai z, s) − ψi (y)ei∞ dµ.
DT i=1 S ∞ (T )
Similarly, by working with other cusps, we obtain
Z h "
X
F(z)Eκ (z, s)dµ + F(Ai z)Eκ (Ai z, s) − ψi (y)eiκ dµ
DT i=1 S ∞ (T )
(5.76)
h
X
= Rκ (F, s) + hκT (s) + φiκ hiT (1 − s).
i=1
Put all these in a vector form, what we just said then yields
Z Xh "
F(z)E(z, s)dµ + F(Ai z)E(Ai z, s) − ψi (y)ei (y, s) dµ
DT i=1 S ∞ (T )
Take the spacial case when F ≡ 1 is the constant function, this gives
Proof. The first equality is a direct consequence of (5.74). To prove the second
equality, it suffices to note that, for F ≡ 1,
(i) F(Ai z)Eκ (Ai z, s) − ψi (y)eiκ = Eκ (Ai z, s) − eiκ , whose integration over S ∞ (T ) is
simply zero:
(ii) Rκ (F, s) is also zero by definition.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 139
Definition 5.8. Fix T ∈ R>0 . We define a T -version of rank two zeta function
ζQ,2
b T
(s) of Q by
Z
ζQ,2
b T
(s) := b s) dx ∧ dy
E(z, <(s) > 1.
≤ 21 log T
MQ ,2 1 y2
ζQ,2
This T -version of rank two zeta function b T
(s) can be calculated explicitly
in terms of the classical Riemann zeta function.
ζQ,2
Proposition 5.9. For the generalized zeta function b T
(s), we have
Let H be the three dimensional hyperbolic upper half space. For the hyperbolic
dx2 + dy2 + dr2
metric ds2 := , its volume form, resp. its Laplace operator, is
r2
given by
2 ∂ ∂2 ∂2 ∂
2
!
dx ∧ dy ∧ dr
dµ := resp. ∆ := r + + − r . (5.79)
r3 ∂x2 ∂y2 ∂r2 ∂r
Lemma 5.11. (Stokes’ Formula for Hyperbolic Geometry) With the same nota-
tions as above, we have
$
dx ∧ dy ∧ dr
((∆ − 2s(2s − 2)) f · g − f (∆ − 2s(2s − 2)g)
D r3
"
∂ ∂ ∂ ∂
! !
dr dr
= f · g − f · g dy ∧ + f ·g− f · g ∧ dx
∂D ∂x ∂x r ∂y ∂y r
1 ∂ 1 ∂ !
+ f ·g− f · g dx ∧ dy.
r ∂r r ∂r
b s) − 1 · 0 dx ∧ dy ∧ dr
= −2s(2s − 2) · E(P,
r3
$D
b s) dx ∧ dy ∧ dr
= (∆ − 2s(2s − 2))1 · E(P,
b s) − 1 · (∆ − 2s(2s − 2))E(P,
D r3
"
∂ b ∂ b 1 ∂ b
!
dr dr
= − E(P, s) dy ∧ − E(P, s) ∧ dx − E(P, s) dx ∧ dy.
∂D ∂x r ∂y r r ∂r
For a fixed T > 0, introduce a truncated domain D = DT ⊂ DT , a compact
subset of D, by cutting off the neighborhoods of cusps defined by r > T for the
cusp ∞, and its Ai -transform for other cusps κi , where Ai (∞) = κi , i = 1, . . . , h. In
terms of the notations in §4.4.4.2,
[h
DT := D r F
fη (T ),
i
(5.82)
i=1
and, we may and hence will assume that T is sufficiently large to guarantee
i=1 Fηi (T ) =
Sh f Fh f
i=1 Fηi (T ). Then, the boundary of DT may be classified into two
different types. Namely, the first, denoted as the Pi (T )’s, which are compact sur-
faces obtained from the intersection of the cuspidal neighborhood for ∞ with the
hyperplane defined by r = T , and its association for other cusps; and the second,
denoted by S σ,µ ’s, which are defined by the conditions µ(σ, P) = µ(τ, P) for pairs
of inequivalent Γ-cusps (σ, µ). It is well known that, see e.g. [25], since the group
Γ is finitely generated, we can and hence will pair the boundary surfaces S σ,µ ’s
to satisfying the condition that, within each pair, the oriented norm directions for
their components are opposite to each other, using the fact that the angles mea-
sured in hyperbolic geometry and in Euclidean geometry are the same. Therefore,
when calculating the integrations in Example 5.1, after cancelling the contribu-
tions from two boundary surfaces in every pair appeared in the S σ,µ ’s, what is left
on the right hand side of Example 5.1 is merely the part of the integrations over
the first type of the boundaries Pi (T )’s. Furthermore, for the induced integrations
over Pi (T )’s, we have the following:
Lemma
" 5.12. With the same notation as " above,
∂ b dr ∂ b dr
E(P, s) dy ∧ = 0 and E(P, s) ∧ dx = 0. (5.83)
Pi (T ) ∂x r Pi (T ) ∂y r
In particular,
$ h "
∂ b
!
dx ∧ dy ∧ dr 1 X
E(P, s)
b = r E(P, s) dx ∧ dy. (5.84)
D r3 2s(2s−2) i=1 Pi (T ) ∂r
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 143
Proof. We use the Fourier expansion for E(P, s) given in Theorem 5.7. to divide
these integrations into two parts, one for the constant terms one and the other for
all the non-constant terms.
(1) The part resulting from the constant terms is simply zero. Indeed, by The-
orem 5.7, the constant terms are functions in r (and s) and hence are independent
of x and y. Therefore, the corresponding derivatives of these constant terms in the
x-direction and hence the integrations vanish.
(2) The part resulting from all the non-constant terms is zero so well, even be-
cause of totally different reasons. Indeed, for the non-constant terms, the averages
of exponential functions e2πihω ,zi over z = x + iy ∈ Pi (T ) are all zero.
0
To finally evaluate the integrals (on the right hand side) of (5.84), by (5.47) in
Theorem 5.6, the constant term of E ba (P, s) at the cusp ηi = B(∞) is given by
2π
N(aB )2s ζa,aB (2s)r2s + √ N(aB )2−2s ζa,a−1 (2s − 1)r2−2s
∆K · (2s − 1) B
Let K be an algebraic number field of degree n over Q with OK its ring of integers
of K and DK the absolute value of its discriminant of K.1 Denote by r1 , resp. r2 ,
the number of real, resp. complex, places of K and denote by S ∞ the collection of
all Archimedean places of K. For convenience, we use σ, resp. τ, for real, resp.
complex places in S ∞ . By an abuse of notation, let ∆K be the hyperbolic Laplace
operator for the space H r1 × Hr2 . That is,
X X
∆K := ∆σ + ∆τ , (5.85)
σ:R τ:C
1 Differentfrom the rest of this book, here we use DK instead of ∆K , for the purpose of saving ∆ to
represent Laplacian operators below.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 144
where
∂2 ∂2 ∂2 ∂2 ∂2 ∂
! !
∆σ := y2σ + and ∆τ := rτ2 + + −r . (5.86)
∂xσ2 ∂y2σ ∂xτ2 ∂y2τ ∂rτ2 ∂rτ
Let a1 , . . . , ah be OK -ideals such that their associated ideal classes gives the ideal
group CLK of K. In this case, we use a as a running symbol for the ai ’s. Let
Γ := SL(OK ⊕ a). For the action of Γ on H r1 × Hr2 , denote by CΓ the set of Γ-
cusps, and write CΓ := [η1 ], . . . , [ηh ] for some η1 , . . . , ηh ∈ P 1 (K).
By §5.2.2 and §5.3.3, to obtain an explicit formula for the rank two non-abelian
ζK,2 (s) of K, it suffices to calculate the following integration
zeta function b
$
b2,a (τ, s)dµ(τ).
E (5.89)
DT
Here Eb2,a (τ, s) denotes the complete Eisenstein series in Definition 5.7.
Motivated by what we have done for totally real fields in the previous subsec-
tion, we use the following facts to treat the integration (5.89).
(1) The Eisenstein series Eb2,a (τ, s) is an ‘eigenfunction’ of the associated Laplace
operator.
(2) The Stokes’ formula would transform the integration domain from DT to its
boundary ∂DT .
(3) We divide the boundary ∂DT into two groups. Namely, the one consisting of
the so-called horo-spheres and the others for the rest. For the first group, since
horo-spheres boundaries are paired with opposite normal directions, there are
no contributes to our integration from such boundary components because of
cancellation within each pair.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 145
(4) On the remaining boundary components, the Fourier expansion of the Eisen-
stein series Eb2,a (τ, s) can be used to simplify the calculation. Indeed, in this
case, there are two parts of the variable directions involved, namely, one in
the ReZ direction and another in the ImJ direction. It is easy to see that the
averages of higher degree terms (in the Fourier expansion for the Eisenstein
series E
b2,a (τ, s)) on the boundary ReZ directions are zero.
∂
Here denotes the outer normal derivative and dµ denotes the volume element
∂ν
of the boundary ∂DT , by an abuse of notations.
∂ b
To go further, we next examine the integrand E 2,a (τ, s) in the ImZ directions
" ∂ν
∂ b
and study how the integration E2,a (τ, s) dµ can be simplified. For this,
∂D(T ) ∂ν
as usual, we first proceed our discussions by transforming the cusps η to ∞ using
some special transformations introduced in Lemma 4.2. Clearly, they do not really
change the integrations, since special transformations are elements of SL(2, K).
Thus, without loss of generality, we may assume that η = ∞.
To calculate our integrations more effectively, we need to simply write down
For this purpose, we next introduce some variable changes for the ImJ directions.
Recall that the ImJ directions are simply the yσ -directions for the real places
σ’s, and the vτ -directions for the complex places τ’s. (Here to avoid notational
confusion, we use P = z + v j, instead of P = z + r j, to denote point P ∈ H,
since r may be confused with r1 and r2 for number of real and complex places.)
Moreover, for a point (z1 , . . . , zr1 , P1 , . . . , Pr2 ) in H r1 × Hr2 , its ImJ coordinates
(yσ1 , . . . , yσr1 , vτ1 , . . . , vτr2 ) ∈ Rr+1 +r2 yield a natural norm
Y Y
N(yσ1 , . . . , yσr1 , vτ1 , . . . , vτr2 ) = yσ1 · . . . · yσr1 · vτ1 · . . . · vτr2 2 = v2τ .
yσ ·
σ: R τ: C
1
j =
(i) The entries of the first row is given by e(0) , j = 1, 2, . . . , r1 + r2 :
r1 + r2
1 +r2
rX
(ii) j = 0 for all i = 1, . . . , r1 + r2 − 1: and
e(i)
j=1
1 +r2
rX
( j)
(iii) j log |εk | = δik for all i, k = 1, . . . , r1 + r2 − 1.
e(i)
j=1
2 By the construction of the fundamental domain D in § 4.4.4.2, near cusps, D coincides with funda-
Set accordingly
1 +r2 −1
r1Y
r +r2 1 + j) 2 2Yq
t j := v2τ j = Y0 1 ε(r
q j = 1, . . . , r2 . (5.97)
q=1
Note here, for the complex places τ, the power 2 is added since Nτ = 2.
With respect to this change of variables, from the precise construction in
§ 4.3.4.2, the fundamental domain for the action of Γη ⊂ SL(OK ⊕ a) on H r1 × Hr2
becomes simply the regular domain characterized by
1 1
0 < Y0 < ∞, − ≤ Y1 , . . . , Yr1 +r2 −1 ≤ ,
2 2 (5.98)
(xσ1 , . . . , xσr1 : zτ1 , . . . , zτr2 ) ∈ Fη (ab ),
−2
First we treat the ReZ directions. Since, under the above change of variables,
the changes in such directions are given by a parallel transformation along the
lattices ab−2 , there is no change of volume here. Next we treat the ImJ directions.
Recall that, in general, to find the matrix (g̃i j ) obtained from (gi j ) by the change
of variables, we need to calculate the partial derivatives so as to get g̃i j from the
formula
X ∂xα ∂xβ
g̃i j = gαβ . (5.101)
α,β
∂ x̃i ∂ x̃ j
where
yσr1 tτr2
!
yσ1 tτ1
X0 := ,··· , , ,··· ,
(r1 + r2 )Y0 (r1 + r2 )Y0 (r1 + r2 )Y0 (r1 + r2 )Y0
and, for p, q = 1, 2, · · · , r1 + r2 − 1,
2 (r1 ) 2 (r1 +1) 2 2 (r1 +r2 ) 2 2
X p = log ε(1)
p yσ 1
, · · · , log ε p yσr1
, 2 log ε p vτ1 , · · · , 2 log ε p vτr .
2
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 149
r1 r2
log ε(rp 1 + j) = 0, we have
2
X X
This is rather direct. Indeed, since log ε(i)
p +
i=1 j=1
1
g̃11 = , g̃1 j = g̃ j1 = 0, j = 2, · · · , r1 + r2 . (5.104)
(r1 + r2 )Y02
And, for i, j = 1, . . . , r1 + r2 − 1,
r1 1 +r2
rX
2 2
X
g̃(i+1)( j+1) = 4 log ε(p)
i log ε(p)
j + 4 log ε(p)
i log ε(p)
j . (5.105)
p=1 p=r1 +1
r1 +r2 −1
Consequently, if we set R := log ε(p)
q be the regulator of K, then
p,q=1
4r1 +r2 −1 2
det(g̃i j ) = R. (5.106)
(r1 + r2 )Y02
Therefore, by combining this with contributions in the ReZ directions, we have
s
1 Y Y
dω = det(g̃i j ) · 2 dY0 dY1 · · · dYr1 +r2 −1
dxσ dzτ
Y0 σ:R τ:C
(5.107)
2r1 +r2 −1 dY0 Y Y
= √ R dY1 · · · dYr1 +r2 −1 dxσ dzτ .
r1 + r2 Y02 σ:R τ:C
To make all this useful, we need to find the induced volume form on the bound-
ary ∂DT as well. This is now rather direct. Indeed, by our constructions, the
boundary ∂DT of DT consists of
we have
!
1
ν= √ , 0, . . . , 0 . (5.109)
r1 + r2 T 0
3 The factor N(ab−2 ) is added here, because, in the definition of D , what we used is the distance to
T
cusp, not simply Y0 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 150
∂f ∂f
!
1 √
= √ , 0, . . . , 0 · grad f = r1 + r2 · T 0 . (5.110)
∂ν r1 + r2 T 0 ∂Y0
With the volume form issue settled, we now examine in more details the
boundary ∂DT . By (the fact that the group SL(OK ⊕ a) is finitely generated and)
our concrete construction of the fundamental domain D for Γ, the boundary ∂DT
of DT consists of finitely many of surfaces which are either the parts of the horo-
spheres or the parts Xi (T ) of the planes cutting out by the equation Y0 = T i0 , where
T i0 = N(ab−2i ) · T (with bi the fractional ideal associated to the cusp ηi ). Moreover,
besides the hyperplanes associated to the Y0 = T 0 ’s, the part of the horospheres
(appeared on the boundary) is divided into the sets of inequivalent pairs, within
each of which, the integration of the outer normal derivative along one surface is
equal to the integration of the inner normal derivative along the remaining sur-
face. (In terms of Y p≥1 , in each of the pairs, the two members are determined
by the equations Y p = ± 12 .) Thus, after such calculations, we are left with the
integrations on the Xi (T )’s, where Xi (T ) denotes the part of the boundary of DT
coming from the pull back of the intersection of the hyper-surface Y0 = T i0 with
Fηi , i = 1, 2, . . . , h. That is,
$ $
b2,a (τ, s)dµ(τ) = 1 1
E ∆K E b2,a (τ, s)dµ(τ)
DT r1 + 4r2 s(s − 1) DT
"
∂E
b2,a (τ, s)
= dµ (5.111)
∂DT ∂ν
h "
1 1 X ∂Eb2,a (τ, s)
= dτ.
r1 + 4r2 s(s − 1) i=1 Xi (T ) ∂ν
(Here we used the fact that for T ≥ 1, Xi (T ) are disjoint from each other, proved
in Proposition 5.2.)
Finally, to complete our calculation, we use the Fourier expansion of the
Eisenstein series Eb2,a (τ, s) listed in Theorem 5.7. First, we consider the non- con-
stant terms there. Easily, the average for e2πit (together with its derivative) over an
! ∂Eb (τ,s)
interval of length 1 is zero. Thus, in X (T ) 2,a∂ν dτ, contributions from higher
i
degree terms (in the Fourier expansion) is simple zero. Hence, in the integration
of (5.111), we are only left with contributions resulting from the constant terms
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 151
Here, in the last equality, we have used the fact that the lattice corresponding to
αi
cusp ηi = i with bi = OK αi + aβi and Y p ∈ − 2 , 2 . Therefore,
1 1
is given by ab−2
βi
with the precisely formula we have obtained for A0i (s) and the functional equation
ζ(s), we have proved the following:
for b
Proof. Indeed, by the functional equation, we only need to calculate the coeffi-
T s−1
cient of . By Theorem 5.7, the partial constant term A0,i for the completed
s−1
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 152
T s−1
Hence, up to a constant factor depending only on K, the coefficient of in the
$ s−1
integration b2,a (τ, s)dµ(τ) is simply the summation hi=1 of N(ab−2 )A0,i
E
P
i
DT
twisted by a multiple factor N(ab−2 i )
s−1
, resulting from the discrepancy between T
T s−1
and T i0 . That is to say, the coefficient of is nothing but
s−1
h
X
ΓR (2s)r1 ΓC (2s)r2 N(a)∆K s ζ[a−1 bi ] (2s)N(ab−1 −2s
· N(ab−2 −2 s−1
i ) i ) · N(abi )
i=1
h
X
= ∆Ks · ΓR (2s)r1 ΓC (2s)r2 ζ[a−1 bi ] (2s)
i=1
We end this discussion by pointing out that the method proving Theorem 5.8
is first adopted in [24] to establish the trace formulas related to totally real fields.
In this section, following Suzuki ([111]), we show that all zeros of rank two zeta
functions lie on the central line.
Let n(R) denote the number of zeros of f (z) inside CR , the circle of the radius
R centered at the origin. It is well known that, for an entire function f (z) of order
one, we have, see e.g. [116],
α
(1) n(R) = O(RX ), ∀α > 1,
(2) The series |ρn |−α converges. In particular,
ρn , f (ρn )=0
! ! z 2 !
z z
1− exp =1+O n → ∞. (5.113)
ρn ρn ρn
! !
Y z z f (z)
(3) The infinite product P(z) := 1− · exp converges, and is an
ρn
ρ n ρ n P(z)
entire function of order 1 without zeros, hence is of the form exp(A + Bz) for
some constants A, B.
ζ(s), we
Example 5.2. (See e.g. [23]) For the complete Riemann zeta function b
have
! !
1 Y z z
ζ(s) = eA+Bz ·
s(s − 1) · b 1− · exp . (5.115)
2 ρ
ρ ρ
γ
!
1 1 1
Here A = − log 2, B = − − 1 + log 4π with γ = lim 1 + + · · · + − log n
2 2 n→∞ 2 n
denotes the Euler constant.
Proposition 5.11. ([111]) Assume that the Riemann Hypothesis holds for the Rie-
ζQ,2 (s) of Q lie on
mann zeta function. All the zeros of the rank two zeta function b
the central line <(s) = 21 .
Proof. Let ξ(s) := s(1 − s)b
ζ(s) and F(z) := −ξ i
2 + 2iz .
F(z + i/4) − F(z − i/4) = (1 + 2iz)(−2iz)b ζ(1 + 2iz) − 2iz(1 − 2iz)b ζ(2iz)
ζ(1 + 2iz) b ζ(2iz)
= 2iz(1 − 2iz)(1 + 2iz) ·
b
−
1 + 2iz
2iz − 1
ξ 2(i/2 + iz) ζ(2(i/2 + iz) − 1)
= iz(1 + 4z ) · = iz(1 + 4z2 ) b
ζQ,2 (i/2 + iz),
2
b
−
(i/2 + iz) − 1 i/2 + iz
as desired.
Claim 5.6. Assume that the Riemann hypothesis holds for the Riemann zeta func-
i i
tion. Then, all zeros of F z + −F z− are real. In particular, the Riemann
4 4
hypothesis holds for bζQ,2 (s).
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 155
Proof. By Example 5.2, F(z) is an entire function of order 1. Hence, there are
constants A, B such that ! !
Y z z
F(z) = eA+Bz · 1− · exp .
ρ:F(ρ)=0
ρ ρ
Since the ρ’s are the zeros of complete Riemann zeta function, by the RH for ξ(s),
all the ρ’s are real.
Moreover, since F(z) = −ξ(i/2 + 2iz), we have, for x ∈ R,
F(x) = −ξ 1/2 + 2ix = −ξ 1/2 + 2ix = −ξ(i/2 − 2ix .
That is to say, F(x) takes only real values on x-axis. This implies that both the
constants A and B are real.
If z0 = x0 + iy0 be a zero of F(z + i/4) − F(z − i/4) = iz(1 + 4z2 ) b ζQ,2 (i/2 + iz),
z0 = 0 and/or z0 is a zero of b ζQ,2 (i/2 + iz) since b
ζQ,2 (i/2 + iz) admits simple poles at
z = ±1/2i. Consequently, F(z0 + i/4) = F(z0 − i/4). Hence, by taking the absolute
values for both sides, we arrive at
z0 + i/4 z0 + i/4
Y ! !
A+B(z0 +i/4)
e · 1− · exp
ρ ρ
! ! (5.116)
Y z0 − i/4 z0 − i/4
= eA+B(z0 −i/4) · 1− · exp .
ρ ρ
Since, from above, B ∈ R and ρ ∈ R, we get
Y∞
(x0 − ρn )2 + (y0 − 14 )2
1= .
n=1 (x0 − ρn ) + (y0 + 4 )
2 1 2
Thus, for y0 > 0, resp. for y0 < 0, the right hand side is < 1, resp. > 1. This is a
contradiction. Therefore, we must have y0 = 0.
In this proof, the original RH for the Riemann zeta function
! is used to ensure
z0 + i/4
!
z0 − i/4
that ρ are real. This then forces the relation exp = exp .
ρ ρ
However, as pointed out by Lagarias, by the functional equation, we may pair
the roots ρ and 1 − ρ of the complete Riemann zeta function, or better, to
group the roots ρ, 1 − ρ, ρ and 1 − ρ altogether in the Hadamard product in-
volved and assume further that ! this Hadamard product is taken in the form
Y0 z
F(z) = eA+Bz · 1 − , where 0 means that ρ’s are paired or grouped
Q
ρ:F(ρ)=0 ρ
as above. In this way, we obtain the following unconditional:
ζQ,2 (s) of Q
Proposition 5.12. ([65]) All the zeros of the rank two zeta function b
lie on the central line <(s) = 12 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 156
Therefore, with the same arguments as in the proof for Proposition ref52318, we
have y0 = 0.
ζQ,2
Similar result holds for the functions b T
(s) as well, provided that T ≥ 1. Indeed,
by Theorem 5.8, we have
ζ(2s) ζ(2s − 1) −s
ζQ,2 (s) = ·T .
T
b b
b · T s−1 −
s−1 s
Consequently,
F(z + i/4) · T −i/2−iz − F(z − i/4) · T −i/2+iz = iz(1 + 4z2 ) b
ζQ,2
T
(i/2 + zi).
Therefore, using the same arguments as in the proof above, we have
Y z0 + i/4 z0 + i/4
eA+B(z0 +i/4) ·
1
1− · exp · T −iz− 2
ρ ρ
Y z 0 − i/4 z0 − i/4 1
= eA+B(z0 −i/4) · · T iz− 2 .
1− · exp
ρ ρ
This implies, similarly, that :
Y∞
(x0 − ρn )2 + (y0 − 41 )2 T −y0
1= · y .
n=1 (x0 − ρn ) + (y0 + 4 )
2 1 2 T 0
Or equivalently,
∞
Y (x0 − ρn )2 + (y0 − 14 )2
T 2y0
= . (5.117)
n=1 (x0 − ρn )2 + (y0 + 14 )2
Consequently, for T ≥ 1, we have
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 157
(1) If y0 > 0, the left hand side of (5.117) is > 1. But the right hand side is < 1, a
contradiction.
(2) If y0 < 0, the left hand side of (5.117) is < 1. But the right hand side is > 1, a
contradiction.
ζQ,2
Theorem 5.11. Let T ≥ 1 be a fixed constant. All the zeros of b T
(s) lie on the
central line <(s) = 2 .
1
Obviously, the arguments above works even if the Riemann zeta function is
replaced by a classical zeta or L-function. For example, we have the following:
Theorem 5.12. Let K be a number field. All the zeros of rank two non-abelian
ζK,2 (s) of K lie on the central line <(s) = 12 .
zeta functions b
Hence, all the arguments in the proof of Theorem 5.12 work well.
We end this elementary part by mentioning that, in the rest parts of this book,
the classical constructions and theories appeared here will be generalized (from
the group GL2 ) to all reductive groups, in particular, to the groups GLn . Unlike
the case for GL2 , where there is essentially only one proper parabolic subgroups,
for general reductive groups, the theories and constructions become very sophisti-
cated, because of the complications of the system of parabolic subgroups involved.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 159
PART 3
159
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 161
Chapter 6
Multiple L-Functions
161
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 162
if its unipotent radical over the algebraic closure F of F is trivial. And a sub-
group P of (a reductive group) G is called parabolic if P is closed in the Zariski
topology and the induced quotient space G/P is a complete algebraic variety. It
is well known that a closed subgroup of G is a parabolic subgroup if and only if
it coincides with the normalizer of its unipotent radical. In the sequel, let G be
a connected reductive group defined over F, and call all closed subgroups of G
defined over F as F-subgroups of G. Then, see e.g. [14] and [103], any two mini-
mal parabolic F-subgroups of G are conjugate over F, two parabolic F-subgroups
of G which are conjugate over some extension of the field F are conjugate over
F, and the set of conjugacy classes of parabolic subgroups, resp. of parabolic
F-subgroups, of G has 2r , resp. 2rF elements, where r, resp. rF , denotes the di-
mension of a maximal torus of the commutator subgroup [G, G] of G, resp. the
dimension of a maximal torus of the commutator subgroup [G, G] of G which
splits over F. In addition, each parabolic subgroup P of G admits a Levi decom-
position, namely, it can be represented as a semi-direct product of its unipotent
radical and an F-closed reductive subgroup, called a Levi subgroup of P, and two
Levi subgroups in P are conjugate by an element of P which is rational over F.
For a fixed minimal F-parabolic subgroup P0 of G, denote its unipotent radical
by N0 = NP0 and fix an F-Levi subgroup M0 = MP0 of P0 . Then P0 = M0 N0 is
a natural Levi decomposition of P0 . By definition, an F-parabolic subgroup P of
G is called standard if P contains P0 . For a standard parabolic subgroups P, there
exists a unique Levi subgroup M = MP of P which contains M0 . We call MP
the standard Levi subgroup of P. Similarly, there is a natural Levi decomposition
P = MP NP of P, where N = NP denotes the unipotent radical of P. From now
on, for simplicity, we use the terms parabolic subgroups and Levi subgroups to
denote standard F-parabolic subgroups and standard Levi subgroups respectively,
unless otherwise is stated.
Let P be a parabolic subgroup of G. Write T P for the maximal split torus in
the center ZP of MP and T P0 for the maximal split quotient torus of MP . Then
T P = Hom(Gm , ZP ) ⊗ Gm T P0 = Hom Hom(MPab , Gm ), Gm
and (6.1)
Here MPab denotes the maximal abelian quotient of MP . Moreover, the composition
T P ,−→ ZP ,−→ MP − MPab − T P0 (6.2)
is an isogeny, i.e. a morphism with finite kernel and finite cokernel. Consequently,
there is an injective morphism of free abelian groups of the same finite rank
X∗ (T P ) ,→ X∗ (T P0 ) (6.3)
where X∗ (T ) = Hom(Gm , T ) is the lattice of 1-parameter subgroups in the torus T .
Denote by X ∗ (T P ) := Hom(MP , Gm ) = Hom(MPab , Gm ) the space of characters of
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 163
MP . Easily,
X∗ (T P ) ' Hom(Gm , ZP ) X∗ (T P0 ) ' Hom X ∗ (T P0 ), Z .
and (6.4)
Following [3, 4], we set aP := X∗ (T P ) ⊗ R. Accordingly,
aP = X∗ (T P ) ⊗ R = X∗ (T P0 ) ⊗ R. (6.5)
For later use, denote the R-dimension of aP by d(P).
More generally, if P ⊆ Q are two standard parabolic subgroups of G, there are
canonical morphisms
T Q ,−→ T P ,−→ T P0 − T Q0 . (6.6)
The morphisms T Q ,→ T P and T P0 T Q0 induce a canonical embedding aQ ,→ aP
and a canonical retraction aP aQ . Hence, there exists a canonical splitting
aP = aQ
P ⊕ aQ (6.7)
where aQ
P denotes the kernel of the restriction. Taking the dual R-vector spaces,
we have a∗P = X ∗ (MP ) ⊗ R = X ∗ (T P ) ⊗ R, and hence obtain a canonical splittings
a∗P = aQ∗
P ⊕ aQ .
∗
(6.8)
Similarly, if P ⊆ Q ⊆ R are standard parabolic subgroups of G, there are canonical
splittings
aP = aQ R
P ⊕ aQ ⊕ aR and a∗P = aQ∗
P ⊕ aQ ⊕ aR .
R∗ ∗
(6.9)
P , aQ
We shall denote by [·]Q , [·]RQ and [·]R the canonical projections of aP onto aQ R
P P
and aR respectively. In particular, for a0 := aP0 and a0 := aP0 ,
a0 = a0P ⊕ aQ
P ⊕ aQ and a∗0 = a0P∗ ⊕ aQ∗
P ⊕ aQ .
∗
(6.10)
We have Φ0 := ΦP0 is a root system with Φ+0 := Φ+P0 the set of positive roots and
Φ0 = Φ+0 t Φ−0 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 164
(3) Let ∆0 := ∆P0 ⊂ Φ+0 be the set of simple roots, and let b
∆0 be the subsets of
simple weights in a∗0 .
+
Then ∆0 forms a basis of aG∗ α∈∆0 Z≥0 α, and the elements of ∆0 are non-
P
P0 , P0 ⊂
b
negative linear combinations of elements in ∆0 . In addition, for each α ∈ Φ0 , there
exists a corresponding coroot α∨ .
Being the dual of the collection of simple weights and of simple roots, the sets ∆∨0
∆∨0 are the sets of simple coroots and of simple coweights, respectively, and
and b
are R-bases of the space aG0 .
For a standard parabolic subgroup P, ΦP may not be a root system, even many
constructions for a root system (Φ0 , Φ+0 , ∆0 ) can be parallel introduced.
(5) Let ∆P ⊂ Φ+P be the set of non-trivial restrictions to T P of simple roots (in ∆0 ),
or the same, be the set of non-trivial restrictions to aP of simple roots (in ∆0 ).
Namely, ∆P = α|aP ∈ a∗P : α ∈ ∆0 }r{0}.
(6) For each α ∈ ∆P , choose β ∈ ∆0 such that α = β|aP and set α∨ := [β∨ ]GP be the
projection of β∨ ∈ aG0 to aGP .
Example 6.1. There exists a canonical inner product (·, ·) on aG0 . Since the map
0 . Under this identification, for α ∈ Φ0 ,
a 7→ (a, ·) defines an isomorphism aG0 ' aG∗
2 4
α =
∨
α and (α , α ) =
∨ ∨
. In addition, for α ∈ ∆0 , the corresponding
(α, α) (α, α)
fundamental dominant weight $Gα is characterized by the following properties.
(α, α)
(a) $Gα , $Gα = .
4 !
2 2 G
(b) For all β ∈ ∆0 , δαβ = $Gα , β = $ ,β .
(β, β) (β, β) α
More generally, in stead of (G, P), the above constructions also admit a relative
version for each pair (P, Q) of standard parabolic subgroups of G satisfying P ⊆ Q,
which are compatible with the natural decomposition aP = aQ P ⊕ aQ .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 165
+
(8) Let ΦQP = ΦP∩MQ , resp. ΦP = ΦP∩MQ , resp. ∆P = ∆P∩MQ , be the set of
Q+ Q
+
α ∈ ΦP , resp. α ∈ ΦP , resp. α ∈ ∆P , which occurs in the Lie algebra mQ of
the Levi factor MQ of Q.1
Note that ∆Q P may also be understood as the subset of the elements α ∈ ∆P which
appear in the action of T P on the unipotent radical of P ∩ MQ . Indeed, MQ ∩ P is
a parabolic subgroup of MQ with the nilpotent radical NPQ := NP ∩ MQ . Thus ∆Q P
is simply the set of roots of the parabolic subgroup (MQ ∩ P, T P ).
As the notation suggested, ∆Q Q∗
P forms an R-basis of aP . Furthermore, being an
element of ∆P , the element α ∈ ∆P admits a companion α∨ , the projection of the
Q
For G = SLn (R), let P0 = B be the Borel subgroup of G consisting of all the
upper triangular matrices, and let M0 be the Levi component consisting of the
diagonal matrices. Easily, the unipotent radical N0 of P0 consists of the upper
triangular matrices whose diagonal entries are all 1. Let A0+ be the ‘positive’ part
of the split torus T 0 consisting of the diagonal matrices diag a1 , a2 , . . . , an , where
a1 , a2 , . . . , an ∈ R>0 and i=1 ai = 1. Then,
Qn
n
X
a0 = , , . . . , n
=
H := (H H H ) ∈ : H 0 (6.11)
1 2 n R i
i=1
and the logarithmic map in (6.35) is given by
g 7→ HP0 a(g) = log a1 + log a2 + . . . + log an ,
HP0 : P0 → a0 (6.12)
where, for an element g ∈ G, g = n · m · a(g) · k denotes
its Iwasawa decomposition
with n ∈ N0 , m ∈ M01 , a(g) = diag a1 , a2 , . . . , an ∈ A0+ and k ∈ K = SO(n).
1 To understand (8) and (9), please refer to the following canonical maps A ←- A ,→ M ,→ P =
0 P P
MP NP ,→ Q = MQ NQ ←- MQ . Elements of ∆P are given as the restriction of simple roots of α’s
on A0 to T P . Or equivalently, they are non-trivial characters of T P appearing in nP . Since there are
naturally morphisms nP ,→ p ,→ q ←- mQ , hence nP ∩ mQ makes sense in q, and so does ∆Q P . Here,
as to be easily guessed, p is the Lie algebra of P, etc...
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 166
Write ei ∈ a∗0 for the element satisfying ei (H) = Hi . The set of the associated
simple roots becomes simply
∆0 = e1 − e2 , e2 − e3 , . . . , en−1 − en .
(6.13)
Sometimes, we identify ∆0 with the index set ∆ := {1, 2, . . . , n − 1} via the assign-
ment αi := ei − ei+1 7→ i (i ∈ ∆).
There is a well-known bijection between the ordered partitions of n and the
standard parabolic subgroups of G. More precisely, for an ordered partition
I := n1 , n2 , . . . , nk of n (so that n = n1 + n2 + . . . + nk and ni ∈ Z>0 ), the as-
sociated standard parabolic subgroup PI , or the same, Pn1 ,n2 ,...,nk , is defined to be
the one consisting of the blocked upper triangular matrices whose diagonal entries
are matrices A11 , A22 , . . . , Akk of sizes n1 , n2 , . . . , nk , respectively. Obviously, the
unipotent radical NI of PI is given by the blocked upper triangular matrices whose
diagonal blocks are the identity matrices of sizes n1 , n2 , . . . , nk . Namely, the ma-
trices of the type
A11 ∗ . . . ∗ ∗ In1 ∗ . . . ∗ ∗
0 0 . . . 0 Akk 0 0 . . . 0 Ink
where Ini denotes the identity matrix of size ni . Similarly, the (standard) Levi
component MI of PI consists of the blocked diagonal matrices whose diagonal
blocks are given by the block matrices of sizes n1 , n2 , . . . , nk , and the correspond-
ing AI+ consists of the matrices in the form diag a1 In1, a2 In2 , · · ·, ak Ink . Namely,
the matrices of the type
A11 0 . . . 0 0 a1 In1 0 . . . 0 0
Therefore, the bijection between the parabolic subgroups of G and the subsets
of ∆0 , or better of ∆, is given explicitly as follows: The parabolic subgroup P = PI
corresponds to the subset I(P) of ∆ such that ∆ − I(P) = N1 , N2 , . . . , Nk . That is,
n oGn o
I(P) = 1, 2, . . . , N1 − 1 N1 + 1, N1 + 2, . . . , N2 − 1
G Gn o
··· Nk−1 + 1, Nk−1 + 2, . . . , Nk − 1 .
For simplicity, from now on, we will use I to indicate both the partition and its
corresponding subset I(PI ). For example, the subset I = ∅ corresponds to the
partition (1, 1, . . . , 1) of n = 1 + 1 + · · · + 1 (since ∆ − ∅ = {1, 2, . . . , n − 1}), and the
corresponding parabolic subgroup is simply P0 the minimal parabolic subgroup.
Let J be the set associated to the partition n = f1 + f2 + . . . + fm , where
f1 = n1 +n2 +. . .+nk1 , f2 = nk1 +1 +nk1 +2 +. . .+nk2 , . . . , fm = nkm−1 +1 +nkm−1 +2 +. . .+nkm .
Then, ∅ ⊆ I ⊆ J, and from ∆ − J ⊆ ∆ − I ⊆ ∆, we get P0 = P∅ ⊆ PI ⊆ P J . Then
a J is the subspace of a0 defined by
n o
a J := H1( f1 ) , H2( f2 ) , . . . , Hm( fm ) ∈ a0 . (6.16)
Clearly a J is in aI since aI is a subspace of a0 defined by
n o
aI := A(n 1 , A2 , . . . , An
1) (n2 ) (nk )
∈ a0 (6.17)
and we have the map
aJ → aI
(nk ) (n )
(6.18)
H1( f1 ) , H2( f2 ) , . . . , Hm( fm ) 7→ H1 , H1 , . . . , H1 1 , . . . , Hm km .
(n1 ) (n2 )
where H = H1 , H2 , . . . , Hn ∈ a0 . This implies that b
∆I = $i : i ∈ ∆ − I , since
n o
∆I = αi a : i ∈ ∆ − I . (6.21)
I
∆IJ which is not a subset of ∆∅ in general, even ∆IJ ⊂ ∆I consists of the restriction
of α` to aIJ with ` ∈ J − I = (∆ − I) − (∆ − J). This confirms the definition that
n o
∆IJ = α` aJ : ` ∈ J − I = (∆ − I) − (∆ − J) . In particular, if J = ∆, we recover the
I
relations a∆∅ = a∅ , a∆I = aI and ∆∆∅ = ∆∅ , ∆∆I = ∆I .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 169
Let F be a number field with A = AF its associated adelic ring. For a fixed
connected reductive group G defined over F, denote by ZG the center of G. Let
P0 be a minimal parabolic subgroup of G and denote a Levi decomposition of P0
by P0 = M0 N0 , where, as usual, we fix once and for all the Levi M0 , and N0 is the
unipotent radical of P0 . Recall that a parabolic subgroup P is G is called standard
if P ⊇ P0 , and that for such parabolic subgroups, there exists one and only one
Levi decomposition P = MN satisfying the condition that M ⊇ M0 and N is the
unipotent radical of P. We call M the standard Levi subgroup of P. Denote by
Rat(M) the group of rational characters of M, i.e. the morphism M → Gm , where
Gm denotes the multiplicative group. Obviously, Rat(M) = X ∗ (T P ). Set
a∗M := Rat(M) ⊗Z R = a∗P , a M := HomZ (Rat(M), R) = aP ,
(6.27)
a∗M,C := Rat(M) ⊗Z C, a M,C := HomZ (Rat(M), C).
For convenience, sometimes, we also write
|mv |χv v
Y
|χ| : M(A) → R∗ m = (mv ) 7→ m|χ| := (6.29)
v∈S
Clearly,
There is always a maximal compact subgroup K of G(A) such that, for all standard
parabolic subgroups P = MN of G,
P(A) ∩ K = M(A) ∩ K · N(A) ∩ K . (6.36)
In the sequel, fix such a maximal compact subgroup K of G(A). Accordingly, we
obtain the so-called Langlands decomposition
G(A) = M(A) · N(A) · K, (6.37)
and hence a natural map
mP : G(A) → M(A)/M(A)1 g = m · n · k 7→ M(A)1 · m (6.38)
where g ∈ G(A), m ∈ M(A), n ∈ N(A) and k ∈ K.
Moreover, we may hence will choose some Haar measures on M0 (A), N0 (A),
K such that
(a) The induced measure on M(F) is the counting measure and the volume of
the induced measure on M(F)\M(A)1 is 1. (Recall that, from the classical
reduction theory, M(F)\M(A)1 is of finite volume.)
(b) The induced measure on N0 (F) is the counting measure and the volume of
N(F)\N0 (A) is 1. (Recall that being unipotent radical, N(F)\N0 (A) is com-
pact.)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 171
Accordingly, There are induced Haar measures via log M on the spaces
a M0 , a∗M0 , etc... Furthermore, if we denote by ρ0 the half of the sum of the positive
roots of the maximal split torus T 0 (in the central Z M0 ) of M0 ,
Z
f 7→ f (mnk) dk dn m−2ρ0 dm (6.39)
M0 (A)·N0 (A)·K
defined for continuous functions with compact supports on G(A) induces a Haar
measure dg on G(A). This in turn yields induced measures on M(A), N(A) and
hence on a M , a∗M , P(A), etc..., for all parabolic subgroups P. It is well known that
the following compatibility condition holds
Z Z
f (mnk) dk dn m−2ρ0 dm = f (mnk) dk dn m−2ρP dm
M0 (A)·N0 (A)·K M(A)·N(A)·K
for all continuous functions K with compact supports on G(A), where ρP denotes
one half of the sum of all positive roots of the maximal split torus T P of the central
Z M of M.
functions, which are often divergent. For this, since the space M(F)\M(A)1 is of finite volume, what
should be considered is over the affine part A.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 172
Then, for a fixed compact subset ω ⊂ P0 (A), we can form an associated Siegel set
n o
S (ω; t0 ) := p · a · k : p ∈ ω, a ∈ A M0 (A) (t0 ), k ∈ K . (6.43)
In particular, for a sufficiently big ω and a sufficiently small t0 ∈ M0 (A) (in the
sense that t0α is very close to 0 for all α ∈ ∆0 ), the classical reduction theory
recalled in Chapter 4 may be generalized as
n o (6.45)
S P (ω; t0 ) := p · a · k : p ∈ ω, a ∈ APM0 (A) (t0 ), k ∈ K .
Here, as in §6.1.2, ∆0P denotes the set of positive roots for (P0 ∩ M, T 0 ).
Let F be a number field with A its adelic ring, and let G be a connected reductive
group over F. Denote by G(A) the associated adelic group of G. Since every
algebraic reductive group is a linear algebraic group, we fix, once and for all, an
embedding iG : G ,→ SLn which sends g to (gi j ). Accordingly, introduce a height
function on G(A) by setting
Y
sup |gi j |v ∀i, j .
kgk := (6.47)
v∈S
It is well known that, up to O(1), these height functions are essentially unique.
Because of this uniqueness, all the growth conditions introduced in the sequel are
well-defined, i.e. they do not depend on the height function we use.
By definition, a function f : G(A) → C is said to have moderate growth if
there exist some constants c, r ∈ R such that
if there exist some constants c, r ∈ R, and a certain λ ∈ <X M0 such that, for all
a ∈ A M(A) and k ∈ K,
f (amk) ≤ c · kakr · mP0 (m)λ ∀m ∈ M(A)1 ∩ S P (ω; t0 ). (6.49)
In addition, a complex valued function f : G(A) → C is said to be smooth if, for
every g = gfin · g∞ ∈ G(Afin ) × G(A∞ ), there exist an open neighborhood Vfin , resp.
V∞ , of gfin , resp. g∞ , in G(A) and a C ∞ -function f1 : V∞ → C such that
f (g0fin · g0∞ ) = f1 (g0∞ ) ∀g0fin ∈ Vfin , g0∞ ∈ V∞ . (6.50)
By contrast, a function f : S (ω; t0 ) → C is said to be rapidly decreasing if
there exists a constant r > 0 and, for all λ ∈ <X M0 , there exists a constant c > 0
such that
|φ(ag)| ≤ c · kak · mP0 (g)λ ∀a ∈ A M(A) , g ∈ G(A)1 ∩ S (ω; t0 ). (6.51)
More generally, a function f : G(F)\G(A) → C is said to be rapidly decreasing if
the restriction f S (ω;t ) of f to a Siegel set S (ω; t0 ) in (6.44) is rapidly decreasing.
0
Definition 6.2.
From the special structure of A M(A) -finite functions and the Fourier analysis over
the compact space A M(A) \Z M(A) , we obtain a natural isomorphism
'
C[a M ] ⊗ A(N(A)M(F)\G(A))Z −→ A(N(A)M(F)\G(A))
(6.59)
(Q, φ) 7→ g 7→ Q(log M (mP (g)) · φ(g)
and an isomorphism
'
C[a M ] ⊗ A0 (N(A)M(F)\G(A))Z −→ A0 (N(A)M(F)\G(A))ξ . (6.60)
(See below for a more concrete and useful form.) Accordingly, let Π0 (M(A))ξ be
isomorphism classes of irreducible representations of M(A) occurring in the space
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 175
(1) There is an open cone C ⊂ <XGM such that, if <π ∈ C, the Eisenstein series
E(λ · ψ, π0 ⊗ λ)(g), namely E(ψ; λ; g), converges uniformly for g in a compact
subset of G(A) and λ in an open neighborhood of 0 in XGM . In particular,
nif P = [π] is cuspidal, we may eveno take C to be the positive Weyl cone
λ ∈ <XGM : hλ − ρP , α∨ i > 0, ∀α ∈ ∆GP .
(2) The Eisenstein series E(ψ; λ; g) is an automorphic forms on G(F)\G(A).
For an automorphic form ψ ∈ A N(A)M(F)\G(A) of P-level, we may write
the associated Eisenstein series as a convergence series
X
E ψ, λ, g := mP (δg)λ+ρ · φ(δg)
(6.68)
δ∈P(F)\G(F)
for λ ∈ a∗P,C
satisfying <λ belongs to a certain positive cone CP .
Set P = [π]. For the elements w ∈ W, the associated Weyl group of (G, P0 ),
we fix once and for all some representatives in G(F) of w, which we denote by w
as well, with an abuse of notation. Set M 0 := wMw−1 and denote by P0 = N 0 M 0
the associated parabolic subgroup. Then W acts naturally on the automorphic
representations, from which we obtain an equivalence classes wP of automorphic
representations of M 0 (A).
Proposition 6.1. (§§ II.1.6-7 of [75]) Assume that h<π, α∨ i 0 for all α ∈ ∆GP .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 177
(1) For an automorphic φ, M(w, π)φ depends only on double coset M 0 (F)wM(F).
In particular, M(w, π)φ is well-defined for w ∈ W.
(2) The integral (6.69) converges absolutely and uniformly for g varying in a
compact subset of G(A).
(3) M(w, π)φ ∈ A(N 0 (A)M 0 (F)\G(A))wπ . And if φ is L2 , which from now on we
always assume, so is M(w, π)φ.
(4) Assume that w = wαn · · · wα1 satisfying the condition that, for wi := wαi · · · wα1
and Mi = wi Mw−1 i , we have αi+1 is a simple root of ∆G , w−1 (αi+1 ) > 0 and
−1 P i
the set of positive multiple roots of an element of wi (αi+1 ) : i = 0, . . . , n − 1
coincides with α ∈ ∆GP : α > 0, wα < 0 . Then
where
Example 6.2. Let G = SL3 and take φ to be the constant function 1 on M0 . Being
cuspidal, from above, the associated Eisenstein series E(1, λ, g) and its constant
term along with P0 are given by
E(1, λ, g) =
X
mP0 (δg)λ+ρ0 ∀< λ ∈ C0 ,
δ∈P0 (F)\S L3 (F)
(6.74)
E P0 (1, λ, g) =
X
M(w, λ) · mP0 (g)w(λ+ρ0 ) .
w∈W(M0 )
Lemma 6.1. The morphism Φ induces a bijection between the set of isomorphism
classes of vector bundles of rank n on the curve Spec OF and the double coset
Q
space GLn (F) GLn (Afin ) v∈S fin GLn (Ov ).
(1) The morphism ΦP induces a bijection between the set of isomorphism classes
of filtrations of vector bundles of successive ranks (n1 , n1 + n2 , . . . , n) over X
Q
and the double coset space PI (F)\PI (Afin )/ v∈S fin PI (Ov ).
(2) The natural embedding PI (Afin ) ,→ PI (A) admits a modular interpretation
(E∗ ; α∗,F : (α∗,Ov )v∈S fin ) 7→ (Ek ; αk,F : (αk,Ov )v∈S fin ). (6.77)
(3) The canonical projection PI (Afin ) → MI (Afin ) → GLn j (Afin ) for j = 1, . . . , k,
where MI denotes the standard Levi of PI , admits a modular interpretation
(E∗ ; α∗,F : (α∗,Ov )v∈S fin ) 7→ (gr j (E∗ ); gr j (α∗,F ), gr j (α∗,Ov )v∈S fin ), (6.78)
where gr j (E∗ ) := E j /E j−1 , gr j (α∗,F ) : F n j ' gr j (E∗ )F and gr j (α∗,Ov ) : Ov n j '
gr j (E∗ )Ov , v ∈ S fin are induced by α∗,F and α∗,Ov .
Definition 6.5. Let g = (gfin ; g∞ ) ∈ GLn (A) and let Q be a parabolic subgroup of
GLn . We define pgQ : [0, n] → R to be a polygon characterized by the conditions
Proposition 6.2. Let K = v∈S fin GLn (Ov )×O(n)r1 ×U(n)r2 . For any fixed polygon
Q
p : [0, n] → R satisfying p(0) = p(n) = 0 and d ∈ R, the subspace
M≤p
F,n (d) := g ∈ GLn (F)\GLn (A) K : degar (g) = d, p̄ ≤ p
g
is compact.
Proof. This is a restatement of Theorem 1.3 on the moduli spaces of semi- stable
OK -lattices, based on the classical reduction theory. More precisely, it consists of
two parts, namely, compactness at the lattice level, and compactness of fibers for
the canonical projection Φ.
For the compactness at the lattice level, note that with the degree fixed, by the
Arakelov-Riemann-Roch formula stated at 1.5, the volumes of OF -lattices of rank
n corresponding to the (Eg , ρg )’s for all g ∈ GLn (F)\GLn (A) K are fixed. Hence
the condition pg ≤ p implies that there is an upper bound for the volumes of all
the sublattices of (Eg , ρg ). This further implies that there is a lower bound for the
Minkowski successive minimums for the lattices involved. Therefore, as proved
in Theorem 1.3, at the lattice level, the corresponding space is compact.
On the other hand, fibers of the projection from GLn (F)\GLn (A)/GLn (OF )
to the space of isometry classes of OF -lattices are simply given by the following
product v∈S fin SLn (Ov ) × SOn (R)r1 × SUn (C)r2 , namely K. But it is well-known
Q
that K itself is compact.
More generally, similarly to what L. Lafforgue does for the function fields
in [66], we have the following:
(1) For a fixed parabolic subgroup P of GLn and an element g ∈ GLn (A), there
is a unique maximal element p̄gP among all the pgQ ’s, where Q runs over all
parabolic subgroups of GLn which are contained in P.
(2) For any fixed polygon p : [0, n] → R, a real number d ∈ R, and a standard
parabolic subgroup P of GLn , the subset
n o
MP;≤p
F,n (d) := g ∈ GLn (F)\GL n (A)/K : degar (g) = d, p̄g
P ≤ p, pg
P ≥ −p
is compact.
More generally, for a connected reductive reductive group over F, a similar theory
can be developed. We will leave this to Part 6.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 181
Eisenstein series are analytic invariants of reductive groups over number fields
and the associated theory for the so-called Eisenstein system in [70], see also [90]
and [75], is one of the highest achievements in mathematics. In this section, we
introduce multiple L-functions as integrations of these Eisenstein series over some
intrinsically defined geometric spaces.
For simplicity, we assume that our reductive group G is the general linear
group GLn . As in §6.1.3, we let P0 be the minimal parabolic subgroup P0 consist-
ing of upper triangular matrices corresponding to the partition (1, . . . , 1). Then its
(standard) Levi component M0 consists of diagonal matrices. Let P = PI = U I MI
be the standard parabolic subgroup associated to the partition I = (n1 , . . . , n|P| ) of
n, and denote by MI , resp. NI , the standard Levi subgroup, resp. the unipotent
radical, of PI . (See e.g. §6.1.3) For a fixed irreducible automorphic representation
π0 of MI (A), if π is equivalent to π0 , there exists a λ ∈ XGM such that π ' π0 ⊗ λ,
and π also can be realized as a submodule of the space of automorphic forms of
MI (A). Indeed, A(M(F)\M(A))π = λA(M(F)\M(A))π0 where λ is considered as
a function on M(A). Therefore,
A(N(A)M(F)\G(A))π = λ ◦ mP A(N(A)M(F)\G(A))π0
choose an automorphic form
φ ∈ A(NI (A)MI (F)\G(A))π ∩ L2 (NI (A)MI (F)\G(A)) := A2 (NI (A)MI (F)\G(A))π .
Here, as usual, L2 (NI (A)MI (F)\G(A)) denotes the space of L2 functions on
ZG(A) NI (A)MI (F)\G(A). Following Definition 6.3, we denote by E(φ, π)(g)
the associated Eisenstein series. By Theorem 6.1, E(φ, π)(g) belongs to
A(G(F)\G(A)), namely, E(φ, π)(g) is an automorphic form on G(A).
Being automorphic forms, Eisenstein series are of moderate growth. In par-
ticular, they are generally not integrable over G(F)\G(A)1 . On the other hand,
these series are smooth and hence are integrable over any compact subset of
G(F)\G(A)1 . To make the integration meaningful, we need to find intrinsically
defined compact domains in G(F)\G(A)1 . For this, we fix a convex polygon
p : [0, n] → R. Hence, it makes
h i sense to talk about the p- semi-stable OF -lattices
of rank n. Denote by M≤p F,n 1 the moduli spaces of p-semi-stable OF -lattices of
rank n and volume 1. Then, by Proposition 6.2, there is a natural identification
n o
M≤p
F,n [1] = g ∈ GLn (F)\GLn (A) : degar (g) = ∆F , p̄ ≤ p .
n/2 g
(6.80)
Remark 6.1. (1) When defining multiple L-functions, we assume that φ comes
from a single irreducible automorphic representation. However, such a restriction
is artificial, even it simplifies constructions and particularly results below.
(2) Discussions below for the multiple L-functions L≤p F,n work well for general
Q, ≤ν
multiple L-functions LF, µ (φ, π). In the sequel, for simplicity, we only work with
≤p
LF,n .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 183
With the same notation as in §6.2.3, let G = GLn , set P = [π], and for
w ∈ W = Sn , set M 0 := wMw−1 , and denote by P0 = N 0 M 0 the associated
parabolic subgroup, and hence the intertwining operator M(w, π). By the fun-
damental work of Langlands, these intertwining operators not only satisfy nice
properties as explained in Proposition 6.1, but play an important role in estab-
lishing the function equations of Eisenstein series. Consequently. we obtain the
following basic properties of the multiple L-functions.
Proof. (1), resp. (2), comes from Theorem IV.1.8, resp. Theorem IV.1.10 of [75].
This certainly explains why only L2 -automorphic forms are used in original defi-
nition of multiple L-functions.
(3) (Holomorphicity)
(a) When <(π) ∈ C, L≤p F,n (φ, π) is holomorphic.
(b) L≤p
F,n (φ, π) is holomorphic at π where <(π) = 0.
(4) (Singularities) Assume that φ is a cusp form. Then
(a) There is a locally finite set of root hyperplanes D such that the singularities
of L≤p
F,n (φ, π) are carried out by D.
(b) All singularities of L≤pF,n (φ, π) are without multiplicities at π provided that
h<(π), α i ≥ 0 ∀α ∈ ∆GM .
∨
(c) There are only finitely many of singular hyperplanes of L≤p F,n (φ, π) which
n o
intersect π ∈ P : h<(π), α i ≥ 0, ∀α ∈ ∆ M .
∨
We end this chapter with a comment that the above discussion works for gen-
eral reductive groups G as well, if we replace the semi-stable OK -lattices with the
semi-stable arithmetic principal G-torsors to be introduced in Part 6.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 185
Chapter 7
In this section, we recall the constructions and basic properties of Arthur’s analytic
truncation following [3, 4, 6].
+ G
Definition 7.1. The (acute) Weyl chamber aG+
P and the (obtuse) Weyl cone aP of
G
aP are defined by
P := H ∈ aP : hα, Hi > 0 ∀α ∈ ∆P ,
aG+
G
+ G
(7.2)
aP := H ∈ aGP : h$Gα , Hi > 0 ∀α ∈ ∆P .
185
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 186
Lemma 7.1.
+ G
(1) (see e.g. [44]) When P = P0 , the spaces aG+
0 and a0 coincides with the usual
positive Weyl chamber and the positive cone. In particular, + aG0 is the dual
cone of aG+
0 and
[
aG0 = 0 .
w aG+ (7.3)
w∈W
Similarly, as above, since all elements of b∆PQ are non-negative linear combina-
tions of the elements in ∆PQ , we obtain the following:
Lemma 7.2. Let P ⊆ Q be the standard parabolic subgroups of G. For the char-
acteristic function τPQ , resp. b
τPQ , of the positive chamber, resp. the positive cone,
we have
τPQ ≥ τPQ .
b (7.5)
Corollary 7.1. ( [4]) For fixed N ≥ 0 and T ∈ a0 , there exist constants c0 and N 0
such that, for any function φ on P(F)\G(A)1 , and x, y ∈ G(A)1 ,
!
X
N0 N0 |φ(u)|
φ(δx) ·b
τP H(δx) − H(y) − T ≤ c kxk · kyk · sup
0
N
. (7.7)
δ∈P(F)\G(F) u∈G(A)1 kuk
Lemma 7.4. (Lemma 6.1 in [3]) The function σ21 on a0 coincides with the char-
acteristic function of the following subset (of a0 )
α(H) > 0 ∀α ∈ ∆21
β(H) ∆ 2
.
H ∈ a ≤ 0 ∀β ∈ \∆ (7.9)
1 1 1
$(H) > 0 ∀$ ∈ ∆2
b
1 When δ runs over P(F)\G(F), the collection of all δPδ−1 ’s consists of all parabolic subgroups of G
conjugating to P.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 188
∈ a0 α
∀α ∈ ∆PQ .
H
$α (H) ≤ 0, if Λ(α ) > 0
∨
Here the $α ’s are the fundamental weights defined by h$α , βi = δαβ ∀α, β ∈ ∆0 ,
and, for a subset S ⊆ a0 , 1S denotes the characteristic function of S in a0 . As an
application of Langland’ combinatorial lemma, we have the following:
n o
sP1 (T 0 , ω) := ω · a · K : α H0 (a) − T 0 > 0 ∀α ∈ ∆0P .
(7.13)
Let F P (x, T ) be the characteristic function of the set of x ∈ G(A) such that δx
belongs to sP1 (T 0 , T, ω) for some δ ∈ P1 (F). Namely,
F P (x, T ) := 1n o (7.16)
x ∈ G(A) : δx ∈ sP1 (T 0 , T, ω) ∃δ ∈ P1 (F) .
Definition 7.3. ( [3,4]) Let T ∈ a+0 be sufficiently regular and let φ be a continuous
function on G(F)\G(A)1 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 190
Here and in the sequel, the measure on N(A) is chosen to be the one satisfying
Z
dn = 1. (7.20)
N(F)\N(A)
where the sum runs over all parabolic subgroups, standard or not.
As it stand, Arthur’s analytic truncation is indeed a well-designed device, in
which all the constant terms are tackled in such a way that the contributions from
all different levels of the parabolic subgroups are not only counted at the cor-
responding cuspidal region, but there is no overdone totally for these parabolic
subgroups, since each parabolic subgroup is essentially contributed once and only
once.
(3) Let Kfin be an open compact subgroup of G(Afin ), and let r, r0 be two positive
real numbers. Then there exists a finite subset Xi : i = 1, 2, . . . , N ⊂ U,
the universal enveloping algebra of g∞ , such that, for any right Kfin -invariant,
smooth function φ on G(F)\G(A),
N
X 0
n o
ΛT φ(ag) ≤ kgk−r sup |δ(Xi )φ(ag0 )| kg0 k−r : g0 ∈ G(A)1 , (7.25)
i=1
The key to prove of this result is that the nilpotent radicals of parabolic sub-
groups have relatively simple structures. Namely, they all admit filtrations of sub-
spaces with affine successive quotients. For details, please refer to Arthur’s origi-
nal papers mentioned above.
Obviously, ΛT,G = ΛT .
(1) comes from Langlands’ combinatorial lemma. For a detailed proof, please
refer [4].
Definition 7.5. Let T ∈ a+0 be sufficiently regular. The truncated domain Σ(T ) :=
ZG(A)G(F)\G(A) of the space G(F)\G(A)1 is defined by
T
All these can be used to prove the following result, by applying the Langland
combinatoric lemma and the associated inverse formula. For details, please refer
to [5].
Corollary 7.4. ([5]) For a sufficiently regular T ∈ a+0 , the characteristic function
F Q (g, T ) introduced in (7.16) coincides with the partial truncation along P for the
constant function 1. That is to say,
F P (g, T ) = ΛP,T 1(g) =
X P
X
(−1)|∆R | τRP H(δg) − T .
b
R:P0 ⊆R⊆P δ∈R(F)\P(F)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 193
Note that the right hand side makes sense even when T is not sufficiently
regular. This then leads to the definition of F P (g, T ) for general T . Namely,
Proposition 7.4. ( [3–5], see also [41]) The function F P (g, T ) on the space
N(A)M(F)\G(A)1 is bounded and compactly supported, uniformly for T varying
in a compact subset.
and
Q
X
(−1)|∆R | ΓRP (H, X)b
ΓRQ (H, X) = δPQ ,
R:P⊆R⊆Q
X R
(7.35)
ΓRP (H, X)ΓRQ (H, X) = δPQ .
(−1)|∆P |b
R:P⊆R⊆Q
Q
∆P R
ΓQ
Consequently, since, by definition, b P (H, X) = (−1) ΓP (H − X, −X),
Q
X
τQ
P (H − X) = (−1)|∆R | ΓRP (H − X, −X)τRQ (H),
R:P⊆R⊆Q
X X (7.36)
F (g, T + X) =
P
F R (δg, T )ΓRP (H(δg) − T, X).
R:P0 ⊆R⊆P δ∈R(F)\P(F)
Even, via the theory of the so-called Eisenstein systems, we may clarify some
preliminary structures of these Eisenstein periods ωG;T
F (φ, π), in general, it is very
difficulty to understand them completely, due to the complications of Eisenstein
series and their associated constant terms. However, when φ is a cusp form, the
situation changes dramatically as we will see below, essentially because the con-
stant terms of the associated Eisenstein series have very simple structures.
Let V be a real n dimensional vector space, denote its dual space by V ∗ . Then VC∗
is the space of complex linear forms on V. By a cone in V we shall mean a closed
subset of the form
n o
C := x ∈ V : hµi , xi ≥ 0 ∀i (7.43)
where µi }ni=1 is a basis of V ∗ . Let e j be the dual basis of V. We shall say that
λ ∈ VC∗ is negative, resp. non-degenerate, with respect to C if <(hλ, e j i) < 0, resp.
hλ, e j i , 0, for each j = 1, . . . , n.
Denote by S (V ∗ ) := Sym(V ∗ ) the symmetric algebra of V ∗ . In the sequel,
we view elements of this space as polynomial functions on V. By definition, an
exponential polynomial function on V is a function of the form
r
X
f (x) = ehλi ,xi Pi (x) (7.44)
i=1
where the λi are distinct elements of VC∗ and the Pi (x) are non-zero elements of
S (VC∗ ). We call λi ’s the exponents of f . Consider the integral
Z
IC f ; λ := f (x)ehλ,xi dx, (7.45)
C
and set
Hk,i := λ : hλ, ek i = hλi , ek i
n o
(1 ≤ i ≤ r). (7.46)
r
X
Lemma 7.9. ([47]) Let f (s) = ehλi ,xi Pi (x) be an exponential polynomial func-
i=1
tion on V, and let C be a cone in V.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 196
(1) The λi ’s are uniquely determined. And f is integrable over C if and only if λi
are negative with respect
to C for all i.
(2) The function IC f ; λ has analytic continuation to V ∗ with (only) hyperplane
singularities along Hi,k , 1 ≤ k, i ≤ r.
(3) Suppose that f is absolutely integrable over C. Then IC f ; λ is holomorphic
R
at 0 and IC f ; 0 = C f (x) dx.
(4) Function IC f ; λ is holomorphic at 0 if and only if for all i, λi are non-
degenerate with respect to C, i.e. hλi , ek i , 0 for all pairs (i, k).
Remark 7.1. The assumptions above are motivated by the micro structure for the
constant terms of Eisenstein series. On the other hand, by the reduction theory,
the quotient M(F)\M(A)1 is of finite volume. Hence, when taking integration, we
should deal with the A-part, which is affine, over which the exponential polyno-
mial function can be controlled easily if the exponents are negative.
where n ∈ N(A), a ∈ AP , m ∈ M(A)1 , k ∈ K, and that, for all j, the α j (X)’s are
polynomials, and the φ j (g)’s are automorphic forms in AP (G) such that φ j (ag) =
φ j (g) for a ∈ AP . Following [47], we define
Z ∗
f (g) · τP H(g) − T dg
P(F)\G(A)1
l Z
X Z ∗ Z # (7.66)
:= φ j (mk) dm dk × α j (X)ehλ j ,Xi τP (X − T ) dX .
j=1 K M(F)\M(A)1 aP
Accordingly, let A(G)∗∗ be the space of φ ∈ A(G) such that (a) and hence also (b)
are satisfied for all P. As a direct consequence of the constructions and results in
the previous subsection, we have the following:
All these, together with Example 7.2, implies the following result on the Eisen-
stein periods associated to cusp forms.
ehwλ−ρ,T i
Z
1
X
ωG;T ( M0 , λ) = vol Λ
b ∨
· M(w, λ) dm (7.69)
α∈∆0 hwλ − ρ, α i
F 0 Q ∨
w∈W T0 (F)\T0 (A)1
Here in the last two equalities, we have used the fact that, for the nilpotent rad-
ical N0 of M0 , the volume of N0 (F)\N0 (A) is one (see e.g. §6.1.5), and that
vol T0 (F)\T0 (A)1 = vol F ∗ \I1F n = b ζF (1)r where bζF (1)r := Res s=1b
ζF (s) (see
e.g. [66], or [113] where different normalization of the volume form is used). On
the other hand, by the well-known Gindikin-Karpelevich formula, we can write
down the associated intertwining operator M(w, λ) explicitly.
Lemma 7.12. (See e.g. [La3]) For the Eisenstein series EG/M0 (1, λ, g),
ζF hλ, α∨ i
Y b
M(w, λ) = . (7.71)
α>0,wα<0 ζF hλ, α i + 1
b ∨
By all these, particularly, Lemma 7.12 and the relation (7.69), we finally obtain
F (1 M0 , λ)
ωG;T ehwλ−ρ,T i ζF hλ, α∨ i
X Y b
= ζ
b F (1)n
. (7.72)
α∈∆0 hwλ − ρ, α i α>0,wα<0 b
∨
Q
vol Λ
b∨
0 w∈W ζ F hλ, α∨i + 1
Note that the right hand side of (7.72) makes sense for all T ∈ a0 , even the
right hand side is only defined for sufficiently positive T .
Definition 7.8. ( [126]) For a reductive group G over a number field F and T ∈ a0 ,
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 203
ehwλ−ρ,T i ζF (hλ, α∨ i)
X Y !
ωG;T .
b
F (λ) := · (7.73)
α∈∆0 hwλ − ρ, α i α>0,wα<0 b
∨
Q
w∈W ζF (hλ, α∨ i + 1)
ζF (hλ, α∨ i)
!
X 1 Y
ωGF (λ)
b
:= · (7.74)
α∈∆0 hwλ − ρ, α i α>0,wα<0 b
∨
Q
w∈W ζF (hλ, α∨ i + 1)
Proposition 7.6. ([57]) For a sufficiently positive T ∈ a+0 , the volume of the trun-
cated domain Σ(T ) ⊂ G(F)\G(A)1 is given by
vol Σ(T )
Y
ζF (1)nw i>1 b
ζF (i)ni,w
X Q
= ζ · ehwρ−ρ,T i . (7.78)
−n
b
b F (i) i
vol Λ0 α ∨i − 1
Q
b ∨
i>1
w∈Wspa α∈∆0 −wBw hwρ,
Our strategy is to take the residues for both side of this equation at λ = ρ, because,
by the result of Langlands ( [68]), the residue of E(g, 1, λ) at λ = ρ is constant.
Indeed, since Σ(T ) is compact for every sufficiently regular T ,
Z !
Resλ=ρ ∧ E(g, 1, λ) dg
T
Σ(T )
Z (7.80)
= Resλ=ρ E(g, 1, λ) dg = vol Σ(T ) · Resλ=ρ E(1 M0 ; λ; g).
Σ(T )
Hence it suffices to calculate the residues Resλ=ρ E(1 M0 ; λ; g) and Resλ=ρ ωG;T
F (λ).
We first treat Resλ=ρ E(1 M0 ; λ; g).
It is well-known that Resλ=ρ E(1 M0 ; λ; g) is obtained from the residue of con-
stant term along the Borel subgroup P0 , and, by Proposition 6.1, or better, (6.73),
the constant term of E(1, g, λ) along P0 of G is given by
X
E B (g, 1, λ) = M(w, λ)ehwλ+ρ,H(g)i . (7.81)
w∈W
Hence, from hρ, α∨ i = 1 for α ∈ ∆0 and hρ, α∨ i > 1 if α < ∆0 , α > 0,
Y
Resλ=ρ E(g, 1, λ) = lim hλ, α i − 1 · M(w0 , λ)
∨
(7.82)
λ→ρ
α∈∆0
where w0 is the longest Weyl group element, which transforms all the positive
ζF hλ, α∨ i
Y b
roots to negative. Now, by Lemma 7.12, M(w0 , λ) = , we have
α>0 ζF hλ, α i + 1
b ∨
Y
Resλ=ρ E(g, 1, λ) = b
ζF (1)n · ζF (i)ni .
b (7.83)
i>1
Therefore, by (7.80),
Z !
E(1 M0 ; λ; g) dg = b
Y
ζF (1)n ζF (i)ni · vol Σ(T ) .
Resλ=ρ b (7.84)
Σ(T ) i>1
Next we treat the residue of the right hand side of (7.79) at λ = ρ. Since
hwλ, α∨ i = hλ, (w−1 α)∨ i, with nw = # α ∈ ∆0 : wα < 0 ,
X Y hλ, α∨ i − 1
Resλ=ρ ωG;T
F (λ) = λ→ρ
lim ehwλ−ρ,T i M(w, λ)
w∈W α∈∆0
hwλ, α ∨i − 1
(7.85)
ζF (1)nw i>1 b ζF (i)ni,w
X Q
= .
hwρ−ρ,T i
b
e
α∈∆0 ,w−1 α<∆0 hwρ, α i − 1
∨
Q
w∈W spa
as desired.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 205
For many purposes, the formula (7.78) is sufficient. However, to expose the
hidden structures related to standard parabolic subgroups, we prove the following:
Lemma 7.13.
(1) There is a bijection between Wspa and the collections of standard parabolic
subgroups of G.
(2) If w ∈ Wspa , hwρ, α∨ i < 0 for all α ∈ ∆ − wBw .
Proof. (1) For each w ∈ Wspa , we obtain a subset Bw ⊆ ∆ and hence also an
associated standard parabolic subgroup Pw . Conversely, any standard parabolic
subgroup P of G is associated to a unique subset J ⊆ ∆. Set W J ⊆ W be the
subgroup of W generated by the reflections of α ∈ J, and introduce a subset
W(J) := w ∈ W/W J : w(J) ⊆ Φ+ ⊆ W/W J . It is known that the set W(J) gives
all the distinguished coset representatives of W/W J . Let w J be the element of the
maximal length in W(J). Then w J ∈ Wspa and in fact Wspa = w J : J ⊆ ∆ .
(2) To start with, we let w0 be the element of the maximal length in W and
let w0,J be the element of the maximal length in W J . By [20], for w J in (1), we
have w J = wl wl,J . Assume that α ∈ ∆ − wBw . Obviously, there exists a certain
β ∈ ∆ such that w0 (α) = −β. On the other hand, by definition, w0,Bw w0 = w−1 .
Consequently, w−1 (α) = −w0,Bw (β). This implies that β < Bw . Indeed, otherwise,
−w0,Bw (β) ∈ Bw and w−1 (α) ∈ Bw . Since α < wBw , we arrive at w−1 (α) < Bw , a
contradiction. But with β < Bw , we have wl,Bw (β) > 0, and hence w−1 (α) < 0.
To finally write the summation above as one over standard parabolic subgroups,
we need to give a group theoretic interpretations for other terms. For this, we
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 206
Corollary 7.5. ([68], see also [57]) Let G be a split, semi-stable group on F. For
the volume of (a fundamental domain of) G(F)\G(A)1 , we have
Y
vol G(F)\G(A)1 =vol Λ b∨
0 ζF (i)−ni .
b
(7.87)
i>1
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 207
PART 4
For reductive groups over number fields, new type of zeta functions are introduced
as the residues of the associated periods. They were initially discovered for SL3
as a by-product of the study of the rank three non-abelian zeta function. With a
result of Diehl on the residues of Siegel Eisenstein series associated to Sp2n , we
are able to locate the singular hyperplanes of the SLn and Sp2n -periods and fur-
ther discover the single variable zeta functions associated. To understand what
happens for other groups, G2 naturally becomes a target, being exceptional and
of rank two. The result is fascinating. Not only two different zeta functions are
discovered for G2 , the role played by the maximal parabolic subgroups is also
exposed. This finally leads to the zeta functions for reductive groups and their
maximal reductive groups, which are conjecturally to satisfy the functional equa-
tion and the Riemann Hypothesis. In this part, we first give a detailed account of
the above process, then present Suzuki’s proof for the Riemann Hypothesis of the
G2 -zeta functions based on our explicit formulas, and more importantly, Komori’s
beautiful Lie theoretic proof of the functional equations for all these new zeta
functions based on an involution of the Weyl elements involved. For the case of
(SLn , Pn−1,1 ), the functional equation was first jointly proved by Kim and myself.
Even our original proof is not produced, we instead use a joint work with Zagier
to write down the (SLn , Pn−1,1 )-zeta functions explicitly.
207
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 209
Chapter 8
In the following two sections, we, following [126], explain how a study of high
rank non-abelian zeta functions leads to a new type of zeta functions associated to
reductive groups and their maximal parabolic groups. For simplicity, we assume
that the base field involved is the field of rationals.
For a fixed n ∈ Z>0 , denote by b ζQ,n (s) the rank n non-abelian zeta function of
Q. By Proposition 3.2, b ζQ,n (s) can be calculated as an integration of Eisenstein
series over a moduli space of semi-stable lattices. Namely, (up to a constant factor
depending only on n),
Z
b Λ(g), n s dg.
ζQ,n (s) =
b E (8.1)
MQ,n [1] 2
Here MQ,n [1] ⊂ SLn (Z)\SLn (R)/SOn denotes the moduli spaces of semi-stable
(Z-)lattices of rank n and volume one, and
X 1
s) := π−s Γ(s) E(Λ(g), s) := π−s Γ(s)
E(Λ(g),
b
2s
(8.2)
x∈Λ(g)r{0}
kxk
denotes the complete Epstein zeta function for the lattice Λg = (Zn , ρ(g)), where
ρ(g) denotes the metric on Zn ⊗R = Rn induced by to the positive definitive matrix
gt · g for some g ∈ SLn (R).
Identify MQ,n [1] with the semi-stable part of a fundamental DSLn (Z) of
SLn (Z)\SLn (R)/SOn . Denote by 1MQ,n [1] the characteristic function of MQ,n [1].
By (8.1), we have
Z n
ζQ,n (s) =
b 1MQ,n [1] (g) · Eb Λ(g), s dg. (8.3)
SLn (Z)\SLn (R)/SOn 2
209
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 210
Up to this point, it is natural to ask whether 1MQ ,n [1] is related with Arthur’s
analytic truncation. This is indeed the case.
Theorem 8.1 (Theorem 14.2). Denote by Λ0SLn , or simply Λ0 , the Arthur analytic
truncation associated to SLn . Let 1 be the constant function of the value one on
the space SLn (Z)\SLn (R)/SOn . Then
1MQ,n [1] = Λ0SLn 1. (8.4)
In addition, Λ0SLn is unipotent and self-adjoint.
Recall that to make the original Arthur analytic truncation ΛT work, the pa-
rameter T has to be sufficiently regular T , a condition claiming that (T, α) 0
for all simple roots α. Our result significantly generalize this truncation theory of
Arthur. Indeed, as we will see in Theorem 14.2 below, it suffices for us to assume
that the positivity of the (T, α)’s.
We will prove this theorem in Part 6, where a much more general result will
be given. Instead, based on the basic properties of Arthur’s analytic truncation in
§7.1.6,1 we give the following direct consequence of Theorem 7.2.
Corollary 8.1 (Theorem 15.1). All non-abelian zeta functions are Eisenstein pe-
riods. More precisely, up to a constant factor depending only on n,
Z
b Λ(g), n s dg.
ζQ,n (s) =
b Λ0 E (8.5)
SLn (Z)\SLn (R)/SOn 2
This clearly indicates that to study non-abelian zeta functions, we should apply
the theory of Arthur periods, or better, Eisenstein periods. For this, we next expose
some explicit relation between the Epstein zeta function and Eisenstein series.
Let SPn be the space of positive definitive matrices of determinant one and size
n, and let Γn := diag(±1, . . . , ±1) ∩ SLn (Z) SLn (Z). Denote by Qn1 ,...,nk the
subgroup of Γn induced from the standard parabolic subgroup Pn1 ,...,nk of SLn (R)
associated to the partition n = n1 + · · · + nk , that is, the subgroup of Γn consisting
H1 ∗ · · · ∗
0 H2 · · · ∗
of matrices in SLn (Z) of the form . . . . for suitable H j ∈ SLn j (Z) (1 ≤
.. .. . . ..
0 0 · · · Hk
Definition 8.1. Let Y ∈ SPn be a positive definitive matrix of size n and let
n = n1 +2 + . . . + nk be a fixed partition of n.
(1) The Siegel zeta functions Zn∗1 (Y; sn1 , . . . , sn−1 ) associated to Y and Qn1 ,1,...,1 is
defined by
n−1
−s j
X Y
Zn∗1 (Y; sn1 , . . . , sn−1 ) := Y[N] j ∀1 ≤ n1 ≤ n − 1. (8.6)
N∈Qn1 ,1,...,1 \Γn j=n1
(2) The Siegel Eisenstein series En1 ,n2 ,...,nk s; Y associated to Y and Qn1 ,...,nk is
defined by
X k
Y −s j n j +n j+1
En1 ,...,nk Y; s1 , . . . , sk := <(s j ) > . (8.7)
Y[A j ]
(A j ∗)=A∈Qn1 ,...,nk \GLn j=1
2
A j ∈Zn×N j
(3) For every m ≤ n, the Koecher zeta function associated to Y and Qm,n−m is
defined by
X −s n
Zm,n−m (Y, s) := Y[A] <(s) > . (8.8)
2
n×m
A∈Z /GLm (Z)
rankA=m
Lemma 8.1. (See e.g. [115]) With the same notation as above,
(1) Zn∗1 (Y; sn1 , . . . , sn−1 ), En1 ,n2 ,...,nk (Y; s1 , . . . , sk ) and Zm,n−m (X, s) are well-
defined, and admit meromorphic continuations to the whole parameter
spaces.
(2) ([22]2 ) There exists a constant c, depending only on n1 , such that
n1
Res sn = n1+1 Zn∗1 (Y; sn1 , . . . , sn−1 ) = cn1 Zn∗1 +1 Y; sn1 +1+ , sn1 +2 , . . . , sn−1 . (8.9)
1 2 2
In particular, by taking r = 1 and repeating this process, we obtain, up to a
constant factor,
n−1
Res sn−1 =1 · · · Res s2 =1 Res s1 =1 Z1∗ (Y; s1 , s2 , . . . , sn−1 ) = |Y|− 2 . (8.10)
Yn−1
(3) Zn,0 (Y, s) = |Y|−s · ζ(2s − j), and
j=0
Z1,n−1 (Y; s)
|Y|−s · En−1,1 (Y −1 ; s) = E1,n−1 (Y; s) = . (8.11)
Z1,0 (I; s)
2 Please correct a misprint in [22] for (8.9).
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 212
Proposition 8.1. Let g ∈ SLn (R) such that Y = gt g ∈ SPn . Denote by Λ(g) the
Z-lattices (Zn , ρg ). We have
1 X 1 s
∗
Zn−1 (Y −1 ; s) = · |Y[x]|−s = · E Λ(g); , (8.16)
ζ(2s) x∈Zn \{0} ζ(2s) n/2
Z1∗ (Y; s1 , s2 , . . . , sn−1 ) = E(n) Y; s1 , s2 , . . . , sn .
(8.17)
In particular,
n−2
E Λ(g); s = Restn−2 =1,...,t2 =1, t1 =1 Z1∗ Y; ns − , tn−2 , tn−3 , . . . , t2 , t1 . (8.18)
2
That is to say, Epstein zeta functions are residues of Siegel Eisenstein series.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 213
Here, in the last step, we have used the fact that nj=1 a j = |Y| = 1.
Q
On the other hand, since Y is positive definite, there exists a matrix g such that
Y = gt g. Indeed, we may take g = T (g)N where T (g) = diag(t1 , t2 , . . . , tn ) with
a j = t2j . Consequently, by definition,
mB (g)λ+ρB = T (g)λ+ρB . (8.19)
Hence, if we write λ = (z1 , z2 , . . . , zn ) ∈ C , then j=1 z j = 0. In these coordinates,
n Pn
by §6.1.3, the Weyl vector ρ becomes
n − 1 n − 1 n − 1 n − 1
ρ = ρB = , − 1, . . . , 1 − ,− . (8.20)
2 2 2 2
Thus, by a direct calculation, we obtain
mB (g)λ+ρB = t1−[(n−1)+(2z1 +z2 +···+zn−1 )] t2−[(n−2)+(z1 +2z2 +···+zn−1 )] · · · tn−1
−[1+(z1 +z2 +···+2zn−1 )]
. (8.21)
Recall now that the relative Langlands Eisenstein series on SLn induced from
the constant function 1 on the Levi part of theX
Borel B = P1,1,...,1 is defined by
E(1; λ; g) := E SLn /B (1; λ; g) := mB (δg)λ+ρB . (8.22)
γ∈S Ln (Z)/P1,1,...,1 =B
Hence, if we make the coordinate changes from λ = (zi ) to s = (si ) by the linear
transformations
2s1 = 1 + (z1 − z2 )
= 1 + (z2 − z3 )
2s2
(8.23)
··· ···
= 1 + (z − z ),
2s
n−1 n−1 n
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 214
what we have said then proves the following elementary relations between the
Epstein zeta functions and the Eisenstein series.
By Corollary 8.1 and Proposition 8.2, to obtain an explicit expression for the rank
ζQ,n (s), it suffices to evaluate the integration
n zeta function b
Z
Resz2 −z3 =1 · · · Resz3 −z4 =1 · · · Reszn−1 −zn =1 E(1; z1 , z2 , . . . , zn ; g) dµ(g).
MQ,n [1]
Moreover, by Theorem 8.2, f the moduli spaces MQ,n [1] (of semi-stable lattices
of rank n and volume one) is compact, we can freely interchange the orders of
R
(a) the operation of taking the integration on M [1] , and
Q,n
(b) the operation of taking the residues Resz2 −z3 =1 · · · Resz3 −z4 =1 · · · Reszn−1 −zn =1 .
Here in the last step, we have used the fact that Λ0 is unipotent and self-adjoint.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 215
Up to this point, it is only natural to recall that, for T ∈ a+0 sufficiently regular,
we have, by Corollary 8.2 and Theorem 7.5,
Z
E(1; z1 , z2 , . . . , zn ; g) dµ(g)
Σ(T )
Z
= ΛT E(1; z1 , z2 , . . . , zn ; g) dµ(g) (8.25)
SLn (Z)\SLn (R)/SOn
ehwλ−ρ,T i ζ hλ, α∨ i
X Y b
= .
α∈∆ hwλ − ρ, α iα>0,wα<0 b
∨
Q
w∈W ζ hλ, α∨ i + 1
Here Σ(T ) denotes the compact subset in (a fundamental domain of the space)
SLn (Z)\SLn (R)/SOn with characteristic function ΛT 1. Therefore, by Proposi-
tion 7.4 and Theorem 8.2, (8.25) holed for all T ∈ a+0 t {0}. All these then prove
the following:
Theorem 8.2. For all T ∈ a+0 t {0} and λ = (z1 , . . . , zn ) satisfying ni=1 zi = 0,
P
ehwλ−ρ,T i ζ hλ, α∨ i
Z
E(1; λ; g) dµ(g) =
X Y b
. (8.26)
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
Σ(T ) w∈W ζ hλ, α∨ i + 1
ehwλ−ρ,T i ζ hλ, α∨ i
X Y b
ωQ,T
SLn
(z1 , z2 , . . . , zn ) = .
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
w∈W ζ hλ, α∨ i + 1
SLn
(2) The T -version zeta function ZQ,T (z1 ) of SLn over Q is defined by
SL
SLn
ZQ,T (z1 ) := Resz2 −z3 =1 Resz3 −z4 =1 . . . Reszn−1 −zn =1 ωQ,T
n
(λ)
= Resz2 −z3 =1 Resz3 −z4 =1 . . . Reszn−1 −zn =1
(8.27)
ehwλ−ρ,T i ζ hλ, α∨ i
X Y b !
.
α∈∆ hwλ − ρ, α iα>0,wα<0 b
∨
Q
w∈W ζ hλ, α∨ i + 1
In particular, ωQ,0
SLn SLn
(λ) and ZQ,0 (z1 ) are called the period and the zeta function,
of SLn over Q respectively, and denote by ωQSLn (z1 ) and ZQSLn (z1 ), respectively.
Normally, after taking residues in (8.27), there are still zeta factors left in the
denominators for terms corresponding to w ∈ W, despite of the cancelations of
the same zeta factors in both denominator and numerators (with possible uses of
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 216
the functional equation of Riemann zeta function). We want to eliminate all such
reminding zeta factors (in the denominators). For this, we can form the ‘least
common multiple’ of them, based on the following observations.
(i) There are finitely many of non-trivial zeta functions, say, I(SLn ) of them,
SL SL
ζ a1SLn z1 + b1SLn , b
b ζ a2 n z1 + b2SLn , . . . , b
ζ aI(SLn
n)
z1 + bI(SL
SLn
n)
,
QI(SL ) SL
satisfying the condition that the product i=1
n
ζ ai n z1 + biSLn ·ZQSLn (z1 )
b
admits only finitely many singularities.
(ii) There are finitely many factors of special zeta values, say J(SLn ) of them,
SL SL
bζ c1SLn , b
ζ c2 n , . . . , bζ c J(SL
n
n)
,
These new zeta functions are extremely beautiful, since they satisfy the func-
tional equation and a weak Riemann Hypothesis.
Theorem 8.3 (Functional Equation). There exists a constant c SLn such that
ζQ;o
SLn
c SLn − s = b
ζQ;o s.
SLn
b (8.29)
This result in the case n = 2, 3, 4, 5 was first verified by the author in [126],
and was jointly proved by Kim and myself ([58]). In the next chapter, we will
ζQ;o
explicitly write done bSLn
s and consequently obtain a new proof.
ζQSLn s of SLn over Q is defined by
Definition 8.5. The zeta function b
c SLn − 1
ζQSLn (s) := b
b ζQ;o
SLn
s+ . (8.30)
2
Accordingly, the functional equation becomes
ζQSLn (1 − s) = b
b ζQSLn (s). (8.31)
ζQSLn(s)
Theorem 8.4 (Weak Riemann Hypothesis). All but finitely many zeros of b
1
lie on the central line <(s) = .
2
This result will be proved in Part 6. In the case of n = 2, resp. 3, resp. 4, 5, the
Riemann Hypothesis for b ζQSLn (s) is proved by [65], resp. [109], resp. [55], earlier.
8.2.1 Sp2n-Periods
Motivated by the discussion in the previous subsection, we next use the periods
to introduce zeta functions for other types of reductive groups. We start with the
symplectic groups. Let G = Sp2n be the symplectic group of size n. Denote √ by Sn
to be the Siegel upper half space of rank n. For any Z ∈ Sn , write Z = X + −1Y
with X its real part and Y its imaginary parts. By definition, Y ∈ Sn if and only if
Y = Im Z > 0 and Z t = Z is symmetric. There is a natural action of Sp2n on Sn
defined as follows:
!
A B
(M, Z) 7→ MhZi := (AZ + B) · (CZ + D)−1 ∀M = ∈ Sp2n (R), ∀Z ∈ Sn .
CD
Set also Y(M) := ImMhZi. It is well known that the √ action of Spn (R) on Sn is
transitive and the stabilizer group of the element −1 I in Sp2n (R) is equal to
Sp2n (R) ∩ SO(2n). Consequently, there is a natural isomorphism
Sp2n (R) SO(2n) ∩ Sp2n (R) ' Sn .
(8.32)
By an abuse of notations, set Γn := diag(±1, ±1, · · · , ±1) ∩Sp2n (Z) Sp2n (Z)
( ! )
∗∗
be the Siegel modular group, and let P = Pn := ∈ Γn be the associated
0∗
maximal parabolic subgroup. Moreover, when Z ∈ Sn , we define the Siegel-Maaβ
Eisenstein series, or more correctly, the Siegel-Epstein zeta function by
X |Y|−s
En (Z; s) := . (8.33)
γ∈P\Γ
kCZ + Dk−2s
n
Motivated by our discussions on the periods and zeta functions associated to SLn ,
we introduce the following:
Definition 8.6. For a T in the positive cone of the root space associated to Sp2n ,
the T -version period of a single variable for Sp2n over Q is defined by
Z
Sp
zQ,T2n (s) := ΛT En (Z; s) dµ(Z). (8.34)
Γn \Sn
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 218
Sp Sp
In particular, we call zQ,02n (s) := ζQ,T2n (s) T =0 the principal zeta function for Sp2n
Sp
over Q and denote it simply by zQ 2n (s).
Similar to the case of SLn , to write down these integral periods explicitly, we
make some preparations.
These Siegel Eisenstein series are closed related to the Siegel zeta functions asso-
ciated to the standard parabolic subgroup Qn1 ,1,...,1 of SLn .
To use our formulas for the classical Siegel Eisenstein periods above, similar to
the case for SLn , we next write them in terms of Langlands’ language.
√ λ = (z1 , z2 , . . . , zn ) be an element in the root space a0 of Sp2n . For Z =
Let
X + −1 Y ∈ Sn , set av = |Yv |/|Yv−1 | and a := (av ). Then
Yn
aλ (Z) = v = |Y1 |
a−z v −z1 +z2
|Y2 |−z2 +23 · · · |Yn−1 |−zn−1 +zn |Yn |−zn (8.39)
v=1
for s1 = z1 − z2 , s2 = z2 − z3 , . . . , sn−1 = zn−1 − zn , sn = zn . Accordingly, define
the associated power function by
n
∀s = (s1 , s2 , . . . , sn ).
Y
p−s (Y) := |Yµ |−sµ (8.40)
µ=1
Then, by (8.39) and (8.38), we obtain (1) and (2) below respectively.
By Theorem 7.5 in Part 3, and the Gindikin-Karpelevich formula for the associated
intertwining operator, we arrive at
ehwλ−ρ,T i ζ hλ, α∨ i
Z
ΛT E(1; λ; M) dµ(M) =
X Y b
· .
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
Sp(n,Z)\Sn w∈W ζ hλ, α∨ i + 1
Sp
This is nothing but the period ωQ,T2n (λ) of Sp2n introduced in Definition 7.7. Similar
to the group of SLn , by Proposition 8.4, we are led to the following:
Sp
Definition 8.7. The T -version zeta function ZQ,T2n (z1 ) of Sp2n over Q is defined by
Sp
ZQ,T2n (z1 ) := Resz1 −z2 =1 Resz2 −z3 =1 · · · Reszn−1 −zn =1
ehwλ−ρ,T i ζ hλ, α∨ i (8.42)
X Y b !
· .
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
w∈W ζ hλ, α∨ i + 1
Here λ = (z1 , z2 , . . . , zn ).
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 220
Sp
By Proposition 8.4, we have up to an affine change of variables, ZQ,T2n (z1 ) co-
Sp
incides with the period of single variable zQ,T2n (z1 ) of Sp2n in Definition 8.6.
Similar to the discussion after Definition
n Sp 8.3, for Sp2n , we obtain finitely many
Sp
o
ζ ai 2n z1 + bi 2n : ai , 0, i = 1, . . . , I(Sp2n ) and finitely
factors of zeta functions b
n Sp o
many factors of special zeta values b ζ c 2n : j = 1, . . . J(Sp2n ) .
j
Sp
Definition 8.8. The zeta function bζQ;o2n of Sp2n over Q is defined by
Y2n ) Sp
I(Sp Y2n ) Sp Sp
J(Sp
Sp2n Sp2n
ζQ,T ;o s :=
b ζ ai s + bi
b 2n
· ζ c j · ZQ 2n s ,
b 2n
<(s) 0.
i=1 j=1
Initial examples and results strongly suggest these zeta functions satisfy the
functional equation and the Riemann Hypothesis. In particular, we have
Theorem 8.5. ([110, 126]) Both zeta functions associated to Sp4 over Q satisfy
the functional equation and the Riemann Hypothesis.
To introduce new genuine zeta functions for a general reductive group G, we need
to take residues along some singular hyperplanes for the period
ζ hλ, α∨ i
X 1 Y b
ωQ (λ) :=
G
· .
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
w∈W ζ hλ, α∨ i + 1
In the case when G = SLn or SP2n , this is done with the help of the so-called
Epstein zeta functions, a special type of single variable Eisenstein series. For a
general G, this proves to be too much to be expected. Above all, we should guar-
antee that the final singular variable zeta functions for reductive groups obtained
should satisfy the functional equation and the Riemann Hypothesis numerically.
Obviously, among all classical groups, we should start with the groups with
very lower ranks and smaller Weyl groups. By examining groups of types Bn , Dn ,
E6,7,8 , F4 and G2 , it is then only natural for us to focus on the exceptional and
interesting G2 , which is of rank two and admits only 12 Weyl elements.
Accordingly, in the sequel till the end of this subsection, let G be the ex-
ceptional group G2 . Fix a maximal split torus T in G and a Borel subgroup B
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 221
containing T . There are only two simple roots in ∆, namely, the short root αshort
and the long root αlong . For simplicity, denote them by α and β respectively. Then
Φ+ = α, β, α + β, 2α + β, 3α + β, 3α + 2β .
n o
(8.43)
Let X(T ) be the character group of T and a∗C = X(T )⊗C. Introduce coordinates
in a∗C with the basis 2α + β, α + β. Then the point (z1 , z2 ) ∈ C2 corresponds to
the character λ = z1 (2α + β) + z2 (α + β), so that λ(t(a, b)) = |a|z1 |b|z2 . As such,
ρ0 = 5α + 3β and C of the positive Weyl chamber in a∗C is characterized by
n o n o
C := λ ∈ a∗C : < hλ, γ∨ i > 0∀γ > 0 = z1 (2α + β) + z2 (α + β) : <z1 > <z2 > 0 .
By a direct calculation, we have the following table on wλ and {γ > 0 | wγ < 0}.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 222
we have
hλ, α∨ i = z1 − z2 ,
hλ, β∨ i = z2 ,
hλ, (3α + β)∨ i = z1 ,
(8.48)
hλ, (2α + β)∨ i = 2z1 + z2 ,
hλ, (3α + 2β)∨ i = z1 + z2 ,
hλ, (α + β)∨ i = z1 + 2z2 .
w hwλ, α∨ i − 1 hwλ, β∨ i − 1
e z1 − z2 − 1 z2 − 1
wα z2 − z1 − 1 z1 − 1
wβ z1 + 2z2 − 1 −z2 − 1
w3α+β −2z1 − z2 − 1 z1 + z2 − 1
w2α+β −z1 − 2z2 − 1 z2 − 1
w3α+2β z1 − z2 − 1 −z1 − 1
wα+β 2z1 + z2 − 1 −z1 − z2 − 1
σ( π3 ) −z1 − 2z2 − 1 z1 + z2 − 1
σ( 2π3 ) −2z1 − z2 − 1 z1 − 1
σ(π) −z1 + z2 − 1 −z2 − 1
σ( 4π3 ) z1 + 2z2 − 1 −z1 − z2 − 1
σ( 3 )
5π
2z1 + z2 − 1 −z1 − 1
Q ζ(hλ,γ∨ i)
b
w γ>0,wγ<0 b
ζ(hλ,γ∨ i+1)
e 1
ζ(z1 −z2 )
b
wα ζ(z1 −z2 +1)
b
bζ(z2 )
wβ bζ(z2 +1)
bζ(z1 −z2 ) b ζ(z1 ) b ζ(2z1 +z2 )
w3α+β bζ(z1 −z2 +1) b ζ(z1 +1) b ζ(2z1 +z2 +1)
ζ(z1 −z2 ) b
b ζ(z1 ) b ζ(2z1 +z2 ) b ζ(z1 +z2 ) b ζ(z1 +2z2 )
w2α+β ζ(z1 −z2 +1) b
b ζ(z1 +1) bζ(2z1 +z2 +1) b ζ(z1 +z2 +1) b ζ(z1 +2z2 +1)
ζ(z1 ) b
b ζ(2z1 +z2 ) b ζ(z1 +z2 ) b ζ(z1 +2z2 ) b ζ(z2 )
w3α+2β ζ(z1 +1) b
b ζ(2z1 +z2 +1) b ζ(z1 +z2 +1) b ζ(z1 +2z2 +1) b ζ(z2 +1)
bζ(z1 +z2 ) b ζ(z1 +2z2 ) b ζ(z2 )
wα+β bζ(z1 +z2 +1) b ζ(z1 +2z2 +1) b ζ(z2 +1)
ζ(z1 +2z2 ) b ζ(z2 )
σ( π3 )
b
ζ(z1 +2z2 +1) b
b ζ(z2 +1)
ζ(2z1 +z2 ) b ζ(z1 +z2 ) b ζ(z1 +2z2 ) b ζ(z2 )
σ( 2π
3 )
b
ζ(2z1 +z2 +1) b
b ζ(z1 +z2 +1) b ζ(z1 +2z2 +1) b ζ(z2 +1)
ζ(z1 −z2 ) b ζ(z1 ) b ζ(2z1 +z2 ) b ζ(z1 +z2 ) b ζ(z1 +2z2 ) b ζ(z2 )
σ(π)
b
ζ(z1 −z2 +1) b
b ζ(z1 +1) b ζ(2z1 +z2 +1) b ζ(z1 +z2 +1) b ζ(z1 +2z2 +1) b ζ(z2 +1)
ζ(z1 −z2 ) b ζ(z1 ) b ζ(2z1 +z2 ) b ζ(z1 +z2 )
σ( 4π
3 )
b
bζ(z1 −z2 +1) b ζ(z1 +1) b ζ(2z1 +z2 +1) b ζ(z1 +z2 +1)
ζ(z1 −z2 ) b ζ(z1 )
σ( 5π
3 )
b
ζ(z1 −z2 +1) b
b ζ(z1 +1)
Q ζ(hλ,γ∨ i)
Table 8.3: Factors
b
γ>0,wγ<0 b
ζ(hλ,γ∨ i+1)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 224
1 1 Q ζ(hλ,γ∨ i)
·
b
w hwλ,α∨ i−1 hwλ,β∨ i−1 γ>0,wγ<0 b ζ(hλ,γ∨ i+1)
1 1
e z1 −z2 −1 z2 −1
1 1 ζ(z1 −z2 )
z2 −z1 −1 z1 −1 · b
b
wα ζ(z1 −z2 +1)
1 1 ζ(z2 )
z1 +2z2 −1 −z2 −1 · b
b
wβ ζ(z2 +1)
1 1 ζ(z1 −z2 ) b ζ(z1 ) b ζ(2z1 +z2 )
−2z1 −z2 −1 z1 +z2 −1 · b
b
w3α+β ζ(z1 −z2 +1) b ζ(z1 +1) bζ(2z1 +z2 +1)
1 1 ζ(z1 −z2 ) b ζ(z1 ) b ζ(2z1 +z2 ) b ζ(z1 +z2 ) b ζ(z1 +2z2 )
−z1 −2z2 −1 z2 −1 · b
b
w2α+β ζ(z1 −z2 +1) b ζ(z1 +1) bζ(2z1 +z2 +1) b ζ(z1 +z2 +1) b ζ(z1 +2z2 +1)
1 1 ζ(z1 ) b ζ(2z1 +z2 ) b ζ(z1 +z2 ) b ζ(z1 +2z2 ) b ζ(z2 )
z1 −z2 −1 −z1 −1 · b
b
w3α+2β ζ(z1 +1) b ζ(2z1 +z2 +1) b ζ(z1 +z2 +1) b ζ(z1 +2z2 +1) b ζ(z2 +1)
1 1 ζ(z1 +z2 ) b ζ(z1 +2z2 ) b ζ(z2 )
2z1 +z2 −1 −z1 −z2 −1 · b
b
wα+β ζ(z1 +z2 +1) b ζ(z1 +2z2 +1) b ζ(z2 +1)
ζ(z1 +2z2 ) b ζ(z2 )
σ( π3 ) 1 1
−z1 −2z2 −1 z1 +z2 −1 · b
b
ζ(z1 +2z2 +1) b ζ(z2 +1)
ζ(2z1 +z2 ) b ζ(z1 +z2 ) b ζ(z1 +2z2 ) b ζ(z2 )
σ( 2π
3 )
1 1
−2z1 −z2 −1 z1 −1 · b
b
ζ(2z1 +z2 +1) b ζ(z1 +z2 +1) b ζ(z1 +2z2 +1) b ζ(z2 +1)
ζ(z1 −z2 ) b ζ(z1 ) b ζ(2z1 +z2 ) b ζ(z1 +z2 ) b ζ(z1 +2z2 ) b ζ(z2 )
σ(π) 1 1
−z1 +z2 −1 −z2 −1 · b
b
ζ(z1 −z2 +1) b ζ(z1 +1) bζ(2z1 +z2 +1) b ζ(z1 +z2 +1) b ζ(z1 +2z2 +1) b ζ(z2 +1)
ζ(z1 −z2 ) b ζ(z1 ) b ζ(2z1 +z2 ) b ζ(z1 +z2 )
σ( 4π
3 )
1 1
z1 +2z2 −1 −z1 −z2 −1 · b
b
ζ(z1 −z2 +1) b ζ(z1 +1) bζ(2z1 +z2 +1) b ζ(z1 +z2 +1)
ζ(z1 −z2 ) b ζ(z1 )
σ( 5π
3 )
1 1
2z1 +z2 −1 −z1 −1 · b
b
ζ(z1 −z2 +1) b ζ(z1 +1)
1 1 Q ζ(hλ,γ∨ i)
Table 8.4: Terms ·
b
hwλ,α∨ i−1 hwλ,β∨ i−1 γ>0,wγ<0 b
ζ(hλ,γ∨ i+1)
By taking the summations of all terms in Table 8.4, we obtain the period
ωGQ2 (z1 , z2 ) for G2 over Q. However, it still looks very difficult to determine along
which line we should take the residue, since there are too many to determine.
Back to the general structure, recall that for G2 , the associated period is a two
variable function given by
ζ hλ, α∨ i
X 1 Y b
ωGQ2 (z1 , z2 ) = · , (8.49)
w∈W
hwλ − ρ, α∨short i hwλ − ρ, α∨long i α>0,wα<0 bζ hλ, α∨ i + 1
where (z1 , z2 ) = λ ∈ a0 , and ∆ = {αshort , αlong } consisting only two roots, i.e.
the short root α = αshort and the long root β = αlong . In particular, being two
variables, what we should find is merely one singular hyperplane in the (z1 , z2 )-
plane, namely, a line of the form az1 + bz2 + c = 0. To find the singular line,
we first concentrate on the rational factor hwλ−ρ,α∨ i1hwλ−ρ,α∨ i , up to a change of
short long
variables λ 7→ wλ, there are only two of them, namely,
hλ − ρ, α∨short i = 0 and hwλ − ρ, α∨long i = 0. (8.50)
January 18, 2018 9:46 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 225
At this point, by examining what we have done for the groups SLn and Sp2n ,
it is not difficult for us to detect that the singular hyper-planes along which the
residues are
(a) taken in the denominators of the rational functional factors in ωGQ (λ), i.e. in
1
, (8.51)
α∈∆ hwλ − ρ, α i
∨
Q
(b) concentrated on these associated to the identity Weyl element Id, i.e. on
1
. (8.52)
α∈∆ hλ − ρ, α i
∨
Q
In this way, we obtain two new zeta functions for the group G2 , namely,
Reshλ−ρ,α∨short i=0 ωGQ2 (λ) , and Reshλ−ρ,α∨long i=0 ωGQ2 (λ) .
Since there is a one-to-one correspondence between the set of maximal (stan-
dard) parabolic subgroups and the set of simple roots, if we denote by Pshort and
Plong for the maximal parabolic subgroup of G2 corresponding to the (set of) sim-
ple root αlong and αshort , respectively, it is then only natural for us to name the
corresponding zeta functions (after the normalization process similarly to that
G /P
for zeta functions of SLn and Sp2n ) as b ζQ 2 long (s) and bζQG2 /Pshort (s), or the same,
/P
ζ G2 /Pα (s), respectively.
G
ζ 2 β (s) and b
b
Q Q
From the period ωGQ2 (z1 , z2 ) of G2 over Q listed in Table 8.4, to obtain a zeta func-
ζ G2 /Pα (s) for G2 /Pα , we first take the residue along with the singular hyper-
tion b Q
plane hλ+ρ0 , α∨ i = 0 of ωGQ2 (z1 , z2 ), namely the line z1 −z2 −1 = 0. Set s = z2 (then
z1 = 1+s, z2 −z1 = −1, 2z1 +z2 = 3s+2, z1 +z2 = 2s+1, z1 +2z2 = 3s+1, z1 −1 = s
and z2 + 1 = s + 1). In such a way, we get the following (single variable) period
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 226
ωGQ2 /Pα (s) associated to G2 /Pα over Q (with the same order as in Table8.4).
ωGQ2 /Pα (s) := Resz1 −s−1=0 ωGQ2 (z1 , s)
1 1 1 1 1 1 1 b ζ(s + 1) bζ(3s + 2)
= + · +0+ ·
s − 1 −2 s b ζ(2) −3s − 3 2s ζ(2) ζ(s + 2) ζ(3s + 3)
b b b
1 1 1 ζ(s + 1) ζ(3s + 2) ζ(2s + 1) b ζ(3s + 1)
+
b b b
·
−3s − 2 s − 1 b ζ(2) bζ(s + 2) b ζ(3s + 3) bζ(2s + 2) b ζ(3s + 2)
1 ζ(s + 1) b ζ(3s + 2) b ζ(2s + 1) bζ(3s + 1) b ζ(s)
+ +0+0+0
b
·
−s − 2 b ζ(s + 2) b ζ(3s + 3) b ζ(2s + 2) bζ(3s + 2) b ζ(s + 1)
1 1 1 bζ(s + 1) b ζ(3s + 2) bζ(2s + 1) b ζ(3s + 1) b ζ(s)
+ ·
−2 −s − 1 b ζ(2) b
ζ(s + 2) b ζ(3s + 3) bζ(2s + 2) b ζ(3s + 2) b ζ(s + 1)
1 1 1 ζ(s + 1) ζ(3s + 2) ζ(2s + 1) 1 1 1 b ζ(s + 1)
+ + .
b b b
· ·
3s −2s − 2 b ζ(2) ζ(s + 2) ζ(3s + 3) ζ(2s + 2)
b b b 3s + 1 −s − 2 ζ(2) ζ(s + 2)
b b
Here, we have used the functional equation of b ζ(s) to change b ζ(az + b) with a < 0
or a = 0, b < 0 to ζ(−az − b + 1). Taking the cancellations of the common
b
factors in both the denominator and numerators, and multiplying with the factor
ζ(2) · b
b ζ(s + 2)bζ(2s + 2)b ζ(3s + 3), to clear out all the zeta factors in the determinators
of all terms in ωGQ2 /Pα (s), we arrive at
G2 /Pα 1 b
ζQ,o
b (s) := ζ(2) · b
ζ(s + 2)b ζ(2s + 2)b ζ(3s + 3)
s−1
1 b 1 b
− ζ(2) · bζ(s)b ζ(2s + 1)b ζ(3s + 1) − · ζ(s + 2)b ζ(2s + 2)bζ(3s + 3)
s+2 2s
1 1 1 b
+ ·bζ(s)b ζ(2s + 1)b ζ(3s + 1) − · ζ(s + 1)bζ(2s + 2)b ζ(3s + 2)
2(s + 1) 3s + 3 2s
1 1 b 1 1 b
− ζ(s + 1)b ζ(2s + 1)b ζ(3s + 2)− ζ(s + 1)bζ(2s + 1)bζ(3s + 1)
3s 2s + 2 3s + 2 s − 1
1 1 b
− · ζ(s + 1)bζ(2s + 2)b ζ(3s + 3).
3s + 1 s + 2
G2 /Pα G2 /Pα
One checks easily that b ζQ,o (−1 − s) = b ζQ,o (s).
G /P G /P
ζQ,o2 β (s) := Reshλ+ρ0 ,β∨ i=0 ωQ2 β (z1 , z2 )
b
1 1 1 1
= +0+ · +0
s−2 s + 1 −2 b ζ(2)
1 ζ(s − 1) b ζ(s) b ζ(2s + 1) b
ζ(s + 1) b
ζ(s + 2)
+
b
·
−s − 3 ζ(s) ζ(s + 1) ζ(2s + 2) ζ(s + 2) ζ(s + 3)
b b b b b
1 1 ζ(s) b ζ(2s + 1) b
ζ(s + 1) b
ζ(s + 2) 1
+
b
·
s − 2 −s − 1 b ζ(s + 1) b
ζ(2s + 2) b
ζ(s + 2) b
ζ(s + 3) bζ(2)
1 1 ζ(s + 1) b
ζ(s + 2) 1 1 1 b ζ(s + 2) 1
+ +
b
· ·
2s −s − 2 b ζ(s + 2) b
ζ(s + 3) b
ζ(2) −s − 3 s b ζ(s + 3) b
ζ(2)
1 1 b ζ(s − 1) b ζ(s) b ζ(2s + 1) b
ζ(s + 1) b
ζ(s + 2) 1
+ · + 0 + 0.
−s −2 ζ(s) ζ(s + 1) ζ(2s + 2) ζ(s + 2) ζ(s + 3) b
b b b b b ζ(2)
ζ(2) · b
Multiplying with b ζ(s + 3)b
ζ(2s + 2), and shifting from s to s − 1 we arrive at
the following:
G /Pβ
ζQ 2
Definition 8.10. The zeta function b (s) for G2 /Pβ is defined by
G /P 1 b 1 b
ζQ 2 β (s) :=
b ζ(2) · b
ζ(s + 2)b
ζ(2s) − ζ(2) · b
ζ(s − 2)b
ζ(2s − 1)
s−3 s+2
1 1 b
+ ·bζ(s − 2)b
ζ(2s − 1) − · ζ(s + 2)b
ζ(2s)
2s − 2 2s
1 1
− ζ(s − 1)b
·b ζ(2s − 1) − ·bζ(s + 1)b
ζ(2s)
s(s − 3) (s − 1)(s + 2)
1 1
− ζ(s)b
·b ζ(2s) − ζ(s)b
·b ζ(2s − 1).
(2s − 2)(s + 1) (2s)(s − 2)
G /Pβ
ζQ 2
Proposition 8.6 (Functional Equation). The zeta function b (s) for G2 /Pβ
satisfies the functional equation
G /P G /P
ζQ 2 β (1 − s) = b
b ζQ 2 β (s).
G /P
In addition, bζQ 2 β (s) admits 4 singularities, namely, four simple poles located at
s = −2, 0, 1, 3.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 228
The first is about a refine statement of the Hadamard products for entire functions.
Lemma 8.3 ([110]). Let F(s) be an entire function of genus zero or one. Assume
Proof. This is rather standard. We only treat the case that F(s) is of genus one,
since the genus zero case can be proved similarly. Obviously, F(s) is of genus one
if and only if the Hadamard product factorization
Y s
F(s) = eA+Bs sm 1− exp(s/ρ) (m ∈ Z≥0 ) (8.57)
ρ
ρ
where the outer product on the right-hand side converges absolutely and uni-
formly on any compact subsets of C. Indeed, for each B(ρ), the convergence
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 229
factor exp(c(B(ρ))s) coincides with c(B(ρ)) = 2β|ρ|−2 for γ > 0. But (b) implies
|β − 1/2| < σ0 , and hence
X X X
|c(B(ρ))| ≤ |ρ|−1 + (2σ0 + 1) |ρ|−2 < ∞.
B(ρ) 0,ρ: real ρ
Proof. We use conditions (c), (d) and (e) to verify it. Indeed, by (e) we have
F(1 − σ)
!
R 3 log → −∞ as σ → +∞. (8.59)
F(σ)
On the other hand, by (8.58) we have
F(1 − σ) 0
σ − 1 m Y σ − 1 + β Y (σ − 1 + β)2 + γ2
= eB (1−2σ) .
F(σ) σ ρ=β∈R
σ − β ρ=β+iγ (σ − β)2 + γ2
γ>0
since the number of real zeros is also finite by (b) and (c). But, for these same σ,
X (1 − 2β)(2σ − 1) X (1 − 2σ0 )(2σ − 1)
log 1 − ≤ log 1 −
ρ=β+iγ
(σ − β)2 + γ2 ρ=β+iγ
(σ − 1/2)2 + γ2
γ>0 γ>0
X 1
(2σ − 1) ,
ρ=β+iγ
(σ − 1/2)2 + γ2
γ>0
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 230
Z ∞
dN(t)
which is nothing but the Stieltjes integral . Thus, by (d),
γ0 (σ − 1/2)2 + t2
∞ ∞
log(σ + γ0 )
Z Z
dN(t) (log t) dt
.
γ0 (σ − 1/2)2 + t2 γ0 (σ − 1/2) + t
2 2 σ − 1/2
X (1 − 2β)(2σ − 1)
This implies that log 1 − log(σ + γ0 ), which contra-
ρ=β+iγ
(σ − β)2 + γ2
γ>0
dicts with (8.62). Therefore we have verified the claim and the lemma.
To end this discussion, we recall the following lemma for later use.
Lemma 8.4. ( [65]) Let f be an entire function of genus zero or one such that
Proof. The first inequality, resp. the second equality, of (11.26) is a direct con-
sequence of (8.63), resp. the functional equation for ξ(s), it suffices to prove the
assertions in 11.26. The arguments before Claim 8.1 in the proof of the previous
lemma are valid for f (s) as well. Hence, we have the Hadamard product for f (s)
in the form
A+B0 s
Y Y s
f (s) = e 1 −
ρ
B(ρ) ρ∈B(ρ)
To complete the proof, it suffices to use the standard properties for the Möbius
s + (c − ρ)
transformation s 7→ τ = , which transforms the half plane <(s) > 21 ,
s − (c + ρ)
resp. <(s) < 12 , to the outside, resp. inside, of the circle |τ| = 1.
Proposition 8.7. ( [112]) There are constants B0i ≥ 0 (i = 1, 2) such that the
following Hadamard products for the entire functions fi (s) (i = 1, 2) in (8.68) are
convergent absolutely on any compact subset of C
! !
0 s Y s
fi (s) = ci smi eBi s 1 − 1−
β0,i βi <1/2 βi
0,βi ∈R
! ! Y " ! !#
s s s s
× 1− 1− 1− 1− .
ρ0,i ρ0,i <(ρi )<1/2
ρi ρi
=(ρ)i >0
By Lemma 8.3, to prove this proposition, it suffices to show that both f1 (s)
and f2 (s) satisfy all the five conditions in Lemma 8.3. Obviously, (a) is satisfied.
In addition, if (b) holds, (d) can be easily verified using well-known estimate
|ξ(s)| ≤ exp(C|s| log |s|)
and Jensen’s formula (see §4.1 of [109], for example). Hence it suffices to verify
the conditions (b), (c) and (e) for both f1 (s) and f2 (s), since
f1 (0) = 0, f1 (s) = f10 (0)s + O(s2 ), f10 (0) ' −2.176 , 0, f2 (0) ' −6.283 , 0.
Lemma 8.6. There exist constants σ1 and σ1 such that all zeros of f1 (s) and f2 (s)
are contained in the strip σ1 < <(s) < σ2 .
In addition, since all |arg((1−3s)/2)| and |arg(2−3s)/2|; −s/2, (1− s)/2, (1−3s)/2
and (2 − 3s)/2; and arg(−1 − s)/2, arg(1 − s)/2 are strictly less than π/2. Hence
we can apply the Stirling formula for <(s) < 0, and then
r
Γ((1 − 3s)/2) 2 −1/2
= |s| (1 + O(|s|−1 )),
Γ((2 − 3s)/2) 3
Γ(−s/2) √ Γ((−1 − s)/2)
= 2 |s|−1/2 (1 + O(|s|−1 )), = 2 |s|−1 (1 + O(|s|−1 )).
Γ((1 − s)/2) Γ((1 − s)/2)
On the other hand, as σ → −∞,
ζ(1 − 3σ)ζ(2 − 3σ) → 1, ζ(−σ)ζ(1 − σ)ζ(1 − 3σ)ζ(2 − 3σ) → 1,
ζ(−1 − σ)ζ(1 − σ) → 1, ζ(−σ)ζ(1 − σ) → 1.
Therefore, if σ = <(s) < 0, and |s|, |σ| are both large,
r
8π √
|R11 (s)| ≤ · |s|−1/2 · (1 + O(|s|−1 )), |R12 (s)| ≤ 2 3πA · (1 + O(|s|−1 )),
3
√
|R21 (s)| ≤ 8π · |s|−1/2 · (1 + O(|s|−1 )), |R22 (s)| ≤ 2πA · (1 + O(|s|−1 )),
as desired.
Lemma 8.7. There are at most finitely many zeros of the entire function f1 (s),
resp. f2 (s), in the region <(s) > 1/3, resp. <(s) > 0. In particular, there are at
most finitely many zeros of f1 (s) and f2 (s) in the right-half plane <(s) ≥ 1/2.
Proof. We have
f1 (s) = (s − 1)2 (3s − 2)(As − A + 1)ξ(s + 1)ξ(3s) [ 1 − Q11 (s) − Q12 (s) ] ,
(8.79)
f2 (s) = (As + 3)(s − 1)2 (s − 2)ξ(s + 2) [ 1 − Q21 (s) − Q22 (s) ] ,
where
(s + 1)(s − 2) ξ(s)ξ(3s − 1)
Q11 (s) = · ,
(s − 1)(3s − 2)(As − A + 1) ξ(s + 1)ξ(3s)
2(s − 2) ξ(s)
Q12 (s) = · ,
(3s − 2)(As − A + 1) ξ(s + 1)
2(s − 3) ξ(s + 1) (s + 2)(s − 3) ξ(s)
Q21 (s) = · , Q22 (s) = · .
(As + 3)(s − 1) ξ(s + 2) (As + 3)(s − 1) ξ(s + 2)
2
(s + 2)(s − 3)
( )
2(s − 3)
D2 := s ∈ C <(s) ≥ 0, + ≥ 1 .
(As + 3)(s − 1) (As + 3)(s − 1)2
Then, by (8.79) and (8.80), f1 (s) , 0, resp. f2 (s) , 0, if s < D1 and <(s) ≥ 1/3,
resp. if s < D2 and <(s) ≥ 0. The region D1 is bounded, since
(s + 1)(s − 2) 2(s − 2)
+ <1
(s − 1)(3s − 2)(As − A + 1) (3s − 2)(As − A + 1)
for large |s|. Similarly, the region D2 is bounded. Therefore, there are at most
finitely many of zeros for f1 (s), resp. f2 (s), in <(s) ≥ 1/3, resp. in <(s) ≥ 0.
8.3.2.5 Positivity
By the results above, we can apply Lemma 8.3 to the functions fi (s) (i = 1, 2).
Hence to complete the proof of Proposition 8.7, it suffices to prove the following:
Lemma 8.8. There are only three zeros for each of f1 (s) and f2 (s) in the right
half plane <(s) ≥ 1/2. More precisely, for f1 (s), resp. for f2 (s), these are given
by one real zero at s = 1, resp. s = 2, and other two non-real conjugate zeros
approximated located at s ' 0.927 ± i · 2.09, resp. s ' 1.17 ± i · 3.43.
Theorem 8.7.
(1) All zeros of Z1 (s) lie on the line Re(s) = 1/2 except for four simple zeros
s = 0, 1/3, 2/3, 1.
(2) All zeros of Z2 (s) lie on the line Re(s) = 1/2 except for four simple zeros
s = −1, 0, 1, 2.
To prove Theorem 8.7, we first analysis the zeros of Zi (s) in the right half plane.
Proposition 8.8. There is no zero for the functions Z1 (s) and Z2 (s) in the right-half
plane <(s) ≥ 20.
Proof. We have
where
(s + 1)(s − 2) ξ(s) ξ(3s − 1)
R11 (s) = ,
(s − 1)(3s − 2)(As − A + 1) ξ(s + 1) ξ(3s)
2(s − 1)(s − 2) ξ(s)
R12 (s) = ,
(s − 1)(3s − 2)(As − A + 1) ξ(s + 1)
s2 (3s − 1)(As − 1) ξ(s − 1) ξ(3s − 2) ξ(2s − 1)
R13 (s) = ,
(s − 1) (3s − 2)(As − A + 1) ξ(s + 1) ξ(3s)
2 ξ(2s)
s(s + 1)(s − 2) ξ(s) ξ(3s − 1) ξ(2s − 1)
R14 (s) = ,
(s − 1)2 (3s − 2)(As − A + 1) ξ(s + 1) ξ(3s) ξ(2s)
2s2 (s + 1) ξ(s) ξ(3s − 2) ξ(2s − 1)
R15 (s) = ,
(s − 1) (3s − 2)(As − A + 1) ξ(s + 1) ξ(3s)
2 ξ(2s)
2(s − 3) ξ(s + 1)
R21 (s) = ,
(s − 1)(As + 3) ξ(s + 2)
(s + 2)(s − 3) ξ(s)
R22 (s) = ,
(s − 1)2 (As + 3) ξ(s + 2)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 238
|R11 (s)| ≤ 0.1, |R12 (s)| ≤ 0.3, |R13 (s)| ≤ 0.05 |R14 (s)| ≤ 0.1, |R15 (s)| ≤ 0.1,
|R21 (s)| ≤ 0.3, |R22 (s)| ≤ 0.13, |R23 (s)| ≤ 0.15 |R24 (s)| ≤ 0.2, |R25 (s)| ≤ 0.1
for <(s) ≥ 20 (in fact, these bounds already hold for <(s) ≥ 10). These estimates
imply Z1 (s) , 0 and Z2 (s) , 0 for <(s) ≥ 20 by (8.81), since (s − 1)2 (3s − 2)(As −
A + 1)ξ(s + 1)ξ(3s)ξ(2s), resp. (s − 1)2 (s − 2)(As + 3)ξ(s + 2)ξ(2s), has no zero in
the right-half plane <(s) ≥ 20.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 239
Proposition 8.9. There is no zero for Z1 (s), resp. Z2 (s), in the part of the strip
1/2 < σ < 20 characterized by |t| ≥ 25, resp. |t| ≥ 36.
√
Proof. Let ρi0 = βi0 + −1γi0 (i = 1, 2, γi0 > 0) be the complex zero of fi (s) in
Lemma 8.8. By Proposition 8.7, f1 (s) and f2 (s) has the factorization
0
s s
fi (s) = fi0 (0) eBi s s(1 − s) 1 − 1− ·Πi (s) (i = 1, 2, B0i ≥ 0),
ρi0 ρi0
where
" #
Y s Y s s
Πi (s) = 1− 1− 1− .
0,β∈R
β ρ=β+iγ ρ ρ
β<1/2, γ>0
Note that all zeros of Πi (s) (i = 1, 2) lie in σi0 < <(s) < 1/2 for some σi0 . Hence
g1 (1 − s)
Z1 (s) =g1 (s) · 1 − (g1 (s) = f1 (s) · ξ(2s)),
g1 (s)
(8.82)
g2 (1 − s)
Z2 (s) =g2 (s) · 1 − (g2 (s) = f2 (s) · ξ(2s))
g2 (s)
and
g1 (1 − s) 0 Π1 (1 − s) s − 1 + ρ0 s − 1 + ρ0 ξ(2s − 1)
=eB1 (1−2σ) ,
g1 (s) Π1 (s) s − ρ0 s − ρ0 ξ(2s)
(8.83)
g2 (1 − s) 0 Π2 (1− s) s s−1+ρ0 s−1+ρ0 ξ(2s−1)
=eB2 (1−2σ) .
g2 (s) Π2 (s) 1− s s − ρ0 s − ρ0 ξ(2s)
Because B0i ≥ 0, we have
0
eBi (1−2σ) ≤ 1 (<(s) > 1/2). (8.84)
Lemma 8.9. Let ρ10 = β10 + iγ10 ' 0.927 + i · 2.09, resp. ρ20 = β20 + iγ20 '
1.17 + i · 3.43, be the complex zero of f1 (s), resp. f2 (s), in Lemma 8.8. Let
s = σ + it be a point in the part of the strip 1/2 < σ ≤ 20 characterized by t ≥ 25,
resp. t ≥ 36. Then there exists at least two distinct zeros ρ = β + iγ of b
ζ(s) such
that 0 < β ≤ 1/2, |t − γ| ≤ 22,
s − 1 + ρ0 2s − 1 − (1 − ρ) s − 1 + ρ0 2s − 1 − (1 − ρ)
< 1, < 1, (8.87)
s − ρ0 2s − ρ s − ρ0 2s − ρ
and
s − 1 + ρ0 2s − 1 − (1 − ρ) s 2s − 1 − (1 − ρ)
· < 1, · < 1. (8.88)
s − ρ0 2s − ρ s−1 2s − ρ
√
Proof. By an abuse of notations, we write ρ0 = β0 + −1γ0 for ρi0 . By squaring
(8.87) and (8.88), we have
(σ + β0 − 1)2 + (t ± γ0 )2 (2σ + β − 2)2 + (t − γ)2
· < 1. (8.89)
(σ − β0 )2 + (t ± γ0 )2 (2σ − β)2 + (t − γ)2
Claim 8.2. ([64]) For any real value of t there exists at least three distinct zeros
ρ = β + iγ of b
ζ(s) such that 0 < β ≤ 1/2 and
|t − γ| ≤ 22. (8.90)
Proof. Suppose |t| ≥ 25. Then there exists at least three distinct zeros ρ = β + iγ
of bζ(s) satisfying 0 < β ≤ 1/2 and |t − γ| < 15.1 by applying Lemma 5 in [109]
to t + 10.1 and t − 10.1 (Lemma 5 in [109] is essentially Lemma 3.5 of [64]). For
|t| < 25, estimate (8.90) also holds for three distinct zeros because b
ζ(s) has zeros
at s = ±14.13, ±21.02, ±25.01.
Therefore, it suffices to show that (8.89) holds for 0 < β ≤ 1/2, |t − γ| < 22,
1/2 < σ ≤ 20 and t ≥ 25, namely to show
(σ + β0 − 1)2 + (t ± γ0 )2 (2σ − 2 )2 + 222
3
· <1
(σ − β0 )2 + (t ± γ0 )2 (2σ − 12 )2 + 222
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 241
by the similar reasons in the later half of section 4.3 in [109]. Obviously, the
inequality (8.3.3.2) is equivalent to
(2σ − 1) 8(t ± γ0 )2 − P(σ) > 0, (8.91)
where P(σ) = 8(4β0 − 3)σ2 − 8(4β0 − 3)σ − 8β20 + 3890β0 − 1945.
Using the value β0 ' 0.927, resp. 1.17, we see that P(σ) < 3807, resp.
< 7777, for 1/2 < σ < 20. On the other hand, using the value γ0 ' 2.09, resp.
3.43, we see that 8(t ± γ0 )2 > 3872, resp. > 8192, for t ≥ 25, resp. ≥ 36 since
|t ± γ0 | = t ± γ0 > 22, resp. > 32, for t ≥ 25, resp. ≥ 36. Hence (8.91) hold, and it
implies (8.89).
Proposition 8.10 (Riemann Hypothesis). There are only two zeros for Z1 (s),
resp. Z2 (s), namely, the simple zeros at s = 2/3, 1, resp. s = 1, 2 in the box
1/2 < σ < 20, |t| ≤ 25, resp. 1/2 < σ < 20, |t| ≤ 36.
Proof. Because the boxes characterized by 1/2 < σ ≤ 20, |t| ≤ 25 or 36 are finite,
we can computationally check the assertions of Proposition 8.10 as in the proof of
Lemma 8.8, say using Mathematica. The outcome is as stated.
Proof of Theorems 8.7 and 8.6. As a direct consequence of Propositions 8.8, 8.9,
8.10, and the functional equation of Z1 (s), resp. Z2 (s), it is easily to conclude that
all zeros of Z1 (s), resp. Z2 (s), lie on the line <(s) = 1/2 except for simple zeros
located at s = 0, 1/3, 1/2, 2/3, 1, resp. s = −1, 0, 1/2, 1, 2. This completes the
proof of Theorems 8.7 and 8.6. That is, the Riemann hypothesis holds for the zeta
functions associated to the exceptional group of type G2 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 242
From above, starting from the period ωGQ (λ) of G governed by huge symmetries
ζ hλ, α∨ i
!
X 1 Y b
ωGQ (λ) = · , (8.93)
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
w∈W ζ hλ, α∨ i + 1
to find the singular hyperplanes along which the residues are taken, our success in
introducing the zeta functions for SLn , Sp2n and G2 leads to the rational function
1
. (8.94)
α∈∆ hλ − ρ, α i
∨
Q
(a) There are exactly (r − 1)-singular hyperplanes along which we should take
residues, and
(b) all these (r − 1)-singular hyperplanes are of the forms hλ − ρ, α∨ i, taking from
the total r-factors in the denominator of (8.94).
This then exposes the hidden role played by the maximal parabolic subgroup
P of G. More precisely, for a standard maximal parabolic subgroup P of G, there
exists one and only one simple root αP such that, under the inverse order one-to-
one correspondence between parabolic subgroups of G and subsets of the set of
simple roots ∆, the maximal parabolic subgroup P of G is associated to the subset
∆ r {αP } of ∆. Therefore, we have the following interpretations of (a) and (b).
(?) The (r − 1) singular hyperplanes for taking the residues are given by
hλ − ρ, α∨ i = 0 α ∈ ∆ r {αP }. (8.95)
Upon this point, we are very sure how a new type of zeta functions for G, or
better, for the pairs (G, P), should be introduced.
Let G be a (split) reductive group of rank r over a number field F. Fix a mini-
mal parabolic subgroup P0 of G and let M0 be the Levi subgroup of P0 and let
T 0 be the maximal split torus of the center of M0 . Since T 0 contains only semi-
simple elements (which all commute with each other), it follows that, with respect
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 243
n o
αP ∈ ∆. Hence it makes sense to write ∆ r {αP } =: βP,1 , βP,2 , . . . , βP,r−1 . For
simplicity, we sometimes write αP as α p for a certain 1 ≤ p ≤ r. Accordingly, for
a meromorphic function f (λ), after taking the residues along (r − 1) hyperplanes
hλ, β∨P,k i = 0 (1 ≤ k ≤ r − 1), the resulting function
Reshλ−ρ,β∨P,r(G)−1 i=0 · · · Reshλ−ρ,β∨P,2 i=0 Reshλ−ρ,β∨P,1 i=0 f (λ)
or the same,
Res sr =0 . . . Res s p =0 . . . Res s1 =0 f (λ)
[
Obviously, after taking the residues, there are still zeta factors left in the de-
nominators of the terms in ωG/P F (s), even after the cancelations of the common
zeta factors in both the numerators and the denominators. It is not too difficult to
see that there are a certain minimal integer I(G/P) ≥ 0 and some finitely many
non-trivial zeta function factors
ζF aG/P
b
1 s P + b1
G/P
, bζF aG/P
2 s P + b2
G/P
, ..., bζF aG/P
I(G/P) sP + bI(G/P) ,
G/P
Theorem 8.8 (Functional Equation). There exists a constant cG/P ∈ Q such that
ζF;o
G/P
− s + cG/P = b
ζF;o
G/P
s.
b (8.104)
This theorem was first verified for the lower rank groups SLn (n = 2, 3, 4, 5),
Sp4 and G2 , by myself, then proved for (G, P) = (SLn , Pn−1,1 ) by Kim-Weng
in [58], and finally for all (G, P) by Komori [59]. For details, please refer to the
next two chapters.
Definition 8.13. The zeta function b ζFG/P s for (G, P) over F is defined by
!
cG/P − 1
bζ s := ζF;o s +
G/P b G/P
. (8.105)
2
By Theorem 8.8, b ζ G/P s admits the standard functional equation
ζFG/P 1 − s = bζFG/P s .
b (8.106)
Chapter 9
247
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 248
notations simply as p. In the sequel till the end of this chapter, unless otherwise is
stated, P stands for a maximal standard parabolic subgroup of (G, P0 ), and if there
is another maximal standard parabolic subgroup Q, we use the index q, 1 ≤ q ≤ n
instead.
For α ∈ Φ, denote by σα : V → V, v 7→ v − hv, αiα∨ the reflection with
respect to the α-perpendicular hyperplane in V to σ. Let W be the associated
Weyl group generated by n simple reflections σ j := σα j associated to simple roots
α j (1 ≤ j ≤ n). It is well known that W is also generated by reflections associated
to all the positive roots α ∈ Φ+ , or equivalently, to all the roots α ∈ Φ, and an
element w ∈ W admits a shortest decomposition as a product of σ j ’s. For a Weyl
element w ∈ W, let
Φw := Φ+ ∩ w−1 (Φ− ). (9.2)
Then, Φw is the set of positive roots whose w-image is negative, and |ΦW | co-
incides with the length of w, defined to be the number of elements in a shortest
decomposition of w interns of the σi ’s. Let w0 be the longest element of W. By
the theory of Coxeter groups, there is one and only one such an element w0 in W
such that w20 = Id, w0 (∆) = −∆ and hence w0 (Φ+ ) = Φ− .
Let P be a maximal standard parabolic subgroup of G. For simplicity, we
write ∆ p = β1 , . . . , βn−1 . Let Φ p be the collection of elements in Φ which are
perpendicular to λ p and set Φ±p := Φ p ∩ Φ± . It is not difficult to show that Φ p is
a root system. Accordingly, Φ+p is generated by ∆ p with Z≥0 coefficients, and, all
elements of the associated Weyl group W p , generated by σα for all σ ∈ Φ p , admit
shortest decompositions in terms of the σ j ’s with j , p. In particular, there exists
a unique longest element w p in W p such that
w p (∆P ) = −∆ p and w p (Φ+p ) = Φ−p . (9.3)
1 X X r
Similarly, set ρ p := α. In general, ρ p , λ j.
2 α∈Φ+ j=1,,p
p
ωG/P
F (s; T ) := Reshλ−ρ, β1 i=0 . . . Reshλ−ρ, βn−1 i=0 ωF (λ; T ).
∨ ∨
G
(9.7)
ωG/P
F (s) := ωF (s; 0).
G/P
(9.8)
Here, ωGF (λ; T ) denotes the T -period of G over F in Definition 7.8, namely,
X Y 1 Y b ζ F (hλ, α ∨
i)
ωF (λ; T ) :=
G
ehwλ−ρ,T i . (9.9)
hwλ − ρ, α i α∈Φ b
∨
ζ (hλ, α ∨ i + 1)
w∈W α∈∆ w F
where
Y 1 Y b ζF (hλ, α∨ i)
Aw (λ) := . (9.11)
α∈∆
hwλ − ρ, α∨ i α∈Φ b ζF (hλ, α∨ i + 1)
w
∆ p = α ∈ ∆ p : wα ∈ ∆ ∪ Φ− .
(9.12)
Proposition 9.1. The T -period of (G, P) and hence also the period of (G, P) over
F are given by the summations over special Weyl elements. More precisely,
X Y
hw(sλ p +ρ)−ρ,T i |Φw ∩∆ p | 1
ωF (s; T ) :=
G/P
e cF
, α ∨ is+hρ, α∨ i−1
hλ p
w∈W 0 α∈w−1 (∆)r∆ p
(9.13)
Y Y 1
ζF hλ p , α∨ is + hρ, α∨ i .
× b
α∈Φw r∆ p α∈(−Φw )
bζ F hλ p , α ∨ is + hρ, α∨ i
ζF (s) at s = 1.
Here cF denotes the residue of b
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 250
The essentials for this proof come from [57], or better, [58] on functional
equation of group zeta functions associated to (G, P) = (SLn , Pn−1,1 ). See also
[59].
Here Res or lim under b means the corresponding term is omitted as usual.
Therefore, to get a non-zero contribution for the particular w ∈ W, there should
be cancellations between either the common factors appeared in both numerator
and denominator in the first big bracket, or the rational factors in the numerator of
the first big bracket and the same number of poles from the zeta function factors
in the second big bracket. Recall that the complete Dedekind zeta function b ζF (s)
of F admits only two singularities, namely, two simple poles at s = 0, 1. Set
cF := Res s=1 ζF (s). Then, using (9.14) again, naturally, we should rewrite the
b
Y
product hλ − ρ, β∨ i · Aw (λ) as
β∈∆ p
Y Y 1 Y 1
β∨
hλ, i−1
β∈∆ p
hλ, α∨ i−1 hλ, α∨ i−1
α∈(w (∆)∪Φw )∩∆ p
−1 α∈(w (∆))r∆ p
−1
(9.15)
Y hλ, α∨ i − 1 b ζF (hλ, α∨ i) Y ζ (hλ, α ∨
i)
F .
b
×
α∈Φw ∩∆ p ζF (hλ, α∨ i + 1)
b
α∈Φw r∆ p ζF (hλ, α i + 1)
b ∨
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 251
In particular, when s j → 0 for all 1 ≤ j ≤ n, j , p, all the factors in the first big
bracket should be cancelled out with those in either the second and or the fourth
big brackets. Indeed, we have the following:
hλ, α∨ i − 1 = hλ p , α∨ i s + hρ, α∨ i − 1 . 0.
(c) For α ∈ Φw ∩ ∆ p , within the fourth group of factors,
hλ, α∨ i − 1 bζF (hλ, α∨ i)
cF
−→ + o(sk ).
ζF (hλ, α∨ i + 1)
b ζF (2)
b
ζF (hλ, α∨ i)
(d) For α ∈ Φw r ∆ p , within the fifth group of factors, the ratio
b
ζF (hλ, α∨ i + 1)
b
is a well-defined meromorphic function.
Consequently,
for a fixed w ∈ W, to have a non-zero constant contribution
Y
from hλ − ρ, β∨ i · Aw (λ) in the case s p = s and all sk = 0 for k , p, all the
β∈∆ p
hλ − ρ, β∨ i’s in the first group of factors in (9.15) should be cancelled by either the
factors in (a) or the factors in (c) completely. This implies that, for this particular
w ∈ W, we must have that ∆ p ⊂ w−1 (∆) ∩ Φw , or equivalently, that w satisfies
the condition that ∆ p = α ∈ ∆ p : wα ∈ ∆ ∪ Φ− . Namely, w ∈ W 0 is special.
Moreover, by Lemma 9.1, we conclude that
X
w(sλ p +ρ)−ρ,T i
Y 1
ωF (s; T ) =
G/P
e
, α ∨ i s + hρ, α∨ i − 1
hλ p
w∈W 0 α∈(w−1 (∆))r∆ p
Y c F
Y ζ
b F (hλ p , α ∨
i s + hρ, α∨
i)
×
α∈Φw ∩∆ p ζF (2) α∈Φw r∆ p ζF (hλ p , α i s + hρ, α i + 1)
b b ∨ ∨
X Y 1
ew(sλ p +ρ)−ρ,T i cF w p
|Φ ∩∆ |
=
hλ p , α∨ i s + hρ, α∨ i − 1
w∈W 0 α∈(w (∆))r∆ p
−1
(9.18)
Y Y 1
× ζF (hλ p , α i s + hρ, α i)
∨ ∨
b
α∈Φw ζF (hλ p , α is+hρ, α i+1)
∨ ∨
α∈Φw r∆ p
b
X Y 1
ew(sλ p +ρ)−ρ,T i cF
|Φ ∩∆ |
= w p
hλ p , α∨ i s + hρ, α∨ i − 1
w∈W 0 α∈(w (∆)r∆ p
−1
Y Y 1
× ζF (hλ p , α i s + hρ, α i)
∨ ∨ .
b
α∈Φw r∆ p α∈(−Φw ) Fζ
b (hλ p , α ∨ i s+hρ, α∨ i)
Hence, to complete the proof of the proposition, it suffices to apply the functional
ζF (1 − s) = b
equation b ζF (s) to the final group of factors.
From now on till the end of this chapter, we will specialize our discussion to
the case that G is the special linear group SLn and P is the maximal parabolic
subgroup P consisting of matrices whose final row vanishes except for its last
entry, corresponding to the ordered partition (n − 1) + 1 of n.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 253
Since we are working on SLn , let V e = (Rn , h·, ·i) denote the canonical Eu-
n
clidean space R , namely, the n-dimensional real vector space equipped with the
n
X
standard Euclidean metric h(xi ), (yi )i := xi yi . Denote the associated canon-
i=1
n
e1 , . . . , en so that Rn 3 (x1 , . . . , xn ) = xi ei . Then,
X
ical orthonormal basis by
i=1
by §6.1.3, for the semi-simple Lie group SLn , its root space is given by
Xn
V = (x1 , . . . , xn ) ∈ V xi = 1 , and the set ∆ of simple roots is given by
e:
i=1
∆ = α1 , . . . , αn−1 , where
respectively. In addition, with an identification of V and its dual space V ∨ by h·, ·i,
the associated fundamental weights λ j nj=1 are characterized by
j n
j X
ek − ei
X
λj = 1 ≤ j ≤ n − 1. (9.20)
k=1
n i=1
In particular,
n−1
1 X X 1
ρ= α= λ j = n − 1, n − 3, . . . , −(n − 1) (9.21)
2 α∈Φ+ j=1
2
For later use, we will also write ∆P , Φ+P , ρP and λP simply as ∆0 , Φ0 + and ρ0 , $0
respectively. Easily, ∆0 = α1 , . . . , αn−2 , Φ0 + = {ei − e j : 1 ≤ i < j ≤ n − 1},
Let S0n be the subset of Sn , which, under the identification Sn ' W, corresponds
to W 0 . We call elements of S0n special permutations. Since |∆P | = |∆| − 1 = n − 1,
by definition, σ ∈ S0n is special if and only if
α ∈ ∆P : wα < 0 ∪ α ∈ ∆ : σ(α) ∈ ∆P has cardinality n − 2.
(1) The permutation σ is special if and only if σ satisfies the following condition:
{σ ∈ ∆P : σα < 0 or σα ∈ ∆ = ∆P . (9.29)
(2) ([58]) For special permutation σ ∈ S0n , if σ(ei − en ) > 0, then σ(ei+1 − en ) > 0.
Lemma 9.4. ([135]) We have I1 < · · · < Im (in the sense that all elements of Iν
are less than all elements of Iν+1 if 1 ≤ ν ≤ m − 1) and I10 > · · · > Im0 (in the same
sense).
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 256
Proof. Indeed, let A denote the set of indices i ∈ {1, . . . , n − 2} with σ(i + 1) =
σ(i) + 1. Then σ(i) − i is constant when we pass from any i ∈ A to i + 1, so each
set Iν is a connected interval that is contained in A except for its right end-point i0 ,
which satisfies σ(i0 + 1) < σ(i0 ), so that i0 + 1 belongs to an Iµ satisfying tµ < tν
and hence µ > ν. But then Iµ contains a point that is bigger than one of the points
of Iν and that has an image under σ that is smaller than the image of that point,
and since all of these sets are connected intervals this means that all of Iµ lies to
the right of all of Iν and that all of Iµ0 lies to the left of all of Iν0 .
σ(i)
k1
k2
..
.
kp
a
l1
..
.
lr
i
k1 k2 · · · kp l1 ··· lr n
n−a a−1
p
X r
X
Obviously, ki = n − a, ` j = a − 1, and p + r = m. Accordingly, for
i=1 j=1
later use, we introduce the numbers 0 = K0 < K1 < · · · < K p = n − a and
0 = L0 < L1 < · · · < Lr = a − 1 by
Ki = k1 + · · · + ki (1 ≤ i ≤ p), L j = l1 + · · · + l j (1 ≤ j ≤ r). (9.31)
We end this discussion on special permutations by determining the totality of
S0n . A direct computation shows |S02 | = 2, |S03 | = 5, |S04 | = 12. More generally,
we have the following:
In order for Lσ (s) to be different from the constant function zero, by the first
two relations of (9.33), there should be a complete cancellation of all factors hλ −
ρ, αi in the numerator of the first term in (9.34) that vanish at λ = λ s with either
(i) the factors hσλ s −ρ, βi appearing in the denominator of the first term in (9.34),
or else
(ii) the poles at λ = λ s of the factors b
ζ hλ, αi appearing in the numerator of the
second term in (9.34) for which hλ s , αi = 1.
ζ(hλ, αi j i)
Y Y b !
ζσ[<n] (s) :=
b 1−q −hλ−ρ,αk i
.
1≤k≤n−2 1≤i< j≤n−1 ζ(hλ, αi j i + 1)
b λ=λ s
σ(k)>σ(k+1) σ(i)>σ( j)
1 = # α ∈ Φ0 + : σα < 0, hρ, αi = k ,
X n o
mσ (k) :=
1≤i< j≤n−1
σ(i)>σ( j), j−i=k (9.39)
nσ (k) := mσ (k) − mσ (k − 1) , nσ (1) = mσ (1) = #Bσ .
Proof. To prove (2), we regroup the terms of (9.34) for special permutation σ ∈
S0n . First we cancel the terms in the numerator of the first factor in (9.34) for
α ∈ Aσ with the corresponding terms in the denominator for β = σα. The first
factor 1/Rσ (s) in (9.37) is the value at λ = λσ of the product of the remaining
terms β ∈ D r σAσ in this denominator. The second factor b ζσ[n] (s) in (9.37) is
the value at λ = λσ of the product of the terms in the second factor in (9.34) for
α < Φ0 + , i.e. α = ei − en > 0. The third factor b ζσ[<n] (s) in (9.37), which can also be
written as
ζ(hλ, αi)
Y Y b !
ζσ (s) =
b[<n]
(hλ − ρ, αi) · ,
α∈Bσ α∈Φ0 + ζ(hλ, αi + 1) λ=λ s
b
σ(α)<0
is obtained by collecting all the remaining zeta factors and rational factors appear-
ing in the numerator.
To prove (3), we note that the terms occurring in b ζσ[<n] (s) are of two types: for
ζ(hλ, αi j i)
b
α ∈ Bσ we must combine the quantities hλ−ρ, αk i and before taking
bζ(hλ, αi j i + 1)
the limit as λ → λ s because the first has a zero and the second has a pole, while
in the remaining zeta-quotients from the second term in (9.37), corresponding to
α ∈ Φ0 + r Bσ , we could simply substitute λ = λ s instead of taking a limit. Since
the residue of bζ(s) at s = 1 is 1, using the first equation in (9.28), we verify the
relation (3), whose right hand side, as one sees, does not depend on s at all.
In (3), the numbers nσ (k) in (9.38) may well be negative. To obtain the zeta
functions, we should remove them according to §8.4.2 by a certain normalization
process. For details, please see the next subsection.
In this subsection, following [135], we apply the results in §9.1.5 for special per-
mutations to calculate the rational factor Rσ (s) and the zeta factors b ζσ[n] (s) and
ζσ[<n] (s) appearing in Lemma 9.6(2) explicitly. We begin with Rσ (s).
b
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 261
Proof. By definition, Y Y
Rσ (s) = hσλ s − ρ, αk i = (hλ s , σ−1 αk i − 1).
1≤k≤n−1 1≤k≤n−1
σ−1 (αk )<∆0 σ−1 (αk )<∆0
For each k occurring in this product, write ei − e j =: αi j . Then the
σ−1 (αk ) =
condition αi j < ∆ says that the points (i, σ(i) = k) and ( j, σ( j) = k + 1) do not
0
belong to the same square block in Figure 9.1 of σ. Furthermore, we see that the
k’s occurring in the product, in decreasing order, together with the corresponding
values of i and j, are given by the first three colums of the following table
ζσ[n] (s).
We next consider the zeta factor b
Proof. This is much easier. From λ s = s$+ρ, we get hλ s , ei −en i = s+n−i. More-
over, from Figure 9.1, for the special permutation σ = σ(k1 , . . . , k p ; a; l1 , . . . , lr ),
we have
ei − en : 1 ≤ i < n, σ(i) > σ(n) = e1 − en , e2 − en , . . . , en−a − en .
n o n o
Before treating the zeta factor bζσ[<n] (s), recall that we need to make some nor-
malizations by clearing out the zeta factors appeared in the denominators. For
this, introduce
# Bσ if k = 1,
rσ (k) :=
nσ (k) − n(k) if k ≥ 2,
where, in analogy with the numbers nσ (k) (equation (9.39), we have set
n o
m(k) = # α > 0 : hρ, αi = k and n(k) = m(k) − m(k − 1) . (9.40)
Clearly m(k) = n − k for 1 ≤ k ≤ n and n(k) = −1 for 2 ≤ k ≤ n. Moreover, by
Lemma 9.6(3),
Y Y Y
bζσ[<n] (s) · ζ(i)−n(i) = b
b ζ(1)#Bσ ζ(i)nσ (i)−n(i) =
b ζ(i)rσ (i) .
b (9.41)
i≥2 i≥2 i≥1
Y
We claim that ζσ[<n] (s)
b · ζ(i)
b −n(i)
would do this job. Indeed, if we set, for k ∈
i≥2
Z≥1 , the Siegel number b
vk by
k
Y
vk :=
b ζ(`),
b (9.42)
`=1
In particular, rσ (k) ≥ 0.
= 1 + # α ∈ Φ0 + : σα > 0, hρ, αi = k ,
n o
i < j ≤ n − 1 and σ(i) < σ( j)) if and only if i and j belong to the same block,
say Iµ for some µ, associated to σ(k1 , . . . , k p ; a; l1 , . . . , lr ), and also σ( j) ∈ Iµ (or
equivalently j + 1 ∈ Iµ ), since otherwise σ(αi j ) < 0.
Denote by (m(k) − mσ (k))µ (resp. rσ,µ (k)) the contribution to m(k) − mσ (k)
(resp. to rσ (k)) of the block Iµ . With the discussion above, we have
X X
m(k) − mσ (k) = (m(k) − mσ (k))µ and rσ (k) = rσ,µ (k).
µ µ
Fix some µ and let Iµ := a+1, a+2, . . . , a+b with a, b ∈ Z>0 . Clearly, when
k = 1, rσ,µ (1) = # (a + b − 1, a + b) = 1, since, for other (i, i + 1), σ(i) < σ(i + 1).
Moreover, when k ≥ 2, by (9.27) and the characterization of the graph again,
n o
(m(k) − mσ (k))µ = # (i, j) : i, j + 1 ∈ Iµ , i < j, j = i + k
n o
= # (i, j) : a + 1 ≤ i < j < a + b, j = i + k .
Note that, for each fixed i (with a + 1 ≤ i < a + b),
n o 1 i + k < a + b
# (i, j) : a + 1 ≤ i < j < a + b, j = i + k = .
0 i + k ≥ a + b
Hence, (m(k) − mσ (k))µ = b − (k + 1). This implies that, for all k ≥ 1, we have
rσ,µ (k) = (m(k − 1) − mσ (k − 1))µ − (m(k) − mσ (k))µ = 1. Consequently,
Y
ζ(k)rσ,µ (k) = b
b ζ(1) b
ζ(2) · · · b
ζ(b) .
i≥1
(1) We have
SL /P SL /P
ζF n n−1,1 (s) = b
b ζF,on n−1,1 (ns − n). (9.46)
Here
n
SL /P
X X X
ζF,on n−1,1 (s) =
b ζF (s+a) (9.47)
(−1) p+r−1Rk1 ,...,k p ;a;l1 ,...,lr (s) b
a=1 k1 ,...,k p >0 l1 ,...,lr >0
k1 +···+k p =n−a l1 +···+lr =a−1
with
vk1 . . .b
b vk p 1
Rk1 ,...,k p ;a;l1 ,...,lr (s) := ·
(k1 + k2 ) . . . (k p−1 + k p ) s + a + k p
(9.48)
1 vl1 . . .b
v lr
,
b
× ·
−s − a + 1 + l1 (l1 + l2 ) . . . (lr−1 + lr )
k
Y
ζF (1) = cF , and b
v1 := Res s=1b
where b vk = ζF (i) (k ≥ 2).
b
i=1
(2) (Functional Equation) We have
S L /P S L /P
ζQ,on n−1,1 (−n − s) = b
b ζQ,on n−1,1 (s), (9.49)
or equivalently,
S L /P S L /P
ζQ n n−1,1 (1 − s) = b
b ζQ n n−1,1 (s). (9.50)
SL /P SL /P
Proof. The formula for b ζQ,on n−1,1 (s) and hence for bζF,on n−1,1 (s), in (1) is a sim-
ple combination of Lemmas 9.6, 9.7, 9.8, 9.9 and 9.10. With the formula (9.47)
obtained, the functional equation (9.49) in (2) comes directly from the classical
ζ(1 − s) = b
relation b ζ(s). This also justifies (9.46) in (1) and the functional equation
(9.50) in (2).
An analogue of (1) for the function field is founded in [135] jointly with Za-
gier, and (2) is first jointly proved with Kim in [58] with a different method. In the
next chapter, we follow Komori [59] to prove the functional equation for the zeta
functions associated to general reductive groups G.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 265
Chapter 10
Functional Equation
In §8.4.2, for connected reductive groups G defined over number fields F and
their maximal parabolic subgroups P, the zeta functions b ζFG/P (s) are introduced.
In this chapter, we shall establish functional equations for these zeta functions
following [59]. As mentioned earlier, in the case (G, P) = (SLn , Pn−1,1 ), the same
result was first proved by Kim and myself jointly in [58].
10.1 Preparations
The first step is to write down explicitly the zeta factors appeared in normalization
process in §8.4.2 to define the zeta function for (G, P)/F. We continuously use
the notations in the first section in §9.1.1.
10.1.1 Statement
We start with some notational preparations. For a pair of integer (k, h) ∈ Z2 , and a
special Weyl element w ∈ W 0 , set
n o
n p,w (k, h) := # α ∈ w−1 (Φ− ) : hλ p , α∨ i = k, hρ, α∨ i = h
n o (10.1)
M p (k, h) := max n p,w (k, h − 1) − n p,w (k, h) .
w∈W 0
Obviously, n p,w (k, h) = 0 for all but finitely many pairs (k, h), and the same holds
true for M p (k, h). Since Φ = Φ+ t Φ− , when k, h ≥ 1,
265
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 266
W ⊂ Aut(Φ) and Aut(Γ) ⊂ Aut(Φ) in a natural way. Indeed, from the simple
transitivity of W on positive roots and the fact that for γ ∈ Aut(Γ), γ(∆) = ∆, we
have Aut(Φ) = Aut(Γ) n W.
Finally, introduce the constant
cG,P := cP := c p := 2 hλ p − ρ p , α∨p i ∈ Z>0 . (10.3)
With all this, following [59], we are ready to state the normalization factor and
the functional equation of zeta function associated to (G, P) over F.
ζF;o
G/P
− c p − s; w0 (T ) = bζF;o
G/P
(s; T ) = b
ζF;o
G/Q
(s; γ(T )).
b
and
I(G/P)
Y Y∞ Y
∞
ζF aG/P
b
i s + bG/P
i = ζF (ks + h) M p (k,h) ,
b
i=1 k=1 h=2
J(G/P)
Y Y∞
ζF cG/P
b
j = ζF (ks + h) M p (0,h) .
b
j=1 h=2
In the sequel till the end of this chapter, we give a proof of Theorem 10.1. We start
with two (groups of) elementary relations on the Lie theoretic structures involved.
This result will be used in the proof of micro functional equations in §10.2.1.
and
X X
w p (ρ) = ρ − α=ρ− α = ρ − 2ρ p . (10.7)
α∈Φw p α∈Φ+p
On the other hand, since wk (Φ+p ) = Φ+p r {αk } ∪ {−αk } for k , p, we have then
wk (ρ p ) = ρ p − αk = ρ p − hρ p , α∨k iαk , and hence hρ p , α∨k i = 1. Therefore,
n
X n
X
ρp = hρ p , α∨k iλk = λk + hρ p , α∨p iλ p = ρ + hρ p − λ p , α∨p iλ p .
k=1 k=1,,p
To state the second group of relations, similar to n p,w , we introduce the count-
ing functions n p (k, h) for (k, h) ∈ Z2 by
n o
n p (k, h) := # α ∈ Φ : hλ p , α∨ i = k, hρ, α∨ i = h . (10.8)
Since Φ = Φ+ t Φ− , when k, h ≥ 1,
n p (k, h) := # α ∈ Φ+ : hλ p , α∨ i = k, hρ, α∨ i = h .
n o
(10.9)
These counting functions can be understood through a special generating function.
Indeed, if we define the character of the dual Lie algebra, by omitting the Cartan
sub-algebra contributions, by
X ∨
X(v) := eα (v) ∀v ∈ V, (10.10)
α∈Φ
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 268
X X
ehλ p ,α it+hρ,α i =
∨ ∨
X(tλ p + ρ) = n p (k, h)ekt+h .
α∈Φ k,h=−∞
Lemma 10.2. Let w0 , resp. w p , be the longest element of W, resp. of W p , and for
γ ∈ Aut(Γ), set q ∈ γ(p). Then, for (k, h) ∈ Z2 , we have
Proof. All these can be deduced from some manuuplications of the generating
series X(v). We start with
Obviously, by Proposition 9.1, since H p (s) is a product runs through all α ∈ Φ− , all
zeta factors appeared in the denominators of the Aw (s)’s in ωG/PF (s; T ) are cleared
out after multiplying by H p (s). On the other hand, as to be seen later, H p (s) is
an overdone factor, since it introduce some extra zeta function factors. Despite of
this, the function Z p (s) is very nice, since we can prove a functional equation for
it. To state it, we introduce an element γ0 ∈ Aut(Γ) characterized as follows: For
the longest element w0 of W, we have w0 (∆) = −∆. Hence, there is an element
γ0 ∈ Aut(Γ) such that −Id = γ0 w0 . Obviously, γ02 = Id.
= (w−1 (Φ− )) r ∆ p .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 270
we have
X Y 1
ehw(sλ p +ρ)−ρ,T i cF w p
|Φ ∩∆ |
Z p (s; T ) =
hλ p , α∨ is + hρ, α∨ i − 1
w∈W 0 α∈(w (∆))r∆P
−1
Y
ζF hλ p , α is + hρ, α i
∨ ∨
× b
α∈(w−1 (Φ− ))r∆ p
X r
= eh p,w (s;T ) cF f p,w (s)g p,w (s),
p,w (10.17)
w∈W 0
Here in the last step, we have used w0 (ei − e j ) = ei j− ei . Therefore r p,w0 ww p = r p,w ,
the first equality.
Similarly, by definition,
r p,γ−1 wγ = |Φγ−1 wγ ∩ ∆ p | = # α ∈ ∆P : (γ−1 wγ)(α) < 0
= # β ∈ ∆q : w(β) < 0 ,
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 271
since γ(Φ± ) = Φ± , γ(∆) = ∆ and γ(∆ p ) = ∆ p . This gives the second equality
r p,γ−1 wγ = rq,w .
We next treat f p,w and g p,w . Fix w ∈ W and γ ∈ Aut(Γ). For a subset A ⊂ Φ
with A = ∆ or Φ− , we set
S p;w (A; s) := hλ p , α∨ is + hρ, α∨ i : α ∈ w−1 (A) r ∆ p
Therefore, we have
S p,w (A; −c p − s)
= hλ p , α∨ i(−c p − s) + hρ, α∨ i : α ∈ w−1 (A) r ∆ p
= h−w(λ p )s + w(−c p λ p + ρ) − ρ, γ0 (T )i
= h−(ww p )(λ p )s − (ww p )(ρ) − ρ, γ0 (T )i
= h(w0 ww p )(λ p s + ρ) − ρ, T i
= h p,w0 ww p (s, T ).
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 272
Here, in the 3rd, resp. the 4th, equality, we have used the relation w p (ρ) = c p λ p −ρ
in Lemma 10.1, resp. the relation γ0 w0 = −Id and the fact that h·, ·i is W-invariant.
Similarly, for the sixth equality, we have
With all these, it is very easy to prove the functional equations for Z p . Indeed,
by (10.17), we have
X h p,w −c p −s;γ0 (T ) r p,w
Z p − c p − s; γ0 (T ) = e cF f p,w (−c p − s)g p,w (−c p − s)
w∈W 0
X r p,w0 ww p
= eh p,w0 ww p s;T
cF f p,w0 ww p (s)g p,w0 ww p (s)
w∈W
w(∆ p )⊂(∆∪Φ− )
X
r
= eh p,v s;T
cFp,v f p,v (s)g p,v (s)
v∈W
(vw−1 −
p )(∆ p )⊂w0 (∆∪Φ )
= Z p (s; T ).
Here, in the second to the last, resp, the last step, we have used Lemma 10.2, resp.
the facts that w0 (A) = −A for A = ∆ or Φ± and that w p (∆ p ) = −∆ p .
Similarly,
X r
Z p s; γ(T ) = ehq,w (s;γ(T )) cFq,w fq,w (s)gq,w (s)
w∈W
w(∆q )⊂(∆∪Φ− )
X r
=
−1 wγ
eh p,γ−1 wγ (s;T ) cFp,γ f p,γ−1 wγ (s)g p,γ−1 wγ (s)
w∈W
∆q ⊂w−1 (∆∪Φ− )
X r
= eh p,v (s;T ) cFp,v f p,v (s)g p,v (s)
v∈W
∆q ⊂(γv−1 γ−1 )(∆∪Φ− )
= Z p (s; T ).
Here, in the last step, we have used the facts that γ(A) = A for A = ∆ or Φ± , and
that γ(∆ p ) = ∆ p .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 273
10.2.2 Normalization
As mentioned earlier, even there is no zeta factor left in the denominator of each
w term for Z p (s; T ), it is overdone. In order to make a correction so as to arrive at
the zeta function for (G, P) over F, we need to single ut the extra factors used. For
this, following [59], we divide out extra zeta function factors appeared in Z p (s; T ).
For a fixed w ∈ W 0 , let
Y Y 1
H p,w (s) := ζF (hλ p , α is + hρ, α i)
∨ ∨
b
α∈Φw ζF (hλ p , α is + hρ, α i + 1)
∨ ∨
α∈Φw r∆ p
b
be the zeta factors appeared in (9.13) corresponding to w. In terms of pairing
hλ p , α∨ i, hρ, α∨ i = (k, h), and hence n p,w (k, h), we have
Y Y
ζF (hλ p , α∨ is + hρ, α∨ i) = b
b ζF (s + 1)n p,w (1,1) ζF (hλ p , α∨ is + hρ, α∨ i)
b
α∈Φw r∆ p α∈Φw r∆
∞ Y
Y ∞
=b
ζF (s + 1)n p,w (1,1) ζF (ks + h)n p,w (k,h) ,
b
k=0 h=2
Y ∞ Y
Y ∞
ζF (hλ p , α∨ is + hρ, α∨ + 1i) =
b ζF (ks + h + 1)n p,w (k,h)
b
α∈Φw k=0 h=1
Y∞ Y ∞
= ζF (ks + h)n p,w (k,h−1) .
b
k=0 h=2
This implies that
∞ Y
Y ∞
H p,w (s) = b
ζF (s + 1)n p,w (1,1) ζF (ks + h)n p,w (k,h)−n p,w (k,h−1) .
b (10.20)
k=0 h=2
Clearly, when n p,w (k, h) − n p,w (k, h − 1) < 0, the corresponding zeta function factor
should be understood as terms in denominator of H p,w (s). To get ride of these
a a > 0
factors, we introduce a test function δ(a) := , a (non-negative)
0 otherwise
counting function
n o
N p (k, h) := max δ n p,w (k, h − 1) − n p,w (k, h) (k, h) ∈ Z2 (10.21)
w∈W 0
and an auxiliary function
∞ Y
Y ∞
DF,p (s) := D p (s) := ζF (ks + h)n p (k,h−1)−N p (k,h) .
b (10.22)
k=0 h=2
∞ Y ∞
H p (s) Y
Consequently, = ζF (ks + h)N p (k,h) is the minimal zeta factors
b
D p (s) k=0 h=2
which should be used in the normalization process of §8.4.2 so as to get the zeta
ζFG/P (s).
function b
Hence, n p,Id (k, h − 1) − n p,Id (k, h) = 0. Consequently, in N p (k, h), δ can be removed
without changing the outcome.
In this subsection, we prove the functional equation. Recall that, by (10.14) and
(10.25), we have
!
H p (s) G/P 1
ζQ;o
bG/P
(s; T ) = ω (s; T ) = Z p (s; T ). (10.26)
D p (s) F D p (s)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 275
Proof. The last equality comes directly form the definition. To prove the first, we
decompose D p (s) into two parts, one consisting of all special zeta values and the
other for the rest. That is,
D p (s) = D0p D1p (s), (10.27)
where
∞
Y
D0p := bζ(h)n p (0,h−1)−M p (0,h) ,
h=2
∞ Y
∞
(10.28)
Y
D1p (s) := ζF (ks + h)
b n p (k,h−1)−M p (k,h)
.
k=1 h=2
Since, for k ≥ 1 and h ≤ 1, n p,w (k, h − 1) = 0 and M p (k, h) = 0, we have
Y∞ Y ∞
D1p (s) = bζF (ks + h)n p (k,h−1)−M p (k,h) . (10.29)
k=1 h=−∞
Hence to complete our proof, it suffices to show that
D1p (−c p − s) = D1p (s). (10.30)
For this, we need to treat counting functions M p (k, h) and n p (k, h).
(2) By the second equality in (10.1), it suffices to prove the equality for
n p,w (k, h). But this is a direct consequence of the third equality of Lemma 10.2
and the fact that ∆ p ⊂ w−1 (∆ ∪ Φ− ) if and only if ∆q ⊂ (γw−1 γ−1 )(∆ ∪ Φ− ),
which itself can be deduced from trivial relations γ(Φ± ) = Φ± , γ(∆) = ∆ and
γ(∆ p ) = ∆ p .
Consequently,
∞ Y
Y ∞
D1p (−c p − s) = ζF (−kc p − ks + h)n p (k,h−1)−M p (k,h)
b
k=1 h=−∞
∞ Y
Y ∞
= ζF (kc p + ks − h + 1)n p (k,h−1)−M p (k,h)
b
k=1 h=−∞
ζF (1 − s) = b
(by the functional equation b ζF (s))
∞
Y ∞
Y
= ζF (ks + h)n p (k,kc p −h)−M p (k,kc p −h+1)
b
k=1 h=−∞
∞ Y
Y ∞
= ζF (ks + h)n p (k,h−1)−M p (k,h)
b (by Lemma 10.5) = D1p (s).
k=1 h=−∞
To end this chapter, following [56], for each pair (G, P), we give the exact value
of the constant cG,P .
Lemma 10.6. (see e.g., Appendix 2 of [56]) Let G be a semi-simple Lie group.
Let α p be the simple root of G corresponding to the maximal parabolic group P.
Then
(A) cSLn ,P = n,
2n − p
p,n
=
(B) cSO2n ,P (n ≥ 2),
2n
p=n
(C) cSp ,P = 2n − p + 1 (n ≥ 3),
n
2n − p − 1
p , n − 1, n
=
(D) cSO2n+1 ,P (n ≥ 4),
2n − 2
p = n − 1, n
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 277
p=1
29
p=1
17
p=1 p=2
12
19
p=2
11
p=2 p=3
9 14
p=3
8
p=3 p=4
7 11
= = p=4 =
(E) cE6 ,P cE7 ,P cE8 ,P
10
9 p=4
9 p=5
p=5
13
p=5 p=6
12
13
p=6
18
p = 6, p=7
11
23
p = 7,
14
p = 8,
17
11 p = 1
p=2
7
(F) cF4 ,P =
5 p=3
p = 4,
8
5 p = long
(G) cG2 ,P =
3 p = short.
This can be verified by writing down the room system in concrete terms.
We omit the details of the calculations, but offer the following numbering of the
Dynkin diagrams.
residue of ωSL
Q
3 ;T
(λ) along z2 − z3 = 1 and making a normalization, we get
SL /P ;T 1 b
ζQ,o3 2,1 (t) := −
b ζ(2) · b
ζ(3s + 1) · e−3sx+x+2y + 0
3s + 3
1 1 1 1
− ζ(3s + 3) · e(3s+3)(x+y) +
·b ζ(3s + 1) · e−3sx
·b
2 3s + 1 2 3s + 2
1 1 1
− ζ(3s + 2) · e3sx+3sy+4x+2y + b
·b ζ(2)b
ζ(3s + 3)e−3sy+x−y .
3s 3s + 3 3s
Here we write z3 = s and hence z2 = s + 1, z1 = −2s − 1 Clearly, there is no
functional equation for these two functions. However, if we set y = 0 in T =
(x, y, −x − y) so that T = (x, 0, −x) ∈ C · ρ sits on the line spanned by ρ, then
SL /P ;xρ 1 1 1 b
bζQ,o3 1,2 (t) = b ζ(2)b
ζ(3t + 3)e3tx+4x − ζ(3t + 3)e(3t+3)x
3t 2 3t + 1
1 1 b 1 b b 1 1 b
+ ζ(3t + 1)e−3tx − ζ(2)ζ(3t + 1)e x − ζ(3t + 2)e−3tx+x ,
2 3t + 2 3t + 3 3t 3t + 3
SL /P ;xρ 1 b b 1 1 b
bζQ,o3 2,1 (t) = − ζ(2)ζ(3s + 1)e−3sx+x − ζ(3s + 3)e(3s+3)x
3s + 3 2 3s + 1
1 1 b 1 1 b 1
+ ζ(3s + 1)e−3sx − ζ(3s + 2)e3sx+4x + b ζ(2)bζ(3s + 3)e x .
2 3s + 2 3s 3s + 3 3s
It is rather easy to verify that these two functions satisfy the functional equation
SL3 /P1,2 ;xρ SL3 /P2,1 ;xρ
ωQ (−1 − s) = ωQ (x). (10.31)
To get the standard function equation, we introduce
SL /P 1 b b 1
ζQ;T3 1,2 (s) :=
b ζ(2)ζ(3s)T 3s+1 − b ζ(2)b
ζ(3s − 2) · T
3s − 3 3s
1 1 b 1 1 b 1 1b
− ζ(3s)T 3s + ζ(3s − 2)T −3s+3 − ζ(3t − 1)T −3s+4 ,
2 3s − 2 2 3s − 1 3s − 3 3s
SL )/P 1 1 b b
ζQ;T3 2,1 (t) := − b
b ζ(2)b
ζ(3s − 2) · T −3s+4 + ζ(2)ζ(3s)T
3s 3s − 3
1 1 b 1 1 b 1 1b
− ζ(3s)T 3s + ζ(3s − 2)T −3s+3 − ζ(3s − 1)T 3s+1 .
2 3s − 2 2 3s − 1 3s − 3 3s
Then the functional equation above becomes
SL )/P SL )/P
ζQ;T3 1,2 (1 − s) = b
b ζQ;T3 2,1 (s). (10.32)
On the other hand, if T , 0, it is easy to check with examples that there are zeros
of these functions off the central line, numerically. Since T = 0 corresponds to
semi-stable condition as we will see in Part 6, in some sense, we can say that the
Riemann hypothesis comes from the stability condition.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 279
PART 5
The Riemann hypothesis for the 10 concrete zeta functions of SLn (n = 2, 3, 4, 5),
Sp4 and G2 over Q are proved by Suzuki-Lagarias, Suzuki and Ki. The common
features are the meromorphic decompositions M(s)+M(−c−s) for these zeta func-
tions determined by the refined symmetries, good controls for the zero free regions
of M(s), and an elementary but useful lemma of Ki. Following this strategy, the
author proves a weak Riemann hypothesis for zeta functions of (SLn , Pn−1,1 )/Q.
More generally, Ki-Komori-Suzuki show that a weak Riemann hypothesis would
hold for the zeta functions of Chevalley groups over Q, if their discriminants are
not vanished. In this part, we first give a new proof for the case (SLn , Pn−1,1 )/Q us-
ing explicit formulas for the related zeta functions obtained in Part 4, then present
the works of Ki-Komori-Suzuki on zeta functions of Chevalley groups over Q.
Their work is very technical, because of the complications of the algebraic and
analytic structures involved. We end this part with our confirmation of the weak
Riemann hypothesis for the zeta functions of the type E exceptional groups over
Q, by verifying that the associated discriminants are positive using Mathematica.
279
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 281
Chapter 11
SL n/P n−1,1
ζ
Conditional Weak RH for b (s)
Q
SL /P
ζQ n n−1,1(s), we first introduce its discriminant, then show that, if this discrim-
For b
inant is not zero, a weak Riemann hypothesis would hold.
281
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 282
Therefore, for n ≥ 1,
n n−2
X vi X X vn1 · · ·b vnk
Bn (s) = (−1)k
b b
i=1
s − i k=1 n1 ,...,nk >0
(n1 + n2 ) · · · (nk−1 + nk )
n1 +...+nk =n−i
!
1
×
n − (n1 + . . . + nk−1 )
n−1
n−2
X vn−i X X vn1 · · ·b vnk 1
= (−1)k .
b b
i=0
s − n + i k=1 n1 ,...,nk >0
(n1 + n2 ) · · · (nk−1 + nk ) nk + n − i
n1 +...+nk =i
ζ(s + a),
In parallel, for the remaining zeta factors b
SLn /Pn−1,1
ζQ
Conditional Weak RH for b (s) 283
SLn /Pn−1,1
ξQ
Definition 11.1. We define the xi-function b (s) of (SLn , Pn−1,1 ) over Q by
n !
SL /P SL /P
Y
ξQ n n−1,1(s) :=
b (s + h) bζQ n n−1,1(s). (11.6)
h=0
Definition 11.2.
n+1
≥ SL /P
ζQ 2 (s), called the upper half of b
(1) We define a function b ζQ n n−1,1(s), by
n−a X
a−1
≥ n+1 X0 X −bn−a,k b
vkba−1,i b
vi b
ζQ 2 (s) :=
b ζ(s + a) (11.7)
a≥ n+1
2
k=1 i=1
s+a+k s+a−1−i
where
P
X0
a≥ n+1
n even
.
2
:= (11.8)
a≥ n+1
2 Pa≥ n+1 − 1 Pa= n+1
n odd
2 2 2
n+1
≤ SL /P
ζQ 2 (s), called the lower half of b
(2) We define a function b ζQ n n−1,1(s), by
n−a X
a−1
≤ n+1 X0 X −bn−a,k b
vkba−1,i b
vi b
ζQ 2 (s) :=
b ζ(s + a) (11.9)
a≤ n+1
2
k=1 i=1
s+a+k s+a−1−i
where
P
X0
a≤ n+1
n even
.
2
:= (11.10)
a≤ n+1
2 Pa≤ n+1 − 1 Pa= n+1
n odd
2 2 2
n+1 n+1
≥ ≤
ξQ 2 (s) and b
(3) We introduce the functions b ξQ 2 (s) by
n ! n !
n+1
≥ n+1 n+1
≥ n+1
≥
Y ≤
Y
ξQ 2 (s) :=
b (s + h) bζQ 2 (s) ξQ 2 (s) :=
and b (s + h) bζQ 2 (s).
h=0 h=0
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 284
Definition 11.3. For all 1 ≤ a ≤ n, we define the partial zeta functions bζQ(a) (s) and
the rational functions ra (s) by
n−a
X −bn−a,k b vk
a=1
s+a+k
k=1
ζQ(a) (s)
n−a a−1
X X −bn−a,k b vk ba−1,i bvi
b
1 < a < n . (11.15)
:= ra (s) :=
ζ(s + a)
b
k=1 i=1
s + a + k s + a − 1 −i
n−1
−bn−1,i b
vi
X
a=n
s+n−1−i
i=1
Accordingly, we define the partial xi-function b ξQ(a) (s) and the polynomial pa (s) by
n
Y
ξQ(a) (s) := (s + h) · b
(a)
b ζQ (s) =: pa (s) ξ(s + a) 1≤a≤n (11.16)
h=0
SLn /Pn−1,1
ζQ
Conditional Weak RH for b (s) 285
Definition 11.4. We define the invariants Bk (1 ≤ k ≤ n), also called the discrim-
SL /P
ζQ n n−1,1(s), by
inants of b
k
X X vk1 · · ·b vk p
Bk = (−1) p−1 .
b
(11.18)
p=1 k1 ,...,b p >0
(k1 + k2 ) · · · (k p−1 + k p )
k1 +...+k p =k
In addition, set B0 = 1.
Proposition 11.2. If Bn−1 is not zero, the polynomial pn (s) dominates all the
polynomials pa (s) with a ≥ n+1
2 . Moreover, if Bn−a Ba−1 , 0,
n−1 a=n
deg pa (s)) = n−2 n−1≥a≥2.
(11.19)
n − 1 a = 1
Proof. Denote by d(a) the right hand side of (11.19), and let La be the coefficient
of sd(a) in pa (s). By definition,
n
Y 1
pn (s) = (s + h) · b
(n)
ζQ (s)
h=0
ξ(s + n)
n
Y 1
= (s + h) b ζ(s + n)Bn−1 (s + n − 1)
h=0
ξ(s + n) (11.20)
n−2
n−1
Y X X vk1 · · ·bvk p 1
= (s+h) (−1) .
p−1
b
+k +k +
h=0
p=1
k1 ,...,k p >0
(k1 2 ) · · · (k p−1 p ) s n−1−k p
k1 +...+k p =n−1
n
Y 1
pa (s) = (s + h) · b
(a)
ζQ (s)
h=0
ξ(s + a)
n
Y 1
= (s + h) Bn−a (−s − a)b
ζ(s + a)Ba−1 (s + a − 1)
h=0
ξ(s + a)
n
n−a
Y X X v k · · ·b
v k 1
(11.22)
= (s+h)
(−1) p−1 b 1 p
+k +k
h=0
p=1 k1 ,...,k p >0
(k1 2 ) · · · (k p−1 p ) −s−a+k p
k1 +...+k p =n−a
a−1
v ` · · ·b
v` 1 ζ(s + a) .
X X b
1 q
× (−1)q−1
b
q=1 `1 ,...,`q >0
(`1 + `2 ) · · · (`q−1 + `q ) s + a − 1 − `q ξ(s + a)
`1 +...+`q =a−1
ζ(s + a) 1
=
b
Since , we have d(a) ≤ n − 2 and
ξ(s + a) (s + a)(s + a − 1)
n−a
X X vk1 · · ·b vk p
La = − (−1) p−1
b
p=1 k1 ,...,k p >0
(k1 + k2 ) · · · (k p−1 + k p )
k1 +...+k p =n−a
a−1
(11.23)
X X v`1 · · ·b v`q
× (−1)q−1
b
q=1 `1 ,...,`q >0
(`1 + `2 ) · · · (`q−1 + `q )
`1 +...+`q =a−1
= − Bn−a Ba−1 .
Hence d(a) = n − 2 if Bn−a Ba−1 , 0. This proves the proposition for a ≥ n+1
2 , and
hence all the cases by the functional equation.
ξQ(a) (s)
b pa (s) ξ(s + a) ra (s) ζ(s + a) ζQ(a) (s)
b
= = = (n) .
b b
(11.24)
ξQ(n) (s)
b pn (s) ξ(s + n)
b r n (s) ζ(s + n)
b ζQ (s)
b
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 287
SLn /Pn−1,1
ζQ
Conditional Weak RH for b (s) 287
Moreover,
n−a
ra (s) X X vk1 · · ·b vk p 1
= (−1) p−1
b
rn (s) p=1 k1 ,...,b p >0
(k1 + k2 ) · · · (k p−1 + k p ) s + a + k p
k1 +...+k p =n−a
a−1
X X v`1 · · ·b v`r 1
× (−1)r−1
b
r=1 `1 ,...,`r >0
(`1 + `2 ) · · · (`r−1 + `r ) s + a − 1 − `r
`1 +...+`r =a−1
−1
n−1
X
k−1
X bvn1 · · ·b vnk 1
× (−1)
(n1 + n2 ) · · · (nk−1 + nk ) s − nk
n1 ,...,nk >0
k=1
n1 +...+nk =n−1
n−a (11.25)
1 X X vk1 · · ·b vk p
= (−1) p−1
b
s p=1 k1 ,...,b p >0
(k1 + k2 ) · · · (k p−1 + k p )
k1 +...+k p =n−a
a−1
X X v`1 · · ·b v`r
× (−1)r−1
b
r=1 `1 ,...,`r >0
(`1 + `2 ) · · · (`r−1 + `r )
`1 +...+`r =a−1
n−1
X X vn1 · · ·b vnk
× (−1)k−1
b
k=1 n1 ,...,nk >0
(n1 + n2 ) · · · (nk−1 + nk )
n1 +...+nk =n−1
(1 + (a + k p )/s)(1 + (a − 1 − `r )/s)
!−1
.
×
1 − nk /s
ra (s)
Hence, is bounded when |s| → ∞, if Bn−1 , 0. On the other hand, for the
rn (s)
zeta factors, by Lemma 8.4,
ξ(s + a) ξ(s + a)
! !
a+n−1 a+n−1
< 1 <(s) > − = 1 <(s) = − .
b b
&
ξ(s + n)
b 2 ξ(s + n)
b 2
(11.26)
Thus, for all a ≥ n+12 ,
ξ(s + a) n − 1 + n+1 3n − 1
<1 <(s) > − =− .
b 2
(11.27)
bξ(s + n) 2 4
3n − 1 5 n
Nevertheless, if n = 2, − = − < −1 = − . Hence, (11.27) implies that
4 4 2
ξ(s + a) n
<1 <(s) > − .
b
(11.28)
ξ(s + n)
b 2
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 288
Proposition 11.3. If Bn−1 is not zero, there are at most finitely many zeros
≥ n+1 n
ξQ 2 (s) lying in the right half- plane <(s) > − for all n ≥ 2, and
of b
2
3n − 1
<(s) > − when n ≥ 3.
4
Proof. Indeed, by definition,
ζ (a) (s)
X0 n−1 b
n+1
≥
ξQ ξQ(n) (s) 1 +
(s) = b
2
b (11.29)
a= n+1
2 bζ (n) (s)
where
X0 n−1 b ζ (a) (s) X0 n−1 ra (s) (s + n) (s + n − 1) b ξ(s + a)
= . (11.30)
n+1
ζ (s)
a= 2 b(n) a= n+1
2 r n (s) (s + a) (s + a − 1) ξ(s + n)
b
Therefore, by (11.28), for <(s) > − n2 , we have
X0 n−1 b ζ (a) (s) X0 n−1 ra (s) (s + n) (s + n − 1) b ξ(s + a)
≤
n+1
ζ (s)
a= 2 b(n) a= n+1
2 r n (s) (s + a) (s + a − 1) ξ(s + n) (11.31)
b
X0 n−1 ra (s) (s + n) (s + n − 1)
≤ .
a= n+1
2 rn (s)(s + a)(s + a − 1)
Set now
X0 n−1 ra (s) (s + n) (s + n − 1)
φn (s) = (11.32)
a= n+1
2 rn (s)(s + a)(s + a − 1)
by (11.25) and the conclusion followed, we conclude that, if Bn−1 , 0, for
|s| → ∞,
X0 n−1 ra (s)(s+n)(s+n−1) 1 X0 n−1 Bn−a Ba−1 1
|φn (s)| ≤ = +O 2 . (11.33)
a= n+1
2 rn (s)(s+a)(s+a−1) s a= n+1
2 Bn s
Let then Dn ⊂ C be the set consisting of points s such that
X0 n−1 ra (s) (s + n) (s + n − 1)
1> ≥ φn (s) . (11.34)
a= n+1
2 rn (s) (s + a) (s + a − 1)
By (11.33), obviously, Dn is a finite subset. Therefore, for the second factor in
(11.29), we have, by (11.28) and (11.34), if Bn−1 , 0,
X0 n−1 b ζ (a) (s) n
1+ n+1
>0 ∀s < Dn and <(s) ≥ − . (11.35)
a= 2 b(n)
ζ (s) 4
As discussed above, when n ≥ 3, we may take s from the region s < Dn and
3n − 1
<(s) > − . Finally, let us treat the first factor of (11.29). Note that
4
ξQ(n) (s) = pn (s) b
b ξ(s + n) (11.36)
where pn (s) is a polynomial degree n−1 provided Bn−1 , 0 and ξ(s) = s(s+1)b ζ(s).
Therefore, to complete the proof, it suffices to note that ξ(s + n) , 0 in the region
b
<(s) > −n + 1 and that there are at most n − 1 zeros for pn (s).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 289
SLn /Pn−1,1
ζQ
Conditional Weak RH for b (s) 289
As usual, a slight better result can be obtained. Indeed, with the same structure
as in the proof above, by a known result stated at page 135 of [116] obtained
by Vinogradov’s method applying to the factors of the Riemann zeta function
involved, we conclude the following:
n+1
≥
ξQ 2 (s)
Corollary 11.1. If Bn−1 is not zero, there are at most finitely many zeros of b
n
lying in the region <(s) ≥ − − δ(t), where s = σ + it and σ(t) is a positive
2
function on the real line satisfying δ(t) log |t| → ∞ as |t| → ∞.
ζ(s + a)
b
We first treat the quotient . Since we are working on <(s) → −∞, we use
ζ(s + n)
b
ζ(s) to get
the functional equation for b
a−n Γ
−s−a+1
ζ(s + a) b ζ(−s − a + 1) ζ(−s − a + 1)
= =π .
b 2
2
−s−n+1 ζ(−s − n + 1)
(11.38)
ζ(s + n) ζ(−s − n + 1)
b b Γ 2
For the Γ factors, by Stirling’s formula, for arg(s) < π − ε,
√ s s 1 1 1 1 !
Γ(s) = 2πs 1+ + + . . . + + O . (11.39)
e 12s 288s2 am sm |s|m+1
So there are some real constants c j such that, for <(s) → −∞,
Γ −s−a+1 −s−a+1 −s−a+1
r
2 s+a−1( 2 ) 2 a−n
= e 2
+ −s−n+1 −s−n+1
Γ −s−n+1
2
s n − 1 ( ) 22
c1 (a, n) cm (a, n) 1 !
× 1+ + ... + + O m+1 (11.40)
s sm |s|
! a−n 1 !
2 2 c1 (a, n) cm (a, n)
= − 1+ + ... + + O
s s sm |s|m+1
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 290
−s − a + 1 π
!
since arg < < π − ε. Here, in above equalities, the c j ’s are
2 2
changed from line to line, with an abuse of notations. Next, we treat the quo-
ζ(−s − a − 1)
tient . Recall that for <(s) < 0,
ζ(−s − n + 1)
∞
X µ(k)
1
= , (11.41)
ζ(−s) k=1 k−s
where µ(k) denotes the Mobius numbers. So there exists real constants dk such that
ζ(−s − a + 1) X dk (a, n)
=1+ <(s) → −∞. (11.42)
ζ(−s − n + 1) k≥2
k−s
X0 n−1 ζ(s + a)
Consequently, the second factor pn (s) +
b
pa (s) in (11.37)
a= n+1
2 ζ(s + n)
b
becomes
X0 n−1 s n−a
2
pn (s)+ p (s) −
n+1 a
a= 2 2π
X dk (a, n) (11.44)
c1 (a, n) cm (a, n) 1 !
× 1+ + ... + + O m+1 1 +
−s
.
s sm |s| k≥2
k
1
Assume that Bn−a Ba−1 , 0. Then pa (s) is a polynomial in t = s 2 of degree 2n − 2
and 2n − 4 when a = n and a < n, respectively. Write
s n−a 2
qa (s) := pa (s) − . (11.45)
2π
Then
n−( n22 +1)
s
p 2 +1 (s) − 2π n even
n
q[ n2 +1] (s) =
n− n+1
2
(11.46)
p n+1 (s) − 2πs 2
n even,
2
1
and, as polynomials in z = s 2 ,
n
degz q[ 2n +1] (s) ≥ degt (qa (s)) + 1 ∀ + 1 < a ≤ n. (11.47)
2
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 291
SLn /Pn−1,1
ζQ
Conditional Weak RH for b (s) 291
X dk (a, n) ! qn (s)
! X dk ([ n + 1], n) !−1
× 1+ + 1+ 2
.
k≥2
k−s q[ n2 +1] (s) k≥2
k−s
Since, as <(s) → −∞, the polynomial q[ n2 +1] (s) is dominating and the series
X dk (a, n)
1+ are always strictly positive and uniformly bounded, there exists a
k≥2
k−s
bounded subset D0n ⊂ C such that
|ϕn (s)| < 1 and hence 1 + ϕa (s) > 0 ∀s < D0n . (11.50)
Therefore, by (11.37) and (11.48), it suffices to prove the non-vanishing statement
ζ(s + n) q[ n2 +1] (s) of a zeta function and a poly-
in the proposition for the product b
nomial. But this is rather obvious, since it is well-known that, for <(s) < −n, the
1
ζ(s + n) has no zero, and that for a fixed polynomial of s 2 , its zeros
zeta function b
are always contained in a bounded domain, depending on the coefficients.
≥ n+1
11.3 ξ
Refined Hadamard Product for b 2
(s)
Q
≥ n+1
ξ
11.3.1 Distributions of Zeros for b 2
(s)
Q
n
As usual, write s = σ + it, and let T > 1 and σ > . Denote by Nn (T ; σ) the
n+1
2
≥
number of zeros of bξQ 2 (s) in the region
n n o
− σ < <(s) < − and 0 < =(s) < T . (11.51)
2
Lemma 11.2. Assume that Bn−1 and B[ n2 ] B[ n−1 2 ]
are not zero. There exists real
constants c1 > 0, c2 and c3 such that, with the constant σn as in Proposition 11.4,
Nn (T ; +∞) = Nn (T ; σn ) =c1 (n) T log T + c2 (n) T + O(log T ). (11.52)
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 292
Proof. This can be proved rather directly using some well-known techniques
in analytic number theory, based on Proposition 11.3 for the zero free region of
≥ n+1 ≥ n+1
ξ 2 (s) on the right, Proposition 11.4 for the zero free region of b
b
Q ξ 2 (s) on the Q
left, and the relations (11.29), (11.35) and (11.48), (11.50) in the proofs of these
two propositions. Indeed, by the Cauchy residue formula,
Z
1 ≥ n+1
Nn (T ; σn ) = ξQ 2 (s)
d log b
2πi RT
(11.53)
1 ≥ n+1
= (∆1 + ∆2 + ∆3 + ∆4 ) arg ξQ (s)
b 2
2πi
where RT denotes a rectangle with vertices −σL + c i, σR + c i, σR + iT, −σL + iT ,
associated to some sufficient positive real numbers σL , σR and a certain small
enough positive constant c satisfying the conditions that
n+1
≥
ξQ
(i) there is no zeros of b 2
(s) on the boundary of RT , and
(ii) |φn (−σL + i t)| < 1 and |ϕn (σR + i t)| < 1 for |t| ≥ c
Lemma 11.3 (Lemma in §9.4 of [116]). Let σ0 > 0, T 0 and take 0 ≤ α <
β < σ0 , which may depend on T . Assume that f (s) is an analytic function, taking
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 293
SLn /Pn−1,1
ζQ
Conditional Weak RH for b (s) 293
real value on real line, and regular for σ ≥ α, except possibly finitely many poles
on the real line, and takes no zero on the line t = T . Then, for σ ≥ β,
!−1
σ0 − α 3
arg f (σ + iT ) ≤ π log m Mα,T +2 log + π,
(11.54)
σ0 − β 2
where m, Mσ,t are positive constants satisfying
m ≤ < f (σ + it) f (σ0 + it0 ) ≤ Mσ,t (∀σ0 ≥ σ, 1 ≤ t0 ≤ t). (11.55)
and
n+1
≥
ξ
11.3.2 Refined Hadamard Product for b 2
(s)
Q
≥ n+1
ξQ
There is a natural structure of Hadamard product for the function b 2
(s). To
explain this, set, for simplicity, that
n
≥ n+1 ≥ n+1
ξQ 2 (s) := b
e ξQ 2 − + i s . (11.56)
2
Proposition 11.5. Assume that Bn−1 and B[ n2 ] B[ n−1
2 ]
are not zero. The function
n+1
≥
ξQ
e 2
(s) admits a natural Hadamard product
n+1
(s) = κn eαn s Pn (s) Hn (s).
≥
ξQ
e 2
(11.57)
Here
where, for all m, <(ρm ) > 0 and 0 < δ(t) < =(ρm ) < cn with δ and cn are the
function and constant in Corollary 11.1 and Proposition 11.4, respectively.
n+1
≥
Proof. Since ξ(ks + h) are entire functions of order one, e
ξQ 2
(s) is an entire
≥ n+1
function of order at most one. Moreover, ξQ 2 (s)
since b takes real values over real
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 294
n+1
≥
axis, by the functional equation, if ρ is a zero of e
ξQ 2 (s), so is −ρ. Therefore, by
standard theory for entire functions of finite orders, using Propositions 11.3 and
≥ n+1
11.4, we conclude that eξ 2 (s) admits a Hadamard product in the form
Q
n+1
(s) = κn eβn s Pn (s) hn (s),
≥
ξQ
e 2
(11.59)
where κn , 0 and βn are complex numbers, Pn (z), is a non-zero polynomial having
no zero in =(z) > 0 except for purely imaginary zero, and
∞ !
Y z z z z
hn (z) = 1− 1+ exp − , (11.60)
m=1
ρm ρm ρm ρm
where <(ρm ) > 0 for all m. and the infinite products converge uniformly on every
compact subset in C. Moreover, if set ρm = am + ibm with am ∈ R and bm ∈ R>0 ,
∞ ∞ ∞
X 1 1 X bm X 1
− ≤2 ≤ 2 cn . (11.61)
m=1 m
ρ ρm |ρ
m=1 m
|2 |ρ
m=1 m
|2
Here in the last step we have used Proposition 11.4. Since the sums on the
right hand sides of (11.60) are all finite, we can and hence will move the fac-
z z
tors exp − out of the infinite product. This then implies the existence of
ρm ρm
structural product in the statement of the proposition, if we set
∞ !
X 1 1
αn = βn + − . (11.62)
m=1
ρm ρm
Hence, to complete the proof, it suffices to prove the following
Pn (−it)
Obviously, log = o(1) as t → + ∞. Moreover, by Propositions 11.3 and
Pn (it)
11.4, as t → +∞, we have 0 < bm < |cn | + 1, and hence
∞ ∞
∞
a2m + (t + bm )2 X
X 4bm t X t log m
log 2 ≤ = O = O(log t). (11.64)
m=1
a +(t−b )2
m m a2 +(t − b )2
m=1 m m2 +t2
m m=1
≥ n+1 ≥ n+1
So, log ξQ 2 (−it) e
e ξQ 2 (it) = 2t=(αn ) + O(log t) as t → +∞. Consequently, to
prove =(αn ) = 0 and hence αn ∈ R, it is sufficient to prove the following
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 295
SLn /Pn−1,1
ζQ
Conditional Weak RH for b (s) 295
ζ 2n + t
b (11.67)
= 1 + o(1)
ζ 1 − 2 + t q[ n2 +1] − 2 − t
n n
b
t n−1
2 1
= n 1 + o(1) .
2π q[ 2 +1] − 2 − t
n
Here, in the last step, we have used (11.43). This proves Lemma 11.5, and hence
Lemma 11.4 and Proposition 11.5.
SL n /P n−1,1
11.4 ζ
Conditional Weak Riemann Hypothesis for b (s)
Q
Proof. Otherwise, assume that, outside the real line, F = limn→∞ Fn has M(> 2N)
roots z1 , z2 , . . . , z` with multiplicities m1 , m2 , . . . , m` . Obviously, m1 + m2 + · · · +
m` = M. By Hurwitzs theorem, for each i = 1, 2, . . . , `, there exists R j > 0 such
that for any 0 < r j < R j , Fn has exactly m j roots in the disk |z − z j | < r j (0 <
r j < R j ) counting with multiplicities for all n ≥ N j , where N j may depend on r j .
Choose r > 0 such that 0 < r < R j and {z : |z − z j | ≤ r} are all away from the
real line, i.e. the closed disk |z − z j | ≤ r are entirely contained in the upper/lower
half plane for all 1 ≤ j ≤ `. Hence, for a sufficiently large n, Fn has exactly m j
zeros in the disk |z − z j | < r counting with multiplicities for all j. This implies
that, away from the real line, Fn admits at least M = m1 + m2 + · · · + m` zeros.
This is a contradiction.
Lemma 11.6 (Proposition 3.1 in [56]). Let W(z) be a function in C which admits
a Hadamard product of the form
∞ " ! !#
Y z z
W(z) = H(z)eαz 1− 1+ , (11.68)
k=1
ρk ρk
where H(z) is a nonzero polynomial, α ∈ R, =(ρk ) ≥ 0 for all k, and the product
in (11.68) is uniformly convergent in any compact subset of C. Assume that H(z)
has exactly N zeros, counted with multiplicities, in the lower half-plane =(z) < 0.
Then, away from the real line, there are at most 2N, counted with multiplicities
for the function W(z) + W(z), resp. W(z) − W(z)). In particular, there are at
most N zeros of W(z) + W(z) in the lower-half plane and in the upper-half plane,
respectively.
SLn /Pn−1,1
ζQ
Conditional Weak RH for b (s) 297
Claim 11.1. Let U(z) and V(z) be polynomials with real coefficients. Assume
that U(z) . 0, resp. V(z) . 0 and that there are exactly N zeros, counted with
multiplicities, of W(z) = U(z) + iV(z) in the lower half-plane =(z) < 0. Then,
outside the real line, there are at most 2N zeros, counted with multiplicities, for
U(z) and V(z), respectively.
Proof. We prove this following Lemma 2 at page 215 of [18]. We may assume
that U(z) and V(z) have no common real roots. This implies that W(z) admits
no real roots. We also assume that the degree m of W(z) satisfies m ≥ 2N, as
otherwise the result is an easy consequence of the relations 2U(z) = W(z) + W(z)
and 2iV(z) = W(z) − W(z).
By the argument principle and the assumption on the zeros of W(z), if z runs
over the real axis from −∞ to ∞, the argument of W(z) increases by (m − 2N)π.
Hence there are at least m − 2N different points on the real axis where W(z) be-
comes purely imaginary, resp. real. Hence, if z = ∞ is not one of these points,
there are m − 2N real roots for U(z), resp. V(z) at least, and otherwise there are
at least m − 2N − 1 real roots. But in the last case, the degree of U(z), resp. V(z)
must be < m. Therefore, in both cases, there are at most 2N non-real zeros for
U(z), resp. V(z).
Since U(z) and V(z) are polynomials with real coefficients, W(z) = U(z) −
iV(z). Hence,
2U(z) = W(z) + W(z) and 2iV(z) = W(z) − W(z). (11.69)
Conversely, if W(z) is an arbitrary polynomial, by (11.69), the coefficients of U(z)
and V(z) must be real. Hence, we may rewrite Claim 11.1 as follows:
Claim 11.2. Let W(z) be a polynomial satisfying W(z) , −W(z), resp. W(z) ,
−W(z). Assume that there are exactly N zeros, counted with multiplicities, of
W(z) has in the lower half-plane =(z) < 0. Then, outside the real line, there are at
most 2N zeros, counted with multiplicities, of W(z) + W(z), resp. W(z) − W(z).
Back to the proof of the lemma. Following page 2379 of [55], we introduce
the wn (z)’s by
n "
αz n Y
! !#
z z
wn (z) = H(z) 1 + 1− 1+ . (11.70)
n k=1 ρk ρk
Obviously, limn→∞ wn (z) = W(z) and wn (z) has exactly n zeros in the lower half-
plane =(z) < 0. By Claim 11.2, wn (z) + wn (z) has at most 2n zeros outside of the
real axes. Since wn (z) + wn (z) converges to W(z) + W(z) uniformly in any compact
subset of C, by Corollary 11.2, W(z) + W(z) has at most 2n zeros outside the real
axes. The same arguments work for W(z) − W(z).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 298
SL n /P n−1,1
ζ
11.4.2 Conditional Weak Riemann Hypothesis for b (s)
Q
fices to show that θ(t) is strictly increasing as t → +∞. By Propositions 11.2 and
n
11.3, on the central line <(s) = − , we have, as |s| → ∞
2
≥ n+1
ξQ 2 (s) = pn (s)b ζ(s + n) 1 + o(1) .
b (11.73)
By taking logarithmic derivative, as t → +∞, Gamma factors give log t and
ζ0
!
log t
(1 + it) = O = o(log t). (11.74)
ζ log log t
Therefore, there exists a suitable constant c such that
d
θ − c p /2 + it = c log t 1 + o(1) .
(11.75)
dt
So θ(t) is strictly increasing as t → +∞ as desired.
Theorem 11.3 (Theorem 15.2 and Proposition 15.1). Let n ≥ 1. For the moduli
space MQ,n [1] of semi-stable lattices of rank n and volume one,
n
X X vn1 · · ·b vnk
0 < vol MssQ,n [1] = Bn+1 = (−1)k−1 .
b
k=1 n ,...,n >0
(n1 + n 2 ) · · · (nk−1 + nk )
1 k
n1 +...+nk =n
We will prove this and hence complete our proof of the weak Riemann hy-
SL ,P
ζQ n n−1,1 (s) (n ≥ 2) in Chapter 14.
pothesis for the zeta function b
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 299
Chapter 12
The main purpose of this chapter is to prove a conditional weak Riemann hy-
pothesis for the zeta functions associated to reductive groups and their maximal
parabolic subgroups, essentially following [56]. We first expose some fine sym-
metric structures for the zeta functions associated to reductive groups and their
maximal parabolic subgroups, then, based on a detailed analysis of the Lie struc-
tures and some standard analytic number theoretic estimations involved, for the
Chevalley groups over Q, we show that, if the discriminants are not vanishing, all
but finitely many zeros of these zeta functions lie on the central line.
Let G be a reductive group over a number field F. Fix a minimal parabolic sub-
group of G and let P be a standard maximal parabolic subgroup of G. Denote by
(V, h·, ·i); Φ, Φ± , ∆; W; ρ the associated root system. That is to say, (V, h·, ·i) is
the metrized root space, and Φ, resp. Φ± , resp. ∆, is the set of roots, resp. the
subset of ±-roots, Xresp. the set of simple roots, W = hσα : α ∈ ∆i is the Weyl
group, and ρ = α is the Weyl vector. For each w ∈ W, set Φw := Φ+ ∩ w−1 Φ−
α>0
be the set of positive roots such that their w-images become negative, and denote
by w0 ∈ W the longest element in W. It is well known that the length `(w) of
w coincides with Φw , w20 = id, and w0 (∆) = −∆ and hence w0 (Φ± ) = Φ∓ . For
later use, write ∆ =: α1 , . . . , αn and let λ1 , . . . , λn be the set of fundamental
X n
dominant weights. Then, hλ j , α∨i i = δi j for all 1 ≤ i, j ≤ n, and ρ = λ j . Here
j=1
as usual, for α ∈ Φ, denote by α∨ the associated co-root.
299
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 300
(1) For w ∈ W 0 and (k, h) ∈ Z2 , the counting functions n p,w (k, h), n p (k, h) and
M p (k, h) are defined by
n p,w (k, h) := # α ∈ w−1 (Φ− ) : hλ p , α∨ i = k, hρ, α∨ i ,
n p (k, h) := # α ∈ Φ : hλ p , α∨ i = k, hρ, α∨ i ,
(12.2)
M p (k, h) := max n p,w (k, h − 1) − n p,w (k, h) .
w∈W 0
(2) The discriminant of (G, P) over F is defined by
X r
∆G/P
F := cFp,w c p,w d p,w . (12.3)
w∈W p
Here
W p := w ∈ W 0 : w(Φ+ r Φ+p ) ⊂ Φ+ ,
where ωG/P
F (s) is the period of (G, P) over F defined by
X Y 1 Y b ζ F (hλ, α ∨
i)
Res sn =0 . . . Res
[ s p =0 . . . Res s1 =0
.
wλ − ρ, α i σ∈Φ b
∨
ζ (hλ, α ∨ i + 1)
w∈W 0 α∈∆ w F
where
Y 1
f p,w (s) = ,
hλ p , α∨ is + hρ, α∨ i − 1
α∈(w−1 (∆))r∆ p
Y (12.9)
g p,w (s) = ζF hλ p , α∨ is + hρ, α∨ i .
b
α∈(w−1 (Φ− ))r∆ p
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 302
Since the coefficients hλ p , α∨ i of the variable s appeared in g p,w (s) may well
be negative, by the functional equation b ζF (1 − s) = bζF (s),
Y
g p,w (s) = ζF 1−hλ p , α is−hρ, α i
∨ ∨
b
α∈((w−1 (Φ− ))∩Φ−
Y
ζF hλ p , α∨ is + hρ, α∨ i
× b
α∈((w−1 (Φ− ))∩Φ+ )r∆ p
Y
= ζF hλ p , α∨ is+hρ, α∨ i+1
b
(12.10)
α∈((w−1 (Φ+ ))∩Φ+
Y
ζF hλ p , α∨ is+hρ, α∨ i
× b
α∈((w−1 (Φ− ))∩Φ+ )r∆ p
Y Y b
= ζF hλ p , α∨ is + hρ, α∨ i + 1 ζF hλ p , α∨ is + hρ, α∨ i .
b
α∈Φ+ rΦw α∈Φw r∆ p
1 w(α) > 0
Introduce, for w ∈ W , a characteristic function δα,w by δα,w :=
0
.
0 w(α) < 0
Accordingly, we can combine two products in (12.10) to write g p,w (s) as
+
−1
Y
g p,w (s) = b
ζF (2)|∆ p ∩w (Φ )| ζF hλ p , α∨ is + hρ, α∨ i + δα,w ,
b (12.11)
α∈Φ+ r∆ p
for which all the coefficients of s on the right hand are non-negative. Here the ad-
−1 +
ζF (2)|∆ p ∩w (Φ )| comes from the terms in the first product of (12.10)
ditional factor b
satisfying α ∈ ∆ p and w(α) ∈ Φ+ (so that hλ p , α∨ i = 0 and hρ, α∨ i = 1).
Definition 12.2.
Directly from the definition, we obtain the first two equalities below.
Proof. The first two come directly from the definition. Moreover, the second
equality implies that, for the third one, it suffices to treat `−p,w + `−p,w0 ww p . Easily,
(Φ+ r Φ+p ) ∩ (w0 ww p )−1 Φ− = w p (Φ+ r Φ+p ) ∩ w−1 Φ+ = (Φ+ r Φ+p ) ∩ w−1 Φ+ .
Hence, for all w ∈ W 0 ,
`−p,w + `−p,w0 ww p = (Φ+ r Φ+p ) ∩ w−1 (Φ− ) + (Φ+ r Φ+p ) ∩ (w0 ww p )−1 (Φ− )
= (Φ+ r Φ+p ) ∩ w−1 (Φ− ) + (Φ+ r Φ+p ) ∩ w−1 Φ+
= Φ+ r Φ+p ,
as desired.
By Proposition 10.2, there exists a fixed ε p ∈ {±1} such that, for all w ∈ W 0 ,
there are the micro functional equations
r p,w0 ww p = r p,w ,
f p,w0 ww p (−c p − s) = ε p f p,w (s), (12.16)
g p,w0 ww p (−c p − s) = g p,w (s).
Hence, by Lemma 12.1, if we set1
X 1 X
E p (s) := + cr p,w f (s) g (s), (12.17)
2 F p,w p,w
+ = w∈W p
w∈W p
then what we just said proves the following
Corollary 12.1. There exists a fixed ε p ∈ {±1}, depending only on p, such that
Z p (s) = E p (s) + ε p E p (−c p − s).
Note that all the terms in this decomposition of Z p (s) (in terms of E p (s)) are
meromorphic functions in s. This is very different from the decomposition for the
Riemann zeta function used in the study of Riemann-Siegel formula.
r p,w
1 When W p= is empty, the second summation
P
w∈W p= cF f p,w (s) g p,w (s) is defined to be zero.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 304
where
(Φ+ )|
−1
Y
c p,w := ξ(2)|∆ p ∩w ξF hρ, α∨ i + δα,w ,
α∈Φ+ r∆ p
Y (12.20)
ξF hλ p , α∨ is + hρ, α∨ i + δα,w .
z p,w (s) :=
α∈Φ+ rΦ+p
g̃ p,w (s)
In addition, by a direct calculation, is a polynomial. given by
g p,w (s)
g̃ p,w (s) −1 +
Y
= 2|∆ p ∩w (Φ )| hλ p , α∨ is+hρ, α∨ i+δα,w hλ p , α∨ is+hρ, α∨ i+δα,w −1 .
g p,w (s) α∈Φ+ r∆ p
Hence, if we set
Y g̃ p,v (s) 1 !
Q p (s) := and X p (s) := Q p (s) Z p (s), (12.21)
0
g p,v (s) f p,v (s)
v∈W
Q p (s) and Q p (s) f p,w (s) are polynomials and X p (s) is entire, since, by (12.9),
Y 1
f p,w (s) = . (12.22)
hλ p , α∨ is + hρ, α∨ i − 1
α∈w−1 (∆)r∆P
!
Y g̃ p,v (s) 1
where a p,w (s) := is the polynomial given by
g p,v (s) f p,v (s)
v∈W 0 r{w}
(Φ+ )|
Y −1
Y
a p,w (s) = 2|∆ p ∩v hλ p , α∨ is + hρ, α∨ i − 1
v∈W p0 r{w} α∈(v−1 (∆)r∆ p
Y !
× hλ p , α is + hρ, α i + δα,v hλ p , α is + hρ, α i + δα,v − 1 .
∨ ∨ ∨ ∨
α∈Φ+ r∆ p
by (12.17) and Lemma 12.1, what we have just said proves the following:
Definition 12.3. We define the polynomial R p (s), the ξ-function factor ∆ p (s) and
the ξ-function ξG/P
F (s) associated to (G, P) over F by
n o
R p (s) := g.c.d a p,w (s) : w ∈ W 0 ,
∞ Y
Y ∞
∆ p (s) := ξF (ks + h)N p (k,h−1)−M p (k,h) ,
(12.27)
k=1 h=2
X p (s)
ξG/P
F (s) := ,
R p (s) ∆ p (s)
respectively. Here ‘g.c.d’ means the monic polynomial of maximal degree which
divides a p,w (s)’s in the polynomial ring C[s].
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 306
Proof. (2) is a direct consequence of the previous proposition, and (1) comes
directly from the definition and the discussion in §10, in particular, §10.2.1 for
Theorem 10.1(1) on the normalization of our zeta function bζFG/P (s).
Therefore, to prove Theorem 12.1, it suffices to prove the same statement for
the entire function ξG/P
F (s) for a Chevalley group G over Q.
Lemma 12.2.
This establishes the first equality. To prove the second, note that, for α ∈ Φ+ ,
hρ, αi ≥ 1. Hence, with n p (k, h) = α ∈ Φ : hλ p , α∨ i = k, hρ, α∨ i = h ,
Y ∞ ∞
YY
z p (s) = ξ hλ p , α∨ is + hρ, α∨ i + 1 = ξ(ks + h)N p (k,h−1) ,
α∈Φ+ rΦ+p k=1 h=2
as desired.
Lemma 12.4. With the same notation as above, for a fixed w ∈ W 0 , as |s| → ∞,
Y 1
d p,v
hλ p , α∨ is + hρ, α∨ i − 1
a p,v (s) α∈v−1 (∆)rΦ p
= X . (12.34)
a p (s) r p,w
Y 1 !
cF c p,w d p,w 1+o(1)
hλ p , α∨ is+hρ, α∨ i−1
w∈W p α∈w (∆)rΦ p
−1
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 308
Proof. We first write down the terms of a p,w (s) explicitly. From the definition, for
w ∈ W p0 ,
Y " −1 +
Y
a p,w (s) = 2|∆ p ∩v (Φ )| hλ p , α∨ is + hρ, α∨ i − 1
v∈W 0 r{w} α∈(v−1 (∆))r∆ p
Y #
hλ p , α is + hρ, α i + δα,v
∨ ∨
hλ p , α is + hρ, α i + δα,v − 1
∨ ∨
×
α∈Φ+ r∆ p
Y " 1 Y
= hλ p , α∨ is + hρ, α∨ i − 1
d p,v
0
v∈W r{w} α∈(v−1 (∆))rΦ p
Y #
hλ p , α is + hρ, α i + δα,v
∨ ∨
hλ p , α is + hρ, α i + δα,v − 1
∨ ∨
×
α∈Φ+ rΦ+p
" Y 1
= a∗p (s) d p,w
hλ p , α∨ is + hρ, α∨ i − 1
α∈w−1 (∆)rΦ p
#
Y 1
× .
hλ p , α∨ is + hρ, α∨ i + δα,w hλ p , α∨ is + hρ, α∨ i + δα,w − 1
α∈Φ+ rΦ+p
Consequently,
X r
a p (s) = cFp,w c p,w a p,w (s)
w∈W p
X " r
Y 1
= a∗p (s) cFp,w c p,w d p,w
hλ p , α∨ is + hρ, α∨ i − 1
w∈W p α∈w−1 (∆)rΦ p
#
Y 1
× .
hλ p , α∨ is + hρ, α∨ i + δα,w hλ p , α∨ is + hρ, α∨ i + δα,w − 1
α∈Φ+ rΦ+p
Therefore, as |s| → ∞,
Y 1
d p,v
hλ p , α∨ is + hρ, α∨ i − 1
a p,v (s) α∈v−1 (∆)rΦ p
= X " ,
a p (s) r
Y 1 #
cFp,w c p,w d p,w 1 + o(1)
hλ p , α∨ is + hρ, α∨ i − 1
w∈W p α∈w−1 (∆)rΦ p
as desired.
We next show that, as |s| → ∞, a p (s) dominants a p,w (s) for w ∈ W ≥ r W p , using
(12.34). For this, we first examine the Lie structures of W p ⊂ W 0 .
The discussions so far can be used to study the denominator of (12.34) since
w−1 (∆) r Φ p = w−1 (α p ) for w ∈ W p . To treat the numerator, let w ∈ W 0 be an
element satisfying the condition w−1 (∆) r Φ p = 1. Accordingly, denote by βw
the unique element of ∆ ∩ w(Φ p ), and let ∆±w := ∆ ∩ w(Φ±p ). Then
∆ = ∆+w t ∆−w t { βw }. (12.36)
But the root α is simple, so only the second sum remains. That is, α ∈ ∆ ∩ w(∆ p ).
This then implies that ∆+w ⊆ ∆ ∩ w(∆ p ).
Definition 12.7. We define, for integers k and h, the subsets Σ p (k) and Σ p (k, h) of
Φ by
Σ p (k) := α ∈ Φ : hλ p , α∨ i = k ,
(12.37)
Σ p (k, h) := α ∈ Σ p (k) : hρ, α∨ i = h .
Obviously, for all but finitely many k and (k, h), Σ p (k) and hence Σ p (k, h)’s are
empty. Moreover, we have
G∞ G∞
Φ+ r Φ+p = Σ p (k) and Σ p (k) = Σ p (k, h). (12.38)
k=1 h=1
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 311
Case (A.1.i) Assume αi0 ∈ ∆+w . Then, αi0 ∈ ∆ ∩ w(∆ p ) ⊂ Φ+ ∩ w(Φ+ ). This
contradicts with (12.39).
Case (A.1.ii) Assume αi0 ∈ ∆−w . Then, αi0 ∈ Φ+ ∩ w(Φ−p ). This again contradicts
with (12.39).
Lemma 12.7. Let w ∈ W 0 and α ∈ Φ+ . Assume that α ∈ (Φ+ rΦ+p ) ∩ w−1 (Φ− ) and
α + α j ∈ Φ+ rΦ+p for some α j ∈ ∆ p . Then α + α j ∈ (Φ+ rΦ+p ) ∩ w−1 (Φ− ).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 312
Proposition 12.4. Assume the discriminant of (G, P) over F is not zero. We have
deg a p (s) ≥ deg a p,v (s) + 1 ∀v ∈ W p≥ r W p . (12.40)
That is to say, a p (s) is a dominant polynomial, comparing with polynomials a p,v (s)
for all v ∈ W p≥ r W p .
Y 1
By Corollary 12.3, v−1 (∆) r Φ p ≥ 2. Hence, is
hλ p , α∨ is + hρ, α∨ i − 1
α∈v−1 (∆)rΦ p
at least of order s , since, for all a ∈ v (∆) r Φ p , hλ p , α∨ i , 0. Therefore,
−2 −1
We end this section by a remark that, unlike the rest of this chapter, the re-
sults in this and the previous subsection are mainly supplied by the author, as the
corresponding arguments in [56] do not work.
Hence, by (12.20),
z p,v (s) Y ξF (hλ p , α∨ i + hρ, α∨ i)
= . (12.42)
z p (s) ξF (hλ p , α∨ i + hρ, α∨ i + 1)
α∈(Φ+ rΦ+p )∩v−1 (Φ− )
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 314
To estimate these quotients, as above, we need to understand what are the struc-
tures involved for (Φ+ r Φ+p ) ∩ v−1 (Φ− ). This naturally leads to the decomposition
∞
G ∞
G
Φ+ r Φ+p = Σ p (k) and Σ p (k) = Σ p (k, h), (12.43)
k=1 h=1
Obviously, for all but finitely many k, resp. (k, h), the set Σ p (k), resp. Σ p (k, h),
is empty. More precisely, we have the following
Lemma 12.8. (Lemma 1.4.5 of [28]) Let ni=1 ki αi (ki > 0) be the highest root of
P
Φ+ . Then, for k ∈ Z>0 , the set Σ p (k) is not empty if and only if 1 ≤ k ≤ k p .
Definition 12.8. For a positive integer k with Σ p (k) , ∅, we define the lowest
root α−p (k), resp. the highest root α+p (k), of Σ p (k) to be the root of Σ p (k) such
that β − α−p (k), resp. α+p (k) − β, is (possibly empty) sum of simple roots for every
β ∈ Σ p (k).
Lemma 12.9. (Proposition 1.4.2 of [28]) For k ∈ Z>0 , if Σ p (k) , ∅, there exists a
unique lowest root α−p (k), resp. unique highest root α+p (k), in Σ p (k).
Proof. The first two are simply given in Lemma 1.4.6 of [28]. For the third, by
definition, w p α−p (k)∨ = α+p (k)∨ = kα∨p + β+p (k)∨ . Hence, by Lemma 12.9 (2),
But α−p (k)∨ = kα∨p + β−p (k)∨ . Therefore, β−p (k) − w p β+p (k) = kβ+p (1).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 315
(1) the invariants h±k are defined by h±k := hρ, β±p (k)∨ i.
(2) the p-level of α ∈ Σ p (k) is defined to be hρ, (α − kα p )∨ i.
Obviously, h±k ≥ 0 and h+k , resp. h−k , is the highest level , resp. the lowest level,
of (the elements in) Σ p (k).
Moreover, by Corollary of [53], we have α+p (k2 )∨ −α+p (k1 )∨ ∈ Φ∨ . Since k2 −k1 > 0
and β+p (k)∨ ∈ j: j,p Z≥0 α∨j , we conclude that β+p (k2 ) − β+p (k1 ) ∈ j: j,p Z≥0 α j .
P P
Therefore, by definition, h+k + h−k = kh+1 .
(2) It suffices to prove h+k + h−k = kh+1 , since others are direct consequences of
this relation. By the proof of Lemma 10.1, hρ p , α∨i i = 1 for i , p. Hence, for
β∨ ∈ j: j,p Z≥0 α∨j ,
P
Therefore, by definition,
k + h−k = hρ, α−p (k)∨ i = hρ, w p α+p (k)∨ i = hρ, w p kα∨p + β+p (k)∨ i
= khρ, w p α∨p i + hρ, w p β+p (k)∨ i = khρ, w p α∨p i − hρ, β+p (k)∨ i
Now we are ready to decompose Σ p (k). We start with a property of Σ p (k, h).
Lemma 12.12. Let k and h be positive integers such that Σ p (k, h) is not empty.
Then, for Σ p (k, h) =: β1 , . . . , βn p (k,h) , there exists simple roots α j1 , . . . , α jn p (k,h) in
∆ p (not necessary distinct) satisfying the conditions
With this, we can effectively construct Σ p (k) starting from α−p (k). For example,
when k = 1, we have the following
(1) L1 (1) 3 α±p (1) and L1 (1) = h ρ, α+p (1) − α−p (1) i + 1;
(2) L1 (1) > L2 (1) , L3 (1) , . . . , L M(1) (1) ≥ 1;
(3) There exist some βm (1) ∈ Σ p (1) with β1 (1) = α−p (1) and some αm, j (1) ∈ ∆ p
(1 ≤ m ≤ M(1), 1 ≤ j ≤ Lm (1) − 1) satisfying
n o
(3.1) Lm (1)∨ := { βm (1)∨ } t βm (1)∨ + lj=1 αm, j (1)∨ : 1 ≤ l ≤ Lm (1) − 1 ,
P
(3.2) 1 + h−1 = hρ, β1 (1)∨ i, hρ, β M(1) (1)∨ i ≤ c p /2 and
hρ, β1 (1)∨ i < hρ, β2 (1)∨ i ≤ hρ, β3 (1)∨ i ≤ . . . ≤ hρ, β M(1) (1)∨ i,
D E
(3.3) For all 1 ≤ k ≤ M(1), ρ , 2βm (1)∨ + |Lj=1
P m (1)|−1
αm, j (1)∨ = c p .
We will delay a detailed proof of this lemma to the appendix of this chapter. In-
stead, we give a similar result for all the k’s.
(1) L1 (k) 3 α±p (k) and L1 (k) = h ρ, α+p (k) − α−p (k) i + 1;
(2) L1 (k) > L2 (k) , L3 (k) , . . . , L M(k) (k) ≥ 1;
(3) There exist some βm (k) ∈ Σ p (k) with β1 (k) = α−p (k) and some αm, j (k) ∈ ∆ p
(1 ≤ m ≤ M(k), 1 ≤ j ≤ Lm (k) − 1) satisfying
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 317
n o
(3.1) Lm (k)∨ := { βm (k)∨ } t βm (k)∨ + lj=1 αm, j (k)∨ : 1 ≤ l ≤ Lm (k) − 1 ,
P
(3.2) k + h−k = hρ, β1 (k)∨ i, hρ, β M(k) (k)∨ i ≤ kc p /2 and
hρ, β1 (k)∨ i < hρ, β2 (k)∨ i ≤ hρ, β3 (k)∨ i ≤ . . . ≤ hρ, β M(k) (k)∨ i,
D E
(3.3) For all 1 ≤ k ≤ M(k), ρ , 2βm (k)∨ + |Lj=1
P m (k)|−1
αm, j (k)∨ = kc p .
Proof. By §1.4 of [28], we may reduce the problem to decompose Σ p (1) for irre-
ducible root systems. Hence it suffices to apply the previous lemma to complete
the proof.
and let
hm (k) := max hρ, α∨ i : α ∈ Lm (k) ,
(12.48)
hm,w (k) := max hρ, α∨ i : α ∈ Lm (k)w .
Here in the last definition for hm,w (k) we assume that m ∈ Λw (k).
Lemma 12.14. The value hm (k) is attained by one of α ∈ Lm (k)w . More strongly,
Proof. It suffices to prove (1) and (2). For (1), from the condition hρ p , α∨ i < 0,
there exists α j ∈ ∆ p such that hα j , α∨ i < 0, since
X X
2ρ p = α= n jα j (n j ∈ Z>0 ).
α∈Φ+p 1≤ j,p≤n
In particular,
> 0 h > kc p /2,
hρ p , α i
∨
(12.50)
< 0 h < kc p /2.
As the first application of the Lie theoretic structures we exposed, we now can
treat the quotient z p,w (s) z p (s) of (12.42) for w ∈ W 0 r W p . From now on till the
end of this chapter, we assume that F = Q.
Here in the last step, we have used (3.1) of Lemma 12.5 and Lemma 12.14. Con-
sequently, by (11.26), we have, for every m ∈ Λw (k),
ξ(ks + hm,w (k)) hm,w (k) + hm (k)
<1 <(s) > − .
ξ(ks + hm (k) + 1) 2k
Now, by Lemma 12.5(3.3), hm,w (k) ≥ hρ, βm (k)∨ i = kc p − hm (k). Hence, for m ∈
hm,w (k) + hm (k) c p
Λw (k), ≥ . This proves the first inequality of (12.51), unless
2k 2
hm,w (k) + hm (k) = kc p . Clearly, in the exceptional case when hm,w (k) + hm (k) = kc p ,
by (11.26) again, we have
ξ(ks + hm,w (k)) cp
=1 <(s) = − ,
ξ(ks + hm + 1) 2
as desired.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 319
As the second application of the Lie structure exposed in §12.4.1, we here explic-
itly determine M p (k, h) and ∆ p (s) used in the normalization process.
Definition 12.10. We define the constant D0p and the function D1p (s) by
∞
Y ∞ Y
Y ∞
D0p := ζ(h)n p (0,h−1)−M p (0,h) , D1p (s) :=
b ζF (ks+h)n p (k,h−1)−M p (k,h) , (12.54)
b
h=2 k=1 h=2
Proof. (1) is proved in Lemma 10.2. As for (2) (and in fact also (1)), please refer
to Proposition 1 of [73].
Proof. Since n p (k, h) = n p,w0 (k, h), this is a direct consequence of Lemmas 12.16
and 12.17.
With the normalization factor treated, we are now ready to study zero free regions
of the function A p (s). We start with a simple case over a half plane on the right.
z p (s) cp
(1) There exists no zero for in the right half plane <(s) ≥ − .
∆ p (s) 2
z p (s) cp
(2) There exists no zero for in the region <(s) ≥ − − δ(t), where δ(t)
∆ p (s) 2
is a certain positive function satisfying δ(t) log |t| → ∞ (|t| → ∞).
Proof. Recall that by Lemma 12.3 and Corollary 12.4, we have, respectively,
∞ Y
Y ∞
z p (s) = ξ(ks + h)N p (k,h−1) ,
k=1 h=2
∞
(12.57)
Y Y Y
∆ p (s) = ξ(ks + h) ξ(ks + h) .
n p (k,h−1) n p (k,h)
k=1 h≤(kc p +1)/2 h>(kc p +1)/2
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 321
This implies
∞
z p (s) Y Y
= ξ(ks + h)n p (k,h−1)−n p (k,h) . (12.58)
∆ p (s) k=1 h>(kc +1)/2
p
In particular, the right hand contains no zeta factor in the denominator, since
n p (k, h − 1) − n p (k, h) ≥ 0 by Lemma, 12.17.
1−h
Recall that ξ(s) , 0 for <(s) ≥ 1. Hence ξ(ks + h) , 0 for <(s) ≥ .
k
kc p + 1 cp 1 − h
Therefore, it suffices to show that when k ≥ 1, h > implies ≥ .
2 2 k
kc p + 1 1−h c p
But this is rather trivial, since, if kc p is even, h > implies ≤− ,
2 k 2
1−h cp 1 cp
and if kc p is odd, ≤− − < − . This proves (1) and hence also
k 2 2k 2
(2) by a standard work on zero free region of ξ(s) in [116]. In fact, follow-
ing [116], with the Vinogradov method, δ(t) can be taken to satisfy the condition
1
δ when |t| → ∞.
(log t)2/3 (log log t)1/3
B p (s)
Recall that, in (12.29), A p (s) = is an entire function defined using
R p (s) ∆ p (s)
X 1 X r
B p (s) := + c p,w c a (s) z (s),
F p,w p,w p,w
w∈W >
2 w∈W =
(12.59)
p p
n o
R p (s) := g.c.d a p,w (s) : w ∈ W 0 .
As a direct consequence of Propositions 12.4, 12.6 and Lemma 12.18, we have
(1) There are at most finitely many zeros of the function A p (s) lying in the right
cp
half-plane <(s) ≥ − .
2
(2) There are at most finitely many zeros of the function A p (s) lying in the re-
cp
gion <(s) ≥ − − δ(t), where δ(t) is a certain positive function satisfying
2
δ(t) log |t| → ∞ ( |t| → ∞).
where
1 X a p,w (s) z p,w (s)
X !
r p (s) := + · .
>
w∈W p rW p
2 w∈W = a p (s) z p (s)
p
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 322
Thus, if we set
X 1 X a p,w (s)
+ ,
D p := s ∈ : ≥ 1
C
2 a (s)
>
w∈W rW p w∈W p=
p
p
We next treat the zero free region on the left. This is more crucial and complicated.
Proposition 12.7. There exists no zero for the function B p (s) in the region on the
left defined by <(s) ≤ −κ p log |=(s)| + 10 , where κ p is a certain positive real
number.
Since z p (s) is a finite product of zeta functions and hence can be treated rather
easily as above using the functional equation. Hence it suffices to find the zero
free region on the left for the function a p (s) + b p (s).
As in the proof of Proposition 12.6, we use
k(p) Y
z p,w (s) Y ξ(ks + hm,w (k))
= , (12.61)
z p (s) k=1 m∈Λw (k)
ξ(ks + hm (k) + 1)
(2) As σ → −∞, for any fixed n ≥ 0, there exists real constants c j (a, b) such that
! a−b−1
ξ(ks + a)
!
2π 2
c1 (a, b) cn (a, b)
= − s 1+ + ... + + O |s| −n−1
ξ(ks + b + 1) k s sn
X Cm (k; a, b)
× 1 + .
m≥2
m−s
Proof. (1) To write down the left hand side as a Dirichlet series, it suffices to use
∞
1 X
the well-known formula = µ(n) n−s , where, as usual, µ(n) denotes the
ζ(s) n=1
n-th Mobius number.
(2) is a direct consequence (1). Indeed by the well-known Stirling formula,
we have, for | arg(s) | < π − ε,
√
!
1 1 1
Γ(s) = 2πss e 1 +
s −s
+ + ... + + O |s| −n−1
. (12.62)
12s 288s2 cn sn
Therefore, with | arg(s)| < π−ε, as |s| → ∞, there are real coefficients polynomials
a j (x) such that
Γ(s + λ)
!
a1 (λ) an (λ)
= sλ 1 + + . . . + n + O |s|−n−1
(12.63)
Γ(s) s s
where O |s|−n−1 depends on λ and ε. Applying this, resp. (1), to the Gamma
ξ(ks + a)
factors, resp. the zeta factors, of , we obtain (2) with an obvious
ξ(ks + b + 1)
change of variables.
z p,w (s)
Directly applying this result to each term of in (12.61), we obtain
z p (s)
!
z p,w (s) c1 (w) cn (w) X Cm (w)
= (−2πs)a p,w eb p,w 1+ + . . . + n + O |s|−n−1 1 +
z p (s) s s m−s
m≥2
for some real constants ci (w) and C j (w). Here
k(p)
1X X
hm (k) − hm,w (k) + 1 ,
a p,w :=
2 k=1 m∈Λ (k)
w
(12.64)
k(p)
1 X X
hm (k) − hm,w (k) + 1 .
b p,w := log k
2 k=1 m∈Λw (k)
Since, for 1 ≤ k ≤ k(p) and m ∈ Λw (k), by definition, hm (k) − hm,w (k) ≥ 0, we
1
have, a p,w ∈ Z>0 . Hence, for sufficiently large n,
2
∞ n
X Qµ (s1/2 ) X ck (µ)
a p (s) + b p (s) = 1 + + O |s|
−(n+1)/2
(12.65)
µ
−s k/2
µ=1 k=1
s
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 324
n o
where Qµ (s) are polynomials and M := max deg(Qµ (s)) : µ = 1, 2, . . . < ∞.
Let µ0 be the smallest positive integer X = M. Then, we can
XsuchX
that deg
X Qµ0 (s)
rewrite the summation in (12.65) as = + + so as to get
µ≥1 µ=µ0 µ≤µ0 −1 µ≥µ0 +1
n
1/2
Qµ0 (s ) X c1 (µ 0 )
a p (s) + b p (s) = 1 + + O |s|−(n+1)/2
µ0
−s k/2
k=1
s
µ −1 ∞
n
X0 X 1 Qµ (s1/2 ) X c1 (µ)
+ + 1 + + O |s| .
−(n+1)/2
µ=1 µ=µ +1
(µ/µ0 )−s Qµ0 (s1/2 ) k=1
sk/2
0
Now we take σ0 > 0 such that, for <(s) < −σ0 and |=(s)| > 1,
∞ n
1 Qµ (s1/2 ) (µ)
c < 1 .
X X 1
1 + + O |s|−(n+1)/2
(µ/µ )−σ Q (s 1/2 ) s k/2 2
µ=µ +1 0 µ0 k=1
0
Hence, for <(s) < −σ0 , when |=(s)| → ∞, with our choice of µ0 , we have
µX0 −1
n
1 Qµ (s1/2 ) X c1 (µ) = O µσ0 |s|−1/2 .
1 + + O |s|
−(n+1)/2
(µ/µ0 )−σ 1/2
Qµ0 (s ) s k/2
µ=1 k=1
Therefore, for sufficiently large R > 0, when <(s) < σ0 and |=(s)| > R,
Qµ0 (s1/2 )
a p (s) + b p (s) = 1 + ϕ(s; 1/2) + O µ−σ
0 |s|
−1/2
(12.66)
µ0 −s
for a certain function ϕ(s; 1/2) satisfying ϕ(s; 1/2) < 1/2. Consequently, when
|s| → ∞ with σ < κ p log(|t| + 10),
B p (s) = z p (s) a p (s) + b p (s) = z p (s) Qµ0 (s1/2 ) µ0s 1 + φ(s)
(12.67)
for a certain function φ(s) satisfying |φ(s)| < 1. If necessary, by enlarging κ p , we
deduce the assertion for zeros of B p (s) using (12.67).
Corollary 12.7. All but finitely many zeros of the function A p (s) lie in the region
cp
s = σ + it ∈ C : −κ p log |t| + 10 < σ < − .
(12.68)
2
Proof. Indeed, form the proof above, particularly, (12.67), we have
z p (s) 1 z p (s) 1
A p (s) = a p (s) + b p (s) = Qµ0 (s1/2 ) µ0s 1 + φ(s) .
∆ p (s) R p (s) ∆ p (s) R p (s)
If necessary, by enlarging κ p , this expression of A p (s) takes care of the lower
bounded of σ. As for the upper bound of σ, we apply Corollary 12.6 and Propo-
sition 12.7.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 325
We first recall a modification of Lemma 11.3, taken from [116], see e.g. page 213.
Lemma 12.20. Fix σ0 > 0, T > 10. Let 0 ≤ α < β < σ0 , where α and β may
depend on T . Let f (s) be an analytic function admitting no zero on the line t = T ,
real for real variables, and regular for σ ≥ α except for possibly finitely many
poles on the real line. Choose constants m, Mσ,t satisfying
Then, for σ ≥ β,
π 3
arg f (σ + iT ) ≤ log Mα,T +2 + log m + π. (12.69)
log (σ0 − α)/(σ0 − β) 2
We use this lemma to analysis the distributions for the zeros of A p (s).
cp
Proposition 12.8. Let T > 1 and σ > . Denote by N(T ; σ) be the number of
2
zeros of A p (s) in the region
n o
s ∈ C : −σ < <(s) < −c p /2 − δ(t), 0 < =(s) < T . (12.70)
Assume that the discriminant of b ζQG/P (s) is not zero. There exist real constants
σL > 0, and c1 > 0, c2 , c3 such that
N(T ; σL ) = c1 T log T + c2 T + O(log T ),
(12.71)
N(T ; +∞) = c1 T log T + c3 T + O(log2 T ).
In particular,
Proof. This is rather standard. Indeed, from the proof of Proposition 12.7, there
exist a positive integer µ0 and a suitable function φ(s) together with a positive real
number σL such that
(1) a p (s) + b p (s) = Qµ0 (s1/2 ) µ0s 1 + φ(s) .
(2) | φ(−σL + it) | < 1 for some fixed σL > 0 as |t| → ∞.
(3) | φ(s) | < 1 for <(s) < κ p log |=(s)| + 10 as |s| → ∞.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 326
Moreover, using the function r p (s) introduced in the proof of Corollary 12.6, i.e.
X 1 X a p,w (s) z p,w (s) !
r p (s) := + a p (s) · z p (s) ,
(12.73)
>
w∈W rW
2 w∈W =
p p p
we obtain, by the definition of R p (s) in (12.59) and the proof of Corollary 12.6, in
particular, (12.60),
a p (s1/2 )
(4) is a polynomial,
R p (s)
z p (s) a p (s1/2 )
(5) A p (s) = 1 + r p (s) , and
∆ p (s) R p (s)
(6) |r p (s)| < 1 for <(s) > −c p /2 if s < D p , where
X 1 X a p,w (s)
+ .
D p := s ∈ : ≥ 1 (12.74)
C
2 a (s)
>
w∈W rW w∈W =
p
p p p
∞
Y Y
Γ∗ (s) := γ(ks + h)n p (k,h−1)−n p (k,h) ,
k=1 h>(kc p +1)/2
∞
a p (s) Y Y
L(s) := ζ(ks + h)n p (k,h−1)−n p (k,h) ,
R p (s) k=1 h>(kc +1)/2
p
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 327
Obviously, it suffices to treat the term ∆2 arg L(s) 1+φ p (s) , since a standard cal-
culation using the Stirling formula easily evaluates ∆2 arg Γ∗ (s) . By definition,
"
X 1 X a p,w (s)
L(s) φ p (s) = +
>
w∈W p rW p
2 w∈W = R p (s)
p
# (12.75)
γ hλ p , α is + hρ, α∨ i
∨
Y
× ζ hλ p , α is + hρ, α i .
∨ ∨
γ hλ p , α∨ is + hρ, α∨ i + 1
α∈(Φ+ rΦ+p )∩w−1 (Φ− )
a p,w (s)
Since is a polynomial, it is also quite standard to estimate this term. In-
R p (s)
deed, by Chapter V of [116], there exists a positive number M such that, when
−σL ≤ <(s) ≤ σR , we have L(s) 1 + φ p (s) T M . Therefore, by Lemma 12.20,
∆2 arg L(s) 1 + φ p (s) = O log T .
This proves the asymptotic formula for N(T ; ∞) and hence the proposition as
well.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 328
We here show that A p (z) admits a refined structure of the Hadamard product, using
Proposition 12.8. Set for simplicity
c
ep (z) := A p − p + i z ,
A (12.76)
2
The important parts are that α is a real number and the locations of the zeros
of G(z) H1 (z) H2 (z).
Proof. The function A p (z) is a linear combination of the products of finite zeta
function factors b ζ(ks + h), or better ξ(ks + h), with the coefficients of rational
function. Hence A ep (z) is an entire function of order at most one, since the ξ(ks +
h)’s are entire functions of order one (and maximal type). In addition, A p (z) takes
real values over real axis. Hence, by the functional equation, if ρ is a zero of
ep (z), so is −ρ. Therefore, by the theory of entire functions, using Corollary 12.6
A
and Proposition 12.7 on the locations of zeros of A p (z), we conclude that A ep (z)
admits an Hadamard product of the form
ep (z) = κ eβz G(z) h1 (z) h2 (z),
A (12.79)
where κ, resp. β, resp. G(z), is a non-zero real number, resp. a complex num-
ber, resp. a non-zero polynomial having no zero in =(z) > 0 except for purely
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 329
Lemma 12.21. With the same notation as above, α ∈ R, or the same =(α) = 0.
B p − c p /2 + y
= .
B p − c p /2 − y
Thus it suffices to verify the statement in the latest lemma for the function BP (s)
instead of A p (s). Now, by Propositions 12.4 and 12.6, we have, as y → +∞,
B p − c p /2 + y = a p − c p /2 + y · z p − c p /2 + y · 1 + o(1) .
(12.83)
Claim 12.1. As y → +∞, there exists some constant A , 0 and m ≥ 0 such that
B p − c p /2 − y
= A−1 y−m 1 + o(1) . (12.84)
a p − c p /2 + y · z p − c p /2 + y
B p − c p /2 − y
Proof. By definition, is given by
a p − c p /2 + y z p − c p /2 + y
1 X a p,w − c p /2 − y z p,w − c p /2 − y
X
+ .
2 w∈W = a p − c p /2 + y z p − c p /2 + y
w∈W > rW
p p p
where δ0w = 1 if w ∈ W p and w−1 (∆) r Φ p = 1, and 0 otherwise. As for the zeta
factor, we write
z p,w − c p /2 − y z p,w − c p /2 + y z p,w − c p /2 − y
= · .
z p − c p /2 + y z p − c p /2 + y z p,w − c p /2 + y
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 331
z p,w − c p /2 − y
Hence, it suffices to analyze . Recall that, by Lemma 12.19, as
z p,w − c p /2 + y
σ → +∞,
ξ k(−c p − s) + h + m
ξ ks + h + m
!− kc p −2h−2m+1 !
2π 2
c1 (k, h) cn (k, h)
= 1+ + ... + n
+ O(|s|−n−1
) 1+O(2−σ ) .
ks s s
Thus, by §12.4.1, there exist real numbers b01 (w), . . . , b0n (w) such that, as σ → +∞,
z p,w (−c p − s) Y ξ hλ p , α∨ i(−c p − s) + hρ, α∨ i + δα,w
=
ξ hλ p , α∨ is + hρ, α∨ i + δα,w
z p,w (s) α∈Φ+ rΦ+ p
+
p +hk
k(p) kcY
ξ k(−c p − s)+h n p,w (k,h) ξ k(−c p − s)+h+1 n p (k,h)−n p,w (k,h) (12.87)
Y ! !
=
ξ ks + h ξ ks + h + 1
k=1 h=1
2π −a0p,w 0 !
b1 (w) bn (w)
= eb p,w 1 + + . . . + n + O |s|−n−1 1 + O(2−σ .
s s s
0 0
Here, a p,w and b p,w are real constants with
+
k(p) k+h
X Xk
a0p,w := n p,w (k, h)(kc p − 2h + 1) + (n p (k, h) − n p,w (k, h))(kc p − 2h − 1) .
k=1 h=k+h−k
This implies
+
p +hk
k(p) kcX
X
a0p,w = 2n p,w (k, h) − n p (k, h)
k=1 h=k+h−k
+ +
p +hk
k(p) kcX
X p +hk
k(p) kcX
X
=2 2n p,w (k, h) − n p (k, h) + n p (k, h)
k=1 h=k+h−k k=1 h=k+h−k
>
> 0 w ∈ W p
Therefore, −a p,w =
0
Consequently, by (12.85) and (12.87), we
0
w ∈ W p= .
have, as y → +∞, there exist real numbers b∗1 (w), . . . , b∗n (w) such that
B p − c p /2 − y
a p − c p /2 − y z p − c p /2 − y
(12.88)
b∗1 (w) b∗n (w)
!
−(a p,w −a0p,w ) b∗p,w
=y e δw +
0
+ . . . + n + O |y| −n−1
1 + O(2 .
−y
y y
This completes the proof of claim and hence two lemmas and the proposition.
Proof. Now we are ready to complete a proof of Theorem 12.1 following [56].
By Proposition 12.2,
In addition, if the discriminant ∆G/P Q of the zeta function bζQG/P (s) is not zero, by
Theorem 12.9, A p (s) admits a refined Hadamard product (12.77). Hence, by the
conditions (1), (2) and (3) listed in that theorem, it is possible to apply Lemma 11.6
for W(z) = A ep (z) = A p (−c p /2 + iz) to conclude that all but finitely many zeros of
cp
ζQG/P (s) lie on the central line <(s) = − .
b
2
To prove the related zeros are simple, we use the algebraic structures of A p (s)
again. Indeed, from (12.89), we have
ξQG/P − c p /2 + it = A p − c p /2 + it eiθ(t) ± eiθ(t) .
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 333
Here θ(t) := arg A p − c p /2 + it . Thus, by a standard argument, it suffices to
show that θ(t) is strictly increasing as t → +∞. By Propositions 12.4 and 12.6, on
c
the central line <(s) = − 2p , we have, as |s| → ∞
a p (s) z p (s)
B p (s) = a p (s) z p (s) 1 + o(1) and hence A p (s) = 1 + o(1) .
R p (s) ∆ p (s)
Thus, by (12.58), we have, as t → +∞, there exists a suitable constant c such that
d
θ − c p /2 + it = c log t · 1 + o(1) .
dt
z (s)
Indeed, the factor log t comes from the Gamma factors of ∆pp (s) in (12.58), and the
following standard estimation (see e.g. Theorem 5.7 of [116])
ζ0
!
log t
(1 + it) = O = o(log t),
ζ log log t
in the case3 kc p is even and n p (k, h − 1) − n p (k, h) > 0 for some 1 ≤ k ≤ k(p) with
h = 1 + kc p /2. Therefore, we complete a proof of Theorem 12.1.
Proposition 12.10. The residue of ωGF (λ) at the Weyl vector λ = ρ is given by
X r −1 −
Resλ=ρ ωGF (λ) = ζF (2)−|∆∩w (Φ )|
cFp,w b
w∈W o
Y 1 Y ζF (hρ, α∨ i) (12.90)
.
b
×
hρ, α∨ i − 1 ζF (hρ, α∨ i + 1)
α∈w−1 (∆)r∆ α∈(Φ+ r∆)∩w−1 (Φ− )
b
Proof. This result is first proved in [57] at page 1684 (see also (10.4) of [56]). By
the proof of Proposition 7.6 we have
X r −1
(Φ− )|
ωG/P
F (s) = ζF (2)−|∆ p ∩w
cFp,w · b
w∈W 0
Y 1 Y ζF (hρ, α∨ i)
b
×
hρ, α∨ i −1 ζF (hρ, α∨ i + 1)
α∈w−1 (∆)∩(Φ p r∆ p ) α∈(Φ+p r∆ p )∩w−1 (Φ− )
b
Y 1 Y ζF (hλ p , α∨ is + hρ, α∨ i)
b
× .
hλ p , α∨ is + hρ, α∨ i − 1 ζF (hλ p , α∨ is + hρ, α∨ i + 1)
α∈w−1 (∆)∩Φ p α∈(Φ+ rΦ+p )∩w−1 (Φ− )
b
On the other hand, by the Lie structures exposed in §12.4.1, we have either
{α p } α p ∈ w (∆),
−1
(1) ∆ ∩ w (∆) r Φ p = {α p } ∩ w (∆) =
−1 −1
or
∅
otherwise,
{α p } α p ∈ w (Φ ),
−1 −
+ +
(2) ∆ ∩ (Φ r Φ p ) ∩ w (Φ ) =
−1 −
∅
otherwise,
and two cases cannot happen at the same time. Therefore, using hλ p , α∨p i = 1, we
conclude that the residue Res s=0 ωG/P
F (s) is equal to
X r −1
(Φ− )|
Res s=0 ωG/P
F (s) = ζF (2)−|∆ p ∩w
cFp,wb
w∈W 0
α p ∈w−1 (∆)
Y 1 Y ζF (hρ, α∨ i)
b
×
hρ, α i − 1
∨
ζF (hρ, α∨ i + 1)
α∈w−1 (∆)∩(Φ p r∆ p ) α∈(Φ+p r∆ p )∩w−1 (Φ− )
b
Y 1 Y ζF (hρ, α∨ i)
b
×
hρ, α i − 1
∨
ζF (hhρ, α∨ i + 1)
α∈w−1 (∆)r(Φ p ∪{α p }) α∈(Φ+ rΦ+p )∩w−1 (Φ− )
b
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 335
X r cF b −1 −
+ cFp,w ζF (2)−|∆ p ∩w (Φ )|
w∈W 0
ζF (2)
b
α p ∈w−1 (Φ− )
Y 1 Y ζ(hρ, α∨ i)
b
×
hρ, α∨ i −1 ζ(hρ, α∨ i + 1)
α∈w−1 (∆)∩(Φ p r∆ p ) α∈(Φ+p r∆ p )∩w−1 (Φ− )
b
Y 1 Y ζ(hhρ, α∨ i)
b
×
hρ, α∨ i −1 ζ(hhρ, α∨ i + 1)
α∈w−1 (∆)rΦ p α∈(Φ+p r(Φ+p ∪{α p }))∩w−1 (Φ− )
b
X r −1
(Φ− )|
Y 1 Y ζ(hρ, α∨ i)
= ζ(2)−|∆ p ∩w
b
cFp,wb
hρ, α∨ i −1 ζ(hρ, α∨ i + 1)
α∈w−1 (∆)r∆ α∈(Φ+p r∆)∩w−1 (Φ− )
b
w∈W 0
α p ∈w−1 (∆)
X r +1 −1 −
Y 1 Y ζ(hρ, α∨ i)
+ ζ(2)−|∆∩w (Φ )|
b
cFp,w b
hρ, α i − 1
∨
ζ(hρ, α∨ i + 1)
α∈w−1 (∆)r∆ α∈(Φ+ r∆)∩w−1 (Φ− )
b
w∈W 0
α p ∈w−1 (Φ− )
as desired.
Remark 12.1. We will prove in Part 6 that this residue coincides with the total
mass of semi-stable principle G-lattices with certain fixed invariants, as predicted
in the conjecture on stability, parabolic induction and masses.
Recall that, for the definition of periods of G and (G, P) associated to a reduc-
tive group G and its a standard maximal group P, what really matters is their
associated root system (V, h·, ·i; Φ, Φ± , ∆; W; ρ) and the induced data (∆ p , W p ).
In this sense, we may also call these periods as the periods for the root system
(V, (·, ·); Φ, Φ± , ∆; W; ρ) and (∆ p , W p ), respectively. We claim that more general
type periods can be introduced for what we call pseudo root systems.
Definition 12.11. By a pseudo root system (Φ, ∆), we mean the data
(V, (·, ·); Φ, Φ± , ∆; W; ρ) consisting of a metrized R-vector space (V, h·, ·i), four fi-
nite sets Φ, Φ± , ∆, a finite Coxeter group W and a vector ρ ∈ V, satisfying
X 1X
(1) Φ = Φ+ t Φ− , ∆ is a base of V, Φ− = −Φ+ ⊂ Z≥0 α, ρ = α and
α∈∆
2 α>0
W := h σα : α ∈ Φ+ i = h σα : α ∈ ∆ i and W(Φ) = Φ
(2) there is a decomposition Φ = tm Φ
i=1 i such that (V i , (·, ·) i , Φi , ∆i , Wi ) satisfies
the condition (1) with Vi ⊥ V j (i , j), V = ⊕i=1 Vi , W(Φ) = m
m Q
i=1 W(Φi ), and
the data (Φ+i = Φ+ ∩ Φi , ∆i = ∆ ∩ Φi ) is reduced in an obvious sense.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 336
Example 12.1. (1) Recall that a finite set Φ is called a root system if
(a) Φ is a finite subset in a metrized R-vector space, and does not contain 0;
(b) For α ∈ Φ, Zα ∩ Φ = {±α};
(c) For α ∈ Φ, σα (Φ) = Φ; and
2
(d) h β, α i := (β, α∨ ) = (β, α) ∈ Z for all α, β ∈ Φ.
(α, α)
Hence a restricted pseudo root system is a root system. Moreover, for a Φ satis-
fying (a, c, d) but (b), then, by Exercise 9 of §9.4 in [44], for α ∈ Φ, among all
multiples of α, only ± 21 α, ±α, ±2α ∈ Φ. These Φ’s are clearly pseudo root sys-
tems. For a concrete example of pseudo root system which is not a root system,
please see the same exercise cited above.
(2) Let (V, (·, ·); Φ, ∆) be a root system and write ∆ = α1 , α2 , . . . , αn . For each
fixed 1 ≤ p ≤ n, set ∆ p = {α1 , . . . , αbp , . . . , αn and let Φ p be the subset of Φ
consisting of the roots which are perpendicular to the p-th fundamental dominate
weight λ p . Set Φ±p := Φ± ∩ Φ p . Then (Φ p , Φ±p , ∆ p ) is certainly a restricted pseudo
root system, hence a (sub) root system (of Φ).
(3) Let (V, (·, ·); Φ, ∆) be a root system and write ∆ = α1 , α2 , . . . , αn . For each
fixed 1 ≤ p ≤ n, set ∆ p = {α1 , . . . , αbp , . . . , αn and denote by H p := α⊥p ⊂ V be the
cxdimension one hyperplane in V which is perpendicular to λ p . Obviously, ∆ p is
a R-basis of H p . Denote by Φ fp , resp. Φ f±p , be the collection of the projections of
the elements of Φ, resp. Φ , on H p . Then (Φ
± fp , Φ
f±p , ∆ p ) is a pseudo root system.
σ∈∆
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 337
Proposition 12.11. Let (Φ, ∆) be a pseudo root system of rank n and fix 1 ≤ p ≤ n.
Proof. (1) comes directly from Proposition 12.10. Moreover the proof there is
obviously valid for (2) as well.
The residue (2) in Proposition 12.11 above is closely related with the leading coef-
ficient of the polynomial a p (s) in (12.31) of §12.3.2. Recall that, for each w ∈ W 0 ,
the invariant d p,w is defined by
1 −1 +
Y
:= 2|∆ p ∩w (Φ )| hρ, α∨ i − 1
d p,w
α∈w−1 (∆)∩(Φ p r∆ p )
Y (12.94)
hρ, α∨ i + δα,w hρ, α∨ i + δα,w − 1 .
×
α∈Φ+p r∆ p
Proposition 12.12. Let (Φ, ∆) be a rank n root system. Then for each 1 ≤ p ≤ n,
(Φ ,∆ ) 1 X 1
Resλ=ρ p ωQ p p (λ) = Q · c p,w d p,w ,
p αw i
, ∨
α∈Φ+ r∆ ζ(hρ, α i + 1)
b ∨
w∈W
hλ
p p p
|w−1 (∆)rΦ p |=1
w∈W p
hλ p , α∨w i p,w w∈W p
hλ p , α∨w i
|w−1 (∆)rΦ p |=1 ∆ p ⊂w−1 (∆ p ∪Φ−p )
Y 1 Y Y
× ζ(hρ∨ , αi+1)
b ζ(hρ∨ , αi).
b
hρ p , α∨ i−1
α∈w−1 (∆)∩(Φ p r∆ p ) α∈(Φ+p r∆ p )∩w−1 (Φ+ ) α∈(Φ+p r∆ p )∩w−1 (Φ− )
Furthermore, for w ∈ W p , easily,
w−1 (∆) ∩ (Φ p r ∆ p ) = w−1 (∆ p ) r ∆ p and ∆ p ∩ w−1 (∆) = ∆ p ∩ w−1 (∆ p ).
Also, for w ∈ W p , by the facts that,
(Φ+p r ∆ p ) ∩ w−1 (Φ± ) = (Φ+p r∆ p ) ∩ w−1 (Φ±p ) ∪ (Φ+p r∆ p )∩w−1 (Φ± rΦ±p )
Therefore,
−1
X 1 X ζ(2)−|∆ p ∩w (∆ p )|
=
b
C p,w D p,w
w∈W
hλ p , α∨w i w∈W
hλ p , w−1 (α∨p )i
p p
|w−1 (∆)rΦ p |=1 ∆ p ⊂w−1 (∆ p ∪Φ−p )
Y 1 Y Y
× ζ(hρ∨ , α∨ i+1)
b ζ(hρ∨ , αi)
b
hρ p , α∨ i−1
α∈w−1 (∆ p )r∆ p α∈(Φ+p r∆ p )∩w−1 (Φ+p ) α∈(Φ+p r∆ p )∩w−1 (Φ−p )
X −1
Y 1
= ζ(2)−|∆ p ∩w (∆ p )|
b
hρ p , α∨ i − 1
w∈W p α∈w−1 (∆ p )r∆ p
∆ p ⊂w−1 (∆ p ∪Φ−p )
Y Y
× ζ(hρ∨ , α∨ i + 1)
b ζ(hρ∨ , αi).
b
α∈(Φ+p r∆ p )∩w−1 (Φ+p ) α∈(Φ+p r∆ p )∩w−1 (Φ−p )
Here, in the last step, we have used relations hλ p , w−1 (α∨p )i = hw(λ p ), α∨p i = 1.
This then proves the proposition by using (12.93).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 339
We use the same numbering as in [49]. We write α∨ = ni=1 ai α∨i as the sequence
P
a1 , . . . , an with a p = 1. Then, same as in [56], we have:
Type An
By the symmetry of root system of An type, it suffices to consider the case
Pj
when 1 ≤ p ≤ [n/2] + 1. Let α∨i, j := k=i αk . Then for 1 ≤ j ≤ p, we have
n o
L j (1)∨ = α∨j, n , α∨j, n−1 , . . . , α∨j, p+ j−1 , α∨j+1, p+ j−1 , . . . , α∨p, p+ j−1 . (12.96)
• p=1
L1 (1)∨ : 111111 111110 111100 111000 110000 100000
• p=2
L1 (1)∨ : 111111 111110 111100 111000 110000 010000
L2 (1)∨ : 011111 011110 011100 011000
• p=3
L1 (1)∨ : 111111 111110 111100 111000 011000 001000
L2 (1)∨ : 011111 011110 011100 001100
L3 (1)∨ : 001111 001110
• p=4
L1 (1)∨ : 111111 111110 111100 011100 001100 000100
L2 (1)∨ : 011111 011110 001110 000110
L3 (1)∨ : 001111 000111
• p=5
L1 (1)∨ : 111111 111110 011110 001110 000110 000010
L2 (1)∨ : 011111 001111 000111 000011
• p=6
L1 (1)∨ : 111111 011111 001111 000111 000011 000001
Type Bn
Let
j
X j
X k
X
α∨i, j := α∨i and β∨i, j, k := α∨i + 2 α∨i .
l=i l=i l= j+1
Then, for 1 ≤ j ≤ min{p, n − p + 1}, let
n
L j (1)∨ = β∨j, p, n , β∨j, p+1, n , . . . , β∨j, n−1, n ,
o
α∨j, n , α∨j, n−1 , . . . , α∨j, p+ j−1 , α∨j+1, p+ j−1 , . . . , α∨p, p+ j−1 .
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 340
• p=1
L1 (1)∨ : 122222 112222 111222 111122 111112 111111 111110 111100
111000 110000 100000
• p=2
L1 (1)∨ : 112222 111222 111122 111112 111111 111110 111100 111000
110000 010000
L2 (1)∨ : 012222 011222 011122 011112 011111 011110 011100 011000
• p=3
L1 (1)∨ : 111222 111122 111112 111111 111110 111100 111000 011000
001000
L2 (1)∨ : 011222 011122 011112 011111 011110 011100 001100
L3 (1)∨ : 001222 001122 001112 001111 001110
• p=4
L1 (1)∨ : 111122 111112 111111 111110 111100 011100 001100 000100
L2 (1)∨ : 011122 011112 011111 011110 001110 000110
L3 (1)∨ : 001122 001112 001111 000111
L4 (1)∨ : 000122 000112
• p=5
L1 (1)∨ : 111112 111111 111110 011110 001110 000110 000010
L2 (1)∨ : 011112 011111 001111 000111 000011
L3 (1)∨ : 001112 000112 000012
• p=6
L1 (1)∨ : 111111 011111 001111 000111 000011 000001
Type Cn
j
X j
X n−1
X
Let α∨i, j := α∨i and β∨i, j := α∨i + 2 α∨i + α∨n .
l=i l=i l= j+1
Then, for 1 ≤ j ≤ min{p, n − p + 1},
n
L j (1)∨ = β∨j, p , β∨j, p+1 , . . . , β∨j, n−2 ,
o
α∨j, n , α∨j, n−1 , . . . , α∨j, p+ j−1 , α∨j+1, p+ j−1 , . . . , α∨p, p+ j−1 .
Moreover, if n − p + 1 < p ≤ n − 1, then, for n − p + 1 < j ≤ min{p, 2n − 2p},
n o
L j (1)∨ = β∨j, p , β∨j, p+1 , . . . , β∨j, 2n−p− j , β∨j+1, 2n−p− j , . . . , β∨p, 2n−p− j .
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 341
• p=1
L1 (1)∨ : 122221 112221 111221 111121 111111 111110 111100 111000
110000 100000
• p=2
L1 (1)∨ : 112221 111221 111121 111111 111110 111100 111000 110000
010000
L2 (1)∨ : 012221 011221 011121 011111 011110 011100 011000
• p=3
L1 (1)∨ : 111221 111121 111111 111110 111100 111000 011000 001000
L2 (1)∨ : 011221 011121 011111 011110 011100 001100
L3 (1)∨ : 001221 001121 001111 001110
• p=4
L1 (1)∨ : 111121 111111 111110 111100 011100 001100 000100
L2 (1)∨ : 011121 011111 011110 001110 000110
L3 (1)∨ : 001121 001111 000111
L4 (1)∨ : 000121
• p=5
L1 (1)∨ : 111111 111110 011110 001110 000110 000010
L2 (1)∨ : 011111 001111 000111 000011
• p=6
L1 (1)∨ : 222221 122221 112221 111221 111121 111111 011111 001111
000111 000011 000001
L2 (1)∨ : 022221 012221 011221 011121 001121 000121 000021
L3 (1)∨ : 002221 001221 000221
Type Dn
Let
j
X j
X n−2
X n−2
X
α∨i, j := α∨i , β∨i, j := α∨i + 2 α∨i + α∨n−1 + α∨n , γi∨ := α∨i + α∨n .
l=i l=i l= j+1 l=i
Then, in the case 1 ≤ p ≤ n − 2, for 1 ≤ j ≤ min{p, n − p + 1},
n o
L j (1)∨ = β∨j, p , β∨j, p+1 , . . . , β∨j, n−2 , α∨j, n−1 , . . . , α∨j, p+ j−1 , α∨j+1, p+ j−1 , . . . , α∨p, p+ j−1 .
Moreover, if n − p + 1 < p ≤ n − 2, then, for n − p + 1 < j ≤ min{p, 2n − 2p − 1},
n o
L j (1)∨ = β∨j, p , β∨j, p+1 , . . . , β∨j, 2n−p− j−1 , β∨j+1, 2n−p− j−1 , . . . , β∨p, 2n−p− j−1 .
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 342
• p=1
L1 (1)∨ : 122211 112211 111211 111111 111110 111100 111000 110000
100000
L2 (1)∨ : 111101
• p=2
L1 (1)∨ : 112211 111211 111111 111110 111100 111000 110000 010000
L2 (1)∨ : 012211 011211 011111 011110 011100 011000
L3 (1)∨ : 111101 011101
• p=3
L1 (1)∨ : 111211 111111 111110 111100 111000 011000 001000
L2 (1)∨ : 011211 011111 011110 011100 001100
L3 (1)∨ : 001211 001111 001110
L4 (1)∨ : 111101 011101 001101
• p=4
L1 (1)∨ : 111111 111110 111100 011100 001100 000100
L2 (1)∨ : 011111 011110 001110 000110
L3 (1)∨ : 001111 000111
L4 (1)∨ : 111101 011101 001101 000101
• p=5
L1 (1)∨ : 122211 112211 111211 111111 111110 011110 001110 000110
000010
L2 (1)∨ : 012211 011211 011111 001111 000111
L3 (1)∨ : 001211
• p=6
L1 (1)∨ : 122211 112211 111211 111111 111101 011101 001101 000101
000001
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 343
Type E6
• p=1
L1 (1)∨ : 123212 123211 122211 122111 122101 112101 111101 111001
111000 110000 100000
L2 (1)∨ : 112211 112111 111111 111110 111100
• p=2
L1 (1)∨ : 112211 112111 111111 111110 111100 111000 110000 010000
L2 (1)∨ : 012211 012111 012101 011101 011100 011000
L3 (1)∨ : 112101 111101 111001 011001
L4 (1)∨ : 011111 011110
• p=3
L1 (1)∨ : 111111 111110 111100 111000 011000 001000
L2 (1)∨ : 111101 111001 011001 001001
L3 (1)∨ : 011111 011110 011100 001100
L4 (1)∨ : 011101 001101
L5 (1)∨ : 001111 001110
• p=4
L1 (1)∨ : 122111 122101 112101 111101 111100 011100 001100 000100
L2 (1)∨ : 112111 111111 111110 011110 001110 000110
L3 (1)∨ : 012111 012101 011101 001101
L4 (1)∨ : 011111 001111
• p=5
L1 (1)∨ : 123212 123211 122211 122111 112111 111111 111110 011110
001110 000110 000010
L2 (1)∨ : 112211 012211 012111 011111 001111
• p=6
L1 (1)∨ : 123211 122211 122111 122101 112101 111101 111001 011001
001001 000001
L2 (1)∨ : 112211 112111 111111 011111 001111 001101
L3 (1)∨ : 012211 012111 012101 011101
Type E7
• p=1
L1 (1)∨ : 1343212 1243212 1233212 1233211 1232211 1222211 1122211
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 344
Type E8
• p=1
L1 (1)∨ : 13456423 12456423 12356423 12346423 12345423 12345422
12345322 12344322 12334322 12334312 12234312 12234212 12233212
12223212 12223211 12222211 12222111 12222101 11222101 11122101
11112101 11111101 11111001 11111000 11110000 11100000 11000000
10000000
L2 (1)∨ : 12345323 12345313 12345312 12344312 12344212 12334212
12333212 12333211 12233211 11233211 11223211 11222211 11222111
11122111 11112111 11111111 11111110 11111100
L3 (1)∨ : 12234322 11234322 11234312 11234212 11233212 11223212
11123212 11123211 11122211 11112211
• p=2
L1 (1)∨ : 11234322 11234312 11234212 11233212 11223212 11223211
11222211 11222111 11222101 11122101 11112101 11111101 11111001
11111000 11110000 11100000 11000000 01000000
L2 (1)∨ : 01234322 01234312 01234212 01233212 01223212 01123212
01123211 01122211 01112211 01112111 01111111 01111110 01111100
01111000 01110000 01100000
L3 (1)∨ : 11233211 01233211 01223211 01222211 01222111 01222101
01122101 01112101 01111101 01111001
L4 (1)∨ : 11123212 11123211 11122211 11112211 11112111 11111111
11111110 11111100
L5 (1)∨ : 11122111 01122111
• p=3
L1 (1)∨ : 11123212 11123211 11122211 11122111 11122101 11112101
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 346
Type F4
• p=1
L1 (1)∨ : 1342 1242 1232 1231 1221 1220 1120 1110 1100 1000
L2 (1)∨ : 1222 1122 1121 1111
• p=2
L1 (1)∨ : 1122 1121 1120 1110 1100 0100
L2 (1)∨ : 0122 0121 0120 0110
L3 (1)∨ : 1111 0111
• p=3
L1 (1)∨ : 1111 1110 0110 0010
L2 (1)∨ : 0111 0011
L3 (1)∨ :
• p=4
L1 (1)∨ : 1231 1221 1121 1111 0111 0011 0001
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 348
L2 (1)∨ : 0121
Type G2
• p=1
L1 (1)∨ : 13 12 11 10
• p=2
L1 (1)∨ : 11 01
For simplicity, we work over C. Let G be a reductive group over C with P its
maximal parabolic subgroup. Then the quotient G/P is proper. Denote by λ p the
fundamental dominant weight associated to P, TXP the tangent bundle of G/P, and
Pic(G/P) the Picard group of G/P. Then by [102],
Pic(G/P) ' Zλ p and det(TX/P ) = c p Lλ p , (12.97)
where LλP denotes the ample line bundle generating Pic(G/P).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 349
Chapter 13
As we have seen in Theorem 8.6, the zeta functions for the exceptional group
G2 satisfy the Riemann hypothesis. The same method can also be applied to the
exceptional group F4 . In this chapter, we follow our work [133] to prove that all
the zeta functions for the exceptional groups of type E, namely E6 , E7 and E8 ,
satisfy a weak Riemann hypothesis.
Following the structures of the associated Weyl groups, to realize the set of simple
roots for the exceptional group of types En (n = 6, 7, 8), we choose a basis of R8
as follows.
α1 = (0, 0, 0, −1, 1, 0, 0, 0), α2 = (0, 0, −1, 1, 0, 0, 0, 0),
α3 = (0, −1, 1, 0, 0, 0, 0, 0), α4 = (1, 1, 0, 0, 0, 0, 0, 0),
1 (13.1)
α5 = (−1, 1, 0, 0, 0, 0, 0, 0), α6 = (1, −1, −1, −1, −1, −1, −1, 1),
2
α7 = (0, 0, 0, 0, −1, 1, 0, 0), α8 = (0, 0, 0, 0, 0, −1, 1, 0).
Accordingly, introduce the finite sets
∆6 = {α1 , α2 , α3 , α4 , α5 , α6 },
∆7 = {α1 , α2 , α3 , α4 , α5 , α6 , α7 }, (13.2)
∆8 = {α1 , α2 , α3 , α4 , α5 , α6 , α7 , α8 }.
It is not too difficult to calculate their related Cartan matrices and hence to confirm
that ∆n is indeed the set of the simple roots for the root system Φn associated to
the exceptional group En (n = 6, 7, 8), with the Dynkin diagrams given by
349
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 350
α4
E6 (13.3)
α1 α2 α3 α5 α6
α4
E7 (13.4)
α7 α1 α2 α3 α5 α6
α4
E8 (13.5)
α8 α7 α1 α2 α3 α5 α6
Note that the numbering here for the simple roots are very different from some
standard numbering used in [16] and [44]. We will use this numbering in this
chapter, for the purpose of writing down the Weyl groups of En more effectively.
α4
D5 (13.6)
α1 α2 α3 α5
Denote its associated roots system and the Weyl group by Φ5 and W5 , respec-
tively. We obtain two chains
and Signev consists of even sign changes for the first 5 coordinates. In particular,
W5 = S5 × 2(0)+(2)+(4) = 1920.
5 5 5
(− 2 , − 2 , − 2 , − 2 , 2 , − 2 , − 2 , 2 ), (− 12 , − 21 , − 12 , 12 , − 12 , − 12 , − 21 , 12 ),
1 1 1 1 1 1 1 1
1
, ,
1 1
, 1
, 1
, 1
, ,
1 1 1
, ,
1 1 1 1
, , , 1
, ,
1 1
(− 2 − 2 2 − 2 − 2 − 2 − 2 2 ), (− 2 − 2 2 2 2 − 2 − 2 2 ),
,
1 1
, 1
, 1
, 1
, 1
, ,
1 1
,
1 1
, ,
1 1 1
, , 1
, ,
1 1
(− 2 2 − 2 − 2 − 2 − 2 − 2 2 ), (− 2 2 − 2 2 2 − 2 − 2 2 ),
,
1 1 1
, , ,
1 1
, 1
, ,
1 1
,
1 1 1 1
, , , 1
, 1
, ,
1 1
+
(− − − − ), (− − − − ),
Φ6,5 = 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2
( 2 , − 2 , − 2 , − 2 , − 2 , − 2 , − 2 , 2 ), ( 2 , − 2 , − 2 , 2 , 2 , − 2 , − 2 , 2 ),
1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1
1
, ,
1 1
, ,
1 1
, 1
, ,
1 1 1
, ,
1 1 1
, , 1
, 1
, ,
1 1
( 2 − 2 2 − 2 2 − 2 − 2 2 ), ( 2 − 2 2 2 − 2 − 2 − 2 2 ),
,
1 1
, 1
, ,
1 1
, 1
, ,
1 1
,
1 1
, ,
1 1
, 1
, 1
, ,
1 1
( 2 2 − 2 − 2 2 − 2 − 2 2 ), ( 2 2 − 2 2 − 2 − 2 − 2 2 ),
( 1 , 1 , 1 , − 1 , − 1 , − 1 , − 1 , 1 ),
, ,
1 1 1 1 1
, , , 1
, ,
1 1
2 2 2 2 2 2 2 2 ( 2 2 2 2 2 − 2 − 2 2 )
(13.9)
which we denote by {α6,1 , α6,2 , . . . , α6,16 }, in the order of 1st left-1st right-2nd
left-2nd right, .... Denote by σ6,i (1 ≤ i ≤ 16) the reflections associated to α6,i .
Our first result, jointly with Katayama ([51]), is obtained using Mathematica.
Lemma 13.1. A set of complete representatives for W6 W5 can be chosen to be
Id, σ6,1 , σ6,2 , σ6,3 , σ6,4 , σ6,5 , σ6,6 , σ6,7 , σ6,8 , σ6,9 , σ6,10 ,
σ , σ , σ , σ , σ , σ ,
6,11 6,12 6,13 6,14 6,15 6,16
Σ6 = .
(13.10)
σ σ , σ σ , σ σ , σ σ , σ σ ,
6,1 6,8 6,1 6,12 6,1 6,14 6,1 6,15 6,1 6,16
σ σ , σ σ , σ σ , σ σ , σ σ
6,2 6,15 6,2 6,16 6,3 6,16 6,5 6,16 6,9 6,16
where τ6,i stands for the (i + 1)-st element in the display (13.10) for the set Σ6 .
Our calculations, using Mathematica, is not very difficult. On the other hand,
the choices we made in §§ 13.1.1 and 13.1.2 for the orders of the simple roots
proves to be rather crucial.
To construct the Weyl group W7 for the exceptional group of type E7 , we use
the Weyl group W6 of type E6 as presented in the previous lemma. Similarly, by
(13.7), W6 is a subgroup of W7 , it suffices to find the 56 complete representatives
for the cosets in W7 W6 .
For this, we first introduce a subset Φ+7,6 := {α ∈ Φ+7 : α < Φ+6 } consisting of
27 positive roots in Φ7 which do not belong to Φ6 . Namely,
(− 21 , − 12 , − 12 , − 12 , − 21 , 12 , − 21 , 12 ),
(−1, 0, 0, 0, 0, 1, 0, 0),
1
, 1
, ,
1 1 1 1
, , , ,
1 1 1
, ,
1 1
, ,
1 1 1
, , ,
1 1
(− 2 − 2 − 2 2 2 2 − 2 2 ), (− 2 − 2 2 − 2 2 2 − 2 2 ),
1
, ,
1 1 1
, , ,
1 1
, ,
1 1
,
1 1
, 1
, ,
1 1 1
, , ,
1 1
(− 2 − 2 2 2 − 2 2 − 2 2 ), (− 2 2 − 2 − 2 2 2 − 2 2 ),
,
1 1
, ,
1 1
, ,
1 1
, ,
1 1
,
1 1 1
, , 1
, ,
1 1
, ,
1 1
(− 2 2 − 2 2 − 2 2 − 2 2 ), (− 2 2 2 − 2 − 2 2 − 2 2 ),
, ,
1 1 1 1 1 1
, , , , ,
1 1
(− 2 2 2 2 2 2 − 2 2 ), (0, −1, 0, 0, 0, 1, 0, 0),
(0, 0, −1, 0, 0, 1, 0, 0), (0, 0, 0, −1, 0, 1, 0, 0),
+
(0, 0, 0, 0, −1, 1, 0, 0), (0, 0, 0, 0, 0, 0, −1, 1),
Φ7,6 = . (13.12)
(0, 0, 0, 0, 1, 1, 0, 0), (0, 0, 0, 1, 0, 1, 0, 0),
(0, 0, 1, 0, 0, 1, 0, 0), (0, 1, 0, 0, 0, 1, 0, 0),
1
, 1
, 1
, ,
1 1 1
, , ,
1 1 1
, 1
, ,
1 1
, ,
1 1
, ,
1 1
( 2 − 2 − 2 − 2 2 2 − 2 2 ), ( 2 − 2 − 2 2 − 2 2 − 2 2 ),
1
, 1
, 1
, 1
, 1
, 1
, 1
, 1 1
, 1
, 1
, 1
, 1
, 1
, 1
, 1
( 2 − 2 2 − 2 − 2 2 − 2 2 ), ( 2 − 2 2 2 2 2 − 2 2 ),
1
, 1
, 1
, 1
, 1
, 1
, 1
, 1 1
, 1
, 1
, 1
, 1
, 1
, 1
, 1
( 2 2 − 2 − 2 − 2 2 − 2 2 ), ( 2 2 − 2 2 2 2 − 2 2 ),
1
, 1
, 1
, 1
, 1
, 1
, 1
, 1 1
, 1
, 1
, 1
, 1
, 1
, 1
, 1
( 2 2 2 − 2 2 2 − 2 2 ), ( 2 2 2 2 − 2 2 − 2 2 ),
(1, 0, 0, 0, 0, 1, 0, 0)
We denote them, with above displayed order, by {α7,1 , α7,2 , . . . , α7,27 }, and denote
by σ7, j the reflection associated to α7, j (1 ≤ j ≤ 27).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 353
Proposition 13.1. A set of complete representatives for W7 W6 can be chosen to
be
Id, σ7,1 , σ7,2 , σ7,3 , σ7,4 , σ7,5 , σ7,6 , σ7,7 , σ7,8 , σ7,9 , σ7,10 ,
σ , σ , σ , σ , σ , σ , σ , σ , σ , σ ,
7,11 7,12 7,13 7,14 7,15 7,16 7,17 7,18 7,19 7,20
σ , σ , σ , σ , σ , σ , σ ,
7,21 7,22 7,23 7,24 7,25 7,26 7,27
σ7,1 σ7,14 , σ7,1 σ7,19 , σ7,1 σ7,20 , σ7,1 σ7,21 , σ7,1 σ7,22 ,
Σ7 := σ σ , σ σ , σ σ , σ σ , σ σ , .
7,1 7,23 7,1 7,24 7,1 7,25 7,1 7,26 7,1 7,27
σ7,2 σ7,22 , σ7,2 σ7,24 , σ7,2 σ7,25 , σ7,2 σ7,26 , σ7,2 σ7,27 ,
σ σ , σ σ , σ σ ,
7,3 7,25 7,3 7,26 7,3 7,27
σ σ , σ σ , σ σ , σ σ , σ σ ,
7,4 7,26 7,4 7,27 7,5 7,27 7,6 7,26 7,6 7,27
σ σ , σ σ , σ σ , σ σ , σ σ σ
7,7 7,27 7,8 7,27 7,9 7,27 7,14 7,27 7,1 7,14 7,27
(13.13)
We denote them, with the above displayed order, by α8,1 , α8,2 , . . . , α8,57 , and
denote by σ8,k the reflection associated to α8,k (1 ≤ k ≤ 57). This then implies our
first main result on the construction of W8 .
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 355
Id, σ8,1 , σ8,2 , σ8,3 , σ8,4 , σ8,5 , σ8,6 , σ8,7 , σ8,8 , σ8,9 , σ8,10 ,
σ , σ , σ , σ , σ , σ , σ , σ , σ , σ ,
8,11 8,12 13 8,14 8,15 8,16 8,17 8,18 8,19 8,20
σ , σ , σ , σ , σ , σ , σ , σ , σ , σ ,
8,21 8,22 8,23 8,24 8,25 8,26 8,27 8,28 8,29 8,30
σ , σ , σ , σ , σ , σ , σ , σ , σ , σ ,
8,31 8,32 8,33 8,34 8,35 8,36 8,37 8,38 8,39 8,40
σ , σ , σ , σ , σ , σ , σ , σ , σ , σ ,
8,41 8,42 8,43 8,44 8,45 8,46 8,47 8,48 8,49 8,50
σ , σ , σ , σ , σ , σ , σ ,
8,51 8,52 8,53 8,54 8,55 8,56 57
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,1 8,20 8,1 8,22 8,1 8,24 8,1 8,26 8,1 8,28 8,1 8,29
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,1 8,31 8,1 8,33 8,1 8,35 8,1 8,37 8,1 8,39 8,1 8,40
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,1 8,41 8,1 8,42 8,1 8,43 8,1 8,44 8,1 8,45 8,1 8,46
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,1 8,47 8,1 8,48 8,1 8,49 8,1 8,50 8,1 8,51 8,1 8,52
σ σ , σ σ , σ σ , σ σ , σ σ ,
8,1 8,53 8,1 8,54 8,1 8,55 8,1 8,56 8,1 8,57
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,2 8,29 8,2 8,40 8,2 8,41 8,2 8,42 8,2 8,43 8,2 8,44
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,2 8,45 8,2 8,46 8,2 8,47 8,2 8,48 8,2 8,49 8,2 8,50
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,2 8,51 8,2 8,52 8,2 8,53 8,2 8,54 8,2 8,55 8,2 8,56 8,2 8,57
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,3 8,29 8,3 8,43 8,3 8,45 8,3 8,46 8,3 8,47 8,3 8,49 8,3 8,50
σ8,3 σ8,51 , σ8,3 σ8,52 , σ8,3 σ8,53 , σ8,3 σ8,54 , σ8,3 σ8,55 , σ8,3 σ8,56 , σ8,3 σ8,57 ,
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,4 8,29 8,4 8,46 8,4 8,47 8,4 8,50 8,4 8,51 8,4 8,52
σ σ , σ σ , σ σ , σ σ , σ σ ,
8,4 8,53 8,4 8,54 8,4 8,55 8,4 8,56 8,4 8,57
σ8,5 σ8,29 , σ8,5 σ8,47 , σ8,5 σ8,51 , σ8,5 σ8,52 , σ8,5 σ8,53 ,
.
σ8,5 σ8,54 , σ8,5 σ8,55 , σ8,5 σ8,56 , σ8,5 σ8,57 ,
σ8,6 σ8,29 , σ8,6 σ8,52 , σ8,6 σ8,53 , σ8,6 σ8,54 , σ8,6 σ8,55 , σ8,6 σ8,56 , σ8,6 σ8,57 ,
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,7 8,29 8,7 8,47 8,7 8,51 8,7 8,53 8,7 8,54 8,7 8,55 8,7 8,56
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
%
8,8 8,29 8,8 8,50 8,8 8,54 8,8 8,55 8,8 8,56 8,8 8,57 8,7 8,57
σ σ , σ σ , σ σ , σ σ , σ σ ,
8,9 8,29 8,9 8,49 8,9 8,55 8,9 8,56 8,9 8,57
σ8,10 σ8,29 , σ8,10 σ8,48 , σ8,10 σ8,56 , σ8,10 σ8,57 , & σ8,11 σ8,57 ,
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,11 8,29 8,11 8,47 8,11 8,51 8,11 8,53 8,11 8,54 8,11 8,55 8,11 8,56
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,12 8,29 8,12 8,46 8,12 8,54 8,12 8,55 8,12 8,56 8,12 8,57
σ σ , σ σ , σ σ , σ σ , σ σ ,
8,13 8,29 8,13 8,45 8,13 8,55 8,13 8,56 8,13 8,57
σ σ , σ σ , σ σ , σ σ ,
8,14 8,29 8,14 8,44 8,14 8,56 8,14 8,57
σ σ , σ σ , σ σ , σ σ , σ σ ,
8,15 8,29 8,15 8,43 8,15 8,55 8,15 8,56 8,15 8,57
σ σ , σ σ , σ σ , σ σ ,
8,16 8,29 8,16 8,42 8,16 8,56 8,16 8,57
σ σ , σ σ , σ σ , σ σ ,
8,17 8,29 8,17 8,41 8,17 8,56 8,17 8,57
σ σ , σ σ , σ σ , σ σ ,
8,18 8,29 8,18 8,40 8,18 8,56 8,18 8,57
σ σ , σ σ , σ σ , σ σ , σ σ ,
8,19 8,29 8,19 8,39 8,19 8,57 8,20 8,29 8,20 8,38
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
&
8,21 8,29 8,21 8,37 8,21 8,57 8,22 8,29 8,22 8,36 8,25 8,57
σ σ , σ σ , σ σ , σ σ , σ σ , σ σ , σ σ ,
8,23 8,29 8,23 8,35 8,23 8,57 8,24 8,29 8,24 8,34 8,25 8,29 8,25 8,33
σ σ , σ σ , σ σ , σ σ , σ σ ,
8,26 8,29 8,26 8,32 8,27 8,29 8,27 8,31 8,27 8,57
σ σ ,σ σ ,σ σ ,σ σ ,σ σ ,σ σ ,σ σ
8,28 8,29 8,28 8,30 8,30 8,57 8,32 8,57 8,34 8,57 8,36 8,57 8,38 8,57
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 356
Theorem 13.2. With the same notation as above, the Weyl groups for the excep-
tional groups of types E6 , E7 , E8 are given respectively by
G26
W6 = τ6,i W5 ,
i=0
55 G
G 26
W7 = τ7, j τ6,i W5 , (13.18)
j=0 i=0
239 G
G 55 G
26
W8 = τ8,k τ7, j τ6,i W5 .
k=0 j=0 i=0
As mentioned above, for this theorem to be hold, the special orders for the
simple roots made in §13.1.1 and §13.1.2 play an extremely crucial role.
Our calculations of W8 in the winter 2015 and the spring of 2016 were rather
long and complicated, with hundreds bugs searching and programs fixing. In
addition, there was a huge problem about how to store the W8 data used in our
Mathematica programs with a Mac Power. This forced our work stations stoping
calculations repeatedly. At some point, we realized the main reason causing all
these stops was merely the lack of storage. Soon after, we offered a not-perfect-
but-practical solution. To be more precisely, instead, we introduced
G239 G
55
W8,i := τ8,k τ7, j τ6,i W5 , (0 ≤ i ≤ 26) (13.19)
k=0 j=0
and hence divided our calculations into 27 parallel units mentioned. Consequently,
for W8 , we were able to start our second phrase of calculations on zeta functions
themselves and on the verifications of the weak Riemann Hypothesis for these
Weng zetas associated to the exceptional groups of types E6 , E7 , E8 . This lasted
for another 2 months. One of the up-shot is the above theorem.
and define X
kP (w) := (1 − δα.w ) = (Φ+ r Φ+P ) ∩ w−1 (Φ− ) ,
α∈Φ+ rΦ+P (13.23)
W p := w ∈ WP : kP (w) = 0 .
1 −1
Φ+ |
Y
:= 2|∆P ∩w (hρ, α∨ i − 1) (13.24)
dP,w
α∈(w−1 ∆)∩(Φ P r∆P )
Y
× (hρ, α∨ i + δα,w )(hρ, α∨ i + δα,w − 1).
α∈ΦP r∆P
where αw is the only element of w−1 (∆) r ΦP . By Theorem 12.1, to show that
ζQG,P (s), it suffices to verify that
the weak Riemann Hypothesis holds for b
∆G,P
Q > 0. (13.26)
Set, accordingly,
−1
Φ+ |
Y 1 Y
ζ(2)|∆P ∩w
BP,w := b ζ(hρ, α∨ i + δα,w ).
b
hρ, α∨ i − 1 α∈Φ+ r∆
α∈(w−1 ∆)∩(ΦP r∆P ) P P
(13.27)
By (12.95), we have
X BP,w
∆G,P
Q = . (13.28)
w∈WP
hλP , α∨w i
|(w−1 ∆)rΦP |=1
Here λ1 , λ2 , . . . , λn is the dual bases of {α∨1 , α∨2 , . . . , α∨n } for n = 6, 7, 8.
December 26, 2017 9:48 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 358
WP† := w ∈ W sp : |∆ r wΦP | = 1 ,
(13.29)
WP‡ := w ∈ W † : |w(Φ+ r Φ+P ) ∩ Φ− | = 0 .
Aw t Bw = ∆P ⇐⇒ ∆P ⊂ w−1 (∆ t Φ− ). (13.30)
WP† = w ∈ W sp : |∆ ∩ wΦP | = n − 1 .
(13.31)
To prove ∆QEn ,P , 0 for every n = 6, 7, 8 and for all the associated maximal
parabolic subgroups, we first search on the selections of Weyl elements belonging
to W sp , W † and W ‡ respectively. Accordingly, we carry out our calculations in
Mathematica programs into three steps.
sp
Step 1. Calculate special Weyl elements in WP , denoted as ‘sp’ in the program.
This takes very very long time.
Step 2. Calculate elements of WP† , denoted as ‘sps’ in the program. This takes
rather long time. Finally,
Step 3. Calculate elements of WP‡ , called very special in the program. This is very
pleasant as we have the following very beautiful
Example 13.1. When (G, P) = (E8 , P1 ), for simplicity, we denote the Weyl ele-
ment w5,s τ6,i τ7, j τ8,k simply by (s + 1, i + 1, j + 1, k + 1), where w5,s ∈ W5 is with
the index s predesigned by our program. Then, from our Mathematica program
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 359
Very special Weyl elements are these which appeared in WP‡ giving non-trivial
contributions to ∆G,PQ . To calculate ∆Q for G = E 6,7,8 , we here calculate the con-
G,P
stant hλP , α∨w i. Here, for each very special Weyl element w, αw is the only element
in w−1 (∆)rΦP . Our calculation using Mathematica confirms the following known
result.
Proof. Fix a very special Weyl element w ∈ WP‡ . Note that h·, ·i is w-invariant, it
is sufficient to calculate hw(λP ), β∨w i where βw is the unique element in ∆ r w(ΦP ).
Our calculation then shows that
‡
Lemma 13.2. For all w ∈ Wn,P , when applicable,
w(λ1 ) = (0, 0, 0, 0, 1, 1, 1, 3), w(λ2 ) = (0, 0, 0, 1, 1, 1, 1, 4),
1 1 1 1 1 1 1 5
w(λ3 ) = (0, 0, 1, 1, 1, 1, 1, 5), w(λ4 ) = ( , , , , , , , ),
2 2 2 2 2 2 2 2
(13.34)
1 1 1 1 1 1 1 7
w(λ5 ) = (− , , , , , , , ), w(λ6 ) = (0, 0, 0, 0, 0, 0, 0, 2),
2 2 2 2 2 2 2 2
w(λ7 ) = (0, 0, 0, 0, 0, 1, 1, 2), w(λ8 ) = (0, 0, 0, 0, 0, 0, 1, 1).
BP,w
Recall that ∆QEn ,P =
X
where
w∈WP
hλP , α∨w i
|(w−1 ∆)rΦP |=1
−1
Φ+ |
Y 1 Y
ζ(2)|∆P ∩w
BP,w := b ζ(hρ, α∨ i + δα,w ).
b
hρ, α∨ i − 1 α∈Φ+ r∆
α∈(w−1 ∆)∩(ΦP r∆P ) P P
(13.35)
By Proposition 13.3, hλP , α∨w i = 1. Hence, only BP,w becomes essential now. Thus
to calculate the constants ∆QEn ,P (n = 6, 7, 8) for all associated maximal parabolic
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 361
We have calculated all these. The reader may find some details in our website.
The up-shot is the following explicit formulas for the discriminants ∆QEn ,P (n =
6, 7, 8), obtained using Mathematica.
(E8 , P1 )
1 5 4 3 2
− ζ(2) b
b ζ(3) (3 + 4b
ζ(2)(−2 + 3b
ζ(3)))b
ζ(4) bζ(5) bζ(6)b
ζ(7)
3628800
3 2
(9450 + 5040b
ζ(2) (−9 + 10b ζ(3)) + 60b
ζ(2) (1435 + 72b ζ(3)(−21 + 10b
ζ(4)))
+ 7b
ζ(2)(27b
ζ(3)(225 + 16b
ζ(4)(−19 + 25b
ζ(5)))
+ 200(−37 + 27b
ζ(4)b
ζ(6)(b
ζ(4) − 8b
ζ(5)b
ζ(8)))))
(E8 , P2 )
1 b 5b 3b 2
ζ(2) ζ(3) ζ(4)(9 + 18b
ζ(2) − 4bζ(2)(8 + 9b
ζ(3)(−1 + 2b
ζ(4))))
51840
(−45 + b
ζ(2)(200 + 8b
ζ(2)(−25 + 36bζ(3)) − 9b
ζ(3)(25 + 16b
ζ(4)(−2 + 5b
ζ(5)))))
(E8 , P3 )
1 b 4 2
− ζ(2) (−1 + 2b
ζ(2))b
ζ(3) (3 + 4b
ζ(2)(−2 + 3b
ζ(3)))b
ζ(4)
17280
(−45 + b
ζ(2)(200 + 8b
ζ(2)(−25 + 36bζ(3)) − 9b
ζ(3)(25 + 16b
ζ(4)(−2 + 5b
ζ(5)))))
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 362
(E8 , P4 )
1 6 5 4 3 2
− ζ(2) b
b ζ(3) bζ(4) b ζ(5) bζ(6) bζ(7)
12700800
3
(99225 + 2b
ζ(2)(99225b ζ(2) + 44100(−8 + 9b ζ(3))
2
ζ(2) (425 + 18b
− 1568b ζ(3)(−46 + 24b
ζ(3) + 25b
ζ(4)))
2
+ 42b
ζ(2)(19600 + 45b
ζ(3) (385 + 12b
ζ(4)(−64 + 35b
ζ(4) + 70b
ζ(5)))
+ 24b
ζ(3)(−1540 + b
ζ(4)(1568 + 25b
ζ(5)(−64 + 63b
ζ(6)))))
ζ(3)b
− 144b ζ(4)(3528 + 25b
ζ(5)(−196 + 9b
ζ(6)(32 + 49b
ζ(7)(−1 + 4b
ζ(8)))))))
(E8 , P5 )
1 5 4 3 2
ζ(2) (−1 + 2b
b ζ(2))b
ζ(3) bζ(4) bζ(5) bζ(6)
1814400
2
(14175 + b
ζ(2)(11025(−8 + 9bζ(3)) + 4(882b ζ(2) (−25 + 36b
ζ(3))
+ 5b
ζ(2)(8330 − 9b
ζ(3)(15b
ζ(3)(−49 + 96b
ζ(4)) + 2(833 + 16b
ζ(4)(−49 + 45b
ζ(5)))))
+ 36b
ζ(3)b
ζ(4)(−882 + 25b
ζ(5)(49 + 36b
ζ(6)(−2 + 7b
ζ(7)))))))
(E8 , P6 )
1 6 6 5 5 4 3 2 2
ζ(2) b
b ζ(3) bζ(4) bζ(5) bζ(6) bζ(7) bζ(8) bζ(9) bζ(10)bζ(11)
4191264000
4 3
ζ(2) (−215+264b
(1164240b ζ(3))+41580b ζ(2) (14980 + 3b ζ(3)(−6139+3840b
ζ(4)))
2
− 10914750(3 + 16b
ζ(3) bζ(4)b
ζ(5))
2 2
+ 140b
ζ(2) (−14850b
ζ(3) (126 + b
ζ(4)(−387 + 224b
ζ(4) + 252b
ζ(5)))
ζ(3)(−58135 + 48b
− 99b ζ(4)(1487 − 225b
ζ(5) + 630b
ζ(4)b
ζ(6)))
+ 175(−24431 + 72b
ζ(4)b
ζ(6)(187b
ζ(4) − 216b
ζ(5)b
ζ(8))))
2
+ 99b
ζ(2)(6048000b
ζ(3) bζ(4)b
ζ(5)
+b
ζ(3)(−2149875 + 16b
ζ(4)(202419 − 25b
ζ(5)(2695 + 72b
ζ(6)(−72 + 245b
ζ(7)))))
− 9800(−245 + 9b
ζ(4)b
ζ(6)(15b
ζ(4)
ζ(6)b
− 8(−5b ζ(9)b
ζ(10) + 3b
ζ(8)(b
ζ(5) + 20b
ζ(7)b
ζ(10)b
ζ(12)))))))
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 363
(E8 , P7 )
1 5 5 5 4 3 3 2
− bζ(2) (−1 + 2b
ζ(2))b
ζ(3) bζ(4) bζ(5) bζ(6) bζ(7) bζ(8) bζ(9)b
ζ(10)b
ζ(11)
6652800
3
(−51975 + 44bζ(2) (10255 + 48b
ζ(3)(−448 + 225b ζ(3)))
2
ζ(2) (1540 + 3b
− 440b ζ(3)(−819 + 210b
ζ(3) + 8b
ζ(4)(39 − 70b
ζ(5))))
+ 7b
ζ(2)(47300 + 9(b
ζ(3)(−4950 + 32b
ζ(4)(132 + 25b
ζ(5)(−11 + 6b
ζ(6))))
2
ζ(6)(3b
− 275b ζ(4) − 32b
ζ(4)b
ζ(5)b
ζ(8) + 192b
ζ(5)b
ζ(8)b
ζ(9)b
ζ(12)))))
(E8 , P8 )
1 6 6 6 6 5 5 4
ζ(2) b
b ζ(3) b ζ(4) bζ(5) bζ(6) bζ(7) bζ(8)
22054032000
4 3 3 2 2
ζ(9) b
b ζ(10) bζ(11) b ζ(12) b ζ(13) b ζ(14)bζ(15)bζ(16)bζ(17)
4
(−172297125 + 91891800b
ζ(2) (−9 + 10b
ζ(3))
3
ζ(2) (−122605 + 36b
−24310b ζ(3)(6153 − 3200b
ζ(4) + 150b
ζ(3)(−20 + 21b
ζ(4))))
2 2
+ 170b
ζ(2) (−1081080b
ζ(3) (10 + 3b
ζ(4)(−5 + 4b
ζ(5)))−
ζ(3)(−252175 + 16b
− 117b ζ(4)(13046 + 175b
ζ(5)(−55 + 36b
ζ(6))))
+ 3850(−4745 + 36b
ζ(4)b
ζ(6)(13b
ζ(4) − 72b
ζ(5)b
ζ(8))))
ζ(2)(51b
− 39b ζ(3)(606375 + 16b
ζ(4)(−40887 + 25b
ζ(5)(1925+
ζ(6)(−119 + 110b
18b ζ(7)))))
2
+ 1925(−16660 + 3b
ζ(6)(765b
ζ(4) + 288b
ζ(4)b
ζ(8)(−17b
ζ(5) + 20b
ζ(6)b
ζ(10))
+ 5440b
ζ(8)b
ζ(12)(b
ζ(5)b
ζ(9) − 18b
ζ(10)b
ζ(14)b
ζ(18))))))
(E7 , P1 )
1 b 4 4 3 2
− ζ(2) (−1 + 2b
ζ(2))b
ζ(3) bζ(4) bζ(5) bζ(6)bζ(7)
604800
3 2
ζ(2) (−9 + 10b
(5040b ζ(3)) + 378(25 − 32b ζ(3) b ζ(4))
2
+ 60b
ζ(2) (1435 + 72b
ζ(3)(−21 + 10b
ζ(4)))
+ 35b
ζ(2)(27b
ζ(3)(45 + 16b
ζ(4)(−3 + 5b
ζ(5))) + 40(−37
+ 27b
ζ(4)b
ζ(6)(b
ζ(4) − 8b
ζ(5)b
ζ(8)))))
(E7 , P2 )
1 b 4b 2
− ζ(2) ζ(3) (3 + 4b
ζ(2)(−2 + 3b
ζ(3)))b
ζ(4)
8640
(−45 + b
ζ(2)(200 + 8bζ(2)(−25 + 36b
ζ(3)) − 9b
ζ(3)(25 + 16b
ζ(4)(−2 + 5b
ζ(5)))))
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 364
(E7 , P3 )
1 b 3
− ζ(2) (−1 + 2bζ(2))b
ζ(3)(3 + 4b
ζ(2)(−2 + 3b
ζ(3)))
1728
2
(9 + 18b
ζ(2) − 4bζ(2)(8 + 9b
ζ(3)(−1 + 2b
ζ(4))))
(E7 , P4 )
1 b 5b 4b 3b 2b
ζ(2) ζ(3) ζ(4) ζ(5) ζ(6)
907200
2
(14175 + b ζ(2)(11025(−8 + 9bζ(3)) + 4(882b
ζ(2) (−25 + 36b
ζ(3))
+ 5b
ζ(2)(8330−9b
ζ(3)(15b
ζ(3)(−49+96b
ζ(4))+2(833 + 16b
ζ(4)(−49+45b
ζ(5)))))
+ 36b
ζ(3)b
ζ(4)(−882 + 25b
ζ(5)(49 + 36b
ζ(6)(−2 + 7b
ζ(7)))))))
(E7 , P5 )
1 b 4 3 2
ζ(2) (−1 + 2b
ζ(2))b
ζ(3) bζ(4) bζ(5)
43200
2
(−675 + 2b
ζ(2)(675b
ζ(2) − 40b ζ(2)(65 + 9bζ(3)(−12 + 5b
ζ(3) + 10b
ζ(4)))
+ 9(200 + b
ζ(3)(−225 + 16b
ζ(4)(18 + 25b
ζ(5)(−1 + 3b
ζ(6)))))))
(E7 , P6 )
1 5 5 4 4 2 2
ζ(2) b
b ζ(3) bζ(4) b ζ(5) bζ(6) bζ(7) bζ(8)b
ζ(9)
9072000
4 3
(141750 + 453600b ζ(2) − 1800b ζ(2) (833 + 15b ζ(3)(−35 + 48b
ζ(4)))
2
ζ(2) (−3773 + 18b
− 500b ζ(3)(245 − 328b
ζ(4) + 42b
ζ(3)(−2 + 3b
ζ(4)))
2
+ 2016b
ζ(4) bζ(6))
ζ(2)(43000 + b
− 21b ζ(3)(−37125 + 48b
ζ(4)(1197 + 25b
ζ(5)(−25 + 72b
ζ(6))))
ζ(4)b
− 5400b ζ(6)(5b
ζ(4) − 8b
ζ(8)(b
ζ(5) − 10b
ζ(6)b
ζ(10)))))
(E7 , P7 )
1 5 5 5 4 3 3 2
ζ(2) b
b ζ(3) bζ(4) bζ(5) bζ(6) bζ(7) bζ(8) bζ(9)b
ζ(10)b
ζ(11)
3326400
2
(51975 + b
ζ(2)(3850(−86 + 81b ζ(3)) − 44b ζ(2) (10255
+ 48b
ζ(3)(−448 + 225b
ζ(3)))
+ 440b
ζ(2)(1540 + 3b
ζ(3)(−819 + 210b
ζ(3) + 8b
ζ(4)(39 − 70b
ζ(5))))
+ 63(−32b
ζ(3)b
ζ(4)(132 + 25b
ζ(5)(−11 + 6b
ζ(6)))
2
+ 275b
ζ(6)(3b
ζ(4) − 32b
ζ(4)b
ζ(5)b
ζ(8) + 192b
ζ(5)b
ζ(8)b
ζ(9)b
ζ(12)))))
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 365
(E6 , P1 )
1 b 4b 4b 3b 2b b
− ζ(2) ζ(3) ζ(4) ζ(5) ζ(6)ζ(7)
302400
3 2
(9450 + 5040bζ(2) (−9 + 10b
ζ(3)) + 60b
ζ(2) (1435 + 72b
ζ(3)(−21 + 10b
ζ(4)))
+ 7b
ζ(2)(27b
ζ(3)(225 + 16b
ζ(4)(−19 + 25b
ζ(5)))
+ 200(−37 + 27b
ζ(4)b
ζ(6)(b
ζ(4) − 8b
ζ(5)b
ζ(8)))))
(E6 , P2 )
1 b 3 2
− ζ(2) (−1 + 2b
ζ(2))b
ζ(3) bζ(4)
1440
(−45 + b
ζ(2)(200 + 8b
ζ(2)(−25 + 36b ζ(3)) − 9b
ζ(3)(25 + 16b
ζ(4)(−2 + 5b
ζ(5)))))
(E6 , P3 )
1 b 2
ζ(2) (−1 + 2b
ζ(2))(3 + 4b
ζ(2)(−2 + 3b
ζ(3)))2
288
(E6 , P4 )
1 b 4b 3b 2b
ζ(2) ζ(3) ζ(4) ζ(5)
21600
2
(−675 + 2b
ζ(2)(675bζ(2) − 40b
ζ(2)(65 + 9b
ζ(3)(−12 + 5b
ζ(3) + 10b
ζ(4)))
+ 9(200 + b
ζ(3)(−225 + 16b
ζ(4)(18 + 25b
ζ(5)(−1 + 3b
ζ(6)))))))
(E6 , P5 )
1 b 3 2
− ζ(2) (−1 + 2b
ζ(2))b
ζ(3) bζ(4)
1440
(−45 + b
ζ(2)(200 + 8b
ζ(2)(−25 + 36b ζ(3)) − 9b
ζ(3)(25 + 16b
ζ(4)(−2 + 5b
ζ(5)))))
(E6 , P6 )
1 b 4b 4b 3b 2b b
− ζ(2) ζ(3) ζ(4) ζ(5) ζ(6)ζ(7)
302400
3 2
(9450 + 5040bζ(2) (−9 + 10b
ζ(3)) + 60b
ζ(2) (1435 + 72b
ζ(3)(−21 + 10b
ζ(4)))
+ 7b
ζ(2)(27b
ζ(3)(225 + 16b
ζ(4)(−19 + 25b
ζ(5)))
+ 200(−37 + 27b
ζ(4)b
ζ(6)(b
ζ(4) − 8b
ζ(5)b
ζ(8)))))
Corollary 13.1. For all n = 6, 7, 8 and for all their maximal parabolic subgroups
of En , the discriminants ∆QEn ,P are (strictly) positive.
December 26, 2017 9:55 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 366
Proof. This is a direct consequence of the theorem above. For example, with
(E8 , P1 ), we have
(π3 ζ(3)4 · (4π − 3(3 + ζ(3))) · ζ(5)2 · ζ(7))
∆QE8 ,P1 = −
75243740696496046080000000
· (−567000π3 (−63 + 2ζ(3)) + 7938000π(185 + 27ζ(3))
− 79380π2 (5125 + 136ζ(3)) + 8π8 (−35 + 12ζ(5))
− 4465125(360 + ζ(3)(135 + 2ζ(5)))).
Approximately,
206392.
∆QE8 ,P1 ≈ > 0. (13.36)
75243740696496046080000000
Our verifications for other cases are the same.
We mention in passing that, even positive, all the ∆QEn ,P ’s are indeed very very
small. This phenomenon is also observed for the case of S Ln .
PART 6
Stability plays a key role in various branches of mathematics and is used here to
complete a proof of the weak Riemann Hypothesis for our zeta functions. We
first show that, for special linear groups and more generally for reductive groups,
the geometric truncations of stability on the moduli spaces of arithmetic principal
torsors essentially coincide with Arthur’s analytic truncations (on the associated
adelic spaces). Then, we confirm the conjecture on Stability, Parabolic Reduction
and Masses, calculating the semi-stable masses (of arithmetic principal torsors)
in terms of parabolic reduction of stability, as a natural continuation of the works
of Atiyah-Bott, Harder-Narasimhan, Mumford in geometry and Arthur, Lafforgue
in arithmetic. Based on these, we establish the special uniformity of zeta func-
tions on the equivalence of the high rank non-abelian zeta functions and the Weng
zeta functions for special linear groups. Finally, we complete a proof of the weak
Riemann hypothesis for our zeta functions over the field of rationals, by showing
that their discriminants are positive, being the semi-stable masses of the arith-
metic principal torsors for which the canonical types coincide with the maximal
parabolic subgroups used. All the constructions and results of this part are new.
367
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 369
Chapter 14
Let F be a number field with OF its ring of integers. Denote by S = S ∞ t S fin the
collection of inequivalent places of F with S ∞ , resp. S fin , the set of infinite, resp.
finite, places. For v ∈ S , denote by Fv the v-completion of F. For example, if
v ∈ S ∞ , Fv = R or C. Accordingly, we call v real or complex. Denote by r1 , resp.
r2 , the number of real, resp. complex, places in S ∞ . In addition, for v ∈ S fin , Fv is a
non-Archimedean local field. Denote by Ov its ring of integers. Let A be the adelic
ring of F, and let GLn (A), resp. SLn (A), be the adelic group associated to GLn ,
Q Q Q
resp. SLn . Let K := v∈S fin GLn (Ov ) × v∈S ∞ On (R) × v∈S ∞ U n (C), resp.
Q Q Q v:real v:complex
SK := v∈S fin SLn (Ov ) × v∈S ∞ SOn (R) × v∈S ∞ SUn (C), be the maximal
v:real v:complex
compact subgroup of GLn (A), resp. of SLn (A). Here On , resp. Un , denotes the
orthogonal group, resp. the unitary group, of size n, and set SOn := SLn ∩ On and
SUn := SLn ∩ Un . Accordingly, we obtain the quotient spaces GLn (F)\GLn (A)/K
and SLn (F)\SLn (A)/SK. By §§1.4.1-2 and §6.2.3, we have
Proposition 14.1. There are natural identifications among the following spaces.
(1) The quotient space GLn (F)\GLn (A)/K, resp. SLn (F)\SLn (A)/SK.
(2) The quotient space GLn (OF )\GLn (R)r1 × GLn (C)r2 On (R)r1 × Un (C)r2 , resp.
SLn (OF )\SLn (R)r1 × SLn (C)r2 SOn (R)r1 × SUn (C)r2 .
(3) The moduli space Mtot tot
F,n , resp. MF,n [1], of OF -lattices of rank n, resp. of rank
n and volume one. And
(4) The moduli space Mtot OF ,n
, resp. MtotOF ,n
(d), of rank n metrized vector bundles
of all degrees, resp. of degree d, on X = Spec OF .
369
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 370
Proposition 14.2. Let g ∈ GLn (A) and let P be a parabolic subgroup of GLn .
(1) Among all the pgQ , where Q runs over all parabolic subgroup of GLn , there is a
unique maximal convex polygon pg . Moreover there exists a unique parabolic
g
subgroup Q such that pg g = pg .
Q
(2) Let d ∈ R. Forn p : [0, n] → R be a convex polygon satisfying
o p(0) = p(n) =
0. The space g ∈ GLn (F)\GLn (A) : degar g = d, pg ≤ p is compact. More
generally,
(3) Among all the pgQ , where Q runs over all parabolic subgroups contained in P
of GLn , there is a unique maximal element pgP . Moreover there exists a unique
g
parabolic subgroup QP ⊆ P of GLn such that pg g = pgP .
QP
n o
(4) The space g ∈ GLn (F)\GLn (A) : degar g = d, p̄gP ≤ p, pgP ≥ −p is compact.
Definition 14.1. Let p be a normalized convex polygon on [0, n]. The character-
istic function 1(p∗ ≤ p) and 1(p∗P ≤P p) associated to a parabolic subgroup P, on
GLn (F)\GLn (A)1 are defined by
g
g
1, if p ≤ p, 1, if pP ≤P p,
1(p ≤ p) = and 1(pP ≤P p) =
g
g
(14.1)
0, otherwise 0, otherwise
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 371
respectively. Here pg and pgP , denote the canonical polygon and the polygon asso-
ciated to P of g, respectively.
Remark 14.1. (1) This result and its proof below are arithmetic analogues of a
result and its proof obtained by Lafforgue ([63]) on vector bundles of curves over
finite fields.
(2) In algebraic geometry, or better in geometric invariant theory of Mumford
(see e.g. [80]), a fundamental principle claims that, for an action of a reductive
group over a space, if a point is not GIT stable, there should exist a parabolic
subgroup which destroys the corresponding stability. Our result above gives a
concrete realization of this principle in our arithmetic setting.
Proof. Since all parabolic subgroups of G which conjugate to P are exactly the
δ-conjugates of P, where δ ∈ P(F)\G(F), by the easy relation pδg g
P = pδPδ−1 , it
suffices to prove the following equivalence relation
The proof below is rather involved. To start with, we recall that, for a parabolic
subgroup P, by Proposition 14.2, there are a P-canonical polygon pgP of g and a
g g,P
P-canonical parabolic subgroup QP ⊆ P of G. Denote by Λ∗ the P-canonical
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 372
g
filtration of Λg induced by QP . Certainly, for different P and P0 , it may happen
g g
that QP = QP0 .
Lafforgue’s idea is to regroup the P’s in (14.4) according to their associated
P-canonical parabolic subgroups Q and then to show that for each fixed Q, the
summations on all involved P results exactly zero. Namely, after rewriting
g
where P runs over all parabolic subgroups of GLn satisfying QP = Q.
where Λg,P
j is the filtration of Λ induced from the parabolic subgroup P, and
g
g,P
Λ∗ is the P-canonical filtration associated with Λg .
Proof. The first inclusion is rather trivial. Indeed, for the canonical
g .filtration
g g g . g g
Λ∗ of Λ , for all 1 ≤ j ≤ |Q | − 1, Λ j+1 Λ j are semi-stable and µ Λ j Λ j−1 >
g
g . g nµ g o g
µ Λ j+1 Λ j . On the other hand, the filtration Λ∗ of Λg associated to µ Q is a
partial filtration of the canonical one, obtained
g . g by selecting
gonly
. gthe
special pieces
for which some stronger conditions µ Λ j Λ j−1 > µ Λ j+1 Λ j + µ should be
g g
satisfied. Therefore, Q ⊆ µ Q .
g
Next we show that µ QP ⊆ P. This is simple, since by our definition, the
µ g
filtration {µ Λg,P
∗ } associated to QP is a refinement of the filtration of Λ associated
g
to P.
g g
Finally, we prove QP ⊆ µ QP . This also quite direct. Indeed, by definition,
g,P
for the P-canonical filtration Λ∗ of Λg , the P-Harder-Narasimhan conditions
g,P . g,P g,P . g,P g,P . g,P
hold. Namely, Λ j Λ j−1 are semi-stable and µ Λ j Λ j−1 > µ Λ j+1 Λ j for
all the j’s, except for possibly the indices j induced directly from P. But up to
g
the indices j induced from P, all other indices for µ QP is, by definition, character-
ized as the parts of the P-canonical filtration for which much stronger conditions
g,P . g,P g,P . g,P
µ Λ j Λ j−1 > µ Λ j+1 Λ j + µ are assumed to be satisfied.
characterized by
n o
P 7−→ I(P) := rkΛg,P
i : i = 1, 2, . . . , |P| − 1 (14.12)
is a well-defined bijection. Here P denotes the set of parabolic subgroups in GLn .
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 374
Proof. We first show that the assignment is well-defined, by examining what are
the conditions satisfied by P in the set
g g
P ∈ P : QP = Q1 , µ QP = Q2 , pgP >P p .
n o
(14.13)
g
Since the convex polygon pgQ1 of QP = Q1 associated to the filtration
{0} = Λg,Q
0
1
⊂ Λg,Q
1
1
⊂ Λ2g,Q1 ⊂ . . . ⊂ Λ|Q
g,Q1
1 |−1
⊂ Λg,Q
|Q1 |
1
= Λg
n g o
is the unique maximal element among all pQ : Q ⊆ P , Q1 ⊆ P, and except for
the indices associated to P,
g,Q1 . g,Q1 g,Q1 . g,Q1
µ Λk Λk−1 > µ Λk+1 Λk , (14.14)
g
since Q1 = QP is P-canonical. Therefore,
I(Q1 ) − J0 (Q1 ) ⊂ I(P). (14.15)
g
µ
In parallel, from QP = Q2 , by definition, the filtration
{0} = Λg,Q
0
2
⊂ Λg,Q
1
2
⊂ Λ2g,Q2 ⊂ . . . ⊂ Λ|Q
g,Q2
2 |−1
⊂ Λg,Q
|Q2 |
2
= Λg
coincides with the filtration
n g,P o n g,Q1 . g,Q1 g,Q1 . g,Q1 o
Λ j : j = 1, 2, . . . , |P| − 1 ∪ Λkg,Q1 : µ Λk Λk−1 > µ Λk+1 Λk +µ
Hence, by examining the ranks involved in all the filtrations, we have
I(Q2 ) = I(P) ∪ Jµ (Q1 ).
Consequently,
I(P) ⊂ I(Q2 ) and I(Q2 ) − Jµ (Q1 ) ⊂ I(P). (14.16)
In addition, by definition, the relation pgP >P p is equivalent to the following
group of inequalities
g
pP − p rkΛg,Pj > 0. (14.17)
Hence I(P) ⊆ K p (Q2 ) since I(P) ⊆ I(Q2 ). This together with (14.15) and (14.16)
implies that P 7→ I(P) is well-defined.
The injectivity is rather clear since with a fixed partition I, there exists only
one unique standard parabolic subgroup PI such that I(PI ) = ∆ − I.
n o
Finally, we verify the surjectivity. Let Σ be a subset of 1, 2, . . . , r satisfying
Σ ⊆ I(Q2 ), Jµ (Q1 ) ⊆ I(Q2 ), I(Q1 ) − J0 (Q1 ) ∪ I(Q2 ) − Jµ (Q1 ) ⊆ Σ ⊆ K p (Q2 )
Since there obviously exists a parabolic subgroup P = PΣ , it suffices to check
whether there exists parabolic subgroups Q1 ⊆ Q2 of GLn contained in P satisfy-
g g
ing the conditions that QP = Q1 , µ QP = Q2 and pgP >P p.
g
It is clear that pgP >P p since Σ ⊂ K p (Q2 ). On the other hand, QP = Q1 and
µ g
QP = Q2 are direct consequences of the definition
of P-canonical
filtration and
the conditions that I(Q1 ) − J0 (Q1 ) ⊆ Σ and I(Q2 ) − Jµ (Q1 ) ⊆ Σ as explained in
the proof above of the well-defined statement.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 375
We are now ready to complete our proof of Lemma 14.1 and hence the theo-
rem. Indeed, from Lemma 14.2,
To prove the theorem, assume otherwise that the value 1 does appear in the case
pg p. In particular,
I(Q1 ) − J0 (Q1 ) ∪ I(Q2 ) − Jµ (Q1 ) = K p (Q2 ). (14.18)
Let then µ = 0. Obviously, this implies that Q1 = Q2 = Q. Consequently,
(−1)|P| 1(pP >P p) = (−1)|P| 1(pP >P p)
X X
g g g g
P:QP =Q1 , µ QP =Q2 P:QP =Q= 0 QP
This is nothing but the summation appeared in the Proposition. Hence, by defi-
nition and the convexity property of p, we have, from (14.18), I(Q1 ) = J0 (Q1 ),
or better K pg (Q2 ) = ∅. Consequently, pg ≤ p, a contradiction. This completes the
proof of Lemma 14.1 and hence the theorem.
For simplicity, we, in this subsection, work only with Q, the field of rationals, and
use mixed languages of adeles and lattices. Without loss of generality, we assume
that all the Z-lattices here are of volume one. Accordingly, set G = SLn .
Lemma 14.3. Let T (p) be the character of P1,...,1 associated to a normalized con-
vex polygon p : [0, n] → R. Then
T (p) = p(1) − p(0), p(2) − p(1), . . . , p(i) − p(i − 1),
(14.20)
. . . , p(n − 1) − p(n − 2), p(n) − p(n − 1) .
pgP is affine on the closed intervals [ni , ni+1 ] and pgP (ni ) = degar Λg,P .
i
g
We can write down these values of pP more precisely. Indeed, if P = NP MP
denotes the Levi decomposition of P, and g = n · m · a(g) · k denotes the Iwasawa
decomposition of g with n ∈ NP (A), m ∈ MP (A)1 , k ∈ K := p SL(OQ p ) × SOn (R)
Q
and a = a(g) = diagar a1 In1 , a2 In2 , . . . , ak Ink ∈ A+ , we have, for all i = 1, . . . , k,
i
Y i
X
nj
degar Λg,P
i = − log a j = − n j log a j . (14.27)
j=1 j=1
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 377
pgP (n1 ) − p(n1 ) > 0, pgP (n2 ) − p(n2 ) > 0, . . . , pgP (nk−1 ) − p(nk−1 ) > 0.
Lemma 14.4. The condition 1(pgP >P p) = 1 is equivalent to the following system
of inequalities
− n1 log a1 − p(r1 ) > 0,
− n1 log a1 − n2 log a2 − p(n2 ) > 0,
(14.29)
... ... ... ...
− n1 log a1 − n2 log a2 − . . . − nk−1 log ak−1 − p(nk−1 ) > 0.
On the other hand, for H = H1 , H2 , . . . , Hr ∈ a0 , the projections of H to aI , resp.
a0I , is given by
(n1 ) (n2 ) N|P|
(n|P| )
n1 N2
1 1 X
X 1 X
H j , H j , . . . , H j ∈ aI
n1 n2 n|P|
j=N1 +1 j=N|P|−1 +1
j=1
resp.
N1 N1
H1 − 1
X 1 X H1 + H2 + . . . + Hr1
H j , H2 − H j , . . . , Hn1 − , ...,
n1 j=1 n1 j=1 d1
N|P| N|P| N|P|
1 X 1 X 1 X
HN|P|−1+1 − H j , HN|P|−1+2 − H j , . . . , HN|P| − H j
n|P| j=N +1 n|P| j=N +1 n|P| j=N +1
|P|−1 |P|−1 |P|−1
= H1 , H2 , . . . , Hr
(n1 ) (n2 ) N|P|
(n|P| )
n1 N2
1 1 X
X 1 X
− H j , H j , . . . , H j ∈ a0I .
n1 n2 j=N +1 n|P| j=N|P|−1 +1
j=1 1
H + H2 + . . . + HN1 > 0,
1
H1 + H2 + . . . + HN1 + HN1 +1 + HN1 +2 + . . . + HN2 > 0,
........................ ............
(14.31)
H1 + H2 + . . . + HN1 + HN1 +1 + HN1 +2 + . . . + HN2
+ . . . + HN +1 + HN +2 + . . . + HN > 0.
|P|−1 |P|−1 |P|−1
(n|P| )
) (n )
where a(g) = diag b1 , b2 , . . . , bn = diag a(n
1 , a2 , . . . , a|P|
1 2
and
T (p) = t1 , t2 , . . . , tn = p(1), p(2) − p(1), . . . , p(n − 1) − p(n − 2), −p(n − 1) .
As such, the system of inequalities (14.31), after replacing the Hi ’s with the pre-
cise coordinates of −H0 (g) − T (p), are then changed to the following system of
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 379
inequalities
−log b1 − p(1) + −log b2 − p(2)+ p(1) + . . . + −log bN1 − p(N1 )+ p(N1−1) > 0,
−log b1 − p(1) + −log b2 − p(2)+ p(1) + . . . + −log bN1 − p(N1 )+ p(N1−1)
+ − log bN1 +1 − p(N1 + 1) + p(N1 ) + − log bN1 +2 − p(N1 + 2) + p(N1 + 1)
+ . . . + − log bN2 − p(N2 ) + p(N2 − 1) > 0,
..........................................
+ + . . . +
−log b1 − p(1) −log b2 − p(2)+ p(1) −log b N1
− p(N1 )+ p(N 1−1)
+ + + + + + +
− log b N 1 +1 − p(N1 1) p(N 1 ) − log b N 1 +2 − p(N 1 2) p(N 1 1)
+ . . . + +
− log b N 2
− p(N2 ) p(N 2 − 1)
..........................................
+ − log bN|P|−2 +1 − p(N|P|−2 + 1) + p(N|P|−1 )
+ − log bN|P|−2 +2 − p(N|P|−2 + 2) + p(N|P|−2 + 1)
+ . . . + − log bN − p(N|P|−1 ) + p(N|P|−1 − 1) > 0.
|P|−1
Therefore, to complete our proof, it suffices to use Lemmas 14.4 and 14.5.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 380
We are finally ready to expose the following beautiful intrinsic relation between
geometric truncations using stability and analytic truncations using positivity.
We end this discussion by mentioning that the above theorem makes it possible
to identify the geometric Eisenstein periods and the analytic Eisenstein periods,
defined as the integrations of Eisenstein series over the moduli spaces of vector
bundles and the integrations of truncated Eisenstein series over suitable funda-
mental domains for SLn , respectively. In later chapters, we will develop a parallel
theory for general reductive groups.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 381
Chapter 15
15.1.1.1 Idempotence
Lemma 15.1. Let T ∈ C t {0} and let P be a parabolic subgroup of SLn . Then
ΛP,T ΛP,T 1 = ΛP,T 1,
(15.1)
ΛP,T (p) 1(p∗ ≤ p) = 1(p∗ ≤ p).
Proof. By Theorem 14.2, we may identify the geometric truncation 1(p∗ ≤ p) and
the analytic truncation ΛP,T 1, and hence it suffices to prove the second equality.
But for geometric truncations, this is simply a direct consequence of the proof of
Theorem 14.1. Indeed, for g ∈ SLn (A). when pgP ≤P p, by definition, the right
381
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 382
hand side is one, and the left hand side becomes ΛP,T (p) 1 (g) = 1(pg ≤ p) = 1 by
Theorem 14.2 again. On the other hand, if pgP P p, by the proof of Theorem 14.1,
the right hand side becomes 0, while the right hand side gives ΛP,T (p) 0, and hence
becomes 0 as well.
Next, we show that ΛP,T is self-adjoint with respect to natural inner product.
(1) Let φ1 and φ2 be two continuous functions on SLn (F)\SLn (A). Assume that
φ1 is slowly increasing, and that φ2 is rapidly decreasing. Then
(ΛT φ1 , φ2 ) = (φ1 , ΛT φ2 ). (15.2)
(2) Let P be a parabolic subgroup of SLn with Levi decomposition P = N M. Let
φ1 and φ2 be two continuous functions on P(F)N(A)\SLn (A). Assume that φ1
is slowly increasing, and that φ2 is rapidly decreasing. Then
(ΛP,T φ1 , φ2 ) = (φ1 , ΛP,T φ2 ). (15.3)
Proof. Since a similar proof of (1) below works for (2) as well, we only give the
details of a proof for (1). For this, we use the proof of the self adjoint property
in [3, 4]. Indeed, since the inner product (ΛT φ1 , φ2 ) is defined by the absolutely
convergent integrals, we have
Z X
(ΛT φ1 , φ2 ) = (−1)n−1−|P|
G(F)\G(A)1 P
X Z
× φ1 (nδg)b
τP (H(δg) − T )φ2 (g) dn dg
δ∈P(F)\G(F) N(F)\N(A)
X Z Z
= (−1)n−1−|P| φ1 (ng)φ2 (g)b
τP (H(g) − T ) dg dn
P N(F)\N(A) P(F)\SLn (A)
X Z Z (15.4)
= (−1)n−1−|P|
φ1 (g)φ2 (ng)b
τP (H(g) − T ) dg dn
N(F)\N(A) P(F)\SLn (A)
ZP X
= (−1)n−1−|P|
SLn (F)\SLn (A) P
X Z
× φ1 (g)b
τP (H(δg) − T )φ2 (nδg) dn dg.
δ∈P(F)\SLn (F) N(F)\N(A)
Lemma 15.3. Let T ∈ Ct{0} and denote by Σ(T ) ⊂ SLn (F)\SLn (A)/SK the com-
pact subset with ΛT 1 as its characteristic function. Then, for any Eisenstein series
E SLn /P (φ, λ, g) induced from an L2 -automorphic form φ on the Levi subgroup of a
parabolic subgroup P of SLn ,
Z Z
SLn /P
(φ, λ, g)dµ(g) = ΛT E SLn /P (φ, λ, g) dµ(g). (15.5)
E
Σ(T ) SLn (F)\SLn (A)/SK
Proof. We first need to justify that Σ(T ) makes sense. By Theorem 14.2, for all
g ∈ SLn (F)\SLn (A)/SK, 1(pg ≤ p) = ΛT (p) 1 (g). The right hand side, be-
ing taking values either zero or one, is easily a characteristic function of a com-
pact subset of DSLn ,F := SLn (F)\SLn (A)/SK. Hence Σ(T ) makes sense. Conse-
quently, by Lemmas 15.1 and 15.2, we have
Z Z
ΛT E SLn /P (φ, λ, g) dµ(g) = ΛT ΛT E SLn /P (φ, λ, g) dµ(g)
D DSLn ,F
Z SLn ,F Z
Λ 1 (g) · Λ E SLn /P
E SLn /P (φ, λ, g)dµ(g)
= T T
(φ, λ, g) dµ(g) =
DSLn ,F Σ(T )
as desired.
(2) Let (G, P) be a pair consisting of a reductive group of rank n and its maximal
parabolic subgroup. Denote by α p its associated simple root. The T-version
of the (G, P)-zeta function of F is defined by
G,T
ζF,T
b G/P
(ns − n) := Norm(s) · Res s1 =0,..., s[
p =0,...sn = 0
ωF (λ) (15.8)
where
X ehwλ−ρ,T i Y ζF (hλ, α∨ i)
ωG,T
F (λ) = .
b
(15.9)
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
w∈W ζF (hλ, α∨ i + 1)
In the study of rank n zeta functions, the author formulates the following:
Theorem 15.2. The conjecture on parabolic reduction, stability, and the volumes
holds.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 385
Proof. By (9.46) and (9.47), and more importantly, Theorem 15.1, we have
n
X X vF,k1 . . .b
b vF,k p 1
ζF,n (s) =
b (−1) p+r−1 ·
a=1 k1 ,...,k p >0
(k1 + k2 ) . . . (k p−1 + k p ) ns − n + a + k p
k1 +···+k p =n−a
ζF (ns − n + a)
×b
X 1 vF,l1 . . .b vF,lr
.
b
× ·
l1 ,...,lr >0
−ns + n − a + 1 + l1 (l1 + l2 ) . . . (lr−1 + lr ) (15.12)
l1 +···+lr =a−1
ζF,n (s)
Therefore, by Theorem 3.2(3), for the residue of non-abelian zeta function b
at s = 1, we have
vol MF,n [1] = Res s=1 bζF,n (s)
n
X X vF,k1 . . .b
b vF,k p 1
= (−1) p+r−1 · ζF (a)b
·b vF,a−1
a=2 k1 ,...,k p >0
(k1 + k2 ) . . . (k p−1 + k p ) a + k p
k1 +...+k p =n−a
X vF,k1 . . .b
b vF,k p 1
+ (−1) p+r−1 · ζF (1)
·b
k1 ,...,k p >0
(k1 + k2 ) . . . (k p−1 + k p ) 1 + k p
k1 +...+k p =n−1
n
X X vF,k1 . . .b
b vF,a
vF,k pb 1 (15.13)
= (−1) p+r−1 ·
a=2 k1 ,...,k p ,a>0
(k1 + k2 ) . . . (k p−1 + k p ) k p + a
k1 +...+k p +a=n
X vF,k1 . . .b
b vF,1
vF,k pb 1
+ (−1) p+r−1 ·
k1 ,...,k p ,1>0
(k1 + k2 ) . . . (k p−1 + k p ) k p + 1
k1 +...+k p +1=n
X X vF,n1 . . .b
vF,nk
=
b
(−1)k−1
k≥1 n1 ,...,nk >0,
(n1 + n2 ) . . . (nk−1 + nk )
n1 +...+nk =n
as desired.
Proof. The second half is a direct of the first since the stability condition is open,
it suffices to prove the existence. As the proof below works for F as well, without
loss generality, we work only over Q. Then, it suffices to show that there is a stable
lattice in MQ,n [T ], for one and hence for all T . We use an induction on the rank
n. When n = 1, every rank one lattice is stable. Assume that all Λ ∈ MQ,n [1] are
not stable. Choose stable lattices Λi (i = 1, 2) of rank ni and degree di . Consider a
non-trivial extension of Λ2 by Λ1
0 → Λ1 → Λ → Λ2 → 0. (15.14)
Assume that the degree of Λ equals zero. Since Λ is non-stable, d1 > 0. Let Λ0 be
a sub lattice of Λ such that 0 ⊂ Λ0 ⊂ Λ is the canonical filtration of Λ. We obtain
an induced filtration
0 → Λ1 ∩ Λ0 → Λ0 → Λ0 /(Λ1 ∩ Λ0 ) → 0. (15.15)
Since Λ1 ∩ Λ0 ,→ Λ1 and Λ0 /(Λ1 ∩ Λ0 ) ,→ Λ2 , we have, using the stability of
the Λi ’s, the slope of Λ0 is bounded from below, resp. from above, by −d2 /n1 ,
resp. by d2 /n2 . Therefore, the composition of morphisms Λ0 ,→ Λ Λ2 is an
isomorphism. This implies that (15.14) splits, a contradiction.
(1) All but finite many zeros of the non-abelian zeta function bζQ,n (s) lie on the
1
central line <(s) = .
2
SL /P
(2) All but finite many zeros of the Weng zeta function bζQ n n−1,1 (s) lie on the
n
central line <(s) = − .
2
Proof. By Theorems 11.2 and 15.1, it suffices to verify the non-vanishing of the
volume of MQ,k [1] for 1 ≤ k ≤ n − 1. This comes from Theorem 15.2 and
Proposition 15.1.
Chapter 16
387
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 388
(1) X∗ (AP ), resp. X∗ (A0P ), is the group of one-parameter subgroup in AP , resp. A0P .
(2) X∗ (AP ) and X∗ (A0P ) are free abelian groups of the same finite rank.
(3) Induced from (16.3), there is a natural injective morphism
X∗ (AP ) ,−→ X∗ (A0P ). (16.5)
Accordingly, set
X∗ (AP )R = X∗ (A0P )R =: aP . (16.6)
Q
and aR , respectively. For simplicity, if P is fixed, we sometimes use [·] instead
of [·]Q
P.
(3) Let ∆0 := ∆P0 ⊂ Φ+0 be the set of simple roots of (Φ0 , Φ+0 ).
On the other hand, (even it may not be a root system,) many constructions for
a root system may be in parallel introduced for ΦP . To be more precise,
(5) For each α ∈ ∆P , there exists one and only one β ∈ ∆0 such that α = β a .
P
∆∨P := α∨ : α ∈ ∆P .
(16.15)
∆∨P forms an R-basis of aGP , and denote its dual basis in aG∗
P by
∆P := $Gα : α ∈ ∆P .
b (16.16)
(6) There exists a canonical inner product (·, ·)P on aGP . Let WP be the Weyl group
generated by reflections σα on aGP associated to α ∈ ΦP defined by
(α, β)P
σα (β) := β − 2 α ∀β ∈ aGP . (16.17)
(α, α)P
X
∆P ⊂
One checks that b Z≥0 α, that WP is generated by σα with α ∈ ∆P , and that
α∈∆P
(·, ·)P is WP -invariant.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 390
Definition 16.1.
Lemma 16.1.
Proof. (1) comes from the fact that every $Gα ∈ b ∆P belongs to α∈∆P Z≥0 α. (2)
P
is a standard fact. For details, please refer to §10.1 of [44]. (3) comes from the
definition of fundamental weights $α . (4) is rather standard as well. For details,
please see Proposition 18 of Chapter VI of Bourbaki [16].
x ∈ aG+
0 t {0} if and only if x ≥ w(x) ∀w ∈ W. (16.24)
n n
X o
0 t {0} := (x1 , x2 , . . . , xn ) ∈ R : x1 ≥ x2 ≥ . . . ≥ xn ,
aG+ xj = 0
n
j=1
Proposition 16.1. ( [7]) For x, y ∈ aG0 , the conditions below are equivalent.
(1) x ≤ y.
dx ⊆ Wy.
(2) W c Here W x denotes the W-orbit of x, and b
S denotes the convex cone
of a subset S ⊂ aG0 .
(3) φ(x) ≤ φ(y) for all W-invariant convex function φ : aG0 → R.
c ⊆ Gy.
(4) Gx c Here Gx denote the G-orbit of x for the associated adjoint action.
Proof. (2) ⇒ (1) This is clear. Indeed, W dx ⊆ Wy c implies that x ∈ Wy.c Hence,
x = w∈W aw w(y) with aw ≥ 0 and w∈W aw = 1. Thus, by (16.24), we have
P P
x = w∈W aw w(y) ≤ ( w∈W aw )y = y.
P P
Proof. Let αi be the simple roots normalized to have length one. Then the αi ’s
are the unit normal vectors to the faces of aG0 and form a basis of + aG0 . Since y is
assumed interior to aG0 , we have (x, αi ) > 0 for all i. But y − z is assumed interior
to + aG0 , we have y − z = ri=1 ai αi with ai > 0. Thus, if denote by σi = σαi the
P
tai
reflection associated to αi , and, for 0 ≤ t ≤ 1, introduce the numbers bi :=
2(y, αi )
Xr
and set b := 1 − bi , we have, since σi (y) = y − 2(y, αi )αi ,
i=1
r
X r
X r
X
by + bi σi (y) = y + bi (σi (y) − y) = y − 2 bi (yi , αi )αi
i=1 i=1 i=1
r
(16.25)
X
= y−t ai αi = tz + (1 − t)y = x.
i=1
Hence x ∈ Wy c provided bi ≥ 0 and b ≥ 0, and this will hold if, for all i, we
(y, αi ) (y, αi )
have 0 ≤ t ≤ 2 . Set now ci (y, z) := 2 . We want to know the behavior
rai rai
of ci (y, z)’s when y runs over points in the line segment [y, z]. Obviously, when
replacing y by the variable point x = tz + (1 − t)y, ai changes to (1 − t)ai and
(x, αi ) = t(z, αi ) + (1 − t)(y, αi ) ≥ (1 − t)(y, αi ) (16.26)
since z ∈ aG0 and t ≥ 0. Therefore ci (x, z) ≥ ci (y, z) and it suffices for us to take
c(y, z) := mini ci (y, z).
Lemma 16.2.
(1) Let x ∈ aG0 . Denote by Gx the G-orbit of x under the adjoint action. Then,
with respect to the induced orthogonal projection π : g → aG0 ,
π(Gx) = W
dx. (16.27)
(2) Let φ be a W-invariant convex function on t and ϕ be the corresponding G-
invariant function on g. Then ϕ is also convex.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 393
Proof. We only prove (2) while refer (1) to [59]. For a function on Rn , denote
by Γ( f ) be the region above its graph, i.e. the collection of all points (x, y) with
x ∈ Rn and y ∈ R such that y > f (x). Then the convexity of the function f is
equivalent to the convexity of the grape Γ( f ). Recall that Γ is convex if every
boundary point a has a supporting hyperplane Ga , i.e. Γ is contained in one of the
two half-spaces complementary to Ha . Apply this to the functions φ, ϕ and their
associated regions Γ(φ) ⊂ aG0 ⊕ R and Γ(ϕ) ⊂ g ⊕ R. Because ϕ is G-invariant, so is
Γ(ϕ). Hence it suffices to prove the existence of a supporting hyperplane to Γ(ϕ)
at boundary points (λ, y) of Γ(φ). Since γ(φ) is assumed to be convex, there exists
a supporting hyperplane H ⊂ aG0 ⊕ R. Let then H 0 = π−1 (H) ⊂ g ⊕ R where π is
the orthogonal projection. Take then (x, y) ∈ Γ(ϕ) satisfying y > ϕ(x). By (1), or
more precisely, by π(Gx) ⊆ W dx, we have ϕ(x) ≥ φ(π(x)) since φ is convex. Hence
y > φ(pi(x)) and (π, x) ∈ Γ(φ). Therefore (π, x) is in one side of H and hence (x, y)
is on the corresponding side of H 0 . This means that H 0 is the required supporting
hyperplane for Γ(ϕ).
Obviously, (1) is equivalent to (5), and (2) implies (3) and (4), and (3) implies
(5) and (4) implies (5). By Lemma 16.2(1), resp. (2), we have (4) implies (2),
resp. (5) implies (3). It remains to show that (3) implies (2). For this, we take
φ(x) := w∈W exp(w(tx), ei ) where t > 0 and ei are a basis of aG0 , choosing from
P
the edges of this cone. Note that for x, y ∈ aG0 , by Lemma 16.24(3), (4), we have
(x, ei ) ≥ (w(x), ei ) and (y, ei ) ≥ (w(y), ei ). Thus, by the continuity, when x and
y are interior to aG0 , we have (x, ei ) < (w(x), ei ) and (y, ei ) < (w(y), ei ) if w , 1.
Consequently, for sufficiently large t, the first term with w = 1 in the sum defining
φ is dominant. Hence by φ(x) ≤ φ(y) we have (y, ei ) ≥ (x, ei ) for all i. Therefore,
y − x ∈ + aG0 and x ≤ y.
0 , since there are r basic integral wights ei that lie in the edges of a0 and
span aG+ G+
generate it.
(2) If φ(x) = φ(y) for all W-invariant convex functions φ, then W dx = Wy.c
Hence, the extreme points of these two convex polyhedra must coincide. But the
extreme points of W dx are certainly among the finite set W x. Hence W x and Wy
intersect and so coincide.
In the case when G = Un , the partial order above can be described explicitly in
concrete terms, say, following Atiyah-Bott [7].
Let µ = (µ1 , µ2 , . . . , µn ) and λ := (λ1 , λ2 , . . . , λn ) ∈ Rn . By definition, µ ≤ λ,
if, after arranging each sequence in decreasing order, we have
P
j=1 µ j ≤ j=1 λ j
i Pi
1≤i≤n−1
(0)
n µj = n λj
P P
j=1 j=1
and that M is called doubly stochastic, if both M and its transpose t M are
stochastic.
Lemma 16.4. ([11]) The subset of doubly stochastic matrices of size n is the
convex hull of the permutation matrices.
Consequently, we have
(3) S[n (µ) ⊆ Sn (λ). Here, as usual, Sn is the symmetric group acting on the co-
[
ordinates of Rn via σ : (x1 , x2 , . . . , xn ) 7→ (xσ(1) , xσ(2) , . . . , xσ(n) ).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 395
On the other hand, the geometric notion of convexity can be transformed into
the dual statement on convex functions using the following well-known:
Indeed, by the equivalence of (a) and (b), we see easily that (3) implies (4).
On the other hand, if we take φ(x1 , x2 , . . . , xn ) := ni=1 f (xi ), then (4) implies (1).
P
Consequently, (0) and hence (1), (2), (3), (4) are equivalent to
(5) The λ j ’s are eigen values of a hermitian matrix with diagonal entries the µ j ’s.
Theorem 16.1. For any two elements λ, µ ∈ Rn , the condition µ ≤ λ and hence
(1), (2), (3), (4) (5) are all equivalent to either of the following two conditions.
c ⊆ [λ],
(6) [µ] c where [λ] denotes the conjugacy class of hermitian matrices whose
eigenvalues are the λ j ’s.
(7) For all A ∈ [µ] and B ∈ [λ], and for all convex Sn -invariant function ϕ on the
space of hermitian matrices, ϕ(A) ≤ ϕ(B).
But by (c) above, this means that µ ∈ Rn is in the convex hull of Sn (λ) for the
eigenvalues λ j of A.
(6) ⇒ (7) is trivial. Furthermore, by setting ϕ(B) = φ(λ), then (4) also implies
(7). So what is left is that (7) implies (4).! This is classical. Indeed, for every con-
α0
!
A B
vex invariant function ϕ, we have ϕ ≥ϕ for a blocked skew-hermitian
CD 0δ
matrix with α and δ the central components of A and ! D respectively. This follows
α0
from Horn’s theorem and the observation that is the convex hull of the Sn -
0δ
!
A B
orbit of the diagonal part of . As the same argument works for all types of
CD
blocked matrices, we are done.
The constructions above admit some natural generalizations to the pairs (P, Q)
consisting of standard parabolic subgroups P, Q of G satisfying P ⊆ Q. To start
with, recall that there are canonical morphisms
A0 ←- AP ,→ MP ,→ P = MP NP ,→ Q = MQ NQ ←- MQ , (16.29)
and that the elements of ΦP , resp. Φ+P , resp. ∆P are the restrictions of the roots in
Φ, resp. the positive roots in Φ+ , resp. the simple roots in ∆, to AP . Accordingly,
we let ΦQ P = ΦP∩MQ , resp. ΦP
Q+
= Φ+P∩MQ , resp. ∆Q P = ∆P∩MQ , be the set of
α ∈ ΦP , resp. α ∈ Φ+P , resp. α ∈ ∆P , which appear in the Lie algebra mQ of the
Levi factor MQ of Q. Indeed, the elements of ∆P are non-trivial characters of AP
which appear in nP , and there are natural morphisms nP ,→ p ,→ q ←- mQ , hence
nP ∩ mQ makes sense in q, and so does ∆Q P . Here, as the notation stands, p is the
Lie algebra of P, etc....
In terms of Lie spaces, there is a natural decomposition aP = aQ P ⊕ aQ , and the
set ∆Q
P forms an R-basis of a Q∗
P . In addition, being elements of ∆ P , each α ∈ ∆P
Q
Q∨
P with respect to ∆P .
be the dual R-basis of aQ∗
Induced from the W-invariant inner product (·, ·) on a, we obtain a natural inner
product (·, ·)Q Q Q
P on aP by taking the restriction. Set WP be the group generated by
the reflections σa on the space aP defined by
Q
(β, α)
σα : β 7−→ β − 2 α ∀α ∈ ΦQ
P , ∀β ∈ aP
Q
(16.32)
(α, α)
since (β, α) = (β, α)QP and (α, α) = (α, α)P . Obviously, (·, ·)P is WP -invariant and
Q Q Q
(1) If α ∈ ∆Q
P , by definition,
§7.1, many structures for the root systems work for the relative setting as well.
Definition 16.3 (Definition 7.2). (1) The relative (acute) Weyl chamber aQ+
P is
defined by
P := H ∈ aP : hα, Hi > 0 ∀α ∈ ∆P .
aQ+
Q Q
(16.36)
(2) The relative (obtuse) Weyl cone + aQ
P is defined by
+ Q
aP := H ∈ aP : h$αQ , Hi > 0 ∀α ∈ ∆Q
Q
P .
(16.37)
+ Q
(3) The characteristic functions of the subsets aQ+ Q
P and aP in aP are denoted by
τP and b
Q
τP , respectively.
Q
(1) The positive Weyl chamber is contained in the positive Weyl cone. That is,
+ Q
P ⊆ aP .
aQ+ (16.38)
(2) The positive Weyl cone is the dual cone of positive Weyl chamber. That is, + aQ
P
is the dual cone of aQ+ Q
P with respect to the inner product (·, ·) on aP
+ Q
n o
aP = x ∈ aQ P : (x, aP ) ⊂ R>0 .
Q+
(16.39)
P , for every α ∈ ∆P ,
Proof. (1) First we note that, for a fixed H ∈ aQ Q
(a) if either (α, H) > 0 or ($αQ , H) > 0, then, ($αQ , H) > 0 for all α ∈ ∆QP.
(b) if either (α, H) ≤ 0 or ($αQ , H) ≤ 0, then, $αQ , H ≤ 0 for all α ∈ ∆QP .
Since for H ∈ aQ+ P , hα, Hi > 0 for every α ∈ ∆P . Hence, from (a) above,
Q
Claim 16.2. For two elements α, β of ∆Q P , there exists a strictly positive constant
cαβ such that
α, $βQ QP = cαβ · δαβ .
(16.40)
Here, as usual δαβ denotes the Weierstrass delta symbol.
n o
Proof. By definition, $αQ : α ∈ ∆Q P is the dual basis of aQ P for the basis
n o
α : α ∈ ∆P of aP . Furthermore, β = [γ ]P ∈ aP ⊆ a0 ⊕ aP ⊕ aG∗
∨ Q Q∗ ∨ ∨ Q Q∗ P∗ Q∗
Q = a0 for
G∗
Moreover, we have the following standard facts for the Coxeter groups WPQ .
Proposition 16.2.
(1) Let σ ∈ WPQ , then σ aQ+ aQ+
S T
σ∈WPQ P P is a union of finite number of hy-
perplanes of aQ
P , all passing through the origin. We call them the walls.
[
P =
(2) aQ Q σ aP
Q+
.
σ∈WP
In particular, aQ+
P is a ‘fundamental domain’ for the action of WPQ on the metrized
R linear space aP , (·, ·)P .
Q Q
Proposition 16.3.
x ∈ aQ+
P if and only if x − w(x) ∈ + aQ
P ∀ w ∈ WPQ . (16.43)
In particular, if x ∈ aQ+ Q
P , then x ≥ w(x) for all w ∈ WP .
Q
(2) Let x, y be two elements in aP . Then
Q
x≤y if and only if x∈W
d
P y. (16.44)
Here WPQ y
denotes the WPQ -orbit
of the element y, and for any subset S ∈ aQ
P,
Sb denotes the convex cone of S in aQ
P .
Proof. (1) When (P, Q) = (P0 , G), this is recalled in Lemma 16.1, a standard fact
given in Bourbaki [16], Chapter VI, Proposition 18. However, the proof there
works for general Coxeter groups.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 400
(2) This may be proved following that of Proposition 16.1. Indeed, for a fixed
y ∈ aQ+
P , we may consider the set
+
P;≤y := x ∈ aP : x ≤ y .
aQ
Q+
(16.45)
P;≤y = aP ∩ Wy.
aQ Q+ c (16.46)
Hence, by (1),
X X
y= aw y ≥ aw w(y) = y. (16.48)
w∈WPQ w∈WPQ
If so, a finite number N of repetitions, where N −1 < c(y, z) then proves that the
whole interval [z, y] belongs to Wy.
c
aα > 0 for all α ∈ ∆QP . Now let wα ∈ WP be the reflection in the wall (x, α) = 0 so
Q
X X X (y, α)Q
by + bα wα (y) = y + bα (wα (y) − y) = y − 2 ba P
α α α (α, α)Q
P
X
= y−t aα α = tz + (1 − t)y = x.
α
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 401
In this section, we first recall some basic facts on maximal compact subgroups
and the Carton involution associated to algebraic groups, mainly following [15]
and [31]. Then we introduce a notion of canonical metrics associated to the Carton
involution. When the groups involved are semi-simple, all these are classical.
Let G be a real Lie group with finitely many connected components. Then any
compact subgroup of G is contained in a maximal one. If K is a maximal one, G
is diffeomorphic to the direct product of K with a euclidean space, G/G0 = K/K 0 ,
and any two maximal compact subgroups are conjugate in G.
If G1 is a closed normal subgroup of G, with finitely many connected compo-
nents, then the maximal compact subgroups of G1 are the intersections of G1 with
the maximal compact subgroups of G. Similarly, if G1 is a closed subgroup of
G with finitely many connected components such that all maximal compact sub-
groups of G are conjugate by elements of G1 , the maximal compact subgroups of
G1 are the intersections of G1 with the maximal compact subgroups of G. Conse-
quently, in both cases, by taking a maximal compact subgroup K of G containing
a maximal one of G1 , we see that G1 ∩ K is maximal compact in G1 for at least
one and hence for all maximal compact subgroups K of G by conjugacy.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 402
Note that in the case when G is semi-simple, θK is the usual Cartan involution.
Motivate by this, we call θK the Cartan involution of G(R) with respect to K.
Moreover, the existence of the Cartan involution θK implies an existence of a non-
degenerate symmetric bilinear form B on gC × gC satisfying the follows.
Lemma 16.9. Let H be a positive definite real symmetric bilinear form on g. Let
θ := H −1 B. Assume that H is K-compatible. Then
When G is semi-simple, this definition and lemma are due to Grayson ([34]).
Definition 16.6. The morphisms θ in the previous lemma are called Cartan invo-
lutions associated to K.
1 Here as usual, we view the bilinear H as a linear map from g to g. Hence H −1 B is indeed a linear
endomorphism of g.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 405
Proof. Recall that, for φ ∈ AutLie (g), we have t φBφ = B since B is characteristic.
g;K , we have (− φHφ)B = φ θφ. This implies that φHφ
−1
Hence, for any H ∈ Mtot t t
belongs to Mtot
g;K as well. Consequently, the assignment
(H, g) 7−→ t (Ad g)H(Ad g) g;K , ∀g ∈ G
∀H ∈ Mtot (16.52)
defines a natural action of G on Mtot
Here, as usual, for g ∈ G, Ad g denotes its
g;K .
adjoint. This proves (1).
In terms of Cartan involutions θ, the action above is equivalent to
(θ, g) 7−→ t (Ad g)θ(Ad g) ∀g ∈ G. (16.53)
Recall that, for a general G, if we set oθ := { X ∈ c : θX = −X }. Then Vθ := exp oθ
is a split component of G. Moreover, by Proposition 2.1.10 of [31], the assignment
θ 7→ oθ gives a bijection from the set of Cartan involutions associated to K to the
set of split components of G. In particular, when G = G1 , this map gives a bijec-
tion from the set of Cartan involutions to the set of maximal compact subgroups
of G. Hence, in our case, since G is connected and reductive, all maximal compact
subgroups of G are conjugate to each other. This implies that G acts transitively
on Mtotg;K . Thus, to complete the proof, it suffices to show that the stabilizer group
of θK in G is exactly K itself. For this we choose ΘK : G(R) → G(R) to be a Car-
tan involution satisfying dΘK = θK . By definition, K = { g ∈ G(R) : Θ(g) = g }.
On the other hand, for g ∈ G, θK = (Adg)−1 θK (Adg) if and only if g−1 ΘK g is in
the center of G(R)0 . But this center is trivial by our assumption, hence g belongs
to the stabilizer group of θK if and only if g ∈ K.
Definition 16.7. Let F be a number field with OF its ring of integers. Let G be
a connected (split) reductive group over a number F. By a metrized G-torsor, or
equivalently, a principal G-lattice (over OF ), we mean a pair gOF , (Hσ ) consist-
ing of the canonical infinitesimal integral structure gOF of G/F and an element
(Hσ ) of Mtot
g∞ ;K∞ .
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 407
Q
Set now B := σ∈S ∞ Bσ , where Bσ denotes the canonical form on gFσ with
respect to Kσ . By our construction,
Lemma 16.10. Let G be a reductive group over R with g its associated Lie al-
gebra. Then a sub-algebra a ⊂ g is a parabolic sub-algebra if and only if it is
adg -nilpotent and the orthogonal component a⊥ with respect to a canonical form
on g is contained in a.
When G is semi-stable, all the lemmas in this subsection are due to [33, 34].
The proofs there work in our general setting with some modifications when
necessary.
Proof. With an extension of the coefficients to C, we can and hence will work
over C. Let b be a Borel sub-algebra of g. We must show that b ⊆ a.
Being a canonical form, B is associated to a maximal compact subgroup K
of G. Consequently, using the (±1) eigenspaces within b of the Cartan involution
θK , we obtain an
Xorthogonal decomposition b = k0 ⊕ p0 with respect to B. De-
note by p0 = gα where gα denotes the associated root space, and fix a base
α∈k∗, α>0
of b obtaining from gα ’s compatible with the order determined by positive roots
involved.
Let x ∈ a⊥ , the perpendicular subspace of a. Since it is solvable, a⊥ is con-
tained in b. Consequently, if we write x = k + p with k ∈ k0 and p ∈ p0 , then
adb (p) is an upper-triangle matrix and adb (k) is a diagonal matrix. Thus the nilpo-
tent property of adg a implies that adb (x) is nilpotent. Hence adb (k) = 0. This
implies that α(k) = adgα (k) = 0 for all α ∈ Φ+ . Thus k = 0 since simple roots form
a basis of the dual space of b. Therefore a⊥ ⊆ p0 .
To go further, note that B(k0 , gα ) = 0, hence k0 ⊆ p⊥0 . Moreover, since
B(gα , gb ) = 0 when α + β , 0. Hence p0 ⊂ p⊥0 . Therefore, p⊥0 = b since
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 408
dim b + dim[b, b] = g and p0 = [b, b]. This then completes the proof since
b = p⊥0 ⊆ (a⊥ )⊥ = a.
Conversely, if a is a parabolic sub-algebra, then we may choose a Borel sub-
algebra b such that b ⊆ a. With this, it is rather obvious to track back to show the
only if direction left.
φ : W ⊗ W → Λ/W . Since
0 0 0
To go further, consider next the Lie bracket
inf W 0 ⊗ W 0 = inf W 0 + inf W 0 = 0 and sup Λ/W 0 < 0, we have φ = 0. This
implies that W 0 is a Lie sub-algebra lattice of gOF . For the similar reason, since
[W − j , W −1 ] ⊆ W − j−1 for all j ≥ 0, W −1 is a nilpotent ideal. Consequently, W −1 is
the nilpotent radical of W 0 and W 0 /W −1 is a reductive quotient of W 0 . Therefore,
by Lemma 16.10, W 0 is a parabolic sub-algebra lattice over OF , since we have:
(1) W 0 = W,
(2) W j = (W − j+1 )⊥ within g∞ for all i,
(3) [W i , W j ] ⊆ W i+ j for all −r ≤ i, j ≤ r − 1.
as desired.
Finally, from the parabolic sub-algebra lattice W 0 over OF , we obtain the cor-
responding parabolic subgroup of G and this is its own normalizer. Therefore, the
sub-algebra lattice W 0 over OF determines a reduction of the structural group G
of Λ to this parabolic subgroup Q. We denote this new principal P-lattice over OF
by ΛQ and call it the canonical parabolic reduction of Λ. All these then prove the
following:
Proposition 16.5. Let G be a connected (split) reductive group over F. Set gOF
be the canonical infinitesimal structure of G over the ring of integers OF of F,
Q Q
G∞ := σ∈S G(Fσ ) and g∞ := Lie(G∞ ). Assume that Kσ := σ∈S Kσ is a
maximal compact subgroup of G∞ . Then, for each (Hσ ) ∈ Mtot g∞ ;K∞ , the induced
principle G-lattice Λ = (gOF , (Hσ )) admits a canonical parabolic reduction ΛQ
for a unique F-parabolic subgroup Q of G.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 410
Let F be an algebraic number field with OF its ring of integers. By a group scheme
G on X = Spec OF , we mean G is a (smooth affine) algebraic group scheme over
X with connected fibers. Similarly, by a G-torsor E on X with a structural map
π : E → X, we means that π is faithfully flat and locally finite type (surjective)
projection, and there is a group action G × E → E satisfying the properties that it
is compatible with π and that the induced morphism G ×X E → E ×X E, charac-
terized by (g, x) 7→ (g x, x), is an isomorphism. Dente by G, resp. E, the generic
fiber GF and EF of G and E, respectively. We always assume that G is split over
F. Naturally, we obtain an induced F-action G × E → E. In the sequel, we will
be interested in the reductive group schemes G defined over F, that is, the alge-
braic group schemes such that all of its geometric fibers are (connected) reductive
algebraic groups.
Let now B be a dense open subscheme of X. Then for pairs (Gi , Ei ), i = 1, 2,
consisting of a reductive group scheme Gi and a Gi -torsor Ei on X, we say that
(G1 , E1 ) is B-isomorphic to (G2 , E2 ), denoted by (G1 , E1 ) 'B (G2 , E2 ), if G1 ' G2
as group schemes over B and there exists a B-isomorphism E1 ' E2 which is
compatible with the actions of G1 and G2 . We have the following recent result of
Javanpeykar-Loughran.
Lemma 16.12. ([48]) For a fixed n ∈ Z>0 and a dense open subscheme B of X,
the set of B-isomorphism classes of pairs (G, E) is finite, provided that G is an
algebraic reductive group scheme of rank n.
Definition 16.8.
Lemma 16.13. There exists a bijection between the moduli space of arithmetic G-
torsors over OF and the quotient space G(F)\G(A)/K, where K = v∈S fin G(Ov ) ×
Q
K∞ denotes the maximal compact subgroup of G(A) induced by the OF structure
on G and K∞ .
With the infinite places discussed, with Lemma 16.13, we are ready to give
an adelic version of compatible metrics, the compatible measures on G(A) with
respect to K = v∈S fin G(Ov ) × K∞ . Our basic principle is that all the compati-
Q
ble measures with respect to K on G(A) should be compatible with the Iwasawa
decomposition
G(A) = N0 (A) · A M0 (A) · M0 (A)1 · K = N(A) · A M(A) · M(A)1 · K. (16.63)
Here P0 = N0 M0 is the Levi decomposition of a minimal parabolic subgroup
of G and P = N M is the Levi decompositions as above for standard parabolic
subgroups P of G.
To be more precisely, for the compact group K∞ of G∞ , its associated compat-
ible measures on G∞ are discussed above. This induces naturally the compatible
measures dm on M0 (A) with respect to K ∩ M0 (A) with an induction argument,
since the works obviously for all split tori. Furthermore, since, for the unipotent
radical N of P, N0 (F) is discrete and N0 (F)\N0 (A) is compact, we can define the
compatible measures dn on N0 (A) to be the Haar measures on N0 (A) such that the
volume of N0 (F)\N0 (A) is one. With this, for all compactly supported continuous
functions f on G(A),
Z
f (uamk) dk a−2ρ0 da dm du
f 7→ (16.64)
N0 (A)·A M0 (A) ·M0 (A)1 ·K
and P = N M its Levi decomposition. Set p := Lie(P), n = Lie (N) and form
p∞ := σ∈S ∞ p ⊗ Fσ and n∞ := σ∈S ∞ n ⊗ Fσ . Denote by Mtot
Q Q
n∞ ;K∞ the moduli
space of the n∞ and K∞ -admissible metric on p∞ .
Lemma 16.14. There exists a bijection between the moduli space of admissible
arithmetic parabolic group schemes over Spec OF associated to K∞ and the quo-
tient space P(F)\P(A)/K, where K = Kfin × K∞ with Kfin the maximal compact
subgroup of P(A) defined by the associated OF -structure.
Let G be a connected reductive group scheme over Spec OF with G its generic
fiber. We assume that G is split over F. Fix a maximal compact subgroup scheme
K of G and a minimal parabolic subgroup scheme P0 of G over Spec OF . Denote
by N0 the unipotent radical of P0 . Let P be a standard parabolic subgroup scheme
over Spec OF with N the unipotent radical. As above, denotes the associated
generic fibers by P0 and P respectively.
Definition 16.10.
Proof. As the proof for the second one is a direct generalization of the first, we
here only prove (1) to illustrate. The existence is a direct consequence of the
constructions in §16.2.3, in particular, that of §16.2.3.2. As for the uniqueness,
if there exists two, say P and P0 . Let T be the maximal torus of G satisfying
T ⊂ P ∩ P0 . If T is not split, by passing to an extension of F, we may assume
that T is split over the extension field. On the other hand, it is clear that this does
not affect the positivity of the slopes involved. Consequently, there exists an even
larger parabolic subgroup scheme satisfying all the properties. This contradicts to
the maximality we start with.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 415
Remark 16.1. (1) As noticed above, the conditions EP ' EQ ×Q P and EP '
ν
EνQ × Q P imply automatically that ν(EQ )]P = νP and ν(EνQ )]P = νP , respectively.
N0 ,K∞
Accordingly, we denote by MG,F (νG ), or simply MG,F (νG ), to be the moduli
space of semi-stable admissible arithmetic G-torsors E over OF of slope νG with
respect to N0 and K∞ . Similarly, for each parabolic subgroup scheme P of G,
0 ,K∞
we denote by MN F;G
(P, νG ), or simply MF;G (P, νG ), the (sub) moduli space of
G-torsors EG over OF of slope νG which admit (P, νP ) as their canonical types
and EP as their parabolic reductions. We denote such P by P and write E := EP
and ν := νP . More generally, for a parabolic subgroup scheme P of G, we denote
0 ,K∞
by MN P,F
(νP ), or simply MP,F (νP ) the moduli space of semi-stable admissible
arithmetic P-torsors EP over OF of slope νP . And, for each parabolic subgroup
0 ,K∞
scheme Q contained in P, denote by MN P,F
(Q, νQ ), or simply MP,F (QνQ ), the
sub moduli space corresponding to the admissible arithmetic P-torsors EP over
OF of slope νP with respect to N0 and K∞ which admit (Q, νQ ) as their canonical
types, and EQ as their parabolic reductions. We denote such Q by QP , and write
EP := EQ and νP := νQ .
P P
Recall that, in the theory of automorphic forms and trace formula, when working
with Arthur’s analytic truncation ΛP,T , we always assume that T is sufficiently
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 416
regular. This is mainly because, in the proof of Theorem 7.1 on the basic prop-
erties for Arthur’s analytic truncation ΛP,T , used are the classical reductions and
the Langland partition for G(A) via Langlands combinatorial lemma, for them the
sufficiently regular condition is crucial. As we will see below, for these funda-
mental properties of Arthur’s analytic truncation, by our result on the equivalence
between the analytic truncation and geometric truncation using stability (to be es-
tablished below), this sufficiently regular condition becomes artificial and hence
will be removed. What matters is the positive condition
T ∈ a0P+ ∪ {0}.
Let F be a number field. Denote by OF its ring of integers and denote by A its
adelic ring. Let G be a reductive group over F and K be a maximal compact
subgroup of G(A). Fixed once and for all an integral model G of G, a group
of reductive group scheme over X = Spec OF with G its generic fiber. (Recall
that by Lemma 16.12, there are only finitely many of them with a fixed B-type.)
For a element g ∈ G(F)\G(A)/K, let Eg be the associated G-torsor. Denote by
(P, νg ) the canonical type of Eg and by Eg,P the canonical reduction of Eg . More
generally, if P is a parabolic scheme with P its generic fiber, we let Eg,P be the P-
torsor obtained from Eg via the parabolic reduction associated to P ,→ G. Denote
by νg,P the slope of Eg,P . From the definitions, we have
ν p := νP,g ∈ aGP ⊆ a0P ⊕ aGP = aG0 and νP,p ∈ aGP ⊆ a0P ⊕ aGP = aG0 . (16.72)
When G = GLn , this is essentially Theorem 14.2, for which the canonical
convex polygons and a natural partial order on a0 of type An−1 is used. However,
for general reductive groups, this language of polygons cannot be applied. Instead,
we use naturally defined partial orders recalled above.
Proof. Since a similar proof below for (1) works for (2) as well, we here only
work put the details for (1).
First assume that νg ≤ T . Then the left hand side is equal to 1. Hence it
suffices to show that the right hand side is equal to 1 as well. For this purpose,
we analysis the terms on the right hand side. For each fixed parabolic subgroup
P of G, its associated contribution (to the right hand side) is given by, up to sign,
1 νg,P >P νP,g . Easily, from the definition,
νP,g ≥ νP,g .
as desired.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 418
Consequently, we have
νP,g − νP,g ∈ + aGP or the same νP,g ≥P νP,g ∀P. (16.75)
This is a contradiction unless P = G. That is to say,
1 νg,P >P νP,g = 0
unless P = G. (16.76)
Therefore, the right hand side is simply 1, since 1(νg,G >G νP,g ) ≡ 1. This verifies
the claim and hence also the theorem under the assumption that νg ≤ T .
Next we assume that νg T . Then the left hand side becomes zero from the
definition. Hence it suffices to show that the right hand side equal zero as well,
namely,
(−1)dim aP 1(νP,g >P T ) ≡ 0.
X G
(16.77)
P
For this purpose, recall that, for each parabolic subgroup P and the arithmetic
P-torsor associated to g, there exists a unique P-canonical subgroup QP,g . Ac-
cordingly, following an idea of Lafforgue used in the proof of Theorem 14.2, we
regroup the summation in (16.77) according to Q = QP,g . In this language, since,
for a fixed parabolic subgroup P of G, the P-canonical parabolic subgroup of g is
unique, to prove the theorem, it suffices to establish the following:
Proof. We first point out that, on the right hand side of (16.79), the sum is taken
over parabolic subgroups P such that, with respect to fixed parabolic subgroups Q
and ν Q, the canonical parabolic reduction within P of the admissible arithmetic
G-torsor Eg is given by Q and the ν-parabolic reduction within P of the admissible
G-torsor Eg is given by ν Q.
For a standard parabolic subgroup P of G, denote by P = MP NP its Levi
decomposition with MP , resp. NP , the standard Levi factor, resp. the nilpotent
radical, of P. It is well known, see e.g. [36] or, more recently, [21], that there exist
the simple factors of reductive groups MP,1 , . . . , MP,|P| , ordered according to Φ+0 ,
or the same ∆0 , such that
For this, associated to g ∈ G(A), or better, to the principal P-bundle EP,g with
slope νP,g , we introduce the following sets:
g n o
∆0 := β j ∈ ∆ : hβ j , νQ,g i > 0 ,
g n o
∆ν := β j ∈ ∆ : hβ j , νQ,g i > hβ j , νi ≥ 0 , (16.84)
ν g ν
n o
∆T := αi ∈ ∆ : hαi , νν Q,g − T i > 0 .
Consequently,
g g g
∆ν ⊆ ν ∆ ⊆ ∆ and (∆ r ∆0 ) ∪ ( ν ∆ r ∆ν ) ⊆ ∆P ⊆ ν∆gT . (16.85)
Proof. Both injectivity and surjectivity come directly from the fact that there
exists an order reversed bijection between the subsets of ∆0 and the standard
parabolic subgroups of G. Indeed, the infectivity is rather obvious. As for the sub-
jectivity, the existence of a parabolic subgroup comes from J ⊆ ν ∆. Moreover, by
reversing (a,. . . ,e) above, we see that QP,g = Q, ν QP,g = ν Q, and νP,g >P T.
Lemma 16.17. Let Y be a finite set and denote by Σ(Y) its power set, i.e. the set
consists of all the subsets of Y. Then
X 0 Y , ∅,
(−1) =
|S |
(16.86)
1 Y = ∅.
S ∈Σ(Y)
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 421
Consequently, for the fixed Q and ν Q as above, the condition that the right
hand of (16.79) is not zero implies that
g g g
∆ν ⊆ ν ∆ ⊆ ∆, ∆ r ∆0 ∪ ν ∆ r ∆ν = ν∆gT .
(16.87)
g ∈ Q(F)\G(A)/K : Qg = Q, ν Qg = ν Q, νg ≤ T .
n o
(16.89)
Claim 16.6. Assume that T dominates ν. Then the set ν∆gT is empty provided that
g g g
∆ν ⊆ ν ∆ ⊆ ∆, ∆ r ∆0 ∪ ν ∆ r ∆ν = ν∆gT .
(16.90)
Proof. Assume otherwise that ν∆gT is not empty. Let k be the smallest number such
that βk of ∆ attains maximal among the hβ j , νg −T i’s for β j ∈ ∆. Since T dominates
g
ν, we certainly have βk ∈ ∆ν . On the other hand, βk ∈ ν ∆ and hence βk ∈ ν∆gT since,
by our assumption, ν∆gT , ∅. For the same reason, β j ∈ ∆0 . This contradicts with
g g
the condition that ∆ r ∆0 ∪ ν ∆ r ∆ν = ν∆gT . Therefore, ν∆gT = ∅.
g g g g
Since ∅ = ν∆gT = (∆ r ∆0 ) ∪ ( ν ∆ r ∆ν ), we have ∆ = ∆0 and ν ∆ = ∆ν . This to
say, Qg = Q and ν Qg = ν Q. Moreover, under the assumption that T dominates ν,
the relations νν Q,g ≤ T and ν Qg = ν Q implies that νg = νQ,g ≤ T . This verifies the
first claim, i.e, Claim 16.5.
Hence the map (16.79) is the constant map g 7→ 0. This proves Lemma 16.16 and
hence also Theorem 16.2.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 422
Now we are ready to expose the following fundamental result on an intrinsic rela-
tion between Arthur’s analytic truncation and the geometric truncation of stability.
In this subsection, we will establish some fundamental properties for analytic and
hence geometric truncations. These properties are first proved by Arthur for ana-
lytic truncations ΛT under the condition that the parameter T is sufficiently regular
(see e.g. Theorem 7.1). Based on Theorem 16.3, we here show that all these prop-
erties can be generalized under a mild positivity condition
T ∈ a0P+ ∪ {0}. (16.95)
In the sequel, we call such a T positive.
16.3.3.1 Compactness
(2) Let T ∈ a0P+ ∪ {0}. For any T 0 ∈ a0P+ , ΛP,T 1 is the characteristic function of
the compact subset
n o
g ∈ P(F)N(A)\G(A)1 /K : −T 0 ≤ νP,g ≤ T . (16.97)
Proof. Since a similar proof below for (2) works for (1), we only prove (2). For
this, by Theorem 16.3, ΛP,T 1 is the characteristic function of the moduli space
n o
g ∈ P(F)N(A)\G(A)/K : −T 0 ≤ νP,g ≤ T . (16.98)
show that (ΛP,T − ΛP,T )1 is the characteristic function of a compact subset. But
0
This then leads to the problem whether there exist semi-stable compatible
arithmetic G-torsors of slope ν. For our limited purpose, we give the following:
Proposition 16.8. Assume that G is semi-simple. Then the moduli space MG,F (0)
is non-empty and hence with a non-zero volume.
Proof. This is a direct consequence of the works of Grayson in [33, 34], particu-
larly, §4 and §5 of [34]. Indeed, in the space G(F∞ )/K∞ , there exists a non-empty
manifold with boundary Xss such that its quotient under G(OF ) gives the moduli
space MG,F (0). Therefore, it suffices to prove that Xss is not empty. This is a
direct consequence of Properties (A1∼ A9) of [34]. Indeed, Xss is not only non-
empty, but a topological space whose boundary is homotopy equivalent to the Tits
building by a homotopy equivalence which respects the action of G(OF ).
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 424
16.3.3.2 Idempotence
We next show that the operator ΛP,T is idempotent for all positive T .
Proof. The proof is similar to that of Lemma 15.1. Indeed, by Theorem 16.3, we
may identify the geometric truncation 1(νP,g ≤P T ) and the analytic truncation
ΛP,T 1, and hence it suffices to prove the second equality. But for geometric trun-
cations, this is a direct consequence of the proof of Theorem 16.2. Indeed, for
g ∈ G(A), when νP,g ≤P T , from the definitions, the right hand side is one, and the
left hand side becomes ΛP,T 1 (g) = 1(νP,g ≤P T ) = 1 by Theorem 16.3 again.
On the other hand, if νP,g P T , by the proof of Theorem 16.2, the right hand
side becomes 0, while the right hand side gives ΛP,T 0, and hence becomes 0 as
well.
Next, we show that ΛP,T is self-adjoint with respect to the natural inner product.
(1) Let φ1 and φ2 be two continuous functions on G(F)\G(A)1 /K. Assume that
φ1 is slowly increasing, and that φ2 is rapidly decreasing. Then
ΛT φ1 , φ2 = φ1 , ΛT φ2 . (16.100)
Proof. When G = SLn , this is simply Lemma 15.2. Even the group G now is
more general, the proof for Lemma 15.2 works here as well, after changing C and
SLn in (15.4) to aG+
0 and G, respectively.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 425
As a direct consequence to the above basic properties for analytic and geometric
truncations, similar to Lemma 15.3, we have the following:
Theorem 16.4 (Analytic Periods are Geometric Periods). Let T ∈ aG+ 0 t {0}
and denote by Σ(T ) ⊂ G(F)\G(A)1 /K be the compact subset with ΛT 1 as its
characteristic function. Then, for an Eisenstein series EG/P (φ, λ, g) induced from
an L2 -automorphic form on the Levi subgroup of a parabolic subgroup P of G,
Z Z
E G/P (φ, λ, g)dµ(g) = ΛT E G/P (φ, λ, g) dµ(g).
(16.102)
Σ(T ) G(F)\G(A)1 /K
Proof. This is a direct consequence of Theorem 16.3 and Propositions 16.7, 16.9
and 16.10, since if we denote by DG,F a fundamental domain of G(F)\G(A)1 /K,
we have
Z Z
ΛT E G/P (φ, λ, g) dµ(g) = ΛT ΛT E G/P (φ, λ, g) dµ(g)
D DG,F
Z G,F Z
ΛT 1 (g) · ΛT E G/P (φ, λ, g) dµ(g) =
= E G/P (φ, λ, g)dµ(g)
DG,F Σ(T )
as desired.
In the next chapter, we will use the results above to prove a weak Riemann
hypothesis for the Weng zeta functions of reductive groups over number fields.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 427
Chapter 17
Let F be a number field and let G be a (split) reductive group of rank r over F.
Fix a minimal parabolic subgroup P0 of G with a Levi decomposition P0 = M0 N0 ,
where M0 , resp. N0 , denotes a Levi subgroup, resp. the nilpotent radical, of P0 .
For the constant function 1 on M0 (A), denote by E G/P0 (1, g, λ) the induced Eisen-
stein series on G(F)\G(A). In addition, for every non-negative T ∈ aG0 t{0}, denote
by Σ(T ) the compact subset in G(F)\G(A)1 /K defined by the characteristic func-
tion ΛT 1. By Theorems 16.3 and 16.4, Σ(T ) admits also a geometric interpretation
≤T
as the moduli spaces MG,F of K-admissible arithmetic G-torsors Eg satisfying the
condition that their associated canonical types νEg are ≤ T . In particular, Σ(0)
is the moduli space MG,F of semi-stable K-admissible arithmetic G-torsors. Our
≤T
main purpose is to calculate the volume of the moduli spaces MG,F .
Let P be a standard maximal parabolic subgroup and denote by α p ∈ ∆
be the corresponding simple root and set ∆ p = ∆ r {α p }. And, for the Weyl
group W associated to (G, P), introduce the set Wspa , resp. W 0 , of standard, resp.
(p-)special, elements by
n o
Wspa := w ∈ W : ∆ = α ∈ ∆ : wα ∈ ∆ ∨ wα < 0 ,
n o (17.1)
W 0 := w ∈ W : ∆ p = α ∈ ∆ p : wα ∈ ∆ ∨ wα < 0 .
427
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 428
(1) The integration of E G/P0 (1; g, λ) over the truncated domain Σ(T ) is given by
ehwρ−ρ,T i ζ(hλ, α∨ i)
Z
E G/P0 (1; g, λ)dµ(g) =
X Y
.
b
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
≤T
MG,F w∈W ζ(hλ, α∨ i + 1)
In particular,
ζ(hλ, α∨ i)
Z
1
E G/P0 (1; g, λ)dµ(g) =
X Y
.
b
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
MG,F w∈W ζ(hλ, α∨ i + 1)
≤T
(2) The volume of MG,F is given by
X Q bni,w −ni (i)
i≥1 ζF
vol ∆ ∨ r2 r
· (2π) (−1) r−|Aw |
ehwρ−ρ,T i .
− hwρ, α∨ i)
Q
w∈Wspa α∈∆rwAw (1
Z
Therefore, lim E G/P0 (1; g, λ)dµ(g) is equal to the product of the volume of
λ→ρ Σ(T )
Σ(T ) and b
ζF (1)r k≥2 b
ζF (k)nk . This leads to lim ωG,F
T
Q
(λ) where
λ→ρ
X e hwρ−ρ,T i Y ζF (hλ, α∨ i)
ωG,F .
T
b
(λ) := (17.4)
α∈∆ hwλ − ρ, α i α>0,wα<0 b
∨
Q
w∈W ζF (hλ, α∨ i + 1)
So the method of [57] can be used to give the formula above.
Motivated by Theorem 7.6, we can write the above result in terms of parabolic
reduction. To explain this, we make some preparations first.
Recall that, if E is an arithmetic G-torsor of slope ν with (P, ν) its canonical
type of E, by Proposition 16.6, there exists a semi-stable arithmetic p-torsor EP
such that
E = EP ×P G and P , [ν]G = ν.
[νP ]GP ∈ aG+ (17.5)
Denote by MG,F (P, ν) the moduli space of arithmetic G-torsors with canonical
type (P, ν). In the sequel till the end of this subsection, for simplicity, we write P
for P.
Let P = MP NP be the Levi decomposition of P with MP , resp. NP , a Levi
subgroup, resp. the unipotent radical, of P. Denote by M MP ,F (ν) the moduli
spaces of semi-stable arithmetic MP -torsors of slope ν. There is a natural fibration
structure
π : MG,F (P, ν) −→ M MP ,F (ν)
. (17.6)
E 7−→ N\EP
Since NP (A)/NP (F), or better NP (R)/NP (Z) is a compact torus and NP admits
a filtration of one-dimensional affine spaces, by the conditions (a,b, c) for the
associated Haar measures in §6.1.5, the volume of the moduli space MG,F (P, ν) is
equal to the volume of the moduli space N\EP .
Moreover, from the theory of reductive groups explained in [36], if MP is a
connected semi-simple group over a field F, then the set {MP,i }i∈I of the minimal
non-trivial normal smooth connected F-subgroups of MP is finite, where, each
MP,i is F-simple,
Y the MP,i ’s are pairwise commutative, and the multiplication ho-
momorphism MP,i → MP is a central isogeny. More generally, for a connected
i
reductive group MP over a field F with Z its maximal central F- torus, if we de-
note by MP0 := D(MP ) its semi-simple derived group and let {MP,i } be the F-simple
factors of MP0 , then for the associated MP,i ’s, the multiplication map
Y
MP,i → MP (17.7)
i∈I
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 430
≤
(1) The volume of MG,F T can be calculated in terms of parabolic reduction.
Namely,
X 1
≤
vol MG,F T (ν) = (−1)|∆P | · Q
P:P⊆Q α∈∆\wP JP (1 − hwP ρ, α∨ i)
standard parabolic
(17.9)
Y tot hw ρ−ρ,T i
× vol M MP.i (0) e J .
i
Q
(2) The volume of i∈I M Mi ,F (0), and hence the volumes of M M,F (0) and
MG,F (P, 0) are not zero.
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 431
As a preparation for our proof of the weak Riemann hypothesis for zeta functions
of reductive groups, we expose the following ket connection.
Proposition 17.2. Let G be a reductive group over a number field F and let P be
a maximal subgroup of G. Then the discriminant of b ζFG/P (s) is (up to a non-zero
factor,) the volume of the moduli space MG,F (P, 0) of admissible arithmetic G-
torsors of slope zero and canonical type (P, 0).
Proof. By the proof of Proposition 17.1, the volume of the moduli space
MG,F (P, 0) is the same as the product of the volumes of the moduli spaces
M Mi ,F (0) of semi-stable admissible arithmetic Mi -torsors, where the Mi ’s are the
simple factors of the Levi subgroup of P. Hence it suffices to show that the dis-
criminant of b ζQG/P (s) coincides with this product. For this, we make a decomposi-
tion of the discriminant of b ζ G/P (s). We start with the following:
Q
Lemma 17.1. Let (Φ, Φ± , ∆, W, ρ) be the complete data of a reduced and irre-
ducible root system. Denote by λα : α ∈ ∆ be the set of fundamental dominate
weights. For a simple root α ∈ ∆, denote by Φα the subset of Φ which is orthog-
onal to λα . Then Φα , together with its intersections with Φ± and ∆, forms a root
system.
α1 (k), . . . , αrk (k) and b ∆r (k) = λ1 (k), . . . , λrk (k) , then α1 , α2 , . . . , αr−1 =
α1 (1), . . . , αr1 (1), α1 (2), . . . , αr2 (2), . . . , α1 (`), . . . , αr` (`) , and
`
Y 1 X
∆r = t`k=1 ∆r (k), Wr = Wr (k) and ρr (k) = α.
k=1
2 α∈Φ (k)+
r
Consequently, if we rewrite the period ωGF (λ) as ωΦ ∆,F (λ) based on the fact that
ωΦ∆,F (λ) is essentially determined by the root system (Φ, ∆), then it makes sense for
Φr (k)
us to introduce the period ωΦ ω∆r (k);F λ(k) for all i = 1, . . . , `,
r
(λ
∆r r ) and the periods
where λ = i=1 (si +1)λi and λ(k) = j=1 (s j (k)+1)λi (k) with λ1 (k), . . . , λrk (k) =
Pr−1 Prk
∆r (k) for all k = 1, , . . . , `. Assume that αr is the simple root corresponding to the
b
maximal parabolic subgroup P. Then, by Proposition,12.12 and the definition
of the discriminant ∆G/P F of the zeta function b ζFG/P (s), up to the multiple factor
α∈Φ+r r∆r ζF (hρ, α i + 1),
Q b ∨
∆G/P
F = lim ωΦ r
∆r ,F (λr ).
λr →ρr
In addition, from the orthogonality of the Φr (k)’s, we have easily
` `
Φr (k)
ωΦ Φr
lim ωΦ
Y Y
= ω λ(k) ω = ∆r (k),F λ(k) .
r (k)
r
∆r ;F (λ) ∆r (k);F and lim (λ
∆r ,F r )
λr →ρr λ(k)→ρr (k)
k=1 k=1
Moreover, since for each 1 ≤ k ≤ `, (Φr (k), ∆r (k)) is naturally associated to the
semi-simple algebraic group Mk , by Theorem 17.1, lim ωΦ ∆r (k),Q λ(k) coin-
r (k)
λ(k)→ρr (k)
cides with the volume of the moduli space M Mk ,Q (0) of the semi-stable admissible
arithmetic Mk -torsors of slope zero. This, together with Proposition 17.1, then im-
ζQG/P (s) is the volume of the moduli space MG,F (P, 0)
plies that the discriminant of b
of admissible arithmetic G-torsors of slope zero and canonical type (P, 0).
Chapter 18
Following our work [131], we expose two levels of the distribution structures for
ζQ,n (s) and b
the zeros of b ζQG/P(s).
433
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 434
while the classical theta leads to the Dirac distributions, the secondary big Theta
recovers the Sato-Tate type distributions for high rank zeta zeros (associated to
non-CM elliptic curves defined over Q). All these in turn motivate the current
works on a parallel structure for the zeros of bζQ,n (s).
√
ζQ,n (s) as ρ = 12 + −1 γ. Arrange the γ’s
To explain this, write the zeros of b
in an increasing order
0 ≤ γn,1 ≤ γn,2 ≤ · · · ≤ γn,3 ≤ . . . , (18.1)
and, as usual, denote by
Nn (T ) := # k : 0 < γn,k < T
(18.2)
ζQ,n (s) whose imaginary parts lie between 0 and T .
the number of zeros of b
ζQ,n (s), as T → ∞,
Theorem 18.1. Let n ≥ 21 . For the zeros of b
n n 2πe
(1) Nn (T ) = T log T − T log + O(1).
2π 2π n
2π k log log k
(2) γn,k = 1+O .
n log k log k
2π 1 log log k
(3) γn,k+1 − γn,k = +O .
n log k k log k
The structures presented in this theorem are pretty much similar to the one for
the Riemann zeros ( [116]). To go further, motivated by [76], based on (2) and (3)
above, we introduce
Being of Dirac type, in particular, the distributions for the zeros of the non-abelian
zeta functions are very different from that of Riemann zeros, which conjecturally
coincide with that of GUE in the theory of random matrix. However, it turns out
1 Here and in the sequel, when n = 2, stronger results hold. For details, please see Proposition 18.1.
December 21, 2017 13:46 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 435
that there is yet another refined level of structure for the zeros of non-abelian zeta
functions. To explain this, motivated by our studies for non-abelian zeta functions
of function fields ([132]), we introduce the following:
Definition 18.2. The big Delta pair correlation functions for the zeros of high
rank non-abelian zeta function
ζQ,n (s) are defined by
n
Δn,k := δn,k − 1 · log γn,k . (18.5)
2π
The distributions of Δn,k and δn for the high rank zeta zeros and the Riemann
zeros, respectively, are expected to be closely related ([83]). We, motivated by the
conjectural connection between Riemann zeros and GUE, make the following:
Conjecture 18.1. Denote by μ(Δn ) and μ(GUE) the probability measures associ-
ated to Δn,k and the Gaussian unitary ensemble, respectively. Then
limn→∞ Discrep μ(Δn,k ), μ(GUE) = 0. (18.6)
Our method to prove the above results works for the zeta functions b ζQG/P (s)
as well. To explain this, let (G, P) be a pair consisting of a Chevalley group and
its maximal parabolic subgroup over Q. By Theorem 17.2, if the rank of G is
ζQG/P (s) satisfies a weak Riemann hypothesis. Consider the zeros
at least one, b
cP cP
ρ = − + γ i of b ζQG/P (s) on the central line <(s) = − . Arranging the γ’s in an
2 2
increasing order
0 ≤ γG/P
1 ≤ γG/P
2 ≤ · · · ≤ γG/P
k ≤ ..., (18.7)
and denote by
n o
N G/P (T ) := # k : 0 < γG/P
k <T (18.8)
the number of the zeros with imaginary parts lying between 0 and T . Set
∞
1X
dP := k · NP (k, [(kcP − 1)/2]),
2 k=1
∞
(18.9)
1X
eP := NP (k, [(kcP − 1)/2]) k log k
2 k=1
Theorem 18.3. As T → ∞,
dP eP − dP log(2πe)
(1) N G/P (T ) = T log T + T + O(log T ).
π !! π
π n 1
(2) γG/P
n = 1+O .
dP log n log n
π 1
!
1
(3) γn+1 − γn =
G/P G/P
+O .
dP log n log2 n
Accordingly, introduce the pair correlation function, the small delta, for these
zeta zeros by
dP G/P ! dP G/P
!
δG/P := γ − γG/P
· log γ . (18.10)
k π k+1 k π k
Theorem 18.4. As k → ∞,
!
1
δG/P
k =1+O . (18.11)
log k
ζ G/P
In other words, the pair correlation distributions for the zeros of b Q (s) are
simply of Dirac type. Motivated by ∆n,k , we introduce the following:
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 437
ζ G/P
Definition 18.3. The big Delta pair correlation functions for the zeros of b Q (s)
are defined by
!
G/P dP G/P
∆G/P := δ − 1 · log γ . (18.12)
k k π k
To start with, we prove a stronger result for rank two zeta zeros, obtained inde-
pendently by Suzuki and the author.
Proof. It suffices to prove the first three since (4) comes directly from (2) and (3).
1
We start with (1). Since bζ(2s − 1) = b ζ(2s) when s = + i t, by Theorem 5.9,
2
ζ(1 + 2it)
!
1 1 b
ζQ,2
b + it = · cos θ2 (t), (18.13)
2 2 − 12 + i t
where θ2 (t) denotes the argument of b ζ(1 + 2it)/(− 12 + it). Recall that in the proof
of the simplicity of the zeros, θ2 (t) is an increasing function in t. Hence,
θ2 γ2,n+1 − θ2 γ2,n = π.
(18.14)
Furthermore, it is not difficult to understand the asymptotic of
θ2 (t) = arg Γ 1/2 + it + arg ζ 1/2 + it − arg π 2 +it − arg − 1/2 + it . (18.15)
1
Consequently,
!
log t
θ2 (t) = t log t − t (1 + log π) + O . (18.18)
log log t
This, together with (18.14), implies that N2 (T ) = θ2 (T )/π, asymptotically, and
hence also (1). To prove (2), we use the relation
This implies that γ2,n log γ2,n ∼ πn, and hence γn,2 ∼ π n/ log n. Consequently,
!
log(γ2,n ± 1)
π N2 (γ2,n ±1) = (γ2,n ±1) log(γ2,n ±1)−log π−1 +O . (18.20)
log log(γ2,n ± 1)
From this, with a simple manipulation, we obtain (2). To prove (3), we use (18.18),
or equivalently, (18.16) and (18.17) to conclude that, for sufficiently small h,
!! !
1 1
θ2 (t + h) − θ2 (t) = h log t 1 + O +O . (18.21)
log log t t
Applying this to t = γ2,n and h = γ2,n+1 − γ2,n , then (3) comes from (18.14).
ζQ,2 (s)
Definition 18.4. The big Delta pair correlation functions for the zeros of b
are defined by
γ2,k
∆2,k = δ2,k − 1 · log log .
(18.22)
π
ζQG/P (s).
Step 1. Fine symmetric structures of b
Without loss of generality, let P be a standard parabolic subgroup of G. De-
note by P = MP NP the Levi decomposition of P, nP the Lie algebra of NP , AP the
maximal central subgroup of MP , and aP := X ∗ (AP )R . Let ∆P be the set of roots for
(P, AP ), i.e. the finite subset of non-zero elements in X(AP )Q parametrizing the de-
composition nP = ⊕α∈ΦP nα of the eigenspace under the adjoint action Ad : AP →
GL(nP ) of AP , where nα := {Xα ∈ nP : Ad(a)Xα = aα Xα ∀a ∈ AP }. Note that
1X
ΦP ⊂ X(AP )Q ⊂ X(AP )Q ⊗ R ' a∗P . Accordingly, set ρP := (dim nα ) α.
2 α∈ΦP
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 439
By the theory of algebraic groups (see e.g. [14, 103]), there is an order revers-
ing bijection
where
α∈∆ hλ − ρ, α i ζ (hλ, α∨ i)
Q ∨ Y
.
b
T w (s) := lim Q P
α∈∆ hwλ − ρ, α i
∨
λ→λP
α>0, wα<0 ζ (hλ, α∨ i + 1)
b
Even after taking the residues, there are still some zeta factors left in the de-
nominators of T w . To get ride of them, following (12.7), we first introduce an
‘overdone’ normalization for ωG/P
F (s) and set
where
−1 +
Y Y
2|∆P ∩w Φ |
qP,w (s) := hλ s + ρ, α∨ i − 1
w∈W 0 α∈(w−1 ∆)∩∆P
Y
× hλ s + ρ, α∨ i + δα,w hλ s + ρ, α∨ i + δα,w − 1 ,
α∈Φ+ \∆P
where
(Φ+ )|
−1
Y
c p,w := ξ(2)|∆ p ∩w ξF hρ, α∨ i + δα,w ,
α∈Φ+ r∆ p
(Φ+ )|
Y −1
Y
a p,w (s) = 2|∆ p ∩v hλ p , α∨ is + hρ, α∨ i − 1
v∈W p0 r{w} α∈(v−1 (∆)r∆ p
Y !
× hλ p , α is + hρ, α i + δα,v hλ p , α is + hρ, α i + δα,v − 1 ,
∨ ∨ ∨ ∨
α∈Φ+ r∆ p
Y
ξF hλ p , α∨ is + hρ, α∨ i + δα,w .
z p,w (s) :=
α∈Φ+ rΦ+p
Therefore, as in Definition 12.2, for w ∈ W 0 , set `±p,w := (Φ+ r Φ+p ) ∩ w−1 (Φ± ) ,
and
W p< := w ∈ W p0 : `+p,w < |Φ+ r Φ+p |/2 ,
where
∞ Y
Y ∞
DP (s) := ξ(ks + h)NP (k,h−1)−MP (k,h) ,
k=1 h=2
BP (s)
Then, by Proposition 12.2, ξQG/P (s) is entire, and, for AP (s) := ,
RP (s)DP (s)
− cP /2 + i t .
Step 2. Asymptotic behaviors for the arguments of AP
Let θP (t) be the argument of AP − cP /2 + it . By (18.31),
c c
ξQG/P − + i t = AP − + i t · ei θP (t) ± e−i θP (t) ,
P P
(18.32)
2 2
The first term is simply O(1), since a p (s), RP (s) are polynomials. To treat the
second term, we use the formula (12.58) that
∞
z p (s) Y Y
= ξ(ks + h)NP (k,h−1)−NP (k,h) .
D p (s) k=1 h>(kc +1)/2
P
December 21, 2017 12:32 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 442
Note that 1 ≥ < (ks + h) = − c2P k + h > 12 for s = − c2P + it. We can assume
cP
that arg ζ − k + h + ikt = O(log |t|) obtained under the Riemann Hypothesis.
2
Thus, from the Stirling formula,
∞
z p (s) X X
arg cP = NP (k, h − 1) − NP (k, h)
DP (s) s=− 2 +it k=1 h>(kc +1)/2
P
× arg(ks + h)(ks + h − 1) s=− cP +it
2
! !
cP h ikt cP h ikt c
P
+ arg π 4 k− 2 − 2
+ arg Γ − k + + + arg ζ − k + h + ikt
4 2 2 2
X∞ X
= NP (k, h − 1) − NP (k, h)
k=1 h>(kcP +1)/2
! ! ! !
kt kt kt 1
× O(1) − log π + log −1 +O + O(log t)
2 2 2 t
∞ " #! !
X kcP − 1 kt
= log t − log(2πe) + log k + O(log t) .
NP k, ·
k=1
2 2
All these then prove the following proposition and hence also Theorem
18.3(1).
ζQG/P (s).
Step 3. Distributions of zeros for b
Then !! !
1 n 1 1 1 1
γG/P
n = 1+O and γG/P
n+1 − γG/P
n = +O .
C1 log n log n C1 log n log2 n
1 n
As above, then n = C1 γn log γn + O(γn ), or better, since γn ∼ ,
C1 log n
1
C1 γn log γn = n · 1 + O . (18.38)
log n
This verifies the first assertion.
For the second, we shift our attention to θ = θP . Then, for T 0, ∆T > 0,
θ(T +∆T ) − θ(T ) = C 0 N G/P (T + ∆T ) − C 0 N G/P (T )
T + ∆T T + ∆T
!
= C C1 T log
0
+ C C1 ∆T log(T + ∆T ) + C C2 ∆T + O log
0 0
T T
!
1
= C 0C1 ∆T log T + 1 + O ,
T
!T
T + ∆T ∆T
since T log = log 1 + = O(1). In particular, by taking T = γn and
T T
∆T = γn+1 − γn , we get
C = θ(γn+1 ) − θ(γn ) = C 0C1 (γn+1 − γn ) log γn + 1 + O 1/γn .
This implies
!
C 1 1
γn+1 −γn ∼ 0 and C = C C1 (γn+1 −γn ) log γn +O
0
. (18.39)
C C1 log γn log γn
Therefore,
!
C 1 1
γn+1 − γn = + O .
C 0C1 log n log2 n
This then completes the proof of the lemma and hence Theorems 18.3 and 18.4,
since we can take
dP eP − dP log(2πe)
C = C 0 = π, C1 = and C2 = (18.40)
π π
ζ G/P (s). Indeed, by the proof of the simplicity of zeros of
for the zeta function b Q
ζQG/P (s), θP (t) is monotone and hence θP (γG/P
b
k+1 ) − θP (γk ) = π. This proves Theo-
G/P
We end this discussion with the comment that Theorems 18.1 and 18.2 can
also be proved directly using Theorem 9.1.
∞ ∞ !k
X i k Λk (n) k X Λk (n) ζ0
Here λz (n) := − z with Λk (n) defined by := (s) .
k=0
2 k! n=1
ns ζ
The M-function M(z, σ) is real valued, decays rapidly as |z| → ∞. Since
λz (m)λz (n) = λz (mn) for (m, n) = 1, M(z, σ), being a Fourier transform, admits a
natural convolution Euler product.
Theorem 18.5. Let ϕ(x) be a function in C 1 (R)! such that ϕ0 (x) 1 for |x| < 1,
d ζ0 n
ϕ0 (x) x−2 for |x| ≥ 1, and ϕ < + int is bounded on R. Then for n ≥ 2,
dt ζ 2
Z ∞
1 X 1 n
ϕ ∆n,k = ϕ(u) du.
lim m u, (18.42)
T →∞ Nn (T ) 2π −∞ 2
0<γk ≤T
Z ∞
Here m(u, σ) := M(u + iv, σ) dv.
−∞
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 445
APPENDICES
This part presents a series of joint works with K. Sugahara [106–108] on adelic
arithmetic cohomology groups over arithmetic varieties and applications. To be
more precise, in Appendix A, we start with a general construction for these co-
homology groups, based on the pioneer works of Parshin and Beilinson on adelic
cohomology groups for quasi-coherent sheaves over Noetherian schemes, a gener-
alization of the Osipov-Parshin tensor products in dimension one ind-pro topology
to all dimensions and a general uniformity principal. Then in Appendix B, we shift
our attention to arithmetic surfaces. Here we construct the arithmetic cohomol-
ogy groups for Weil divisors concretely and establish some of the basic algebraic
properties for them. In Appendix C, we give an intensive study of the ind-pro
topology in arithmetic dimension two. In Appendix D, we prove the topological
duality for the associated cohomology groups and show that the middle arithmetic
cohomology groups are always finite. Finally, in Appendix E, as an application of
these arithmetic cohomology constructions, we construct some arithmetic central
extensions and hence prove several reciprocity laws for arithmetic surfaces, using
some ideas of Osipov and Parshin for algebraic surfaces and the Arakelov theory.
445
b2530 International Strategic Relations and China’s National Security: World at the Crossroads
Appendix A
We first recall some basic constructions of adelic cohomology theory for algebraic
varieties following Parshin and Beilinson ([10, 91, 92], see also [43]).
447
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 448
(i) X1,α1 , being an integral subvariety of the normal scheme X00 , defines a discrete
valuation of the field of rational functions on X0 , whose residue field coincides
0
with the field of rational functions on the normal scheme X1,α 1
.
(ii) More generally, for a fixed i, 1 ≤ i ≤ n, Xi,αi , being an integral subvariety of
0
the normal scheme Xi−1,α i−1
, defines a discrete valuation of the field of rational
0
functions on Xi−1,αi−1 , whose residue field coincides with the field of rational
0
functions on the normal scheme Xi,α i
.
Accordingly, for each collection (α1 , α2 , . . . , αn ) of indices, the chain of field of ra-
tional functions K0 , K1,α1 , . . . , Kn,αn defines an n-dimensional local field K(α1 ,...,αn )
and hence an Artin ring
Theorem A.1. ([91, 92, 136]) Let δ = (p0 , p1 , . . . , pn ) be a reduced flag on X, and
K(α1 ,...,αn ) be the n-dimensional local field associated to the collection of indices
(α1 , α2 , . . . , αn ) above. Then
either Fv0 ((tn−1 )) · · · ((t1 )), or Fv0 {{tn }} · · · {tm+2 }}((tm )) . . . ((t1 ))
Let X be a Noetherian scheme, and let P(X) be the set of (integral) points of X (in
the sense of Grothendieck). For p, q ∈ P(X), if q ∈ {p}, we write p ≥ q. Let S (X)
be the simplicial set induced by (P(X), ≥), i.e. the set of m-simplices of S (X) is
defined by
S (X)m := (p0 , . . . , pm ) | pi ∈ P(X), pi ≥ pi+1 ,
(A.3)
the natural boundary maps δni are defined by deleting the i-th point, and the de-
generacy maps σni are defined by duplicating the i-th point:
δm
i : S (X)m → S (X)m−1 , (p0 , . . . , pi , . . . , pm ) 7→ (p0 , . . . , p̌i , . . . , pm ),
(A.4)
σm
i : S (X)m → S (X)m+1 , (p0 , . . . , pi , . . . , pm ) 7→ (p0 , . . . , pi , pi , . . . , pm ).
m = (p0 , . . . , pm ) ∈ S (X)m pi , p j ∀i , j .
S (X)red
(A.5)
For p ∈ P(X) and M an O p -module, set [M] p := (i p )∗ M, where i p : Spec(O p ) ,→
X denotes the natural induced morphism. Moreover, for K ⊆ S (X)m and a point
p ∈ P(X), introduce p K ⊆ S (X)m−1 by
p K := (p1 , . . . , pm ) ∈ S (X)m−1 | (p, p1 , . . . , pm ) ∈ K ∩ S (X)m .
red red
(A.6)
Then, we have the following
Proposition A.1. ([10, 91, 92], see also Proposition 2.1.1 of [43]) There exists
a unique system of functors {A(K, ∗)}K⊆S (X) from the category of quasi-coherent
sheaves on X to the category of abelian groups, such that
m ,F .
AmX (F ) := A S (X)red
(A.8)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 450
Indeed, for coherent sheaves F , we can uniquely define {A(K, F )}K⊆S (X) by (2);
and, for quasi-coherent sheaves F , in addition, we use (1), since any such F can
be obtained as direct limits of coherent sheaves on X. Clearly, if we introduce
Ki0 ,...,im = (p0 , . . . , pm ) ∈ S (X)m | codim {pr } = ir ∀ 0 ≤ r ≤ m , and define
AX;i0 ,...,im (F ) := AX Ki0 ,...,im , F , then
M
AmX (F ) = AX;i0 ,...,im (F ). (A.9)
0≤i0 <···<im ≤dim X
p∈P(X) lim←−l ϕ p ;
Q l
Theorem A.2. ([10,91,92], see also Theorem 4.2.3 of [43]) Let X be a Noetherian
scheme. Then, for any quasi-coherent sheaf F over X,
H i A∗X (F ), d∗ = H i X, F
∀i. (A.12)
A.1.3 Example
Let X be an integral regular projective curve defined over a field k. Denote its
generic point by η and its field of rational functions by k(X). For a divisor D on X,
let OX (D) be the associated invertible sheaf. Then, from definition, the associated
adelic spaces can be calculated as follows:
AX;0 (OX (D)) = A {η}, OX (D)
(A.13)
= lim OX (D)η mlη OX (D)η = lim k(X)/{0} = k(X),
←−l ←−l
and
AX;01 (OX (D)) = AX {η, p | p ∈ X : closed point}, OX (D)
Remark A.1. To calculate AX;01 (OX (D)), when dealing with the constant sheaf
[k(X)]η , we cannot use Proposition A.1(2) directly, since [k(X)]η is not coherent.
Instead, above, we first expressed it as an inductive limit of coherent sheaves
OX (E) associated to divisors E, then get the result from the inductive limit of
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 452
adelic spaces for OX (E)’s. Indeed, if we had used Proposition A.1(2) directly,
then we would have obtained simply A {p} | p ∈ X : closed point}, [k(X)]η = {0},
a wrong claim.
Note that AX (OX (D)) is simply AX (D) of [98]. We have proved the following:
Proposition A.2. For a divisor D over an integral regular projective curve de-
fined over a field k, we have
Let K be a number field with OK the ring of integers. Denote by S fin , resp. S ∞ ,
the collection of finite, resp. infinite, places of K. Write S = S fin ∪ S ∞ . Let
π : X → Spec OK be an integral arithmetic variety of pure dimension n + 1. That
is, an integral Noetherian scheme X, a proper morphism π with generic fiber XF
a projective variety of dimension n over K. For each v ∈ S , we write Fv the v-
completion of K, and for each σ ∈ S ∞ , we write Xv := X ×OK SpecFv and write
ϕσ : Xσ → XF for the map induced from the natural embedding F ,→ Fσ . In
particular, an arithmetic variety X consists of two parts, the finite one, which we
also denote by X, and an infinite one, which we denote by X∞ . These two parts
are closely interconnected.
The part of our theory on arithmetic adelic complexes for finite places now be-
comes very simple. Indeed, our arithmetic variety X is assumed to be Noetherian,
so we can apply the theory recalled in §A.1 directly. In particular, for a quasi-
coherent sheaf F on X, we have well-defined adelic spaces
X; i0 ,...,im (F ) := AX (Ki0 ,...,im , F ).
Afin (A.19)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 453
(1) (Finite adelic ring) From the Parshin-Beilinson theory for the Noetherian
scheme X, we define
X := AX;012 (OX ) = lim
Afin
−→
lim
←−
AX;12 (D1 ) AX;12 (D1 ). (A.20)
D1 D2 : D2 ≤D1
Here D∗ ’s are divisors on X and AX;12 (D) := AX;12 (OX (D∗ )) for ∗ = 1, 2;
(2) (∞-adelic ring) Associated to the regular integral curve XF over K, we obtain
the adelic ring
AXF := AXF ;01 (OXF ) = lim
−→
lim
←−
AXF ;1 (D1 ) AXF ;1 (D1 ). (A.21)
D1 D2 : D2 ≤D1
Here D∗ ’s are divisors on XF and AXF ;1 (D) := AXF ;1 (OXF (D∗ )) for ∗ = 1, 2.
By definition,
A∞ ⊗Q R . (A.22)
X := AXF ⊗ Q R := lim lim AXF ;1 (D1 ) AXF ;1 (D1 ) b
b
−→ ←−
D1 D2 :D2 ≤D1
(3) (Arithmetic adelic ring) The arithmetic adelic ring of an arithmetic surface X
is defined by
M
Aar A∞X.
ar fin
X := AX;012 := AX (A.23)
The essential point here is, for divisors Di , i = 1, 2, over the curve XF , when
D2 ≤ D1 , the quotient AX;1 (D1 ) AX;1 (D1 ) is a finite dimensional K- and hence
Q-vector space.
To help the reader understand the above formal definition in concrete terms,
we add the following examples.
Now we are ready treat adelic spaces at infinite places for general arithmetic vari-
eties. Motivated by the discussion above, we make the following
As mentioned at the beginning, for arithmetic varieties, the finite and infinite parts
are closely interconnected. Therefore, when developing an arithmetic cohomol-
ogy theory, we should treat them as an unify one. For arithmetic curves, this is
done in §2.1.2 by an uniformity condition. To treat general arithmetic varieties,
we go as follows.
Let F be a quasi-coherent sheaf on X, denote its induced sheaf on the generic
fiber XF by FF . It is well-known that F is quasi-coherent as well. Moreover,
for a point p of XF , we denote its associated Zariski closure in X by Epi . Then,
motivated by §2.1.2, we introduce the following
(1) The finite, resp. infinite, adelic space of type (i0 , . . . , im ) associated to F is
defined by
X; i0 ,...,im (F ) := AX KX; i0 ,...,im , F ,
Afin
(A.29)
A∞ , .
resp. X; i0 ,...,im (F ) := A X K X F ; i 0 ,...,im
F F
Remark A.2. (1) For any p ∈ P(XF ), OX,Ep = OXF ,p and k(X)Ep = k(XF )p . Conse-
quently, for any (p0 , . . . , pn ) ∈ S (XF )n , we have a natural morphism
A (Ep0 , . . . , Epn ), F = A (p0 , . . . , pn ), FF .
(A.32)
It is in this sense we use the relation fEp0 ,Ep1 ,...,Epn = fp0 ,p1 ,...,pn above. (In particu-
lar, if pi ’s are vertical, there are no conditions on the corresponding components.)
Clearly, this uniformity condition is an essential one, since it characters the nat-
ural interconnection between finite and infinite components of arithmetic adelic
elements.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 456
(2) When n = 0, the index set {0, . . . , î, . . . , n} is empty, so the space A∞ X; ∅ (FF )
for arithmetic curves has yet been defined. More generally, as to be seen below,
we need the spaces A∞ X; ∅ (FF ) for all arithmetic varieties. Here, to complete our
definition, for an arithmetic variety X, we view A∞ X; ∅ (FF ) as the (−1)-level of the
adelic complex for its generic fiber XF . That is to say, we define it as follows. By
p. 63 of citeY, we have the (-1)-simplex 1U for open U ⊆ X. Set then S (XF )−1 =
{1U | U ⊆ X : open}, and, for K ⊆ S (XF )−1 , let
A∞
X; ∅ (K, FF ) := { sF ∈ FF (U K,F ) ⊗ R | s ∈ F (U K ) } (A.33)
where U K := ∪1U ∈K U and sF denotes the section induced by s.
With all this, we are ready to introduce the reduced arithmetic adelic spaces.
(1) For I := iα1 <···<iαk iα1 , . . . , iαk , define the type I reduced arithmetic adelic
T
space Aar
X; I (F ) of F by
\
Aar
X; I (F ) := Aar
X; iα ,...,iα (F ); (A.34)
1 k
iα1 <···<iαk
(2) For m ≥ 0, define the m-th reduced arithmetic adelic space Aar
X; m (F ) of F by
M
ar, red ar
AX; m (F ) := AX; I (F ). (A.35)
I: |I|=m+1
Lemma A.1. For any subset I ⊂ {0, 1, . . . , n}, the type I reduced arithmetic adelic
space Aar
X; I (F ) for F above is well defined.
and dm = m
L L ar
i m
Aar
P
i=0 (−1) di : X; k0 ,...,lm−1 (F ) −→ AX; k0 ,...,km (F ).
All in all, we are now ready to introduce the following main definition.
unless i = 0, 1, . . . , n + 1.
We here give an example of the above theory for arithmetic curves, which was
previously developed in §2.1.2 based on Tate’s thesis ([113]).
Let D = ri=1 ni pi be a divisor on X = Spec OK . Write ni = ordpi (D). For
P
simplicity, we use A∗• (D) instead of A∗• (OX (D)). Then, same as in §A.1.3,
Y
X;01 (D) = (ap ) ∈
Afin ap ∈ Op ∀p ∈ S fin .
p∈S fin
(A.40)
And, since Dη = 0 is trivial,
Y
A∞ X;0 (D) = lim OXF ,η mη OXF ,η ⊗Q R = F/{0} ⊗Q R = F ⊗Q R = Fσ ,
l
←l
σ∈S ∞
Y
Aar OX (D) = Afin 01 (D) ⊕ A0 (D) = (ap ) ∈
∞
Fp ap ∈ Op ∀p ∈ S fin .
X;01 p∈S
In particular, Aar
X;01 OX (D) ' AF is independent of D.
To understand Aar fin
0 (D), we first calculate A0 (D). Same as in §A.1.3 again,
AX;0 (D) = F. Note that, from above, AX;0 (D) = F ⊗Q R. Thus, by definition,
fin ∞
X;0 (D) = (av ; aσ ) ∈ AX;0 (D) ⊕ AX;0 (D) (av ) = ifin ( f ), (aσ ) = i∞ ( f ) ∃ f ∈ F .
∞
Aar fin
In this way, we conclude that the following two arithmetic adelic complexes
coincide
d1
0 −→ Aar ar ar
X;0 (OX (D)) ⊕ AX;1 (OX (D)) −→ AX;01 (OX (D)) −→ 0
d1 (A.41)
0 −→ F ⊕ Aar
X;1 (OX (D)) −→ AF −→ 0
d1 : (a0 , a1 ) 7→ a1 − a0 .
Therefore,
Har0 F, OX (D) = F ∩ Aar
X;1 (OX (D)),
(A.42)
Har1 F, OX (D) = AF F + Aar .
X;1 (O X (D))
This coincides with the definition of Har0 F, g in §2.1.2, where g = (gp ; gσ ) ∈
GL1 (AF ) such that gσ = 1 ∀σ ∈ S ∞ and ordp (gp ) = ordp (D) ∀p ∈ S fin , i.e. the
P
associated divisor D(g) := p∈S fin ordp (gp ) p coincides with D.
As it stands, all above is compatible with cohomology theory for arithmetic
curves developed in §2.1.2, even a much more refined theory is developed there.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 459
Appendix B
459
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 460
resF : Ωcts
F/kF −→ kF , ω = f dt 7→ coeftt−1 ( f ). (B.2)
By [78], this is well defined, i.e. independent of the choice of t. Moreover, for a
finite field extension F 0 /F, we have the following commutative diagram
resF 0
Ωcts
F 0 /kF −
−−−−→ kF 0
TrF 0 /F ↓ ↓ TrkF0 /kF (B.3)
resF
Ωcts −−−→ kF .
F/kF −
which takes the values in Fπ(x) , the local field of K at the place π(x). Recall also
that, following [113], we have the canonical character
TrFπ(x) /Q p
λπ(x) : Fπ(x) −−−−−−−→ Q p −→ Q p /Z p −→ Q/Z ,→ R/Z ' S1 . (B.10)
Introduce accordingly
is a finite direct sum of local fields R((t)) and C((t)). Hence, similarly, for each
σ ∈ S ∞ , we have the associated residue maps resP,σ . Define
Here, as in [113], to make all compatible, we set λσ (x) = −TrFσ /R , i.e. with a
minus sign added.
Proof. Write the summation C⊆X,x∈C: π(x)∈S fin ResC,x ( fC,x gC,x ω) as double sum-
P
mations C⊆X x∈C: π(x)∈S fin ResC,x ( fC,x gC,x ω). Then, by Example A.2, for all but
P P
finitely many curves C, ResC,x ( fC,x gC,x ω) = 0. So it suffices to show that for a
fixed curve C, x∈C: π(x)∈S fin ResC,x ( fC,x gC,x ω) is finite. This is well known.
P
B.2.2 Non-Degeneracy
Proof. Let g ∈ Aar X be an adelic element such that, for all f ∈ AX , h f, giω = 0. We
ar
show that g = 0.
Rewrite the summation in the definition of h·, ·iω according to prime horizontal
curves E Par associated to closed points P ∈ XF and prime vertical curves V ⊆ X
appeared in the fibers of π. Namely, P∈XF x∈EPar + V∈X x∈V . Then, note that,
P P P P
for a fixed adelic element g = (gC,x , gP,σ ) ∈ Aar
X,
Now assume, otherwise, that g , 0. There exists either some vertical curve
C such that 0 , gC,x ∈ k(X)C,x , or, a certain algebraic point P of the generic
ˆ
fiber XF such that 0 , LgP,σ ∈ k(XF )P ⊗F Fσ . In case gC,x , 0, by defini-
tion, gC,x = (gCi ,x ) ∈ i k(X) Ci ,x = k(X) C,x , where C i runs over all branches
of C, and k(X)Ci ,x is a two-dimensional local field. Fix a branch Ci0 such that
gCi0 ,x , 0. By definition, ResCi ,x := λπ(x) ◦ resCi ,x . So we can choose an el-
ement h ∈ k(X)Ci0 ,x such that ResCi0 ,x (hω) , 0. For such h, we then take
fCi0 ,x ∈ k(X)Ci0 ,x such that fCi0 ,x gCi0 ,x = h. Accordingly, if we construct an adelic
element K by taking all other components fCi ,x to be zero but the fCi0 ,x , then we
have h f, giω = ResCi0 ,x ( fCi0 ,x gCi0 ,x ω) = ResCi0 ,x (hω) , 0. This contradicts to
our original assumption that h f, giω = 0. Hence, all the components of g corre-
sponding to vertical curves are zero. Since we can use the same argument for the
components of g corresponding to algebraic points of XF to conclude that all the
related components are zero as well, so g = 0. This completes the proof.
Let K be a number field with OK the ring of integers, and let π : X → Spec OK
be an arithmetic surface with XF its generic fiber. Then X is a regular integral
2-dimensional Noetherian scheme, π is proper, and XF is a geometrically con-
nected, regular, integral projective curve defined over K. Our purpose here is
to introduce certain level two intrinsic subspaces of the arithmetic adelic space
∞
Aar ar
X := AX,012 := AX ⊕ AX .
fin
To start with, we analyze the structures of Afin X,01 , one of the level two subspaces
of Afin
X = A fin
X,012 . By definition, see e.g. [91], an element ( fC,x )C,x ∈ Afin
X belongs
to Afin
X,01 , if, for all curves C, the partial components ( f )
C,x x∈C are independent of
fin
x. So we may simply write elements of AX,01 as ( fC )C . On the other hand, with
respect to π, curves on X may be classified as being either vertical or horizontal.
Therefore, we may and will write ( fC )C = ( fC )C: ver × ( fC )C: hor . Accordingly, set
X,01 , resp. AX,01 denotes the collections of ( fC )C: ver , resp. ( fC )C: hor . Fur-
where Afin,v fin,h
M M
Aar =A fin ar fin
⊗K R ,
X,02 X,02 k(X F ), and A X,12 (D) := A X,12 (D) AXF (DF )b
where
Here we have used the natural imbedding k(X) = k(XF ) ,→ AXF ,→ Aar
X.
Accordingly, we then also obtain three level one subspaces
X,0 = k(X);
(1) Aar
X,1 (D) = {( fC )C,x × ( fP ) ∈ AX : fP ∈ AX,1 (D), fE P = fP ∀P ∈ XF };
(2) Aar ar
0 → Aar ar ar ar ar ar ar
X,0 ⊕ AX,1 (D) ⊕ AX,2 (D) → AX,01 ⊕ AX,02 ⊕ AX,12 (D) → AX,012 → 0
Proof. The first three are direct consequences of the construction, while (4) is
standard from homotopy theory. Indeed, we only need to write down the boundary
maps: the first is the diagonal embedding, the second is given by (x0 , x1 , x2 ) 7→
(x0 − x1 , x1 − x2 , x2 − x0 ) and the final one is given by (x01 , x02 , x12 ) 7→ x01 + x02 −
x12 .
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 465
One checks easily that our definitions here are specializations of the definitions
in §A.2.3 for the line bundle F = OX (D) over arithmetic surface X.
For late use, we here establish a fundamental property for the level two arithmetic
adelic subspaces introduced above. In fact, as we will see below, this property, in
turn, characterizes these subspaces.
Fix a non-zero rational differential ω on the arithmetic surface X. Then, by
Proposition B.1, we have a natural non-degenerate pairing h·, ·iω on Aar X . For a
ar
subspace V of AX , set
X hw, viω = 0 ∀v ∈ V
V ⊥ := w ∈ Aar
(B.21)
Proposition B.2. Let X be an arithmetic surface, let D be a Weil divisor and let ω
be a non-zero rational differential on X. Denote by (ω) the divisor on X associated
to ω. Then
⊥
(1) Aar
X,01 = Aar
X,01 .
⊥
(2) A ar
=A .
ar
X,02 ⊥ X,02
(3) AX,12 (D) = Aar
ar
X,12 ((ω) − D).
Proof. By an abuse of notation, we will write elements of Afin X as ( fC,x )C,x or even
( fC,x ), and elements of A∞ X as ( f P ) P or even ( f P ).
⊥
We begin with a proof of Aar ar ar
X,01 ⊆ (AX,01 ) . Let f, g ∈ AX,01 . By definition,
f = ( fC )C,x × ( fP )P , g = (gC )C,x × (gP )P , and, for all algebraic points P of XF ,
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 466
Here E P denote the complete arithmetic curve associated to the algebraic point
P ∈ XF and in the last step, we have used our definition condition fEP = fP and
gEP = gP for elements f, g in Aar X,01 . Now, by the residue theorem for horizontal
curves E P (Theorem 5.4 of [79]), Q∈E P ResQ ( fP gP ω) = 0. Therefore, hf, giω = 0,
P
⊥
and Aar ar
X,01 ⊆ (AX,01 ) .
⊥
Next, we show that Aar ar
X,01 ⊃ (AX,01 ) , based on the following ind-pro structure
ar
on AX,01 in Corollary B.1
Let C be an irreducible curve C. Then, induced by the perfect pairing h·, ·iω :
Aar ar
X × AX → R/Z, for any divisor D on X, we obtain a pairing
X,1 (D) AX,1 (D − C) ' AX,1 ((ω) + C − D) AX,1 ((ω) − D) ' AC,0 = k(C), (B.24)
Aar
ar ar ar ar
ar ar
we conclude that h·, ·iωC ,a : AC,01 ×AC,01 → F p annihilates k(C)×k(C). But h·, ·iωC ,a
ar ar
can be identified with the canonical residue pairing h·, ·i : AC,01 × AC,01 → Fp
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 467
ar ⊥
where, to save space, we set A := Aar X,1 , B := AX,12 and G := AX,01 . Con-
ar
sequently, the vertical map in the middle is surjective. On the other hand, since
ar ⊥ 0
Aar
X,01 ⊆ AX,01 , this same map is also injective. Therefore, for any D ≤ D,
0 ⊥
0 ar
Aar
ar
X,1 (D) AX,1 (D ) ' Aar
X,1 ((ω) − D ) AX,1 ((ω) − D)
Consequently, we have
⊥
Aar
X,01 = Aar
X,01 ,
X,01 = lim0 lim0 AX,1 (D) AX,1 (D ) = lim lim0 AX,1 ((ω) − D ) AX,1 ((ω) − D).
0 0 ar
Aar ar ar ar
←−D −→D −→D ←−D
D0 ≤D D0 ≤D
for some f ∈ k(XF ) (since, by definition, the 02 type adeles are independent of
one dimensional curves). Thus, for f, g ∈ Aar X,02 , we have
X X
hf, gi = ResC,x ( fC,x gC,x ω) + ResP ( fP gP ω)
C,x P
X X X
= ResC,x ( f x g x ω) + ResP ( f gω).
x C:C3x P
3 Itis well-known that, see e.g. [46], if χ is a non-zero character on AC,01 ar such that χ(k(C)) = {0},
then the induced pairing h·, ·iχ : AC,01 × AC,01 → Fq ; (f, g) 7→ χ(f · g) is perfect and k(C)⊥ = k(C).
ar ar
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 468
Note that for a fixed x, f x and g x are fixed. Thus, by the residue theorem for the
point x (Theorem 4.1 of [78]), we have C:C3x ResC,x ( f x g x ω) = 0. So
P
X X X
hf, gi = 0+ ResP ( f gω) = ResP ( f gω).
x P P
hf, gi = 0, X,02 .
∀ f, g ∈ Aar
⊥
Therefore, Aar ar
X,02 ⊂ (AX,02 ) .
⊥
For the opposite direction Aar ar
X,02 ⊃ (AX,02 ) , similarly as in (1), we, in theory,
ar
can use the following ind-pro structure on AX,02 :
However, due to the lack of details for horizontal differential theory in literature,
we decide to first use this ind-pro structure to merely prove the part that if f =
X,02 ) , for vertical C’s, fC,x = f x ; and then to take a more
⊥
( fC,x )C,x × ( fP )P ∈ (Aar
classical approach for the rest.
Choose f = ( fC,x )C,x × ( fP )P ∈ (Aar ⊥
X,02 ) . Then by our assumption, for any
element g = (g x )C,x × (g)P ∈ Aar X,02 , we have
X X X
0 = hf, gi = ResC,x ( fC,x g x ω) + ResP ( fP gω). (B.26)
x C:C3x P
Now note that the element (g x )C,x ∈ Aar X,02 and the element g ∈ k(XF ) can
be changed totally independently, we conclude that both of the summations,
x C:C3x ResC,x ( fC,x g x ω) and
P P P
P ResP ( fP gω), are constants independent of g.
P
This then implies that both of them are 0. Indeed, since P ResP ( fP gω) is a con-
stant independent of g, by choosing g to be constant function, we conclude that
P
P ResP ( fP gω) ≡ 0, ∀g ∈ k(XK ). Consequently, by (B.26),
X X
ResC,x ( fC,x g x ω) ≡ 0, ∀(g x ) ∈ Aar
X,02 . (B.27)
x C:C3x
To end the proof, we need to show that fC,x ’s are independent of C and fP are
independent of P. First, we treat the case when C is vertical. As said above, we
will use the associated ind-pro structures. So, assume for now that C is vertical.
Then, for any divisor D on X, we have
Aar
ar
X,2 (D) AX,2 (D − C) 'AC,1 (D|C ),
X,02 = lim0 lim0 AX,2 (D) AX,2 (D ) = lim lim0 AX,2 ((ω) − D ) AX,2 ((ω) − D),
0 0 ar
Aar ar ar ar
←−D −→D −→D ←−D
D0 ≤D D0 ≤D
The first sum is zero, since we have (B.27) by our choice of f. The second sum,
being taken over all prime curves passing through x0 , is zero as well, since we can
apply the residue theorem for the point x0 as above (with fH,x0 being independent
of C). That is to say,
X
0
ResC,x0 ( fC,x g ω) = 0.
0 x0
C:C3x0
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 470
Now, applying the above existence to obtain a g satisfying that for any fixed
rational function f0 and any fixed curve C0 3 x0 , we have
ordC ( fC,x0 ) + ordC ( f0 − g) + ordC (ω) ≥ 0, C = C0
0
Since the last quantity is always zero for all f0 , this then implies that fC,x 0
0
= 0,
namely, fC0 ,x0 = fH,x0 .
To end the proof of (2), we still need to show that fP = fP0 for a fixed P0 ∈ XK
and all P ∈ XK . But this is amount to a use of a similar argument just said again,
based on Chinese remainder theorem. We leave the details to the reader. Thus, if
f = ( fC,x )C,x × ( fP )P ∈ (Aar
X,02 ) , then fC,x = fC0 ,x and fP = fP0 for fixed C 0 and P0 .
⊥
⊥
Finally, we prove (3). The inclusion Aar ar
X,12 ((ω) − D) ⊆ (AX,12 (D)) is easy.
Indeed, for elements f = ( fC,x )×( fP ) ∈ AX,12 ((ω)−D), g = (gC,x ) × (gP ) ∈ Aar
ar
X,12 (D),
we have hf, gi = C,x ResC,x ( fC,x gC,x ω) + P ResP ( fP gP ω) = 0, since every term
P P
in each of these two summations is zero by definition.
⊥
To prove the other direction Aar ar
X,12 ((ω) − D) ⊃ (AX,12 (D)) , we make the fol-
lowing preparations. Set e π : X → Spec OF → Spec Z. Then, by Theorem 5.7
of [78], the dualizing sheaf ωeπ of eπ is given by, for an open subset U ⊆ X,
ωeπ (U) = ω ∈ Ωk(X)/Q | ResC,x ( f ω) = 0 ∀x ∈ C(⊆ U), ∀ f ∈ OX,C .
By a similar argument as in [78] (used to prove the above result), we have, for a
fixed curve C0 ,
ωeπ,C0 = ω ∈ Ωk(X)/Q | ResC0 ,x ( f ω) = 0 ∀ x ∈ C0 , f ∈ OX,C0 .
X X
0 = hf, gi = ResC,x ( fC,x gC,x ω) + ResP ( fP gP ω).
C,x P
Thus, by the independence of components of adeles, we see that, for fixed (C, x)
and P, ResC,x ( fC,x gC,x ω) = 0 and ResP ( fP gP ω) = 0. This, together with the equiv-
alence between (a) and (b) and the equivalence between (i) and (ii), we conclude
that f = ( fC,x ) × ( fP ) ∈ Aar
X,12 ((ω) − D), as desired.
B.4.1 Definitions
Proposition B.3.
.
Har2 (X, D) = Aar
X,012 AX,01 + AX,02 + AX,12 (D) .
ar ar ar
. ar
X,02 ∩ AX,01 +AX,12 (D)
' Aar X,02 +AX,02 ∩AX,12 (D) .
ar ar
AX,01 ∩Aar ar ar
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 472
Proof. This is an arithmetic analogue of [91]. Indeed, (1) comes directly from the
definition, and via the first and second isomorphism theorems in elementary group
theory, (2) is obtained using a direct group theoretic calculation for our arithmetic
adelic complex.
Cohomology groups for the divisor D defined here coincide with the adelic
global cohomology defined in the previous section associated to the invertible
sheaf OX (D). That is, Har0 (X, D) = Har0 (X, OX (D)). In fact, our general construction
in the previous section is a higher dimensional generalization of our constructions
for arithmetic surfaces here in this section, which itself is motivated by that in
§B.27 for arithmetic curves.
Just like the classical cohomology theory, our arithmetic cohomology groups also
admit an inductive structure related to vertical geometric curves on arithmetic
surfaces. More precisely, we have the following
(b) AX,1 (D + C)/AX,1 (D) = k(C) as there are neither changes along horizontal
ar ar
(x, y) 7→ x − y,
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 473
we conclude that
(d) Ker φ is given by
X,1 (D + C)/AX,1 (D) ∩ AX,2 (D + C)/AX,2 (D) = Har (C, (D + C)|C ),
Aar ar ar ar 0
= Afin
X,12 (D + E P )/AX,12 (D) ⊕ AXF (DF + P)/AXF (DF ) ⊗Q R
fin
= AEP (Dfin + E P )|EP ⊕ H 0 (XF , DF + E P ) ⊗ R/H 0 (XF , DF ) ⊗ R.
Similarly, we have the corresponding morphism
ϕ
X,1 (D + E P )/AX,1 (D) ⊕ AX,2 (D + E P )/AX,2 (D) −→ AX,12 (D + E P )/AX,12 (D)
Aar ar ar ar ar ar
(x, y) 7→ x − y,
and hence obtain the following proposition in parallel.
However, unlike for vertical curves, we do not have the group isomorphisms
between Ker ϕ, reps. Coker ϕ, and Har0 (E P , (D+E P )|EP ), reps. Har1 (E P , (D+E P )|EP ).
This is in fact not surprising: different from vertical curves, for the arithmetic
cohomology, there is no simple additive law with respect to horizontal curves
when count these arithmetic groups: In Arakelov theory, we only have
1
χar (X, D + E P ) = χar (X, D) + χar (E P , (D + E P )|EP ) − dλ (E) (B.28)
2
with discrepancy − 21 dλ (E) resulting from the Green functions. (See e.g. p.114
of [67].)
On the other hand, recall that, on generic fiber XF , we have the long exact
sequence of cohomology groups
and that, in Arakelov theory, see e.g. Chapter VI, particularly, p.140 of [67],
what really used is the much rough version λ(DF + P) ' λ(DF ) ⊗ OXF (D + P)|P
where λ denotes the Grothendieck-Mumford determinant. One checks that the
exact sequence (B.29) does appear in our calculation (B.28) above. Indeed, for
the curve XF /F,
H 0 (XF , DF ) = k(XF ) ∩ AXF (DF ) & H 1 (XF , DF ) = AXF /k(XF ) + AXF (DF ).
(Note that AXF (DF + P)/AXF (DF ) is supported only on P.) Clearly, all this can be
read from the calculations in (a,b,c) and the morphism ϕ above. So our construc-
tion offers a much more refined structure topologically.
X = lim
Aar
lim AX,12 (D) AX,12 (E). (B.30)
−→D ←−E:E≤D
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 475
Note that AX,12 (D) AX,12 (E)’s are locally compact topological spaces. Conse-
quently, as explained in the appendix, induced from the projective limit, we get a
natural final topology on
AX,12 (D) = lim AX,12 (D) AX,12 (E).
(B.31)
←−E:E≤D
Similarly, induced from the projective limit, we get a natural initial topology on
Moreover, by Theorem II, we know that all three level two subspaces Aar 01 , A02
ar
ar ar
and A12 (D) are closed in AX . Consequently, we obtain natural topological struc-
tures on arithmetic cohomology groups Hari (X, OX (D)) induced from the canonical
topology of AarX . Recall that, for a topology space, its topological dual is defined
by bT := {φ : T → S1 continuous} together with compact-open topology.
This theorem is proved in the following chapter, after we expose some basic
structural results for the ind-pro topology on Aar
X.
b2530 International Strategic Relations and China’s National Security: World at the Crossroads
Appendix C
Theorem C.1. Let X be an arithmetic surface. Then with respect to the canonical
ind-pro topology on Aar
X , we have
(1) Level two subspaces Aar X,01 , AX,02 and AX,12 (D) are closed in AX ;
ar ar ar
ar
(2) AX is a Hausdorff, complete, and compact oriented topological group;
(3) Aar
X is self-dual. That is, as topological groups,
X ' AX .
ar
Aar
d
To begin with, let us recall some basic topological constructions for inductive
limits and projective limits of topological spaces.
477
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 478
Projective topology is also called the initial topology since it is the coarsest
topology on G such that πn : G → Gn are continuous.
Lemma C.1.
πn,m are surjective. So fn (xn ) = gn (xn ) for all xn ∈ Pn . This means that fn = gn for
sufficiently large n. Consequently, lim fn ≡ lim gn , and hence ϕ is injective.
−→n −→n
To show that ϕ is surjective, let f : lim Pn → S1 be a continuous map. Then,
←−n
for any open subset U ⊂ S1 containing 1, f −1 (U) is an open neighborhood of
0 in lim Pn . Hence, there exists an open subgroup V of lim Pn such that V ⊆
←−n ←−nY
f −1 (U). Being a subgroup, we may write V as V = lim Pn ∩ Kn where Kn ⊆
←−n
α
Pn are open subgroups and Kn = Pn for almost all n. However, by the continuity,
f (V) = 1. So, there exists a sufficiently large N such that f ( lim Pn ) = 1 for
←−n:n≤N
all n ≥ N. Built on this, we define, for n ≥ N, the maps fn : Pn → S1 , xn 7→
f (x) if πn (x) = xn . Note that f (x) always make sense, since πn : lim Pn → Pn
←−n
is surjective. Moreover, fn ’s are well defined. Indeed, if y ∈ lim Pn such that
←−n
πn (y) = xn , then πn (y) = xn = πn (x) for n ≥ N. Hence x − y ∈ V. This implies that
f (y) = f (x). Clearly, by definition, ϕ(lim fn ) = f . So ϕ is surjective.
−→n
(3) and (4) are direct consequences of the bijectivity of ϕ. Indeed, to prove
that ϕ is continuous, it suffices to show that for open subsets of lim
[ Pn in the
←−n
form U = W(K, 0), ϕ−1 (U) is open in lim d Pn , where K is a compact subset of
−→n
lim Pn . Since πn,m are continuous, Kn = πn (K) are compact. In this way, we get an
←−n
inductive system of open subsets {W(Kn , 0)}n . Set U = lim W(Kn , 0). Note that,
−→n
by the bijectivity of ϕ, f (U) = U. This shows that K is continuous.
To prove ϕ is open, let U be a subset of lim cPn . By definition, there exists n
−→n
such that U = ιn W(Kn , 0) for a compact subset Kn of Pn . But πn : lim Pn → Pn
−1
←−n
is open, K := π−1
n (Kn ) is compact in lim Pn . Consequently, W(K, 0) is open in
←−n
[
lim Pn . Note that, from the bijectivity of ϕ, we have ϕ(lim W(Kn , 0)) = W(K, 0).
←−n −→n
So ϕ is open. This proves the lemma and hence also the proposition.
Lemma C.2.
Proof. (1) Note that for n < n0 , fn = fn0 ◦ ιn,n0 and xn0 = ιn,n0 (xn ). Consequently,
fn (xn ) = fn0 ◦ ιn,n0 (xn ) = fn0 (ιn,n0 (xn )) = fn0 (xn0 ). So K is well defined.
To prove that K is continuous, let U ⊂ S1 be an open subset and
lim xn ∈ f −1 (U). By definition, xn ∈ fn−1 (U) =: Un and Un are open. Hence
−→n
U := lim Un is open and x = lim xn ∈ U. Moreover, f (U) ⊆ U. So K s continuous.
−→n −→n
(2) Assume that ψ(lim fn ) = f ≡ 0. Then, for any x = lim xn , f (x) = 0. This
←−n −→n
means that for all n, and xn ∈ Dn , fn (xn ) = f (x) = 0 where x is determined by the
condition that for all n0 ≥ n, xn0 = ιn,n0 xn . (Since ιn are injective, this is possible.)
Thus fn ≡ 0 and hence lim fn = 0.
←−n
Dn , let fn = f ◦ ιn : Dn → S1 . Clearly, fn is continuous.
(3) For any f ∈ lim d
←−n
So fn ∈ D cn . Moreover, for all n0 ≥ n, fn = f ◦ ιn = f ◦ ιn0 ◦ ιn,n0 = fn0 ◦ ιn,n0 . That
is, { fn }n forms a projective limit. Obviously, ψ(lim fn ) = f .
←−n
(4) This is rather involved: not only the just proved bijectivity of ψ, but
all assumptions for our injective system are used here. Let U = W(K, 0) be an
[
open subset of lim Dn , where K ⊆ lim Dn is compact. Since lim Dn is Haus-
−→n −→n −→n
dorff, and Dn are complete. Dn ⊆ lim Dn are closed. So Kn := K ∩ Dn
−→n
are compact. If lim fn is an element in ψ−1 (U) ⊆ lim Dn , we have fn ∈
←−n −→n
W(Kn , 0), and lim fn ∈ lim W(Kn , 0) ⊆ W(K, 0). But, by the bijectivity of ψ,
←−n ←−n
lim W(Kn , 0) = W(lim Kn , 0). So it suffices to show that W(lim Kn , 0) is open,
←−n −→n −→n
or the same, lim Kn is compact. This is a direct consequence of our assump-
−→n
tions. Indeed, since our inductive system if compact oriented, there exists a cer-
tain n0 such that lim Kn ⊆ Dn0 . Hence, Kn = Kn0 for all n ≥ n0 . Consequently,
−→n
lim Kn = Kn0 ⊆ Dn0 . So lim Kn is compact, since Kn0 is a compact subset of the
−→n −→n
closed subspace Dα0 .
(5) This is a direct consequence of the bijectivity of ψ. Indeed, let U ⊆ lim D
cn
←−n
be an open subset. By definition, without loss of generality, we may assume that
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 481
aiC ∈ AC,01 ,
∞
X YY
hC (aiC )πCiC aiC = 0 (iC 0); .
:= ∈ k(X)C,x
C
iC =−∞
C x:x∈C
min{iC : aiC , 0} = 0 (∀ C)
0
This gives the finite adelic space for X. Moreover, from Parshin-Osipov, see e.g.
Definition A.1, we have the infinite adelic space A∞ X := AXF ⊗Q R, and hence the
b
total arithmetic adelic space Aar
X for the arithmetic surface X:
X := AX ⊕ AX .
∞
Aar fin
(C.2)
Moreover, there are natural ind-pro structures on and hence on Afin
X , A∞
X Aar
X, since
AX = lim lim AX,12 (D) AX,12 (E),
fin
−→D ←−E:E≤D
(C.3)
AX = lim lim AXF ,1 (D) AXF ,1 (E)b
∞
⊗Q R.
−→D ←−E:E≤D
Consequently, induced from the locally compact topologies on AX,12 (D) AX,12 (E)
∞
⊗Q R, we get canonical ind-pro topologies on Afin
and AXF ,1 (D) AXF ,1 (E)b X , AX , and
ar
hence on AX . For example, by [72], a fundamental system of open neighborhood
of 0 in Afin
X is given by
∞
aiC ∈ UiC ⊆ AC,01 open subgroup
X
−iC fin
= .
h (a )π ∈ : U ∀ i 0 (C.4)
C iC C C A X iC
AC,01 C
iC =−∞
min{iC : UiC = AC,01 } ≤ 0 (∀ C)
0
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 482
Proposition C.3. For a fixed a ∈ ACar , the scalar product by a, namely, the map
a×
X −→ AX , x 7→ ax, is continuous.
Aar ar
Remark C.1. This result can be used to establish a similar result for two-dimen-
sional local fields. Indeed, since k(X)C,x = lim lim AX,12 (nC) AX,12 (mC), as a
−→n ←−m:m≤n
X = lim
subspace of Afin lim AX,12 (D) AX,12 (E), there is a natural ind-pro topol-
−→D ←−E:E≤D
ogy on k(X)C,x , induced from the ind-pro topology on AX,012 . On the other hand,
for a two-dimensional local field K, we have F = lim lim m−n
−m
F mF , where mF
−→n ←−m:m≤n
denotes the maximal ideal of K. So from the natural locally compact topologies on
the quotient spaces m−n
−m
F mF , we obtain yet another ind-pro topology on K, and
hence on k(X)C,x , since k(X)C,x is also a direct sum of two-dimensional local fields.
Induced from the same roots of locally compact topology on one-dimensional lo-
cal fields, these two topologies on k(X)C,x are equivalent. This, with the above
proposition, then proves the following
Corollary C.1. For a fixed aC,x ∈ k(X)C,x , the scalar product by aC,x , namely,
aC,x ·
the map k(X)C,x −→ k(X)C,x , α 7→ aC,x α is continuous. In particular, the scalar
product of a fixed element on a two-dimensional local field is continuous.
This result can be used to prove that, with respect to the canonical ind-pro
topology, two-dimensional local fields are self-dual as topological groups.
∞
Proposition C.4. The subspaces AX,12 (D) ⊂ Afin X and AXF ,1 (D) ⊂ AX , and the
level two subspace AX,01 , AX,02 and AX,12 (D) of AX are complete and hence closed.
ar ar ar ar
Proof. As our proof below works for all other types as well, we only
treat AX,12 (D) ⊂ Afin X to demonstrate how our arguments work. Since
AX,12 (D) AX,12 (E)’s are finite dimensional vector spaces over one-dimensional
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 483
Proposition C.5. For an arithmetic surface X, its associated adelic spaces Afin
X
and A∞
X are complete.
Proof. We will give a uniform proof for both finite and infinite cases. For this
∞
reason, we use simply A to denote both Afin X and AX , and A(D) for both AX,12 (D)
and AXF (D)⊗Q R. Clearly, it suffices to prove the following:
b
Lemma C.3. For a strictly increasing sequence {A(Dn )}n in A, lim A(Dn ) is
−→n
complete.
Proof. Let {an }n be a Cauchy sequence of A. We will show that these exists a
divisor D such that {an }n ⊆ A(D). Assume that, on the contrary, for all divisors D,
{an }n * A(D). We claim that there exists a subsequence {akn }n of {an }, a (strictly
increasing) subsequence {Dkn }n of {Dn }n , and an open neighborhood U of 0 in
lim A(Dn ) such that
−→n
If so, since akm < U + A(Dm ) and akn ∈ A(Dm ), for any n > m, akn − akm <
U. So, {akn }n is not a Cauchy sequence of lim A(Dn ). But, since {an }n is a
−→n
Cauchy sequence in A, with respect to the natural induced topology on lim A(Dn ),
−→n
{akn }n ⊆ lim A(Dn ) is a Cauchy sequence of lim A(Dn ). This is a contradiction.
−→n −→n
Therefore, there exists a divisor D such that {an }n ⊆ A(D). By Proposition 40,
A(D) is complete. So the Cauchy sequence {an }n is convergent in A(D) and hence
in A as well.
To prove the above claim, we select {akn }n and the corresponding Dkn ’s as
follows. To begin, let ak1 = a1 . Being an element of A, there always a divisor Dk1
such that ak1 ∈ A(Dk1 ). Since for all D, {an }n * A(D), there exists k2 and a divisor
Dk2 such that Dk2 > Dk1 and ak2 ∈ A(Dk2 ) − A(Dk1 ). By repeating this process, we
obtain a subsequence {akn }n of {an }, a (strictly increasing) subsequence {Dkn }n of
{Dn }n such that (i) above holds. Hence to verify the above claim, it suffices to find
an open subset U satisfying the conditions (ii) and (iii) above. This is the contents
of the following:
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 484
Claim C.1. Let {A(Dn )}n be a strictly increasing sequence and {an }n be a se-
quence of elements of A. Assume that an ∈ A(Dn ) − A(Dn−1 ) for all n ≥ 1.
Then there exists a open subset U of lim A(Dn ) such that a1 , . . . , an , · · · < U and
−→n
am+1 , . . . , an , · · · < U + A(Dm ) for all m < n.
Finite Adeles Since A(D1 )/A(D0 ) is Hausdorff, there exists an open, and hence
closed, subgroup U1 ⊆ A(D1 ) such that a1 < U1 and U1 ⊃ A(D0 ). Since A(D1 )
is complete and U1 is closed in A(D1 ), U1 is complete as well. Now, viewing in
A(D2 ), since A(D2 ) is Hausdorff, U1 is a complete subspace, so U1 is closed in
A(D2 ). Hence A(D2 )/U1 is Hausdorff too. Therefore, there exists an open and
hence closed subgroup V2,0 of A(D2 ) such that a1 , a2 < V2,0 and V2,0 ⊃ U1 . In
addition, A(D2 ) A(D1 ) is Hausdorff, there exists an open subgroup V2,1 such that
a2 < V2,1 and V2,1 ⊃ A(D2 ). Consequently, if we set U2 = V2,0 ∩V2,1 , U2 is an open
hence closed subgroup of A(D2 ) such that a1 , a2 < U2 , a2 < U2 + A(D1 ) and U2 ⊃
U1 . So, inductively, we may assume that there exists an increasing sequence of
open subgroups U1 , . . . , Un−1 satisfying the properties required. In particular, the
following quotient spaces A(Dn ) Un−1 +A(D0 ) = A(Dn ) Un−1 , . . . , A(Dn ) Un−1 +
A(Dm ), . . . , A(Dn ) Un−1 +A(Dn−1 ) = A(Dn ) A(Dn−1 ) are Hausdorff. Hence, there
are open subgroups Vn,m , 0 ≤ m ≤ n − 1 of A(Dn ) such that am+1 , . . . , an < Vn,m
and Vn,m ⊃ Un−1 + A(Dm ). Define Un := ∩n−1 m=1 Vn,m . Then U n is an open subgroup
of A(Dn ) satisfying a1 , . . . , an < Un , am+1 , . . . , an < Un + A(Dm ), 1 ≤ m ≤ n −
1 and Un ⊃ Un−1 . Accordingly, if we let U = lim Un , by definition, U is an
−→n
open subgroup of lim A(Dn ), and from our construction, a1 , . . . , an , · · · < U and
−→n
am+1 , . . . , an , · · · < U + A(Dm ), m ≥ 1.
Proposition C.6. For an arithmetic surface X, its associated adelic spaces Afin X
and A∞X are compactly oriented. That is to say, for any compact subgroup, resp.
∞
a compact subset, K in Afin
X , resp. AX , there exists a divisor D on X, resp. on XF ,
such that K ⊆ AX,12 (D), resp. K ⊆ AXF (D).
Proof. As in the previous subsection, we treat both finite and infinite places uni-
formly. Assume that for all divisor D, K * A(D). Fix a suitable D0 . By our
assumption, K * A(D0 ). Since A is Hausdorff and A(D0 ) is closed, there exists a
certain divisor D1 and an element a1 ∈ K, such that D1 > D0 , a1 ∈ A(D1 )\A(D0 ).
Similarly, since A(D1 ) is closed, A A(D0 ) is Hausdorff, we can find an open sub-
group U10 ⊆ A A(D0 ) such that a1 + A(D0 ) * A A(D0 ). Consequently, there exists
an open subgroup U1 of A such that a1 < U1 and U1 ⊃ A(D0 ). Now use (D1 , A)
instead of (D0 , A), by repeating the above construction, we can find a divisor D2 ,
an open subgroup U2 ⊂ A and an element a2 ∈ K ∩ A(D2 )\A(D1 ) such that
D2 > D1 , a2 < U2 , U2 ⊃ U1 . In this way, we obtain a sequence of divisors Dn ,
a sequence of elements an ∈ K ∩ A(Dn )\A(Dn−1 ) and a sequence of open sub-
groups Un such that Dα > Dn−1 , an < Un , Un ⊃ Un−1 . Let U = lim Un . Then U is
−→n
an open, and hence closed, subgroup of A. Consequently, K ∩ U is compact. This
is a contradiction. Indeed, since a1 , . . . , an < Un for all n, the open covering {Un }n
of K ∩ U admits no finite sub-covering.
Proof. We apply Proposition C.1, resp. Proposition C.2 to prove the first, resp.
the second, group of two isomorphisms. We need to check the conditions listed
there.
As above, we treat finite adeles and infinite adeles simultaneously. So we use
A and A(D) as in §C.1.3. Then A(D) = lim A(D) A(E). Now, for E < E 0 ,
←−E:E≤D
A(D) A(E) = A(D) A(E 0 ) ⊕ A(D0 ) A(E) −→ A(D) A(E 0 ), and the first projections
coincide with structural maps πD/E,D/E 0 : A(D) A(E) → A(D) A(E 0 ). So πD/E,D/E 0
are surjective and open. Similarly, being a natural quotient map, πD,D/E : A(D) →
A(D) A(E) are surjective and open. So, by Proposition C.1, we get the first group
of two homeomorphisms for topological groups.
To treat the second group, recall that A(D) A(E) are complete. So, their pro-
jective limits A(D)’s are complete. This implies that A(D) are closed in A, since
A is Hausdorff. On the other hand, for D < D0 , AX,12 (D) ⊆ AX,12 (D0 ). So the
structural maps ιD,D0 : A(D) → A(D0 ) and ι : A(D) → A are injective and closed.
Thus, by Proposition C.2, it suffices to show that the inductive system {A(D)}D is
compact oriented. This is simply the contents of § C.1.3 and §C.1.4. All this then
completes our proof, since the last two homeomorphisms are direct consequences
of previous four.
Proof. Since A(D)/A(E) are locally compact and hence they are self dual. Thus,
to prove this double dual properties for our spaces, it suffices to check the con-
ditions listed in Proposition C.2 for inductive systems { lim A(D) [
A(E)}E and
−→E:E≤D
in Proposition C.1 for the projective system {A(D)}
[ D . With the above lengthy dis-
cussions, all this now becomes rather routine. For example, to verify that c A is
Hausdorff, we only need to recall that S1 is compact. Still, as careful examinations
would help understand the essences of our proof above, we suggest ambitious
readers to supply omitted details.
Now we prove that the scalar product maps on adelic spaces are continuous, even
adelic spaces are not topological rings.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 487
∞
Proposition C.8. For a fixed element a of Afin
X , resp. of AX , the induced scalar
a× a×
product map φfin
a : AX −→ AX , resp. φa : AX −→ AX , is continuous.
fin fin ∞ ∞ ∞
Proof. If a = 0, there is nothing to prove. Assume, from now on, that a , 0 and
X∞
write a = (aC )C = hC (aiC )πCiC C ∈ Afin
X . Here, for each C, we assume that
iC =iC,0
aiC,0 , 0. To prove that ϕa is continuous, by our description (C.4) of the ind-pro
X , it suffices to show that for an open subgroup U = (UC )C , as
topology on Afin
an open neighborhood of 0, its inverse image ϕ−1 a (U) contains an open subgroup.
X∞ X∞
Write U = (UC ) = = hC (AC,1 (D jC ) πCiC + hC (AC,01 )πCiC ∩ Afin
X , and for
jC =−∞ jC =rC
later use, set IC := rC − iC,0 .
X∞
Let b = (bC )C = hC (bkC )πCkC C ∈ ϕ−1
a (U) ⊂ AX . Then, for each fixed C,
fin
kC =−∞
∞
X ∞
X
aC bC = hC (aiC )hC (blC −iC ) πClC . Recall that hC is simply the lifting map
lC =−∞ iC =iC,0
Y0 Y0 Y0
hC : AC,01 ' bC,x πC
O bC,x −lifting
O −−−→ bC,x . Thus if bk
O ∈ AC,01 , we al-
C
x:x∈C x:x∈C x:x∈C
X∞
hC AC,01 πCmC . Moreover, if
ways have hC (aiC )hC (blC −iC ) ∈ we write, as we
mC =0
can, that aiC ∈ AC,1 (FiC ), bkC ∈ AC,1 (EkC ) for some divisors FiC and EkC , then we
∞
X
hC AC,1 (FiC + ElC −iC ) πCmC .
have hC (aiC )hC (blC −iC ) ∈
mC =0
Now write bC = kC =−∞ + kC =IC hC (bkC )πC . We will construct the required
PIC −1
P∞ kC
(ii) To extend the range including also the degree IC − 1, choose a divisor E IC −1
such that hC AC,1 (FiC,0 + E IC −1 ) ⊆ hC AC,1 (DrC −1 ) . Then, if we choose bC ∈
∞
X Y0
hC AC,1 (E IC −1 )πCIC −1 + hC (AC,01 )πCkC ∩
k(X)C,x , aC bC ∈ UC as well.
kC =IC x:x∈C
(iii) Similarly, to extend the range including the degree IC − 2, choose a divi-
sor E IC −2 such that hC AC,1 (FiC,0 + E IC −2 ) ⊆ hC AC,1 (DrC −2 ) ∩ hC AC,1 (DrC −1 )
and hC AC,1 (FiC,0 +1 + E IC −2 ) ⊆ hC AC,1 (DrC −1 ) . Consequently, if we choose
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 488
C −1
IX ∞
X Y0
hC AC,1 (EkC )πCkC + hC (AC,01 )πCkC ∩
element bC ∈ k(X)C,x , we
kC =IC −2 kC =IC x:x∈C
have aC bC ∈ UC .
Continuing this process repeatedly, we then obtain divisors EkC ’s such that,
C −1
IX ∞
X Y0
hC AC,1 (EkC )πCkC + hC (AC,01 )πCkC ∩
for bC ∈ VC := k(X)C,x , we
kC =−∞ kC =IC x:x∈C
have aC bC ∈ UC .
Since, for all but finitely many C, rC ≤ 0 and iC,0 ≥ 0, or better, IC = rC −iC,0 ≤
0. Therefore, from above discussions, we conclude that C VC ∩ Afin
Q
X is an open
subgroup of Afin fin
φ
Q
X and a V
C C ∩ A X ⊆ U. In particular, a is continuous.
A similar proof works for φ∞ a . We leave details to the reader.
We prove next that the residue map is continuous. Fix a non-zero rational differ-
ential ω on X. Then, for an element a of Afin ∞
X , resp. AX , induced from the natural
residue pairing h·, ·iω , we get a natural map ϕa := ha, ·iω : Afin
fin
X −→ R/Z, resp.
ϕ∞a := ha, ·iω : A∞
X −→ R/Z.
∞
Lemma C.4. Let a be a fix element in Afin X , resp. AX . Then the induced map
ϕa := ha, ·iω : AX −→ R/Z, resp. ϕa := ha, ·iω : AX −→ R/Z, is continuous. In
fin fin ∞ ∞
(1) ϕ is continuous;
(2) ϕ is injective;
(3) The image of ϕ is dense;
(4) ϕ is open.
∞
Proof. (1) Let A be Afin
X , resp. AX . For an open subset W(K, 0) of A , where
c K
is a compact subgroup, resp. a compact subset, of A, let U := ϕ W(K, 0) . By
−1
A = lim lim A(D)/A(E).
Proposition C.7, c [ So we may write χ0 := h1, ·iω
←−D −→E:E≤D
as lim lim χD/E with χD/E ∈ A(D)/A(E).
[ Accordingly, write AD/E :=
←−D −→E:E≤D
A(D)/A(E), KD/E := K ∩ A(D) K ∩ A(E) and let U D/E := aD/E ∈ AD/E :
χD/E aD/E KD/E = {0} resp. an open subset V . Since, for a fixed divisor D, A(D)
is closed in A, K ∩ A(D) is a subgroup, resp. a subset, of A(D). So, for E ≤ D,
KD/E is compact in AD/E . Consequently, from the non-degeneracy of χD/E on
locally compact spaces, U D/E is an open subgroup, resp. an open subset, of A,
and U = lim lim U D/E . We claim that U is open. Indeed, by Proposition C.6,
←−D −→E:E≤D
A is compact oriented. So, for compact K, there exists a divisor D1 such that
K ⊆ A(D1 ). On the other hand, since χ0 is continuous, there exists a divisor D2
such that A(D1 + D2 ) ⊆ Ker(χ0 ). Hence U ⊃ A(D2 ). Thus, for a fixed D, with
respect to sufficiently small E ≤ D, we have U D/E = AD/E . This verifies that U
is open, and hence proves (1), since the topology of c
A is generated by the open
subsets of the form W(K, 0).
(4) This is the dual of (2). Indeed, let U ⊆ A be an open subgroup, resp.
an open subset, of A. Then U ∩ A(D) is open in A(D). Since A(D) is closed,
U D/E := U ∩ A(D) U ∩ A(E) is open in AD/E . This, together with the fact that
χ is non-degenerate on its locally compact base space, implies that KD/E :=
D/E
aD/E ∈ AD/E : χD/E aD/E · U D/E = {0} resp. an open subset V is a compact
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 490
subset, resp. a compact subset. Let K := lim lim KD/E . Since U is open,
←−D −→E:E≤D
there exists a divisor E such that A(E) ⊆ U. This implies that there exists a
divisor D such that K ⊆ A(D). Otherwise, assume that, for any D, K * A(D).
Then, there exists an element k ∈ K such that k < A(ω) − E). Hence we have
χ(k · A(E)) , {0}, a contradiction. This then completes the proof of (4), and hence
the proposition.
We end this long discussions on ind-pro topology over adelic space Aar
X with
the following main theorem.
X ' AX . X ' AX .
fin ∞ ar
[
Afin
X ' AX and A∞
d In particular, Aar
d
Proof. With all the preparations above, this now becomes rather direct. Indeed,
by Proposition C.9, we have an injective continuous open morphism ϕ : A → c A.
So it suffices to show that ϕ is surjective. But this is a direct consequence of the
fact that ϕ is dense, since both A and cA are complete and Hausdorff.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 491
Appendix D
For an arithmetic surface X and a Weil divisor D, there are natural arithmetic
cohomology groups Hari (X, OX (D)) (i = 0, 1, 2). Using ind-pro topology on adelic
space Aar 0 1
X,012 , we show that Har (X, OX (D)) is discrete, Har (X, OX (D)) is finite, and
2
Har (X, OX (D)) is compact. Moreover, we prove that all possible summations of
canonical subspaces Aar X,i (D), AX,kl (D) (i, k, l = 0, 1, 2) are closed in AX,012 , and
ar ar
X,0 , AX,1 (D), AX,01 (D), for a Weil divisor D on X. These spaces can be
spaces Afin fin fin
Afin
X,01 := {( fC )C,x ∈
AX,012 }, Afin:= {( f x )C,x ∈ AX,012 },
X,02
AX,12 (D) := ( fC,x )C,x ∈ AX,012 ordC ( fC,x ) + ordC (D) ≥ 0 ∀C ⊆ X ,
fin
X,0 := AX,01 ∩ AX,02 , AX,1(D) := AX,01 ∩ AX,12 (D), AX,2 (D) := AX,02 ∩ AX,12 (D).
Afin fin fin fin fin fin fin fin fin
491
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 492
X,01 = lim
Aar lim Aar
ar
−→ ←− X,1 (D) AX,1 (E),
D E: E≤D
(D.7)
Aar = lim lim Aar
ar
X,02 −→ ←− X,2 (D) AX,2 (E).
D E: E≤D
.
Har2 (X, D) = Aar
X,012 AX,01 + AX,02 + AX,12 (D) .
ar ar ar
(1) Aar ar
01 A1(E) is discrete;
(2) A12 (E) A1 (E)ar is compact.
ar
01 A1 (E) = lim
Aar ar
Aar ar
−→ 1 (D) A1 (E)
D: D≥E
where D runs over all Weil divisors on X such that D − E is effective. Note that
if D = E + C for an integral curve on X, either horizontal or vertical, we have
1 (D) A1 (E) ' k(C) = AC,0 , which
Aar
ar ar ar
is discrete (say, in AC,01 ). Hence, for any
ar ar
D ≥ E, the quotient space A1 (D) A1 (E) is discrete as there exist finitely many
Ci ’s such that D − E = i Ci . To complete proof, recall that topologies on our
P
adelic spaces are induced from the ind-pro one, that is, the strongest topology
making all inductive limits continuous. Consequently, as an inductive limit of
discrete spaces, Aar ar
01 A1 (E) is discrete.
(2) From exact sequence
ar 0
Aar
12 (E) A12 (E )
A1 (E) A1 (E ) → A12 (E) A12 (E ) ar ar 0 ,
ar ar 0 ar ar 0
(D.8)
A1 (E) A1 (E )
by taking projective limit on E 0 , we obtain an exact sequence
Aar ar ar ar
1 (E) → A12 (E) A12 (E) A1 (E). (D.9)
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 494
Aar (E) Aar (E 0 ) Aar
12 (E)
(One may also see this using first 12ar 12ar 0 ' Aar (E)+A ar 0 and then
A1 (E) A1 (E ) 1 12 (E )
ar
A12 (E)
12 (E) A1 (E) = lim
Aar
ar ar ar
ar ar 0 .) Hence to prove A12 (E) A1 (E) is compact,
←− A1 (E)+A12 (E )
0 0
E :E ≤E
Aar (E) Aar (E 0 )
being the projective limit, it suffices to show that the quotient spaces 12ar 12ar 0
A1 (E) A1 (E )
are compact. On the other hand, as above, for any integral curve C on X, there is
a canonical exact sequence
ar ar ar ar
0 → AC,0 → AC,01 → AC,01 AC,0 → 0 (D.10)
ar ar Aar (E) Aar (E−C)
with AC,01 AC,0 compact. That is to say, 12ar ar12 is compact, since exact
A1 (E) A1 (E−C)
(3) By definition, Har0 (X, OX (E)) = Aar 01 ∩ A02 ∩ A12 (E) = A01 ∩ A12 (E) ∩
ar ar ar ar
closed.
definition, AX,2 (E) = AX,2 (E) + H (X, OX (E)) ⊗Z R and Har0 (X, OX (E)) + Afin
ar fin 0
2 (E) =
ar fin 0
(E) A (E)+H (X,O X (E))⊗Z R
H 0 (X, OX (E)) + Afin = X,2Afin (E)+H 0 (X,O (E))
A 2
2 (E). Hence, we have Har 0
(X,OX (E))+Afin
2 (E) 2 X
0
is naturally isomorphic to H H(X,O X (E))⊗Z R
0 (X,O (E))
X
, a compact torus.
0 + A2 (E) is closed.
Corollary D.1. For a Weil divisor E on X, Aar ar
02 ∩ A01 +A12 (E)
Aar ar ar
1
Proof. Note that H (X, OX (E)) can also be written as Aar +A 2 (E) ar . By Propo-
0
01 + A12 (E) is closed. Hence, the numerator is also closed. But Har
sition D.2(1), Aar ar 1
0 + A2 (E) is closed.
is finite and Hausdorff, hence the denominator, namely, Aar ar
sition D.2(1), resp. Proposition D.3(3) below, whose proof is independent of this
Aar +Aar (D)
argument, we conclude that this latest quotient space ar ar 01 12ar ar ar
A01 +A12 (D) ∩ A01 +A02 +A12 (E)
is nothing but the topological dual of Aar (D) + H 0 (X, OX (E)) Aar
1 1 (D). So
Aar + 0 ar ar
+ 0
1 (D) H (X, O X (E)) A 1 (D) is discrete. Therefore, A 1 (D) H (X, OX (E))
is closed, as required.
To complete the proof, we next treat general E. Fix a Weil divisor D0 on X
such that D0 ≥ D, D0 ≥ E. By the special case just proved, we have Aar 1 (D) +
Aar ar 0 ar
+ ar ar
= ar
+ ar
02 ∩ A 12 (D ) is closed. Since A 1 (D) A 02 ∩ A12 (E) A 1 (D) A 02 ∩
12 (D ) ∩ A12 (E), A1 (D) + A02 ∩ A12 (E) is closed.
0
Aar ar ar ar ar
(3) By Lemma D.2, Aar (E) Har0 (X, OX (E)) is compact. With the natural surjection
2
2 (E) Har (X, OX (E)) A2 (E) + A01 A01 , this latest space is also compact.
Aar
0 ar ar ar
01 + A02 is closed;
Aar ar
(1)
(2) A02 + Aar
ar
12 (E) is closed;
(3) A01 + Aar
ar
02 + A12 (E) is closed;
ar
Aar (E)∩ Aar +Aar
Proof. (1) By Theorem D.1, H 1 (X, OX (E)) = A121 (E)ar +A012 (E)02ar , is finite. Hence, the
numerator Aar 12 (E) ∩ A01 + A12 (E) is closed since the denominator A1 (E) +
ar ar ar
A2 (E)ar is closed by Proposition D.2(3). Consequently, for any two Weil di-
visors D, E with D ≥ E, viewing as subspaces of Aar
ar
12 (D) A12 (E), the space
12 (D)∩ A01 +A02
Aar ar ar
is closed. On the other hand, by definition,
A12 (E)∩ A01 +A02
ar ar ar
is closed.
(3) By Lemma D.1(2), Aar (E) Aar
12 1 (E) is compact. With natural surjection
01 +A02 +A12 (E)
Aar ar ar
A12 (E) A1 (E) Aar +Aar +Aar (E) , the latest space is compact as well. But Aar
ar ar
01 02 1
01 ⊃
Aar
1 (E), hence the denominator is simply A ar
01 + A ar
02 . By (1), it is closed. Conse-
quently, as above, the numerator Aar 01 + A02 + A12 (E) is closed.
ar ar
(1) Aar 0
0 is discrete. In particular, Har (X, OX (D)) is discrete;
2
(2) H (X, OX (D)) is compact.
H (XF , OXF (E F )), where OXF (E F ) denotes the restriction of OX (E) to the generic
0
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 498
Here ResC,x , resp. ResP,σ , denotes the local pairing on X, resp. on Xσ . Moreover,
we have
(1) If W1 and W2 are closed subgroups of W such that the spaces W1 + W2 and
W1⊥ + W2⊥ are closed, then
(W1 + W2 )⊥ = W1⊥ ∩ W2⊥ and (W1 ∩ W2 )⊥ = W1⊥ + W2⊥ ;
(2) If W is a closed subgroup of W, then, algebraically and topologically,
(W ⊥ )⊥ = W W .
[ ⊥
and W' W
= Aar ar ar 0
X,01 ∩ AX,02 ∩ AX,12 (D) ' Har (D).
. ar
' AarX,02 ∩ AX,01 + AX,12 (D) X,02 + AX,02 ∩ AX,12 (D) .
ar ar
AX,01 ∩ Aar ar ar
Consequently,
X,02 ∩ AX,01 + AX,12 ((ω) − D)
Aar
ar ar
c
Har1 (X,[
(ω) − D) =
Aar∩ Aar X,02 + AX,02 ∩ AX,12 ((ω) − D)
ar ar
X,01
⊥ ⊥
Aar∩ Aar ∩ Aar ar
X,01 X,02 X,02 ∩ AX,12 ((ω) − D)
' ⊥ ⊥
Aar
X,02 + Aar X,01 + AX,12 ((ω) − D)
ar
Proposition D.3(3), namely, Aar X,01 + AX,02 + AX,12 ((ω) − D) is closed. As for
ar ar
the fourth morphisms, both associated quotients spaces are Hausdorff, since de-
nominators and numerators are all closed and Aar X,012 is Hausdorff. They are also
discrete since Har1 is always so. Therefore, this fourth morphism is a topological
homeomorphism, being a group isomorphism.
January 26, 2018 9:4 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 501
Appendix E
About 50 years ago, Tate [114] developed a theory of residues for curves using
traces and adelic cohomologies. Tate’s work was integrated with the K2 cen-
tral extensions by Arbarello-De Concini-Kac [2]. Reciprocity laws for algebraic
surfaces were established by Parshin in [92]. Later, these reciprocity laws were
reproved by Osipov [85], based on Kapranov’s dimension theory [50]. Osipov
constructs dimension two central extensions and hence establishes the reciprocity
law for algebraic surfaces using Parshin’s adelic theory [91]. More recently, a
categorical proof of Parshin’s reciprocity law was found by Osipov-Zhu [88]. In
essence, Osipov’s construction may be viewed as K2 type theory of central exten-
sions developed by Brylinski-Deligne [19].
However, to establish reciprocity laws for arithmetic surfaces, algebraic K2
theory of central extensions is not sufficient: while it works for reciprocity laws
around points and along vertical curves, it fails when we treat horizontal curves.
To remedy this, instead, we first develop a new theory of arithmetic central ex-
tensions, based on the fact that for exact sequence of metrized R-vector spaces of
finite dimensional
V∗ : 0 → V 1 → V 2 → V 3 → 0,
there is a volume discrepancy γ(V ∗ ). Accordingly, for an R-vector space V, not
necessary to be finite dimensional, a subspace A, and commensurable subspaces
B, C, following [2], we have the group
GL(V, A) := {g ∈ Aut(V) : A ∼ gA},
the R-line
(A|B) := λ(A/A ∩ B)∗ ⊗ λ(B/A ∩ B),
and the natural contraction isomorphism
α := αA,B,C : (A|B) ⊗ (B|C) ' (A|C).
501
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 502
Note that A/A ∩ B and B/A ∩ B are finite dimensional R-spaces, we hence can
introduce metrics on them and hence obtaining the metrized R-line (A|B). In
general, the contraction map αA,B,C of (4.3) of [2] does not give an isometry.
We denote the corresponding discrepancy by γ(αA,B,C ) and hence the isometry
αA,B,C := αA,B,C · γ(αA,B,C ) so that we obtain a canonical isometry
α := αA,B,C : (A|B) ⊗ (B|C) (A|C).
In parallel, for a pair of commensurable (metrized) subspaces A, B and A0 , B0 , the
natural isomorphism β introduced in §4.4) of [2] induced a canonical isometry
βA,B;A0 ,B0 : (A|B) ⊗ (A0 |B0 ) −→ (A ∩ A0 |B ∩ B0 ) ⊗ (A + A0 |B + B0 ).
The up-shot of this consideration then leads to the following construction of arith-
c Ar of GL is given by
metic central extension GL
c Ar := (g, a) : g ∈ GL, a ∈ (A|gA), a , 0
n o
GL
together with the multiplication
(g, a) ◦ (g0 , a0 ) := (gg0 , a ◦ g(a0 ))
where
ab := a ◦ b := αA,B,C (a ⊗ b) ∈ (A|C) ∀a ∈ (A|B), b ∈ (B|C).
Our first result is the following analogue of
Proposition I
As a special case, if we assume that the metric system is rigid, that is to say, ,
we then recover the constructions of [2].
To go further, we consider the local field R((t)) of Laurent series with R
coefficients. For each finite dimensional sub-quotient space of R((t)), we may
identify it with a standard form Vm,n := ni=m Rti for suitable integer m, n ∈ Z.
P
We assign a metric on Vm,n based on the standard Euclidean metric of R with
(ti , t j ) = δi j , ∀m ≤ i, j ≤ n. Then, for any f (t), g(t) ∈ R((t)), if we write
f (t) = tνt ( f ) f0 (t), g(t) = tνt (g) g0 (t), then f (t) R[[t]] g(t) R[[t]] can be explicitly
calculated. As a direct consequence, we have
| f0 (0)νt (g) |
νR((t)) ( f, g) := logh f, giR[[t] = log .
|g0 (0)νt ( f ) |
The pairing h f, giR[[t] is in fact the reciprocity symbol along horizontal curve
at infinity. To introduce reciprocity symbol at finite places, we start with corre-
sponding 2 dimensional local fields L, kL ((u)), kL {{u}} with kL finite extensions of
Q P . On L, there exists a discrete valuation of rank 2: (ν1 , ν2 ) : L∗ → Z ⊕ Z. With
the help of this, then we can define a reciprocity symbol
ν1 ( f ) ν1 (g)
νL : K2 (L) → Z, ( f, g) 7→ .
ν2 ( f ) ν2 (g)
With all this, we are now ready to state our reciprocity laws. Let π : X →
SpecOF be a regular arithmetic surface defined over a number field F with generic
fiber XF . Let C be an irreducible curve on X and x a closed point of X. As usual,
see e.g., [92], we obtain an Artinian ring KC,x which is a finite direct sum of two-
dimensional local fields Li . Accordingly, we define
Our proof of this theorem is as follows: for the first two, similar as in [84],
we first construct a K2 -central extension of k(X)∗ with the help of Kapranov’s
dimension theory, then we interpret νC,x as a commutator of lifting of elements in
some K2 central extension of the group K(X)∗ , and finally use the splitting of the
central extension of adeles.
During this process, we discover that the fundamental reason for our reci-
procity law along a vertical curve is the Riemann-Roch in dimension one and
the intersection in dimension two. To establish the reciprocity law for horizontal
curves, we need to construct a new arithmetic central extension, using Arakelov
intersection in dimension two and a refined arithmetic Riemann-Roch theorem for
arithmetic curves and our arithmetic adelic theory built up in [106], or the same,
Appendices A and B of this book.
For algebraic surfaces, a similar result is proved in [85]. The arguments there
works here since the objects we are treating are algebraic. Instead of a verbatim
copy, we will develop a much refined theory for the reciprocity laws along hor-
izontal curves, where both algebraic and arithmetic structures should be treated
simultaneously.
B A
subgroups A and B, we define [A|B]1 := log # A∩B − log # A∩B . One checks eas-
ily that, for open compact subgroups A, B, C, we have [A|B]1 + [B|C]1 = [A|C]1 .
Motivated by this, we call a real valued function num on the set of open com-
pact subgroups of G a numeration if for any two open compact subgroups A, B,
num(B) = num(A) + [A|B]1 . Denote by Num(G) the collection of all numerations
on G. Clearly, for a fixed real number α, if num ∈ Num(G), num + α, defined by
(num + α)(A) = num(A) + α, also belongs to Num(G). Consequently, Num(G)
admits a natural R-torsor structure. Furthermore, for a short exact sequence of
locally compact subgroups 0 → G1 → G2 → G3 → 0, there exists a canonical
isomorphisms of R-torsors Num(G1 ) ⊗R Num(G3 ) ' Num(G2 ). For a proof, see,
e.g. Proposition 2 of [85].
Next, we work with numerations for two-dimensional local fields. We will
use the natural ind-pro structure to build our constructions from that of dimension
one. Let M = i∈I Li with I a certain index set and Li some two-dimensional
Q
local fields. Induced by pro-ind topology, there is a natural product topology
on M. Let W1 and W2 be two closed subspaces of M satisfying the condi-
tion that there exists a closed subspace W ⊆ W1 ∩ W2 such that Wi /W’s are
locally compact. Accordingly, we define an R-torsor by [W1 |W2 : W]2 :=
HomZ (Num(W1 /W), Num(W2 /W)). By Lemmas 4,5 of [85], for Wi , i = 1, 2, 3
and W as above, there is a canonical isomorphisms of R-torsor [W1 |W2 : W]2 ⊗
[W2 |W3 : W]2 ' [W1 |W3 : W]2 . Consequently, using the ind-pro topology, we
define [W1 |W2 ]2 := lim [W1 |W2 : W]2 , an R-torsor. Moreover, by Proposition D.2
←−W
of [85], there is a canonical isomorphism [W1 |W2 ]2 ⊗ [W2 |W3 ]2 ' [W1 |W3 ]2 .
We apply this construction a single two-dimensional local field L appeared
above. For a fixed n, set On := un OL where OL denotes the valuation ring of
L w.r.t. ν2 , and u a local parameter w.r.t. νc . Denote Autc (L) the automorphism
group of L with respect to the ind-pro topology. Similarly, as Corollary 2 of Propo-
sition 1 in [85], we have L∗ ⊆ Autc (L). Hence, if W is a subspace of L such that
On1 ⊆ W ⊂ On2 for n2 , n1 ∈ Z, then, for g ∈ Autc (L), there exists m2 ≤ m1 such
that Om1 ⊆ gW ⊆ Om2 , since g and its inverse are continuous w.r.t. the ind-pro
topology. In particular, for any g1 , g2 , we have an R-torsor [g1 W|g2 W]2 satis-
fying g[g1 W|g2 W]2 ' [gg1 W|gg2 W]2 for g ∈ Autc (L). Consequently, there are
canonical isomorphisms [W|g1 W]2 ⊗ g1 [W|g2 W]2 ' [W|g1 W]2 ⊗ [g1 W|g1 g2 W]2 '
[W|g1 g2 W]2 . For ni ∈ [W|gi W]2 , i = 1, 2, denote by n1 g1 (n2 ) ∈ [W|g1 g2 W]2 the
image of n1 ⊗ (g1 (n2 )).
With numerations in dimension two introduced, we are ready to construct a K2
center extension. By definition, a center extension Aut
dc (L)W of Autc (L) w.r.t W is
given by the following data: set theoretically, Autc (L) = (g, n) | g ∈ Autc (L), n ∈
d
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 507
[W|gW]2 ; group theoretically, (g1 , n1 )(g2 , c2 ) := (g1 g2 , n1 g1 (n2 )). Obviously, this
π
is a central extension 0 → R → Aut dc (L) → Autc (L) → 1. Denote the associated
2-cocycle by c. Moreover, by Proposition 4 of [85], if H is a subgroup of Autc (L)
such that HW ⊆ W, the central extension splits over H.
This central extension can be used to introduce a reciprocity symbol. For gi ∈
Autc (L) (i = 1, 2), let b dc (L) such that π(b
gi ∈ Aut gi ) = gi . Set hg1 , g2 iL := [b
g1 ,b
g2 ],
since it is well known that [b g1 ,bg2 ] depends only on g1 , g2 , and can be written as
hg1 , g2 iL = c(g1 , g2 ) − c(g2 , g1 ). (See e.g. Proposition 6 of [85].) The last relation
implies that h·, ·iL is bi-multiplicative and skew-symmetric. Hence if we use the
decomposition L∗ = L∗ × uZ × UL1 with UL1 the special unit group 1 + uOL , we
have, for f, g ∈ k(X)∗ ⊆ Autc (L), h f, gi = [k(p) : k p ] · vL (K, g), where k denotes
the residue field of the prime π(p) ∈ Spec OK and νL is the one introduced in §1.
For details, see e.g. Theorem 1 of [85].
We end this discussion on K2 -central extensions with a global additivity of
reciprocity pairing. As above, let M = i∈I Li with Li two-dimensional local
Q
fields. For each i, we have similarly the subspace Oi,n . Fix a subspace Wi satisfying
Q
Oi,n1 ⊆ Wi ⊆ Oi,n2 and set W := i∈I Wi . On the other hand, let H be a group
admitting embeddings H ⊆ Autc (Li ) for all i. Assume that, for each h ∈ H, there
exists m1 , m2 ∈ Z such that Oi,m1 ⊆ hWi ⊆ Oi,m2 for all i ∈ I, and hence R-torsors
[W|hW]2 for all h ∈ H. Similarly, with this data, we obtain a central extension
0 → R → H bW → H → 1 and hence h·, ·i M on H. Moreover, if I = I1 t I2 ,
Mk := i∈Ik Lk , Wk := i∈Ik Wi , k = 1, 2, and H := k(X)∗ , we have h f, gi M =
Q Q
h f, gi M1 + h f, gi M2 . (See e.g. Proposition 7 of [85].)
We start with (1). Denote by OX,x the (completed) local ring of X at x and
KX,x := Frac OX,x (its fractional field). Clearly, there is a canonical correspon-
dence between prime ideals of OX,x and the branches of (normalizations of)
curves passing through x. For f, g ∈ k(X)∗ , we may and will choose free OX,x -
module Dk ⊆ KX,x (k = 1, 2) such that D1 ⊆ OX,x ⊆ D2 , D1 ⊆ f OX,x ⊆ D2 ,
and D1 ⊆ gOX,x ⊆ D2 . Furthermore, let I be the (finite) set of prime ideals
which are the supports of OX,x -module D2 /D1 and divisors of K and g. Let
M1 := P∈I KP,x , W1 := P∈I OP,x , M2 = P<I KP,x , W2 = P<I OP,x and
Q Q Q Q
M := M1 × M2 , W := W1 × W2 . From this data, we obtain three central ex-
tensions and reciprocity symbols. By additivity, for f, g ∈ k(X)∗ , h f, gi M =
h f, gi M1 + h f, gi M2 . Moreover, since f M2 = M2 and gM2 = M2 are invariant,
h f, gi M2 = 0. Thus, h f, gi M = h f, gi M1 = P∈I h f, giKP,x , by additivity again.
P
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 508
Similarly, for (2), let S f,g be set of closed points of C in supports of divisors
of K and g, M1 := x∈S f,g KC,x , W1 := x∈S f,g OC,x , M2 := x∈(CrS f,g ) KC,x , W2 :=
Q Q Q
Here mEP ,x denotes the maximal ideal of the local field of E P at x, and we remind
the reader that OXF (P) can be identified with m−1
XF ,P .
Definition E.1. We define the canonical additive numerations for the spaces
above by
(1) num0 mEP ,x := − log q x with q x the size of residue field of E P at x;
(2) num0 OXF (Q)|P := g∞ (Q, P), the Arakelov-Green function on X∞ evaluated
at the points of X∞ corresponding to P and Q;
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 510
(3) num0 A[[u]] un A[[u]] := n · num0 (A) for n ∈ Z and A locally compact and
numerable;
(4) num0 OXF (Q)k |P [[u]] um OXF (Q)k |P [[u]] := mk · g∞ (P, Q) for k, m ∈ Z.
Proof. With constructions above, our proofs for two isomorphisms are similar.
Hence we only give details for the first. By definition, it suffices to identify the
canonical numerations num0 for both sides. Since canonical numerations of Lau-
rent series parts for both sides are the same, we are led to identify num0 for the
coefficient part of both sides, i.e. to verify that
Y Y Z
m−n
num0 x
E ,x × OXF (Q)⊗mQ P
⊗Q R × e
b − log kskdµ
P
x∈E P Q∈XF X∞
Y −n0x
Y ⊗m0Q
Z
= num0 m E P ,x × OXF (Q) ⊗ R ×e
P Q
b − log ks0 kdµ .
x∈E P Q∈XF X∞
Arakelov intersection theory [67], is simply degar (L|E P ), the Arakelov degree of
metrized line bundle L|E P on E P . Similarly, the right hand side is the logarithm of
n0x Q
R
Q∈XF G ∞ (P, div(sF )) · e X − log ks kdµ , namely, degar (L|E P ).
Q 0 0
x∈E P q x
∞
Definition E.2. For metrized line bundles Li on X with non-zero rational sec-
tions
si , s0i (i = 1, 2), which do not vanish along E P , we define a R-torsor
W ar W ar by
L1 ,s1 L2 ,s2 2
WLar ,s WLar ,s := lim
←−
HomR Num W ar ar
L ,s
WL,s
, Num W ar ar
L ,s
WL,s
.
1 1 2 2 2 1 1 1 2
(L,s)
L,→L1 ,L2
In §E.2.2.1, the condition that s does not vanish on E P plays a crucial role, even it
does not hold in general. We next develop a general theory using residues.
Let L be a metrized line bundle on X. If s , 0 is a rational section of L, let f s
be a rational function such that s0 := s · f s−1 does not vanish along E P .
To construct W ar , we write s = s0 · uνEP (s) · f0 with f0 a rational func-
L,s
tion, and u a local parameter defining E P in X. Then s0 is a section of the
line bundle L0 := L(−νEP (s) · E P ). By [67], there is a natural metric on L0
obtained from L ⊗ OX (E P )νEP (s) . Here, as usual, OX (E P ) denotes the line bun-
dle OX (E P ) equipped with the Arakelov metric defined by g∞ (P, ·) ( [67]). Our
idea of constructing W ar is to introduce a term-to-be-defined decomposition
L,s
(νEP (s))
W ar = W ar ⊗ W ar ⊗ W ar . Since s0 and f0 do not vanish along
L,s L0 ,s0 OX (E P ),u O X , f0
E P , the spaces W ar and WOX , f0 have already been defined. Therefore, we have
L0 ,s0
to define WOX (E P ),u , or WE P ,u shortly, and justify the decomposition.
To start with, we recall the Arakelov adjunction formula. Let Kπ be the canon-
ical divisor of arithmetic surface π : X → Y := Spac OK . Set E P = Spec A for a
certain order in Frac (A), a finite extension of K and let ω0 , 0 be a rational sec-
tion of Kπ such that ω0 (E P ) . 0. Then, see e.g. Corollary 5.5 at page 99 of [67],
2
globally, K π · E P + E P = dEP /Y + dλ (E P ). Here,
Moreover, by the reside theorem, i.e. Theorem 4.1 of Chapter IV in [67], the
residue map induces a canonical isomorphism res : Kπ (E P )|EP ' WeEP /Y , where
WeEP /Y denotes the sheaf on affine curve E P associated to the A-module WEP /Y . Set
Y b x −ν x (ω0 |E P )
WEfinP ,u,ω0 := u · mE ((u)),
P ,x
x∈E P
Y ⊗−νQ (ω0 |X )
WE∞P ,u,ω0 := u · OXF Q F
P
⊗Q R,
((u))b
Q∈XF
Z
cE P ,ω := log kω0 kdµ − dλ (E P ).
0
X∞
Since, by the residue theorem above, Kπ (E P )|EP = WeEP /Y . This is the reason that
we define WEfinP ,u as above, and, as will be seen below, then gives a nice decompo-
sition for W ar .
E P ,u,ω0
= − dE P /Y + K π · E P − dλ (E P ) = K π · E P − K π + E P · E P
= EP · EP.
From this definition, canonical numeration num0 for the coefficient part is
given by
⊗−νP ( f s )
νP ( f s ) E P + c1,ar L ⊗ E P + c1,ar (OX ) · E P = c1,ar (L) · E P = degar L|E P .
Here, as usual, c1,ar denotes the first arithmetic Chern class of a metrized line
bundle. This latest degree is independent of s, s0 , f0 and ω0 . Hence we have
proved the following
Proposition E.3. For non-zero rational sections s and s0 of L, and non-zero ra-
tional sections ω0 and ω00 of Kπ such that ω0 (E P ) . 0, ω00 (E P ) . 0, we have
Num W ar Num W ar 0 0 . In particular, we may write Num W ar
simply
L,s,ω0 L,s ,ω0 L,s,ω0
as Num W ar
or Num W .
ar
L,s L
Definition E.4. For metrized line bundles Li with non-zero rational sections si , s0i
of Li (i = 1, 2), we define an R-torsor W ar W ar 2 by
L1 ,s1 L2 ,s2
ar
, Num WLar ,s WL,s .
ar
WL ,s WLar ,s 2 := lim HomR Num WLar ,s WL,s
ar
1 1 2 2 ←−:(L,s) 1 1 1 2
L,→L1 ,L2
Same proof as that for Proposition D.2 in §E.2.2.1 then proves the following:
This then already implies H 0 A Ar (L) = Har0 E P , L|EP ((u)). On the other hand,
E P ,∗
to treat the 1-st cohomology group, note that Parshin-Beilinson’s H 1 (E P , ·) is al-
ways 0, since E P is affine. Hence, Parshin-Beilinson’s H 0 (E P , ·) is exact. This
implies that for horizontal curve E P , the maps
(1) num0 Har0 E P , L|EP := h0ar E P , L|EP , the 0-th arithmetic cohomology for the
metrized invertible sheaf L|EP on the curve E P ;
(2) num0 Har1 E P , L|EP := h1ar E P , L|E P , the 1-st arithmetic cohomology for the
With all this, we are now ready to state one of our main results. We will use it
in our proof of the splitness of an arithmetic centra extension below.
Proof. It suffices to identify canonical numerations of both sides. For this, first,
by Definitions E.1 and E.5, we conclude that the Laurent series parts for both sides
are numerated in the same way. Therefore, it suffices to treat the corresponding
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 515
coefficients for both sides. By Definition E.3 and the discussion after it, we know
that num0 for the left hand side gives
On the other hand, by Definitions E.5 and E.6, num0 for the coefficient of
Num A Ar (L1 ) equals
E P ,∗
= [W ar : W ar ]2 .
OX OX (divar ( f g))
With all this, we are ready to introduce the following main definition.
f, b
Furthermore, for f, g ∈ k(X)∗ , let b g be their inverse images in k(X)∗W ar with
OX
f ,b
respect to the natural projection. Then [ b g] = b
f ◦b
g◦ bf −1 ◦ b
g−1 depends only on
f, g, and belongs to the center R since f g = g f . Define a reciprocity symbol by
f ,b
[ f, g]W ar := [ b g].
OX
E.5 Splitness
0 → R −→ k(X)∗W ar −→ k(X)∗ → 1.
OX
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 517
(2) The exact sequence (1) splits, and hence the reciprocity symbol degenerates.
Proof. (1) follows directly from the definition. As for (2), we introduce one more
0
central extension k(X)∗W ar of the group k(X)∗ by R as follows:
OX
(a) As a set, its elements are given by pairs ( f, α0 ), where f ∈ k(X)∗ and α0 ∈
HomR Num(A Ar (OX ), Num(A Ar ( f OX )) ;
E P ,∗ E P ,∗
(b) For the group law, ( f, α0 ) ◦ (g, β0 ) := f g, α0 ◦ f (β0 ) .
0
By Theorem E.2, k(X)∗W ar ' k(X)∗W ar . In addition, since h ∈ k(X)∗ takes
OX OX
Num(A Ar (OX )) to Num(A Ar (OX (divar (h)))), and hence the exact sequence ad-
E P ,∗ E P ,∗
mits a natural section induced by the element h ∈ HomR Num(A Ar (OX )),
E P ,∗
0
Num(A Ar (OX (divar (h)))) . Therefore, the central extension k(X)∗W ar splits.
E P ,∗ OX
introduce central extensions of k(X)∗ with respect to M arf,g and Marf,g to get first
the groups k(X)∗Mar and k(X)∗ f,g and hence the associated reciprocity symbols
f,g Mar
[∗, ∗] Marf,g and [∗, ∗] Marf,g . Based on the techniques in §E.3, it is rather direct to show
[∗, ∗] Mar = [∗, ∗] Marf,g + [∗, ∗] Marf,g . (E.5)
EP
pairing [∗, ∗] M such that [∗, ∗] is additive, and, if f, g ∈ k(X)∗ keep M stable, then,
with the proof of Proposition 4(4) of [85], we have [ f, g] M = 0. Consequently,
by (E.5), [ f, g] Marf,g = [ f, g] Marf,g + [ f, g] Marf,g = [ f, g]W ar = 0. Using the additivity
OX
again, we have 0 = x∈EP [ f, g]KEP ,x + Q∈XF [ f, g]KX ,Qb⊗Q R . This then proves
P P
F
A proof may be given using discrepancy for exact sequence of metrized local free
sheaves based on §6 of [2]. In fact, an experienced reader should already see this
f (P)νP (g)
from the facts that standard tame symbol on XF is given by (−1)νP ( f )νP (g) g(P) νP ( f )
and that gσ (P, div( f )) + X − log k f kdµ = − log k f (P)k. This formula is also one of
R
σ R
the reasons whey we add the factor e X − log kskdµ in the definition of W ar .
∞ L,s
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 519
Bibliography
519
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 520
520 Bibliography
[19] J.-L. Brylinski and P. Deligne, Central extensions of reductive groups by K2, Publ.
Math. Inst. Hautes Etudes Sci. 94 (2001) 5-85.
[20] W. Casselman, Introduction to the Theory of Admissible Representations of p-adic
Reductive Groups, unpublished notes.
[21] B. Conrad, Reductive group schemes, manuscripts, 2011.
[22] B. Diehl, Die analytische Fortsetzung der Eisensteinreihe zur Siegelschen Modul-
gruppe, J. reine angew. Math. 317 (1980) 40-73.
[23] H.M. Edwards, Riemann’s Zeta Function, Dover Pub., 1974.
[24] I.Y. Efrat, The Selberg Trace Formula for PSL2 (R)n , Memoirs of AMS, 359, 1987.
[25] J. Elstrodt, F. Grunewald and J.L. Mennicke, Groups Acting on Hyperbolic Space:
Harmonic Analysis and Number Theory, Springer, 1997.
[26] G.B. Folland, A course in abstract harmonic analysis, Studies in advanced mathe-
matics, CRC Press, 1995.
[27] P. Francini, The size function h0 for quadratic number fields. Journal de Theorie des
Nombres de Bordeaux 13 (2001) 125-135.
[28] R. Friedman and J.W. Morgan, Holomorphic principal bundles over elliptic curves.
II. The parabolic construction. J. Differ. Geom. 56 (2000) 301-379.
[29] A. Fröhlich and M.J. Taylor, Algebraic Number Theory, Cambridge studies in ad-
vanced mathematics 27, Cambridge Univ. Press, 1991.
[30] S.A. Gaal, Point set topology, Pure and Applied Mathematics 16, Academic Press,
1964.
[31] R. Gangolli and V.S. Varadarajan, Harmonic analysis of spherical functions on real
reductive groups, Springer, 1988.
[32] G. van der Geer & R. Schoof, Effectivity of Arakelov Divisors and the Theta Divisor
of a Number Field, Sel. Math., New ser. 6 (2000) 377-398.
[33] D.R. Grayson, Reduction theory using semistability. Comment. Math. Helv. 59
(1984) 600-634.
[34] D.R. Grayson, Reduction theory using semistability. II. Comment. Math. Helv. 61
(1986) 661-676.
[35] R.P. Groenewegen, An arithmetic analogue of Clifford’s Theorem. Journal de The-
orie des Nombres de Bordeaux 13 (2001) 143-156.
[36] SGA3, Seminaire de Geometrie Algebrique du Bois Marie, Schemas en groupes
1962-1964, Lecture Notes in Mathematics 151, 152 and 153, 1970.
[37] S.D. Gupta, On the Rankin-Selberg Method for functions not of rapid decay on
congruence subgroups, J Number Theory 120 (1997) 95-103.
[38] T. Hayashi, Computation of Weng’s rank 2 zeta function over an algebraic number
fields, J. number Theory 125 (2007) 473-527.
[39] D.A. Hejhal, On a result of Selberg concerning zeros of linear combinations of
Lfunctions. Int. Math. Res. Not. 11 (2000) 551-577.
[40] D.A. Hejhal, On the horizontal distribution of zeros of linear combinations of Euler
products. C. R. Math. Acad. Sci. Paris 338 (2004) 755-758.
[41] W. Hoffmann, Geometric estimates for the trace formula, Ann. Global Analysis and
Geometry 34 (2008) 233-261.
[42] A. Horn, Doubly stochastic matrices and the diagonal of a rotation matrix, Amer. J.
Math. 76 (1954) 620-630.
[43] A. Huber, On the Parshin-Beilinson adeles for schemes, Abh. Math. Sem. Univ.
Hamburg, 61 (1991) 249-273.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 521
Bibliography 521
[44] J.E. Humphreys, Introduction to Lie Algebras and Representation Theory, GTM 9
Springer, 1978.
[45] K. Iwasawa, Letter to Dieudonné (April 8, 1952), Advanced Studies in Pure Math.
21 (1992) 445-450.
[46] K. Iwasawa, Lecture notes at Princeton university, notes in Japanese, 1973.
[47] H. Jacquet, E. Lapid, and J. Rogawski, Periods of automorphic forms. J. Am. Math.
Soc. 12 (1999) 173-240.
[48] A. Javanpeykar and D. Loughran, Good reduction of algebraic groups and flag va-
rieties, Archiv Mathematik 104 (2015), 133-143.
[49] V.G. Kac, Infinite-Dimensional Lie Algebras, Cambridge University Press, 1990.
[50] M. Kapranov, Semi-infinite symmetric powers, arXiv:math.QA/0107089.
[51] S. Katayama, Weyl group of E6 and Weng zeta function, Thesis, Kyushu Univ. (in
Japanese) 2016.
[52] K. Kato, Existence theorem for higher local fields, Geom. Top. Monograph. 3,
Geom. Topol. Publ. (2000) 165-195.
[53] V.G. Kazakevich and A.K. Stavrova, Subgroups normalized by the commutator
group of the Levi subgroup, Zapiski Nauchnykh Seminarov POMI 319 (2004) 199-
215.
[54] G. Kempf, F. Knudsen, D. Mumford and B. Saint-Donat, Toroidal Embeddings I,
Lect. Notes in Math. 339, Springer, 1972.
[55] H. Ki, On the zeros of Weng’s zeta functions, IMRN 13 (2010) 2367-2393.
[56] H. Ki, Y. Komori and M. Suzuki, On the zeros of Weng zeta functions for Chevalley
groups, Manuscript Math. 148 (2015) 119-176.
[57] H. Kim and L. Weng, Volume of truncated fundamental domains, Proc. AMS 135
(2007) 1681-1688.
[58] H. Kim and L. Weng, Functional equation for zeta functions of (SLn , Pn−1,1 ),
preprint.
[59] Y. Komori, Functional equations for Weng’s zeta functions for (G, P)/Q, Amer. J.
Math. 135 (2013) 1019-1038.
[60] B. Konstant, On convexity, the Weyl group and the Iwasawa decomposition, Ann.
Sci. Ec. Norm. Super. 6 (1973) 413-455.
[61] M. Kontsevich and Y. Soibelman, Stability structures, motivic Donaldson-Thomas
invariants and cluster transformations, arXiv:0811.2435.
[62] T. Kubota, Elementary theory of Eisenstein series. Kodansha scientific books, Ko-
dansha, 1973.
[63] L. Lafforgue, Chtoucas de Drinfeld et conjecture de Ramanujan-Petersson. Aster-
isque 243, 1997.
[64] J.C. Lagarias, On a positivity property of the Riemann ξ-function, Acta Arithmetic
89 (1999) 217-234.
[65] J.C. Lagarias and M. Suzuki, The Riemann hypothesis for certain integrals of Eisen-
stein series, J. Number Theory 118 (2006) 98-122.
[66] S. Lang, Algebraic Number Theory, GTM 110, Springer, 1986.
[67] S. Lang, Introduction to Arakelov Theory, Springer, 1988.
[68] R. Langlands, the volume of the fundamental domain for some arithmetical sub-
groups of Chevalley groups, Proc. Symp. Pure Math. 9 (1966) 143-148.
[69] R. Langlands, Euler products, Yale Mathematical Monographs, 1. Yale University
Press, 1971.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 522
522 Bibliography
Bibliography 523
[94] I. Satake, On the compactification of the Siegel space, J. Indian Math. Soc. 20 (1956)
259-281.
[95] I. Satake, On representations and compactifications of symmetric Riemannian
spaces, Ann. of Math. 71 (1960) 77-110.
[96] I. Satake, On compactifications of the quotient spaces for arithmetically defined
discontinuous groups, Ann. of Math. 72 (1960) 555-580.
[97] A. Selberg, Old and new conjectures and results about a class of Dirichlet series,
Proceedings of the Amalfi Conference on Analytic Number Theory (Maiori, 1989)
(Salerno), (1992) 367-385.
[98] J.-P. Serre, Algebraic Groups and Class Fields, GTM 117, Springer, 1988.
[99] U. Shapira and B. Weiss, Stable lattices and the diagonal group, J. Euro. Math. Soc.
18 (2016) 1753-1767.
[100] G. Shimura, Introduction to Arithmetic Theory of Automorphic Functions, Publica-
tions of the Mathematical Society of Japan 11, Math. Soc. Japan, 1971.
[101] C.L. Siegel, Lectures on the geometry of numbers, Springer, 1989.
[102] D.M. Snow, Homogeneous vector bundles. Group actions and invariant theory, CMS
Conference Proceedings 10, (1989) 193-205.
[103] T.A. Springer, Linear Algebraic Groups, Birkhäuser, 1998.
[104] U. Stuhler, Eine Bemerkung zur Reduktionstheorie quadratischer Formen. Arch.
Math. 27 (1976) 604-610.
[105] U. Stuhler, Eine Bemerkung zur Reduktionstheorie quadratischer Formen. II. Arch.
Math. 28 (1977) 611-619.
[106] K. Sugahara and L. Weng, Arithmetic cohomology groups, arXiv: 1507.06074.
[107] K. Sugahara and L. Weng, Har1 for arithmetic surface is finite, arXiv:1603.02353.
[108] K. Sugahara and L. Weng, Arithmetic central extensions and reciprocity laws for
arithmetic surfaces, arXiv:1603.02355.
[109] M. Suzuki, A proof of the Riemann hypothesis for the Weng zeta function of rank
3 for the rationals. in The Conference on L-Functions, 175–199, World Sci. Publ.
2007.
[110] M. Suzuki, The Riemann hypothesis for Weng’s zeta function of Sp(4) over Q, with
an appendix by L. Weng, J. Number Theory 129 (2009) 551-579.
[111] M. Suzuki and L. Weng, private communications.
[112] M. Suzuki and L. Weng, Zeta functions for G2 and their zeros, IMRN (2009) 241-
290.
[113] J. Tate, Fourier analysis in number fields and Hecke’s zeta functions, Thesis, Prince-
ton University, 1950.
[114] J. Tate, Residues of differential forms on curves, Ann. Sci. Eco. Norm, Sup., 4
(1968) 149-159.
[115] A. Terras, Harmonic analysis on symmetric spaces and applications II, Springer,
1988.
[116] E.C. Titchmarsh, The theory of the Riemann zeta-function, The Clarendon Press
Oxford University Press, 1986.
[117] G.N. Watson, A Treatise on the Theory of Bessel Functions, Cambridge University
Press, 1922.
[118] A. Weil, Basic Number Theory, Springer, 1973.
[119] L. Weng, Geometry of numbers, arXiv:1102.1302.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 524
524 Bibliography
[120] L. Weng, Non-abelian zeta functions for function fields. Amer. J. Math. 127 (2005)
973-1017.
[121] L. Weng, Rank Two Non-Abelian Zeta and its Zeros, arXiv:math/0412009.
[122] L. Weng, A rank two zeta and its zeros. J. Ramanujan Math. Soc. 21 (2006) 205-266.
[123] L. Weng, Automorphic Forms, Eisenstein Series and Spectral Decompositions, in
Arithmetic Geometry and Number Theory, 123-210, World Sci. 2006.
[124] L. Weng, Geometric arithmetic: a program, in Arithmetic geometry and number
theory, 211-400, World Sci., 2006.
[125] L. Weng, A geometric approach to L-functions. in The Conference on L-Functions,
219–370, World Sci., 2007.
[126] L. Weng, Symmetries and the Riemann Hypothesis, in Advanced Studies in Pure
Mathematics 58, Japan Math. Soc., (2010) 173-224.
[127] L. Weng, Stability and Arithmetic, in Advanced Studies in Pure Mathematics 58,
Japan Math. Soc., (2010) 225-360.
[128] L. Weng, Parabolic Reduction, Stability and the Masses I: Special Linear Groups, at
https://repository.kulib.kyoto-u.ac.jp/dspace/bitstream/2433/194754/1/1826-16.pdf.
[129] L. Weng, Automorphic Forms and Automorphic L-Functions, RIMS Kokyuroku
1826, (2013) 168-179.
[130] L. Weng, Motivic Euler product and its applications, available at
http://www2.math.kyushu-u.ac.jp/ weng/writings.html.
[131] L. Weng, Distributions of zeros for non-abelian zeta functions, available at
http://www2.math.kyushu-u.ac.jp/ weng/writings.html.
[132] L. Weng, Higher rank zeta functions and Riemann Hypothesis for elliptic curves,
talk at Arithmetic and Algebraic Geometry 2013, Tokyo.
[133] L. Weng, Zeros of zeta functions for exceptional groups of type E, available at
http://www2.math.kyushu-u.ac.jp/ weng/writings.html.
[134] L. Weng and D. Zagier, Higher rank zeta functions for elliptic curves, available at
http://www2.math.kyushu-u.ac.jp/ weng/writings.html.
[135] L. Weng and D. Zagier, Higher rank zeta functions and SLn -zeta functions for
curves, available at http://www2.math.kyushu-u.ac.jp/ weng/writings.html.
[136] A. Yekutieli, An Explicit construction of the Grothendieck residue complex, Aster-
isque 208 (1992).
[137] D. Zagier, The Rankin-Selberg method for automorphic functions which are not of
rapid decay. J. Fac. Sci. Univ. Tokyo 28 (1982) 415-437.
[138] S. Zhang, Positive line bundles on arithmetic surfaces, Ann. Math.136 (1992) 569-
587.
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 525
Index
525
November 15, 2017 15:15 ws-book9x6 BC: 10723 - Zeta Functions of Reductive Groups and Their Zeros WengZeta page 526
526 Index
Index 527
528 Index