An Introduction To Automorphic Representations: Jayce R. Getz, Heekyoung Hahn
An Introduction To Automorphic Representations: Jayce R. Getz, Heekyoung Hahn
An Introduction To Automorphic Representations: Jayce R. Getz, Heekyoung Hahn
An Introduction to Automorphic
Representations
with a view toward Trace Formulae
Springer
To Angela and Adsila, with love
Preface
vii
viii Preface
not vacuous, as the theory of Jacobi forms already plays a role in automorphic
forms [BS98].
The results on orbital integrals in Chapter 17 and the simple relative
trace formula in Chapter 18 are based partially on work of the second author
[Hah09] and both authors [GH15]. However they have not appeared previ-
ously in the generality in which we prove them here. In particular we treat
arbitrary reductive groups over arbitrary global fields. Thus we allow positive
characteristic.
In the interest of full disclosure, we should say a word about set theoretic
foundations. In §17.5 we make use of some more serious algebraic geometry
involving sheaves of sets on a site, and this puts us up against the usual
set-theoretic difficulties. To handle these we assume the universe axiom and
assume (implicitly) that all our categories are small and contained in some
universe. For more details we refer to [Poo17, Appendix A]. It is likely that the
reliance on the universe axiom could be removed at the cost of complicating
the definition of the sets of torsors we consider.
The authors thank Francesc Castella, Andrew Fiori and Cameron Franc for
typsetting the first draft of the lecture notes that formed the germ of this
book, and thank Brian Conrad, Tasho Kaletha, Minhyong Kim, Jason Polak,
Leslie Saper, and Chad Schoen for many useful corrections and comments.
Many people have been generous in answering questions as this book was
written, including Avner Ash, David Ginzburg, Rahul Krishna, Erez Lapid,
Freydoon Shahidi, Michal Zydor, and any others who we may have forgotten.
The authors give special thanks to Alex Youcis for his very careful reading
of the first two chapters, which lead to the clarification of many details. The
authors also thank Ken Ono for suggesting this book project. The first author
wishes to thank the National Science Foundation for support at various times
during the preparation of this book. Part of this book was written while both
authors were members at the Institute for Advanced Study in the spring of
2018, we thank the institute for its hospitality and providing us with excellent
working conditions. Finally, we truly appreciate the encouragement of Madhi
Asgari, Peter Sarnak and many graduate students.
xi
Contents
1 Algebraic Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Affine schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Group schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Extension and restriction of scalars . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Reductive groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.6 Lie Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.7 Tori . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.8 Root data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.9 Parabolic subgroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2 Adeles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1 Adeles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2 Adelic points of affine schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.3 Relationship with restricted direct products . . . . . . . . . . . . . . . . 37
2.4 Hyperspecial subgroups and models . . . . . . . . . . . . . . . . . . . . . . . 39
2.5 Approximation in algebraic groups . . . . . . . . . . . . . . . . . . . . . . . . 42
2.6 The adelic quotient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.7 Reduction theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3 Automorphic Representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.1 Representations of locally compact groups . . . . . . . . . . . . . . . . . 53
3.2 Haar measures on locally compact groups . . . . . . . . . . . . . . . . . . 56
3.3 Convolution algebras of test functions . . . . . . . . . . . . . . . . . . . . . 58
3.4 Haar measures on local fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.5 Haar measures on the points of algebraic groups . . . . . . . . . . . . 60
3.6 Automorphic representations in the L2 -sense . . . . . . . . . . . . . . . 62
3.7 Decomposition of representations . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.8 The Fell topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
xiii
xiv Contents
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433
Chapter 1
Algebraic Groups
Abstract In this chapter we briefly recall some of the basic notions related
to affine algebraic groups. If the reader knows the definition of a reductive
algebraic group (which can be found in §1.5) then the rest of the chapter can
probably be skipped and then referred to later as needed.
1.1 Introduction
1
2 1 Algebraic Groups
That said, at some point it is not a bad idea for any student of automor-
phic representation theory (and really, any mathematician) to come to grips
with the “modern” (e.g. post 1960s) terminology of algebraic geometry and
algebraic groups. It is useful, beautiful, and despite its reputation not really
that counterintuitive or difficult. Thus we will use it in this chapter, mostly
as a means of fixing notation and conventions, and make minor use of it later
in the book.
Much of the material in this chapter can be found in one form or another in
the references in algebraic groups [Bor91, Hum75, Spr09]. Unfortunately they
were all written in the archaic language of Weil used before Grothendieck’s
profound reimagining of the foundations of algebraic geometry. The most
complete reference is [DG74], but it is hard to penetrate if one doesn’t have
solid preparation in algebraic geometry. The reference [Wat79] is more acces-
sible, but covers less. Fortunately J. Milne has reworked the three standard
references in modern language. His book [Mil17] was instrumental in the
preparation of this chapter.
(see Exercise 1.3). In other words, the category of affine schemes over k is
the category of k-algebras “with the arrows reversed.” If k is understood we
often omit explicit mention of it. By way of terminology, if X is an affine
scheme then
X(R)
is called its R-valued points.
Definition 1.2. A functor S : Algk −→ Set is representable by a ring A
if S = Spec(A). In this case we write
O(S) := A
A∼
= k[t1 , . . . , tn ]/(f1 , . . . , fm )
4 1 Algebraic Groups
This is also denoted by An , but we avoid this notation because we will use
the symbol A for the adeles (which will be defined in Definition 2.5).
Here are two other nice properties that the schemes of interest to us will
often enjoy:
Definition 1.4. An affine scheme X is reduced if O(X) has no nilpotent
elements and irreducible if O(X) has a unique minimal prime ideal, or,
equivalently, if its nilradical is prime.
It is important to note that an affine scheme can be both reduced and irre-
ducible (Exercise 1.4).
Assume for the moment that k is a field. An affine scheme Spec(A) of finite
type over k is smooth if the coordinate ring A is formally smooth. Here a
k-algebra A is said to be formally smooth if for every k-algebra B with ideal
I ≤ B of square zero and any k-algebra homomorphism A → B/I one has a
morphism A → B such that the diagram
AO / B/I
O
!
k /B
commutes.
For example, an affine scheme is smooth if it is isomorphic to an affine
scheme
Spec k[t1 , . . . , tn ]/(f1 , . . . , fn−d ) (1.5)
where the ideal in k[t1 , . . . , tn ] generated by the fi and all the (n−d)×(n−d)
minors of the matrix of derivatives (∂fi /∂xi ) is the whole ring k[t1 , . . . , tn ].
We have not and will not define open subschemes and Zariski covers of
schemes, but for those familiar with this language we remark that one can
always cover a smooth affine scheme by open affine subschemes of the form
(1.5).
We now discuss closed subschemes.
Definition 1.5. A morphism of affine schemes
X −→ Y
One checks that the sets V (S) are closed under infinite intersections and
finite unions. Thus we can define the Zariski topology on Spec(A) to be the
topology with V (S) as its closed sets. A morphism of schemes induces a
morphism of the associated topological spaces, and the image of an injective
morphism
Spec(A/I) −→ Spec(A)
is V (I) (see Exercise 1.1).
We close this section with the notion of fiber products. This is a simple
and useful method to create new schemes from existing ones. Let
X ×Y Z := Spec(A ⊗B C)
where the fiber product on the right is that in the category of sets:
By far the case that will be used the most frequently in the sequel is the so-
called absolute product in the category of k-schemes. To define it note that
every k-scheme X = Spec A comes equipped with a morphism X → Spec k,
6 1 Algebraic Groups
X × Z := X ×k Z := X ×Spec k Z.
Algk −→ Group
X(R) −→ Y (R)
Homk (k[x], R) = R.
Gm (R) = R× .
1.3 Group schemes 7
Note that GL1 = Gm . If one wishes to be coordinate free, then for any finite
rank free k-module V one can define
k0 : AffSchk −→ AffSchk0
Resk0 /k X 0 (R) := X 0 (k 0 ⊗k R)
called the Weil restriction of scalars. A priori this is just a set valued
functor on Algk , but if k 0 /k satisfies certain conditions then it is representable,
and hence we obtain a functor
S := ResC/R GL1 .
1.5 Reductive groups 9
of reductive groups over Q. One can choose the embedding so the R-valued
points of its image is a db
b a ∈ GL2 (R)
for Q-algebras R.
We now assume that k is a field and let k sep ≤ k be a separable (resp. alge-
braic) closure of k. Much of the theory simplifies in this case.
One has the following definition:
It turns out that if x ∈ G(k) has the property that r(x) is semisimple
(resp. unipotent) for some faithful representation r then it is semisimple
(resp. unipotent) for any faithful representation r.
Dn G := D(Dn−1 G)
is surjective, but
◦
ZG (k) × Gder (k) → G(k)
need not be. This is false even for k = Q and G = GL2 .
We close this section by recalling the following theorem of Mostow [Mos56],
which states that we can always break an algebraic group in characteristic
zero into a reductive and unipotent part:
Theorem 1.5.4 Let G be an algebraic group over a characteristic zero field.
Then there is a subgroup M ≤ G such that M ◦ is reductive and
G = M Ru (G).
Now that we have defined reductive groups, we could ask for a classification of
them, or more generally for a classification of morphisms H → G of reductive
groups. The first step in this process is to linearize the problem using objects
known as Lie algebras. We will return to the question of classification in §1.7
and Theorem 1.8.3 below.
[·, ·] : g × g −→ g
Morphisms of Lie algebras are simply k-module maps preserving [·, ·].
The pairing [·, ·] is known as the Lie bracket of the Lie algebra g and
the identity (1.8) is known as the Jacobi identity. We remark that if R is
a k-algebra then g ⊗k R inherits the structure of a Lie algebra over R in a
natural manner.
Let LAGk denote the category of linear algebraic groups over k and let
LieAlgk denote the category of Lie algebras over k. There exists a functor
defined by
From the definition just given it is not clear that Lie G is a Lie algebra,
or even a k-algebra, for that matter. At the moment it is only a set-valued
functor. There is one important special case where it is easy to deduce a Lie
algebra structure. Let G = GLV for a free finite rank k-module V . In this
case it is not hard to see that
IV + Xt 7−→ X.
[X, Y ] = X ◦ Y − Y ◦ X
ϕ : O(G) −→ k[t]/t2
such that the composite with the natural map k[t]/t2 → k is . Thus ϕ
maps the ideal ker() into t and hence factors through O(G)/ ker(). Since
O(G)/ ker()2 = k ⊕ ker()/ ker()2 (see Exercise 1.9) we obtain a bijection
Ad : G −→ AutLie G .
To define it, note that for any k-algebra R the morphism R −→ R[t]/t2 giving
R[t]/t2 its R-algebra structure gives rise to a map G(R) −→ G(R[t]/t2 ). Thus
the conjugation action of G(R) on itself gives rise to a conjugation action of
G(R) on G(R[t]/t2 ) which preserves Lie G(R). This is the action denoted Ad
and it is known as the adjoint representation.
14 1 Algebraic Groups
We finally give the reader a definition of the bracket on Lie G. For this we
note that by functoriality of Lie as a k-module valued functor the morphism
Ad gives rise to a morphism of free k-modules of finite rank
We then define
[A, X] = ad(A)(X).
One checks that this is indeed a Lie bracket and that when G = GLV this
recovers the earlier definition of the bracket.
In practice, to compute the Lie algebra of a linear algebraic group G one
just chooses an embedding G −→ GLV and computes Lie G in terms of the
conditions that cut the group G from GLV . One then obtains the bracket for
free; it is just the restriction of the bracket on GLV .
by Taylor expansion, so
1.7 Tori
X ∗ (G) = Hom(G, Gm ).
A co-character is an element of
T 7−→ X ∗ (T )ksep
is continuous where we give Galk the usual profinite topology and GLV (Z)
the discrete topology.
We now record a few examples of tori.
NL/k : ResL/k Gm −→ Gm
Q
be the norm map; it is given on points by x 7→ τ ∈Homk (L,k) τ (x). Then the
kernel of NL/k is an algebraic torus. When L = Q(i) and k = Q this torus is
isomorphic to the group SO2 of the previous example.
16 1 Algebraic Groups
Example 1.8. If L/k is any (separable) field extension then ResL/k (Gm ) is an
algebraic torus. Moreover one can show that:
M
X ∗ (ResL/k (Gm ))L ' Zτ
τ
Any torus T can be decomposed as T = T ani T spl where T spl is the maximal
split subtorus, T ani is the maximal anisotropic subtorus, and T ani ∩ T spl is
finite.
Proof. In the generality we are considering the first result is due to Grothendieck
(see, e.g. [Con14, Appendix A]). The second is [Bor91, IV.11.3]. t
u
In view of the second assertion of Theorem 1.7.2, the rank of a maximal torus
of G is an invariant of G; it is known as the rank of G.
For the remainder of this section G is a reductive group and T ≤ G is a
torus. Let NG (T ) be the normalizer of T in G and ZG (T ) is the centralizer of
T in G. The torus T is maximal if and only if ZG (T ) = T [Mil17, Corollary
17.84].
ResL/k Gm −→ GLn .
r : G −→ GLV
it follows from the fact that T is abelian and reductive that we can decompose
V as
18 1 Algebraic Groups
V = ⊕α∈X ∗ (T ) Vα (1.12)
Ad : G −→ GL(g). (1.13)
For example, when G = GLn this is the usual action of GLn on the space of
n × n matrices gln by conjugation.
Since Ad(T ) consists of commuting semi-simple elements the action of T
on g is diagonalizable. For a character α ∈ X ∗ (T ), let
1.8 Root data 19
where t := Lie T . It turns out that each of the root spaces gα are one dimen-
sional.
One turns the set of roots Φ(G, T ) into a combinatorial gadget as follows.
Let
V := hΦ(G, T )i ⊗Z R,
where hΦ(G, T )i ⊂ X ∗ (T )(k) denotes the (Z-linear) span of Φ(G, T ). The pair
(Φ(G, T ), V ) satisfies remarkable symmetry properties that are axiomatized
in the following:
Definition 1.23. Let V be a finite dimensional R-vector space, and Φ a
subset of V . We say that (Φ, V ) is a root system if the following three
conditions are satisfied:
(a) Φ is finite, does not contain 0, and spans V ;
(b) For each α ∈ Φ there exists a reflection sα relative to α (i.e. an involution
of V with sα (α) = −α and restricting to the identity on a subspace of V
of codimension 1) such that sα (Φ) = Φ;
(c) For every α, β ∈ Φ, sα (β) − β is an integer multiple of α.
A root system (Φ, V ) is said to be of rank dimR V , and is said to be reduced
if for each α ∈ Φ, ±α are the only multiples of α in Φ.
We will not need the notion until §1.9, but a subset
∆⊂Φ (1.15)
where the ci ’s are all integers that have the same sign. We define
Φ+ ⊂ Φ (resp. Φ− ⊂ Φ)
define a system of positive roots a priori and use it to define bases; see, for
example, [Hum78, §10].
The Weyl group of (Φ, V ) is the subgroup of GL(V ) generated by the
reflections sα :
W (Φ, V ) := hsα : α ∈ Φi ⊆ GL(V ).
The following is [Mil17, Corollary 21.38]:
Proposition 1.8.1 If (Φ, V ) is the root system associated with the split max-
imal torus T ≤ G then (Φ, V ) is reduced and
W (Φ, V ) ∼
= W (G, T )(k).
t
u
If Φ ⊂ V is a finite spanning set not containing 0 then (Φ, V ) is a root
system if and only if there is a map
Φ −→ V ∨ := Hom(V, R)
α 7−→ α∨
sα (x) := α∨ (x)α
Ψ (G, T ) = (X ∗ (T ), X∗ (T ), Φ, Φ∨ )
Φ −→ Φ∨
α 7−→ α∨ .
sα (x) := x − hx, α∨ iα
sα∨ (y) := y − hy, αiα∨ .
˜ (G0 , T 0 ) is
is bijective. Moreover, every isomorphism of root data Ψ (G, T )→Ψ
induced by an isomorphism G→G ˜ 0 sending T to T 0 , unique up to the conju-
gation actions of T (k) and T 0 (k). t
u
This theorem tells us that if we can classify root data, then we can classify
split reductive groups. In fact, Killing obtained a classification of the under-
lying root systems (up to some mistakes) in [Kil88, Kil90]. Killing’s work was
revisited and corrected in E. Cartan’s thesis.
If (X, Y, Φ, Φ∨ ) is a root datum, then so is (Y, X, Φ∨ , Φ). The associated
reductive algebraic group over C is denoted G b and is called the complex
dual of G. We note that there is an isomorphism
and !!
tk1
(k1 , . . . , kn ) 7−→ t 7→ ..
. ,
tkn
for every pair of integers (i, j), 1 ≤ i, j ≤ n with i 6= j, and the corresponding
root spaces gln,eij are the linear span of the n × n matrix with all entries zero
except the (i, j)-th component. The coroot e∨ ij associated with eij is the map
sending t to the diagonal matrix with t in the ith entry and t−1 in the jth
entry and 1 in all other entries.
The group Gk trivially has Borel subgroups, and hence G always has at
least one parabolic subgroup, namely G itself. A proper parabolic subgroup
is a parabolic subgroup of G not equal to G. In general, a reductive group
need not have proper parabolic subgroups. For example, if D is a division
1.9 Parabolic subgroups 23
algebra with center k of dimension bigger than 1 over k and G is the algebraic
group defined by
G(R) = g ∈ GL2 (C ⊗R R) : g t −1 −1 g = −1 −1
where the bar denotes complex conjugation. Then the subgroup of upper
triangular matrices in G is a Borel subgroup of G. Thus G is quasi split. It
is not, however, split.
From the optic of finite-dimensional representation theory the behavior of
a split reductive group is essentially as simple as a group over an algebraically
closed field (compare §1.7). Quasi-split groups are a little more technical to
handle, but the existence of the Borel subgroup makes the theory not much
more difficult. The fact that a general reductive group does not have a Borel
subgroup (over the base field) creates more substantial problems. In this
case one has to do with a minimal parabolic subgroup. Of course, in the
quasi-split case, a minimal parabolic subgroup is simply a Borel subgroup.
Basic facts about parabolic subgroups come up constantly in the theory of
automorphic representations. For example, they are used to understand the
structure of adelic quotients at infinity (see §2.7) and are used to describe
the representation theory of a reductive group inductively (see §4.9, §8.2 and
Chapter 10). Thus we record some of the basic structural facts about the set
of parabolic subgroups of G in this section. One reference is [Bor91, §20-21].
It is convenient to start with split tori in G. If there is no split torus con-
tained in G then G is said to be anisotropic. In this case the only parabolic
subgroup of G is G itself. Otherwise G is said to be isotropic. If G is isotropic
then there exists a maximal split torus T ≤ G unique up to conjugation. Just
as in the case when G is split (discussed in §1.8), we can decompose g under
the adjoint action (1.13) into eigenspaces under T :
M
g := m ⊕ gα (1.18)
α∈Φ(G,T )
{bases ∆ ⊆ Φ(G, T )} −→
˜ {P0 ≥ Z(T ) : P0 minimal parabolic subgroups} . t
u
Φ(J) := ZJ ∩ Φ
For the proof of the following theorem see [Bor91, Proposition 21.12]:
Theorem 1.9.2 There is a bijective correspondence
The two bijections in theorems 1.9.1 and 1.9.2 allow us to define the no-
tion of a parabolic subgroup opposite to a given parabolic subgroup. More
precisely, suppose we are given a standard parabolic subgroup P ≤ G with
1.9 Parabolic subgroups 25
−∆ := {α ∈ Φ : −α ∈ ∆}.
−J := {α ∈ Φ : −α ∈ J} ⊆ −∆.
∆ : = {ei,i+1 : 1 ≤ i ≤ n − 1}
Φ+ : = {ei,j : i < j}
where n1 is the first index such that en1 ,n1 +1 is not an element of the subset,
n2 is the second index such that en2 ,n2 +1 is not an element of the subset, etc.
Unwinding these equivalences, we see that the standard parabolic sub-
groups of GLn correspond Pd bijectively to ordered tuples of positive integers
n1 , . . . , nd such that i=1 ni = n. The corresponding parabolic subgroup is
the product of B and the block diagonal matrices of the form
x1
..
M (R) := : x ∈ GL (R)
. i n i
xd
Exercises
Spec(A/I) −→ Spec(A)
for k-algebras R.
1.9 Parabolic subgroups 27
diag : G −→ G × G
denote the diagonal map. We say that G is a group object in the category
of k-schemes if there exist morphisms of k-schemes
m : G × G −→ G
e : Spec k −→ G
inv : G −→ G
G×G×G
Id×m
/G
m×Id m
G
m /G
Spec k × G
e×Id
/ G×G o Id×e
G × Spec k
m
' w
G
(inv,Id) (Id,inv)
G / G×G o G
m
Spec k
e /Go e
Spec k
Prove that G is a group scheme if and only if it is a group object in the
category of k schemes.
1.8. For R-algebras R define
where the bar denotes the action of complex conjugation. Show that Un is
an algebraic group over R, that Un (R) is compact, and that
UnC ∼
= GLnC .
as k-algebras.
then
Lie G = {X ∈ gln : X t J + JX = 0}.
1.11. Let k be a perfect field. Prove that the set of conjugacy classes of
maximal tori T ≤ GLn/k is in natural bijection with étale k-algebras of
degree n.
1.13. Show that for a perfect field k the set of GLn (k)-conjugacy classes of
parabolic subgroups of GLn is in bijective correspondence with the partitions
of n. Here a partition of n is a nonincreasing set of positive integers whose
sum is n.
1.14. Let H ≤ G be affine algebraic groups over a perfect field k. Show that
there is an exact sequence of pointed sets
1 → H(k) → G(k) → G/H(k) → ker H 1 (k, H) → H 1 (k, G) .
Chapter 2
Adeles
2.1 Adeles
The arithmetic objects of interest in this book are constructed using global
fields and their adele rings. We recall the construction of the adeles in this
section; references include [CF86, Neu99, RV99].
To each global field F one can associate an adele ring AF . Before defining
this ring, we recall the related notions of a valuation (or a finite place) of a
global field.
29
30 2 Adeles
| · |v : F −→ R≥0 (2.1)
x 7−→ αv(x)
| · |v : F −→ R≥0
v(x) := max{k ∈ Z : x ∈ pk OF }.
|x|v = qv−v(x)
Here $v is a uniformizer for Fv , that is, a generator for the maximal ideal of
OFv .
Let us make these constructions explicit when F = Q. If p ∈ Z is a prime,
then completing Q at the p-adic absolute value gives rise to the local field
Qp . Its ring of integers is Zp and the maximal ideal is pZp . The residue field
is Zp /pZp ∼= Z/pZ ∼ = Fp , so the normalized absolute is just the usual p-adic
norm. There is only one infinite place of Q, denoted ∞, and Q∞ ∼ = R. The
normalized archimedean norm | · |∞ is the usual Euclidean norm on R.
32 2 Adeles
The normalization allows one to prove that the following product for-
mula holds:
Proposition 2.1.1 For x ∈ F × one has
Y
|x|v = 1
v
Y
FS := AF,S = Fv .
v∈S
(here |∞| is the number of infinite places of S). Consider the open subset of
AF defined by Y Y
U= Uv × $vnv .
v|∞ v-∞
34 2 Adeles
since |x−y|v ≤ |x|v for all finite places v. By the product formula we conclude
that x = y. t
u
We often identify F with its image under the diagonal embedding. Given
Lemma 2.1.3 the following theorem can be surprising the first time one sees
it:
Theorem 2.1.4 (Strong Approximation) If S is any finite nonempty set
of places of F then F is dense in ASF . t
u
Thus omitting one place is enough to move F from being discrete to being
dense. The proof can be found in any standard reference, see [Cas67, §15] for
example.
We close this section by remarking that one can construct analogues of
the finite adeles in more general situations, e.g. schemes of finite type over
the ring of integers of a global field [Hub91]. The construction is quite a bit
more involved than that given above.
In the proof and throughout this book we take the convention that locally
compact spaces are Hausdorff.
If the reader is uncomfortable with fiber products then they can omit the
assertions in the theorem involving them and their proof. The only conse-
quence of these facts we really need in the sequel is that with the definition
of the topology on Gna (R) given above the natural bijection
Gna (R)−→R
˜ n
A := O(X) ∼
= R[t1 , . . . , tn ]/I (2.6)
for an ideal I, and identify X(R) with the subset of Rn on which the elements
of I (thought of as polynomials on Rn ) vanish.
We start with uniqueness. By our assumption on compatibility with fiber
products the natural bijection
˜ n
Spec R[t1 , . . . , tn ](R)−→R
is a homeomorphism provided that we give the right hand side the product
topology. By assumption, this induces a topological embedding X(R) ,→ Rn .
This completes the proof of uniqueness and also shows that X(R) is Hausdorff
if R is Hausdorff. If R is Hausdorff, then 0 ∈ R is closed, so viewing X(R) as
the the vanishing locus of f ∈ I (viewed as polynomials on Rn ) we see that
X(R) is closed. Thus if R is locally compact X(R) then is as well.
We now prove existence. Note that there is a canonical and tautological
injection
We claim that the topology defined using (2.6) as above is the same as the
subspace topology defined by the canonical injection X(R) ,→ RA , so it
is independent of the choice of (2.6). Let a1 , . . . , an ∈ A correspond to t1
(mod I), . . . , tn (mod I) via (2.6), so the injection X(R) ,→ Rn defined by
(2.6) is the composition of the natural injection X(R) ,→ RA and the map
RA → Rn given by projection to the factors indexed by (a1 , . . . , an ). There-
fore every open set in X(R) is induced by an open set in RA because RA → Rn
is continuous. Since every element of A is an R-polynomial in a1 , . . . , an and
R is a topological ring, it follows that the map X(R) → RA is also continuous.
Thus X(R) has been given the subspace topology from RA . This completes
36 2 Adeles
the proof of the claim. It also implies that the formation of the topology on
X(R) is functorial (i.e. morphisms of affine schemes induce continuous maps
on R-points).
Consider a closed immersion
i : X := Spec A ,→ X 0 := Spec A0
X ×Y Z ∼
= (X ×R Z) × Y ×R Y
x ∈ Uα
forms a neighborhood base of the identity in GLn (Fv ). Note that this is
the same topology we would obtain if we just gave GLn (Fv ) ⊂ Mn (Fv ) the
subspace topology.
On the other hand, if S is any finite set of places of F including the infinite
places then
n o
(In + mMn (O bS )) × (1 + mO bS ) : m ⊂ OS
F F F
forms a neighborhood basis for (In , 1) ∈ Mn (ASF ) × ASF . Here m runs over
proper ideals of OFS , and
38 2 Adeles
Y
bFS ) :=
mMn (O $vv(m) Mn (OFv ).
v6∈S
If we intersect one of these neighborhoods with the image of GLn (ASF ) under
(2.8) then we obtain
Y Y
(In + mMn (OFv )) × GLn (OFv ).
v6∈S v6∈S
v(m)6=0 v(m)=0
v(m)
Here mMn (OFv ) = $v Mn (OFv ). This is not the same as the topology
obtained by giving GLn (ASF ) ⊂ Mn (ASF ) the subset topology.
Finally, for any set of places S of F including the infinite places it is not
hard to see by modifying this argument that one has a topological isomor-
phism of locally compact groups
Now assume that G is an algebraic group over the global field F . Choose
a faithful representation
G −→ GLn .
Identify G with its image in GLn and define, for all v - ∞,
where the restricted direct product is defined with respect to the subgroups
Kv . t
u
In fact, in most references G(AF ) is defined using Proposition 2.3.1. This
has the advantage of being concrete, but it makes it awkward to rigorously
prove that the topology satisfies good functorial properties.
Example 2.1. When G = GLn , it is clear that GLn (OF ) is a hyperspecial sub-
group of GLn (F ). It turns out that all maximal compact subgroups of GLn (F )
are conjugate to GLn (OF ) [Ser06, Chapter IV, Appendix 1]. In loc. cit. one
also finds a proof that GLn (OF ) is maximal in GLn (F ).
X −→ Y ; (2.9)
Since O(Y) → O(Y ) is injective and (2.9) is a closed immersion we see that
X := Spec A comes equipped with a closed immersion X → Y and XF = X.
We leave the proof of the following lemma as an exercise (see Exercise 2.4)
Lemma 2.4.2 The scheme X is a model of X and
For the proof of this lemma, see Exercise 2.5. The scheme X earns its moniker
as a schematic closure via a universal property.
If Y = GLn , viewed as an affine group scheme over O, then the condition
that O(GLn ) → O(GLnF ) is injective is clearly satisfied. Thus given any
faithful representation G −→ GLn , we can form the schematic closure G of
G; it is a scheme over O with generic fiber G. In the special case where F is
a local field and O = OF is its ring of integers we obtain
For a global field F and a nonempty finite set S of places of F , the image of F
under the diagonal embedding F → FS is dense. This is fairly easy to prove
and can be viewed as a generalization of the Chinese remainder theorem. If
X is an affine scheme of finite type over F and S is a finite set of places of
F then one can ask if a similar phenomenon holds:
Here when we speak of density we are of course using the canonical topologies
on X(FS ) and X(ASF ) afforded by Theorem 2.2.1.
Despite the relative ease of proving weak approximation when X = Ga ,
establishing whether or not weak approximation holds for a general affine
scheme is very difficult. In general it is false. We refer to [Har04] for a more
detailed discussion.
Our goal in this section is to describe when weak and strong approximation
hold in settings related to algebraic groups. To simplify the discussion we
often assume that F is a number field; additional complications come up in
the general case. Our primary reference is [PR94, Chapter 7]. We start with
the following proposition, the proof of which we leave as an exercise:
Proposition 2.5.1 Let G be a connected algebraic group over a number field
F with Levi decomposition G = M N . Then G admits weak (resp. strong)
approximation with respect to a finite set S of places of F if and only if M
does. t
u
Thus in the number field case studying strong and weak approximation of
algebraic groups is equivalent to studying it for the smaller set of reductive
groups. It turns out that under suitable restrictions on the set of places S
weak approximation always holds:
Theorem 2.5.2 Let G be a connected algebraic group over a number field
F . There is a finite set S0 of finite places of F such that G has weak approx-
imation for any finite set of places of F not containing S0 . t
u
2.5 Approximation in algebraic groups 43
X := G/H := Spec(O(G)H ).
with the natural Galois action (see Proposition 17.1.5). For more details on
algebraic group actions we refer to [MFK94] or §17.1.
To state weak approximation theorems in this context, we recall that an
algebraic torus T is quasi trivial if X ∗ (T ) is a permutation Gal(F /F )-
module. A reductive group G is quasi trivial if the torus G/Gder is quasi
trivial and Gder is simply connected. We record the following theorem of
Borovoi [Bor09, 3.12], which generalizes Theorem 2.5.2:
Theorem 2.5.4 Let G be a reductive quasi trivial group and assume that H
is a connected subgroup. There is a finite set S0 of finite places of the number
field F such that G/H has weak approximation for any finite set of places of
F not containing S0 . t
u
In particular, G/H always has strong approximation with respect to ∞.
Similarly we have a generalization of Theorem 2.5.3:
44 2 Adeles
This is [Bor09, Corollary 3.14]. We note that Borovoi actually works in a more
general context where G is not necessarily reductive, but we have restricted
to the situation above for simplicity.
We now turn to a discussion of strong approximation. We recall that a
connected algebraic group over a global field F is almost simple if Lie G is
a simple Lie algebra, that is, a Lie algebra with no proper ideals. The basic
result on strong approximation is the following. For the proof see [Pra77]:
Theorem 2.5.6 Let G be a connected absolutely almost simple algebraic
group over a global field F and let S be a finite set of places of F . If G
is simply connected and G(FS ) is noncompact then G satisfies strong approx-
imation relative to S. t
u
At least in the number field case there is a converse to this theorem, see
[PR94, Theorem 7.12]. We point out in particular that the semisimplicity
assumption is necessary. Even in the case even in the case G = GL1 strong
approximation is false in general. Indeed,
× b×
F × \(A∞
F ) /OF
Theorem 2.5.7 Assume that H ≤ G are almost simple algebraic groups that
are simply connected and semisimple. Then G/H satisfies strong approxima-
tion relative to S if and only if (G/H)(FS ) is noncompact. t
u
Let G be an affine algebraic group over a global field F (so we do not assume G
to be reductive or connected). Then the subgroup G(F ) ≤ G(AF ) is discrete
(compare Exercise 2.2) and we can consider the quotient G(F )\G(AF ). In this
section we recall basic facts on the topology of G(F )\G(AF ). More precise
results will be recalled in the following section.
The first result we recall is the following (see [Con12a] for the proof):
Theorem 2.6.1 (Borel, Conrad, Oesterlé, Prasad) For any finite set
S of places of F containing the infinite places and for any compact open
subgroup K S ≤ G(ASF ) the quotient
2.6 The adelic quotient 45
is finite. t
u
We note that one does not even have to assume that G is smooth (al-
though this is automatic in the characteristic zero case, see Theorem 1.5.2).
Theorem 2.6.1 is known as finiteness of class numbers (for affine algebraic
groups). Indeed, in the special case G = GL1 , K ∞ := O b× the set above can
F
be identified with the class group of F , and hence the theorem implies the
finiteness of the class group.
It is not hard to see that in general the quotient G(F )\G(AF ) itself is
infinite. However, we could ask when the quotient is compact or finite volume
with respect to a suitable measure (see §3.5), and there is a complete answer
to this question which we recall in Theorem 2.6.2 below. However, it is useful
to first eliminate a trivial obstruction to the quotient G(F )\G(AF ) being
finite volume or compact.
The units F × of F embed into A× F diagonally as a discrete subspace. Since
there the idelic norm | · |AF provides a continuous nontrivial homomorphism
Y
| · |AF := | · |v : F × \A×
F −→ R>0 ,
v
Here |x| := |x|AF = v |x|v . Note that G(F ) is contained G(AF )1 in virtue of
Q
the product formula Proposition 2.1.1. Moreover, G(F ) is discrete in G(AF )1
(see Exercise 2.2).
We now define a subgroup
such that
AG G(AF )1 = G(AF ),
the product being direct.
When F is a number field we let AG be the identity component of the
R-points of the greatest Q-split torus in ResF/Q ZG . When F is a function
field of characteristic p temporarily write Z for the largest Fp (t)-split torus
in the center of
ResF/Fp (t) ZG .
46 2 Adeles
[G]−→G(F
˜ )AG \G(AF ) (2.13)
that intertwines the action of G(AF )1 on the left hand side and the action
of AG \G(AF ) on the left. We will therefore allow ourselves to let the symbol
[G] denote either AG G(F )\G(AF ) or G(F )\G(AF )1 . The group G(AF )1 is
unimodular by Lemma 3.5.4 below, and hence admits a Haar measure (the
notion of a Haar measure and a unimodular group will be recalled in §3.2).
The basic topological and measure theoretic properties of [G] are summa-
rized in the following theorem:
Theorem 2.6.2 (Borel, Conrad, Harder, Oesterlé) Assume that G is
smooth and connected. The quotient [G] defined in (2.12) has finite volume
with respect to the measure induced by a Haar measure on G(AF )1 . The adelic
quotient [G] is compact if and only if for every F -split torus T ≤ G one has
TF ≤ R(GF ).
Proof. The first assertion is [Con12a, Theorem 1.3.6]. See [Con12a, Theorem
A.5.5] for a proof of the last assertion. t
u
For comparison with the classical theory of automophic forms it is con-
venient to relate the adelic quotient to locally symmetric spaces (compare
Chapter 6). For this purpose, if K ∞ ≤ G(A∞ F ) is a compact open subgroup
let
h := h(K ∞ ) = |G(F )\G(A∞ ∞
F )/K |.
h
a
Γi (K ∞ )\G(F∞ ) −→ G(F )\G(AF )/K ∞ (2.14)
i=1
where
Γi (K ∞ ) := G(F ) ∩ ti · G(F∞ )K ∞ · t−1
i .
2.7 Reduction theory 47
Notice that the Γi (K ∞ ) are discrete subgroups of G(F ) and they are moreover
arithmetic in the following sense:
then
Γ0 (N )\GL2 (R) = GL2 (Q)\GL2 (AQ )/K0 (N ).
If we let K∞ = SO2 (R) then :
Γ0 (N )\(C − R)→GL
˜ 2 (Q)\GL2 (AQ )/K∞ K0 (N )
= GL2 (Q)\(C − R) × GL2 (A∞
Q )/K0 (N ),
Reduction theory allows us to control the adelic quotient [G] when it is non-
compact. To be more precise about this assume that G is a reductive group
over a global field F . In the number field case this assumption can be weak-
ened to allow for arbitrary connected affine groups of finite type over F at
the cost of introducing more notation [PR94, §4.3].
The following theorem provides what is known as the Iwasawa decom-
position:
Theorem 2.7.1 Let G be a reductive group over a global field F and let
P ≤ G be a parabolic subgroup. There exists a maximal compact subgroup
48 2 Adeles
Here | · | denotes the usual norm on R when F is a number field and the
canonical norm on Fp ((t)) when F is a function field of characteristic p.
The basic result of reduction theory is the following:
Theorem 2.7.2 (Reduction theory) There exists a Siegel set such that
Proof. See [Spr94]. To aid the reader we note that in loc. cit. the group G(F )
acts on the right. Thus one has to take an inverse to move G(F ) to the left,
and this changes the inequality in the analogue of AT0 (t) in loc. cit. to that
we used in defining AT0 (t). t
u
To understand what is going on in Theorem 2.7.2 it is useful to consider
what is arguably the simplest nontrivial case, that when G = GL2 . In this
case we can take K = O2 (R)GL2 (Z)b and let B = P be the Borel subgroup
of upper triangular matrices (since it is a Borel subgroup it is a minimal
parabolic subgroup). Then ∆ consists of the single root
α : T (R) −→ R×
t1 −1
t2 7−→ t1 t2 .
for small enough t ∈ R>0 . This is the content of Exercise 2.15. We remark
that it is this example that gives reduction theory its name. Indeed, a slight
strengthening of Theorem 2.7.2 in this case yields the familiar fact from the
2.7 Reduction theory 49
Exercises
2.5. Let O be a Dedekind domain with fraction field F and let Y be a flat
O-scheme. Suppose that X → Y := YF is a closed immersion. Let X be the
schematic closure of X in Y. Then for any closed immersion Z → Y whose
generic fiber is an isomorphism onto X there is a unique closed immersion
X → Z such that the following diagram commutes:
Z /Y
2.6. Let O be a Dedekind domain with fraction field F . Let G/O be a flat
group scheme of finite type with generic fiber G := GF . Let H be a group
scheme over F equipped with a morphism of group schemes H −→ G that is
a closed immersion. Show that the schematic closure H of H in G is a group
scheme.
50 2 Adeles
2.7. Let S be a finite set of places of a global field F . Prove that F is dense
in FS .
2.8. Let m > 1 be an integer. Let
A := Z[xij , tij , y]1≤i,j≤n / det(xij )y − 1, (xij ) − 1 + m(tij ) .
Here we view (xij ), (tij ) as matrices in order to make sense out of the poly-
nomial relations. Show that G := Spec(A) is a model of GLn over Z and
that
G(Z) := {g ∈ GLn (Z) : g ≡ 1 (mod mMn (Z))}.
2.9. Let X1 , X2 be affine schemes of finite type over a global field F . Show
that X1 and X2 admit weak (resp. strong) approximation with respect to a
finite set S of places F if and only if X1 × X2 does. Deduce that if G is an
algebraic group over a number field F then G is satisfies weak (resp. strong)
approximation with respect to a finite set S of places F if and only if a Levi
subgroup of G does.
2.10. Let F be a global field. Show that GLn admits weak approximation
with respect to any proper subset of the places of F .
2.11. Assume that the affine algebraic group G over the global field F satisfies
strong approximation with respect to a finite set S of places F . Show that
2.15. Let
Ω = {( 1 1t ) : t ∈ [0, 1]} .
Show that GL2 (AQ ) = GL2 (Q)ΩAGL2 AH (t) K for t sufficiently small.
2.16. Let G be a smooth algebraic group over a global field F with unipotent
radical N . Prove that under the map
2.17. For each integer N ≥ 1 let Γ (N ) ≤ GLn (Z) be the kernel of the
reduction map
GLn (Z) −→ GLn (Z/N ).
Prove that a subgroup of GLn (Q) is a congruence subgroup if and only if it
contains Γ (N ) for some N as a subgroup of finite index.
Chapter 3
Automorphic Representations
End(V )
be the space of all continuous linear maps from V to itself. We let GL(V ) ⊂
End(V ) be the group of invertible continuous linear maps with continuous
inverse.
53
54 3 Automorphic Representations
Often one omits explicit mention of the space of π. If we wish to specify that
π acts on a ?-space (e.g. Fréchet, Hilbert, locally convex) we often to π as a
?-representation.
A morphism of representations (or a G-intertwining map) is a morphism
of topological vector spaces that is G-equivariant; it is an isomorphism if
the underlying morphism of topological vector spaces is an isomorphism.
If two representations are isomorphic then they are also called equivalent.
Two unitary representations (π1 , V1 ) and (π2 , V2 ) are unitarily equivalent
if there is a G-equivariant isomorphism of Hilbert spaces V1 → V2 . Thus
unitary equivalence implies equivalence, but not conversely.
A subrepresentation of (π, V ) is a closed subspace W ≤ V that is pre-
served by π, a quotient of π is a topological vector space W equipped with
an action of G and a G-equivariant continuous linear surjection V → W , and
a subquotient is a subrepresentation of a quotient.
We will mostly be interested in unitary representations, and we need lan-
guage to describe when a representation that is not originally presented as
a unitary representation actually “is” unitary. First, we say that a repre-
sentation on a Hilbert space is unitarizable if it is equivalent to a unitary
representation. Sometimes the natural representations to study are not on
Hilbert spaces, or even complete spaces (see, for example, Chapter 5). How-
ever, one can still discuss unitarity in this situation as follows. Recall that a
pre-Hilbert space is a complex inner product space that admits a countable
orthonormal basis. Thus a Hilbert space is simply a pre-Hilbert space that
is complete with respect to the metric induced by the inner product. We can
still define U (V ) as before. A representation π of G on a pre-Hilbert space V
extends by continuity to the completion V of V (which is a Hilbert space).
With this in mind we say that a representation π on a pre-Hilbert space V
is pre-unitary if π(G) ≤ U(V ). Thus for every pre-unitary representation
of G we obtain, by continuity, a canonical unitary representation of G on V .
Finally, we say that a representation is pre-unitarizable if it is equivalent
3.1 Representations of locally compact groups 55
X × G −→ X (3.1)
(x, g) 7−→ xg
Proof. Let
Cc (X) ≤ L2 (X, dx)
be the subspace of compactly supported continuous functions. Let U ⊂ G
be an open neighborhood of the identity 1 ∈ G with compact closure, let
ϕ ∈ Cc (X), and let W = supp(ϕ)U ⊂ X. Then W is relatively compact, and
π(g)ϕ is supported in W if g ∈ U . Thus
Example 3.1. The points of Borel subgroups are, in general, not unimodular.
For example, if B ≤ GL2 is the Borel subgroup of upper triangular matrices
then for any local field F with normalized valuation | · | we can write
a 1t
B(F ) = : a, b ∈ F × , t ∈ F .
b 1
We note that one can alternately define δG (g) via the relation
d` (g −1 ) = δG (g)d` g (3.6)
Proof. The first assertion is [Kna86, Theorem 8.32] in the case where G is
a Lie group. The proof is the same in general. We also note that in the
notation of loc. cit. one has ∆G := δG , ∆S := δS , and ∆T := δT . For the
second equality we use (3.6). t
u
for f ∈ Cc (G). t
u
dg
The measure dµ in the theorem will be denoted by dh . We remark that in
loc. cit. the author considers right invariant measures on G/H instead, but
the result above follows in the same manner, or from the version with H\G
replaced by G/H.
π(f ) : V −→ V
Z
ϕ 7−→ π(g)f (g)ϕdr g.
G
If π is unitary then this map is bounded, and in fact in the unitary case it is
well-defined for all f ∈ L1 (G) (Exercise 3.16). The definition (3.7) is chosen
so that
π(f1 ∗ f2 ) = π(f1 ) ◦ π(f2 )
(see Exercise 3.17). Thus, in particular, representations π of G give rise to
algebra representations of Cc (G), that is, C-algebra homomorphisms
Cc (G) −→ End(V ).
3.4 Haar measures on local fields 59
Let F be a local field, that is, a field equipped with an absolute value that is
compact with respect to the topology induced by the absolute value (see §2.1).
All infinite locally compact fields arise as the completion of some global field
[Lor08, §25, Theorem 2] with respect to some absolute value. In this section
we define Haar measures on the additive group of F .
Assume first that F is archimedean, so F is isomorphic to R or C. The
standard Haar measure on R is the usual Lebesgue measure, and twice the
standard Haar measure on C. If z = x + iy with x, y ∈ R, then this Haar
measure is |dz ∧ dz̄| = 2dx ∧ dy where dx and dy are the Lebesgue measure
on R.
Let F be a nonarchimedean field with ring of integers OF and let $ be
a uniformizer. Since the additive group of F is a locally compact topologi-
cal group, it admits a Haar measure dx. The set OF ⊂ F , being compact,
has finite measure, so to specify the Haar measure it suffices to specify the
measure dx(OF ) of OF . We often assume dx(OF ) = 1, but sometimes other
normalizations are convenient.
Open sets in F of the form a + $k OF where a ∈ F and k ∈ Z>0 form a
neighborhood base. Using additivity and invariance of the Haar measure one
can compute the measure of such sets:
Lemma 3.4.1 If dx is a Haar measure on F then
dx(a + $k OF ) = q −k dx(OF ).
t
u
Thus for every local field F we have a Haar measure dx. Via the usual
procedure we then obtain a product measure on any F -vector space. It is
again an Haar measure with respect to addition.
Given a Haar measure dx on F we obtain a Haar measure
dx
dx× :=
|x|
Cc (G(Fv )) −→ C
Z
f 7−→ f (g)d|ω` |v (g).
G(Fv )
Since ω` is left G(Fv )-invariant, the measure d|ω` |v is left G(Fv )-invariant,
and hence is a left Haar measure. This is explained in more detail in the
archimedean case in [Kna86, Chapter VIII.2], and in the nonarchimedean
case in [Oes84, §2].
Proof. Let ZG be the center of G and Gder its derived group. We then have
a commutative diagram
3.5 Haar measures on the points of algebraic groups 61
a
ZG × Gder −−−−→ Gm
by yId
c
G −−−−→ Gm
where a and c are given on points by
g 7→ det(Ad(g) : g → g),
the left vertical map b is the product map, and the right vertical map is the
identity. The map a is trivial on ZG and it is trivial on Gder since X ∗ (Gder )
is trivial. Since b is a quotient map, c is trivial as well, and we deduce the
corollary. t
u
This provides us with a description of the left Haar measures dg` on G(Fv )
for every v. We similarly obtain right Haar measures dr (g) := d` (g −1 ). One
now wishes to obtain a right Haar measure on G(AF ). To do this let S be a
finite set of places of F including the archimedean places and let
Y
KS = Kv ≤ G(ASF )
v6∈S
Proof. Since the right Haar measure is constructed as the product of local
Haar measures one has
Y
δG(AF ) (g) = δG(Fv ) (gv ) = | det Ad(g) : g(AF ) → g(AF ) |AF . (3.10)
v
G(F )\G(AF )
(see [Fol95, Theorem 2.49]). This quotient only has finite volume if G(AF ) =
G(AF )1 . Similarly any Haar measure on G(AF )1 induces a right G(AF )-
invariant Radon measure on [G]. The latter quotient always has finite volume
(see Theorem 2.6.2).
The way we obtained the Haar measure on G(A QF ) is a little unnatural.
It would be more natural to just take the product v dωv directly. However,
this product is often not convergent. One can introduce a canonical family of
positive real numbers λv such that
Y
λv d|ω|v
v
is convergent, and in this way one obtains the so-called Tamagawa measure
on G(AF ). Upon choosing a suitable measure on AG we obtain a measure
on G(AF )1 and can thereby obtain the so-called Tamagawa measure on [G].
The measure of [G] with respect to the Tamagawa measure is called the
Tamagawa number of G. This number is intimately connected with the
arithmetic of G. We refer to the reader to [Kot88] and the references therein
for more details.
Now assume that G is an affine algebraic group over a global field F . Consider
the quotient
[G] := G(F )\G(AF )1
3.6 Automorphic representations in the L2 -sense 63
as in (2.12).
By Lemma 3.5.3 a right Haar measure on G(AF ) is left invariant under
G(F ) and with respect to it the space [G] has finite volume (compare Theorem
2.6.2). Thus we have a well-defined Hilbert space
L2 ([G])
with inner product defined as in (3.2) (but with X replaced by [G] and dx
replaced by the right Haar measure dr g).
Definition 3.3. An automorphic representation of G(AF )1 is an irre-
ducible unitary representation π of G(AF )1 that is equivalent to a subquo-
tient of L2 ([G]).
Here a subquotient is a subrepresentation of a quotient representation. The
reason we must consider subquotients, and not just subrepresentations, is
discussed in §3.7 below. We mention a variant of the definition that we will
also use. When G is reductive, an automorphic representation of AG \G(AF )
is an irreducible unitary representation that is equivalent to a subquotient
of L2 (G(F )AG \G(AF )). When G is reductive we can identify automorphic
representations of G(AF )1 and AG \G(AF ) via the commutative diagram
where the horizontal arrows are the action maps and the vertical arrows are
the isomorphisms induced by G(AF )1 →A ˜ G \G(AF ).
Note that contrary to the usual convention we have not assumed that G
is reductive. Even without this assumption the preceding definition makes
sense. It is only when one starts talking about admissible representations
that the assumption of reductivity is really necessary. A concrete example of
this is given right after [BS98, Proposition 3.2.11]. The nonreductive case has
not received as much attention as the reductive case, but see [Sli84].
We end this chapter with §3.7 through 3.10. Sections 3.7 through 3.9
discuss how one can decompose unitary representations of locally compact
groups in general, and §3.10 discusses why we have restricted our attention
to affine algebraic groups instead of certain natural larger classes of groups.
All of these sections can be omitted on a first reading.
Before delving into these sections we note that our definition of an auto-
morphic representation leaves much to be desired in that it is unclear how one
would go about studying these objects and how they are fit into arithmetic
and geometry. In the subsequent two chapters we will develop the basic al-
gebra of convolution operators on automorphic representations that are used
to study them. It is a refinement of the algebra L1 (G(AF )) discussed in §3.3
above. For F a number field and the decomposition
64 3 Automorphic Representations
G(F∞ ) × G(A∞
F )
naturally gives rise to a decomposition of L1 (G(AF )); the first factor is the
archimedean factor which will be discussed in Chapter 4 and the second factor
is the nonarchimedean factor which will be discussed in chapters 5 and 8.
This gives us tools to study automorphic representations, but it still does
not shed much light into how they appear in geometry and arithmetic. The
link with geometry will be made precise in Chapter 6. The primary link with
arithmetic is much more profound, subtle, and conjectural; it is the content
of the Langlands functoriality conjectures, explained in Chapter 12.
V = ⊕i Vi
This is called a direct integral decomposition. We will not make these no-
tions precise here; a nice introduction to these concepts (but without proofs)
is given in [Fol95, §7], for proofs one can consult [Dix77]. Two motivating
examples to keep in mind are the right regular representation of R on L2 (R)
and the representation L2 (Z\R) induced by it. Fourier analysis tells us that
Z ⊕
L2 (Z\R) = ⊕n∈Z Vn and L2 (R) = Vt dt. (3.13)
t∈R
3.8 The Fell topology 65
Vt −→ L2 (R).
L2 (R) −→ Vt
Z
f 7−→ f (x)e−2πitx dx
R
by Fourier theory.
For the converse statement, note that a linear R-intertwining map
Vt −→ L2 (R)
is either zero or has image contained in the C-span of x 7→ e2πitx . On the other
hand x 7→ e2πitx is not L2 (R), so we deduce that any linear R-intertwining
map Vt → L2 (R) is zero. t
u
procedure for defining this topology is very similar to the procedure for defin-
ing the Zariski topology as in §1.2. To define this topology, we recall from
§3.3 every representation (π, V ) of G gives rise to an algebra homomorphism
Cc (G) −→ End(V ) which extends to L1 (G) if π is unitary. We can extend
this homomorphism still further by completing. If we define
C ∗ (G) −→ End(V ).
This is the motivating example of a C∗-algebra, but we will not give the
(elementary) definition of a C∗ -algebra, referring instead to [Fol95, §1.1]. Now
given a unitary representation π of G, we obtain a closed two-sided ideal
By convention V (∅) = ∅. Then one checks that the V (S) just defined form the
closed sets of a topology on Prim(G), called the hull-kernel or Jacobson
topology. Now we have a natural map
b −→ Prim(G)
G
[π] 7−→ ker(π)
We have not (and will not) given the definition of a type I group. However,
we can give a characterization using the following theorem [Fol95, Theorem
7.6]:
Theorem 3.9.1 If G is a second countable locally compact group then the
following are equivalent:
(a) The group G is of type I.
(b) The Fell topology on G
b is T0 .
(c) The map G b −→ Prim(G) is injective. t
u
We point out that the groups of primary concern to us in this book are
fortunately of type I:
Theorem 3.9.2 If G is an algebraic group over a local field F then G(F ) is
of type I. t
u
Proof. For the archimedean case see [Dix77, §13.11.12]. For the nonar-
chimedean case see [Sli84]. t
u
One can find a proof of the following in [Clo07, Appendix]:
Theorem 3.9.3 If G is a reductive group over a global field F then G(AF )
is of type I. t
u
It should be pointed out that if G is not reductive then G(AF ) need not be
of type I [Moo65].
For π ∈ Gb let Vπ be the space of (a particular realization of) π. The basic
refinement of (3.12) is the following (see [Clo07, Theorem 3.3]):
Theorem 3.9.4 Let (ρ, V ) be a unitary representation of G. Then there ex-
ists a Borel measurable multiplicity function
b −→ {1, 2, . . . , ∞}
G
π 7−→ m(π)
We note that since the measure in (3.14) is supported in a countable set this
direct integral is really a direct sum.
Of course in practice it would be better to have an explicit description of
the decomposition of Theorem 3.9.4 for naturally occurring representations
V . Langlands’ profound and foundational work on Eisenstein series provide
such a decomposition in our primary case of interest, namely when V =
L2 ([G]). We will return to this point in §10.4 below.
Exercises
3.1. In the situation of the proof of Theorem 3.1.1 prove that if ϕ ∈ Cc (X)
then kπ(g)ϕ − ϕk∞ → 0 as g → 1.
3.2. Prove that compact (Hausdorff) topological groups are unimodular.
3.3. Show that for a locally compact topological group G the following are
equivalent:
dr (hg) = δ(h)dr g,
d` (gh) = δG (h)−1 d` g,
dr (g −1 ) = δG (g)−1 dr g,
d` (g −1 ) = δG (g)d` g.
dr g = δG (g)d` g.
G(AF ) = P (AF )K
holds (see Theorem 2.7.1). Show that the Haar measures on G(AF ), P (AF ),
and K can be normalized so that
dg = d` pdk.
3.6. Show that if Ω ⊂ G(AF ) is a compact set, then there is a finite set of
places S of F depending on Ω such that the projection of Ω to G(ASF ) is
contained in a maximal compact open subgroup K S ≤ G(ASF ). Deduce that
the products (3.9) are left and right Haar measures on G(AF ), respectively.
70 3 Automorphic Representations
1
dx(OF× ) = dx× (OF× )
1 − q −1
where q is the order of the residue field of OF .
3.8. Let F be a nonarchimedean local field and let dg be the Haar measure
on GL2 (F ) such that dg(GL2 (OF )) = 1. For n ≥ 0 compute
3.9. Suppose that G is a connected reductive group over a global field F and
that AG 6= 1. Prove that G(F )\G(AF ) has infinite volume with respect to
the measure induced by a Haar measure on G(AF ).
3.10. Suppose that G is a connected reductive group over a global field F .
Show that [G] has finite volume using reduction theory (Theorem 2.7.2).
3.11. Prove that
∧i,j dxij
d (xij ) =
| det xij |
is a (right and left) Haar measure on
if R is either a local field or the adeles of a global field and | · | is the norm
on the local field or the adelic norm, respectively.
3.12. Let F be a global field and let R be an F -algebra. For (xij ) ∈ GL2 (R)
let
dx11 ∧ dx21 ∧ dx12 ∧ dx22
ω=
x11 x22 − x21 x12
viewed as a left-invariant top-dimensional differential form on G. For a nonar-
chimedean place v of F , compute
Prove that Y
d|ω|v (GL2 (OFv ))
v
v
ζF (1)
does converge.
3.10 Why affine groups? 71
is not completely reducible; that is, it does not decompose into a direct sum
of irreducible subrepresentations.
3.15. Let G be a locally compact group. Prove that the convolution product
on Cc (G) is associative.
3.17. Let G be a locally compact group and let f, g ∈ Cc (G). Prove that for
any representation π of G one has π(f ∗ g) = π(f ) ∗ π(g).
Chapter 4
Archimedean Representation Theory
For the moment let F be a number field and let G be an affine algebraic
group over F . Consider a Hilbert space representation V of
Y
G(F∞ ) = G(Fv ).
v|∞
73
74 4 Archimedean Representation Theory
exp : g −→ G(F ).
In the case where G = GLn , the Lie algebra gln is the collection of n × n
matrices. The exponential is simply the matrix exponential in this case:
∞
X Xj
exp(X) := .
j=0
j!
d
π(X)ϕ = π(exp(tX))ϕ|t=0
dt
π(exp(t + h)X)ϕ − π(exp(tX))ϕ
= lim |t=0
h→0 h
if the limit exists. We will sometimes simply write Xϕ for π(X)ϕ. We say
that a vector ϕ ∈ V is C 1 if for all X ∈ g, the derivative Xϕ is defined. We
define C k inductively by stipulating that ϕ ∈ V is C k if ϕ is C k−1 and Xϕ
is C k−1 for all X ∈ g. A vector ϕ ∈ V is C ∞ if it is C k for all k ≥ 1.
(see [God15, §25-26] for more details, especially [God15, Theorem 39]). For
this we compute:
d d
(I(ϕ))(g exp(tX)) = hπ(g)π(exp(tX))ϕ, ϕ0 i
dt t=0 dt t=0
76 4 Archimedean Representation Theory
= hπ(g)Xϕ, ϕ0 i
= (I ◦ X)(ϕ)(g).
Since we are assuming that ϕ is smooth, we see that this function is smooth
as well.
In order to verify that Vsm is a representation of g, we must check that
for all X, Y ∈ g and all ϕ ∈ Vsm . By duality, it suffices to prove that this
identity holds after pairing with ϕ0 for all ϕ0 ∈ V , in other words, that
d −1
= π(exp(tX))π(f δG(F )ϕ
dt )
t=0
−1
= Xπ(f δG(F ) )ϕ.
G(F ) × V −→ V
(g, ϕ) 7−→ π(g)ϕ
is continuous. This implies that for all ε > 0 there exists a neighborhood
U ⊆ G(F ) of the identity such that |π(g)ϕ − ϕ| < ε for all g ∈ U . Choose a
nonnegative function f ∈ Cc∞ (G(F )) supported in U such that
Z
f (g)dr g = 1.
G(F )
Then
Z
kπ(f )ϕ − ϕk2 =
f (g)(π(g)ϕ − ϕ)dr g
G(F )
2
Z
≤ f (g)kπ(g)ϕ − ϕk2 dr g ≤ ε
G(F )
The representation theory of compact groups is much simpler than the rep-
resentation theory of noncompact groups. For example, any irreducible rep-
resentation of a compact group is unitarizable and finite-dimensional (see
Theorem 4.3.3 below). A profitable strategy in representation theory is to
78 4 Archimedean Representation Theory
Proof. Let h , i denote the original Hilbert space pairing and k·k2 the original
norm. Define Z
(ϕ1 , ϕ2 ) = hπ(k)ϕ1 , π(k)ϕ2 idk.
K
By construction it is K-invariant so we need only check the claim about the
topology. The maps K → C given by
satisfies
C −2 meas(K)kϕk22 < kϕk22,new < C 2 meas(K)kϕk22 .
This implies the result. t
u
We now prepare to state the Peter-Weyl theorem, which says that all of
the representation theory of a compact Lie group K is contained in L2 (K).
m : G −→ C
4.3 Restriction to compact subgroups 79
g 7−→ (π(g)ϕ1 , ϕ2 )
for some ϕ1 , ϕ2 ∈ V .
The following proposition implies that irreducible representations of K can
be recovered from their matrix coefficients:
Proposition 4.3.2 Suppose K is a compact Lie group and (π1 , V1 ) and
(π2 , V2 ) are two representations of K with π2 unitary. If there exist matrix
coefficients m1 , m2 for π1 , π2 respectively that are not orthogonal in L2 (K)
then there exists a non-trivial intertwining operator I : V1 → V2 .
Proof. Write ( , )i for the Hermitian pairing on Vi . Let x1 , y1 ∈ V1 and
x2 , y2 ∈ V2 be such that
Z
(π1 (k)x1 , y1 )1 (π2 (k)x2 , y2 )2 dk 6= 0.
K
Let Z
I(ϕ) = (π1 (k)ϕ, y1 )1 π2 (k −1 )y2 dk.
K
We claim that I gives an intertwining map
I : V1 −→ V2 .
Indeed, Z
π2 (g) ◦ I(ϕ) = (π1 (k)ϕ, y1 )1 π2 (gk −1 )y2 dk.
K
Thus I is nonzero. t
u
The following theorem tells us that all of the representation theory of a
compact Lie group K is contained in L2 (K):
Theorem 4.3.3 (Peter-Weyl Theorem) Let K be a compact Lie group.
(a) The matrix coefficients of finite dimensional unitary representations of K
are dense in C(K) and Lp (K) for all 1 ≤ p ≤ ∞.
80 4 Archimedean Representation Theory
Let G be an affine algebraic group G over the archimedean field F and let
K ≤ G(F ) be a maximal compact subgroup. Our goals in this section are
to introduce the notion of admissibility, define (g, K)-modules, and attach
a (g, K)-module to each admissible representation. With these concepts in
hand, at the end of the section we will describe the content of the remainder
of the chapter.
Let (π, V ) be a representation of K. For each equivalence class of irre-
ducible representation σ of K we write
V (σ) = {ϕ ∈ V : hπ(k)ϕ : k ∈ Ki ∼
= σ}
where the brackets denote the C-span. This is the σ-isotypic subspace. A
vector in V (resp. a subspace) of V is said to have K-type σ if it is an
element (resp. a subspace) of V (σ).
Definition 4.3. A vector ϕ ∈ V is K-finite if ϕ ∈ V (σ) for some irreducible
representation σ of K. The algebraic direct sum
Vfin := ⊕σ∈K
cV (σ)
d 1
π(X)ϕ = Xϕ = π(exp(tX))ϕ = lim (π(exp(hX))ϕ − ϕ).
dt t=0 h→0 h
g = k(k −1 g) ∈ W U1 ⊂ U.
Therefore
Z Z Z
|f (g)|dr g ≤ |f0 (k)||f1 (k −1 g)|dkdr g
G(F )−U G(F )−U K
Z Z
= |f0 (k)||f1 (k −1 g)|dkdr g
G(F )−U K−W
4.4 (g, K)-modules 83
Z Z
≤ |f0 (k)| |f1 (k −1 g)|dr gdk
K−W G(F )
Z
= f0 (k)dk < ε.
K−W
Thus f satisfies (4.1). Here we have used the fact that δG(F ) (k) = 1 because
K is compact.
We now show that π(f )ϕ is K-finite. Let ρ be a finite dimensional unitary
representation of which f0 is a matrix coefficient. Thus f0 (k) = (ρ(k)ξ, ζ) for
some ξ, ζ in the space of ρ. Then if k1 ∈ K:
Z
−1
f (k1 g) = f0 (k)f1 (k −1 k1−1 g)dk
K
Z
= (ρ(k)ξ, ζ)f1 (k −1 k1−1 g)dk
ZK
= (ρ(k1−1 )ρ(k)ξ, ζ)f1 (k −1 g)dk
K
Z
= (ρ(k)ξ, ρ(k1 )ζ)f1 (k −1 g)dk.
K
−1
Therefore the linear span of the
R functions f (k−11 g) is contained in the linear
span of the functions g 7→ K (ρ(k)ξ, ζ)f1 (k g)dk for varying ξ, ζ in the
space of ρ, and this space isR finite dimensional. Thus
R the space spanned by
the vectors π(k1 )π(f )ϕ = G(F ) f (g)π(k1 g)ϕdg = G(F ) f (k1−1 g)π(g)ϕdg as
k1 varies over K is finite dimensional, so π(f )ϕ ∈ Vfin . Moreover, π(f )ϕ is
smooth for any vector ϕ by Proposition 4.2.3. It follows that V0 is dense in
Vsm , which is dense in V .
Next we prove that Vfin ≤ Vsm if V is admissible. First observe that V0
is K-invariant since Vsm is by Lemma 4.4.3. Let σ be an irreducible unitary
representation of K. Then V0 (σ) ≤ V (σ). Since Vfin is an algebraic direct
sum of the V (σ) it suffices to show that V0 (σ) = V (σ).
Since V (σ) is finite-dimensional by admissibility, V0 (σ) admits a well-
defined orthogonal complement in V (σ) (this is the only part of the proof
where admissibility is used). If ϕ is in this orthogonal complement then ϕ
is orthogonal to all of V0 , because it is orthogonal to V (τ ) for every τ 6= σ.
Therefore ϕ = 0, since V0 is dense. This establishes that V0 (σ) = V (σ), and
hence V0 = Vfin ≤ Vsm .
Finally, must show that Vfin is invariant under g. Let ϕ ∈ Vfin , let W be
the span of ϕ under K, and let
W1 := hY ϕ | Y ∈ g and ϕ ∈ W i,
π(eσ ) : V −→ V (σ)
(4.3)
Z
ϕ 7−→ eσ (k)π(k)ϕdk
K
One also can define the left and right convolution of eσ with any element
f ∈ Cc∞ (G(F )):
Z
f ∗ eσ (g) := f (gk −1 )eσ (k)dk (4.5)
K
Z
eσ ∗ f := eσ (k −1 )f (kg)dk. (4.6)
K
and
V (Ξ) := ⊕σ∈Ξ V (σ).
(algebraic direct sum).
Definition 4.5.1 The Hecke algebra of K-type Ξ is
V (Ξ) = V1 ⊕ V2
where we have used Theorem 4.5.2 for the first equality. It is clear that
eΞ ∗ Cc∞ (G(F ))V1 is a subspace of V (Ξ). If it is a proper subspace then the
inverse image of eΞ ∗ Cc∞ (G(F ))V1 under
eΞ : V 7→ V (Ξ)
gC := g ⊗R C.
gC / U (g)
!
A
Remark 4.2. This is our first time explicitly mentioning ρ, the ubiquitous
half-sum of positive roots. It is incorporated into the definition to make it
compatible with parabolic induction (see §4.9).
Let W := W (G, T )(C) denote the Weyl group of T in G; it acts on tC and
hence U (t). For the proof of the following theorem of Harish-Chandra, see
[Wal88, Theorem 3.2.3]:
Spec(C[(tC )∨ ], C)−→Hom(U
˜ (t), C).
4.7 Classification of (g, K)-modules for GL2R 89
The set on the left can be identified with the maximal ideals of (tC )∨ , which,
by the Hilbert Nullstellensatz, are in natural bijection with elements of (tC )∨ .
Taking W invariants we obtain the proposition. t
u
A g-module V is said to admit an infinitesimal character if Z(g) acts via
a character on V . By Theorem 4.6.1 and Proposition 4.6.2 such a character
can be identified with an element of (tC )∨ .
The following is an immediate corollary of Lemma 4.1:
Corollary 4.6.3 Let V be an irreducible admissible (g, K)-module and let
Λ : V → V be a C-linear map commuting with the actions of g and K. Then
Λ is multiplication by a scalar. t
u
It follows immediately from the corollary that any irreducible (g, K)-module
admits an infinitesimal character. This is a very useful invariant that is sur-
prisingly closely tied to arithmetic.
∆ = (1/4)(H 2 + 2XY + 2Y X)
χ : Z(g) −→ C
Z 7−→ µ
∆ 7−→ s(1 − s).
(c) Y v` = s − 2` v`−2 ,
Let G be a reductive group over the local field F . For the moment we do not
assume that F is archimedean. Let (π, V ) be a representation of G(F ). In
Definition 4.2 above we defined notion of a matrix coefficient. In this section
we explain how matrix coefficients can be used to isolate important classes
of representations of G(F ).
We have already seen in §2.6 that the center of G often causes analytic
problems for somewhat trivial reasons. One often circumvents this difficulty
4.8 Matrix coefficients 91
converges absolutely (this is obvious in the nonarchimedean case, see §8.5, and
in the archimedean case it is proven in [Kna02, Theorem 10.2]). Technically
speaking we have not defined admissibility in the nonarchimedean case; it is
defined in §5.3. The linear functional
Θπ : Cc∞ (G(F )) −→ C
There is also a version of the Plancherel theorem for G(F ) relating spaces of
functions on G(F ) and tempered representations of G(F ). It takes the form
Z
f (g) = Θπ (g)dµPl (π)
\)
G(F
where µPl (π) is the Plancherel measure on the unitary dual. Though this
is called the Plancherel theorem, it is really more like a Fourier inversion
theorem. In the archimedean case this is [Wal92, Theorem 13.4.1]. In the
nonarchimedean case this is [Wal03].
Finally the condition of temperedness is also a representation theoretic
encapsulation of how well-behaved automorphic L-functions are. This will be
discussed in the context of the Ramanujan conjecture in Conjecture 10.6.4.
P = MN
HM : M (F ) −→ aM (4.12)
defined by
hHM (g), χi = log |χ(g)|
∗
for χ ∈ X (M )F and we can identify M (F )1 := ker HM ∩ M (F ). We now
extend HM to HP by extending the characters in X ∗ (M )F to P by declaring
them to be trivial on N . We then obtain HP : P (F ) −→ aM ; this provides
an isomorphism X ∗ (M )F →X
˜ ∗ (P )F .
∗
For each λ ∈ aM C := Hom(aM , C) we therefore obtain a character
P (F ) −→ C×
p 7−→ ehHP (p),λi . (4.13)
Now consider the set of roots of a∗M can be identified with the vector space
underlying the associated root system Φ(G, T ). We let ρ denote half the sum
of the positive roots defined by the parabolic subgroup P (see §1.9). For each
representation (σ, V ) of M (F )1 one can form the representation (σ, λ) of P (F )
by extending σ trivially to P (F ) and twisting by the character hHP (·), λi.
We can then form the (normalized) induced representation
I(σ, λ)
Exercises
4.8. Prove that if g := gl2 then Z(g) = hH, ∆i in the notation of (4.7).
4.9. Prove that the space Vs,µ,ε of §4.7 is indeed a (g, K)-module with the
given action.
4.10. In the notation of §4.7 prove that if πk1 ∼ = πk2 then k1 = k2 , and that
if
πs1 ,µ1 ,ε1 ∼
= πs2 ,µ2 ,ε2
then µ1 = µ2 , ε1 = ε2 , and either s1 = s2 or s1 = 1 − s2 .
4.11. Prove that all finite dimensional (gl2 , O2 (R))-modules are quotients
(resp. subrepresentations) of some Vs,µ,ε for some s, µ, ε, and identify the
Vs,µ,ε of which they are quotients (resp. subrepresentations).
Chapter 5
Representations of Totally Disconnected
Groups
Here and the rest of this chapter, the letters td stand for totally disconnected.
Our basic examples are given by the next lemma, which follows readily
from the definition of the topology on the points of an algebraic group from
§2.2:
Lemma 5.1.1 Let G be an algebraic group over a local nonarchimedean field
F . Then G(F ) is of td-type. Similarly, if F is a global field and S is a finite
set of places of F including the archimedean places if F is a number field,
then G(ASF ) is a td group. t
u
97
98 5 Representations of Totally Disconnected Groups
denote the subalgebra of functions that are right and left K-invariant.
Lemma 5.2.1 Any element f ∈ Cc∞ (G) is in Cc∞ (G // K) for some compact
open subgroup K. If f ∈ Cc∞ (G // K), then f is a finite C-linear combination
of elements of the form
1KγK
for γ ∈ G. Here 1X denotes the characteristic function of a set X.
Proof. By local constancy of f for every γ in the support supp(f ) we can
chose a compact open subgroup K(γ) ≤ G so that f is constant on γK(γ).
Then [
γK(γ)
γ
Sn
is an open cover of supp(f ), so it admits
Tn a finite subcover i=1 γi K(γi ) since
f has compact support. Let Kr := i=1 K(γi ); it is a finite intersection of
compact open subgroups so it is itself a compact open subgroup. We see
that f is right Kr -invariant. Similarly we can find a compact open subgroup
K` ≤ G such that f is left K` -invariant, and letting K := K` ∩ Kr we see
that f ∈ Cc∞ (G // K).
Assume f ∈ Cc∞ (G // K). The union
[
KγK
γ∈supp(f )
1K ∞ γK ∞ = 1KS γS KS ⊗ 1K S
for some finite set of nonarchimedean places S. This reduces the study of
nonarchimedean Hecke algebras to the study of the local Hecke algebras
Cc∞ (G(Fv ))
where the superscript denotes the subspace of fixed vectors and where the
union is over all compact open subgroups K ≤ G.
In this definition we do not assume that the representation V is continuous,
or for that matter even give a topology on V . In fact, V is smooth if and only
if the representation is continuous if with respect to the given td topology on
G and the discrete topology on V . We note that the smooth representations of
5.3 Smooth and admissible representations 101
a1 m1 + · · · + an mn
V K = V1K ⊕ V2K
V K = V1 ⊕ V2
Thus V2 = 0. t
u
where the union is over all compact open subgroups K ≤ G. Then Vsm ≤
V is evidently a smooth subrepresentation of G. We say that the Hilbert
representation V is admissible if Vsm is admissible. We note that in this
case Vsm is a pre-Hilbert representation.
Lemma 5.3.3 Let V be a Hilbert representation of G. The subspace Vsm ≤ V
is dense and G-invariant. Moreover, if V and W are Hilbert representations
and Vsm → Wsm is a G-intertwining morphism then it extends uniquely to a
G-intertwining continuous morphism V → W . If the morphism Vsm → Wsm
is an isomorphism, then so is V → W .
5.4 Contragredients 103
We leave the proof as an exercise (see Exercise 5.11). Thus, by the lemma,
we can detect isomorphisms of Hilbert representations using the underlying
representations in the smooth category. We also have the following analogue
of Theorem 4.4.4:
5.4 Contragredients
V −→ (V ∨ )∨ and V −→ (V ∗ )∗
ι : V −→ V ∗
Proof. The last claim follows from Schur’s lemma (in the form of Exercise 5.5)
and the first claim. Suppose that we have a G-invariant, definite, Hermitian
inner product (·, ·) on V (C-linear in the first variable and C-antilinear in the
second). We then obtain a G-equivariant morphism
V −→ V ∗
ϕ0 7−→ (ϕ 7−→ (ϕ, ϕ0 )). (5.3)
t
u
Here Cc∞ (G(F ) // K) is the subalgebra of functions invariant on the left and
right under K.
The following is arguably the most important fact about this algebra:
Theorem 5.5.1 The spherical Hecke algebra Cc∞ (G(F ) // K) is commuta-
tive.
One way to prove this is via the Satake isomorphism (see Theorem 7.2.1).
In special cases this can also be proven using a trick due to Gelfand (see
Exercise 5.14). Let us describe the algebra in more detail in a special case:
Let
λ = (λ1 , . . . , λn ).
Let T be the maximal torus of diagonal matrices in GLn . Then λ defines a
cocharacter λ : Gm → T given on points by
λ
!
x 1
λ(x) = .. .
.
x λn
Consider
106 5 Representations of Totally Disconnected Groups
{Wv : v ∈ Ξ}
Ξ0 ⊆ S ⊆ Ξ
of finite cardinality set WS := v∈S Wv . If S ⊆ S 0 there is a map
Q
WS −→ WS 0 (5.4)
⊗v∈S ϕv 7−→ ⊗v∈S ϕv ⊗ (⊗v∈S 0 −S ϕ0v ) .
W := ⊗0 Wv := lim WS
−→
S
where the transition maps are given by (5.4). This is the restricted tensor
product of the Wv with respect to the ϕ0v . Thus W is the set of sequences
(ϕv )v∈Ξ ∈ ⊗v Wv
such that wv = ϕ0v for all but finitely many v ∈ Ξ. We note that if we are
given for each v ∈ Ξ a C-linear map
Bv : Wv −→ Wv
such that Bv (w0v ) = w0v for all but finitely many v ∈ Ξ then this gives a
map
B = ⊗v Bv : W −→ W
⊗ϕv 7−→ ⊗Bv (ϕv ).
AS −→ AS 0 (5.5)
⊗v∈S av 7−→ ⊗v∈S av ⊗ (⊗v∈S 0 −S a0v ) .
where the transition maps are given by (5.5). This is the restricted tensor
product of the Av with respect to the a0v . Finally, if Wv is an Av -module
for all v ∈ Ξ such that a0v ϕ0v = ϕ0v for almost all v, then ⊗0v Wv is an
A-module. The isomorphism class of W as an A-module in general depends
on the choice of {ϕ0v }. However, if we replace the ϕ0v by nonzero scalar
multiples we obtain isomorphic A-modules.
An easy example of this construction is the ring of polynomials in infinitely
many variables. It can be given the structure of a restricted tensor product
of algebras
108 5 Representations of Totally Disconnected Groups
Cc∞ (G(A∞ ∼ 0 ∞
F )) = ⊗v-∞ Cc (G(Fv ))
W ∼
= ⊗0v Wv (5.7)
But
V ∼
= V1 ⊗ V2
as Cc∞ (G) ∼
= Cc∞ (G1 × G2 )-modules, where
The Vi are evidently admissible, and they are irreducible by Proposition 5.3.2.
t
u
We now prove Flath’s theorem:
Proof of Theorem 5.7.1: For each finite set S of places of F containing the
infinite places and the set of places v where GFv is ramified let
110 5 Representations of Totally Disconnected Groups
Y
K S := Kv ≤ G(ASF )
v6inS
Cc∞ (G(A∞ ∼ 0 ∞
F )) = ⊗v-∞ Cc (G(Fv )) (5.8)
where the restricted direct product is constructed using the idempotents eKv
for Kv ≤ G(Fv ) with v 6∈ S. Use (5.8) to identify these two algebras for the
remainder of the proof. We then have a well-defined subalgebra
/ eKv is meas(K S ) 1K S .
1
where eK S = ⊗v∈S
By Corollary 5.5.2 and Theorem 5.7.2, as a representation of AS we have
an isomorphism
S
WK ∼ = ⊗v∈S−∞ Wv ⊗ W S
where W S is a one-dimensional C-vector space on which eK S acts trivially.
Hence, by admissibility,
W KS ∼
[
W = = lim ⊗ W ⊗ WS (5.9)
−→ v∈S−∞ v
S S
with respect to the obvious transition maps (compare §5.6). On the other
hand, (5.8) induces an identification
[
Cc∞ (G(A∞ F )) = AS = lim AS (5.10)
−→
S S
where the direct limit is taken with respect to the obvious transition maps
(again, compare §5.6), and it is clear that (5.9) is equivariant with respect to
(5.10). 2
Exercises
5.1. Prove that a group of td-type is Hausdorff, locally compact, and totally
disconnected.
5.2. Define [
Vsm := VK
K
5.7 Flath’s theorem 111
where the union is over all compact open subgroups K ≤ G. Prove that Vsm
is preserved by G and is a smooth subrepresentation of G. The subspace
Vsm ≤ V is the space of smooth vectors in V (in the current setting of td
groups).
5.4. Suppose that V is admissible and preunitary, that is, there is a non-
degenerate inner product on V that is preserved by G. Prove that every
G-invariant subspace V1 ≤ V admits a complement, that is, a G-invariant
subspace V2 ≤ V such that V1 ⊕ V2 = V .
5.6. Assume that G is second countable and V is smooth and irreducible. Let
ϕ1 , . . . , ϕn , ϕ01 , . . . , ϕ0n be two sets of elements of V with ϕ1 , . . . , ϕn linearly
independent. Then there is an f ∈ Cc∞ (G(F )) such that
for all 1 ≤ i ≤ n.
5.7. Assume that G is second countable and V is smooth and irreducible. Let
ZG be the center of G. Show that there is a quasi-character ωπ : ZG → C×
such that π(z) acts via ωπ (z) on V for all z ∈ ZG . The quasi-character ωπ
is called the central quasi-character of π. Show that if χ : G → C× is a
quasi-character, then the central quasi-character of π ⊗ χ is χ|ZG ωπ .
Hom(V K , C) = (V ∨ )K
Show, on the other hand, that f † = f for f ∈ Cc∞ (GLn (F ) // GLn (OF )).
Conclude that Cc∞ (GLn (F ) // GLn (OF )) is commutative.
5.15. Let F be a global field. For all v outside a finite set of places including
the infinite places choose hyperspecial subgroups Kv ≤ G(Fv ). Show that
0
∼
Y
G(A∞
F )= G(Fv )
v-∞
where the restricted tensor product is taken with respect to the Kv . Note that
we allow for infinitely many of the Kv ’s (perhaps all of them) to be different
from those obtained by choosing a model of G as in Proposition 2.3.1.
5.16. Let F be a global field. For all v outside a finite set of places including
the infinite places choose hyperspecial subgroups K1v , K2v ≤ G(Fv ). For
i = 1, 2, let Ai be ⊗0v-∞ Cc∞ (G(Fv )) where the restricted direct product is
taken with respect to eKiv (see (5.6)). Construct an equivalence of categories
between admissible A1 -modules and admissible A2 -modules.
Chapter 6
Automorphic Forms
113
114 6 Automorphic Forms
AG \G(AF ) ∼
= AG \G(F∞ ) × G(A∞
F )
Then the compactly supported smooth functions are simply Cc∞ (G(AF )) =
Cc∞ (G(F∞ )) ⊗ Cc∞ (G(A∞F )).
By definition, a smooth function on G(AF )1 is the restriction of a smooth
function on G(AF )
of (2.14). By Theorem 2.6.1 there are finitely many indices in this disjoint
union. The important point for the current discussion is not the precise defi-
nition of the Γi (K ∞ ), but just that they are discrete arithmetic subgroups of
G(F∞ ) such that the quotient Γi (K ∞ )\G(F∞ ) has finite volume (see Theo-
rem 2.6.2). Clearly automorphic representations are related to functions on
the left side. In this section and the following we make this relationship pre-
cise. We work only at infinity in the current section and then pass to the
adelic setting in the next.
Let
ι0 : G −→ GLn
ϕ : G(F∞ ) −→ C
|ϕ(g)| ≤ cX kgkr .
It turns out that different choices of ι lead to norms that are equivalent in
a sense made precise by Exercise 6.1. In particular, the notion of moderate
growth and uniform moderate growth is independent of the choice of ι.
For the definition of automorphic forms we will require the notion of K∞ -
finite functions and Z(g)-finite functions. The notion of K∞ -finiteness is given
in Definition 4.3. The definition of Z(g)-finite is similar.
116 6 Automorphic Forms
where A(Γ, J, σ) := A(Γ, J)(σ) in the notation of §4.4. We record the follow-
ing important result of Harish-Chandra [HC68]:
Theorem 6.2.1 (Harish-Chandra) For each ideal J ≤ Z(g) of finite codi-
mension the space A(Γ, J) is an admissible (g, K∞ )-module. In particular,
A(Γ, J, σ) is finite-dimensional for each σ. t
u
One might ask about the relationship between A(Γ ) and L2 (Γ AG \G(F∞ )).
There is no obvious relationship between them except under more stringent
assumptions. For example, Eisenstein series (which will be discussed in Chap-
ter 10) provide examples of elements of A(Γ ) need not even be square inte-
grable. It turns out to be more convenient to explain the relationship between
the so-called cuspidal subspaces of A(G) and L2 (Γ AG \G(F∞ )). We will defer
the discussion of this connection to the following section, in which we discuss
automorphic forms adelically.
ϕ : G(AF ) −→ C
|Xϕ(g)| ≤ cX kgkr .
|Xϕ(g)| ≤ cX α(s)r
t
u
Here we are using notation from our discussion of Siegel sets and reduction
theory in §2.7. For a proof of this lemma we refer to [Bor07, §5.4].
By definition, a (g, K∞ ) × G(A∞F )-module (π, V ) is a complex vector space
V that is both a (g, K∞ ) and G(A∞ F )-module (we denote the two action
maps by π) such that the two actions commute. It is admissible if for each
K ∞ ≤ G(A∞ F ) the space of fixed vectors V
K∞
is admissible as a (g, K∞ )-
∞
module. A morphism of (g, K∞ )×G(AF ) is a complex linear map commuting
with the actions of (g, K∞ ) and G(A∞ F ) (no continuity condition is assumed.
We define an irreducible (g, K∞ ) × G(A∞ F )-module to be one that admits no
invariant submodule. There is no topology on V , so we do not assume that
the submodule is closed.
Let Kmax ≤ G(AF ) be a maximal compact subgroup; thus Kmax =
∞ ∞
K∞ × Kmax where K∞ ≤ G(F∞ ) and Kmax ≤ G(A∞ F ) are maximal compact
subgroups. As before, we say that a function ϕ : G(AF ) → C is K-finite if
the span of translates
{x 7→ ϕ(xk) : k ∈ Kmax }
118 6 Automorphic Forms
ϕ : G(AF ) −→ C
(g, K∞ ) × G(A∞
F ) × A −→ A
(X, k, g, ϕ) 7−→ (x 7→ Xϕ(xkg)) . (6.7)
It is important to note that the action of G(A∞ F ) does not extend to an action
of G(AF ). Indeed, let ϕ ∈ A and let g ∈ G(F∞ ). Then x 7→ ϕ(xg) is smooth,
but it is only finite under gK∞ g −1 , not K∞ .
If σ is an irreducible representation of K∞ we denote by A(σ) the subspace
of vectors with K∞ -type σ, and if J ≤ Z(g) is an ideal the subspace of vectors
annihilated by J is denoted by A(J) (resp. A(J, σ)). The action of G(A∞ F )
preserves these spaces, and the action of (g, K∞ ) preserves A(J) (but of
course not A(J, σ) in general).
The following is the adelic analogue of Theorem 6.2.1:
Theorem 6.1. Let J ≤ Z(g) be an ideal of finite codimension. Then A(J)
is an admissible (g, K∞ ) × G(A∞
F )-module.
∞
Proof. We are to show that A(J, σ)K is finite dimensional for each K∞ -type
σ and each compact open subgroup K ∞ ≤ G(A∞ F ). We start by recalling the
description of the adelic quotient [G]/K ∞ given by (2.14). Letting t1 , . . . , th
denote a set of representatives for the finite set G(F )\G(A∞ F )/K
∞
one has a
homeomorphism
h
a
Γi (K ∞ )\G(F∞ ) −→ G(F )\G(AF )/K ∞
i=1
which intertwines the natural action of (g, K∞ ) on both sides. We can now
deduce the theorem by applying Theorem 4.4.1. t
u
ϕ : G(F )\G(AF ) −→ C
{x 7→ ϕ(xg) : g ∈ G(AF )}
is the closure of the space of continuous functions ϕ such that for all proper
parabolic subgroups P < G with unipotent radical N one has
Z
ϕ(ng)dn = 0.
[N ]
where L2cusp (π) is the π-isotypic subspace of L2cusp ([G]) and the sum is over all
isomorphism classes of cuspidal automorphic representations π of AG \G(AF ).
In view of this fact every irreducible subquotient of L2cusp ([G]) is in fact a
subrepresentation.
One justification for our claim that Acusp is the most important subspace
of A is that any element of A can be obtained from elements of Acusp ([M ]) as
M runs over the Levi subgroups of parabolic subgroups of G. We will discuss
the L2 -version of this fact in §10.4 below.
Though the relationship between automorphic representations in the sense
of Definition 6.6 and automorphic representations in the L2 -sense is compli-
cated, they are very closely linked in the cuspidal case. To explain the rela-
tionship, first note that if ϕ is in the space of a cuspidal automorphic repre-
sentation π of G(AF ), then by Schur’s lemma (Lemma 4.1) the subgroup AG
acts via a scalar. Let us refine this a little. Let aG := Hom(X ∗ (G)F , R) and
let
HG : G(F ) −→ aG (6.9)
be defined via hHG (g), λi = log |λ(g)| for λ ∈ X ∗ (G)F (compare §4.9). There
is a unique λ ∈ a∗GC such that the representation
6.5 The cuspidal subspace 121
π ⊗ ehHG (·),λi
AG .
In view of Theorem 6.5.2, for many purposes we can work with either
the definition of automorphic representation in terms of admissible repre-
sentations presented in this chapter, or automorphic representations in the
L2 -sense, whichever is convenient for the situation at hand. Both are often
convenient for different reasons. The L2 definition is convenient for abstract
harmonic analysis, but the definition in terms of admissible representations
has the advantage that one can immediately reduce questions to finite di-
mensional settings, either by passing to K∞ -types or vectors fixed under
a compact open subgroup. We will also explain in the next few sections
how spaces of automorphic forms, though not representations of G(AF ), are
(g, K∞ ) × G(A∞ F )-modules and hence give rise to automorphic representa-
tions.
We also warn the reader that later in this book we will often not be specific
about whether we are viewing an automorphic representation in the L2 sense
or in the sense of an admissible representation of G(AF ) (in the function
field case) or (g, K∞ ) × G(A∞
F ) (in the number field case). Indeed, in practice
one often has to switch back and forth between the two perspectives. The
particular point of view must be deduced from the context.
122 6 Automorphic Forms
For readers familiar with modular forms we make the connection between
modular forms and automorphic forms precise in this section.
Let Γ ≤ GL2 (Z) be a congruence subgroup. For example, we could set Γ
equal to
Γ0 (N ) := ac db ∈ GL2 (Z) : N |c . (6.10)
Definition 6.11. Let k ∈ Z>0 and let H be the complex upper half plane.
The space of weight k modular forms for Γ is the space Mk (Γ ) of holo-
morphic functions f : H → C satisfying the following conditions:
az+b
k ab
(a) f cz+d = (cz + d) f (z) for all ∈ Γ ∩ SL2 (Z) and all z ∈ H,
cd
(b) f extends holomorphically to the cusps.
If f additionally vanishes at the cusps we say that f is a cusp form. The
space of weight k cusp forms is denoted Sk (Γ ).
There are many books on modular forms that the reader can consult for more
details. We mention [DS05, GG12, Kob93, Lan95, Miy06, Ser73, Shi94]. A
recent extension of the theory to larger classes of functions that arise naturally
in applications is explained in [BFOR17].
We remark that if k is even then Mk (Γ ) can be identified with a subspace
of the space of holomorphic differential forms H 0 (Γ ∩SL2 (Z)\H, Ω ⊗k/2 ). This
explains the moniker “modular form.” This is explained in great detail in a
slightly more general setting in [GG12, Chapter 6].
We now relate the space Mk (Γ ) to a space of automorphic forms. The idea
is to pull back an automorphic form on H along the quotient map
In this setting
AGL2 := {( r r ) : r > 0} .
We set
j(g, z) = det(g)−1/2 (cz + d)
for
ab
g= ∈ GL2 (R)+ := {g ∈ GL2 (R) : det g > 0}.
cd
This is an example of an automorphy factor (see [GG12, §6.3] for details).
Set
ϕf (g) = j(g, i)−k f (gi) : GL2 (R)+ −→ C
6.6 From modular forms to automorphic forms 123
t
u
The isomorphism of Proposition 6.6.1 can also be phrased adelically. Let
n o
K0 (N ) := Γ\ a b ∈ GL (Z) b : N |c
0 (N ) = c d 2
One might ask if all cuspidal automorphic forms on GL2 arise from ele-
ments of Sk (Γ ). In fact we have missed many of them. The missing auto-
morphic representations correspond to Maass forms. From the automorphic
perspective it is very simple to define these objects; they are the automorphic
representations of GL2 (AQ ) whose associated (gl2 , O2 (R)) module is in the
principal series in the sense of §4.7. Defining them classically involves at least
as much work as required to formalize the notion of an element of Sk (Γ ).
We refer the reader to [Bum97, §3.2] for more details. The ease with which
we can talk about Maass forms and holomorphic cusp forms on equal footing
illustrates the utility of the language of automorphic representation theory.
124 6 Automorphic Forms
This ∆ should not be confused with the Casimir operator from the previous
section.
One has the following basic theorem:
Theorem 6.7.1 One has
The proof of this fact can be found in any of the standard references [DS05,
Lan95, Miy06, Kob93, Ser73, Shi94].
Define τ (n) ∈ C to be the unique integers such that
∞
X
τ (n)e2πinz = ∆(z).
n=1
In other words the τ (n) could be viewed as the Fourier coefficients on ∆(z).
Ramanujan conjectures that
∞
X τ (n) Y 1
t
= −t
(6.13)
n=1
n p
1 − τ (p)p + p11−2t
[Ram00b, (104)]. For the uninitiated, we pause and explain how fantasti-
cally prescient these two assertions are. In (6.13) Ramanujan has written
down the L-function of the automorphic representation attached to ∆(z)
over 20 years prior to Hecke’s general theory of L-functions for modular
forms [Hec37b, Hec37a] (Ramanujan’s assertion had been proven some time
earlier by Mordell [Mor]). The assertion (6.14) was proven by Deligne as a
consequence of his proof of the Weil conjectures [Del73a], however, the nat-
ural generalization of (6.14) to automorphic representations of GLn remains
open even in the case n = 2; it will be discussed in Conjecture 10.6.4 later.
6.7 Ramanujan’s ∆-function 125
Exercises
Deduce that the notion of moderate growth and uniform moderate growth is
independent of the choice of faithful representation ι0 : G → GLn .
V −→ V
Z
χ (k)dk
ϕ 7−→ π(k)ϕ σ .
K d(σ)
Show that this linear map is an idempotent that projects V onto the σ-
isotypic subspace V (σ).
6.4. Let F be a local field and let G be a connected reductive group over F .
Let K1 , K2 ≤ G(F ) be maximal compact subgroups. If F is nonarchimedean,
show that K1 ∩ K2 is a compact open subgroup of G(F ) and deduce that
a function on G(F ) is K1 -finite if and only if it is K2 -finite. Construct an
example to show that if F is archimedean then a function that is K1 -finite
need not be K2 -finite.
π∼
= ⊗0v πv
where for almost every v the representation πv is unramified in the sense that
it contains a (unique) vector fixed under a hyperspecial subgroup K ≤ G(Fv ).
In this section we discuss the classification of unramified representations. It
turns out that they can be explicitly parametrized in terms of conjugacy
classes in the dual group of G (see Theorem 7.2.1 and Theorem 7.5.1). This
127
128 7 Unramified Representations
Cc∞ (G(F ) // K) −→ C
f 7−→ tr π(f )
V 7−→ V K .
Here we have given the objects of the categories on both sides; morphisms
are simply intertwining maps.
Any irreducible representation is generated by any nonzero subspace, so
an unramified representation is generated by V K . We deduce the proposition.
t
u
7.2 Satake isomorphism 129
If we did not know anything about Cc∞ (G(F ) // K), then we could hardly
regard Proposition 7.1.1 as useful. However, it turns out that Cc∞ (G(F ) // K)
has a simple description:
Theorem 7.2.1 (Satake) Assume that G is split. There is an isomorphism
of algebras
S : Cc∞ (G(F ) // K) −→ C[Tb]W (G,T )(C)
b b
where Gb is the complex connected reductive algebraic group with root datum
dual to that of G and Tb ≤ G
b is a maximal torus. t
u
To gain intuition for the Satake isomorphism, let us consider the special
case of GLn . The Hecke algebra Cc∞ (GLn (F ) // GLn (OF )), as a C-module,
has a basis given by
1λ := 1
λ
$1 1
GLn (OF ) .. GLn (OF )
.
λn
$n
(S −1 )∗
Hom(Cc∞ (G(F ) // K), C) −→ Hom(C[Tb]W (G,T ) , C) −→ Gss (C)/conj
b b
(7.3)
where the first map is pullback along the inverse of the Satake isomorphism.
Here homomorphism means homomorphism of C-algebras, and Gss ⊂ G is the
closed subscheme consisting of semisimple elements. In Proposition 7.1.1 we
saw that unramified representations were in bijection with Hecke characters,
which are precisely elements of Hom(Cc∞ (G(F ) // K), C). We have therefore
proven the following corollary of the Satake isomorphism:
Corollary 7.2.2 Assume that G is split. The composite isomorphism (7.3)
induces a bijection between semi-simple conjugacy classes in G(C)
b and iso-
morphism classes of irreducible unramified representations of G(F ). t
u
From the point of view of automorphic representation theory the fact that
the Satake isomorphism is only valid for split groups is problematic. Indeed,
suppose for the moment that G is a reductive group over a global field F .
Then for all but finitely many places v of F the group GFv is unramified by
Proposition 2.4.3 and hence in particular is quasi-split. However, the group
GFv can be nonsplit for infinitely many v (see Exercise 7.3).
Langlands was able to extend the Satake isomorphism to the quasi-split
case. Historically this was important because it gave crucial hints as to the
structure of the Langlands dual group, which he introduced at the same time
[Lan] (see also [Lan70]). We will define the Langlands dual group in the next
section and use it to extend the Satake isomorphism in §7.5.
For the moment, let G be a connected reductive group over a global or local
field F and let T ≤ G be a maximal torus. To these data we associated in
§1.8 a root datum Ψ (G, T ) = (X ∗ (T ), X∗ (T ), Φ, Φ∨ ). We remind the reader
that X ∗ (T ) and X∗ (T ) are the character groups, Φ ⊂ X ∗ (T ) is the set of
roots of T in g and and Φ∨ ⊂ X∗ (T ) is the set of coroots. We also remind
the reader that the root datum characterizes GF sep where F sep is a separable
closure of F . To ease on notation, we let
If G is not split then the L-group has a more complicated definition involv-
ing a Galois action on G b that records the fact that G is a nonsplit group over
F . To define it we must construct an action of GalF on G(C).b It is obvious
that the root datum should be involved in this process. Indeed, given a Galois
action on GF sep we obtain one on Ψ (GF sep , TF sep ) and hence tautologically
we obtain one on Ψ (G, b Tb). This is not enough, as we really need a Galois
action on G(C). It would suffice to produce a splitting of the surjective map
b
Aut(G(C))
b −→ Aut(Ψ (G,
b Tb))
(X, Y, ∆, ∆∨ ) (7.4)
Proof of Lemma 7.3.1: We just explain the bijection, leaving the details to
the standard references mentioned at the beginning of Chapter 1. Given a
132 7 Unramified Representations
On the other hand, if g ∈ G(k), then the conjugate of this section may be
thought of as Aut(gBg −1 , gT g −1 , {gXα g −1 }α∈∆ ), which is also a pinning of
Gk . t
u
With this preparation complete we can now define the Langlands dual
group. Assume that G is a connected reductive group over a local or global
field F . Choose a Borel subgroup B ≤ GF sep and a maximal torus T ≤ B ≤
GF sep . We obtain a based root datum
(X∗ (T ), X ∗ (T ), ∆∨ , ∆) = Ψ (G,
b B,
b Tb). (7.8)
b → Aut(Ψ (G,
Aut(G) b B,
b Tb))
and hence a map GalF −→ Aut(G).b We define the Langlands dual group
of G or the L-group of G to be the semidirect product
L
G := G(C)
b o GalF
with respect to this action. Notice that the action is canonical up to conju-
gation by G(C)
b by Proposition 7.3.2.
A morphism of L-groups
L
H −→ L G
is simply a homomorphism commuting with the projections to GalF such that
its restriction to the neutral components is induced by a map of algebraic
groups H b −→ G.b Morphisms are also sometimes called L-maps.
Assume for the moment that F is a global field. For every place v upon
choosing an embedding F sep → Fvsep we obtain an embedding
GalFv −→ GalF ,
where GalF acts via permuting the factors. For a slightly more complicated
example let M/F be a quadratic extension and let σ ∈ Gal(M/F ). For an
F -algebra R, consider the quasi-split unitary group
1 1
−t
U (R) := g ∈ GLn (M ⊗F R) : ..
. σ(g) ..
. =g .
1 1
One has
L
U∼
= GLn (C) o GalF (7.11)
where GalF acts via its quotient Gal(M/F ), which acts on GLn (C) via the
isomorphism g 7→ g −t . We mention that there is an L-map
L
U −→ L ResM/F GLn
g o σ 7−→ (g, g −t ) o σ.
(P/G)(k) ←→ L P/L G.
Note that we view (P/G)(k) above as a group scheme opposed to the fact
that L P/L G is a group.
We say that a parabolic subgroup of L G is relevant if its L G-conjugacy
class is in the image of the composite
(X ∗ (T ), X∗ (T ), J(P ), J(P )∨ )
G(C)
b o FrZ
G(C)/conj.
b
G(R)
b × G(R)
b o Fr −→ G(R)
b oσ
(g, (h o Fr)) 7−→ ghFr(g)−1 o Fr.
Tb o Fr/F N −→ G
b o Fr/ ∼ (7.14)
(see [Bor79, Lemma 6.5]). With this isomorphism in mind the following the-
orem is natural:
7.5 The Satake isomorphism for quasi-split groups 137
t
u
As before, we can rephrase this as an isomorphism
t
u
By way of terminology the Fr-semisimple conjugacy class attached to
an isomorphism class of unramified representations is called its Langlands
class. In the split case, the eigenvalues of a representative of the conjugacy
class are called its Satake parameters.
We will not prove either Theorem 7.2.1 or Theorem 7.5.1, though we will
say more about the definition of the map S in a moment. The standard
references are [Car79, Theorem 4.1], [Bor79, §7], [Sat63], which we follow.
There is a categorified version of the Satake correspondence due to Mirković
and K. Vilonen that is known as the geometric Satake correspondence, see
[MV07]. This result is an important tool in the geometric Langlands program.
Moreover it has now been extended to mixed characteristic by work of Zhu
and Bhatt-Scholze.
Let B ≤ G be a Borel subgroup containing a maximal torus T of G and let
K be a hyperspecial subgroup. We let N denote the unipotent radical of B.
We say that K is in good position with respect to (B, T ) if the Iwasawa
decomposition
G(F ) = KB(F )
holds and
B(F ) ∩ K = (K ∩ T (F ))(K ∩ N (F )).
138 7 Unramified Representations
where
Z
f B (t) := δB (t)1/2 f (tn)dn. (7.17)
N (F )
It is not hard to see that this isomorphism is W (G, T )(F )-equivariant. Since
one has an identification
Since Tbs is the fixed points of Fr on Tb, we see that we obtain a map
that is equivariant with respect to the action of F N (which acts via its quo-
tient W (G, Ts )(F ) on the right hand side). It is clearly surjective. To prove
injectivity, assume that t, t0 ∈ Tb(C) and
for some w ∈ W (G, T )(F ). Then ν(t) = ν(w−1 t0 w) since w is fixed by Fr.
Thus t = xw−1 t0 w for some x ∈ ker(ν). By Hilbert’s theorem 90 all x ∈ ker(ν)
are of the form u−1 Fr(u), and we deduce that t o Fr = (wu)−1 (t0 o Fr)wu. t u
Combining lemmas 7.5.3, 7.5.4 and the map of coordinate rings induced
by Lemma 7.5.5 we see that we have constructed the Satake morphism
though we have not proved that it is injective or surjective. The first map in
this factorization of the Satake isomorphism is a special case of a parabolic
descent map. In general, any construction that relates objects on the group
G to objects on Levi subgroups of its parabolic subgroups (in this case the
Levi subgroup T of the Borel subgroup B) is known as parabolic descent. We
will use this idea in the next section to give a more explicit parametrization
of unramified representations.
140 7 Unramified Representations
is
δB (b) := | det(Ad(b))|.
The last ingredient we need to define unramified principal series representa-
tions is unramified quasi-characters.
X ∗ (T )F = X ∗ (Ts )F .
HT : T (F ) −→ aT := Hom(X ∗ (T )F , R)
via
λ ∈ a∗T C := X∗ (T )F ⊗ C
X ∗ (T )F ⊗ C −→ X ∗ (T )F ⊗ C× = Tbs (C)
induced by the map x 7→ e−x (for the last equality see Exercise 7.5). Thus
for every element of Tbs (C) we obtain an unramified quasi-character of T (F ).
Taking T = G in the following lemma we see that all unramified quasi-
characters arise in this manner:
Lemma 7.6.1 One has natural F N -equivariant bijections
provided that we let F N act via its quotient W (G, T )(F ) ≤ W (G, T )(F sep ) =
W (G,
b Tb)(C) on the left two groups.
Proof. For the first isomorphism we start by noting that the inclusion Ts → T
induces a W (G, T )(F )-equivariant isomorphism
Ts (F )/Ks −→T
˜ (F )/KT ,
C× −→ Hom(F × /OF× , C× )
e−λ 7−→ (t 7→ q λ log |t| )
λ ∈ ia∗T
I(λ)∨ ∼
= I(−λ).
(c) Assume that λ ∈ iaT . Under this assumption the representation I(λ) is
irreducible if and only if λ is regular.
t
u
For the proof of the theorem one can refer to [Car79, §3-§4]. These results are
due to Casselman, and one can also find proofs of them in his unpublished
notes [Cas]. For λ ∈ a∗T C we let J(λ) be the unique unramified subquotient
of I(λ). The reason for the apparent asymmetry in item (b) is that J(λ)
is not necessarily an irreducible subrepresentation of I(λ), merely a sub-
quotient. It is however an irreducible subrepresentation of I(w(λ)) for some
w ∈ W (G, T )(F ).
By the theorem, every unramified representation is isomorphic to a J(λ)
where λ is unique up to the action of W (G, T )(F ). This is consonant with
Corollary 7.5.2, the parametrization of unramified representations afforded
by the Satake isomorphism, and it is useful to explicitly relate Corollary
7.5.2 and Theorem 7.6.4:
Proposition 7.6.5 Let f ∈ Cc∞ (G(F ) // K). For λ ∈ a∗T C one has
Since J(λ), being unramified, has a unique spherical line this line its image
must be Cϕ0 under (7.21). t
u
and
Y
c(G) := {(cv ) ∈ L
G◦ o Frv / ∼: cv is Frv -semisimple for all v 6∈ S}/ ∼
v6∈S
where we say that v6∈S cv ∼ v6∈S c0v if and only if cv = c0v for almost every
Q Q
v 6∈ S (i.e. all v outside of a finite set). There is no need to be specific about
the set of places S in the notation here because we can always enlarge S by a
finite set. Any weak L-packet has a representative that is unramified outside
of S, so we have a map
7.7 Weak global L-packets 145
c(H) −→ c(G).
The fact that the class c(Π) is only defined up to a finite set of places
is an irritant, but it is not as horrible as it first appears. To explain, we
will require some results described in more Pddetail in §10.6. Let n1 , . . . , nd
be a collection of positive integers with i=1 ni = n. Let P = M N ≤
GLn be the parabolic subgroup of type (n1 , . . . , nd ) with its standard Levi
decomposition (see Example 1.10). For each i let πi be a cuspidal automorphic
representation of AGLni \GLni (AF ) and let λ ∈ a∗M . We can then form the
induced representation
I(π1 , . . . , πnd , λ)
as in §10.3. In general it may have several irreducible subquotients, all of
which turn out to be automorphic. However it always has a canonical sub-
quotient that will be described in more detail in §10.6. Automorphic repre-
sentations equivalent to this canonical subquotient are known as isobaric
representations. In particular, if I(π1 , . . . , πnd , λ) is irreducible then it is
isobaric; this holds in particular if n1 = n which is to say that the represen-
tation is cuspidal.
We now record the following foundational fact from [JS81b, JS81a]:
146 7 Unramified Representations
Exercises
V 7−→ V K .
7.3. Let E/F be a quadratic extension of number fields with Galois auto-
morphism σ. For any F -algebra R, let
7.5. Let T be a torus over C. Prove that the set valued functors on the
category of C algebras are all naturally equivalent: For C-algebra R,
7.7. Let G be an affine algebraic group over a local field F and let V be a
topological C-vector space (possibly of infinite dimension). Prove that a map
G(F ) → V
is continuous with respect to the natural topology on G(F ) and the given
topology on V if and only if it is locally constant. Conclude that the map
G(F ) → V is continuous if and only if it is continuous when we give V the
discrete topology.
πS0 ⊗ π S
8.1 Introduction
δP := δP (F ) : P (F ) −→ R× (8.1)
is
δP (p) := | det(Adp (p) : p(F ) → p(F ))|.
Let (σ, V ) be a smooth irreducible representation of M (F ). Using the mod-
ular character we define the parabolically induced representation (or
simply induced representation) I(σ) := IndG P (σ) to be the (smooth) rep-
resentation of G(F ) on the space of functions
n o
IndGP (V ) := locally constant ϕ : G(F ) −→ V : ϕ(mng) = δP (m)1/2
σ(m)ϕ(g)
149
150 8 Nonarchimedean Representation Theory
IndG
P : Repsm M (F ) −→ Repsm G(F ), (8.3)
The main result of this chapter is a theorem of Jacquet which states that
every irreducible admissible representation is obtained as a subquotient of
the parabolic induction of a supercuspidal representation for an appropri-
ately chosen parabolic subgroup (see Corollary 8.3.5 for the precise state-
ment). One can profitably compare this result with the Langlands classifica-
tion from §4.9 and the classification of automorphic representations explained
in Chapter 10. These results can all be viewed as manifestations of what
Harish-Chandra called the “philosophy of cusp forms” [HC70a]. Interpreted
broadly, this is a slogan for the statement that all irreducible representations
of reductive groups can all be obtained via parabolic induction from cuspi-
dal representations of parabolic subgroups. Of course, the correct notion of
cuspidal representation varies based on the context.
We will not give a full proof of Theorem 8.3.5, but we will develop much
of the machinery used in the proof because it is useful in many contexts. We
will also discuss the notion of a trace of an admissible representation, and
8.2 Parabolic induction 151
Let us now prove some basic facts about parabolic induction. Let (σ, V ) be
a smooth representation of M (F ).
We start with the following observation:
Lemma 8.2.1 If K ≤ G(F ) is a compact open subgroup and ϕ ∈ IndG
P (V )
K
then −1
ϕ(g) ∈ V M (F )∩gKg .
t
u
for all x ∈ G(F ) (see Exercise 8.3). For x ∈ G(F ) one has
Z
(ϕ1 (kx), ϕ2 (kx))V dk
Kmax
Z Z
= (ϕ1 (kx), ϕ2 (kx))V f (pkx)dr pdk
Kmax P (F )
Z
= (ϕ1 (kx), ϕ2 (kx))V f (pkx)δP−1 (p)d` pdk (see Exercise 3.3)
P (F )×Kmax
Z
= (ϕ1 (pkx), ϕ2 (pkx))V f (pkx)d` pdk
P (F )×Kmax
Z
= (ϕ1 (gx), ϕ2 (gx))V f (gx)dg
G(F )
Z
= (ϕ1 (g), ϕ2 (g))V f (g)dg.
G(F )
Proof. Fix a maximal compact open subgroup Kmax ≤ G(F ) so that the
Iwasawa decomposition G(F ) = P (F )Kmax holds (see Theorem 2.7.1).
Let ϕ ∈ IndG ∨ G ∨
P (V ) and ϕ ∈ IndP (V ). We then have a pairing
Z
∨
(ϕ, ϕ ) := hϕ(k), ϕ∨ (k)idk (8.5)
Kmax
where the pairing on the right is the pairing between V and V ∨ . Arguing as
in the proof of part (b) of Proposition 8.2.2 we deduce that this pairing is
G(F )-invariant.
Since ϕ∨ is smooth the linear form h·, ϕ∨ i on V is smooth and we therefore
obtain a C-linear map
∨ ∨
IndG G
P (V ) −→ IndP (V ) (8.6)
∨ ∨
ϕ 7−→ (·, ϕ ).
8.2 Parabolic induction 153
{ϕx,w : x ∈ X, w ∈ Bx } (8.7)
−1
is a C-basis of I(σ)K . Now for each x let Bx∨ be the basis of V ∨M (F )∩xKx
dual to Bx . We define ϕx,w∨ by replacing V with V ∨ in the construction
above, and we obtain a basis
for IndG ∨ K
P (V ) . It is clear that (8.7) and (8.8) are a basis and dual basis with
respect to the pairing (8.5). t
u
HM : M (F ) −→ aM := Hom(X ∗ (M )F , R) (8.9)
via
hHM (m), λi = log |λ(m)|
for λ ∈ X (M )F . For each λ ∈ a∗M C we then obtain a quasi-character
∗
The point of introducing this extra notation is that it makes clear that the
induced representation I(σ, λ) is part of a continuous family, indexed by a∗M C ,
of representations. In practice this can be very useful. We employed this
154 8 Nonarchimedean Representation Theory
ev1 : IndG
P (W ) −→ W
ϕ 7−→ ϕ(1)
given by composition with ev1 . Here we are using the fact that N acts triv-
ially on IndG G
P (W ) and thus any G(F )-equivariant map V → IndP (W ) fac-
tors through VN . To construct the inverse, suppose we are given an M (F )-
intertwining map Φ : VN → W . We define
e : V −→ IndG (W )
Φ P
ϕ 7−→ (g 7→ Φ(g.ϕ)).
8.3 Jacquet modules 155
The map Φ 7→ Φ
e is inverse to ev1 ◦ (·). t
u
One would hope that the functor N preserves admissibility, and this is
indeed the case:
Theorem 8.3.2 (Jacquet) The functor N takes admissible representations
to admissible representations. t
u
Jacquet also proved the following elegant characterization of quasicuspidal
representations using these functors:
Theorem 8.3.3 (Jacquet) A smooth irreducible representation (π, V ) of
G(F ) is quasicuspidal if and only if VN = 0 for all parabolic subgroups P ≤ G.
t
u
where
1
eK := 1K
meas(K)
is the idempotent attached to K. We must show V K is finite dimensional.
Since π admits a central quasi-character (see Exercise 5.7) if V K is not finite
dimensional then there exists {gn : n ∈ Z>0 } ⊆ G(F ), all inequivalent modulo
ZG (F ), such that π(eK )π(gn )ϕ are linearly independent. Let W ⊆ V K be a
C-vector subspace space such that
for all n and ϕ∨ |W ⊕ker π(eK ) = 0. Then ϕ∨ is fixed by K, hence is smooth, and
hence is an element of V ∨ . On the other hand, by construction the support of
the matrix coefficient hπ(g)ϕ, ϕ∨ i is not compact modulo the center ZG (F ).
This implies the proposition. t
u
Proof. The first assertion implies the second, as induction preserves admis-
sibility by Proposition 8.2.2. For the first, we proceed by induction on the
dimension of G (as an F -algebraic group, say). If G has dimension 1 then it
is a torus, and so the result is trivial.
Assume that for all proper parabolic subgroups P there is no nonzero
intertwining map V → IndG P (W ) where (σ, W ) is a smooth irreducible rep-
resentation of M (F ). Applying Frobenius reciprocity (Proposition 8.3.1) we
see that
HomM (F ) (VN , W ) = 0
for all parabolic subgroups P = M N and all smooth representations σ of
M (F ), which implies π is supercuspidal by Theorem 8.3.3.
Now assume that there is a proper parabolic subgroup P < G, a Levi
subgroup M ≤ P , a smooth representation (σ, W ) of M (F ), and a nonzero
(hence injective) intertwining map V → IndG P (W ). By Frobenius reciprocity,
there is a nonzero intertwining map
VN −→ W
V −→ IndG M ∼ G
P ◦ IndQ (U ) = IndQN (U )
Theorem 8.3.5 provides the first step towards a classification of all irreducible
admissible representations of G(F ) in terms of supercuspidal representations.
The remaining task is to understand isomorphisms between subquotients of
induced representations.
It is useful to start by stepping backwards and recalling the Langlands clas-
sification in the nonarchimedean setting. This classifies irreducible admissible
representations in terms of tempered representations and is the analogue in
the current setting of Theorem 4.9.2.
Fix a minimal parabolic subgroup P0 ≤ G and call a parabolic subgroup
standard if it contains P0 . Let M (F )1 := ker HM where HM is defined as
in (8.9). Let (σ, V ) be an irreducible admissible representation of M (F )1 .
We extend it trivially to M (F ). Given λ ∈ a∗M we can form the induced
representation I(σ, λ) as in (8.10).
The following is the analogue of Theorem 4.9.1 in this setting:
Theorem 8.4.1 If σ is unitary and tempered and λ is in the positive Weyl
chamber then I(σ, λ) admits a unique irreducible quotient J(σ, λ). t
u
The representation J(σ, λ) is known as the Langlands quotient of I(σ, λ).
The only reference for the proof appears to be [Ber, Proposition 42]; the
theorem is stated in [Zel80, §9.1]. There is a sketch of the proof in [JS83,
§1.2].
To proceed we need the notion of linked representations. Following [BZ77]
one calls any set of supercuspidal representations of GLr (F ) for some r of
the form
for some integer d a segment. Two segments are linked if neither is included
in the other and their union is a segment.
To each J(σ a , λa ) we associate the segment
a−1 a−1
{σ ⊗ ehHM (·), 2 i , . . . , σ⊗hHM (·),− 2 i }.
We say that
0
J(σ a , λa ) and J(σ 0a , λa0 ) (8.13)
Proof. In [JS83, §1.2] one finds an argument that proves that an irreducible
pre-unitary tempered representation is generic. In fact any tempered rep-
resentation is pre-unitary (see Definition 4.6), so any irreducible tempered
representation is generic. Now since the J(σiai , λai ) are all square integrable
none of them are linked by Exercise 8.15. Thus we can then use [Zel80, The-
orem 9.7(b)] to conclude the assertion of the theorem. t
u
π(f ) : W −→ W
for any such W . The notion of the trace of a representation will play a crucial
role in the trace formula in later chapters.
In this nonarchimedean setting, a distribution on G(F ) is simply a com-
plex linear functional on Cc∞ (G(F )). Thus the trace map
tr π : Cc∞ (G(F )) −→ C
(see [HC99] for the proof and [Kot05] for a detailed exposition).
In other words there is a locally constant function Θπ on Greg (F ) such
that Z
tr π(f ) = Θπ (g)f (g)dg
G(F )
for all f ∈ Cc∞ (G(F )).This result tells us that we can almost regard tr π as
a function.
The following is a version of linear independence of characters adapted to
this setting:
Proposition 8.5.2 (Linear independence of characters) If π1 , . . . , πn is
a finite set of admissible irreducible representations such that πi ∼
= πj implies
i = j, then the distributions tr πi are linearly independent.
tr πi (f ) = 0 if and only if i 6= 1.
8.5 Traces, characters, coefficients 161
For a refinement of this result at the level of operators see Exercise 8.10.
One can ask for more. Let π be an admissible irreducible representation. A
coefficient of π is a smooth function fπ ∈ Cc∞ (G(F )) such that tr π(fπ ) 6= 0
and tr π 0 (fπ ) = 0 for π 0 ∼
6 π. Thus if a coefficient for π exists, we can use it
=
to isolate π among any set of irreducible admissible representations, finite or
not. For general π, it is not necessarily true that such functions f exist. To put
this in perspective, it is useful to recall the Heizenberg uncertainty principle,
namely that the Fourier transform of a compactly supported smooth function
on a real vector space cannot again be compactly supported. Thus if we
replaced G(F ) with the nonreductive group R there would be no way to
construct coefficients. However, we are not in this setting, and in certain
circumstances we can construct coefficients:
Proposition 8.5.3 Assume that ZG (F ) is compact, and let (π, V ) be an
irreducible supercuspidal representation of G(F ). Then for all f ∈ Cc∞ (G(F ))
there exists a unique fπ ∈ Cc∞ (G(F )) such that
is naturally a G(F ) × G(F )-module, where one copy of G acts via precom-
position and the other via postcomposition. The action is given explicitly by
(g1 , g2 ).A = π(g1 ) ◦ A ◦ π(g2−1 ). We let
Endsm (V ) ≤ End(V )
given by
β(A)(g) = tr(π(g) ◦ A).
Here ρ acts via ρ(g1 , g2 )(f )(h) = f (g1−1 hg2 ). We note that for each A ∈
Endsm (V ) the function β(A) is a sum of matrix coefficients. Indeed, if we
choose a compact open subgroup K ≤ G(F ) such that A is fixed on the left
and right under K, choose a basis B of V K and a dual basis B ∨ of V ∨K then
X
β(A)(g) = hπ(g) ◦ Aϕ, ϕ∨ i. (8.17)
ϕ∈B
hϕ, ϕ∨ i = hπ(1)ϕ, ϕ∨ i =
6 0,
we have that β is not identically zero. Thus since the exterior tensor product
π ⊗ π ∨ is irreducible β is an embedding.
Consider
Parabolic descent is a term for the process by which one passes from objects
(say, representations) on a group to objects on a Levi subgroup of a parabolic
subgroup. In this section we give two examples of this phenomenon. First we
show that the character of a parabolically induced representation is easily
related to the character of the inducing representation (Proposition 8.6.1).
Let P ≤ G be a parabolic subgroup with Levi decomposition P (F ) =
M (F )N (F ) and let K ≤ G(F ) be a maximal compact subgroup.
KM := M (F ) ∩ K
is a maximal compact open subgroup (see Theorem A.4.2); we will also as-
sume this in what follows.
We assume throughout this section that Haar measures are chosen so that
dg = δP (m)dkdmdn (8.20)
tr I(σ)(f ) = tr σ(f P ).
IndG G
P (V ) −→ IndP (V )|K (8.25)
G(F )
that intertwines ResK (I(σ)) with I(σ|K ). Since this is an isomorphism,
I(σ)(f ) induces an endomorphism of IndGP (V )|K , and tr I(σ)(f ) is equal to
the trace of this endomorphism on IndG
P (V ). Define
8.6 Parabolic descent of representations 165
Z
1/2
Φk,k0 (m) := δP (m) f (k −1 mnk 0 )dn
N (F )
It follows that
tr I(σ)(f )
σ(Φk,k0 )ϕ(k 0 )dk 0 ). This en-
R
is the trace of the endomorphism ϕ 7→ (k 7→ K
domorphism has trace Z
tr σ(Φk,k )dk.
K
On the other hand Z
fP = Φk,k dk.
K
The lemma follows. t
u
As an addendum, we prove the following lemma:
Lemma 8.6.2 The constant term provides an algebra homomorphism
This proves our assertion that the constant term is an algebra morphism. t
u
Note that this is not the same as the local version of the orbital integral
appearing in (18.11), it differs in that in (18.11) we use Cγ instead of Cγ◦ .
This is why we there is a superscript ◦ in the notation. The quotient map
8.7 Parabolic descent of orbital integrals 167
Cγ (F )\G(F ) −→ G(F )
Cγ (F )g 7−→ g −1 γg
where (as usual) h := Lie H and g := Lie G and where I is the identity map.
There is a dual relationship between orbital integrals and characters that
can be made precise in many ways; one of the most profound is the absolute
trace formula (see §18.4). The following proposition is consonant with this
duality:
Proposition 8.7.1 Assume that γ ∈ G(F ) is closed, γ ∈ M (F ), and Cγ◦ ≤
M . If DM \G (γ) 6= 0 then one has
Proof. Upon taking a change of variables n 7→ mnm−1 we see that the nor-
malization of the Haar measure in (8.20) is equivalent to
d(mnk) = dmdndk.
N (R) −→ N (R)
(8.29)
n 7−→ (m−1 γm)−1 n−1 (m−1 γm)n.
168 8 Nonarchimedean Representation Theory
We claim that this morphism is an isomorphism. The image is the orbit of the
identity under an action of the unipotent group N , and is therefore closed in
N by [Mil17, Theorem 17.64]. On the other hand the morphism is injective
at the level of points. Indeed if γn−1 γn = γn0−1 γn0 for n, n0 ∈ N (R) then
n0 n−1 ∈ Cγ (R) which implies n0 n−1 = I by our assumption that Cγ◦ ≤ M
(see [Hum95, §2.2, Theorem]). The same is true if we replace γ by m−1 γm
and injectivity follows. In particular the stabilizer of the identity element is
trivial. We concluded using Proposition 17.1.2 that (8.29) is an isomorphism
onto its image, a closed subscheme of N . By considering dimensions we deduce
that (8.29) is an isomorphism.
The Jacobian depends only on γ and is equal to
We note that
|DM \G (γ)| = δP (γ)J(γ)2 .
Therefore we can change variables and deduce that the integral above is
Z Z Z
dn dm
f (k −1 m−1 γmnk)dk
Mγ◦ (F )\M (F ) N (F ) K J(γ) dmγ
{γ ∈ M (F ) : Cγ◦ = M } (8.30)
m\gγ ≤ m\g
Remark 8.2. There are surprisingly few textbook treatments of analytic man-
ifolds in the nonarchimedean setting; one is [Ser06] and some additional topics
are in [Igu00]. Perhaps the reason is that more sophisticated treatments in-
volving rigid analytic spaces, Berkovich spaces, etc. are often needed and
have drawn more attention.
for all γ ∈ M (F ).
The set of γ ∈ M (F ) such that Cγ◦ = M is dense by Proposition 8.7.3, and
for such a γ one has |DM \G (γ)| =
6 0. Let dm be the Haar measure on M (F )
used in (8.20). We then have
Exercises
dg(kmn) = δP (m)dkdmdn
8.5. The functor IndG P is exact (that is, it preserves exact sequences) and
additive (that is, takes direct sums to direct sums).
A −→ C
8.7 Parabolic descent of orbital integrals 171
a 7−→ tr(a : Vi → Vi )
9.1 Introduction
Let G be a reductive group over a global field F . For x ∈ G(AF )1 , one has
the operator
173
174 9 The Cuspidal Spectrum
that this restriction is of trace class. One is then left with understanding the
complement of the cuspidal subspace; this is possible thanks to the theory of
Eisenstein series which will be discussed in Chapter 10.
Before stating the main theorems of this chapter we pause and explain
what one means by a trace in this infinite-dimensional setting.
∞
X
kAϕi k22 < ∞.
i=1
If A is Hilbert-Schmidt we let
∞
X ∞
X
kAkHS = kAϕi k22 and kAktr = kAϕi k2 . (9.3)
i=1 i=1
The Hilbert-Schmidt norm, trace norm, and trace, are independent of the
choice of basis ϕ (Exercise 9.1).
For the reader’s convenience we recall Definition 6.9:
L2cusp ([G]) = ⊕
b π L2cusp (π), (9.4)
e F ) := H(AF ) ∩ G(AF )1 .
H(A
These functions will be revisited in §16.1 below; there we will prove that they
are absolutely convergent and define smooth functions on G(AF )1 . We define
the Poincaré series
Z
Péf (g) := Péf,H,χ (g) := Kf (h, g)χ(h)dh. (9.8)
H(F )\H(A
f F)
P
Since γ∈G(F ) fe(γg) is a compactly supported continuous function on [G],
this integral converges. t
u
In the proof of the next lemma we will encounter for the first time the
trivial, but useful, technique known as unfolding. We formalize this in the
following lemma, the proof of which is an application Theorem 3.2.2:
Lemma 9.2.4 Suppose that G is a Hausdorff, locally compact, second count-
able topological group with right Haar measure dg. If f ∈ L1 (G) and Γ ≤ G
is a discrete subgroup such that the modular character of G is trivial on Γ
then
178 9 The Cuspidal Spectrum
Z X Z
f (γg)dg = f (g)dg.
Γ \G γ∈Γ G
t
u
Here in the lemma we have denoted also by dg the right G-invariant measure
on Γ \G induced by dg. Use this technique, we prove the following lemma.
Lemma 9.2.5 One has
Z
Péf (g)ϕ(g)dg = Pχ (R(f )ϕ).
[G]
Here in the last equality we have unfolded as in Lemma 9.2.4 to deduce that
Z Z
Kf (h, g)ϕ(g)dg = f (g)ϕ(hg)dg. (9.11)
[G] G(AF )1
The integral on the right converges absolutely and defines a continuous func-
tion of h. Combining this with the compactness of H(F )\H(A e F ) and domi-
nated convergence we deduce that our formal manipulations above are justi-
fied. t
u
Proof of Proposition 9.2.2: Each f ∈ Cc∞ (G(AF )1 ) gives rise to a linear form
(·, Péf ) on L2 ([G]). It is continuous by Lemma 9.2.3. In view of Lemma 9.2.5
9.3 Deduction of the discreteness of the spectrum 179
the space V is the intersection over all f ∈ Cc∞ (G(AF )1 ) of the kernels of
these linear forms. 2
for all n.
It is not difficult to see that a Dirac sequence on G(AF )1 exists (Exercise
9.7). They are known as Dirac sequences because for representations (R, V )
of G(AF )1 and ϕ ∈ V one has R(fn )ϕ → ϕ as n → ∞ (Exercise 9.8). We also
note that if fn is a Dirac sequence and (R, V ) is a unitary representation of
G(AF )1 then R(fn ) is self-adjoint for every n by condition (b).
We recall that an operator T on an infinite dimensional Hilbert space V
is compact if and only if it can be written as
∞
X
T = λn hϕn , ·iϕ0n (9.12)
n=1
operator where {ϕn }n>0 , {ϕ0n }n>0 are orthonormal sequences of vectors in
V and {λn }n>0 ⊂ C is a sequence of numbers such that limn→∞ λn = 0. If
T is self-adjoint, then we can take ϕn = ϕ0n , and in particular in this case T
has an orthonormal basis of eigenvectors. A nice reference is [Zhu93, §1.3].
Lemma 9.3.1 Let (R, V ) be a unitary representation of G(AF )1 . Assume
that there is a Dirac sequence {fn } on G(AF )1 such that R(fn ) is compact
for all n. Then (R, V ) decomposes into a Hilbert space direct sum of irre-
ducible closed G(AF )1 -invariant representations of G(AF )1 , each with finite
multiplicities.
In view of Lemma 9.2.1 and Lemma 9.3.1, to prove Theorem 9.1.1 it suffices
to show that Rcusp (f ) is compact for all f ∈ Cc∞ (G(AF )1 ). The following
lemma gives us a criteria for compacity:
Lemma 9.3.2 Let f ∈ Cc∞ (G(AF )1 ). Let V ≤ L2 ([G]) be a closed G(AF )1 -
invariant subspace. Assume one has an estimate
for all ϕ ∈ V . Then the operator R(f )|V is of trace class, hence Hilbert-
Schmidt, hence compact.
The assertions that trace class implies Hilbert-Schmidt implies compact are
left as Exercise 9.3. In the lemma and below k·k∞ denotes the uniform norm.
Proof. This follows from the argument of the proof of [Bor97, Theorem 9.5]
which we will recall here: For given x ∈ G(AF )1 , the assumption implies that
9.3 Deduction of the discreteness of the spectrum 181
the map ϕ 7→ R(f )ϕ(x) is a continuous linear form on V . Hence by the Riesz
representation theorem there exists an element KR(f ),x ∈ V such that
Z
R(f )ϕ(x) = KR(f ),x (y)ϕ(y)dy
[G]
is finite and Z
R(f )ϕ(x) = KR(f ) (x, y)ϕ(y)dy
[G]
for ϕ ∈ L2cusp ([G]). This estimate is really the heart of the proof. In the func-
tion field case we will be able to deduce something much stronger, and in the
number field case we derive it using the argument of [God66]. It is only in
this part of the proof where we use the assumption that G is reductive. We
have been careful in isolating when this assumption is used for two reasons.
First, it is helpful to see at what point the argument passes from abstract
representation theory to something that uses the specific structure of the
adelic quotient in the reductive case. Second, there ought to be a notion of
a cuspidal representation on more general groups than those that are reduc-
182 9 The Cuspidal Spectrum
tive, but we do not know the natural level of generality. We feel that this
is an interesting research problem. One example that has been worked out
reasonably thoroughly is the case where G is the Jacobi group, a semi-direct
product of SL2 and the three-dimensional Heisenberg group [BS98, §4.2]. The
work in [Sli84] is probably relevant here as well.
Assume for this section that F is a function field and that G is reductive.
The asymptotic behavior of cuspidal functions is much simpler in this case
than their behavior in the number field case.
Theorem 9.4.1 (Harder) Let ϕ ∈ L2cusp ([G]) be a function that is invari-
ant under the compact open subgroup K 0 ≤ G(AF )1 . Then ϕ(x) = 0 for x
outside a compact subset of [G] that depends only on K 0 .
Before proving this theorem let us indicate how it can be used to prove
Theorem 9.1.1 in the function field case.
Proof of Theorem 9.1.1 (for function field): Let f ∈ Cc∞ (G(AF )1 ). As ex-
plained in more detail in §16.1 the operator
Now R(f )|L2 (Ω) is Hilbert-Schmidt, so its restriction to the closed subspace
L2 (Ω) ∩ L2cusp ([G]) of L2 (Ω) is also Hilbert-Schmidt (see Exercise 9.5).
Thus we have proven that R(f ) : L2cusp ([G]) → L2cusp ([G]) is Hilbert-
Schmidt. On the other hand, every f ∈ Cc∞ (G(AF )1 ) is a finite linear com-
Pk
bination of convolutions: f = i=1 hi1 ∗ hi2 where hij ∈ Cc∞ (G(AF )1 ) (see
9.4 The function field case 183
Φ+
α ⊂Φ
+
n(Φ+
α ) := ⊕λ∈Φ+ n .
α λ
(9.14)
0
Proof of Theorem 9.4.1: Let ϕ ∈ L2cusp ([G])K . Choose t0 > 0 and a Siegel
set S(t0 ) := ΩAT0 (t0 )K so that AG G(F )S(t0 ) = G(AF ).
Choose t > t0 so that the conclusion of Lemma 9.4.2 holds. Then for
h ∈ S(t) the function
N (Φ+
α )(AF ) −→ C
n 7−→ ϕ(nh)
ϕ : [G] → C
ϕ : SL2 (Z)\H −→ C
(compare (2.14), (6.11)). We can choose a fundamental domain for SL2 (Z)
contained in the Siegel set
9.5 Estimates for the Fourier transform of the kernel 185
1
z ∈ H : Im(z) > , 0 ≤ Re(z) ≤ 1 (9.15)
2
The key point here is that cuspidality implies that only positive n contribute.
Since ϕ is continuous the an are bounded and this allows one to deduce (9.13)
in this case. In the discussion above we have used Fourier analysis on
Z\R ∼
= NB (Z)\NB (R)
to establish (9.13). The general case follows this basic outline, using Fourier
analysis on the unipotent radical of a minimal parabolic subgroup.
We now begin our discussion of the general case, so we let G be a reductive
connected group over a number field F . By applying Weil restriction of scalars
we can and do assume F = Q. Let P be a fixed minimal parabolic subgroup
of G with Levi decomposition P = M N (N being the unipotent radical). Let
T ≤ M be a maximal split torus and let T0 ≤ Gder a split torus such that
T0 ZG = T . The exponential map induces an isomorphism of affine schemes
over Q:
exp : n −→ N.
Here as usual n denotes the Lie algebra of N and we are viewing n as an
affine space over F , isomorphic to Gdima
Fn
. Note that we are not claiming
that the exponential map is an isomorphism with respect to any sort of
group structure.
Any element ϕ ∈ L2 ([G]) is left-invariant under G(Q). Thus for f ∈
∞
Cc (G(AQ )1 ) we can write
Z
R(f )ϕ(x) = f (x−1 y)ϕ(y)dy (9.16)
G(AQ )1
Z X
= f (x−1 γy)dyϕ(m)dm
N (Q)\G(AQ )1 γ∈N (Q)
Z
= Kf,P (x, y)ϕ(y)dy
N (Q)\G(AQ )1
η 7→ f (x−1 exp(η)y)
h , i : nP × nP −→ Ga
x ∈ ΩN ΩM sx K = ΩN sx ΩM K.
for some γ ∈ N (F ) then x−1 sx and s−1x y lie in compact subsets of G(AQ )
1
Proof. Since x ∈ S(t) one has by Lemma 9.5.1 that x ∈ sx ΩG for some
compact subset ΩG of G(AQ )1 depending only on the Siegel set S(t). We can
take ΩG to be invariant under g 7→ g −1 and then it follows that x−1 ∈ ΩG s−1
x .
Now if f (x−1 γy) 6= 0 then x−1 γy lies in a compact subset ΩG of G(AQ )1
depending only on f . By our earlier considerations we see that upon enlarging
ΩG if necessary
s−1 −1 −1
x γy = sx γuy sx sx my sy ky ∈ ΩG (9.19)
k·kn
n 1 Ωn
B,ΩN kηk−B −B
∞ (η)δP (sx )supλ(sx ) ,
hη, ·i|n(λ) 6= 0.
Using Lemma 9.5.2 and integration by parts at the infinite place we conclude
that for all B > 0
Z
x )ηkn 1Ωn
f (x−1 exp(n)y)ψ(hη, ni)dn δP (sx )kAd∨ (s−1 −B
∞ (η).
n(AQ )
Φ+
α ⊂Φ
+
n(Φ+
α ) := ⊕λ∈Φ+ n .
α λ
(9.20)
Lemma 9.5.4 Let α ∈ ∆, and assume that hη, ·i is trivial on n(Φα ). Then
Z
y 7−→ f (x−1 exp(n)y)ψ(hη, ni)dn
n(AQ )
t
u
9.6 The basic estimate 189
for x ∈ S(t).
Combined with Lemma 9.3.1 and Lemma 9.3.2 this completes the proof
of Theorem 9.1.1.
for each α ∈ ∆.
Now if η ∈ n(Q) has the property that hη, ·i is identically zero on n(Φ+
α)
then Z
y 7−→ f (x−1 exp(n)y)ψ(hη, ni)dn
n(AQ )
Z Z !
−1
measdu ([N (Φ+
α )]) f (x exp(n)y)ψ(hη, ni)dn ϕ(y)dy
N (Q)\G(AQ )1 n(AQ )
Z Z Z !
−1
= f (x exp(n)uy)ψ(hη, ni)dn ϕ(y)dȳdu
[N (Φ+
α )] N (Q)\G(AQ )1 n(AQ )
Z Z Z !
= f (x−1 exp(n)y)ψ(hη, ni)dn ϕ(u−1 y)du dȳ
N (Q)\G(AQ )1 n(AQ ) [N (Φ+
α )]
=0
190 9 The Cuspidal Spectrum
for B sufficiently large one can sum over η and apply Lemma 9.5.2 to obtain
a bound Z
−B
R(f )ϕ(x) δP (sx )α(sx ) ϕ(y)dy
sx ΩG
1
where ΩG ⊂ G(AQ ) is a dy measurable compact set. By the Cauchy-
Schwartz inequality this is bounded by
Z 1/2
−B 1/2 2
δP (sx )α(sx ) measdy (ΩN sx ΩG ) |ϕ(y)| dy . (9.22)
ΩN sx ΩG
One has
Z 1/2
2
|ϕ(y)| dy kϕk2 (9.23)
ΩN sx ΩG
and
measdy (ΩN sx ΩG ) = measdy s−1 −1
x ΩN sx ΩG δ(sx ) .
Thus (9.22) is bounded by a constant times
1/2
δP (sx )α−B (sx )kϕk2
as required. t
u
The estimate of Proposition 9.6.1 for cuspidal functions motivates the follow-
ing definition:
Definition 9.3. A function
ϕ : [G] −→ C
9.8 Cuspidal automorphic forms 191
Here we are using notation from our discussion of Siegel sets and reduction
theory in §2.7.
Using this terminology we see that Theorem 9.4.1 and Proposition 9.6.1
imply the following theorem:
Theorem 9.7.1 If f ∈ Cc∞ (G(AF )1 ) and ϕ ∈ L2cusp ([G]) then R(f )ϕ is
rapidly decreasing. t
u
for some fi ∈ Cc∞ (G(AF )1 ) and ϕi ∈ L2cusp ([G]). This follows from Theorem
4.2.4. t
u
AG 2
Since [G] has finite measure we deduce that Acusp ≤ Lcusp ([G]).
We now prove density. For a cuspidal automorphic representation π of
AG \G(AF ) let L2cusp (π) be the π-isotypic subspace of L2cusp ([G]). In view of
Corollary 9.1.2 it suffices to show that AA 2
cusp ∩ L (π) is dense in π.
G
∞
Let K = K∞ K be a maximal compact subgroup of G(AF ). We claim
that for each cuspidal automorphic representation π of AG \G(AF ) (in the L2
sense) the space of K-finite vectors in L2cusp (π) is exactly
AA 2
cusp ∩ Lcusp (π).
G
Exercises
9.4. Let (Y, µ) be a σ-finite measure space. Then L2 (Y, dµ) is a separable
Hilbert space. Let K(x, y) ∈ L2 (Y × Y, dµ × dµ). Prove that the operator
is Hilbert-Schmidt.
9.6. Let G be a connected reductive group over a global field F and let
H ≤ G be a subgroup such that H(F )\H(AF ) ∩ G(AF )1 is compact. Prove
that H(AF ) ∩ G(AF )1 is unimodular.
R(fn )ϕ −→ ϕ
as n → ∞.
9.9. Using Theorem 4.2.5 prove that every f ∈ Cc∞ (G(AF )1 ) can be written
as
r
X
f= f1r ∗ f2r (9.24)
i=1
9.10. With notation as in Lemma 9.5.1 prove that when x varies over S(t)
one has that s−1
x ΩN sx lies in a compact subset of N (AF ).
for n = 1 and n = 2.
Chapter 10
Eisenstein Series
Abstract In this chapter we survey the theory of Eisenstein series. Our main
goal is to state Langlands’ decomposition of L2 ([G]).
195
196 10 Eisenstein Series
HM : M (AF ) −→ aM (10.1)
defined by
hHM (m), λi = log |λ(m)|
∗
for λ ∈ X (M )F . Using the Iwasawa decomposition G(AF ) = N (AF )M (AF )K
we define a morphism
HP : G(AF ) −→ aM
nmk 7−→ HM (m)
Here IndG
P (V ) is the space of measurable functions
m 7→ ϕ(mx)
Thus IndG
P (V ) is a Hilbert space. The action is given by
where ρP ∈ a∗M is half the sum of the positive roots of a maximal split torus
of G contained in P ; of course we use P to denote the notion of positivity
defined by P (see §1.9).
Suppose now that σ is irreducible, so it is an automorphic representation
of AM \M (AF ). Since
Y
ehHM (m),λi = ehHM (mv ),λi
v
I(σ, λ) ∼
Y
= I(σv , λ)
v
where the local factors I(σv , λ) are defined as in §4.9 in the archimedean
setting and §8.2 in the nonarchimedean setting. There is a slight difference
between our conventions in the global setting and in the local setting in terms
of the spaces on which I(σ, λ) acts. In the global setting we are normalizing
things so that the space on which I(σ, λ) acts is independent of λ. In the local
setting we incorporated λ into the definition of the functions themselves. The
two definitions only differ by the character e−hHP (xv ),λ+ρP i at a place v.
P = MN and P 0 = M 0 N 0
198 10 Eisenstein Series
M (w, λ) : IndG G
P (V ) −→ IndP 0 (V )
defined by
Z
−1
M (w, λ)ϕ(x) := e−1 nx)ehHP (we
ϕ(w nx),λ+ρP i hHP 0 (x),−wλ+ρP 0 i
e dn, (10.2)
IndG 0 G
P (V ) ≤ IndP (V )
that consists of K-finite functions in the function field case and automorphic
forms in the sense of Definition 6.5 in the number field case.
When restricted to IndG 0
P (V ) the integral M (w, λ) converges for λ suffi-
ciently large in a sense we now make precise. Let Φ denote the set of roots
of ZM in G and let ∆P be a base for the set of positive roots defined by P
(see §1.9). Set
The subspace iaP ⊆ a∗P C plays a distinguished role, because the represen-
tations I(σ, λ) are unitary for λ ∈ iaP (see Exercise 10.5).
we let
X
E(x, ϕ, λ) := ϕ(δx)ehHP (δx),λ+ρP i . (10.3)
δ∈P (F )\G(F )
(see Exercise 10.4) provided that the series E(x, ϕ, λ) converges absolutely.
The problem is that E(x, ϕ, λ) only converges for λ “sufficiently large”
[Art05, Lemma 7.1]:
Proposition 10.3.1 (Langlands) If Re(λ) ∈ ρP + (a∗P C )+ then E(x, ϕ, λ)
converges absolutely. t
u
On the other hand the representations I(σ, λ) are unitary when λ ∈ iaP ,
which is never in the plane of absolute convergence of the Eisenstein series.
Instead, one has to analytically continue the Eisenstein series as a complex an-
alytic function of λ to the line λ ∈ iaM in order to construct pieces of L2 ([G]).
In the case G = GL2 this was accomplished by Selberg [Sel56, Sel63]. Though
important, it offered only limited insight into the difficulties presented by the
general case.
Langlands was the first to appreciate the difficulties presented by the gen-
eral case, and was able to overcome them [Lan76]. It is hard to overestimate
the impact of Langlands’ work on Eisenstein series to automorphic represen-
tation theory (and indeed much of mathematics, especially number theory).
This work, for example, led to the discovery of Langlands reciprocity as enun-
ciated in Langlands letter to Weil [Lan, Lan70], led to Langlands-Shahidi
theory of automorphic L-functions [Lan71, Sha10], and also is required input
into any general treatment of a trace formula, either in the Arthur-Selberg
sense [Art78, Art80], or in Jacquet’s sense [Jac05a, JLR99].
200 10 Eisenstein Series
for w ∈ W (aP , aP 0 ). t
u
Sometimes in applications one needs more precise information about the
size of Eisenstein series and the intertwining operators. For an example we
refer to [GL06] for an important example where this is crutial. This paper
refers to the equally important paper [M0̈0], which proves that Eisenstein
series the Eisenstein series E(x, ϕ, λ) is a quotient of functions of finite order
under suitable assumptions on ϕ.
Assume for this section that F is a number field. It turns out that there is a
finite direct sum decomposition
where the direct sum is over all association classes P of parabolic subgroups.
Consider the Hilbert space of families F of measurable functions
FP : iaP −→ IndG 2
P (Ldisc ([M ]))
In this section we collect some local preliminaries for our discussion of isobaric
representations in §10.6. Thus for this section we assume that F is a local
field.
We have already discussed in §4.9 and §8.4 how to classify admissible rep-
resentations of G(F ) for reductive F -groups G in terms of tempered repre-
sentations. When G = GLn and F is nonarchimedian, we have also discussed
how this can be refined to a classification of the tempered representations
in terms of the square-integrable representations. We now explain a useful
refinement of these results. Pk
Assume G = GLn . Let n = i=1 ni be a partition of n and for each i let
πi be an irreducible essentially square-integrable representation of GLni (F ).
Let
where
202 10 Eisenstein Series
(
R>0 Ini if F is archimedean and
AGLni =
$Z Ini if F is nonarchimedean.
for a ∈ AGLni .
After changing the order of the partition we can assume that
Re(λ1 ) ≥ · · · ≥ Re(λk ).
! 0
k k
Y
∼
Y 0
J πi1 , λ =J π 1 i , λ0
i=1 i=1
The symbol behaves something like a formal direct sum on the category
of admissible representations, but we emphasize that ki=1 πi is always irre-
ducible.
It is useful to explicitly point out the relationship between the classification
of unramified representations in terms of the Satake correspondence and the
isobaric sum. For this let Tn ≤ GLn be the maximal torus of diagonal matrices
and let Bn ≤ GLn be the Borel subgroup of upper triangular matrices.
Theorem 10.5.3 Let χ1 , . . . χn : F × → C× be n unramified quasicharacters
and let
χ : Tn (F ) −→ C×
n
Y
(tij ) 7−→ χi (tii ).
i=1
Then the isobaric sum ni=1 χi is the unique unramified subquotient of the
GLn
induced representation IndB n
(χ).
Assume now that F is a number field. So far we have explained how the entire
spectrum of L2 ([G]) can be described in terms of automorphic representations
of Levi subgroups of parabolic subgroups of G. We again restrict to the case
G = GLn and discuss refinements of this result.
As a first step in this section we discuss an operation on automorphic
representations of GLn that plays the role in automorphic representation
theory of the direct sum in ordinary representation theory. In fact, under the
204 10 Eisenstein Series
and for each i let σi is a cuspidal automorphic representation of AGLni \GLni (AF ).
Let P be a standard parabolic subgroup of GLn (i.e. a parabolic subgroup
containing the Borel subgroup of upper triangular matrices) with Levi sub-
Qk Qk
group isomorphic to i=1 GLni . Let σ = i=1 σi . We form the induced
representation
I(σ, λ) (10.11)
as in (10.1). For the proof of the following theorem see the appendix to [BJ79]
(see also [Lan79a]):
Theorem 10.6.1 (Langlands) Any irreducible subquotient of I(σ, λ) is an
automorphic representation, and every automorphic representation is of this
form. t
u
This theorem leaves open the question of whether the automorphic rep-
resentation π of GLn (AF ) can arise in two different manners. In fact, it can
arise in essentially only one manner. More precisely, assume that
0
k
X k
X
ni = n0j = n
i=1 j=1
and σi (resp. σj0 ) are cuspidal automorphic representations of AGLni \GLni (AF )
(resp. AGLn0 \GLn0j (AF )) for 1 ≤ i ≤ k and 1 ≤ j ≤ k 0 . We let
j
0
k
Y k
Y
0
σ= σi and σ = σj0 .
i=1 j=1
with the local factors J(σv , λ) defined as in Theorem 4.9.1 and Theorem 8.4.1.
An important open conjecture is that cuspidal representations always have
tempered local factors:
Conjecture 10.6.4 (The Ramanujan conjecture) If π is a cuspidal au-
tomorphic representation of AGLn \GLn (AF ) then πv is tempered for all v.
A discussion of this conjecture and what is known towards it is contained in
[Sar05].
Since the Ramanujan conjecture remains unproven to find a canonical
irreducible subquotient of I(σ, λ) we must proceed differently. The first step
is the following refinement of Theorem 10.6.1:
Theorem 10.6.5 Assume
Re(λ1 ) ≥ · · · ≥ Re(λk ).
The induced representation I(σ, λ) has a unique irreducible quotient J (σ, λ).
If
J (σ, λ) ∼
= J (σ 0 , λ0 )
with Re(λ1 ) ≥ · · · ≥ Re(λk ) and Re(λ01 ) ≥ · · · ≥ Re(λ0k0 ) then k = k 0
and there is a permutation τ of {1, . . . , k} such that σi0 ∼
= στ (i) and hence
λ0i = λτ (i) .
Proof. The uniqueness of the irreducible quotient follows from the corre-
sponding local uniqueness assertion in Theorem 10.5.1 provided that these
quotients can be taken to be spherical outside of finitely many places. This is
implied by Theorem 10.5.3 The second assertion of the theorem then follows
from Theorem 10.6.2. t
u
206 10 Eisenstein Series
This is the global analogue of the local result Theorem 10.5.1. To make
the analogy clearer, it is convenient to adjust our notation slightly. Let πi be
cuspidal automorphic representations of GLni (AF ) for 1 ≤ i ≤ k. Then there
is a unique λ ∈ a∗M such that
⊗ki=1 πi ∼
= σ ⊗ ehHM (·),λi
π1 · · · πk (10.12)
ki=1 πi ∼
Y
= ki=1 πiv
v
L2 (π ⊗m ) ≤ L2 ([M ])
E(x, ϕ, λ).
Here L(s, (σ, m)v ) is the L-function of (σ, m)v (see §11.8). Jacquet then con-
jectured the following theorem, which was proven in [MgW89]:
Theorem 10.7.1 (Moeglin and Waldspurger) Any irreducible subrepre-
sentation of L2 ([GLn ]) is isomorphic to a Speh representation (σ, m) for a
unique factorization md = n and a unique cuspidal automorphic representa-
tion σ of AGLd \GLd (AF ). t
u
208 10 Eisenstein Series
Thus the theorem gives a complete description of L2disc ([GLn ]). This the-
orem forms the basis of much of what we know about the discrete spectrum
for other groups. For example, it together with the Jacquet-Langlands cor-
respondence is used to describe the discrete spectrum of inner forms of GLn
in [Bad08]. As a consequence of the theory of twisted endoscopy, Arthur
[Art13] and Mok [Mok15] have given a description of the discrete spectrum
of quasi-split classical groups in terms of the parametrization of Moeglin and
Waldspurger. This is explained in §13.8 below.
Exercises
10.1. Let G = GL2 and let P be the Borel subgroup of upper triangular
matrices with Levi subgroup M , the diagonal matrices. Let
χ : [M ] −→ C×
( a1 a2 ) 7−→ χ1 (a1 )χ2 (a2 )
In particular show that this series converges absolutely for Re(λ) large
enough.
10.6. Prove theorems 10.5.1 and 10.5.2 using the results on the Bernstein-
Zelevinsky classification stated in §8.4.
For the remaining problems, let triv denote the trivial representation of
GL1 (AF ). Let
π := I(triv, ( 12 , − 12 ));
it is a representation of GL2 (AF ).
10.8. Prove that πv has a composition series of length 2 (the unique irre-
ducible subrepresentation σv is known as the Steinberg representation).
10.9. Prove that the irreducible subquotients of π are in bijection with finite
Q of places
sets of F : the set S of places corresponds to the representation
S
v|S σv ⊗ π
Chapter 11
Rankin-Selberg L-functions
L(s, π × π 0 ).
211
212 11 Rankin-Selberg L-functions
Its image is a unipotent subgroup whose Lie algebra is spanned by {Xα }α∈O(β) .
This unipotent subgroup is in fact stable under Gal(E/F ), and hence is the
E points of a unipotent subgroup N (β) ≤ N that is isomorphic to ResE/F Ga .
Varying β, we see that we have a subgroup
11.3 Generic representations 213
Y
N (β) ≤ N,
β
where the product is over all Gal(E/F )-orbits of simple roots α. We note
that inclusion induces an isomorphism
Y
der
N (β)−→N/N
˜ .
β
Nn (F ) −→ C×
1 x12 ∗ ∗ ∗
1 x23 ∗ ∗
.. ..
(11.1)
7−→ ψ(m1 x12 + · · · + mn−1 x(n−1)n )
. . ∗
1 x(n−1)n
1
λ(π(n)ϕ) = ψ(n)λ(ϕ)
for ϕ ∈ V and n ∈ N (F ).
In this nonarchimedean setting, continuous means locally constant.
214 11 Rankin-Selberg L-functions
λ(π(n)ϕ) = ψ(n)λ(ϕ)
Vsm −→ W(ψ)
W(π, ψ) (11.3)
be the image of any intertwining map Vsm → W(ψ). We write the map
explicitly as
Vsm −→ W(π, ψ)
ϕ 7−→ Wψϕ .
Wψϕ (g) 6= 0
for some ϕ ∈ V and g ∈ G(AF )1 . We remark that a priori this notion depends
on the realization of π in L2cusp ([G]) if it does not occur with multiplicity 1.
It is not hard to check that if π is globally ψ-generic and π ∼
= ⊗0v πv then each
πv is ψv -generic (Exercise 11.3).
The notion of a generic representation is only useful if we have an interest-
ing supply of generic representations. The following theorem implies that all
cuspidal automorphic representations of GLn (AF ) are generic, and moreover
they admit an expansion in terms of Whittaker functionals:
Theorem 11.3.3 Let ϕ ∈ L2cusp ([GLn ]) be a smooth vector in the space of a
cuspidal automorphic representation π of AGLn \GLn (AF ). If ψ : Nn (AF ) →
C× is a generic character then one has
X
ϕ(g) = Wψϕ (( γ 1 ) g) . (11.5)
γ∈Nn−1 (F )\GLn−1 (F )
t
u
The expression for ϕ in (11.5) is called its Whittaker expansion. It is a
generalization (but not the only generalization) of the well-known Fourier
expansion of a modular form.
We now prove this theorem in the special case n = 2. Consider the function
216 11 Rankin-Selberg L-functions
x 7→ ϕ (( 1 x1 ) g) . (11.6)
This is a continuous function on the compact abelian group F \AF , and hence
admits a Fourier expansion. If we fix a nontrivial character ψ : F \AF → C× ,
then all other characters are of the form ψα (x) := ψ(αx) for α ∈ F (see
Lemma B.1.2). Thus the Fourier expansion of the function (11.6) is
XZ
ϕ (( 1 x1 ) g) ψ(αyx)dy. (11.7)
α∈F F \AF
We note that Theorem 11.3.4 is false for essentially every reductive group
that is not a general linear group.
Proof. Let (π1 , V1 ), (π2 , V2 ) be two realizations of a given cuspidal automor-
phic representation (π, V ). Choose equivariant maps Li : V → Vi . We then
obtain Whittaker functionals
λi : Vsm −→ C
L (ϕ)
ϕ 7−→ Wψ i (In ),
= L2 (ϕ)(g).
This implies in particular that V1 and V2 have nonzero intersection, and hence
are equal. t
u
˜ ⊗0v Vv −→ ⊗v W(πv , ψv ).
Vsm −→ (11.9)
Let ϕ ∈ Vsm be a pure tensor, i.e. a vector that maps to a vector of the form
⊗0v ϕv under the first isomorphism. Then, by local uniqueness of Whittaker
models upon normalizing the composite isomorphism Vsm by multiplying by
a suitable nonzero complex number we have
Y ϕ
Wψϕ (g) = Wψvv (g). (11.10)
v
Thus to compute the global Whittaker function Wψϕ (g) it suffices to com-
pute the local Whittaker functions Wψϕvv . When πv and ψv are unramified
this can be accomplished using famous Casselman-Shalika formula [CS80].
For the remainder of this section let F be a nonarchimedean local field and
let ψ : F → C× be a nontrivial unramified character. The Casselman-Shalika
formula is valid for any unramified reductive group over F , but to simplify
our discussion we will assume that G is a split group. Let B ≤ G be the
Borel subgroup with split maximal torus T ≤ B. We let K ≤ G(F ) be a
hyperspecial subgroup in good position with respect to (B, T ). We note that
by the Iwasawa decomposition one has
a
G(F ) = N (F )µ($)K (11.11)
µ∈X∗ (T )
W (µ($)) 1/2
= δB (µ($))χµ (e−λ ).
W (I)
k1 ≥ · · · ≥ kn .
The associated irreducible representation of GLn is Sk1 ,...,kn (Vst ), where Vst
is the standard representation Pn and Sk1 ,...,kn is the Schur functor attached to
the partition k1 , . . . , kn of i=1 kn . Let χk1 ,...,kn be its character.
Corollary 11.4.2 (Shintani) If G = GLn and W ∈ W(J(λ), ψ)GLn (OF ) is
the unique vector satisfying W (In ) = 1 then
! (
$ k1 1/2
.
δBn (µ($))χk1 ,...,kn (e−λ ) if k1 ≥ · · · ≥ kn
W .. =
$ kn
0 otherwise.
t
u
If m < n let
Z
Ψ (s; W, W 0 ) := h
W 0 (h)| det(h)|s−(n−m)/2 dh,
W In−m
Z Z h (11.12)
Ψe(s; W, W 0 ) := W x In−m−1 0
dxW (h)| det h| s−(n−m)/2
dh,
1
where the top integral is over Nm (F )\GLm (F ), the outer integral on the
bottom is over Nm (F )\GLm (F ), and the inner integral on the bottom is over
(n − m − 1) × m matrices with entries in F .
If m = n, then for each Φ ∈ S(F n ) (the Schwartz space) let
Z
Ψ (s; W, W 0 , Φ) := W (g)W 0 (g)Φ(en g)| det g|s dg. (11.13)
Nn (F )\GLn (F )
Here en ∈ F n is the elementary vector with 0’s in the first n − 1 entries and
1 in the last entry.
Proposition 11.5.1 Assume that F is nonarchimedean.
(a) The integrals (11.12) and (11.13) converge for Re(s) sufficiently large.
For π and π 0 unitary they converge absolutely for Re(s) ≥ 1. For π and π 0
tempered, they converge absolutely for Re(s) > 0.
(b) Each integral is a rational function of q −s .
(c) The C-linear span of the integrals in C[q s , q −s ] as W, W 0 (and when m = n
the Schwartz function Φ) is a principal ideal I(π, π 0 ). t
u
In this nonarchimedean case, one proves that there is a unique polynomial
Pπ,π0 ∈ C[x] satisfying Pπ,π0 (0) = 1 such that Pπ,π0 (q −s )−1 is a generator of
I(π, π 0 ). One sets
In the archimedean case one actually defines L(s, π × π 0 ) using the local
Langlands correspondence, which was established very early on in the theory
by Langlands himself. We will explain this in more detail in §12.3 below.
Right now we note the following important bound:
Theorem 11.5.2 If π and π 0 are unitary then L(s, π × π 0 ) is holomorphic
and nonzero for Re(s) ≥ 1. t
u
See [BR94a, §2] for a proof. Technically speaking they only state that the
L-function is holomorphic at this point, but the argument actually proves
that it is nonzero as well.
220 11 Rankin-Selberg L-functions
and for functions f on GLn (F ) or GLn (AF ) let ρ(wm,n )f (x) := f (xwm,n )
and fe(g) := f (wn g −t ).
Theorem 11.5.4 There is a meromorphic function γ(s, π × π 0 , ψ) ∈ C(q −s )
in the nonarchimedean case such that if m < n we have
and if m = n we have
Ψ (1 − s, W f 0 , Φ)
f, W b = ω 0 (−1)n−1 γ(s, π × π 0 , ψ)Ψ (s; W, W 0 , Φ)
for all W ∈ W(π, ψ), W 0 ∈ W(π 0 , ψ). Here ω 0 is the central quasi-character
of π 0 .
We warn the reader that the functional equations in [Jac09] are different. The
root of this is that the definition of the contragredient W
f of W in loc. cit. is
different.
We need one other factor, the local ε-factor:
γ(s, π × π 0 , ψ)L(s, π × π 0 )
ε(s, π × π 0 , ψ) := . (11.15)
L(1 − s, π ∨ × π 0∨ )
If m = n one has
0 −1
Ψ (s, W, W 0 , 1OFn ) = det In2 − q −s e−λ ⊗ e−λ .
Proof. For each m ∈ Z≥1 let T + (m) be the set of the tuples µ = (k1 , · · · , km ) ∈
Zm with k1 ≥ · · · ≥ km ≥ 0. This can be identified with a subset of the dom-
inant weights of Tbm or a cocharacter of Tm . If n > m let
222 11 Rankin-Selberg L-functions
T + (m) ,→ T + (n)
Ψ (s; W, W 0 )
X µ($) −1
= W W 0 (µ($))| det(µ($))|s−(n−m)/2 δB (µ($))
In−m m
µ∈T + (m)
X 0
= χk1 ,...,km (e−λ )χk1 ,...,km ,0,...,0 (e−λ )q −|µ|s
µ∈T + (m)
0
= det(Imn − q −s e−λ ⊗ e−λ )−1 ,
where we let |µ| = k1 +· · ·+km and, as before, χk1 ,...,km ,0,...,0 denotes the char-
acter associated to Sk1 ,...,km ,0,...,0 . This last identity is known as the Cauchy
identity [Bum13, Chapter 38]. Similarly, when n = m we have
Ψ (s; W, W 0 , 1OFn )
W µ($) W 0 (µ($)) 1OFn (en µ($)) | det(µ($))|s δB
X
−1
= n
(µ($))
µ∈T + (n)
X 0
= χk1 ,...,kn (e−λ )χk1 ,...,kn (e−λ )q −|µ|s
µ∈T + (n)
0
= det(In2 − q −s e−λ ⊗ e−λ )−1 .
t
u
0
The calculation above implies that det(Imn − q −s e−λ ⊗ e−λ ) divides L(s, π ×
π 0 )−1 (as a polynomial in q −s )). More is true (see [JPSS83]):
0
Theorem 11.6.2 One has L(s, π × π 0 ) = det(Imn − q −s e−λ ⊗ e−λ )−1 . t
u
The reason that ψ is not encoded in to the left hand side of this equation is
that the right hand side is in fact independent of the choice of ψ.
The following theorem collects the basic facts about Rankin-Selberg L-
functions:
Theorem 11.7.1 The Rankin-Selberg L-function admits a meromorphic con-
tinuation to the plane, holomorphic except for possible simple poles at s = 0, 1.
There are poles at s = 0, 1 if and only if m = n and π ∼ = π 0∨ . One has a
functional equation
Let ϕ(g)
e := ϕ(g −t ) and ϕe0 (g) := ϕ(g −t ). Taking a change of variables h 7→
−t
h we see that
I(s; ϕ, ϕ0 ) = I(1 − s, ϕ,
e ϕe0 ). (11.19)
This simple change of variables is what ultimately powers the proof of theo-
rem 11.7.1.
Define
224 11 Rankin-Selberg L-functions
Z
0 0
Ψ (s; Wψϕ , Wψϕ ) = Wψϕ ( h 1 ) Wψϕ (h)| det h|s−1/2 dh.
Nn−1 (AF )\GLn−1 (AF )
where the inner sum is over γ ∈ Nn−1 (F )\GLn−1 (F ). Since ϕ0 (h) is automor-
phic on GLn−1 (AF ) and | det(γ)| = 1 for γ ∈ GLn−1 (F ), we may interchange
the order of summation and integration and obtain
Z
0
Wψϕ ( h 1 ) ϕ0 (h)| det h|s−1/2 dh.
I(s; ϕ, ϕ ) =
Nn−1 (F )\GLn−1 (AF )
This integral is absolutely convergent for Re(s) 0 which justifies the inter-
change.
Let us first integrate over [Nn−1 ]. View Nn−1 ,→ Nn as matrices of the
form ( u 1 ). Then for u ∈ Nn−1 (AF ) one has Wψϕ (ug) = ψ(u)Wψϕ (g).
I(s; ϕ, ϕ0 )
Z Z
Wψϕ ( u 1 ) ( h 1 ) ϕ0 (uh)du| det h|s−1/2 dh
=
N (A )\GLn−1 (AF ) [Nn−1 ]
Z n−1 F Z
Wψϕ h 1 ψ(u))ϕ0 (uh)du| det h|s−1/2 dh
=
Nn−1 (AF )\GLn−1 (AF ) [Nn−1 ]
Z
0
Wψϕ h 1 Wψϕ (h)| det h|s−1/2 dh
=
Nn−1 (AF )\GLn−1 (AF )
0
= Ψ (s; Wψϕ , Wψϕ ).
t
u
In the theory above crucial use was made of the fact that cuspidal auto-
morphic representations are globally generic. Indeed, our definition of local
L-functions was given in terms of the Whittaker model.
We now explain how to handle the general case. Assume first that F is
a local field. From §10.6 we know that every pair of admissible irreducible
representation π of GLn (F ) and π 0 of GLm (F ) can be written as an isobaric
sum 0
π∼= ki=1 πi and π 0 ∼
= kj=1 πi0
where the πi and πi0 are essentially square integrable and hence generic [JS83,
§1.2].
We then define, for nontrivial characters ψ : F → C× , that
0
k Y
Y k
0
L(s, π × π ) = L(s, πi × πj0 ),
i=1 j=1
0
k Y
Y k
0
γ(s, π × π , ψ) = γ(s, πi × πj0 , ψ),
i=1 j=1
0
k Y
Y k
0
ε(s, π × π , ψ) = ε(s, πi × πj0 , ψ).
i=1 j=1
Of course, one has to check that this is consistent with our earlier definitions.
In other words, we must know that if π and π 0 are generic, then this procedure
gives the same result as if we used our original definition of the local factors.
This can be done.
We then adopt the analogous convention in the global case and obtain
Rankin-Selberg integrals for isobaric automorphic representations. The ana-
lytic properties of these L-functions can be read off from the corresponding
ones for cuspidal representations (see Theorem 11.7.1).
Exercises
11.3. Let G be a split reductive group over a global field F and assume that
π is a cuspidal globally ψ-generic representation of V . If π ∼
= ⊗0v πv prove that
each πv is ψv -generic.
and Proposition 11.5.1 prove that every eigenvalue of eλ lies in (q −1/2 , q 1/2 ).
11.9 Converse theory 227
11.8. Prove Theorem 11.7.1 in the case m = n − 1 using Theorem 11.7.4 and
the local functional equations of Theorem 11.5.4.
Chapter 12
Langlands Functoriality
229
230 12 Langlands Functoriality
Example 12.1. If E is an elliptic curve over Q without CM, then the Tate
module of E gives a representation with image isomorphic to GL2 (Z` ) for
almost all primes ` [Ser72].
χ` : GalQ −→ Z×
`
Definition 12.1. Let F be a local or global field. Then a Weil group for F
is a tuple (WF , φ, {ArtE }) where WF is a topological group,
φ : WF −→ GalF
is a continuous homomorphism with dense image, and, for each finite exten-
sion E/F ,
12.1 The Weil group 231
WE := φ−1 (GalE ),
and
is an isomorphism, where
(
E × if F is local,
CE :=
E × \A×E if F is global.
Here the superscript ab denotes its maximal abelian Hausdorff quotient, that
is, the quotient by the closure of the commutator subgroup. Moreover, these
data are required to satisfy the following assumptions:
(a) For each finite extension E/F the composite
Art induced by φ
E
CE −−−−→ WEab −−−−−−−−→ Galab
E
CE
ArtE
/ W ab
E
σ
CE σ / W abσ
ArtE σ E
ArtE 0
CE 0 / W ab0
E
inclusion transfer
CE / W ab
ArtE E
commutes.
(d) Finally, the map
WF → lim WE/F
←−
is an isomorphism, where WE/F := WF /WE , the bar denoting closure.
Example 12.4. In the global function field case, the Weil group is defined as
in the previous example, but one replaces the residue field above with the
constant field.
For number fields one doesn’t have a nice intrinsic description like in the
above examples.
Almost by definition of the Weil group, one obtains the following corre-
spondence:
Theorem 12.1.2 (The Langlands correspondence for GL1 ) There is a
bijection between isomorphism classes of irreducible automorphic representa-
tions of GL1 (AF ) and continuous representations χ : WF → GL1 (C).
For better or worse the Weil group is still not big enough. The correct en-
largement of WF when F is a number field should be the as yet hypothetical
Langlands group. On the other hand, the correct enlargement of WF when F
is a local field is known, so in this section we restrict to this case. We follow
the treatment in [GR10].
The Weil-Deligne group of a local field F is
(
WF if F is archimedean,
WF0 := (12.2)
WF × SL2 (C) if F is nonarchimedean.
that is just φ in the archimedean case and it the composite of the natural
quotient map WF0 → WF with φ in the nonarchimedean case.
Let G be a group scheme over C whose neutral component is a reductive
group. If F is archimedean, then a representation of WF0 into G(C) is simply
a homomorphism ρ : WF → G(C).
Assume until stated otherwise that F is nonarchimedean. We now prepare
to define a representation of WF0 . Let k be the residue field of F , k a choice
of algebraic closure, and Galk := Gal(k/k). The action of GalF preserves the
ring of integers and the (unique) prime ideal of the ring of integers of any
finite extension field E/F contained in F . There is thus an exact sequence
1 −→ IF −→ GalF −→ Galk −→ 1
where IF is the inertia subgroup, which can be defined as the kernel of the
map GalF → Galk .
The Weil group WF is a subgroup of GalF and contains IF . We thus have
an exact sequence
1 −→ IF −→ WF −→ Galk .
The last map is not surjective. Its image is isomorphic to Z, whereas Galk is
isomorphic to Z,
b the profinite completion of Z. We deduce that
WF ∼
= IF o hFri
ρ : WF0 −→ G(C)
and every representation of WF0 into G(C) b arises in this manner. Thus L-
parameters may be identified with representations into G(C) b in this setting,
and the notion of equivalence of L-parameters corresponds to G(C)-conjugacy
b
of the associated representations.
We now define L-functions, ε-factors, and γ-factors. We start by giving
ourselves a representation ρ : WF0 −→ GL(V ). Recall that the isomorphism
classes of irreducible representations of SL2 are given by the symmetric power
representations Symn of the standard representation. We can therefore de-
compose V as a representation of WF0 = WF × SL2 (C):
∞
M
V = Vn ⊗ Symn .
n=0
Here Vn is a representation of WF (that is zero for all but finitely many n).
We then define
∞
Y −1
L(s, ρ) := det 1 − ρ(Fr)q −(s+n/2) |VnIF . (12.4)
n=0
0 0
in the usual manner then it is a representation of WF0 into GL V ⊕[WF :WE ]
in the sense of Definition 12.2. One checks that when F is nonarchimedean
L(s, ρ) = L s, IndF
E (ρ) . (12.6)
12.2 The Weil-Deligne group and L-parameters 235
χs,m : F × −→ C×
zm (12.8)
z 7−→ |z|s .
(z z̄)m/2
Unless otherwise specified, when we discuss the character χs,m we will always
assume that s ∈ C, that m ∈ Z and if F is real that m ∈ {0, 1}.
Let
( s
π − 2 Γ 2s
if F = R
ΓF (s) := (12.9)
2(2π)−s Γ (s) if F = C.
IndR ∼ R
C (χs,m ) = IndC (χs0 ,m0 ) (12.11)
χ : WF −→ GL1 (C),
ε(s, ρ, ψ)L(1 − s, ρ∨ )
γ(s, ρ, ψ) := . (12.13)
L(s, ρ)
r : L G −→ GL(V )
In the case where G = GLn and the map r : L GLn → GLn (C) is just the
projection to the neutral component r is omitted from notation, i.e.
For this section F is an archimedean local field. Over archimedean local fields
the Langlands correspondence between L-packets of representations of G(F )
and L-parameters into L G was proven by Langlands [Lan89]. We should note
that there is a substantial literature refining his result in various important
ways, but we will not describe it.
Instead we will focus on the G = GLn case and describe the Langlands
correspondence explicitly. Let
K := Kn := {g ∈ GLn (F ) : gg ∗ = In }
g := gn := ResF/R gln .
We recall from §10.5 that any irreducible admissible (g, K)-module π can
be written as an isobaric sum
π∼
= ki=1 πi
where the πi are irreducible and essentially square integrable. It turns out
that the possibilities for essentially square integrable π are very limited. If
G is a semisimple group over R we say that G is equal rank if there is a
maximal torus T ≤ G that is anisotropic. In other words, T ≤ G is maximal
and T (R) is compact.
Harish-Chandra’s fundamental theorem on the existence of discrete series
representations follows:
Theorem 12.3.1 (Harish-Chandra) Let G be a reductive group over R.
There are essentially square-integrable irreducible representations of G(R) if
and only if Gder is equal rank. t
u
We refer to [Wal88, §7.7.1] for the proof.
238 12 Langlands Functoriality
such that
(a) If π ∈ Π(GL1 ) then rec(π) = π ◦ Art−1
F .
(b) If π1 ∈ Π(GLn1 ) and π2 ∈ Π(GLn2 ) then
and
ε(s, π1 × π2 , ψ) = ε(s, rec(π1 ) ⊗ rec(π2 ), ψ).
(c) If π ∈ Π(GLn (F )) and χ ∈ Π(GL1 (F )) then
and
Ψ (G) := {G(C)-conjugacy
b classes of L-parameters into L G} (12.18)
We will state the conjecture for quasi-split groups G in this section. Our
exposition follows that of [Kal16].
The conjectural correspondence is simplest to state in the tempered case
and we will restrict to this setting. We have already defined the notion of a
tempered representation in Definition 4.6. An L-parameter
ρ : WF0 −→ L G
Φt (G) (12.19)
Πt (G)
LL : Πt (G) −→ Φt (G)
b o Fr)Fr−ss (C)
ρ(Fr) ∈ (G
where the disjoint union is over the G(F )-conjugacy classes of hyperspecial
subgroups of G(F ) with K a representative of the class. This is an unrami-
fied L-packet.
Given a tempered L-parameter ρ, the isomorphism classes of representa-
tions in Π(ρ) are by definition a local L-packet. The conjecture, as stated,
leaves open the question of the structure of the packet Π(ρ). To describe a
little of what is expected, let
Sρ := CG(C)
b (ρ(WF0 )) (12.20)
GalF
S ρ := Sρ /ZG(C)
b . (12.21)
Here we are using the fact that Gb comes equipped with an action of GalF ,
the same action used to define L G.
Lemma 12.5.2 For G = GLn the group S ρ is connected.
ι : Π(ρ) −→ Irr(π0 (S ρ ))
Irr(π0 (S ρ )) can be normalized so that the generic element of the packet (which
is unique by Conjecture 12.5.4) is sent to the trivial representation. We call
this normalized injection ιw . Thus we have a pairing
h , i : Π(ρ) × π0 S ρ −→ C
(π, s) 7−→ tr(ιw (π)(s)).
Then for each place v of F there is a GalF -conjugacy class of injective ho-
momorphisms GalFv −→ GalF . Choosing such a homomorphism one obtains
an injection
L
GFv := G(C)
b o GalFv −→ G(C)
b o GalF =: L G.
12.6 Global Langlands functoriality 245
j : Π(ρ) −→ Π(ρ0 )
where the subscript w denotes the weak L-packets of (7.22). The vertical
arrows above are injective, but the horizontal arrows are not in general.
We now return to the general situation; we do not assume anything about
G other than it is reductive over a global field F . Let H be another reductive
group over F and let
r : L H −→ L G
be an L-map. We then obtain a commutative diagram
L r L
HFv −−−−→ GFv
.
y y
L r L
H −−−−→ G
246 12 Langlands Functoriality
r ◦ ρ = (r ◦ ρv ) ∈ Φ(GAF ). (12.24)
Assume for the remainder of this section that G and H are quasi-split
and that conjectures 12.5.1, 12.5.3 and 12.5.4 are known for GFv and HFv
for all places v. This is a reasonable working assumption, as the local Lang-
lands correspondence is known in many cases. The remainder of this section
is devoted to using the local Langlands correspondence to state the global
Langlands functoriality conjecture. We begin with the simplest case. Assume
that G = ResE/F GLn for some field extension E/F of finite degree. In this
case local L-packets are singletons (see §12.5). Thus global L-packets are
singletons in this setting and the following conjecture is natural:
Conjecture 12.6.1 Suppose that G = ResE/F GLn . Let π be an everywhere
tempered cuspidal automorphic representation of H(AF ) with L-parameter
ρ ∈ Φt (HAF ). The unique admissible representation in Π(r ◦ ρ) is automor-
phic.
We invite the reader to compare this with the weaker Conjecture 7.7.1. At
least in the cases where the local Langlands conjecture is known for H Con-
jecture 12.6.1 does not rely on a chain of conjectures for its statement.
For general (even non-quasi-split) H and G the statement of the weaker
Conjecture 7.7.1 doesn’t even require any unknown conjecture in its state-
ment. However, Conjecture 7.7.1 leaves much to be desired. There could be
many L-packets in a weak L-packet, and it does not explain which elements
of an L-packet are automorphic.
Giving a more specific conjecture seems to require the introduction of the
conjectural Langlands group LF . This first appeared in [Lan79a] and is
also discussed in [Art02], where a conjectural construction of the group is
given. The theory of Tannakian categories also indicates that such a group
ought to exist if we assume Langlands functoriality. For a discussion of this
point of view and the obstacles one runs into if one pursues it we refer to
[Ram94].
The Langlands group LF of a global field F is supposed to be an extension
of the Weil group WF by a locally compact group. For reductive F -groups
G one should be able to define global L-parameters LF → L G as in the
local case; they ought to be continuous homomorphisms commuting with
the projection to GalF along the maps LF → WF → GalF . Two global L-
parameters are equivalent if they are G(C)-conjugate.
b For every place v one
12.6 Global Langlands functoriality 247
WF0 v −→ LF ,
Φt (G)
GalF
S ρ := Sρ /ZG
b .
h , iv : Π(r ◦ ρv ) × π0 (S ρv ) −→ C.
Let π 0 be the unique element of Π(r ◦ ρ) such that πv0 is generic for all v.
Assume that π 0 occurs in L2disc ([G]), the largest closed subspace on which the
representation of G(AF ) decomposes discretely. We assume moreover that for
every v there is a homomorphism π0 (S r◦ρ ) −→ π0 (S r◦ρv ). Finally a global
L-parameter
ρ : LF −→ L G
is said to be discrete generic if it is semisimple, its projection to G(C)
b is
bounded and its image is not contained in a proper parabolic subgroup of
L
G.
where ρ runs over the equivalence classes of discrete generic global parameters
satisfying πv ∈ Π(r ◦ ρv ) for all places v.
A few comments are in order. Theoretically, one should be able to re-
duce whatever one wants to know about automorphic representations to the
tempered case. The A-packets of Arthur and his conjectures regarding them
were introduced to make this precise [Art89, Art90] (see also [Clo07] and
[Sha11] where the relationship between A-packets and Conjecture 10.6.4, the
Ramanujan Conjecture, is discussed). However, even when we expect rep-
resentations to be tempered, in most cases we can only prove bounds that
establish that they are close to being tempered. Moreover, the only means
currently known to prove that they are indeed tempered seems to be using
Langlands functoriality (compare [Lan70, Sar05]). To overcome this prob-
lem, as in the local case, instead of just dealing with tempered L-parameters
and representations one enlarges the set of parameters involved in the local
Langlands correspondence to a set of almost tempered parameters and rep-
resentations. This is supposed to be a set that is restrictive enough that the
local Langlands correspondence is still correct, but general enough to include
all representations that can occur as local components of the most tempered
part of the discrete series of L2 ([G]). For examples of such an enlargement
we refer to [Art13].
12.7 Langlands L-functions 249
Assume that we are given a quasi-split reductive G over a local field F such
that the local Langlands correspondence is known for G(F ). Given a repre-
sentation
r : L G −→ GL(V )
an irreducible tempered representation π of G(F ), and a nontrivial character
ψ : F → C× one can then use (12.14) to define the local Langlands L-
function, ε-factor, and γ-factor as follows:
r : L G −→ GL(V ).
GalF −→ L G
commuting with the projection to GalF should give rise to L-packets of au-
tomorphic representations of G(F ). It is natural to try and describe classes
of L-packets that are of this form.
In [Clo90b] isolated a class of automorphic representations on G = GLn
that he called “algebraic” and then conjectured that they are associated to
particular types of Galois representations (not just representations of LF ).
Buzzard and Gee [BG14] later extended Clozel’s conjectures to the case of ar-
bitrary reductive G. We recall Buzzard and Gee’s conjectures in this section,
following the exposition in [BG14] closely.
Until otherwise specified we work locally at an archimedean place v of F
which we omit from notation, writing F := Fv . Let T ≤ G be a maximal
torus, and let BC > TC be a Borel subgroup of GC . The local Langlands
correspondence is known in the archimedean case [Lan89]. It provides us
with a set-theoretic map
LL : Π(G) −→ Φ(G)
π 7−→ LL(π)
The group
has a diagonal action of the Weyl group W (G, T )(C). The action of GalF on
X ∗ (T ) ⊗ C is W (G, T )(C)-semilinear, i.e. for
one has ξ(wx) = ξ(w)ξ(x). The group (12.27) also has an action of GalF de-
fined using both the action of GalF on X ∗ (T ) and the action on HomR (F , C).
Explicitly, if we think of an element of (12.27) as a morphism
φ : HomR (C, C) −→ X ∗ (T ) ⊗ C
(tC )∨ = X∗ (T )C = X ∗ (Tb)C
ι : G(C)
b −→ G(Q
b p)
G(Q
b p ) o GalF (12.28)
ι : L G −→ G(Q
b p ) o GalF .
12.8 Algebraic representations 253
This construction is used in the following conjecture due to Buzzard and Gee.
It is a generalization of a conjecture of Clozel in the case G = GLn .
Conjecture 12.8.2 If π is L-algebraic, then there is a finite set S of places
of F containing all infinite places, all places dividing p, and all ramified
places, and a continuous Galois representation
such that the composite of ρπ and the projection to GalF is the identity and
if v 6∈ S then the Frobenius semisimplification of ρπ |WFv is G(Q
b p )-conjugate
to ι(LL(πv )).
There is additional important information about π at the infinite places and
the places dividing p predicted in the full version of Conjecture 12.8.2 given
in [BG14] but we omit a discussion.
Though in general Conjecture 12.8.2 is open there are important special
cases that are known.
Theorem 12.8.3 Suppose that F is CM or totally real and that π has the
same infinitisimal character as an irreducible representation of ResF/Q GLn
and in particular is L-algebraic. Then Conjecture 12.8.2 is true. t
u
There are in fact two different proofs of Theorem 12.8.3 available, one due to
Harris, Lan, Taylor and Thorn [HLTT16] and another due to Schölze [Sch15].
Exercises
b →G
12.3. For reductive groups H and G, give an example of a morphism H b
that is not induced by a map H → G as in the previous problem.
12.4. Prove that a representation ρ : WF0 −→ GLn (C) (or more properly a
Frobenius semisimple representation) is completely reducible.
Abstract In the number field case the global Langlands functoriality con-
jecture is wide open. Despite this several important cases have established.
We survey some of the hard-earned progress in this chapter.
13.1 Introduction
255
256 13 Known Cases of Global Langlands Functoriality
(X ∗ (T ), X∗ (T ), ∆, ∆∨ )
be a based root datum for G (see §7.3). It follows from the theory of parabolic
subgroups reviewed in §1.9 that there are subsets of ∆1 ⊆ ∆ and ∆∨ 1 ⊆ ∆
∨
such that
(X ∗ (T ), X∗ (T ), ∆1 , ∆∨
1)
is a based root datum for M . Using this fact one can define an L-map
L
M −→ L G. (13.1)
In the special case where G = GLn the conjugacy classes of Levi Pdsubgroups
of parabolics are indexed by tuples n1 , . . . , nd ∈ Z>0 such that i=1 ni = n
as explained in (1.10). The map (13.1) is given by the identity on the Galois
factor and the natural block diagonal embedding
d
Y
GLni (C) −→ GLn (C)
i=1
x1
(x1 , . . . , xd ) 7−→
..
.
xd
Let G be a quasi-split reductive group over F and let E/F be a finite sepa-
rable extension. Suppose we are given an L-map
r : L G −→ L ResE/F GLn
13.4 The strong Artin conjecture 257
for some n. Assume moreover that we know the local Langlands correspon-
dence for G. Then for every cuspidal automorphic representation π of G(AF )
and every place v of F we can associate an L-parameter
ρ(πv ) : WF0 v −→ L G.
By the local Langlands correspondence for ResE/F GLn (Fv ) (which is known,
see §12.3 and §12.4) associated to r ◦ ρ(πv ) there is a unique irreducible
admissible representation r(πv ) with this L-parameter. We denote by
r(π) := ⊗v r(πv ).
r(πv ) ∼
= πv0
defines a representation
where the L-function at the right is the usual Artin L-function. Artin and
Brauer proved that the L-function on the right admits a meromorphic contin-
uation to the plane, satisfies a functional equation, and has a finite number of
poles [MM97]. Artin conjectured the following refinement of this statement:
Conjecture 13.4.1 (Artin) If r is nontrivial and irreducible then L(s, r)
is holomorphic.
The only irreducible representation that not covered by this conjecture is the
trivial representation. If r is the trivial representation then L(s, r) = ΛF (s),
the completed Dedekind zeta function of F . It is certainly not holomorphic;
in fact it has simple poles at s ∈ {0, 1}. Artin’s conjecture implies that this is
the only irreducible Galois representation whose Artin L-function has a pole.
Thus Langlands functoriality implies the following conjecture, which is
often called the strong Artin conjecture:
Conjecture 13.4.2 (Strong Artin) If r is irreducible, then there exist a
cuspidal representation π of GLn (AF ) such that
πv = r(1)v
where GalF acts via its quotient Gal(E/F ), which in turn acts by permuting
the factors. There is an L-map
where the product is over all places w of E dividing v. Here WE0 w is the
Weil-Deligne group as (12.2). In other words, rE/F corresponds to restric-
tion of Galois representations. The L-map rE/F and the functorial trans-
fers it induces (conjectural or not) are known as base change. It con-
jecturally allows us to relate automorphic representations of GLn (AF ) and
ResE/F GLn (AF ) := GLn (AE ). It is customary to write
πE := rE/F (π).
where the first map is the quotient and the second is the regular representa-
tion. We define an L-map
Here rst is the standard representation rst : L GLn −→ GLn (C) and we asso-
ciate to ρ the representation L GL1 −→ GLd(ρ) (C) as in (13.2).
If E/F is a prime degree cyclic extension, then base change and auto-
morphic induction are known. The case n = 2 is due to Langlands [Lan80],
following a crucial special case discovered by Saito [Sai75] and rephrased
representation-theoretically by Shintani [Shi79]. Besides placing the theory
in the correct level of generality, Langlands was able to use it and converse
theory to establish the strong Artin conjecture for 2-dimensional Galois rep-
resentations in many solvable cases (see Theorem 13.4.5).
For GLn the result was established by Arthur and Clozel [AC89]. To state
it, let E/F be a prime degree Galois extension. Let θ be a generator of
Gal(E/F ). It acts on the set of automorphic representations π 0 of GLn (AE )
via
π 0θ (g) := π 0 (θ(g)).
Finally, let NE/F : A× ×
E → AF be the norm map.
be the representation that is the identity on the Galois factor and the sym-
metric kth power representation on the neutral components. By the local
13.6 The Langlands-Shahidi method 263
of GLmn (AF ). Now we know that the standard L-function of this represen-
tation behaves as predicted under the Langlands functoriality conjecture by
Rankin-Selberg theory (see Chapter 11). In general this is very far from the
assertion that π1 π2 is automorphic. Converse theory in the sense of The-
orem 11.9.1 makes the difference precise. For small m and n the difference is
264 13 Known Cases of Global Langlands Functoriality
not as significant and thus using converse theory and the Langlands-Shahidi
method one can prove the following theorem:
Theorem 13.6.2 For 1 ≤ k ≤ 3, let π1 , π2 be cuspidal automorphic repre-
sentations of GL2 (AF ) and GLk (AF ), respectively. Then π1 π2 is automor-
phic. t
u
For the GL2 × GL2 case this was proven in [Ram00a]. For the GL2 × GL3
case this is [KS02, Theorem 5.1]. The paper [Ram00a] was written before the
proof of the local Langlands correspondence. Compatibility of the transfer
with the local Langlands correpondence was checked in [Kim03].
where the bar denotes the action of the nontrivial element of Gal(E/F ).
This is the quasi-split unitary group attached to the extension E/F . The
extension E/F also defines a quasi-split orthogonal group as follows: View E
as a vector space over F equipped with the quadratic form NE/F . Choosing
a basis we obtain a symmetric matrix γE ∈ GL2 (F ) corresponding to the
quadratic form in the usual manner. Then
13.7 Functoriality for the classical groups 265
Jn−1 Jn−1
O∗2n (R) := g ∈ GLn (R) : g γE gt = γE .
Jn−1 Jn−1
We define SOn < On and SO∗n < O∗n to be the subgroups of determinant 1
as usual. We refer to the groups above as the quasi-split classical groups
of rank n defined over R and denote by Gn .
Following [CPSS11] we have the following table of groups and embeddings
of L-groups:
Gn r : L Gn → L HN HN
r : L Gn −→ L HN (13.11)
Then GalF acts on SO2n (C) via its quotient Gal(E/F ), with the nontrivial
element σ of Gal(E/F ) acting by conjugation by w.
b The map
There is a great deal known about Langlands functoriality for this collection
of L-maps. Definitive results were obtained for generic representations in
[CKPSS04, CPSS11]. We focus on this case in the current section. Later
Arthur [Art13] established the same result for arbitrary representations (in
the orthogonal and symplectic cases) under the assumption that the twisted
weighted trace formula can be stabilized (see §13.8 for details). The unitary
case was completed by Mok [Mok15] under this same assumption.
The following theorem is [CKPSS04, Theorem 1.1]. It uses the notation of
the table above. It represents the culmination of an impressive body of work
by Cogdell, Kim, Piatetski-Shapiro and Shahidi.
for all archimedean places v and nonarchimedean places where π, τ and F are
unramified. It is harder to obtain information at the ramified places. For this
purpose one uses stability of γ-factors, which says roughly that certain γ
factors related to those defined in Theorem 11.5.4 depend only on the central
character of πv after taking a suitably ramified by a character of Fv× . At
the end one then applies converse theorems analogous to Theorem 11.9.1 to
deduce the automorphy of r(π).
Assuming Theorem 13.7.1, Ginzburg, Rallis and Soudry [GRS01] had pre-
viously used their descent technique to characterize the image of the functo-
rial transfer attached to r. The characterization is given in terms of poles of
Langlands L-functions. To state it, we define a family of representations r0
and characters χGn : [Gm ]× → C× . The representations r0 listed below are
13.7 Functoriality for the classical groups 267
Gn r0 χGn
SO2n+1 ∧2 1
SO2n , n ≥ 2 Sym2 1
SO∗2n Sym2 ηE/F
Sp2n Sym2 1
U2n AsE/F ⊗ ηE/F 1
U2n+1 AsE/F 1
Here 1 denotes the trivial character, and ηE/F is the character attached to
E/F by class field theory. The only representation that is not self-explanatory
is the Asai representation
LLS (s, πv , r0 )
(13.15)
εLS (s, πv , r0 , ψv )
L(s, πv , r0 )
(13.16)
ε(s, πv , r0 , ψv )
π 0σ (g) := π 0 (g σ )
and let
π 0∗ := (π 0σ )∨ .
This is sometimes known as the conjugate dual. We say that π is conjugate
self-dual if π ∼= π∗ .
For the purposes of stating the following theorem it is convenient to let
π 0∗ := π 0∨ when HN = GLN and let π 0∗ be defined as above otherwise. Let
ωπ0 be the central quasi-character of π 0 (see [GRS01, Sou05] for more details):
the form
π 0 = Ind(π10 ⊗ · · · ⊗ πd0 ) = π10 · · · πd0 ,
where each πi is a unitary cuspidal automorphic representation of HNi (AF )
satisfying
(a) πi0∗ ∼
= πi0
0 ∼ 0
(b) πi = πj if and only if i = j,
(c) LS (s, πi0 , r0 ) has a pole at s = 1 for a sufficiently large finite set of places
S of F including all infinite places.
Moreover, any π 0 as above satisfying ωπ0 |A× = χGn and (a)-(c) is of the
F
form r(π) for some globally generic cuspidal automorphic representation π of
Gn (AF ). t
u
Analogues of 13.7.1 and 13.7.4 are known for spin similitude groups by work
of Asgari and Shahidi [AS06, AS14].
In this section we have concentrated on generic representations. This is be-
cause originally the converse theory approach explained above relied crucially
on this assumption. Recently the genericity assumption has been removed in
work of Cai, Friedberg, Ginzburg, Kaplan. We refer to [CFK18] (which is
based on [CFGK17]) for a precise statement. This provides a new, uncondi-
tional, and independent proof of a portion of the endoscopic classification of
representations on classical groups that we discuss in the next section.
be Cc∞ (Gn (AF )) except in the special case where Gn is SO2n or SO∗2n , in
which case it is the subalgebra of Cc∞ (Gn (AF )) invariant under the outer
automorphism θ.
Let
L2disc ([Gn ]) < L2 ([Gn ])
be the largest closed subspace that decomposes discretely under Cc∞ (Gn (AF ))
(see §3.8 for details). Alternately L2disc ([Gn ]) is the Hilbert space direct sum
of all irreducible subrepresentations of L2 ([Gn ]). In view of Theorem 13.8.1
it is natural to partition L2disc ([Gn ]) into packets consisting of fibers of the
functorial transfer to HN (AF ) and then try to describe the fibers, preferably
in terms of local data. This is precisely what Arthur accomplishes.
We state the main classification result first. The remainder of the section
is devoted to defining the notation that appears.
ec∞ (Gn (AF ))-module isomorphism
Theorem 13.8.2 (Arthur) There is an C
13.8 Endoscopic classification of representations 271
L2disc ([Gn ]) ∼
M M
= π ⊕mψ ,
ψ∈Ψ
e2 (Gn ) π∈Π
fψ (εψ )
where mψ = 1 or 2. t
u
This is [Art13, Theorem 1.5.2]. For a cuspidal automorphic representation
τ of HN (AF ) and m ∈ Z≥1 let (τ, m) be the Speh representation of §10.7.
One says that τ is of orthogonal type if LS (s, τ, Sym2 ) has a pole at s = 1
and of symplectic type if LS (s, τ, ∧2 ) has a pole at s = 1, where S is a finite
set of places containing all infinite places and all places where τ is ramified.
Since
LS (s, τ, Sym2 )LS (s, τ, ∧2 ) = LS (s, τ × τ )
it follows from Theorem 11.7.1 that τ cannot be both orthogonal and sym-
plectic, and it is orthogonal or symplectic if and only if τ ∼
= τ ∨ . We define
the type of the representation (τ, m) is determined as in the table below:
τ m (τ, m)
The set Ψe2 (Gn ) is the set of automorphic representations of HN (AF ) of the
form
di=1 (τi , mi )
where
(1) τi is a cuspidal automorphic representation of HNi (AF )
P d
(2) i=1 Ni mi = N
∨ ∼
(3) τi = τi
(4) τi ∼
= τj if and only if i = j.
(5) If L G◦n is orthogonal (resp. symplectic) then (τi , mi ) is orthogonal (resp. sym-
plectic).
We note that condition (5) actually implies condition (3). The set Ψe2 (Gn ) is
known as the set of discrete global A-packets of Gn (see [Art13, page
33-34 in §1.4]). The multiplicity mψ in Theorem 13.8.2 is defined to be 1
unless N is even, L G◦n = SON (C), and Ni mi is even for all i, in which case it
is 2. The discrete global A-packet is said to be generic if mi = 1 for all i. In
this case we also refer to the packet as a discrete global generic L-packet.
For every ψ ∈ Ψe2 (Gn ) Arthur defines a finite 2-group Sψ and a character
272 13 Known Cases of Global Langlands Functoriality
εψ : Sψ −→ {±1}.
We omit the definition. For every place v of F and every ψ ∈ Ψe2 (Gn ) we
define a representation
by
ψv := ⊕di=1 rec(τiv ) ⊗ Symm
where rec is the local Langlands reciprocity map of (12.15) in §12.4. We note
that ψv is not an L-parameter, but one can obtain an L-parameter from it
as follows: 1/2
ρ(ψv )(g) := ψv g, |g| −1/2 .
|g|
Sψ −→ π0 (S ψv )
where
GalF
S ψv := CGb n (C) (Im(ψv ))/ZG
b (C)n
h·, πv i : π0 (S ψv ) −→ C× .
13.9 The function field case 273
e ψ (εψ ) := {π ∈ Π(ψ)
Π e : h·, πi = εψ }. (13.20)
In the function field case much more is known. Let F be a function field of
characteristic p. For GL2 , the Langlands correspondence for a function field
was proved by Drinfeld [Dd80, Dd87b, Dd87a, Dd88] and for higher rank
GLn , this was proved by L. Lafforgue following Drinfelds’ argument.
In fact, L. Lafforgue, in Fields’ medal winning work [Laf02], proved the
existence of a correspondence between the set of equivalence classes of cus-
pidal automorphic representations of GLn (AF ) and the isomorphism classes
of continuous irreducible `-adic representations
that are unramified almost everywhere. One can view this as establishing
the automorphic to Galois direction of the global Langlands correspondence.
However, for GLn , this implies the both directions and hence establishes the
global Langlands correspondence. More precisely, the other direction can be
deduced from the converse theory of GLn of Weil [Wei67], Piatetski-Shapiro
and Cogdell [CPS94, CPS99] as well as Grothendieck’s functional equation
and the product formula for ε-factors of Laumon [Lau87].
More recently, V. Lafforgue (L. Lafforgue’s brother) gave a decomposition
of the space of cusp forms on an arbitrary reductive group in terms of Lang-
lands parameters that is compatible with the local Langlands correspondence
at almost all places [Laf18].
All of the work above involves an object known as a shtuka (Russian for
“thing”). The very definition of a shtuka involves the scheme
274 13 Known Cases of Global Langlands Functoriality
This sort of self product also comes up in Deligne’s proof of the Riemann
hypothesis in the function field case. The analogue of this scheme does not
exist in the number field case, since the various residue fields of the number
field all have different characteristics.
There has been a great deal of thought about what can be used to replace
(13.21) in the number field case. The idea is that instead of Fp one should use
the field with one element. This remains a very interesting but mostly specu-
lative prospect. The motivation comes not only from the link with Langlands
functoriality but also from the link to the Riemann hypothesis.
Exercises
rk := L GL2 → GL2(k+1)
14.1 Introduction
275
276 14 Distinction and Period Integrals
HomH(F ) (Vπ , Vχ ) 6= 0.
HomH(F ) (Vπ , Vχ )
Thus the first space is just a space of functions, and the second is a space of
smooth functions. The first space can be regarded as the space of sections on
the line bundle over H(F )\G(F ) defined by χ and the second can be regarded
as the space of smooth sections.
Lemma 14.2.1 There is a C-linear isomorphism
Iλ (ϕ)(g) := λ(π(g)ϕ).
t
u
The lemma allows us to describe natural subspaces of the space of relative
characters HomH(F ) (Vπ , Vχ ). For example, let
the inverse image of C ∞ (H(F )\G(F ), χ) under the bijection of Lemma 14.2.1.
We say that the elements of (14.2) are smooth relative characters. We say
that π is smoothly (H, χ)-distinguished if (14.2) is nonzero. More sophisti-
cated notions of smoothness must sometimes be employed in the archimedean
case (see, for example, §11.3).
It is instructive to observe that one can view the theory of generic repre-
sentations discussed in §11.3 as a special case of the more general concept
of a distinguished representation. Assume for simplicity that F is nonar-
chimedean. If G is quasi-split, N ≤ G is the unipotent radical of a Borel
subgroup, and ψ : N (F ) → C× is a generic character then an irreducible
admissible representation π of G(F ) is ψ-generic if and only if it is smoothly
(N, ψ)-distinguished. The space HomN (F ),sm (Vπ , Vψ ) is the space of Whit-
taker functions, and HomG(F ) (Vπ , C ∞ (N (F )\G(F ), ψ)) is the space of Whit-
taker models. Both are at most one-dimensional. For many interesting cases,
the phenomenon that (14.2) is at most one dimensional (or at least finite
dimensional) persists. We discuss this in more detail in §14.4 below.
The global version of distinction involves period integrals of cusp forms. Let
G be a reductive group over a global field F and let H be a subgroup of G.
Let
χ : H(AF ) −→ C×
be a quasi-character trivial on (AG ∩AH )H(F ). If ϕ : [G] → C is a continuous
function then we define the period integral
Z
Pχ (ϕ) := ϕ(h)χ(h)dh (14.3)
(AG ∩AH )H(F )\H(AF )
f := measdg∞ (K ∞ )−1 f∞ 1K ∞ .
Here we have used the absolute convergence of the period integral to justify
switching the order of integration. The above is bounded by
Z
f (g) |Pχ (π(g)ϕ) − Pχ (ϕ)| dg
AG \G(AF )
is closed. For this it suffices to show that AG \G(F )AG H(AF )/H(AF ) is
closed in AG \G(AF )/H(AF ). We in fact show that the image of G(F ) in
the quotient AG \G(AF )/H(AF ) is discrete in Lemma 17.6.4. t
u
A version of the following proposition is proven in [AGR93]:
Proposition 14.3.3 If f ∈ Cc∞ (AG \G(AF )) and ϕ ∈ L2 ([G]) is smooth
then the integral defining Pχ (ϕ) is absolutely convergent.
Proof. We first reduce the proposition to the case where H is connected. Let
K ∞ ≤ H(A∞ F ) be a compact open subgroup such that ϕ and χ are left K -
∞
◦
invariant. We recall that H(AF )/H (AF ) is compact by a theorem of Borel
[Con12a, Proposition 3.2.1], and hence
(see Exercise 2.16). The set [Hu ] is compact by Theorem 2.6.2. Thus applying
the Fubini-Tonelli theorem, we see that it suffices to establish the proposition
in the special case where H is reductive.
Assume first that we are in the function field case. In this setting Theorem
9.4.1 asserts that any smooth function in L2cusp ([G]) is compactly supported.
Thus the proposition follows from Lemma 14.3.2.
Now assume that we are in the number field case. Let TH0 ≤ H der and
T0 ≤ Gder be maximal F -split tori in H der and Gder , respectively. We can
and do assume T0 ∩ H der = TH0 . Let PH be a minimal parabolic subgroup of
H containing TH0 . It admits a Levi decomposition MH NH = PH , where MH
is the centralizer of TH0 in H (it is a Levi subgroup of P [Bor91, Proposition
20.4] or [BT65, Theorem 4.15(b)] ) and NH < PH is the unipotent radical of
PH . By Theorem 2.7.2 there is an 0 < r < 1 such that
d(anmk) = dadndmdk
for (a, n, m, k) ∈ AH ATH0 (r)×NH (AF )×MH (AF )1 ×KH by the Iwasawa de-
composition (Theorem 2.7.1) and Proposition 3.2.1. Thus the integral Pχ (ϕ)
converges absolutely provided that
Z
|ϕ(anmk)|dadndmdk (14.4)
(AG ∩AH )\AH ATH0 (r)×Ω×KH
converges.
Let ∆ be a set of simple roots for T0 in G. We can then form the corre-
sponding set
AT0 (r0 )
for r0 ∈ R>0 . There is a Weyl chamber C ⊂ Lie(AT0 ) such that the closure
of the image of C under the exponential map is AT0 (1). Weyl chambers in
Lie(AT0 ) are permuted simply transitively by W (G, T )(F ) [Bor91, Theorems
14.7, 21.2, and 21.6]. It follows that
[
AT 0 = wAT0 (1)w−1
w∈W (G,T )(F )
and the intersection of any two W (G, T )(F )-conjugates of AT0 (1) is a set of
measure zero with respect to any Haar measure on AT0 . Hence (14.4) is equal
to
14.4 Spherical subgroups 281
X Z
|ϕ(anmk)|dadndmdk.
w∈W (G,T )(F ) (AG ∩AH )\AH ATH0 (r)∩wAT0 (1)w−1 ×Ω×KH
the expectation of the principles laid out in §14.4, although the dual group
of the relevant spherical variety is not defined here as G is not quasi-split.
Similarly, using a variant of the Rankin-Selberg theory we discussed in
§11.7 Flicker and Zinoviev proved the following theorem [Fli88, FZ95]:
Theorem 14.5.3 For G as in Theorem 14.5.2 an automorphic respresenta-
tion π of AG \G(AF ) is distinguished by GLn if and only if the Asai L-function
L(s, π, AsE/F ) has a pole at s = 1. t
u
Here the Asai L-function is the Langlands L-function attached to the Asai
representation. To define it, let E/F be an arbitrary (not just quadratic)
extension with d := [E : F ]. Then the Asai representation
A representation
ρ : GalE −→ GLn (C)
extends uniquely to a homomorphism
commuting with the projections to GalF on the L-group side. Thus to each
such ρ we can associate the representation
⊗τ ∈HomF (E,F ) ρτ
to GalF . Suppose now that E/F is quadratic and that τ is the gen-
erator of Gal(E/F ). Let π be a cuspidal automorphic representation of
AResE/F GLn \ResE/F GLn (AF ). In this case the analytic properties of the Asai
L-function are understood by Theorem 13.7.2. The discussion above (with
global Galois groups replaced by local Weil-Deligne groups) implies that
µs : H(AF ) −→ C×
det a s−1/2
a det a
7−→
χ η(det b),
b det b det b
where χ and η are characters [Gm ] (in particular they are trivial on AGm ). In
this setting Jacquet and Friedberg [FJ93, Theorem 4.1] prove the following:
Theorem 14.5.5 Let π be a cuspidal automorphic representation of GL2n (AF )
with central character ωπ . Assume that η n ωπ = 1. The cuspidal representa-
tion π on GL2n (AF ) is (H, µs0 )-distinguished if and only if L(s, π, ∧2 ⊗ η)
has a pole at s = 1 and L(s0 , π ⊗ χ) 6= 0. t
u
In fact Friedberg and Jacquet prove that for ϕ in the space of π the pe-
riod integral Pµs0 (ϕ) is a holomorphic multiple of L(s0 , π ⊗ χ) (under the
assumption that L(s, π, ∧2 ⊗ η) has a pole at s = 1). Ash and Ginzburg use
Theorem 14.5.5 to construct p-adic L-functions under a technical hypothesis,
see [AG94].
286 14 Distinction and Period Integrals
r : L G −→ ResM/E GLn .
If the partial Asai L-function LS (s, π 0 , AsM E/F ) has a pole at s = 1 then
some cuspidal automorphic representation π 00 of G(AF ) in the L-packet of π 0
is H-distinguished. Moreover, we can take π 00 to be V -cohomological. t
u
Here S is a sufficiently large finite set of places of M including all infinite
places. The notion of V -cohomological is explained in §15.5 below.
The heart of the proof is based on a particular case of a general principle
which we now explain. Let G be a reductive group, let hτ i = Gal(M/F ), and
let
G
e := ResM/F G.
h , i : V × V −→ E
E GV
E=F OV
E =F ⊕F GLV0
E/F quadratic UV
denote the usual Galois cohomology set. The forms of G are in bijection with
H 1 (F, Aut(G)). The so called inner forms are those forms corresponding
to the subgroup of inner automorphisms. This means that the corresponding
cocycles are in the image of the map
The pure inner forms are those corresponding to cocycles in the image of
the map
H 1 (F, G) −→ H 1 (F, G/Gder ) −→ H 1 (F, Aut(G)).
Two pure inner forms are said to be equivalent if the corresponding cocycles
in H 1 (F, G) are equivalent. It turns out that the pure inner forms of GV are
all of the form GV 0 for free E-modules V 0 with E-rank equal to V . More
specifically, if E and V are fixed the side conditions on V 0 that are equivalent
to GV 0 being a pure inner form of GV are given in the following table (see
[GGP12b, §2]):
290 14 Distinction and Period Integrals
E conditions
E =F ⊕F rankE V 0 = rankE V
E/F quadratic V0 ∼
= V as hermitian spaces
Proof. There is a finite set of places S of F such that GFv and G0Fv are
quasisplit for v 6∈ S [Spr79, Lemma 4.9]. On the other hand two inner forms
of the same quasi-split reductive group are necessarily isomorphic [Mil17,
Corollary 23.53]. t
u
G := GV × GW
and GW are general linear groups this simply means that π is a cuspidal
automorphic representation.
Conjecture 14.7.2 (Gan-Gross-Prasad) One has
L( 12 , π, rV ⊗ rW ) 6= 0
Assume for the moment that F is a global field. One should be aware that for
general reductive groups G and connected subgroups H ≤ G there do not ex-
ist cuspidal automorphic representations of G(AF ) that are H-distinguished.
A somewhat trivial example is given by taking H to be the unipotent radi-
cal of a parabolic subgroup. In this case there are no cuspidal automorphic
representations of G(AF ) that are H-distinguished by the very definition of
a cuspidal representation.
14.8 Necessary conditions for distinction 293
σ : G −→ G
of order 2. Let
H := ((Gσ )◦ )der .
In this setting Prasad [Pra19, Theorem 1] gave a beautifully simple criterion
for there to exist generic representations of G(F ) distinguished by H(F ):
Theorem 14.8.2 If there is an irreducible admissible generic representation
of G(F ) distinguished by H(F ) then there is a Borel subgroup B < GF such
that B ∩ σ(B) is a maximal torus of GF . t
u
One also has the following partial converse [Pra19, Proposition 11]:
Proposition 14.8.3 If there is a Borel subgroup B < G such that B ∩ σ(B)
is a maximal torus of G then there exists an irreducible generic representation
of G(F ) distinguished by Gσ (F ). t
u
In this proposition we are assuming B is a subgroup of G, not GF .
Exercises
14.1. Let G be a reductive group over a local or global field F and let
H, H 0 ≤ G be subgroups that are G(F )-conjugate. If F is local prove that an
294 14 Distinction and Period Integrals
14.3. Assume that G is a reductive group over a global field F and that
H ≤ G is a subgroup. Prove that
AG ∩ H(AF ) = AG ∩ AH .
14.4. Let H be a reductive group over the global field F viewed as a subgroup
of G := H × H via the diagonal embedding. Let π1 and π2 be cuspidal
automorphic representations of AH \H(AF ) and let π = π1 ⊗ π2 be their
exterior tensor product, viewed as a cuspidal automorphic representation of
AG \G(AF ). Show that π1 ⊗ π2 is H-distinguished if and only if π1 ∼ = π2∨ .
diverges.
14.7. Assume the notation in the proof of Proposition 14.3.3. Give an ex-
ample of a pair of reductive subgroups H ≤ G such that the closure of
AH − (AH ∩ AT0 (r0 )) is noncompact for any choice of r0 > 0 and any choice
of simple roots ∆ of T0 in G.
15.1 Introduction
295
296 15 The Cohomology of Locally Symmetric Spaces
metric spaces. This was in fact the original motivation for Harder, Langlands,
and Rapoport’s introduction of the notion of a distinguished automorphic
representation [HLR86]. In this paper they related period integrals of au-
tomorphic representations to algebraic cycles on Hilbert modular varieties.
Using this relation they proved cases of the Tate conjecture for these varieties.
There has been work generalizing their results, but even very basic questions
remain. We point out one such question explicitly in §15.6. We hope that in
the near future the knowledge one now has about the étale cohomology of
Shimura varieties can be combined with the relative trace formula to system-
atically study the Tate and Beilinson-Bloch conjectures for these varieties.
It is beyond the scope of this book to make this completely precise, but we
mention some promising results that have been obtained in §15.8.
Understanding the material in this chapter requires some knowledge of
sheaves and cohomology. In §15.8 we even briefly discuss étale cohomology of
Shimura varieties. The results will not be used in the remainder of the book,
so the reader should feel free to skip this chapter on a first reading. However,
we encourage the reader to at some point investigate the fascinating interplay
between arithmetic geometry and automorphic forms that is the subject of
Langlands’ quote in the epigraph.
AG ≤ G(R)
be the identity component in the real topology of the maximal Q-split torus in
the center of G. To ease notation, we write K for K ∞ ≤ G(A∞ ), a compact-
open subgroup. Finally we set
X := AG \G(R)/K∞ .
Γ \X (15.3)
for a sufficiently divisible integer N . The general case follows from this. t
u
298 15 The Cohomology of Locally Symmetric Spaces
G(Q) ∩ g −1 Kg
In fact, one only has to check this condition for all g ∈ G(Q)\G(A∞ )/K,
which is a finite set. An elaboration of the proof of Lemma 15.2.1 implies the
following (see Exercise 15.4):
Lemma 15.2.2 If K ≤ G(A∞ ), then K contains a neat subgroup of finite
index. t
u
Our motivation for introducing this notion is the following:
Lemma 15.2.3 If K is neat then ShK is a smooth manifold.
In §5.2 we discussed the Hecke algebra Cc∞ (G(A∞ )). We recall that this is
the space of C-linear combinations of characteristic functions of double cosets
KgK for g ∈ G(A∞ ). We now explain how these operators can be realized
geometrically as correspondences. Let K and K 0 be compact open subgroups
of G(A∞ ), and g ∈ G(A∞ ) be such that K 0 ⊂ gKg −1 . Then we have a map
0
Tg : ShK −→ ShK (15.4)
0
G(Q)(x, hK ) 7−→ G(Q)(x, hgK).
y TI Tg %
ShK ShK .
Set
where Z
R(1KgK )ϕ(x) := 1KgK (g)ϕ(xg)dg
G(A∞ )
H 1 (Γ \H, C)
{f (z) : f ∈ S2 (Γ )}
V Γ := Γ \(V × X) (15.8)
by the diagonal action of Γ on the product. We say that the diagram given
by the natural projection
V Γ −→ Γ \X (15.9)
V Γ |U := {s : U −→ V Γ } (15.10)
be the abelian group of sections of the map (15.9). Then the functor
{U ⊆ Γ \X} −→ Ab (15.11)
Γ
U 7−→ V |U
H • (Γ \X, V Γ ). (15.12)
We thus have a natural map V K → ShK . This is again a local system, and
we define the associated sheaf of sections as above.
The relationship between the two constructions given above can be de-
scribed as follows. Fix g ∈ G(A∞ ) and let Γ = gKg −1 ∩ G(Q). We then have
an embedding
ι : Γ \X −→ ShK (15.14)
Γ x 7−→ G(Q)(x, g)K.
We first note that for any neat compact open subgroups K, K 0 ≤ G(A∞ )
with K 0 ≤ gKg −1 one has an isomorphism of sheaves
0 0
˜ g−1 V K = V K ×ShK ShK
θ : V K −→T (15.16)
0 0
(v, (x, hK )) 7−→ ((v, (x, hgK)), (x, hK )).
Here Tg−1 V K is the inverse image sheaf and the object on the right is the fiber
product over the canonical projection V K → ShK and the morphism Tg :
15.3 Local systems 301
0
ShK → ShK . The isomorphism θ is called a lift of the correspondence.
This isomorphism canonically induces a morphism (15.15). Concretely the
procedure is to take a cocycle on ShK with values in V K , pull it back to
0 0
ShK , use the isomorphism θ to produce a cocycle on ShK with values in
Tg−1 V K , and then push it down along Tg while summing over the fibers of
Tg−1 V K → V K to obtain a cocycle on ShK .
Alternately, one can work at the level of sheaves as follows. By generalities
on sheaf cohomology [Har11, (4.33)] one has a homomorphism
0 0 0
H • (ShK , V K ) −→ H • (ShK , TI−1 V K ) = H • (ShK , V K ). (15.17)
0
Here the equality results from the identification TI−1 V K = V K . The isomor-
phism θ induces an isomorphism
0 0 0
θ : H • (ShK , V K )−→H
˜ • (ShK , Tg−1 V K ). (15.18)
tr : Tg∗ Tg−1 F −→ F
V K −→ I •
Here the action of k on ∧q (g/k) is the adjoint action. Explicitly Homk (∧q (g/k), A)
is the subspace of HomC (∧q (g/k), A) consisting of functions f satisfying
q
X
f (x1 , . . . , [x, xi ], . . . , xq ) = x.f (x1 , . . . , xq ),
i=1
on the complex by
q
X
(df )(x0 , . . . , xq ) = (−1)i x · f (x0 , . . . , xbi , . . . , xq )
i=1
X
+ (−1)i+j f ([xi , xj ], x0 , . . . , xbi , . . . , xbj , . . . , xq ).
1≤i<j≤q
Here a hat denotes an omitted argument. One checks that this is well-defined
(i.e. independent of the choice of representative of the coset x + k) and that
d ◦ d = 0. We then let H • (g, k; A) denote the cohomology of this complex:
I = {i1 , . . . , iq } ⊂ {1, . . . , m}
ω I := ω i1 ∧ · · · ∧ ω iq .
Let
A := C ∞ (G(Q)\G(A)/K ∞ ).
Denote by
Aq (ShK , V K )
the space of differential q-forms on ShK with coefficients in V K . Any element
of Aq (ShK , V) can be written as
X
η= fI ω I
I
aG := Lie(AG ). (15.23)
H • (ShK , V K ) ∼
= H • (aG \g, K∞ ; (A ⊗ V )AG ). (15.25)
We leave the proof of this Hecke equivariance assertion to the reader (see Ex-
ercise 15.7).
This whole construction motivates the following definition:
Definition 15.5. A vector ϕ ∈ A is cohomological if there exists a repre-
sentation V of G, v ∈ V and ω I on G(R) such that
ξ : AG → R>0
Then
(A ⊗ V K )AG = A(ξ) ⊗ V K .
As usual [G] := AG G(Q)\G(A). There is a G(A)1 -intertwining map
A(ξ) −→ C ∞ ([G])K
(15.26)
ϕ 7−→ (g 7→ ξ 0 (g)ϕ(g))
A(ξ)cusp
be the inverse image of L2cusp ([G])K ∩ C ∞ ([G])K under the map (15.26).
By Corollary 9.1.2
M
A(ξ)cusp = π ⊕m(π)
π
H • (ShK , V K ) ⊃ Hcusp
•
(ShK , V K ) = H • (aG \g, K∞ ; A(ξ)cusp ⊗ V ) (15.27)
M ⊕m(π)
= H • aG \g, K∞ ; (π ⊗ V )K .
π
•
The group Hcusp (ShK , V K ) is known as the cuspidal cohomology. Its com-
plement is described in terms of so called Eisenstein cohomology; for some
15.6 The relation to distinction 305
•
M m(π)
Hcusp (ShK , V K ) = H • (aG \g, K∞ ; (π∞ ⊗ V ) ⊗ (π ∞ )K . (15.28)
π
•
M m(π0 )
Hcusp (ShK , V K )(π ∞ ) := H • (aG \g, K∞ ; (π∞
0
⊗ V ) ⊗ (π ∞ )K
π 0 :π 0∞ ∼
=π ∞
H • (aG \g, K∞ ; π∞ ⊗ V ) 6= 0.
For the remainder of this section we assume for simplicity that ShK is com-
pact , which is to say that Gder has no nontrivial Q-split subtorus by Theorem
2.6.2. We also assume ShK is orientable.
Let V be a representation of G and let V ∨ be the dual representation.
∞
We denote by VH the local system on Sh(H, XH )K∩H(A ) attached to V |H .
K
Let n be the real dimension of Sh and let q be the real codimension of
∞
Sh(H, XH )K∩H(A ) in ShK . There is a cycle class map
∞
H 0 (Sh(H, XH )K∩H(A )
, VH ) −→ H q (ShK , V K ) (15.29)
∞
defined as follows. For a class [Z] ∈ H 0 (Sh(H, XH )K∩H(A )
, VH ) we have a
linear functional
H n−q (ShK , (V K )∨ ) −→ C
(15.30)
Z
η 7−→ η.
Z
for some smooth ϕ ∈ L2 (π), the π-isotypic subspace of L2 ([G]). The point
of the question is that, a priori, it could happen that no such ϕ satisfies the
additional stipulations of the question. For example, it could be that no such
ϕ is V ∨ -cohomological.
We do not know, even conjecturally, the answer to Question 15.6.1 in any
degree of generality. In special cases the question above can be answered in
the affirmative (see [KS15, Sun17] for example).
In this section we have assumed that ShK is compact. The considerations
above are of course still of interest in the noncompact case, but even defining
the cycle class map (15.29) in any level of generality is complicated. Extending
the cycle class map to the noncompact case is another open problem that
has not received the scrutiny it deserves. We refer to [GG12] for an example
where the difficulties caused by noncompactness of ShK are overcome in a
very special case.
We have seen that the question of whether or not a given automorphic rep-
resentation contributes to the cohomology of a Shimura manifold with coeffi-
cients in a local system is completely determined by the (g, K∞ )-cohomology
of its factor at infinity. One thing that makes this so useful is the fact that
(g, K∞ )-cohomology is a very pleasant object with which to work. In this
section we list some properties of these groups; the canonical reference is
[BW00].
Let G1 , G2 be reductive groups over R and let Ki∞ ≤ Gi (R) be maximal
compact subgroups. Moreover let Ai be admissible (gi , Ki∞ )-modules. Then
one has a Künneth forumla:
308 15 The Cohomology of Locally Symmetric Spaces
Theorem 15.7.1 (Künneth formula) For any n ∈ Z≥0 one has natural
isomorphisms
M
H n (g1 ⊕g2 , K1∞ ×K2∞ ; A1 ⊗A2 ) = H p (g1 , K1∞ ; A1 )⊗H q (g2 , K2∞ ; A2 ).
p+q=n
For the rest of this section let G be a reductive group over Q (we only use
the Q-structure to define AG ). Let g be the Lie algebra of GR and let K∞ ≤
G(R) be a maximal compact subgroup. Let A be an admissible (aG \g, K∞ )-
module.
Theorem 15.7.2 (Poincare duality) One has that
H q (a\g, K∞ ; A) ∼
= H (dimR X)−q (a\g, K∞ ; A∨ )∨ .
t
u
Thus, in particular, to compute the (g, K∞ )-cohomology of a (aG \g, K∞ )-
module it suffices to understand the case where g is simple over R and when
q ≤ dimR X/2.
We now consider the cohomology of admissible representations of the form
π∞ ⊗ V
θ : g −→ g (15.33)
be the Cartan involution. Its fixed point subalgebra is k, the Lie algebra of
K. Let p be its −1 eigenspace.
When π∞ is trivial we have (see [BW00, Corollary II.3.2]) the following
proposition:
t
u
We can interpret this proposition as saying that much of the cohomol-
ogy vanishes when V is nondegenerate. This phenomenon persists for other
π∞ . We now explain some vanishing theorems for cohomology in this con-
text. Combined with the Poincaré duality above these theorems indicate that
cohomology of ShK is concentrated around the degree equal to 12 dimR ShK .
For each θ-stable parabolic subalgebra q ≤ g ⊗ C let n(q) be its nilradical.
Let
15.8 Shimura varieties 309
H i (aG \g, K∞ ; π∞ ⊗ V )
In this section we recall the notion of a Shimura variety. Although the founda-
tions of Shimura varieties were established by Shimura, a simpler formulation
of the theory was introduced by Deligne in [Del71]. Deligne’s formulation has
been almost universally adopted in the literature and we will use it here. An
introduction to [Del71] is contained in [Mil05]. Proofs for the results we state
below can be found in these references.
The Deligne torus is
G\X∗ (G)(C)
Example 15.2. Let G = ResF/Q GL2 where F/Q is totally real. Take
aσ bσ
Y
h= −bσ aσ
σ
G(R) := {x ∈ D ⊗Q R : xx∗ ∈ R× }.
h0 : C −→ D ⊗Q R
h : S −→ GR .
M K := M (G, X)K ,
312 15 The Cohomology of Locally Symmetric Spaces
where M (G, X)K is as in Theorem 15.8.1. We can then consider the étale
cohomology groups
• •
Hét (MEK , Q` ) := Hét (M K ×E E, Q` ). (15.35)
in [BR94b]. In many cases these conjectures have now been proven (see
[Shi11, CHLN11] for instance). This whole picture has an analogue when
M K is not assumed to be compact. However the technical complications in
the noncompact setting are enormous and have only recently been overcome
in key cases [Mor08, Mor10].
The emphasis so far has been on using these results to relate Galois rep-
resentations to automorphic representations. As we mentioned in §12.5, one
expects packets of tempered automorphic representations of G(A) to be pa-
rameterized by L-parameters LF → L G. The work on Shimura varieties de-
scribed above can sometimes allow one to establish this correspondence when
the L-parameter is C-algebraic in the sense of §12.8.
Lest the reader be misled, the description of the étale cohomology group
•
Hét (MEK , Q` ) in terms of automorphic representations is useful for more than
15.8 Shimura varieties 313
The image of this map is contained in cl(CH n/2 (X)Q` ). The map is com-
patible with the cycle class map of (15.29) under the relevant comparison
isomorphisms [DMOS82, §I]. Motivated by the Tate conjecture one is lead to
the following question:
Question 15.8.4 What part of the cohomology of
n
Hét (MEK , Q` (n/2))GalEEH
Exercises
The remaining chapters of this book are devoted to stating and proving simple
versions of trace formulae. These formulae all have a geometric side involving
integrals of a test function along certain orbits, and a spectral side involving
period integrals of automorphic forms.
In any of these formulae the first step in this is to employ a fundamental
idea first applied to the study of automorphic forms by Selberg [Sel56] that
we will now describe. Let G be a connected reductive group over a global
field F and as in §2.6 let
315
316 16 Spectral Sides of Trace Formulae
For
f ∈ Cc∞ (AG \G(AF )),
consider the integral operator, the smooth version of the regular representa-
tion,
One has
Z
R(f )ϕ(x) = f (y)R(y)ϕ(x)dy
AG \G(AF )
Z
= f (x)ϕ(xy)dy
AG \G(AF )
Z
= f (x−1 y)ϕ(y)dy
AG \G(AF )
Z X
= f (x−1 γy)ϕ(y)dy.
[G] γ∈G(F )
Here to justify the last step one uses the unfolding lemma 9.2.4. In other
words, R(f ) is an integral operator with kernel
X
Kf (x, y) := f (x−1 γy). (16.4)
γ∈G(F )
Ω1 × Ω2 ⊂ AG \G(AF ) × AG \G(AF )
for Bπ an orthonormal basis of L2cusp (π), the π-isotypical subspace of L2cusp ([G]).
Here π(f ) is just the restriction of R(f ) to the space of π. We note that π(f )
need not be finite rank in general, but it is if f is K∞ -finite by admissibility
of π (see Exercise 16.1).
We thus arrive at the identity that underlies all trace formulae:
X X
f (x−1 γy) = Kπ(f ) (x, y) + ∗ (16.7)
γ∈G(F ) π
For the moment we leave the geometric expansion of the kernel and focus
on the spectral side. We return to the geometric side in §18.2 when we discuss
trace formulae in simple settings.
(see (9.12)).
Lemma 16.2.1 There is a unique function that is smooth as a function of
(x, y) ∈ (AG \G(AF ))2 that is equal to the kernel Kf,V (x, y) almost every-
where.
Given the lemma, we can and do identify the kernel of R(f ) restricted to a
closed subspace of L2cusp ([G]) with a smooth function.
f1 ∗ f2 ∗ f3
for f1 , f2 , f3 ∈ Cc∞ (AG \G(AF )). It clearly suffices to prove the lemma for f of
this special form, so we assume that f = f1 ∗ f2 ∗ f3 . For f ∈ Cc∞ (AG \G(AF ))
let
We deduce that
X X
R(f )ϕ(x)ϕ(y) = ϕ(x)R(f ∨ )ϕ(y). (16.11)
ϕ∈BV ϕ∈BV
Thus
X
Kf,V (x, y) = R(f1 ∗ f2 ∗ f3 )ϕ(x)ϕ(y)
ϕ∈BV
X
= R(f2 ∗ f3 )ϕ(x)R(f1∨ )ϕ(y)
ϕ∈BV
X
= (R(f2 ) × R(f1∨ )) R(f3 )ϕ(x)ϕ(y). (16.12)
ϕ∈BV
H ≤G×G
The right hand side is rapidly decreasing by Theorem 9.7.1 and hence the
integral in the lemma converges by Proposition 14.3.3. t
u
The main result of this chapter relates the kernel function Kf,cusp (x, y) to
these distributions. It amounts to the computation of the cuspidal contribu-
tion to the spectral side of the relative trace formula:
Theorem 16.2.5 Let f ∈ Cc∞ (AG \G(AF )). One has
Z X
Kf,cusp (h` , hr )χ(h` , hr )d(h` , hr ) = rtrH,χ π(f ).
AG,H H(F )\H(AF ) π
Moreover, the integral on the left and the sum on the right are absolutely
convergent. In particular, if R(f ) has cuspidal image then
Z X
Kf (h` , hr )χ(h` , hr )d(h` , hr ) = rtrH,χ π(f ).
AG,H H(F )\H(AF ) π
After the proof of Theorem 16.2.5 we describe the general spectral expansion
of Kf (x, y) and then, in §16.4, describe how to construct function f such that
R(f ) has cuspidal image.
By Proposition 9.6.1 and with the notation of that proposition this is bounded
by a constant depending on B1 , B2 ∈ R>0 and f1 , f2 times
322 16 Spectral Sides of Trace Formulae
X
max (α(sx ))−B1 (α(sy ))−B2 kKπ(f3 ) (x, y)k2
α1 ,α2 ∈∆
π
X (16.16)
= max (α(sx ))−B1 (α(sy ))−B2 tr π(f3∗ ∗ f3 ).
α1 ,α2 ∈∆
π
Since the operator Rcusp (f3∗ ∗f3 ) is trace class by Theorem 9.7.1 this converges
absolutely. We conclude that the sum
X
Kπ (x, y) (16.17)
π
where the sum on P is over all standard parabolic subgroups of G, the sum
on σ is over all irreducible representations of G(AF ) occurring in the discrete
spectrum of HP , and Bσ is over an orthonormal basis of the σ-isotypic sub-
space of σ in HP . This expression, a priori, only converges in L2 . However,
one can make sense of it pointwise by an analogue of the argument proving
Theorem 16.2.5 (see [Art78, §4]).
versions are integrable over the diagonal and can be given spectral interpre-
tations ([Art05] is a nice introduction).
If we place assumptions on the test function f then the expansion can be
made much simpler. One such simplifying assumption is that the operator
R(f ) has cuspidal image. In this case Kf,cusp (x, y) = Kf (x, y). Lindenstrauss
and Venkatesh [LV07] have defined a large class of functions with purely
cuspidal image that essentially have no kernel when restricted to the cuspidal
spectrum, provided one is only interested in functions spherical at infinity. It
would be interesting to remove this assumption.
Before their work, there was a standard example of such functions that
can be used to good effect in studying local factors of automorphic represen-
tations. We recall it now.
Definition 16.1. Let v be a nonarchimedian place of a global field F . A func-
tion fv ∈ Cc∞ (G(Fv )) is said to be F -supercuspidal or simply supercuspidal
if Z
fv (gnh)dn = 0
N (Fv )
for all proper parabolic subgroups P < G (defined over F ) with unipotent
radical N and all g, h ∈ G(Fv ).
Lemma 16.4.1 If f ∈ Cc∞ (AG \G(AF )) is of the form f = fv f v for some
finite place v and fv is supercuspidal then R(f ) has cuspidal image.
Proof. Let P ≤ G be a proper F -rational parabolic subgroup with unipotent
radical N . As usual let [N ] := N (F )\N (AF ). For ϕ ∈ L2 ([G]), R(f )ϕ is
smooth and hence can be integrated over any compact subset. For all x ∈
AG \G(AF ) we have
Z Z Z
R(f )ϕ(nx)dn = f (g)ϕ(nxg)dgdn
[N ] [N ] AG \G(AF )
Z Z
= f (x−1 n−1 g)ϕ(g)dgdn
[N ] AG \G(AF )
Z Z X
= f (x−1 n−1 δg)ϕ(g)dgdn.
[N ] AG N (F )\G(AF ) δ∈N (F )
= 0,
where the last equality follows from the fact that the inner integral
324 16 Spectral Sides of Trace Formulae
Z
f (x−1 n−1 g)ϕ(g)dn
N (AF )
V −→ V /V (N )
Z !
fv 7−→ g 7→ fv (gnh)dn
N (Fv )
Exercises
16.2. Prove that Kf,V (x, y) is independent of the choice of orthonormal basis.
Despite this show that a K1∞ × K2∞ -finite smooth function in (16.19) is a
finite sum of functions of the form ϕ1 ⊗ ϕ2 where ϕi is a Ki∞ -finite smooth
function in L2cusp ([Gi ])(πi ) for 1 ≤ i ≤ 2.
16.4. Let v be a finite place of the number field F and let G be a reductive
group over F . Assume that ZG (Fv ) is compact. Prove that any supercuspidal
function is a finite sum of matrix coefficients of supercuspidal representations
of G(Fv ).
where we take the basis Bπ to consist of K∞ -finite forms (and hence smooth
by Proposition 4.4.2).
Chapter 17
Orbital Integrals
Abstract In this chapter we define and study orbital integrals. We then use
them to describe geometric sides of trace formulae.
This chapter requires more serious algebraic geometry and Galois cohomology
than other chapters in this book. Moving forward, the main concepts that
we will require are various constructions and definitions related to relative
classes contained in the current section and the definition of relative orbital
integrals, given in §17.4 in the local setting and §17.7 in the adelic setting. If
desired, the other sections of this chapter can be omitted, although they will
be used in the proofs of results in §17.4 and §17.7.
We recall the notion of an algebraic group action and some construc-
tions involving it. These show up constantly in the study of trace formu-
lae, implicitly or explicitly, so we have summarized some useful results.
Our basic references for the geometric constructions in this chapter are
[MFK94, Mil17, Poo17].
Let k be a Noetherian ring, let H be a smooth (affine) group scheme over
k, and let X be an affine scheme of finite type over k. A morphism
a : X × H −→ X (17.1)
327
328 17 Orbital Integrals
a×1
X × H × H −−−−H→ X × H
a×my
ay
a
X ×H −−−−→ X
commutes and the composite
1 ×e a
X −−X−−→ X × H −−−−→ X
is the identity. Here m is the multiplication map and e : Spec(k) → H is
the identity section. These assumptions imply, in particular, that for every
k-algebra R
a : X(R) × H(R) −→ X(R)
is a right action. One can formulate the notion of a left action in the analogous
manner. Let a : X × H → X and a0 : X 0 × H → X 0 be right actions of H on
affine schemes X and X 0 of finite type over k, respectively. A morphism of
right actions is a morphism of schemes b : X → X 0 such that
a
X × H −−−−→ X
b×1H y
yb
X 0 × H −−−−
0
→ X0
a
a(γ, ·) : H −→ X.
Hγ = Spec(k) ×X H
where the map H → X is a(γ, ·) and the map from Spec(k) to X is given by
γ. In terms of points, for k-algebras R one has
H(R) −→ I\H(R)
(17.3)
h 7−→ bh.
The fibers of this morphism are the right cosets of I(R) in H(R) [Mil17, §5(c)].
It is important to point out that the map (17.3) need not be surjective for a
given R. For more information we refer to Proposition 17.1.7. Of course we
can also form a right quotient H/I in the analogous manner.
B\GL2 ∼
= P1 .
Hγ \H −→ X
is an immersion. t
u
For affine schemes Y = Spec(A) for a ring A, denote by |Y | the set of
prime ideals of A equipped with the Zariski topology. In other words |Y |
is the underlying topological space of Y . The morphism a(γ, ·) induces a
continuous map of topological spaces |H| → |X|. The image
O(γ) ⊂ X (17.4)
to be the subscheme with this image as its underlying set, given the reduced
induced scheme structure.
Proposition 17.1.2 The map a(γ, ·) induces an isomorphism
Hγ \H −→ O(γ).
H −→ O(γ)
is smooth if Hγ is smooth.
330 17 Orbital Integrals
Proof. See [Mil17, Corollary 7.13 and Proposition 7.17] for the first assertion
and see [Mil17, Proposition 7.4 and Proposition 7.15] for the last two asser-
tions. t
u
I ◦ \H −→ I\H
Example 17.2. Consider the situation when X = H and H acts via conjuga-
tion. Thus for k-algebras R the action is given by
We refer to this as the group case. In this setting O(γ) is the conjugacy
class of γ and Hγ is the centralizer of γ. Assume that k is perfect. Then an
element γ ∈ X(k) is relatively semisimple if and only if it is semisimple in
the usual sense of Definition 1.12 (see [Ste65, Corollary 6.13]). Moreover γ is
relatively regular if and only if γ is regular in the sense of [Ste65]. This is the
reason for the terminology relatively semisimple and relatively regular.
Example 17.3. Let X = h, where h := Lie H, and let H act on X via the
adjoint action. Even if one is primarily interested in the group case, it has
17.1 Group actions, orbits and stabilizers 331
Γ (R) := X(R)/H(R).
where O(Y ) is the coordinate ring of the affine scheme Y (see §1.2). In other
words, a defined in (17.1) is just the map of affine schemes induced by the
map ba of k-algebras. Call an element r ∈ O(X) invariant if b a(r) = r ⊗ 1.
The set of invariant elements is a k-subalgebra, and by abuse of notation we
let
p : X −→ X/H.
332 17 Orbital Integrals
Often in the literature the GIT quotient is denoted by X // H and the stack
theoretic quotient is denoted by X/H or [X/H]. Since we only use GIT
quotients in this work and the doubleslash could be confused with notation
we use for double quotients (see (5.5) for example) we adopt the notation
above.
Let use explain why X/H deserves to be regarded as a quotient. Let p1 :
X ×H → X denote the projection to the first factor. A categorical quotient
of X by H is a k-scheme Y and a morphism p : X → Y such that
a
X × H −−−−→ X
p1 y
p
y (17.10)
p
X −−−−→ Y
commutes and for any morphism q : X → Z from X to another k-scheme Z
such that
a
X × H −−−−→ X
p1 y
q
y (17.11)
q
X −−−−→ Z
commutes there is a unique morphism χ : Y → Z such that q = χ ◦ p.
Concretely, the assertion that (17.11) commutes is a way to make precise the
assertion that q is constant on H-orbits. The universal property states that
any morphism constant on H-orbits factors through the categorical quotient.
As usual with universal properties, the definition immediately implies that
the categorical quotient is unique up to isomorphism if it exists.
As the reader probably has guessed the GIT quotient is a categorical quo-
tient [MFK94, Theorem 1.1], and this explains why it is reasonable to call it
a quotient. In fact, X/H is a uniform categorical quotient, , which im-
plies in particular that for all separable field extensions E/k the base change
(X/H)E is a categorical quotient of XE by HE .
For our purposes we will need to understand the points of X/H. This is
somewhat complicated. Let k ≤ k sep ≤ k be a separable closure and algebraic
closure of k, respectively. We recall that the closed points of the topological
space underlying a scheme Y of finite type over k can be identified with
Autk (k)-orbits of elements of Y (k) [GW10, Proposition 5.4]. Using this stan-
dard fact we can give a fairly concrete description of the points of X/H over
k:
Proposition 17.1.5 Assume that H ◦ is reductive. There is a bijection
Proof of Proposition 17.1.5: Using the action of Galk (see [GW10, §5.2] for
example) one easily reduces the proposition to the case where k = k sep .
We henceforth assume that k = k sep . The morphism X → X/H is surjec-
tive (even universally submersive) [MFK94, Chapter 1.2, Theorem 1.1]. We
claim that the induced map
I\H(k) = {I(k sep )h ∈ I(k sep )\H(k sep ) : hξ(h−1 ) ∈ I(k sep ) for all ξ ∈ Galk }.
Proof. Using the action of Gal(k sep /k) (see [GW10, §5.2] for example) one
easily reduces the proposition to the case where k = k sep .
We henceforth assume that k = k sep . The morphism H → I\H is faithfully
flat and hence surjective [Mil17, Proposition 7.4(b)]. Arguing as in the proof
of Proposition 17.1.5 we deduce that
is also surjective. Since the fibers are the cosets of I(k sep ) in H(k sep ) (see the
discussion around (17.3)) we deduce the proposition. t
u
Still assuming H has reductive neutral component one has a natural sequence
of maps
Assume that
H ◦ := Hr × Hu (17.16)
deal with the case where γ is relatively regular with respect to the action of
Hr .
The subset Z ⊆ X(k) of points that are relatively regular with respect to
the action of Hr can be identified with the set of closed points in an open
subset of the underlying topological space of X. This is a consequence of the
upper semicontinuity of the dimension of stabilizers (see the discussion above
[MFK94, Definition 0.9], but be aware that the notion of regular in loc. cit. is
not the same as ours). Let
X reg ⊂ X
be the open subscheme of X with underlying topological space Z given the
reduced induced subscheme structure. Thus X reg (k) = Z. It is easy to see
that X reg is an H-invariant subscheme.
Assume for the moment that γ ∈ X reg (k). Now consider the GIT quotient
p : X −→ X/Hr
Using the universal property of categorical quotients and the fact that Hu
and Hr commute we see that the action of Hu on X induces an action of Hu
on X/Hr .
Let O(p(γ)) be the Hu orbit of p(γ). By [Mil17, Theorem 17.64] the orbit
O(p(γ)) is closed. Thus p−1 (O(p(γ))) is closed. But using Proposition 17.1.5
and our assumption that γ is regular one sees that p−1 (O(p(γ))) is the orbit
of γ under H.
Thus we have proven the proposition for γ ∈ X reg (k), in other words,
when γ is relatively regular with respect to Hr . We now reduce the general
case to this one. Let X 0 = X, X 1 = X − X reg , and for i > 1 let X i =
X i−1 − (X i−1 )reg . These subschemes are all preserved by H. Thus we have
a sequence of closed subschemes
X = X0 ⊇ X1 ⊇ · · · .
X = X0 ) X1 ) · · · ⊃ Xn ) ∅
where for each i one has X i−1 − X i = X (i−1)reg . Now let γ ∈ X(k) be
relatively semisimple with respect to Hr . Let i be the largest index such that
γ ∈ X i (k). Then γ is relatively regular in X i (k) with respect to the action of
Hr . Let O(γ) be the Hr -orbit of γ in X. Its intersection with X i is O(γ) itself,
and it is closed (being the intersection of two closed subschemes). Moreover
the H-orbit of γ in X is contained in X i . Thus we have reduced to the case
where γ is relatively regular with respect to Hr , as desired. t
u
336 17 Orbital Integrals
We assume until §17.4 below that H is a smooth algebraic group over a field
k with reductive neutral component that acts (on the right) on an affine k-
scheme X. The structure of X/H around sufficiently nice points γ ∈ X(k) has
a simple structure. Luna’s slice theorem, which we will state in this section,
makes this precise. Luna’s slice theorem will only be used in passing below,
so the reader should feel free to omit this section. However, it is useful for
developing an understanding of the quotient X/H.
If H acts on an affine scheme Y of finite type over k we let
X ×H Y := X × Y /H. (17.17)
S × H × Hγ −→ S × H
ϕ : S ×Hγ H −→ X.
is an isomorphism.
Luna’s slice theorem asserts the existence of étale slices under mild hy-
potheses (see [MFK94, Appendix D]):
Theorem 17.2.1 (Luna’s slice theorem) Assume that k is a character-
istic zero field. Étale slices S exist for any γ ∈ X(k) with closed orbit along
which X is normal. Moreover if X is smooth at γ then S can be taken to
be smooth with image a neighborhood of 0 in the normal vector space to the
orbit O(γ) at γ. t
u
One can use this theorem to give a nice description of the quotient X/H
itself. In more detail, for k-algebras R and subgroups H 0 ≤ H let
0
X H (R) := {x ∈ X(R) : xh = x for h ∈ H 0 (R0 ) and R-algebras R0 }.
(17.19)
T /W (H, T )−→H/H,
˜
and the lemma follows in this case. In general we have γ = zγ 0 where (z, γ 0 ) ∈
ZH,t (k̄) × H der (k̄) and one checks that
17.2 Luna’s slice theorem 339
(Cγ,H )k = Cγ 0 ,Hk̄ ,
σ : G −→ G
In this case the set Γ (R) is often known as the set of relative conjugacy
classes (see Example 17.4). They are a generalization of the notion of con-
jugacy classes in a strict sense (see Exercise 17.9).
There is a great deal of helpful geometry available in this special case that
aids in the study of the set of classes Γ (R) of Definition 17.2. It is useful to
introduce the map
Bσ : G −→ G
given on points by g 7→ g −σ g. We denote by Q the scheme-theoretic image of
Bσ . Then Q is a closed affine k-subscheme of G.
We have an isomorphism [Ric82, Lemma 2.4] of k-schemes
Bσ : Gσ \G −→ Q (17.23)
t
u
We then have the following proposition:
Proposition 17.2.10 An element γ ∈ X(k) = G(k) is relatively semisimple
with respect to the action of H if and only if Bσ (γ) is semisimple in the usual
sense.
Proof. See [Ric82, §7]. t
u
We have seen that the notion of a maximal torus in a reductive group is ab-
solutely crucial. In the case of symmetric subgroups, the following definition
is a substitute:
Definition 17.7. A torus T ⊆ G is said to be σ-split if for all k-algebras R
and all t ∈ T (R) one has t−1 = tσ .
Beware that in [Ric82] a σ-split torus is called a “σ-anisotropic torus.”
It is not hard to see that if T is any σ-split torus then T ≤ Q. Indeed,
every element of T (k) is in the image of the isogeny t 7→ t2 = t−σ t. Moreover,
σ-split tori exist (see [Ric82, (2.6)] and [Vus74, §1]).
Let Tσ ≤ Q be a maximal σ-split torus. We have a corresponding Weyl
group
sometimes called the little Weyl group (see §1.7). Here NG (Tσ ) (resp. ZG (Tσ ))
is the normalizer (resp. centralizer) of Tσ in G. This is a finite étale group
scheme over k. We have the following generalization of the Chevalley restric-
tion theorem [Ric82, Corollary 11.5]:
Theorem 17.2.11 (Richardson) Let Tσ ⊆ G be a maximal σ-split torus.
The inclusion Tσ ,→ Q induces an isomorphism
∼
Tσ /W (G, Tσ ) −→ Q/Gσ .
t
u
17.3 Local geometric classes 341
Let k be a field. Suppose that γ ∈ X(k) has closed orbit under H and that
Hγ is smooth (we do not assume that H has reductive neutral component).
We then have a sequence
and we can identify the last object with O(γ)(k) by means of Proposition
17.1.2. We now briefly explain how to compute the image using a little Galois
cohomology. The reader should feel free to omit this discussion and refer back
to it later if necessary. Our primary reference is [Ser02, §I.5].
Recall that for a smooth affine algebraic group H over a field k one defines
Z 1 (k, H), the set of 1-cocycles of Galk in H(k sep ) to be the set of set
theoretic maps
Here the kernel simply means all classes that are sent to the neutral element
of H 1 (k, H).
Now suppose that I ≤ H is a smooth subgroup scheme. We leave the
following lemma as Exercise 17.12:
Lemma 17.3.1 The map
cl : I\H(k) −→ D(k, I, H)
I(k sep )h 7−→ (σ 7→ hσ(h−1 ))
where the disjoint union is over the set of H(k)-orbits in I\H(k) and the
implicit maps are given by
Ih (k)\H(k) −→ I\H(k)
x 7−→ I(k sep )hx.
We note that
cl(Im(Ih (k)\H(k) → I\H(k)))
is simply the class of the cocycle hσ(h−1 ).
The groups Ih are not isomorphic as h varies, but they are related in an
important manner. Two smooth algebraic groups Q and Q e are said to be
forms of each other if there is an isomorphism
ϕ : Qksep −→
˜ Q
e ksep .
over k sep and then classifying all of the groups over k that base change to a
given group over k sep . A form is called an inner form if for all σ ∈ Galk the
automorphism ϕ−1 ◦ σ ◦ ϕ of Qksep is an inner automorphism. It is easy to
check that the Ih are all inner forms of I (see Exercise 17.13).
Now suppose that k is a local field. In keeping with our usual notation for
local and global fields we let k = F . We record the following useful result
(see [Poo17, Proposition 3.5.73]):
Proposition 17.3.2 If X → Y is a smooth map of separated schemes of
finite type over F then X(F ) → Y (F ) is open. t
u
Using this proposition we prove the following lemma:
Lemma 17.3.3 Let I ≤ H be a smooth algebraic subgroup. The map
cl : I\H(F ) −→ D(F, I, H)
Proof. The last assertion follows from the first and Lemma 17.3.1. To prove
the first assertion it suffices to show that the preimage under cl of any point
in D(F, I, H) is open. This is equivalent to the assertion that the maps
Ih (F )\H(F ) −→ I\H(F )
Let F be a local field and let H be a smooth algebraic group defined over F
acting on a smooth affine scheme X. We define local relative orbital integrals
in this section.
Assume we are given a quasi-character χ : H(F ) −→ C× .
344 17 Orbital Integrals
Note that we have used the fact that γ is χ-relevant to define this integral.
This integral depends on the choice of measures, but traditionally they are
not encoded in the notation. In some applications it may be more natural to
replace Hγ by some subgroup of Hγ containing (Hγ )◦ .
Hγ (F )\H(F ) −→ X(F )
(17.32)
Hγ (F )h 7−→ γh.
17.5 Torsors 345
Thus it suffices to show that this map is proper. The orbit O(γ) of γ is closed
in X. Hence O(γ)(F ) is closed in X(F ) by Theorem 2.2.1.
It therefore suffices to treat the case where X = O(γ) = Hγ \H. In this case
we invoke Lemma 17.3.3 to see that the image of Hγ (F )\H(F ) in Hγ \H(F )
is closed. t
u
17.5 Torsors
In this section we discuss torsors, which play a key role in §17.6. It is con-
venient to return to the more general setting of §17.1. Thus H is a smooth
group scheme over a Noetherian ring k acting (on the right) on an affine
scheme X of finite type over k.
For example one could take X = H with the natural right H action. This is
known as the trivial H-torsor. This example can be profitably generalized:
Definition 17.10. Assume X is faithfully flat over a Noetherian ring k. It
is an H-torsor if the map
1X × a : X × H −→ X × X
is an isomorphism.
An equivalent (and more intuitive) way to define a H-torsor is to say that
it is a scheme flat and of finite type over k equipped with an action of H
that becomes isomorphic to the trivial H- torsor after a faithfully flat finite
type base extension. For more details we refer the reader to [Poo17, §6.5.1].
Our definition is the specialization of the definition in [Poo17, §6.5.1] to the
case of affine Noetherian base schemes. The set of all isomorphism classes of
H-torsors is denoted by
q 1 (k, H) = H
H q 1 (Spec(k), H). (17.33)
fppf fppf
It is a pointed set with the class of the trivial torsor as neutral element. The
fppf stands for fidèlment plat de présentation finie, since torsors are required
to be faithfully flat and locally of finite presentation over k. Since we have
assumed H is smooth, H-torsors will in fact be smooth over Spec(k) [Poo17,
Remark 6.5.2]. Assume that k 0 a faithfully flat k-algebra of finite type. Then
one has a natural map
H q 1 (k 0 , H)
q 1 (k, H) −→ H (17.34)
fppf fppf
Y ∧I H := Y × H/I
1X × a : X × H −→ X × X (17.38)
is an isomorphism we see that the map a(x, ·) : H(k sep ) → X(k sep ) is a bi-
jection. In particular for each σ ∈ Galk there is c(σ) ∈ H(k sep ) such that
σ(x) = x.c(σ). One checks that c(σ) is a 1 cocycle, and this is the defini-
tion of the map (17.37). We omit the construction of the inverse map (see
[Poo17, §5.12.4]). Using (17.37) we occasionally allow ourselves to identify
Hq 1 (Spec(k), H) and H 1 (k, H) when k is a field.
fppf
q 1 (k, I) −−−−→ H
H q 1 (k, H)
fppf fppf
∼y
∼
y
H 1 (k, I) −−−−→ H 1 (k, H),
17.5 Torsors 347
where the vertical arrows are given by the bijection (17.37), the top arrow is
given by (17.36) and the bottom arrow is given by observing that a cocycle
with coefficients in I(k sep ) has coefficients in H(k sep ).
I(R) × H(R)−→Y
˜ (R) × H(R)
(i, h) 7−→ (yi−1 , ih).
where the vertical arrows are induced by the quotient maps in the category
of fppf sheaves over Spec(k 0 ) and the horizontal arrows are bijections. Taking
the direct limit over all separable finite extensions k sep ≥ L ≥ k we deduce the
same diagram for L = k sep . In this case the horizontal arrows are not GalL =
Galk equivariant. Despite this, one still deduces the claim from Proposition
17.1.7.
Let y ∈ Y (k sep ) and for each σ ∈ Galk let c(σ) ∈ I(k sep ) be the elements
such that σ(y) = yc(σ). Thus c is the cocycle attached to Y by (17.37). We
now pick (y × 1)I(k sep ) to be our basepoint in Y ∧I H(k sep ). Then
t
u
348 17 Orbital Integrals
is proper.
For γ as in the statement of the theorem one has that O(γ)(AF ) is closed
in X(AF ) by Theorem 2.2.1. It is therefore no loss of generality to assume
that X = O(γ), which is to say that there is a smooth subgroup I of H such
that X := I\H. For any finite set of places S of F including the infinite
places we set
Y
ObFS = OFv and OFS := F ∩ O bFS .
v6∈S
Thus OFS < F is the set of elements of F that are integral outside of the
finite places in S. Recall the notion of a model of an affine scheme reviewed
in §2.4. For a sufficiently large set of places S of F we can choose models H
and X of H and X = I\H over OFS , respectively, such that there is a map
H→X
where the restricted direct product is taken with respect to the open sets
X (OFv ). The isomorphism depends on the choice of model of X. This discus-
sion is also valid with X and X replaced by H (resp. I) and H (resp. I). In
fact we discussed the algebraic group case in §2.3. Moreover, there are again
isomorphisms of topological groups
0 0
H(AF ) ∼ I(AF ) ∼
Y Y
= H(Fv ) and = I(Fv ), (17.43)
v v
where the restricted direct product is taken with respect to the subgroups
H(OFv ) (resp. I(OFv )) for all finite places v. These isomorphisms are induced
by the choice of the model of H and are compatible with the group actions
X(Fv )×H(Fv ) → X(Fv ) and X(AF )×H(AF ) → X(AF ) because we assumed
that there is a morphism H → X whose generic fiber is H → X. It is tempting
to try to define the quotient I\H and relate it to X , but we do not wish to
discuss representability of quotients over Dedekind rings. Fortunately coming
to grips with this issue is unnecessary for our purposes.
Lemma 17.6.2 Assume that I is connected. For a large enough finite set
bS ) in
of places S of F including the infinite places the inverse image of X (O F
S S S
H(AF ) is equal to I(AF )H(OF ). The morphism H → X induces a bijection
b
bS )\H(O
I(O bS )−→X bS ).
˜ (O
F F F
Hγ◦ \H −→ Hγ \H
is proper by [Con12b, Proposition 4.4]. Here we have also used the fact,
mentioned above, that for any smooth algebraic subgroup I ≤ H the scheme
I\H is separated and of finite type over F [Mil17, Theorem 7.18].
Thus we are reduced to showing that for any connected smooth subgroup
I ≤ H the map
I(AF )\H(AF ) → I\H(AF )
is proper. To keep notation consistent with Lemma 17.6.2 we write X = I\H
and let X and H be defined as above that lemma.
Let Ω ⊂ X(AF ) be a compact set. We must show its inverse image in
I(AF )\H(AF ) is compact. Let S be a finite set of places of F including the
infinite places. It suffices to treat the case Ω = ΩS × Ω S where ΩS ⊂ X(FS )
and Ω S = X (O bS ) are compact since any compact subset admits a finite cover
F
by open sets whose closure is of this form.
Taking S sufficiently large and invoking Lemma 17.6.2, we see that the
inverse image of Ω S in H(ASF ) is I(ASF )H(O bS ). We conclude that the inverse
F
image of Ω S in I(ASF )\H(ASF ) is I(ASF )H(O bS ), hence compact. On the other
F
hand the inverse image of ΩS in I(FS )\H(FS ) is compact by Lemma 17.3.3.
2
We observe the following:
Lemma 17.6.3 If Y is an affine scheme of finite type over F then Y (F ) is
discrete and closed in Y (AF ).
Proof. The set F < AF is discrete and closed by Lemma 2.1.3 and hence we
conclude by Exercise 2.2. t
u
Using the techniques above we can prove a result which was used earlier
in Lemma 14.3.2:
Lemma 17.6.4 Let G be a reductive group over F and let H ≤ G be a
reductive subgroup. The image of G(F ) in AG \G(AF )/H(AF ) is discrete and
closed.
Here, of course, we have given AG \G(AF )/H(AF ) the quotient topology.
we deduce that p−1 (x)(AF ) = p−1 (x)(F ) is a single point in G(AF )/H(AF )
hence discrete and closed. Thus the image of G(F ) in G(AF )/H(AF ) is closed.
Define G(AF )1 as in (2.10). The subset
is a closed set containing G(F )(H(AF )∩G(AF )1 ), and hence G(F ) is discrete
and closed in G(AF )1 /(H(AF )∩G(AF )1 ). On the other hand the natural map
q 1 (OF , I) −→ H
Dfppf (OFv , I, H) := ker(H q 1 (OF , H)), (17.44)
fppf v fppf v
q 1 (OF , I) −−−−→ H
H q 1 (OF , H)
fppf v fppf v
y
y (17.45)
where the vertical arrows are given by (17.34), the top arrow is given by
(17.36), and the map on the bottom row corresponds to the obvious morphism
given by observing that a cocycle with coefficients in I(Fvsep ) tautologically
has coefficients in H(Fvsep ) (see Lemma 17.5.1). We let
We then set
0
Y
D(AF , I, H) := D(Fv , I, H)
v
where the prime indicates that we take the restricted direct product with
respect to D(OFv , I, H) ⊆ D(Fv , I, H). We note that if S 0 ⊇ S then the
natural map
D(AF , I, H) −→ D(AF , IOS0 , HOS0 )
F F
is an isomorphism.
Assume now that D(Fv , I, H) is finite for all v. In this case we give
H 1 (AF , H) a topology by stipulating that the open sets are all subsets that
are equal to
Y
ΩS 0 × D(OFv , I, H) (17.47)
v 0 6∈S 0
D(F, I, H) −→ D(AF , I, H)
Y (OFnrv ) 6= ∅.
Proof. Let F be the residue field of OFv and let F be its algebraic closure. It
is the residue field of OFnrv . The surjection OFnrv → F induces a surjection
Y (OFnrv ) → Y (F)
by Hensel’s lemma [BLR90, §2.3, Proposition 5]. On the other hand Y (F) is
nonempty since Y is smooth by Lemma 17.1.6. t
u
Lemma 17.6.7 A class in H 1 (Fv , H) is in the image of the map from
Ȟ 1 (OFv , H) if and only if it is cohomologous to a cocycle with values in
H(OE ) for some finite unramified Galois extension E/Fv .
Proof. Let H be an HOFv -torsor. It is smooth over OFv , and hence there is
a finite unramified extension E/Fv such that Y(OE ) is nonempty by Lemma
17.6.6. The cocycle defined using this point y as in the definition of (17.37)
will take values in H(OE ). This implies the “only if” implication.
Conversely, let c : GalFv → H(OE ) be a cocycle, where E/Fv is an unram-
ified extension. This defines a descent datum on HOE and hence a H-torsor
Y (see [Poo17, Remark 4.4.7]). The cocycle attached to the generic fiber of
this torsor is c, so we deduce the converse assertion. t
u
For each place v we have a class map cl : I\H(Fv ) → D(Fv , I, H) as in
the proof of Lemma 17.3.1.
Lemma 17.6.8 The local class maps induce a map
If H 1 (Fv , I/I ◦ ) is finite for all v it sends compact sets to compact sets.
We do not know if cl is continuous.
354 17 Orbital Integrals
Proof. Let OFnrv be the ring of integers of the maximal unramified extension
Fvnr in Fvsep . By spreading out, for S 0 large enough the map HOS0 → XOS0
F F
is smooth [Poo17, Theorem 3.2.1] and hence the map H(OFnrv ) → X (OFnrv ) is
surjective for v 6∈ S 0 by Lemma 17.6.6 applied to the fibers of the morphism
HOFv → XOFv for v 6∈ S 0 . Thus for v 6∈ S 0 given x ∈ X (OFv ) we can choose
h ∈ H(OFnrv ) such that I(F sep )h = x. Thus cl(x) is contained in D(OFv , I, H)
by Lemma 17.6.7. This implies the first assertion of the lemma.
For the second assertion in view of Lemma 17.3.1 it suffices to observe
that by the argument above cl(x) ∈ Dfppf (OFv , I, H) for v 6∈ S 0 . t
u
Proposition 17.6.9 Let A0 ≤ AH be a subgroup. One has a commutative
diagram
cl
1 −−−−→ I(F )\H(F ) −−−−→ I\H(F ) −−−−→ D(F, I, H)
ay
y y y
cl
1 −−−−→ I(AF )\H(AF )/A0 −−−−→ I\H(AF )/A0 −−−−→ D(AF , I, H).
The top row is an exact sequences of pointed sets, and the map a is proper if
we give D(F, I, H) the discrete topology.
We have incorporated the group A0 for use in the proof of Theorem 18.2.2
below. One checks that cl(x) = cl(xa) for all x ∈ I\H(AF ) and a ∈ AH so
the map cl : I\H(AF )/A0 → D(AF , I, H) is well-defined.
Proof. The commutativity of the diagram is clear, the top line is exact since
(17.29) is exact, and a is proper by Theorem 17.6.5. tu
We now prove that the set of relative classes inside a geometric class is
finite under suitable assumptions:
Theorem 17.6.10 Let A0 ≤ AH be a subgroup. Assume that H 1 (Fv , I/I ◦ ) is
finite for all v. Let Ω ⊆ I\H(AF )/A0 be a compact set. There are only finitely
many H(F )-orbits xH(F ) ⊆ I\H(F ) such that xH(AF )/A0 intersects Ω.
xH(AF )/A0 ∩ Ω 6= ∅
The set cl(Ω) is compact by Lemma 17.6.8. Since a is proper, the set in
(17.49) is finite. t
u
17.7 Global relative orbital integrals 355
In §17.7 we will use Theorem 17.6.10 and the following finiteness result for
geometric classes:
Lemma 17.6.11 Assume H is reductive. Let Ω ⊂ X(AF ) be a compact set.
There are only finitely many relatively semisimple geometric classes in X(F )
whose H(AF )-orbit intersects Ω.
Proof. Let
p : X(AF ) −→ X/H(AF )
be the canonical map. By Proposition 17.1.5 an element of X/H(F ) may
be regarded as a relatively semisimple class. The set of relatively semisimple
geometric classes in X(F ) whose H(AF ) orbit intersects Ω injects into
For use in the relative trace formula we need to have global analogues of the
local orbital integrals (17.31). Before this we must discuss what we mean by
a smooth function in this setting.
For this section we continue to assume that F is a global field. Fix a
model of X over OFS in the sense of §2.4 and denote it again by X by abuse
of notation. Here S is a sufficiently large set of places of F including all
infinite places. Then for all v 6∈ S of F the characteristic function 1X(OFv ) is
an element of Cc∞ (X(Fv )). We define
Cc∞ (X(A∞ 0 ∞
F )) = ⊗v-∞ Cc (X(Fv ))
where the restricted direct product is taken with respect to the functions
1X(OFv ) . It turns out that Cc∞ (X(A∞ F )), defined in this manner, is equal to
the C-vector space of smooth compactly supported functions on X(A∞ F ) (see
Exercise (17.14)).
If F is a global function field we let Cc∞ (X(F∞ )) be the space of locally
constant compactly supported functions on X(F∞ ), and if F is a number
field we let Cc∞ (X(F∞ )) be the usual space of compactly supported smooth
functions on X(F∞ ), viewed as a real manifold. We then set
Assume that
χ : H(AF ) −→ C×
356 17 Orbital Integrals
converges. This is a very delicate question in general and we will only address
it in special cases. We first make the assumption that
H ◦ := Hu × Hr (17.53)
A := ACX (17.54)
where f ∈ Cc∞ (X(AF )). There is an obvious obstruction for the contribution
of a given γ to converge. Letting Hγ denote the stabilizer of γ in H, it is
possible that AHγ◦ is not contained in A, which implies that the contribution
17.7 Global relative orbital integrals 357
is convergent. Here the sum is over γ ∈ X(F ) that are gcf, relatively uni-
modular and relatively elliptic, and such that the Hr -orbit of γ is closed.
Moreover
Z X X0
f (γh)χ(h)dh = τ (Hγ ) ROγ (f ) (17.58)
AH(F )\H(AF ) [γ]∈Γ (F )
where the sum over γ on the left is as before and the prime indicates the sum
on the right is over all classes that are relevant, gcf, relatively unimodular
and elliptic and such that the Hr -orbit of γ is closed.
Several remarks are in order. First, the assumption that the Hr -orbit of
γ is closed implies that γ is relatively semisimple with respect to H (see
Lemma 17.1.9). We will use this fact without further mention below. The
assumption that γ is relatively unimodular is equivalent to the assertion that
Hγ is unimodular, which is implied by our assumption that the Hr -orbit of
γ is closed if H ◦ = Hr by Theorem 17.1.4. In the definition of ROγ (f ) we
dh
use the measure dh γ
on
where dhγ is a Haar measure on A\Hγ (AF ). We use the same measure dhγ in
the definition of τ (Hγ ). Finally, the assumption that γ is gcf is automatically
satisfied if Hγ is connected or F is a number field (Theorem 17.3.4).
We now handle the convergence claim in Theorem 17.7.2:
Proposition 17.7.3 The expression (17.57) is convergent.
358 17 Orbital Integrals
Here the outer sum is over elements of Γgeo,ss (F ) with closed Hr -orbit that are
gcf, relatively unimodular and relatively elliptic. Unfolding the inner integral
for a fixed geometric class [γ0 ] we obtain
Z X
|f (γh)χ(h)|dh
AH(F )\H(AF ) γ∈[γ ]
0
XZ
= |f (γh)χ(h)|dh
[γ] AHγ (F )\H(AF )
where the second sum is over the relative classes [γ] in the geometric class
[γ0 ]. This is equal to
Z
X dh
measdhγ (AHγ (F )\Hγ (AF )) |f (γh)χ(h)| (17.60)
Hγ (AF )\H(AF ) dhγ
[γ]
where for each relatively unimodular γ the measure dhγ is a Haar measure
on A\Hγ (AF ). Here we are using the fact that γ is relatively unimodular to
conclude that dhγ is a Haar measure, not merely a right Haar measure. By
Theorem 2.6.2 the measure
is finite for all γ in (17.60) since all such γ are relatively elliptic. The sum
over [γ] in (17.60) is finite by Theorem 17.6.10. Thus by Proposition 17.7.1
we deduce that (17.60) is finite.
It therefore suffices to show that there are only finitely many relatively
semisimple geometric classes [γ0 ] that can contribute a nonzero summand to
(17.59). Let us return to (17.59). It is equal to
XZ X
|f (γhr hu )χ(hr hu )|dhr dhu
[γ0 ] AHr (F )\Hr (AF )×[Hu ] γ∈[γ ]
0
XZ X
= |fe(γhr )χ(hr )|dhr
[γ0 ] AHr (F )\Hr (AF ) γ∈[γ ]
0
where the inner sum is over Hr (F )-orbits (not H(F )-orbits as above) in
[γ0 ] consisting of elements that are relatively unimodular, elliptic and gcf.
It follows from Lemma 17.6.11 that there are only finitely many closed Hr -
orbits in X(F ) that can contribute a nonzero summand to (17.61). This
implies that there are in fact only finitely many [γ0 ] that can contribute a
nonzero summand to (17.61), and hence to (17.59). t
u
where the sum is over γ with closed Hr -orbit that are relatively semisimple,
elliptic, unimodular and gcf. We unfold the integral to obtain
XZ
f (γh)χ(h)dh
[γ] AHγ (F )\H(AF )
XZ
= χ(h0 )dγ h0 ROγ (f )
[γ] AHγ (F )\Hγ (AF )
360 17 Orbital Integrals
0
X
= τ (Hγ ) ROγ (f )
[γ]
where the first sum over γ is over H(F )-orbits in X(F ) with closed Hr -
orbit that are gcf and relatively elliptic and unimodular. The quotient
AHγ (F )\Hγ (AF ) is of finite volume by Theorem 2.6.2, and the integral over
it in the second expression above is nonzero if and only if γ is relevant. As
in the statement of the theorem, the prime on the sum in third expression
indicates that the sum is over relevant relatively elliptic unimodular classes
in X(F ) with closed Hr -orbit. 2
Exercises
17.4. Give examples to show that the map Γss (k) −→ H(k)\X(k) need not
be surjective and that the map
17.6. Let k be an algebraically closed field. Assume that γ ∈ GLn (k) has
finite order and that either k is of characteristic zero or that the order of
γ is coprime to the characteristic of k. Prove that γ is semisimple. Give a
counterexample to show that if the characteristic of k is positive and divides
the order of γ ∈ GLn (k) then γ need not be semisimple.
17.7. Let k be a field with algebraic closure k. Prove that two elements of
GLn (k) are GLn (k)-conjugate if and only if they are GLn (k)-conjugate.
17.7 Global relative orbital integrals 361
17.8. Let k be a field with separable closure k sep . Prove that a conjugacy
class in GLn (k sep ) that is fixed under Gal(k sep /k) contains an element in
GLn (k).
17.13. Prove that the groups Ih appearing in (17.30) are all inner forms of
I.
σ 7→ gσ gσ0
17.17. Let F be a global field and let H be a smooth algebraic group over
F acting in an affine scheme X of finite type over F on the right. Prove that
γ ∈ X(F ) is relatively semisimple if and only if any element in O(γ) ∩ H(F )
is relatively semisimple. Assuming γ ∈ X(F ) is relatively semisimple, prove
moreover that γ satisfies the following conditions if and only if any element
of the geometric class of γ satisfies the following conditions:
(a) γ is relatively unimodular.
(b) γ is relatively elliptic.
(c) γ is gcf.
Chapter 18
Simple Trace Formulae
363
364 18 Simple Trace Formulae
H ≤G×G
Let
AG,H : = AH ∩ (AG × AG )
(18.2)
A : = AH ∩ ∆(AG )
where
∆ : AG −→ AG × AG
is the diagonal embedding. Thus A ≤ AG,H . Let CG ≤ H be the stabilizer
(in H) of G. It is clear that A ≤ ACG . We assume throughout this section
that in fact A = ACG .
Let
χ : AG,H \H(AF ) −→ C×
be a quasi-character trivial on H(F ) and let f ∈ Cc∞ (AG \G(AF )). As in
(16.14) we define the relative trace of π(f ) with respect to H and χ to be
Z
rtr π(f ) := rtrH,χ π(f ) = Kπ(f ) (h` , hr )χ(h` , hr )d(h` , hr ).
AG,H H(F )\H(AF )
(18.3)
Note that this is slightly different from the definition in (17.51). The root of
this is that it is more convenient for the spectral side of the trace formula to
work with functions on AG \G(AF ) instead of G(AF ). In order for Theorem
366 18 Simple Trace Formulae
18.2.2 below to be valid it is important that we use the same Haar measure
d(h` , hr ) on AG,H \H(AF ) used in the definition of the relative trace, and dhγ
is a Haar measure on AG,H \Hγ (AF ). Thus d(hdh` ,h γ
r)
is a left AG,H \H(AF )-
invariant Radon measure on
It depends on the choice of Haar measure dhγ on A\Hγ (AF ). We now check
that this is well-defined by slightly modifying the proof of Proposition 17.7.1:
Proposition 18.2.1 If γ is relatively unimodular and semisimple then
Z
dh
f (h−1
r γh` )χ(h)
< ∞.
AG,H \H(AF )
A\H (A )
dhγ
γ F
Proof. Let
where
a(γ, (h` , hr )) = h−1
` γhr
is the orbit map. The horizontal arrows from left to right are homeomor-
phisms, and the horizontal arrows from right to left are injections onto closed
subsets. The leftmost vertical arrow is proper by Theorem 17.6.1, and it fol-
lows that the middle, and hence right, vertical arrow is proper. As in the
proof of Theorem 17.4.1 and Proposition 17.7.1 this suffices to establish the
proposition. t
u
Theorem 18.2.2 Let f ∈ Cc∞ (AG \G(AF )) be a function such that R(f ) has
cuspidal image and such that if the H(AF )-orbit of γ intersects the support of
f then the γ is gcf, relatively elliptic, relatively unimodular, and the Hr -orbit
of γ is closed. Then
X 0
X
rtr π(f ) = τ (Hγ )ROγ (f )
π [γ]∈Γ (F )
where the sum on the left is over (H, χ)-distinguished cuspidal automorphic
representations π and the sum on the right is over relevant relative classes
[γ] that are gcf, relatively unimodular, relatively elliptic, and such that the
Hr -orbit of γ is closed. Both sums are absolutely convergent.
We recall that if the Hr -orbit of γ is closed, then γ is relatively semisimple
by Lemma 17.1.9. Following Langlands, the left hand side of the trace for-
mula is called the spectral side and the right hand side of the trace formula
is called the geometric side. We call the trace formula of Theorem 18.2.2
general because its geometric set-up includes all trace formulae studied in
the literature as special cases. Indeed the formula does not hold for a general
connected algebraic subgroup H ≤ G × G without serious modification, so
in some sense it is as general as possible. It is simple because we have im-
posed assumptions on f to eliminate analytic difficulties on the spectral and
geometric sides.
In the proof of Theorem 18.2.2 we require the following analogue of Lemma
17.6.11:
Lemma 18.2.3 Assume that H is reductive and let Ω ⊂ AG \G(AF ) be a
compact set. The set of relatively semisimple geometric classes in G(F ) whose
H(AF )-orbit intersects Ω is finite.
χ ◦ a(1, ·) : H −→ Gm
Then G(F ) is contained in G(AF )0 and the action (18.1) of H(AF ) on G(AF )
preserves G(AF )0 . Moreover the inclusion G(AF )0 → G(AF ) induces a home-
368 18 Simple Trace Formulae
omorphism
(see Exercise 18.1). Thus it suffices to show that given a compact subset Ω ⊂
G(AF )0 /H(AF ) there are only finitely many relatively semisimple geometric
classes in G(F ) whose image in G(AF )0 /H(AF ) intersects Ω. Let
denote the restriction of the natural map G(AF )/H(AF ) → G/H(AF ) to the
GIT quotient. In view of Proposition 17.1.5 it suffices to observe that p(Ω)
is compact and G/H(F ) is discrete and closed by Lemma 17.6.3, hence
p(Ω) ∩ G/H(F )
is finite. t
u
Since we have assumed that R(f ) has cuspidal image one has
where the sum on [γ] is over classes in Γ (F ) that are gcf, relatively uni-
modular, relatively elliptic, and such that the Hr -orbit of γ is closed. The
integral Z
χ((h` , hr ))dhγ
AHγ (F )\Hγ (AF )
18.2 A general simple relative trace formula 369
0
X
= τ (Hγ ) ROγ (f )
[γ]∈Γ (F )
where the sum is over classes [γ] as in the statement of the theorem.
To justify the manipulations above we must check that
Z
X d(h` , hr )
τ (Hγ ) A \H(A ) |f (h` , hr )χ(h` , hr )| (18.8)
G,H F dhγ
[γ] A\H (A ) γ F
is finite, where the sum is over classes [γ] that are gcf, relatively unimod-
ular, relatively elliptic, and such that the Hr -orbit of γ is closed. The in-
tegral in (18.8) converges absolutely by Proposition 18.2.1. It therefore suf-
fices to prove that only finitely many classes [γ] can contribute a nonzero
summand to (18.8). Consider the contribution of a particular gcf, rela-
tively semisimple geometric class [γ0 ]. By Theorem 17.6.10 for any com-
pact set Ω ⊂ Hγ0 \H(AF )/AG,H there are only finitely many H(F )-orbits
xH(F ) ⊂ I\H(F ) such that xH(AF ) intersects Hγ \H(AF )/AG,H . This im-
plies that there are only finitely many relative classes in the geometric class
[γ0 ] that can contribute to the sum.
Lastly we show that there are only finitely many geometric classes that can
contribute to the sum. By the same argument used in the proof of Theorem
17.7.2 we are reduced to showing that for any compact set Ω ⊂ AG \G(AF )
there are only finitely many geometric classes in G with closed Hr -orbit whose
Hr (AF )-orbit intersects Ω. This is the content of Lemma 18.2.3. 2
We now make some comments on the assumptions on f in Theorem 18.2.2.
For more information on the assumption on f involving the geometric side
of the trace formula we refer to exercises 18.2, 18.3 and 18.4.
In practice in order to ensure that f has cuspidal image one assumes
f = fv f v with fv ∈ Cc∞ (G(Fv )) and f v ∈ Cc∞ (AG \G(AvF )) for some finite
place v where fv is supercuspidal. This suffices to ensure R(f ) has cuspidal
image by Lemma 16.4.1. Unfortunately, assuming that fv is supercuspidal
limits the set of π that can be studied using the simple relative trace formula:
Lemma 18.2.4 Let v be a finite place of F . Assume that f ∈ Cc∞ (G(Fv ))
is supercuspidal (not just F -supercuspidal). If πv is an admissible irreducible
representation of G(Fv ) and πv (f ) 6= 0 then πv is supercuspidal.
reduce to this case. This has the advantage of illustrating some techniques
useful for considering restrictions of representations to derived subgroups.
The restriction of π to ZG (F )Gder (F ) decomposes as a finite direct sum of
irreducible representations and hence the same is true of the restriction to
Gder (F ) since ZG (F ) acts via scalars on π:
Let fa (g) := f (ga). For some i the operator πi (fa ) is nonzero. Since f is
supercuspidal so is fa (as a function on Gder (F )) so we deduce that πi is
supercuspidal by Exercise 16.4. As already mentioned, this implies that π is
supercuspidal. t
u
to remove it. This idea was applied again in [GH15], and it is likely that it
could be useful in much greater generality.
The famous simple trace formula of Deligne and Kazhdan [DKV84] (see also
[Rog83, BDKV84]) is a special case of the simple relative trace formula of
Theorem 18.2.2. We explain this in the current section.
For γ ∈ G(F ) let
where the sum on the left is over isomorphism classes of cuspidal automorphic
representations of AG \G(AF ), m(π) is the multiplicity of π in the cuspidal
spectrum of L2 ([G]) and the sum on the right is over gcf, elliptic, semisimple
conjugacy classes in G(F ).
As in the statement of Theorem 18.2.2, some care must be taken with
measures in order for the identity above to be correct. We choose a Haar
measure dg 0 on AG \G(AF ) and for each semisimple γ a Haar measure dgγ0
18.5 The simple twisted trace formula 373
on AG \Cγ (AF ). Then we use dg 0 to define the operator π(f ) (and hence its
trace) and use dgγ0 to define τ (Cγ ). Finally we normalize the Radon measure
dg dg 0
dgγ on Cγ (AF )/G(AF ) used to define Oγ (f ) so that it corresponds to dgγ0
under the natural isomorphism
(see (16.8)). t
u
When θ is the identity this is simply conjugation. Following the lead of §18.4
we let
. We say that γ is θ-gcf if H 1 (Fv , Cγθ /Cγθ )◦ ) is finite for all places v of F .
Similarly we say γ is θ-semisimple if Cγθ has reductive neutral component
and θ-elliptic if Cγθ is elliptic, which is to say that ACγθ ≤ AG .
We now define the spectral counterpart of a twisted orbital integral. Let
374 18 Simple Trace Formulae
Iθ : L2 ([G]) −→ L2 ([G])
(18.17)
ϕ 7−→ (g 7→ ϕ(θ−1 (g)).
This is the twisted trace of π(f ) (with respect to θ). We observe that π(f )◦
Iθ is acting on the whose π-isotypic space, so if π occurs with multiplicity
greater than 1 this is incorporated into the trace. In contrast, we separated
the trace and the multiplicity in the statement of Corollary 18.4.1.
The following is a simple version of the twisted trace formula:
Corollary 18.5.1 Let f ∈ Cc∞ (AG \G(AF )). Assume that R(f ) has cuspidal
image and that any element of G(F ) that is G(AF )-conjugate to the support
of f is θ-gcf, θ-elliptic and θ-semisimple. Then
X 0
X
tr(π(f ) ◦ Iθ ) = τ (Cγθ )TOγ (f ). (18.19)
π [γ]
Historically the twisted trace formula was first introduced by Saito [Sai75]
and Shintani [Shi79]. For a completely general twisted trace formula we refer
to [LW13]. A relative version of the simple twisted trace formula (which is
another special case of Theorem 18.2.2) was introduced in [Hah09]. We discuss
the relationship between the trace formula and the twisted trace formula in
Chapter 19.
18.6 A variant 375
18.6 A variant
Hx := Gx × H ≤ G × G (18.20)
Gx (AF ) × H(AF ) −→ C×
(g, h) 7−→ χ(h)
rtrHx ,χ π(f ).
converges. For conditions ensuring convergence see Theorem 17.7.2. Thus the
integral
Z X
Φ(xh)χ(h)dh (18.21)
AG ∩AH H(F )\H(AF ) x∈X(F )
is well-defined.
Assume that for each x ∈ X(F ) we can choose fx ∈ Cc∞ (AG \G(AF )) such
that
Z
Φ(xy) = fx (g −1 y)dg (18.22)
Gx (AF )
for y ∈ G(AF ). This is not always possible, see Exercise 18.7. Then
376 18 Simple Trace Formulae
Z X
Φ(xh)χ(h)dh
AG ∩AH H(F )\H(AF ) x∈X(F )
X Z X (18.23)
= fx (h−1
1 γh2 )χ(h2 )dh1 dh2 .
x∈X(F )/G(F ) AHx ,χ Hx (F )\Hx (AF ) γ∈G(F )
Thus (18.21) is a sum over x ∈ X(F )/G(F ) of geometric sides (or spectral
sides) of relative trace formula for the various Hx . Studying (18.21) amounts
to studying distinction of automorphic representations by (Hx , χ) as x varies.
In particular, if X is homogeneous, the Hx are all inner forms of eachother.
This method of packaging distinction problems together has proven crucial,
especially in work on the Gan-Gross-Prasad conjecture (see §14.7). We point
out that even when it is not possible to choose fx as in (18.22) it is still
possible to expand the function
X
Φ(xg) ∈ C ∞ ([G])
x∈X(F )
spectrally. More or less it will give something like the right hand side of
(18.23), but the fx involved will not necessarily be compactly supported.
In some sense, this is the key point. Replacing G by X has the effect
of changing the function space S(X(AF )). As an example, let E/F be a
quadratic extension and let G = ResE/F GLn and let X be the space of
Hermitian matrices with respect to E/F . Thus X(F ) is the space of n × n
matrices x with coefficients in E such that θ(x) = xt , where θ is a generator
of Gal(E/F ). Then there is a natural right action
Given two nonzero integers m and n and a positive integer c, the classical
Kloosterman sum is defined as
18.7 The Petersson-Bruggeman-Kuznetsov formula 377
X
K(m, n; c) := e2πi(am+an)/c ,
1≤a≤c
(a,c)=1
for suitable φ ∈ C ∞ (R>0 ). We refer to [IK04, Chapter 16] for more details.
There are many other analytic number theory papers that use the formula,
for example [Iwa02, GS83]. It has also played a role in work on Langlands’
beyond endoscopy proposal [Her11, Her12, Her16, Sar, Ven04].
Following Jacquet, we will derive the Kuznetsov formula as a relative trace
formula (see [KL08, KL13] for a nice exposition for GL2 , and [Ye00] for a
slightly different take on the higher rank case). The beautiful fact about this
trace formula is that the general simple relative trace formula of Theorem
18.2.2 (together with a small additional argument for the spectral side) can
be used to derive it in complete generality, with no additional assumptions on
the test functions involved. We state this precisely in Theorem 18.9.2 below.
In this sense it is analytically the simplest trace formula. The Petersson-
Kuznetsov formula remains an important tool in analytic number theory and
has not been fully exploited in higher rank (see, e.g. [BB15] for recent results
for GL3 ). We have therefore structured our treatment of the formula with
applications to analytic number theory in mind.
We now set notation. Let N := Nr ≤ GLr be the unipotent radical of
the Borel B := Br of upper triangular matrices and let T := Tr ≤ B be the
maximal torus of diagonal matrices. For this trace formula we take
H := N × N ≤ GLr × GLr .
GLr × N × N −→ GLr
Nγ := Hγ ≤ N × N.
deti : GLr −→ Ga
given on points by
deti ( A∗i ∗∗ ) = det(Ai )
are invariant under the action of N × N for 1 ≤ i ≤ r. Here Ai is an i × i
matrix.
For every representative w ∈ GLr (F ) of the Weyl group we define a mor-
phism (of schemes, not group schemes)
by
cr /cr−1
cr−1 /cr−2
Υw (c1 , . . . , cr ) =
.. w.
(18.26)
.
c2 /c1
c1
be (the standard representative of) the long Weyl element. The orbit of w0
is the unique open Bruhat cell [Mil17, Theorem 21.73]. Since it is possible to
choose test functions supported in this cell at a given place, elements of the
open Bruhat cell play a key role.
The map Υw0 gives a useful description of the open Bruhat cell:
Lemma 18.7.2 The Υw0 induces a bijection
Proof. It is clear that any element of N (F )\B(F )w0 B(F )/N (F ) is in the
image of the map Υw0 . On the other hand the partial determinants deti are
all N × N -invariant and
where ci is the ith entry of c and the sign ± depends only on i and r. Thus
Υw0 is injective. Thus Υw0 induces a bijection as claimed.
For the claim on triviality of stabilizers we can assume γ = tw0 for t ∈
T (F ) by what we have already proven. Then if (n1 , n2 ) ∈ Nγ (F ) one has
The left hand side is in the unipotent radical of the Borel opposite to B, so
both sides must equal to I. t
u
via
Qr−1
i=1 mi
..
.
ν(m) := . (18.29)
mr−2 mr−1
mr−1
1
We note that
1 m−1
1 x1 ∗ 1 x1 ∗
.. .. . .. . ..
ν(m)−1
. . ν(m) =
.
(18.30)
1 xr−1 1 m−1
r−1 xr−1
1 1
ψ : F \AF → C× .
For f ∈ Cc∞ (AGLr \GLr (AF )) and γ ∈ GLr (F ), we define the orbital
integral
Z
ROγ (f, m1 , m2 ) := f (n−1
1 γn2 )ψm1 (n1 )ψ m2 (n2 )dn1 dn2 .
Nγ (AF )\N (AF )2
(18.32)
is convergent and
Z
Kf (n1 , n2 )ψm1 (n1 )ψ m2 (n2 )dn1 dn2
[N ]2
0
X
= τ (Nγ )ROγ (f, m1 , m2 )
γ∈N (F )\GLr (F )/N (F )
382 18 Simple Trace Formulae
where the sum is over relevant classes. Here, as usual, [N ] := N (F )\N (AF ).
By Lemma 18.7.1 every γ is relatively unimodular, relatively elliptic, and
relatively semisimple. Thus in the number field case we could invoke Theorem
18.2.2 to prove this proposition. In the function field case we do not know
if every γ ∈ GLr (F ) is gcf, and this is the obstruction to simply invoking
Theorem 18.2.2 in the function field case. Fortunately the gcf assumption
was just used to establish convergence in the proof of Theorem 17.7.2 and in
the current setting we can just proceed directly:
Proof. This follows from the same manipulations used in the proof of The-
orem 18.2.2. The only point one must check is that these manipulations are
justified. This follows from the observation that
Z X
|f (n−1
1 γn2 )|dn1 dn2
[N ]2 γ∈G(F )
Proof. This follows from the fact that the partial determinants deti are in-
variant under the action of N × N on GLr . t
u
Let f ∨ (g) := f (g −1 ).
where the last equality is, by definition, ROΥw (c)−1 (f ∨ , −m2 , −m1 ). t
u
18.8 Kloosterman integrals 383
It turns out that for suitable c the integral ROΥw (c) (f, m1 , m2 ) defined in
(18.32) can be expressed as the product of Kloosterman sums, see Propo-
sition 18.8.6 below. To state this, we first prove a recursive relation for
ROΥw (c) (f, m1 , m2 ). To ease on notation, for m1 , m2 ∈ (F × )r−1 , we let
Then under suitable assumptions the local orbital integral of (18.32) on GLr
can be written as that of GLr−1 in the following way:
Lemma 18.8.5 Assume that c = (c1 , c2 , . . . , cr ) ∈ (OF ∩ F × )r and that
c1 , cr ∈ OF× . Then one has that
where w
e0 is the long Weyl element for GLr−1 ,
and m e 2 ∈ (F × )r−2 are obtained by removing the last entry m1(r−1) from
e 1, m
m1 and by removing the first entry m21 from m2 , respectively.
Proof. To ease notation, let C be the (r − 1) × (r − 1) matrix given by
cr /cr−1
.
C := .. .
c2 /c1
Here
e = c1 α2 α1t + n1 Cn2
C
is an (r−1)×(r−1) matrix, c1 α1 is a column vector, and c1 α2t is a row vector.
Since we assumed that c1 ∈ OF× and cr ∈ OF× , this matrix is in GLr (OF ) if
and only if α1 , α2 ∈ OFr−1 and n1 Cn2 ∈ GLr−1 (OF ). The lemma follows. t u
Proposition 18.8.6 Assume that c = (c1 , . . . , cr ) ∈ OFr ∩ (F × )r where cr ∈
OF× , cj 6∈ OF× , and ci ∈ OF× for all i 6= j. Then
Proof. By Lemma 18.8.5 iterated j−1 times we have that ROΥw0 (c) (1GLr (OF ) , m1 , m2 )
vanishes unless m1(r−1) , . . . , m1(r−j) , m21 , . . . , m2(j−1) ∈ OF . Here m1i and
m2i are the ith entry of m1 and m2 respectively. Assuming this is the case,
the same lemma implies that
Here w
e0 is the long Weyl element of GLr−j , and
in which case it is
X 0
X
rtr(π(f ), m1 , m2 ) = τ (Nγ )ROγ (f, m1 , m2 ) (18.35)
π γ∈N (F )\GLr (F )/N (F )
where the sum on the left is over isomorphism classes of cuspidal automorphic
representations of AGLr \GLr (AF ) and the sum on the right is over relevant
classes γ.
aP := Lie AP .
Moreover
X XZ
n−1
P rtr(I(σ, λ)(f ), m1 , m2 )dλ
P σ iaP
0
X
= τ (Nγ )ROγ (f, m1 , m2 ),
γ∈N (F )\GLr (F )/N (F )
where the prime indicates that the sum on the right is over relevant γ.
Proof. The absolute convergence statement follows from the fact that [N ] is
compact together with the estimates on the terms in the spectral expansion
(18.36) contained in [Art78, §4]. Thus by (18.36) we conclude that
18.9 Sums of Whittaker coefficients 387
Z
Kf (n1 , n2 )ψm1 (n1 )ψ m2 (n2 )dn1 dn2
[N ]2
X XZ
= n−1P rtr(I(σ, λ)(f ), m1 , m2 )dλ.
P σ iaP
v6∈S
fm1S ,m2S (g) := f (ν(m1S )gν(m2 )−1 ) ∈ Cc∞ (AGLr \GLr (AF )) (18.38)
fe := fm1S ,m2S .
We change variables
rtr(π(fe), m1 , m2 )
Z
= KI(σ,λ)(fe) (ν(m1 )−1 n1 ν(m1 ), ν(m2 )−1 n2 ν(m2 ))ψ1 (n1 )ψ 1 (n2 )dn1 dn2
[N ]2
Z
= KI(σ,λ)(fe) (n1 ν(m1 ), n2 ν(m2 ))ψ(n1 )ψ(n2 )dn1 dn2 .
[N ]2
Here we have used the fact that KI(σ,λ)(fe) (g1 , g2 ) is left invariant under
G(F ) × G(F ). This in turn is equal to
Z
KI(σ,λ)(f ) (n1 ν(mS1 ), n2 ν(mS2 ))ψ(n1 )ψ(n2 )dn1 dn2 . (18.39)
[N ]2
The function
Z
(g1 , g2 ) 7−→ KI(σ,λ)(f ) (n1 g1 , n2 g2 )ψ(n1 )ψ(n2 )dn1 dn2
[N ]2
is a global Whittaker function for I(σ, λ)⊗I(σ, λ)∨ for the character (n1 , n2 ) 7→
ψ(n1 )ψ(n2 ) that is unramified outside S. Thus the lemma follows from
uniqueness of Whittaker models in Theorem 11.3.1. t
u
Exercises
For the next three problems we assume that we are in the setting of §18.2.
Let v be a finite place of the global field F and let
18.2. Prove that if fv is supported in the set of elements of G(Fv ) with closed
Hr -orbit and f (x) 6= 0 for some x ∈ Ω then γ is relatively semisimple.
{γ 0 ∈ G(Fv ) : Hγ 0 = Hγ◦0 }
only if
|cr det(ν(m1 m−1
2 ))|v f Y
|deti (ν(m1 )Υw (c)ν(m2 )−1 )|v f |cr det(ν(m1 m−1 i/r
2 ))|v
unless
18.6. Let H be a unipotent algebraic group over a local field F . Prove that
H(F ) is unimodular.
18.7. Let X be the space of n × n matrices, equipped with the natural action
of GLn on the right. Let Φ ∈ Cc∞ (X(AF )/AGLn ). Prove that the restriction
of Φ to GLn (AF )/AGLn is smooth but not compactly supported.
Chapter 19
Applications of Trace Formulae
Selberg’s motivation for proving the trace formula for SL2 over Q was to
prove the existence of cusp form that are fixed by SO2 (R). In the process he
actually estimated the number of such cusp forms in a suitable family and
showed that they exist in abundance. This was the first application of the
trace formula, and it can be viewed as a (quantifiable) existence result. We
discuss this result in §19.2.
Motivated by the Langlands functoriality conjecture, Jacquet and Lang-
lands [JL70] gave a very different type of application of the trace formula.
They compared two different trace formulae on different groups in order to de-
duce a relation between automorphic representations on the two groups. This
idea has been extended to what is known as the theory of twisted endoscopy
and continues to play a key role in automorphic representation theory. We
discuss this in the second part of the chapter. All sections of the chapter are
more or less brief surveys. Our main aim is merely to whet the appetite of
the reader and give them some guidance on further reading.
391
392 19 Applications of Trace Formulae
vol(M )
N (X) = X dim M + o(X dim M ) (19.1)
(4π)dim M/2 Γ ( dim2 M + 1)
as X → ∞. t
u
Consider the Shimura manifolds Sh(G, X)K of §15.2. These are finite unions
of Riemannian manifolds. When they are compact the Weyl law implies the
existence of an abundance of automorphic forms on them. However they are
often noncompact (see Theorem 2.6.2 for a precise statement). Even if one
compactifies Sh(G, X)K , it is still unclear in the noncompact case whether
the eigenforms produced by the Weyl law correspond to the continuous or
cuspidal spectrum. Selberg developed the trace formula for SL2 (over Q) to
address precisely this question [Sel56], and he proved an analogue of (19.2.1)
in this setting.
Müller [M0̈7] established the analogue of the Weyl law for G = SL(n) over
Q. His proof makes use of the full noninvariant trace formula of Arthur, and
can treat cuspforms that are not necessarily fixed by On (R). About the same
time Lindenstrauss and Venkatesh [LV07] proved an analogous result for all
split semisimple adjoint groups over Q. Notably, their method is soft in that
it does not require the full trace formula, only a simple analogue. It is based
on a novel construction of test functions with purely cuspidal image. Their
method currently only treats cuspforms that are fixed by a maximal compact
subgroup at infinity, but it is likely that this can be removed.
Using the same technique Getz and Hahn [GH15] established a relative
analogue of the Weyl law under some assumptions via the general simple
relative trace formula of Theorem 18.2.2. This relative Weyl law gives an
asymptotic formula for the number of Laplacian eigenfunctions on Sh(G, X)K
19.3 The comparison strategy 393
The Langlands’ functoriality conjecture predicts that for two reductive groups
G1 and G1 defined over a global field F and a L-map
r : L G1 −→ L G2 ,
where Hiu is unipotent and Hir is reductive. Let CGi be the centralizer of
Gi in Hi , and assume that ACi ≤ Ai , where Ai is the intersection of AGi ,Hi
and the diagonal copy of AGi . See §18.2 for this notation. Suppose that we
can construct a correspondence
For example, one might have a bijection, but often the situation is more
complicated. We refer to this as a geometric correspondence between
relevant relative classes. We say that the class of δ matches the class of γ
if the two classes correspond. Suppose that for matching δ and γ one can
relate the orbital integrals τ (G2δ )ROδ (f ) and τ (G1γ )ROγ (f G1 ) for suitable
394 19 Applications of Trace Formulae
test functions f ∈ Cc∞ (AG2 \G2 (AF )) and f G1 ∈ Cc∞ (AG1 \G1 (AF )). Then
one can hope to prove an identity of the form
X X
τ (G2δ )ROδ (f ) = τ (G1γ )ROγ (f G1 ). (19.2)
δ∈G2 (F )/H2 (F ) γ∈G1 (F )/H1 (F )
Here the sums over γ and δ are over χ1 and χ2 -relevant elements, respec-
tively, that are additionally relatively regular, relatively elliptic, relatively
unimodular and such that the H1r -orbit of γ (resp. the H2r -orbit of δ) is
closed. We are assuming these things so that sums are convergent by The-
orem 18.2.2. Assuming that an identity like (19.2) can be established for a
sufficiently rich set of test functions one can obtain a relationship between
(H1 , χ1 )-distinguished representations of G1 (AF ) × G1 (AF ) and (H2 , χ2 )-
distinguished representations of G2 (AF ) × G2 (AF ) and thereby deduce an
instance of Langlands functoriality (or its relative analogue due to Jacquet,
Sakellaridis, and Venkatesh). We will explain this in more detail momentar-
ily. Let us point out that in many cases one has to incorporate families of
(Gi , Hi , χi ) on both sides of the formula. We ignore this complication because
we are just speaking in rough terms at the moment.
Let us elaborate what we mean by a relation between τ (G2δ )ROδ (f ) and
τ (G1γ )ROγ (f G1 ). We assume for simplicity that the geometric correspon-
dence is a bijection, though this is rarely the case. One needs to be able
to relate the volumes τ (G1γ ) and τ (G2δ ) for matching γ and δ, but this is
usually not the key issue.
For almost all finite places v there exists a hyperspecial subgroup Kiv ≤
Gi (Fv ) by Proposition 2.4.3. The map
r : L G1 −→ L G2
for all f ∈ Cc∞ (G2 (Fv ) // K2v ). This is often too naı̈ve. The equality will
only hold up to an explicit function of δv and γv called a transfer factor,
but we suppress this difficulty. Formulae of the form (19.4) are known as
the fundamental lemma for the Hecke algebra. The identity (19.4)
for f = 1K2v is known as the fundamental lemma, or the fundamental
lemma for the unit element of the Hecke algebra. In practice one
deduces the fundamental lemma for the Hecke algebra from the fundamental
19.3 The comparison strategy 395
lemma. Finally one needs to know that for all places v there sufficiently large
supply of functions f ∈ Cc∞ (G2 (Fv )) and f G1 ∈ Cc∞ (G1 (Fv )) that match,
meaning that when δ and γ match one has
π0 π
(see Exercise 19.3). Though the set of representations on each side of this
equation is in general infinite, in practice one can use a version of linear
independence of characters to refine it to an identity
π 0 :π 0S ∼
=r(π0S )
(19.7)
tr π S (f G1 S )rtrH1 ,χ1 (π(fSG1 1K2S ))
X
=
π:π S ∼
=π0S
In this section we describe some settings where the general strategy in §19.3
has been successfully applied. They are related to the trace formula and the
twisted trace formula of §18.4 and §18.5. We assume for the remainder of
the chapter that F is a number field because most of the results we will cite
below are proven in this setting (though it seems likely that the proofs would
generalize).
We start by taking G1 to be an inner form of the quasi-split reductive
group G2 over F . We assume that Gder 2 is simply connected so that we can
apply Theorem 19.4.1 below. Then L G1 = L G2 and we take
r : L G1 −→ L G2 (19.8)
G1 /H1 −→G
˜ 2 /H2 . (19.9)
This is almost the first step of the procedure explained in the previous section.
There is a slight wrinkle. Consider the map
This turns out to be enough. In the current setting, the fundamental lemma
for the Hecke algebra is trivial because G1Fv and G2Fv are isomorphic for
almost every place v (see Exercise 19.5). Thus proving transfer of automorphic
representations is reduced to understanding transfer of functions. Jacquet and
Langlands [JL70] were the first to introduce an argument of this type and
they used it to give a complete treatment of the special case where G1 is the
unit group of a quaternion algebra and G2 = GL2 . Because of this transfers
attached to the maps (19.8) are known as Jacquet-Langlands transfers.
Though the full theory of Jacquet-Langlands transfers is still not known,
many results exist in the literature. In particular the theory for GLn is fairly
complete. Let D be a simple algebra of dimension d2 over F . For F -algebras
R let
GLnD (R) := (D ⊗F Mn (R))× .
Let S be the (finite) set of places of F where Dv 6∼
= Md (Fv ). Here Mn (Fv ) is
the space of n × n matrices with entries in Fv .
For every place v we have a diagram
t
GLnD / ∼ (F ) −−−−→ GLnd / ∼ (F )
where the quotients on the bottom are the GIT quotients of GLnD and GLnd
by the conjugation action of GLnD and GLnd , respectively, and the verti-
cal maps are the projections. The map p is surjective (Exercise 19.6), but
pD is not. We say that δ and γ match if t(pD (δ)) = p(γ). An irreducible
admissible representation π 0 of GLn (Fv ) is Dv -compatible if the character
Θπv (γ) is nonzero for some regular semisimple γ ∈ GLnd (Fv ) that matches
a δ ∈ GLnD (Fv ). See Theorem 8.5.1 for the definition of Θπ in the nonar-
chimedean case, the archimedean case is similar [Kna86, Theorem 10.25]. An
automorphic representation π 0 of AGLn \GLn (AF ) is D-compatible if πv is
Dv -compatible for all v ∈ S.
The following theorem is due to Badulescu and Grbac [Bad08] following
several previous works [Rog83, BDKV84], going back to [JL70], where the
n = 2 case was treated.
Theorem 19.4.2 (Jacquet-Langlands correspondence) There exists a
unique injection r from the set of isomorphism classes of discrete automorphic
representations of AGLnD \GLnD (AF ) to the set of isomorphism classes of
discrete automorphic representations of AGLnd \GLnd (AF ) such that r(π)v ∼ =
πv for v 6∈ S. The image consists of discrete automorphic representations π 0
that are D-compatible. t
u
Here we are using the isomorphism GLnD (Fv ) ∼
= GLnd (Fv ) for v 6∈ S to make
sense of the assertion that r(π)v ∼
= πv0 .
398 19 Applications of Trace Formulae
In the setting above the groups G1 and G2 were not that different from
eachother; indeed, they become isomorphic over the algebraic closure. Saito
and Shintani [Sai75, Shi79] discovered the first geometric correspondence in-
volving groups that are not isomorphic over the algebraic closure. It was fully
developed for GL2 by Langlands in [Lan80]. Take G1 = G to be reductive
with simply connected derived group, let E/F be a cyclic field extension, and
let θ be a generator of Gal(E/F ). We then take G2 = ResE/F G, and let
r : L G −→ L ResE/F G
be the map given by the diagonal embedding on the neutral component and
the identity on the Galois factor. To set up our geometric correspondence we
let the χi be trivial, let H1 be the diagonal copy of G in G × G, and let H2
be the θ-twisted diagonal defined by
for F -algebras R. Thus G(F )/H1 (F ) is the space of conjugacy classes and
ResE/F G(F )/H2 (F ) is the space of θ-conjugacy classes in the sense of §18.5.
The geometric correspondence is defined as follows. A reference is [Kot82].
Let δ ∈ ResE/F G(F ) and let
θ(N(δ)) = δN(δ)δ −1
which implies that the G(E)-conjugacy class of N(δ) is fixed under Gal(E/F ).
Thus there is an element of G(F ) in the conjugacy class by Theorem 19.4.1.
We call such an element a norm of δ. This gives us a map
archive of his work online. Some of the fruits of his work are described in
§13.5 and §13.8. We will also not give an overview of the “long march” (in
the words of Ngô, [Ngô10a]) to the fundamental lemma, preferring to leave
this to the survey [Ngô10a] written by the person who completed the proof
in his Fields medal-winning work.
We would be remiss to point out that despite the amazing success of
twisted endoscopy, it has obvious drawbacks. As currently understood it can
only hope to establish cases of functoriality where the image of the functorial
lift consists of representations that are isomorphic to their conjugates under a
single automorphism. A generalization to two automorphisms is proposed in
[Get12, Get14]. These papers are based on interconnected ideas of Braverman,
Kahzdan, L. Lafforgue Langlands, Ngô, and Sakellaridis [BK00, Laf14, Ngô14,
Sak12].
The theory of twisted endoscopy discussed in the previous section has reached
a degree of maturity. One understands how to compare the twisted trace
formulae and the trace formula. Since Jacquet has given us the new tool of
the relative trace formula, we now have a much broader arena to apply the
experience gained in the theory of twisted endoscopy and prove new results.
We mention a few geometric correspondences here and connect them to the
results on distinction reviewed in Chapter 14.
We start by explaining the comparison underlying Theorem 14.5.2 which
characterizes those representations of ResE/F GLn that are distinguished by
a quasi-split unitary group. Here E/F is a quadratic extension. In this case
our G1 = GLn and G2 = ResE/F GLn and
is the base change map, given by the diagonal embedding on the neutral
component and the identity on the Galois factor.
In this case our basic outline from §19.3 must be modified along the lines
of §18.6. Thus for F -algebras R we let
X(R) := {x ∈ Mn (E ⊗F R) : θ(x) = xt }
χ2 : ResE/F Nn (AF ) → C×
u 7−→ ψ(uΘ(u)),
H1 (R) := {(u−t
1 , u2 ) : u1 , u2 ∈ Nn (R)}
T (F )
p2 p1
w &
X(F )/ResE/F Nn (F ) G1 (F )/Nn2 (F )
where the two arrows send an element of T (F ) to its class. We say that
a χ2 -relevant δ and and χ1 -relevant γ match if there is a t ∈ T (F ) such
that δ ∈ p2 (t) and γ ∈ p1 (t). In this setting the fundamental lemma was
conjectured by Jacquet and Ye [JY96] and proven in two different manners
by Ngô [Ngô99b, Ngô99a] and Jacquet [Jac04, Jac05b].
Jacquet and Rallis [JR11] later proposed a related comparison that even-
tually led to the proof of the results on the Gan-Gross-Prasad conjecture
discussed in §14.7. A useful reference is [Cha19]. Let E/F be a quadratic
extension of number fields let Un be the quasi-split unitary group of §13.7.
We take
and let
EM
E L M
θ
σ τ
where ResE/F GLn acts via τ -conjugation on S and Un acts via conjugation
on Q. This is formally similar to the geometric setting underlying base change
discussed in §19.4. One has a set of τ -conjugacy classes and a set of conjugacy
classes that one wishes to relate. For a relatively regular relatively semisimple
δ ∈ ResEM/F GLn (F ) we say that γ is a norm of δ if there is an h ∈ H2 (F )
such that
hγγ −σ h−1 = δδ −θ (δδ −θ ).
404 19 Applications of Trace Formulae
One can show that every relatively regular relatively semisimple δ admits a
norm, and we say that
and
γ ∈ G1 (F ) = ResE/F (UnM/F )E (F ) = UnEM/E (AE )
match if γ is a norm of δ. In this context the fundamental lemma can be
obtained from the fundamental lemma for base change, and transfer for a
sufficiently rich supply of test functions was proven in [GW14]. This provides
enough information to establish Theorem 14.6.1. However there is a relative
analogue of twisted endoscopy lurking here that has yet to be developed.
This theory of twisted relative endoscopy will probably be necessary to ob-
tain a version of Theorem 14.6.1 valid for arbitrary cuspidal automorphic
representations of UnEM/E (AE ).
Exercises
r : L G1 −→ L G2
tr r(π)(f ) = tr π(r(f ))
for all f ∈ Cc∞ (G2 (F ) // K2 ). Here we normalize the Haar measures implicit in
the traces so that the Ki are given measure 1. Observe that r(1K2v ) = 1K1v .
19.3. Let G be a reductive group over the global field F and let H ≤ G × G
and χ : AG,H \H(AF ) → C× be as in §18.2. Let S be a finite set of places of
F including the infinite places and let K S < G(ASF ) be a maximal compact
open subgroup such that Kv is hyperspecial for v 6∈ S. Let
Prove that
rtr(π(f )) = tr π(f S )rtr(π(fS 1K S )).
19.5. Let G1 and G2 be reductive groups over a global field F that are inner
forms of each other. Prove that G1Fv ∼
= G2Fv for all but finitely many places
v of F .
19.7. Let T ≤ GLn be a maximal torus over a field k. Prove that D(k, T, GLn ),
defined as in (17.28), has only one element.
Appendix A
The Iwasawa Decomposition
A.1 Introduction
G(F ) = P (F )K (A.1)
holds and
407
408 A The Iwasawa Decomposition
s := Spec(L) → Spec(k)
where L is the algebraic closure of a field L. Recall from §1.2 that to give
such a morphism is equivalent to giving a morphism of k-algebras k → L.
A geometric fiber of a k-scheme X is the base change Xk where s is a
geometric point of Spec(k)
G(k) −→ (G/P)(k)
Lemma A.2.2 Let G be a smooth affine group scheme of finite type over a
discrete valuation ring O with finite residue field κ and let X be a G-torsor.
If the special fiber Gκ is connected then
X (O) 6= ∅.
Proof. Suppose that g ∈ G(ASF ). Then g ∈ G(OFv ) for all but finitely many
places v, so it suffices to show that for every v 6∈ S one has
The scheme G/P is projective over OFv , hence proper, and thus the arrow
b is a bijection by the the valuative criterion of properness. Thus it suffices
to show that that the arrow a is surjective. The fiber of a over any point of
G/P(O) is a P-torsor, and hence has a point by Lemma A.2.2. t
u
P =MnN (A.4)
λ : Gm −→ G.
A.3 The dynamic method 411
λ : Gm × G −→ G
given on points by
λ(t, g) := λ(t)gλ(t)−1 .
We say that limt→0 λ(t, g) exists if the morphism λ extends to a morphism
λ : Ga × G −→ G.
Then
a t 2 b t3 c
ab c
λ(t) d e f λ(t)−1 = t−2 d e tf .
hg i t−3 g t−1 h i
Thus limt→0 λ(t, g) exists if and only if g is a point of the Borel subgroup of
upper triangular matrices. Moreover limt→0 λ(t, g) = 1, the identity in GL3 ,
if and only if g lies in the group of strictly upper triangular matrices, and g
commutes with the action of λ if and only if g is a diagonal matrix.
Let P (λ) be the group functor on k-algebras given on a k-algebra R by
We also define
and let Z(λ)(R) ≤ G(R) be the subgroup that commute with the action λ.
We then have
Proof. The fact that every P is a P (λ) is [Mil17, Theorem 25.1] and N =
N (λ) for any cocharacter λ such that P = P (λ). Since all Levi subgroups
of P are conjugate under P (k) we deduce that upon conjugating λ by an
element of P (k) we may assume that M = M (λ). t
u
We now complete the proof of Theorem A.1.1. In view of Theorem A.2.3 and
Proposition A.3.1 it suffices to prove Theorem A.1.1 in the local case. Thus
we assume for the remainder of the section that F is a local field.
We will actually prove something more precise than what is required by
Theorem A.1.1. Fix a minimal parabolic subgroup P0 ≤ G and a Levi sub-
group M0 ≤ P0 . We write N0 for the unipotent radical of P0 . We call a
pair (P, M ) consisting of a parabolic subgroup P ≤ G and a Levi subgroup
M ≤ P a standard parabolic pair if P0 ≤ P and M0 ≤ M .
A.4 Proof of Theorem A.1.1 413
where Aff(Ae (G, S)) is the group of affine transformations of Ae (G, S).
B e (M ) −→ B e (G)
Proof. We first check that a Borel subgroup exists. The functor assigning
to an OF scheme S the set of Borel subgroups of GS is representable by a
smooth proper OF -scheme, say X [DG74, XXII, Corollaire 5.8.3(i)]. We have
X (F ) = X (OF ) by the valuative criterion of properness, so there is a Borel
subgroup B ≤ G. The existence of T is [DG74, XXII, Corollaire 5.9.7] t
u
B =T oN
Abstract In this appendix we define the adelic version of the Fourier trans-
form on a vector space and state the corresponding version of Poisson sum-
mation
ψR : R −→ C×
a 7−→ e−2πia
and
ψQp : Qp −→ C×
a 7−→ e2πipr(a) .
Here the principal part pr(a) ∈ Z[p−1 ] is any element such that
a ≡ pr(a) (mod Zp ).
415
416 B Poisson Summation
ψF := ψ ◦ trF/F0 : F −→ C× .
Here the ai are in k (or rather a fixed set of representatives for k = Fp [t]/$ in
Fp [t]). On the right we are implicitly identifying trk/Fp (a−1 ) with an element
of Z mapping to it under the quotient map Z → Fp . If F is a characteristic
p local field then it is an extension of a completion F0 of Fp (t), and we let
ψF := ψ ◦ trF/F0 : F −→ C× .
One checks that all of the ψF we have defined are indeed nontrivial characters.
It is not hard to check the following lemma (see Exercise B.2):
Lemma B.1.1 If F is a local field and ψ : F → C× is any nontrivial char-
acter then there is a bijection
F −→ Fb
α 7−→ (x 7→ ψ(αx)) .
t
u
Here the right hand side is the unitary dual of F , which is simply the set of
characters of F , i.e. the set of continuous homomorphisms F → C× .
If F is a global field we let
Y
ψF := ψF v . (B.1)
v
F \AF −→ F\
\AF
α 7−→ (x 7→ ψ(αx)) .
t
u
Again the right hand side is the unitary dual of F \AF , which is simply the
set of characters of F \AF , i.e. continuous homorphisms F \AF → C× .
h , i : W × W −→ F (B.2)
Let F be a global field and let W be a finite dimensional F -vector space. For
an F -algebra R let W (R) := W ⊗F R. We define
Let S be a finite set of places of F including all the infinite places and all
places where Fv is ramified. If S is large enough then
f = fS 1(Ob S )r .
F
We already know that fbS ∈ S(W (F∞ )) ⊗ S(W (FS∞ )), so it suffices to check
that 1
b b S r = 1 b S . This is the content of Exercise B.7.
(O ) O t
u
F F
t
u
B.3 Global Schwartz spaces and Poisson summation 419
Exercises
B.1. Let F be a local field. Prove that the maps ψF defined in §B.1 are
characters.
B.5. Let F be a nonarchimedean local field and let f ∈ Cc∞ (F ). Prove that
fb ∈ Cc∞ (F ). Here fb can be defined with respect to any Haar measure on F
and nontrivial character ψ : F → C× .
B.6. Prove that the Fourier inversion formula (B.3) holds when F is a nonar-
chimedean local field.
B.7. Let F be a nonarchimedean local field that is unramified over its prime
field. Show that
1bOF = 1OF
where the Fourier transform is defined with respect to the standard additive
character ψF and the Haar measure on F such that dy(OF ) = 1. Conclude
that this Haar measure is self-dual with respect to ψF and the pairing (x, y) 7→
xy.
Hints to selected exercises
Chapter 1
1.12 Use the fact that parabolic subgroups would have unipotent elements.
This would imply that B has zero divisors, contradicting the assumption that
it is a division algebra.
Chapter 2
2.10 Note that GLn (FS ) is an open subset of gln (FS ) for any finite set of
places S and use weak approximation for AF (see Theorem 2.1.4).
2.6 One approach uses the alternate definition of group schemes described
in Exercise 1.7.
2.14 In the archimedian case use the Gram-Schmidt process. In the nonar-
chimedian case one can proceed in an elementary manner using induction on
n. See [Bum97, Proposition 4.5.2].
2.15 Using the description of GL2 (Q)\GL2 (AQ )/K0 (1) given at the end of
§2.6 the exercise can be deduced from the following well-known fact: Every
element of the complex upper half plane is SL2 (Z)-equivalent (under Möbius
transformations) to an element z satifying −1 1
2 ≤ Re(z) ≤ 2 and |z| > 1.
Chapter 3
421
422 Hints to selected exercises
3.3 See [Fol95, §2.4], bearing in mind that δG (g) = ∆(g)−1 in the notation
of loc. cit.
3.13 Prove first that V admits a G-invariant inner product.
Chapter 4
Chapter 8
5.5 Show that V has a countable basis and invoke Lemma 4.1.
5.10 See [Car79].
5.11 To show that Vsm ≤ V is dense mimic the proof of Proposition 4.2.3.
For the other claims one uses the following fact: if X, Y are topological vector
spaces, Y is an F -space, M ⊆ X is a dense subspace, and λ : M → Y is a
continuous map, then Λ admits a (unique) continuous extension to X → Y
(see [Rud91, Chapter 1, Exercise 19]).
5.15 Use the fact that all hyperspecial subgroups are conjugate under
G/ZG (F ) (see §2.4).
5.16 Use the fact that all hyperspecial subgroups are conjugate under
G/ZG (F ) (see §2.4).
Chapter 6
Chapter 8
8.5 Right exactness is the only point that is not obvious. Suppose that
one has a surjection A : V → W of smooth representations of M (F ). Let
ϕ : G → W be an element of IndG
P (W ), and choose a compact open subgroup
K ≤ G(F ) fixing ϕ. By Lemma 8.2.1 for all g ∈ G(F ) one has
−1
ϕ(g) ∈ V M (F )∩gKg .
−1 −1 −1
Since V M (F )∩gKg → W M (F )∩gKg is surjective there is a vg ∈ V M (F )∩gKg
such that a(vg ) = ϕ(g). Fix a minimal set of representatives X for P (F )\G(F )/K.
We then define
e : G −→ V
ϕ
nmxk 7−→ m.vx
0 −→ V1 −→ V2 −→ V3 −→ 0
which implies
1 X
ϕ= 0 ] (π(n)ϕ − ϕ).
[KN : KN 0
n∈KN /KN
Chapter 9
Chapter 10
Chapter 11
g 7−→ λ(π(g)ϕ).
Chapter 12
12.10 Assume first that the πi are all square-integrable. Then since all the πi
are tempered none of them are linked in the sense of §8.4. Hence we conclude
by Theorem 8.4.4. The general case can be reduced to this case by Theorem
8.4.5 and the transitivity of induction.
Chapter 15
15.4 Let
K(N ) := {g ∈ GLn (Z)
b : g ≡ In (mod N Z)}.
Then GLn (Q) ∩ K(N ) is the group in the proof of Lemma 15.2.1 and hence
is neat. Now observe that the set of eigenvalues of matrices in GLn (Q) ∩
g −1 K(N )g is independent of g ∈ GLn (A∞
Q ).
Chapter 17
Chapter 18
18.1 Since G(AF ) = AG G(AF )1 by §2.6 and G(AF )1 ≤ G(AF )0 the map
1 = χ(ah` g1 h−1
r )χ(g2 )
−1
= χ(a)χ(h` h−1
r )χ(g1 )χ(g2 )
−1
= χ(a)
By the duality between tori and their character groups (Theorem 1.7.1) this
implies that a ∈ a(1, AH ), which is to say that a = h0−1 0 0 0
` hr for some (h` , hr ) ∈
AH .
Index
427
428 Index
[AC89] James Arthur and Laurent Clozel. Simple algebras, base change,
and the advanced theory of the trace formula, volume 120 of An-
nals of Mathematics Studies. Princeton University Press, Prince-
ton, NJ, 1989. 258, 261, 262, 283
[AG94] Avner Ash and David Ginzburg. p-adic L-functions for GL(2n).
Invent. Math., 116(1-3):27–73, 1994. 285
[AGR93] Avner Ash, David Ginzburg, and Steven Rallis. Vanishing
periods of cusp forms over modular symbols. Math. Ann.,
296(4):709–723, 1993. 279, 293
[Art78] James G. Arthur. A trace formula for reductive groups. I. Terms
associated to classes in G(Q). Duke Math. J., 45(4):911–952,
1978. 114, 199, 322, 386
[Art80] James Arthur. A trace formula for reductive groups. II. Applica-
tions of a truncation operator. Compositio Math., 40(1):87–121,
1980. 114, 199
[Art89] James Arthur. Unipotent automorphic representations: conjec-
tures. Astérisque, (171-172):13–71, 1989. Orbites unipotentes et
représentations, II. 244, 248
[Art90] James Arthur. Unipotent automorphic representations: global
motivation. In Automorphic forms, Shimura varieties, and L-
functions, Vol. I (Ann Arbor, MI, 1988), volume 10 of Perspect.
Math., pages 1–75. Academic Press, Boston, MA, 1990. 244, 248
[Art02] James Arthur. A note on the automorphic Langlands group.
Canad. Math. Bull., 45(4):466–482, 2002. Dedicated to Robert
V. Moody. 246
[Art05] James Arthur. An introduction to the trace formula. In Har-
monic analysis, the trace formula, and Shimura varieties, vol-
ume 4 of Clay Math. Proc., pages 1–263. Amer. Math. Soc.,
Providence, RI, 2005. 195, 198, 199, 200, 201, 323, 356
[Art13] James Arthur. The endoscopic classification of representations,
volume 61 of American Mathematical Society Colloquium Publi-
433
434 References
[BH06] Colin J. Bushnell and Guy Henniart. The local Langlands con-
jecture for GL(2), volume 335 of Grundlehren der Mathematis-
chen Wissenschaften [Fundamental Principles of Mathematical
Sciences]. Springer-Verlag, Berlin, 2006. 151
[BJ79] A. Borel and H. Jacquet. Automorphic forms and automorphic
representations. In Automorphic forms, representations and L-
functions (Proc. Sympos. Pure Math., Oregon State Univ., Cor-
vallis, Ore., 1977), Part 1, Proc. Sympos. Pure Math., XXXIII,
pages 189–207. Amer. Math. Soc., Providence, R.I., 1979. With
a supplement “On the notion of an automorphic representation”
by R. P. Langlands. 204
[BK00] A. Braverman and D. Kazhdan. γ-functions of representations
and lifting. Geom. Funct. Anal., (Special Volume, Part I):237–
278, 2000. With an appendix by V. Vologodsky, GAFA 2000 (Tel
Aviv, 1999). 376, 400
[BK02] Alexander Braverman and David Kazhdan. Normalized inter-
twining operators and nilpotent elements in the Langlands dual
group. volume 2, pages 533–553. 2002. Dedicated to Yuri I.
Manin on the occasion of his 65th birthday. 376
[BLR90] Siegfried Bosch, Werner Lütkebohmert, and Michel Raynaud.
Néron models, volume 21 of Ergebnisse der Mathematik und ihrer
Grenzgebiete (3) [Results in Mathematics and Related Areas (3)].
Springer-Verlag, Berlin, 1990. 8, 353, 409
[BLS96] A. Borel, J.-P. Labesse, and J. Schwermer. On the cuspidal coho-
mology of S-arithmetic subgroups of reductive groups over num-
ber fields. Compositio Math., 102(1):1–40, 1996. 305
[BMS67] H. Bass, J. Milnor, and J.-P. Serre. Solution of the congruence
subgroup problem for SLn (n ≥ 3) and Sp2n (n ≥ 2). Inst. Hautes
Études Sci. Publ. Math., (33):59–137, 1967. 297
[Bor79] A. Borel. Automorphic L-functions. In Automorphic forms, rep-
resentations and L-functions (Proc. Sympos. Pure Math., Ore-
gon State Univ., Corvallis, Ore., 1977), Part 2, Proc. Sympos.
Pure Math., XXXIII, pages 27–61. Amer. Math. Soc., Provi-
dence, R.I., 1979. 134, 136, 137, 139
[Bor91] Armand Borel. Linear algebraic groups, volume 126 of Gradu-
ate Texts in Mathematics. Springer-Verlag, New York, second
edition, 1991. 2, 16, 23, 24, 43, 280
[Bor97] Armand Borel. Automorphic forms on SL2 (R), volume 130 of
Cambridge Tracts in Mathematics. Cambridge University Press,
Cambridge, 1997. 175, 176, 179, 180
[Bor98] Mikhail Borovoi. Abelian Galois cohomology of reductive groups.
Mem. Amer. Math. Soc., 132(626):viii+50, 1998. 341
[Bor07] Armand Borel. Automorphic forms on reductive groups. In Au-
tomorphic forms and applications, volume 12 of IAS/Park City
436 References
Math. Ser., pages 7–39. Amer. Math. Soc., Providence, RI, 2007.
117
[Bor09] M. V. Borovoi. The defect of weak approximation for homoge-
neous spaces. II. Dal0 nevost. Mat. Zh., 9(1-2):15–23, 2009. 43,
44
[BP16] Raphaël Beuzart-Plessis. La conjecture locale de Gross-Prasad
pour les représentations tempérées des groupes unitaires. Mém.
Soc. Math. Fr. (N.S.), (149):vii+191, 2016. 291
[BR94a] Laure Barthel and Dinakar Ramakrishnan. A nonvanishing result
for twists of L-functions of GL(n). Duke Math. J., 74(3):681–700,
1994. 219
[BR94b] Don Blasius and Jonathan D. Rogawski. Zeta functions of
Shimura varieties. In Motives (Seattle, WA, 1991), volume 55
of Proc. Sympos. Pure Math., pages 525–571. Amer. Math. Soc.,
Providence, RI, 1994. 312
[Bru78] R. W. Bruggeman. Fourier coefficients of cusp forms. Invent.
Math., 45(1):1–18, 1978. 377
[BS98] Rolf Berndt and Ralf Schmidt. Elements of the representa-
tion theory of the Jacobi group. Modern Birkhäuser Classics.
Birkhäuser/Springer Basel AG, Basel, 1998. [2011 reprint of the
1998 original] [MR1634977]. x, 63, 81, 182
[BT65] Armand Borel and Jacques Tits. Groupes réductifs. Inst. Hautes
Études Sci. Publ. Math., (27):55–150, 1965. 280, 339
[BT72] F. Bruhat and J. Tits. Groupes réductifs sur un corps local. Inst.
Hautes Études Sci. Publ. Math., (41):5–251, 1972. 413, 414
[BT84] F. Bruhat and J. Tits. Groupes réductifs sur un corps local. II.
Schémas en groupes. Existence d’une donnée radicielle valuée.
Inst. Hautes Études Sci. Publ. Math., (60):197–376, 1984. 413
[Bum97] Daniel Bump. Automorphic forms and representations, vol-
ume 55 of Cambridge Studies in Advanced Mathematics. Cam-
bridge University Press, Cambridge, 1997. 74, 84, 89, 123, 421,
422
[Bum05] Daniel Bump. The Rankin-Selberg method: an introduction and
survey. In Automorphic representations, L-functions and appli-
cations: progress and prospects, volume 11 of Ohio State Univ.
Math. Res. Inst. Publ., pages 41–73. de Gruyter, Berlin, 2005.
268
[Bum13] Daniel Bump. Lie groups, volume 225 of Graduate Texts in Math-
ematics. Springer, New York, second edition, 2013. 222
[BW00] A. Borel and N. Wallach. Continuous cohomology, discrete sub-
groups, and representations of reductive groups, volume 67 of
Mathematical Surveys and Monographs. American Mathemati-
cal Society, Providence, RI, second edition, 2000. 157, 302, 303,
305, 307, 308, 309
References 437
[GW14] Jayce R. Getz and Eric Wambach. Twisted relative trace for-
mulae with a view towards unitary groups. Amer. J. Math.,
136(1):1–58, 2014. 287, 314, 404
[GZ86] Benedict H. Gross and Don B. Zagier. Heegner points and deriva-
tives of L-series. Invent. Math., 84(2):225–320, 1986. 292
[Hab78] W. J. Haboush. Homogeneous vector bundles and reductive
subgroups of reductive algebraic groups. Amer. J. Math.,
100(6):1123–1137, 1978. 330
[Hah09] Heekyoung Hahn. A simple twisted relative trace formula. Int.
Math. Res. Not. IMRN, (21):3957–3978, 2009. x, 374
[Har74] G. Harder. Chevalley groups over function fields and automor-
phic forms. Ann. of Math. (2), 100:249–306, 1974. 183
[Har77] Robin Hartshorne. Algebraic geometry. Springer-Verlag, New
York-Heidelberg, 1977. Graduate Texts in Mathematics, No. 52.
3
[Har04] David Harari. Weak approximation on algebraic varieties. In
Arithmetic of higher-dimensional algebraic varieties (PaloAlto,
CA, 2002), volume 226 of Progr. Math., pages 43–60. Birkhäuser
Boston, Boston, MA, 2004. 42
[Har11] Günter Harder. Lectures on algebraic geometry I. Aspects of
Mathematics, E35. Vieweg + Teubner Verlag, Wiesbaden, re-
vised edition, 2011. Sheaves, cohomology of sheaves, and appli-
cations to Riemann surfaces. 301, 306
[HC53] Harish-Chandra. Representations of a semisimple Lie group on
a Banach space. I. Trans. Amer. Math. Soc., 75:185–243, 1953.
84, 102
[HC68] Harish-Chandra. Automorphic forms on semisimple Lie groups.
Notes by J. G. M. Mars. Lecture Notes in Mathematics, No. 62.
Springer-Verlag, Berlin, 1968. 116
[HC70a] Harish-Chandra. Eisenstein series over finite fields. In Functional
analysis and related fields (Proc. Conf. M. Stone, Univ. Chicago,
Chicago, Ill., 1968), pages 76–88. Springer, New York, 1970. 150,
195
[HC70b] Harish-Chandra. Harmonic analysis on reductive p-adic groups.
Lecture Notes in Mathematics, Vol. 162. Springer-Verlag, Berlin-
New York, 1970. Notes by G. van Dijk. 103
[HC75] Harish-Chandra. Harmonic analysis on real reductive groups. I.
The theory of the constant term. J. Functional Analysis, 19:104–
204, 1975. 413
[HC99] Harish-Chandra. Admissible invariant distributions on reductive
p-adic groups, volume 16 of University Lecture Series. American
Mathematical Society, Providence, RI, 1999. Preface and notes
by Stephen DeBacker and Paul J. Sally, Jr. 160
444 References