BumpD-Lie Groups
BumpD-Lie Groups
Daniel Bump
Lie Groups
Second Edition
Graduate Texts in Mathematics 225
Graduate Texts in Mathematics
Series Editors:
Sheldon Axler
San Francisco State University, San Francisco, CA, USA
Kenneth Ribet
University of California, Berkeley, CA, USA
Advisory Board:
Graduate Texts in Mathematics bridge the gap between passive study and creative
understanding, offering graduate-level introductions to advanced topics in mathemat-
ics. The volumes are carefully written as teaching aids and highlight characteristic
features of the theory. Although these books are frequently used as textbooks in grad-
uate courses, they are also suitable for individual study.
Lie Groups
Second Edition
123
Daniel Bump
Department of Mathematics
Stanford University
Stanford, CA, USA
ISSN 0072-5285
ISBN 978-1-4614-8023-5 ISBN 978-1-4614-8024-2 (eBook)
DOI 10.1007/978-1-4614-8024-2
Springer New York Heidelberg Dordrecht London
Library of Congress Control Number: 2013944369
This book aims to be both a graduate text and a study resource for Lie groups.
It tries to strike a compromise between accessibility and getting enough depth
to communicate important insights. In discussing the literature, often sec-
ondary sources are preferred: cited works are usually recommended ones.
There are four parts. Parts I, II or IV are all “starting points” where one
could begin reading or lecturing. On the other hand, Part III assumes familiar-
ity with Part II. The following chart indicates the approximate dependencies
of the chapters. There are other dependencies, where a result is used from a
chapter that is not a prerequisite according to this chart: but these are rela-
tively minor. The dashed lines from Chaps. 1 and 2 to the opening chapters of
Parts II and IV indicate that the reader will benefit from knowledge of Schur
orthogonality but may skip or postpone Chaps. 1 and 2 before starting Part II
or Part IV. The other dashed line indicates that the Bruhat decomposition
(Chap. 27) is assumed in the last few chapters of Part IV.
Chapters 1-2
Key
Part II
24-27 23 39 40-41
Part III
The two lines of development in Parts II–IV were kept independent because
it was possible to do so. This has the obvious advantage that one may start
reading with Part IV for an alternative course. This should not obscure the
fact that these two lines are complementary, and shed light on each other. We
hope the reader will study the whole book.
v
vi Preface
Part I treats two basic topics in the analysis of compact Lie groups: Schur
orthogonality and the Peter–Weyl theorem, which says that the irreducible
unitary representations of a compact group are all finite-dimensional.
Usually the study of Lie groups begins with compact Lie groups. It is
attractive to make this the complete content of a short course because it can
be treated as a self-contained subject with well-defined goals, about the right
size for a 10-week class. Indeed, Part II, which covers this theory, could be used
as a traditional course culminating in the Weyl character formula. It covers
the basic facts about compact Lie groups: the fundamental group, conjugacy
of maximal tori, roots and weights, the Weyl group, the Weyl integration
formula, and the Weyl character formula. These are basic tools, and a short
course in Lie theory might end up with the Weyl character formula, though
usually I try to do a bit more in a 10-week course, even at the expense of
skipping a few proofs in the lectures. The last chapter in Part II introduces
the affine Weyl group and computes the fundamental group. It can be skipped
since Part III does not depend on it.
Sage, the free mathematical software system, is capable of doing typical
Lie theory calculations. The student of Part II may want to learn to use it.
An appendix illustrates its use.
The goal of Part I is the Peter–Weyl theorem, but Part II does not depend
on this. Therefore one could skip Part I and start with Part II. Usually when
I teach this material, I do spend one or two lectures on Part I, proving Schur
orthogonality but not the Peter–Weyl formula. In the interests of speed I tend
to skip a few proofs in the lectures. For example, the conjugacy of maximal
tori needs to be proved, and this depends in turn on the surjectivity of the
exponential map for compact groups, that is, Theorem 16.3. This is proved
completely in the text, and I think it should be proved in class but some
of the differential geometry details behind it can be replaced by intuitive
explanations. So in lecturing, I try to explain the intuitive content of the
proof without going back and proving Proposition 16.1 in class. Beginning
with Theorems 16.2–16.4, the results to the end of the chapter, culminating
in various important facts such as the conjugacy of maximal tori and the
connectedness of centralizers can all be done in class. In the lectures I prove the
Weyl integration formula and (if there is time) the local Frobenius theorem.
But I skip a few things like Theorem 13.3. Then it is possible to get to the
Weyl Character formula in under 10 weeks.
Although compact Lie groups are an essential topic that can be treated in
one quarter, noncompact Lie groups are equally important. A key role in much
of mathematics is played by the Borel subgroup of a Lie group. For example,
if G = GL(n, R) or GL(n, C), the Borel subgroup is the subgroup of upper
triangular matrices, or any conjugate of this subgroup. It is involved in two
important results, the Bruhat and Iwasawa decompositions. A noncompact Lie
group has two important classes of homogeneous spaces, namely symmetric
spaces and flag varieties, which are at the heart of a great deal of important
Preface vii
field or an adele ring, and ultimately this “philosophy of cusp forms” leads to
the theory of automorphic forms. Thirdly, the ring R has as a homomorphic
image the cohomology rings of flag varieties, leading to the Schubert calculus.
These topics are surveyed in the final chapters.
What’s New? I felt that the plan of the first edition was a good one, but that
substantial improvements were needed. Some material has been removed, and
a fair amount of new material has been added. Some old material has been
streamlined or rewritten, sometimes extensively. In places what was implicit
in the first edition but not explained well is now carefully explained with at-
tention to the underlying principles. There are more exercises. A few chapters
are little changed, but the majority have some revisions, so the changes are
too numerous to list completely. Highlights in the newly added material in-
clude the affine Weyl group, new material about Coxeter groups, Demazure
characters, Bruhat order, Schubert and Bott–Samelson varieties, the Borel-
Weil theorem the appendix on Sage, Clifford algebras, the Keating–Snaith
theorem, and more.
Notation. The notations GL(n, F ) and GLn (F ) are interchangeable for the
group of n × n matrices with coefficients in F . By Matn (F ) we denote the
ring of n × n matrices, and Matn×m (F ) denotes the vector space of n × m
matrices. In GL(n), I or In denotes the n × n identity matrix and if g is any
matrix, t g denotes its transpose. Omitted entries in a matrix are zero. Thus,
for example,
1 0 1
= .
−1 −1 0
The identity element of a group is usually denoted 1 but also as I, if the
group is GL(n) (or a subgroup), and occasionally as e when it seemed the
other notations could be confusing. The notations ⊂ and ⊆ are synonymous,
but we mostly use X ⊂ Y if X and Y are known to be unequal, although we
make no guarantee that we are completely consistent in this. If X is a finite
set, |X| denotes its cardinality.
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
1 Haar Measure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 Schur Orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3 Compact Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6 Vector Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
11 Extension of Scalars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
12 Representations of sl(2, C) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
xi
xii Contents
15 Tori . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
24 Complexification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
31 Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 535
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 545
Part I
Compact Groups
1
Haar Measure
Such a measure is called a left Haar measure. It has the properties that any
compact set has finite measure and any nonempty open set has measure > 0.
We will not prove the existence and uniqueness of the Haar measure. See
for example Halmos [61], Hewitt and Ross [69], Chap. IV, or Loomis [121] for
a proof of this. Left-invariance of the measure amounts to left-invariance of
the corresponding integral,
f (γg) dμL (g) = f (g) dμL (g), (1.1)
G G
then it is easy to see that the left- and right-invariant measures are, respec-
tively,
dμL = y −2 dx dy, dμR = y −1 dx dy.
They are not the same. However, there are many cases where they do coincide,
and if the left Haar measure is also right-invariant, we call G unimodular .
The measure δ(h)μL (h) is a right Haar measure, and we may write μR (h) =
δ(h)μL (h). The quasicharacter δ is called the modular quasicharacter.
−1
δ(g) f (h) dμL (h) = f (g · g hg) dμL (h) = f (hg) dμL (h).
G G G
Replace f by f δ in this identity and then divide both sides by δ(g) to find
that
f (h) δ(h) dμL (h) = f (hg) δ(h) dμL (h).
G G
Proof. It is easy to see that g −→ g −1 turns a left Haar measure into a right
Haar measure. If left and right Haar measures agree, then g −→ g −1 multiplies
the left Haar measure by a positive constant, which must be 1 since the map
has order 2.
1 Haar Measure 5
Exercises
Exercise 1.1. Let da X denote the Lebesgue measure on Matn (R). It is of course a
Haar measure for the additive group Matn (R). Show that | det(X)|−n da X is both a
left and a right Haar measure on GL(n, R).
Let dg1 and dg2 denote Haar measures on GL(r, R) and GL(s, R), and let da X
denote an additive Haar measure on Matr×s (R). Show that
are (respectively) left and right Haar measures on P , and conclude that the modular
quasicharacter of P is
δ(p) = | det(g1 )|s | det(g2 )|−r .
2
Schur Orthogonality
In this chapter and the next two, we will consider the representation theory
of compact groups. Let us begin with a few observations about this theory
and its relationship to some related theories.
If V is a finite-dimensional complex vector space, or more generally a
Banach space, and π : G −→ GL(V ) a continuous homomorphism, then
(π, V ) is called a representation. Assuming dim(V ) < ∞, the function
χπ (g) = tr π(g) is called the character of π. Also assuming dim(V ) < ∞,
the representation (π, V ) is called irreducible if V has no proper nonzero
invariant subspaces, and a character is called irreducible if it is a character of
an irreducible representation.
[If V is an infinite-dimensional topological vector space, then (π, V ) is
called irreducible if it has no proper nonzero invariant closed subspaces.]
A quasicharacter χ is a character in this sense since we can take V = C
and π(g)v = χ(g)v to obtain a representation whose character is χ.
These facts can be directly generalized in two ways. First, Fourier analy-
sis on locally compact Abelian groups, including Pontriagin duality, Fourier
π(g)v, π(g)w = v, w .
for it is easy to see that this inner product is Hermitian and positive definite.
It is G-invariant by construction.
and
(L1 ⊗ L2 ) (π1 ⊗ π2 )(g)(v1 ⊗ v2 ) = L1 π1 (g)v1 L2 π2 (g)v2 ,
Note that the inverse is needed here so that π̂(g1 g2 ) = π̂(g1 )π̂(g2 ).
If (π, V ) is a representation,
then
by Proposition 2.3 any linear combination
of functions of the form L π(g) v with v
∈ V , L ∈ V ∗ is a matrix coefficient,
though it may be a function L π (g) v where (π , V ) is not (π, V ), but a
larger representation.
Nevertheless, we call any linear combination of functions
of the form L π(g) v a matrix coefficient of the representation (π, V ). Thus,
the matrix coefficients of π form a vector space, which we will denote by Mπ .
Clearly, dim(Mπ ) dim(V )2 .
T ◦ π1 (g) = π2 (g) ◦ T
for g ∈ G. We will denote by HomC (V1 , V2 ) the space of all linear trans-
formations V1 −→ V2 and by HomG (V1 , V2 ) the subspace of those that are
intertwining maps.
For the remainder of this chapter, unless otherwise stated, G will denote
a compact group.
Lemma 2.1. Suppose that (π1 , V1 ) and (π2 , V2 ) are complex representations
of the compact group G. Let , be any inner product on V1 . If vi , wi ∈ Vi ,
then the map T : V1 −→ V2 given by
T (w) = π1 (g)w, v1 π2 (g −1 )v2 dg (2.4)
G
is G-equivariant.
Proof. We have
T π1 (h)w = π1 (gh)w, v1 π2 (g −1 )v2 dg.
G
The variable change g −→ gh−1 shows that this equals π2 (h)T (w), as required.
Proof. We will show that if v1 and v2 are fixed, there exists a constant c(v1 , v2 )
such that
π(g)w1 , v1 π(g)w2 , v2 dg = c(v1 , v2 ) w1, w2 . (2.6)
G
Now the variable change g −→ g −1 and the properties of the inner product
show that this equals the left-hand side of (2.6), proving the identity. The
same argument shows that there exists another constant c (w1 , w2 ) such that
for all v1 and v2 we have
π(g)w1 , v1 π(g)w2 , v2 dg = c (w1 , w2 ) v2, v1 .
G
Combining this with (2.6), we get (2.5). We will compute d later in Proposi-
tion 2.9, but for now we simply note that it is positive since, taking w1 = w2
and v1 = v2 , both the left-hand side of (2.5) and the two inner products on
the right-hand side are positive.
This proves (i). As for (ii), referring to (2.3), π̂(g) is the adjoint of π(g)−1 with
respect to the dual pairing , , so its trace equals the trace of π(g)−1 .
If (π1 , V1 ) and (π2 , V2 ) are irreducible representations, and χ1 and χ2 are their
characters, we have already noted in proving Proposition 2.3 that we may form
representations π1 ⊕ π2 and π1 ⊗ π2 on V1 ⊕ V2 and V1 ⊗ V2 . It is easy to see
that χπ1 ⊕π2 = χπ1 + χπ2 and χπ1 ⊗π2 = χπ1 χπ2 . It is not quite true that
the characters form a ring. Certainly the negative of a matrix coefficient is a
2 Schur Orthogonality 15
matrix coefficient, yet the negative of a character is not a character. The set
of characters is closed under addition and multiplication but not subtraction.
We define a generalized (or virtual ) character to be a function of the form
χ1 − χ2 , where χ1 and χ2 are characters. It is now clear that the generalized
characters form a ring.
Proof. Both tr Ψ (g1 , g2 ) and tr(g1−1 ) tr(g2 ) are continuous, and since diag-
onalizable matrices are dense in GL(n, C) we may assume that both g1
and g2 are diagonalizable. Also if γ is invertible we have Ψ (γg1 γ −1 , g2 ) =
Ψ (γ, 1)Ψ (g1 , g2 )Ψ (γ, 1)−1 so the trace of both tr Ψ (g1 , g2 ) and tr(g1−1 )tr(g2 )
are unchanged if g1 is replaced by γg1 γ −1 . So we may assume that g1 is di-
agonal, and similarly g2 . Now if α1 , . . . , αn and β1 , . . . , βm are the diagonal
entries of g1 and g2−1 , the effect of Ψ (g1 , g2 ) on X ∈ Ω is to multiply the
columns by the α−1 i and the rows by the βj . So the trace is tr(g1−1 )tr(g2 ).
By lemma 2.2 and Proposition 2.6, the character of Π(g) is χ2 (g)χ1 (g). The
space of invariants Ω G exactly of the T which are G-module homomorphisms,
so by Proposition 2.8 we get
χ1 (g) χ2 (g) dg = dim HomG (V1 , V2 ).
G
There are n2 terms on the right, but by (2.5) only the terms with i = j are
nonzero, and those equal d−1 . Thus, d = n.
Proof. Let L ∈ V ∗ and v ∈ V . Define fL,v (g) = L(π(g)v). The map L, v −→
fL,v is bilinear, hence induces a linear map σ : V ∗ ⊗V −→ Mπ . It is surjective
by the definition of Mπ , and it follows from Proposition 2.4 that if Li and vj
run through orthonormal bases, then fLi ,vj are orthonormal, hence linearly
independent. Therefore, σ is a vector space isomorphism. We have
Exercises
(i) Use the variable change h −→ h−1 g to prove the identity of the last two terms.
Prove that this operation is associative, and so C(G) is a ring (without unit)
with respect to covolution.
(ii) Let π be an irreducible representation. Show that the space Mπ of matrix
coefficients of π is a 2-sided ideal in C(G), and explain how this fact implies
Theorem 2.3.
Exercise 2.4. Let G be a compact group, and let G × G act on the space Mπ
by left and right translation: (g, h)f (x) = f (g −1 xh). Show that Mπ ∼
= π̂ ⊗ π as
(G × G)-modules.
Exercise 2.5. Let G be a compact group and let g, h ∈ G. Show that g and h are
conjugate if and only if χ(g) = χ(h) for every irreducible character χ. Show also
that every character is real-valued if and only if every element is conjugate to its
inverse.
Exercise 2.6. Let G be a compact group, and let V, W be irreducible
G-modules.
An invariant bilinear form B : V ×W → C is one that satisfies B g·v, g·w = B(v, w)
for g ∈ G, v ∈ V , w ∈ W . Show that the space of invariant bilinear forms is at most
one-dimensional, and is one-dimensional if and only if V and W are contragredient.
3
Compact Operators
T f, g = f, T g
| T x, x |
|T | = sup . (3.1)
0=x∈H x, x
| T x, x | |T x| · |x| |T | · |x|2 = |T | · x, x ,
so B |T |. We must prove
the
converse. Let λ > 0 be a constant, to be
determined later. Using T 2 x, x = T x, T x, we have
T x, T x
= 14 T (λx + λ−1 T x), λx + λ−1 T x − T (λx − λ−1 T x), λx − λ−1 T x
14 T (λx + λ−1 T x), λx + λ−1 T x + T (λx − λ−1 T x), λx − λ−1 T x
14 B λx + λ−1 T x, λx + λ−1 T x + B λx − λ−1 T x, λx − λ−1 T x
= B2 λ2 x, x + λ−2 T x, T x .
Now taking λ = |T x|/|x|, we obtain
T x, x = x, T x = T x, x
|λ xi − T xi |2 = λ xi − T xi , λ xi − T xi = λ2 |xi |2 + |T xi |2 − 2λ T xi , xi ,
and since |xi | = 1, |T xi | → |λ|, and T xi , xi → λ, this converges to 0. Since
T xi → v, the sequence λxi therefore also converges to v, and xi → λ−1 v.
Now, by continuity, T xi → λ−1 T v, so v = λ−1 T v. This proves that v is
an eigenvector with eigenvalue λ. This completes the proof that N⊥ has an
orthonormal basis consisting of eigenvectors.
Now let {φi } be this orthonormal basis and let λi be the corresponding
eigenvalues. If > 0 is given, only finitely many |λi | > since otherwise we
can find an infinite sequence of φi with |T φi | > . Such a sequence will have
no convergent subsequence, contradicting the compactness of T . Thus, N⊥ is
countable-dimensional, and we may arrange the {φi } in a sequence. If it is
infinite, we see the λi −→ 0.
Proposition 3.1. Let X and Y be compact topological spaces with Y a metric
space with distance function d. Let U be a set of continuous maps X −→ Y
such that for every x ∈ X
and every > 0 there exists a neighborhood N of
x such that d f (x), f (x ) < for all x ∈ N and for all f ∈ U . Then every
sequence in U has a uniformly convergent subsequence.
We refer to the hypothesis on U as equicontinuity.
Proof. Let S0 = {f1 , f2 , f3 , . . .} be a sequence in U . We will show that it has
a convergent subsequence. We will construct a subsequence that is uniformly
Cauchy and hence has a limit. For every n > 1, we will construct a subsequence
Sn = {fn1 , fn2 , fn3 , . . .} of Sn−1 such that supx∈X d fni (x), fnj (x) 1/n.
Assume that Sn−1 is constructed. For each x ∈ X, equicontinuity guaran-
tees the existence of an open neighborhood Nx of x such that d f (y), f (x)
3n for all y ∈ Nx and all f ∈ X. Since X is compact, we can cover X by
1
a finite number of these sets, say Nx1 , . .. , Nxm . Since the fn−1,i take
values
in the compact space Y , the m-tuples fn−1,i (x1 ), . . . , fn−1,i (xm ) have an
accumulation
point, and
we may therefore select the subsequence {fni } such
that d fni (xk ), fnj (xk ) 3n 1
for all i, j and 1 k m. Then for any y,
there exists xk such that y ∈ Nxk and
d fni (y), fnj (y) d fni (y), fni (xk ) + d fni (xk ), fnj (xk )
+d fnj (y), fnj (xk ) 3n
1 1
+ 3n + 3n1
= n1 .
This completes the construction of the sequences {fni }.
The diagonal sequence {f11 , f22 , f33 , . . .} is uniformly Cauchy. Since Y is
a compact metric space, it is complete, and so this sequence is uniformly
convergent.
We topologize C(X) by giving it the L∞ norm | |∞ (sup norm).
Proposition 3.2 (Ascoli and Arzela). Suppose that X is a compact space
and that U ⊂ C(X) is a bounded subset such that for each x ∈ X and > 0
there is a neighborhood N of x such that |f (x) − f (y)| for all y ∈ N and
all f ∈ U . Then every sequence in U has a uniformly convergent subsequence.
22 3 Compact Operators
Exercises
Exercise 3.1. Suppose that T is a bounded operator on the Hilbert space H, and
suppose that for each > 0 there exists a compact operator T such that |T −T | < .
Show that T is compact. (Use a diagonal argument like the proof of Proposition 3.1.)
2
Let φi be an orthonormal basis of L (X). Expand K in a Fourier expansion:
∞
K(x, y) = ψi (x) φi (y), ψi = T φi .
i=1
Show that |ψi |2 = |K(x, y)|2 dμ(x) dμ(y) < ∞. Consider the operator TN with
kernel
N
KN (x, y) = ψi (x) φi (y).
i=1
−1
(Use the variable change h −→ h g to prove the identity of the last two
terms. See Exercise 2.3.) We will sometimes define f1 ∗ f2 by this formula
even if f1 and f2 are not assumed continuous. For example, we will make use
of the convolution defined this way if f1 ∈ L∞ (G) and f2 ∈ L1 (G), or vice
versa.
Since G has total volume 1, we have inequalities (where | |p denotes the
Lp norm, 1 p ∞)
|f |1 |f |2 |f |∞ . (4.1)
|f |1 = |f |, 1 |f |2 · |1|2 = |f |2 .
while
f1 , Tφ f2 = φ(hg −1 ) f1 (h) f2 (g) dg dh.
G G
Theorem 4.1 (Peter and Weyl). The matrix coefficients of G are dense
in C(G).
Proof. Let f ∈ C(G). We will prove that there exists a matrix coefficient f
such that |f − f |∞ < for any given > 0.
Since G is compact, f is uniformly continuous. This means that there exists
an open neighborhood U of the identity
such that if g ∈ U , then |λ(g)f −
f |∞ < /2, where λ : G → End C(G) is the action by left translation:
−1
(λ(g)f )(h) = f (g h). Let φ be a nonnegative function supported in U such
that G φ(g) dg = 1. We may arrange that φ(g) = φ(g −1 ) so that the operator
Tφ is self-adjoint as well as compact. We claim that |Tφ f − f |∞ < /2. Indeed,
if h ∈ G,
(φ ∗ f )(h) − f (h) = φ(g) f (g −1 h) − φ(g)f (h) dg
G
φ(g) f (g −1 h) − f (h) dg
U
φ(g) |λ(g)f − f |∞ dg
U
φ(g) (/2) dg = .
U 2
Let
f = Tφ (f ), f = fλ ,
|λ|>q
26 4 The Peter–Weyl Theorem
where
q > 0 remains to be chosen. We note that f and f are both contained
in |λ|>q V (λ), which is a finite-dimensional vector space, and closed under
right translation by Proposition 4.3, and by Theorem 2.1, it follows that they
are matrix coefficients.
By (4.3), we may choose q so that 0<q<|λ| |fλ |22 is as small as we like.
Using (4.1) may thus arrange that
f f = |fλ |22 < . (4.4)
λ λ 2|φ|∞
0<|λ|<q 0<|λ|<q 0<|λ|<q
1 2
We have
⎛ ⎞ ⎛ ⎞
Tφ (f − f ) = Tφ ⎝f0 + fλ ⎠ = Tφ ⎝ fλ ⎠ .
0<|λ|<q 0<|λ|<q
|f − f |∞ = |f − Tφ f + Tφ (f − f )| |f − Tφ f | + |Tφ f − Tφ f |
2 + 2 = .
Proof. Since C(G) is dense in L2 (G), this follows from the Peter–Weyl
theorem and (4.1).
Theorem 4.2. Let G be a compact group that has no small subgroups. Then
G has a faithful finite-dimensional representation.
and
φ(g) v − π(g)v dg, v |v − π(g)v|dg · |v| |v|2 /2.
N N
Thus, the two terms in (4.5) differ in absolute value and cannot cancel.
Next, using the Peter–Weyl theorem, we may find a matrix coefficient f
such that |f − φ|∞ < , where can be chosen arbitrarily. We have
(f − φ)(g) π(g)v dg |v| ,
G
so if is sufficiently small we have G f (g) π(g)v dg
= 0.
Since f is a matrix coefficient, so is the function g −→ f (g −1 ) by Proposi-
tion 2.4. Thus, let (ρ, W ) be a finite-dimensional representation
w∈W
with
and L : W −→ C a linear functional such that f (g −1 ) = L ρ(g)w . Define a
map T : W −→ H by
T (x) = L ρ(g −1 )x π(g)v dg.
G
28 4 The Peter–Weyl Theorem
Exercises
Exercise 4.1. Let G be totally disconnected, and let π : G −→ GL(n, C) be a
finite-dimensional representation. Show that the kernel of π is open. (Hint: Use the
fact that GL(n, C) has no small subgroups.) Conclude (in contrast with Theorem 4.2)
that the compact group GL(n, Zp ) has no faithful finite-dimensional representation.
Example 5.1. If F is a field, then the general linear group GL(n, F ) is the
group of invertible n × n matrices with coefficients in F . It is a Lie group.
Assuming that F = R or C, the group GL(n, F ) is an open set in Matn (F )
and hence a manifold of dimension n2 if F = R or 2n2 if F = C. The special
linear group is the subgroup SL(n, F ) of matrices with determinant 1. It is a
closed Lie subgroup of GL(n, F ) of dimension n2 − 1 or 2(n2 − 1).
Example 5.2. If F = R or C, let O(n, F ) = {g ∈ GL(n, F ) | g · t g = I}. This
is the n × n orthogonal group. More geometrically, O(n, F ) is the group of
linear transformations preserving the quadratic form Q(x1 , . . . , xn ) = x21 +
x22 + · · · + x2n . To see this, if (x) = t (x1 , . . . , xn ) is represented as a column
vector, we have Q(x) = Q(x1 , . . . , xn ) = t x · x, and it is clear that Q(gx) =
Q(x) if g · t g = I. The group O(n, R) is compact and is usually denoted
simply O(n). The group O(n) contains elements of determinants ±1. The
subgroup of elements of determinant 1 is the special orthogonal group SO(n).
The dimension of O(n) and its subgroup SO(n) of index 2 is 12 (n2 − n). This
will be seen in Proposition 5.6 when we compute their Lie algebra (which is
the same for both groups).
Example 5.3. More generally, over any field, a vector space V on which there
is given a quadratic form q is called a quadratic space, and the set O(V, q)
of linear transformations of V preserving q is an orthogonal group. Over the
complex numbers, it is not hard to prove that all orthogonal groups are iso-
morphic (Exercise 5.4), but over the real numbers, some orthogonal groups are
not isomorphic to O(n). If k + r = n, let O(k, r) be the subgroup of GL(n, R)
preserving the indefinite quadratic form x21 + · · · + x2k − x2k+1 − · · · − x2n . If
r = 0, this is O(n), but otherwise this group is noncompact. The dimensions
of these Lie groups are, like SO(n), equal to 12 (n2 − n).
Example 5.4. The unitary group U(n) = {g ∈ GL(n, C) | g·t g = I}. If g ∈ U(n)
then | det(g)| = 1, and every complex number of absolute value 1 is a possible
determinant of g ∈ U(n). The special unitary group SU(n) = U(n) ∩ SL(n, C).
The dimensions of U(n) and SU(n) are n2 and n2 − 1, just like GL(n, R) and
SL(n, R).
Example 5.5. If F = R or C, let Sp(2n, F ) = {g ∈ GL(2n, F ) | g · J · t g = J},
where
0 −In
J= .
In 0
This is the symplectic group. The compact group Sp(2n, C) ∩ U(2n) will be
denoted as simply Sp(2n).
A Lie algebra over a field F is a vector space g over F endowed with a bilinear
map, the Lie bracket , denoted (X, Y ) −→ [X, Y ] for X, Y ∈ g, that satisfies
[X, Y ] = −[Y, X] and the Jacobi identity
[X, [Y, Z]] + [Y, [Z, X]] + [Z, [X, Y ]] = 0. (5.1)
5 Lie Subgroups of GL(n, C) 33
exp(X) = I + X + 12 X 2 + 16 X 3 + · · · . (5.2)
This function is smooth and has support in the ball {|x| r}. Taking r
sufficiently small, we can make this vanish outside U .
34 5 Lie Subgroups of GL(n, C)
Proposition 5.2. Let G be a closed Lie subgroup of GL(n, C), and let X ∈
Matn (C). Then the path t −→ exp(tX) is tangent to the submanifold G of
GL(n, C) at t = 0 if and only if it is contained in G for all t.
Proof. If exp(tX) is contained in G for all t, then clearly it is tangent to G at
t = 0. We must prove the converse. Suppose that exp(t0 X) ∈ / G for some t0 >
0. Using Proposition 5.1, Let φ0 be a smooth compactly supported function
on GL(n, C) such that φ0 (g) = 0 for all g ∈ G, φ0 0, and φ0 exp(t0 X)
= 0.
Let
f (t) = φ exp(tX) , φ(h) = φ0 (hg) dg, t∈R,
G
in terms of a left Haar measure on G. Clearly, φ is constant on the cosets hG
of G, vanishes on G, but is nonzero at exp(t0 X). For any t,
d
f (t) = φ exp(tX) exp(uX) u=0 = 0
du
since the path u −→ exp(tX) exp(uX) is tangent to the coset exp(tX)G and
φ is constant on such cosets. Moreover, f (0) = 0. Therefore, f (t) = 0 for all
t, which is a contradiction since f (t0 )
= 0.
Proposition 5.3. Let G be a closed Lie subgroup of GL(n, C). The set Lie(G)
of all X ∈ Matn (C) such that exp(tX) ⊂ G is a vector space whose dimension
is equal to the dimension of G as a manifold.
Proof. This is clear from the characterization of Proposition 5.2.
Proposition 5.4. Let G be a closed Lie subgroup of GL(n, C). The map
X −→ exp(X)
We see that Lie(G) is a Lie subalgebra of gl(n, C). Thus, we are able to
associate a Lie algebra with a Lie group.
Example 5.8. The Lie algebra of GL(n, F ) with F = R or C is gl(n, F ).
Example 5.9. Let sl(n, F ) be the subspace of X ∈ gl(n, F ) such that tr(X) =
0. This is a Lie subalgebra, and it is the Lie algebra of SL(n, F ) when F = R
or C. This follows immediately from the fact that det(eX ) = etr(X) for any
matrix X because if x1 , . . . , xn are the eigenvalues of X, then ex1 , . . . , exn are
the eigenvalues of eX .
Example 5.10. Let o(n, F ) be the set of X ∈ gl(n, F ) that are skew-symmetric,
in other words, that satisfy X + t X = 0. It is easy to check that o(n, F ) is
closed under the Lie bracket and hence is a Lie subalgebra.
Proposition 5.6. If F = R or C, the Lie algebra of O(n, F ) is o(n, F ). The
dimension of O(n) is 12 (n2 − n), and the dimension of O(n, C) is n2 − n.
Proof. Let G = O(n, F ), g = Lie(G). Suppose X ∈ o(n, F ). Exponentiate the
identity −tX = tt X to get
exp(tX)−1 = t exp(tX),
whence exp(tX) ∈ O(n, F ) for all t ∈ R. Thus, o(n, F ) ⊆ g. To prove the
converse, suppose that X ∈ g. Then, for all t,
I = exp(tX) · t exp(tX)
= (I + tX + 12 t2 X 2 + · · · )(I + t t X + 1 2
2 t · tX 2 + · · · )
= I + t(X + t X) + 1 2
2 t (X
2
+ 2X · t X + X 2 ) + · · · .
t
Since this is true for all t, each coefficient in this Taylor series must vanish
(except of course the constant one). In particular, X + t X = 0. This proves
that g = o(n, F ).
The dimensions of O(n) and O(n, C) are most easily calculated by comput-
ing the dimension of the Lie algebras. A skew-symmetric matrix is determined
by its upper triangular entries, and there are 12 (n2 − n) of these.
Example 5.11. Let u(n) be the set of X ∈ GL(n, C) such that X + t X = 0.
One checks easily that this is closed under the gl(n, C) Lie bracket [X, Y ] =
XY − Y X. Despite the fact that these matrices have complex entries, this is
a real Lie algebra, for it is only a real vector space, not a complex one. (It is
not closed under multiplication by complex scalars.) It may be checked along
the lines of Proposition 5.6 that u(n) is the Lie algebra of U(n), and similarly
su(n) = {X ∈ u(n) | tr(X) = 0} is the Lie algebra of SU(n).
Example 5.12. Let sp(2n, F ) be the set of matrices X ∈ Mat2n (F ) that satisfy
XJ + J t X = 0, where
0 −In
J= .
In 0
This is the Lie algebra of Sp(2n, F ).
36 5 Lie Subgroups of GL(n, C)
Exercises
Exercise 5.1. Show that O(n, m) is the group of g ∈ GL(n + m, R) such that
g J1 t g = J1 , where
In
J1 = .
−Im
with all entries not on one of the two diagonals equal to zero. If N is odd, the middle
element of this matrix is 1.
Exercise 5.4. Let V1 and V2 be vector spaces over a field F , and let qi be a quadratic
form on Vi for i = 1, 2. The quadratic spaces are called equivalent if there exists an
isomorphism l : V1 −→ V2 such that q1 = q2 ◦ l.
(i) Show that over a field of characteristic not equal to 2, any quadratic form is
equivalent to ai x2i for some constants ai .
(ii) Show that, if F = C, then any quadratic space of dimension n is equivalent to
Cn with the quadratic form x21 + · · · + x2n .
(iii) Show that, if F = R, then any quadratic space of dimension n is equivalent to
Rn with the quadratic form x21 + · · · + x2r − x2r+1 − · · · − x2n for some r.
Exercise 5.5. Compute the Lie algebra of Sp(2n, R) and the dimension of the
group.
5 Lie Subgroups of GL(n, C) 37
Exercise 5.6. Show that there is a ring isomorphism Matn (H) −→ Mat2n (C) with
the following description. Any A ∈ Matn (H) may be written uniquely as A1 + A2 j
with A1 , A2 ∈ Matn (C). The isomorphism in question maps
A1 A2
A1 + A2 j −→ .
−Ā2 Ā1
Exercise 5.7. Show that if A ∈ Matn (H), then A · t Ā = I if and only if the complex
2n × 2n matrix
A1 A2
−Ā2 Ā1
is in both Sp(2n, C) and U(2n). Recall that the intersection of these two groups was
the group denoted Sp(2n).
Exercise 5.8. Show that the groups SO(2) and SU(2) may be identified with the
groups of matrices
a b
a, b ∈ F, |a| + |b| = 1 ,
2 2
−b̄ ā
where F = R or C, respectively.
Exercise 5.9. The group SU(1, 1) is by definition the group of g ∈ SL(2, C) such
that
1
g · J · t g = J, J= .
−1
(i) Show that SU(1, 1) consists of all elements of SL(2, C) of the form
ab
, |a|2 − |b|2 = 1.
b̄ ā
(ii) Show that the Lie algebra su(1, 1) of SU(1, 1) consists of all matrices of the form
ai b
b̄ −ai
with a real.
1 −i
(iii) Let C = √1 ∈ SL(2, C). This element is sometimes called the Cayley
1 i
2i
transform. Show that C·SL(2, R)·C −1 = SU(1, 1) and C·sl(2, R)·C −1 = su(1, 1).
6
Vector Fields
lim x−1
1 f (x1, . . . , xn ) = (∂f /∂x1 )(0, x2 , . . . , xn )
x1 →0
∂ k2 +···+kn f ∂ k2 +···+kn g
and
∂xk22 · · · ∂xknn ∂xk22 · · · ∂xknn
is a local derivation.
Proof. If f and g both vanish at m, then (6.1) implies that a local derivation
X vanishes on M2 , and by Proposition 6.1 it is therefore determined by its
values on x1 , . . . , xn . If these are a1 , . . . , an , then X agrees with the right-hand
side of (6.2).
It follows from the chain rule that this definition is independent of the choice
of local coordinates xi .
Now let A = C ∞ (M, R) be the ring of smooth real-valued functions on M .
Given a vector field X on M , we may obtain a derivation of A as follows.
If f ∈ A, let X(f ) be the smooth function that assigns to m ∈ M the value
Xm (f ), where we are of course applying Xm to the germ of f at m. For
example, if M = U is an open set on Rn with coordinate functions x1 , . . . , xn
on U , given smooth functions ai : U −→ R, we may associate a derivation of
A with the vector field (6.4) by
n
∂f
(Xf )(m) = ai (m) (m). (6.5)
i=1
∂xi
Exercises
The following exercise requires some knowledge of topology.
Exercise 6.1. Let X be a vector field on the sphere S k . Assume that X is nowhere
zero, i.e., Xm
= 0 for all m ∈ S k . Show that the antipodal map a : S k −→ S k and
the identity map S k −→ S k are homotopic. Deduce that k is odd.
Hint: Normalize the vector field so that Xm is a unit tangent vector for all m.
If m ∈ S k consider the great circle θm : [0, 2π] −→ S k tangent to Xm . Then
θm (0) = θm (2π) = m, but m −→ θm (π) is the antipodal map. Also, think about the
effect of the antipodal map on H k (S k ).
7
Left-Invariant Vector Fields
Both f (x) and c0 + c1 (x) + B(x, x) have the same partial derivatives of order
2, so the difference r(x) vanishes to order 3. The fact that B is symmetric
follows from the equality of mixed partials:
∂2f ∂2f
(0) = (0).
∂xi ∂xj ∂xj ∂xi
Here [X, Y ] means XY − Y X computed using matrix operations; that is, the
bracket computed as in Chap. 5. This proposition shows that if X ∈ Matn (C),
and if we associate with X a derivation of C ∞ (G, R), where G = GL(n, C),
using the formula (7.1), then this bracket operation gives the same result as
the bracket operation (6.6) for left-invariant vector fields.
and
(dX ◦ dY f )(g) = c1 (XY ) + 2B(X, Y ). (7.4)
48 7 Left-Invariant Vector Fields
Indeed,
d
(dX f )(g) = f g(I + tX) |t=0
dt
d
= c0 + c1 (tX) + B(tX, tX) + r(tX) .
dt t=0
We may ignore the B and r terms because they vanish to order 2, and since
c1 is linear, this is just c1 (X) proving (7.3). Also
d
(dX ◦ dY f )(g) = (dY f ) g(I + tX)
dt u=0
∂ ∂
= f g(I + tX)(I + uY )
∂t ∂u t=u=0
∂ ∂
= [c0 + c1 (tX + uY + tuXY )
∂t ∂u
+B(tX + uY + tuXY, tX + uY + tuXY )
+r(tX + uY + tuXY )] |t=u=0 .
by (7.3).
We may ask to what extent the Lie algebra homomorphism Lie(φ) contains
complete information about φ. For example, given Lie groups G and H with
Lie algebras g and h, and a homomorphism f : g −→ h, is there a homomor-
phism G −→ H with Lie(φ) = f ?
In general, the answer is no, as the following example will show.
ξ → − det(ξ) = x2 + y 2 + z 2
Exercises
Exercise 7.1. Compute the Lie algebra homomorphism Lie(ψ) : su(2) −→ so(3) of
Example 7.1 explicitly.
Exercise 7.2. Show that no Lie group can be homeomorphic to the sphere S k if k
is even. On the other hand, show that SU(2) ∼
= S 3 . (Hint: Use Exercise 6.1.)
Exercise 7.3. Let J be the matrix (5.3). Let o(N, C) and oJ (C) be the complexified
Lie algebras of the groups O(N ) and OJ (C) in Exercise 5.9. Show that these complex
Lie algebras are isomorphic. Describe o(N, C) explicitly, i.e., write down a typical
matrix.
8
The Exponential Map
The existence of such a solution for sufficiently small |t|, and its uniqueness on
whatever interval it does exist, is guaranteed by a standard result in the theory
of ordinary differential equations, which may be found in most texts. See, for
example, Ince [81], Chap. 3, particularly Sect. 3.3, for a rigorous treatment.
The required Lipschitz condition follows from smoothness of the ai. For the
statement about continuously varying vector fields, one needs to know the
corresponding fact about first-order systems, which is discussed in Sect. 3.31
of [81]. Here Ince imposes an assumption of analyticity on the dependence of
the differential equation on λ, which he allows to be a complex parameter,
because he wants to conclude analyticity of the solutions; if one weakens
this assumption of analyticity to smoothness, one still gets smoothness of the
solution.
In general, the existence of the integral curve of a vector field is only guaran-
teed in a small segment (−, ), as in Proposition 8.1. However, we will now see
that, for left-invariant vector fields on a Lie group, the integral curve extends
to all R. This fact underlies the construction of the exponential map.
Theorem 8.1. Let G be a Lie group and g its Lie algebra. There exists a map
exp : g −→ G that is a local homeomorphism in a neighborhood of the origin
in g such that, for any X ∈ g, t −→ exp(tX) is an
integral curve for the
left-invariant vector field X. Moreover, exp (t + u)X = exp(tX) exp(uX).
Proof. Let X ∈ g. We know that for sufficiently small > 0 there exists
an integral curve p : (−, ) −→ G for the left-invariant vector field X with
p(0) = 1. We show first that if p : (a, b) −→ G is any integral curve for an
open interval (a, b) containing 0, then
p(t) = p(−a/2) p(t + a/2) when −3a/2 t a/2, and it follows from (8.2)
that this definition is consistent on regions of overlap.
Now define exp : g −→ G as follows. Let X ∈ g, and let p : R −→ G be an
integral curve for the left-invariant vector field X with p(0) = 0. We define
exp(X) = p(1). We note that if u ∈ R, then t → p(tu) is an integral curve for
uX, so exp(uX) = p(u).
The exponential map is a smooth map, at least for X near the origin in g,
by the last statement in Proposition 8.1. Identifying the tangent space at the
origin in the vector space g with g itself, exp induces a map T0 (g) −→ Te (G)
(that is g −→ g), and this map is the identity map by construction. Thus, the
Jacobian of exp is nonzero and, by the Inverse Function Theorem, exp is a
local homeomorphism near 0.
Proposition 8.2. Let G, H be Lie groups and let g, h be their respective Lie
algebras. Let f : G → H be a homomorphism. Then the following diagram is
commutative:
df
g −−−−→ h
⏐ ⏐
⏐exp ⏐exp
# #
f
G −−−−→ H
Proof. It is clear from the definitions that f takes an integral curve for a
left-invariant vector field X on G to an integral curve for df (X), and the
statement follows.
since, by the Jacobi identity, both sides equal [x, [y, z]] = [[x, y], z] + [y, [x, z]].
This means that ad(x) is a derivation of g.
Also
ad(x) ad(y) − ad(y) ad(x) = ad [x, y] (8.4)
since applying either side to z ∈ g gives [x, [y, z]] − [y, [x, z]] = [[x, y], z] by the
Jacobi identity. So ad : g −→ End(g) is a Lie algebra representation.
We next explain the geometric origin of ad. To begin with, representations
of Lie algebras arise naturally from representations of Lie groups. Suppose
that G is a Lie group and g is its Lie algebra. If V is a vector space over R
or C, any Lie group homomorphism π : G −→ GL(V ) induces a Lie algebra
homomorphism g −→ End(V ) by Proposition 7.3; that is, a real or complex
representation.
In particular, G acts on itself by conjugation, and so it acts on g = Te (G).
This representation is called the adjoint representation and is denoted Ad :
G −→ GL(g). We show next that the differential of Ad is ad. That is:
Theorem 8.2. Let G be a Lie group, g its Lie algebra, and Ad : G −→ GL(g)
the adjoint representation. Then the Lie group representation g −→ End(g)
corresponding to Ad by Proposition 7.3 is ad.
Proof. It will be most convenient for us to think of elements of the Lie algebra
as tangent vectors at the identity or as local derivations of the local ring there.
Let X, Y ∈ g. If f ∈ C ∞ (G), define c(g)f (h) = f (g −1 hg). Then our definitions
of the adjoint representation amount to
Ad(g)Y f = Y c(g −1 )f .
Applying this, with u fixed to F (t1 , t2 ) = f (et1 X euY e−t2 X ), our last expression
equals
d d d d
f (etX euY ) − f (euY etX ) = XY f (1) − Y Xf (1).
du dt t=u=0 du dt t=u=0
Exercises
Exercise 8.1. Show that the exponential map su(2) → SU(2) is surjective, but the
exponential map sl(2, R) → SL(2, R) is not.
9
Tensors and Universal Properties
We will review the basic properties of the tensor product and use them to
illustrate the basic notion of a universal property, which we will see repeatedly.
If R is a commutative ring and M , N , and P are R-modules, then a bilinear
map f : M × N −→ P is a map satisfying
Lemma 9.1. In any category, any two initial objects are isomorphic. Any two
terminal objects are isomorphic.
Proof. If X0 and X1 are initial objects, there exist unique morphisms f :
X0 −→ X1 (since X0 is initial) and g : X1 −→ X0 (since X1 is initial). Then
g ◦ f : X0 −→ X0 and 1X0 : X0 −→ X0 must coincide since X0 is initial,
and similarly f ◦ g = 1X1 . Thus f and g are inverse isomorphisms. Similarly,
terminal objects are isomorphic.
Theorem 9.1. The tensor product M ⊗R N , if it exists, is determined up to
isomorphism by the universal property.
Proof. Let C be the following category. An object in C is an ordered pair
(P, p), where P is an R-module and p : M × N −→ P is a bilinear map.
If X = (P, p) and Y = (Q, q) are objects, then a morphism X −→ Y consists
of an R-module homomorphism f : P −→ Q such that q = f ◦p. The universal
property of the tensor product means that ⊗ : M × N −→ M ⊗ N is an initial
object in this category and therefore determined up to isomorphism.
Of course, we usually denote ⊗(m, n) as m ⊗ n in M ⊗R N . We have not
proved that M ⊗R N exists. We refer to any text on algebra for this fact, such
as Lang [116], Chap. XVI.
In general, by a universal property we mean any characterization of a
mathematical object that can be expressed by saying that some associated object
is an initial or terminal object in some category. The basic paradigm is that
a universal property characterizes an object up to isomorphism.
A typical application of the universal property of the tensor product is
to make M ⊗R N into a functor. Specifically, if μ : M −→ M and ν :
N −→ N are R-module homomorphisms, then there is a unique R-module
homomorphism μ ⊗ ν : M ⊗R N −→ M ⊗R N such that (μ ⊗ ν)(m ⊗ n) =
μ(m) ⊗ ν(n). We get this by applying the universal property to the R-bilinear
map M × N −→ M ⊗ N defined by (m, n) −→ μ(m) ⊗ ν(n).
As another example of an object that can be defined by a universal prop-
$ let V be a vector space over a field F . Let
erty, $ us ask for an F -algebra
V together with an F -linear map i : V −→ V satisfying the following
condition.
It should be clear from the previous discussion that this universal property
characterizes the tensor algebra up to isomorphism. To prove existence, we can
construct a ring with this exact property as follows. Let unadorned ⊗ mean
⊗F in what follows. By ⊗k V we mean the k-fold tensor product V ⊗ · · · ⊗ V
(k times); if k = 0, then it is natural to take ⊗0 V = F while ⊗1 V = V . If V
has finite dimension d, then ⊗k V has dimension dk . Let
9 Tensors and Universal Properties 59
% ∞
& k
V = ⊗ V .
k=0
$
Then V has the natural structure of a graded F -algebra in which the
multiplication ⊗k V × ⊗l V −→ ⊗k+l V sends
(v1 ⊗ · · · ⊗ vk , u1 ⊗ · · · ⊗ ul ) −→ v1 ⊗ · · · ⊗ vk ⊗ u1 ⊗ · · · ⊗ ul .
$
We regard V as a subset of V embedded onto ⊗1 V = V .
A graded algebra over the field F is an F -algebra A with a direct sum decom-
position
∞
&
A= Ak
k=0
Exercises
Exercise 9.1. Let V be a finite-dimensional vector space over a field F that may
be assumed to be infinite. Let P(V ) be the ring of polynomial functions on V . Note
that an element of the dual space V ∗ is a function on V , so regarding this function
as a polynomial gives an
injection V ∗ −→ P(V ). Show that this linear map extends
∗
to a ring isomorphism V −→ P(V ).
Exercise 9.2. Prove that if V is a vector space, then V ⊗ V ∼ = (V ∧ V ) ⊕ (V ∨ V ).
Exercise 9.3. Use the universal properties of the symmetric and exterior power to
show that if V and W are vector spaces, then there are maps ∨k f : ∨k V −→ ∨k W
and ∧k f : ∧k V −→ ∧k W such that
∨k f (v1 ∨ · · · ∨ vk ) = f (v1 ) ∨ · · · ∨ f (vk ), ∧k f (v1 ∧ · · · ∧ vk ) = f (v1 ) ∧ · · · ∧ f (vk ).
Exercise 9.4. Suppose that V = F 4 . Let f : V −→ V be the linear transformation
with eigenvalues a, b, c, d. Compute the traces of the linear transformations ∨2 f and
∧2 f on ∨2 V and ∧2 V as polynomials in a, b, c, d.
Exercise 9.5. Let A and B be algebras over the field F . Then A ⊗ B is also an
algebra, with multiplication (a ⊗ b)(a ⊗ b ) = aa ⊗ bb . Show that there are ring
homomorphisms i : A → A ⊗ B and j : B → A ⊗ B such that if f : A → C and
g : B → C are ring homomorphisms into a ring C satisfying f (a) g(b) = g(b) f (a)
for a ∈ A and b ∈ B, then there exists a unique ring homomorphism φ : A ⊗ B → C
such that φ ◦ i = f and φ ◦ j = g.
Exercise 9.6. Show that if U and V are finite-dimensional vector spaces over F
then show that
(U ⊕ V ) ∼
= U ⊗ U
and
(U ⊕ V ) ∼
= U ⊗ U .
10
The Universal Enveloping Algebra
We have seen that elements of the Lie algebra of a Lie group G are derivations
of C ∞ (G). They are thus first-order differential operators that are left-
invariant. The universal enveloping algebra is a purely algebraically defined
ring that may be identified with the ring of all left-invariant differential
operators, including higher-order ones.
We recall from Example 5.6 that if A is an associative algebra, then A may
be regarded as a Lie algebra by the rule [a, b] = ab − ba for a, b ∈ A. We will
denote this Lie algebra by Lie(A).
Suppose that g is the Lie algebra of a Lie group G. Consider the ring A of
vector space endomorphisms of C ∞ (G) that commute with left translation
by elements of G. As we have already seen, elements of g are left-invariant
differential operators, by means of the action
d tX
Xf (g) = f ge |t=0 . (10.1)
dt
By the universal property of the universal enveloping algebra, this action
extends to a ring homomorphism U (g) −→ A, the image of which consists
of left-invariant differential operators [Exercise 10.2 (i)]. Let us apply this
observation to give a quick analytic proof of a fact that has a longer purely
algebraic proof.
Proposition 10.1. If g is the Lie algebra of a Lie group G, then the natural
map i : g −→ U (g) is injective.
The center of U (g) is very important. One reason for this is that while
elements of U (g) are realized as differential operators that are invariant under
left-translation, elements of the center are invariant under both left and right
translation. [Exercise 10.2 (ii)]. Moreover, the center acts by scalars on any
irreducible subspace, as we see in the following version of Schur’s lemma.
A representation (π, V ) of a Lie algebra g is irreducible if there is no proper
nonzero subspace U ⊂ V such that π(x)U ⊆ U for all x ∈ g.
the key features of this theory is how the Casimir element becomes the key
ingredient in many proofs (such as that of the Weyl character formula) where
other tools are no longer available. See [92].
Our next task will be to construct the Casimir elements. This requires
a discussion of invariant bilinear forms. If V is a vector space over F and
π : g −→ End(V ) is a representation, then we call a bilinear form B on V
invariant if
B π(X)v, w + B v, π(X)w = 0 (10.2)
for X ∈ g, v, w ∈ V . The following proposition shows that this notion of
invariance is the Lie algebra analog of the more intuitive corresponding notion
for Lie groups.
Proposition 10.3. Suppose that G is a Lie group, g its Lie algebra, and
π : G −→ GL(V ) a representation admitting an invariant bilinear form B.
Then B is invariant for the differential of π.
Proof. Invariance under π means that
B π(etX )v, π(etX )w = B(v, w).
π(y)π(x)π(z) − π(y)π(z)π(x).
while
⎛ ⎞
xi yi z = (−xi [z, yi ] + xi zyi ) = − ⎝ βij xi yj ⎠ + xi zyi ,
i i i,j i
and since βij = −αji , these are equal. Thus Δ commutes with g, and since g
generates U (g) as a ring, it is in the center.
It remains to be shown that Δ is independent of the choice of basis
x1 , . . . , xd . Suppose that x1 , . . . , xd is another basis. Write
x
i = j cij xj , and
if y1 , . . . , yd is the corresponding dual basis, let y
i = j d ij y j . The condition
that B(xi , yj ) = δij (Kronecker δ) implies that k cik djk = δij . Therefore,
the matrices
(cij ) and (dij )
are transposeinverses of each other
and so we have
also k cki dkj = δij . Now k xk yk = i,j,k cki dkj xi yj = k xk yk = Δ.
Although Proposition 10.4 provides us with a supply of invariant bilinear
forms, there is no guarantee that they are nonzero, which is required by
Theorem 10.2. We will not address this point now.
One might wonder, since there may be many irreducible representations,
whether the invariant bilinear forms produced by Proposition 10.4 are all
distinct. Also, since these invariant bilinear forms are all symmetric, one
might wonder if we are missing some invariant bilinear forms that are not
symmetric. The following proposition shows that for simple Lie algebras, there
is essentially a unique invariant bilinear form, and that it is symmetric.
A Lie algebra g is called simple if it has no proper nonzero ideals. An ideal
is just an invariant subspace of g for the adjoint representation, so another way
of saying the same thing is that ad : g −→ End(g) is irreducible. For example,
it is not hard to see that for any field F , the Lie algebra sl(n, F ) is simple.
10 The Universal Enveloping Algebra 65
B([x, y], z) = −B(z, [x, y]) = B([x, z], y) = −B([z, x], y).
Applying this identity three times, B([x, y], z) = −B([x, y], z) and because
the characteristic is not two, we have B([x, y], z) = 0 for all x, y, z. Now we
may assume that g is non-Abelian since otherwise it is one-dimensional and
any bilinear form is symmetric. Then [g, g] is an ideal of g (as follows from
the Jacobi identity) and is nonzero since g is non-Abelian. Since g is simple
[g, g] = g and we have proved that B = 0, a contradiction.
Exercises
Exercise 10.1. Let Xij ∈ gl(n, R) (1 i, j n) be the n × n matrix with a 1 in
the i, j position and 0’s elsewhere. Show that [Xij , Xkl ] = δjk Xil − δil Xkj , where
δjk is the Kronecker δ. From this, show for any positive integer d that
66 10 The Universal Enveloping Algebra
n
n
··· Xi1 i2 Xi2 i3 · · · Xid i1
i1 =1 ir =1
Exercise 10.2. Let G be a connected Lie group and g its Lie algebra. Define an
action of g on the space C ∞ (G) of smooth functions on G by (10.1).
(i) Show that this is a representation of G. Explain why Theorem 10.1 implies that
this action of g on C ∞ (G) can be extended to a representation of the associative
algebra U (g) on C ∞ (G).
(ii) If h ∈ G, let ρ(h) and λ(h) be the endomorphisms of G given by left and right
translation. Thus
Exercise 10.3. Let G = GL(n, R). Let B be the “Borel subgroup” of upper
triangular matrices with positive diagonal entries, and let B0 be the connected
component of the identity, whose matrices have positive diagonal entries. Let
K = O(n).
(i) Show that every element of g ∈ G has a unique decomposition as g = bk with
b ∈ B0 and k ∈ K.
(ii) Let s = (s1 , . . . , sn ) ∈ Cn . By (i), we may define an element φ = φs of C ∞ (G) by
⎛⎛ ⎞ ⎞
y1 ∗ · · · ∗
⎜⎜ 0 y2 · · · ∗ ⎟ ⎟ n
⎜⎜ ⎟ ⎟
φs ⎜⎜ . . . . ⎟ k⎟ = y si , yi > 0, k ∈ K.
⎝⎝ .. .. . . .. ⎠ ⎠ i=1 i
0 0 · · · yn
Exercise 10.4. Give a construction similar to that in Exercise 10.3 for eigenfunc-
tions of the center of U (g) when g = gln (C).
M1 ⊕ M2 −→ M, (m1 , m2 ) −→ i1 m1 + i2 m2 .
These are easily checked to be inverses of each other, and so M = ∼ M1 ⊕M2 . For
example, if M = M1 ⊕M2 , such maps exist—take the inclusion and projection
maps in and out of the direct sum. Now applying the functor M → S ⊗R M to
the maps j1 , j2 , p1 , p2 gives corresponding maps for S ⊗R (M1 ⊗R M2 ) showing
that it is isomorphic to the left-hand side of (11.2).
To prove (11.3), one has an S-bilinear map
(S ⊗R M1 ) × (S ⊗R M2 ) −→ S ⊗R (M1 ⊗R M2 ) (11.5)
such that (s1 ⊗ m1 ), (s2 ⊗ m2 ) → s1 s2 ⊗ (m1 ⊗ m2 ). This map is S-bilinear,
so it induces a homomorphism
(S ⊗R M1 ) ⊗S (S ⊗R M2 ) −→ S ⊗R (M1 ⊗R M2 ). (11.6)
S × (M1 ⊗R M2 ) −→ (S ⊗R M1 ) ⊗S (S ⊗R M2 )
Proposition 11.2.
(i) If V is a real vector space and W is a complex vector space, any R-linear
transformation V −→ W extends uniquely to a C-linear transformation
VC −→ W .
(ii) If V and U are real vector spaces, any R-linear transformation V −→ U
extends uniquely to a C-linear map VC −→ UC .
(iii) If V and U are real vector spaces, any R-bilinear map V × V −→ U
extends uniquely to a C-bilinear map VC × VC −→ UC .
Proof. Part (i) is a special case of (ii) of Proposition 11.1. Part (ii) follows
by taking W = UC in part (i) after composing the given linear map V −→ U
with the inclusion U −→ W . As for (iii), an R-bilinear map V × V −→ U
induces an R-linear map V ⊗R V −→ U and hence by (ii) a C-linear map
(V ⊗R V )C −→ UC . But by (11.3), (V ⊗R V )C is VC ⊗C VC , and a C-linear map
VC ⊗C VC −→ UC is the same thing as a C-bilinear map VC × VC −→ UC .
Proposition 11.3.
(i) The complexification gC of a real Lie algebra g with the bracket extended
as in Proposition 11.2 (iii) is a Lie algebra.
(ii) If g is a real Lie algebra, h is a complex Lie algebra, and ρ : g −→ h is a
real Lie algebra homomorphism, then ρ extends uniquely to a homomor-
phism ρC : gC −→ h of complex Lie algebras. In particular, any complex
representation of g extends uniquely to a complex representation of gC .
(iii) If g is a real Lie subalgebra of the complex Lie algebra h, and if h = g ⊕ ig
(i.e., if g and ig span h but g ∩ ig = {0}), then h ∼ = gC as complex Lie
algebras.
Proof. For (i), the extended bracket satisfies the Jacobi identity since both
sides of (5.1) are trilinear maps on gC × gC × gC −→ gC , which by assumption
vanish on g × g × g. Since g generates gC over the complex numbers, (5.1) is
therefore true on gC .
For (ii), the extension is given by Proposition 11.2 (i), taking W = h.
To see that the extension is a Lie algebra homomorphism, note that both
ρ([x, y]) and ρ(x)ρ(y) − ρ(y)ρ(x) are bilinear maps gC × gC −→ h that agree
on g × g. Since g generates gC over C, they are equal for all x, y ∈ gC .
For (iii), by Proposition 11.2 (i), it will be least confusing to distinguish
between g and its image in h, so we prove instead the following equivalent
statement: if g is a real Lie algebra, h is a complex Lie algebra, f : g −→ h
is an injective homomorphism, and if h = f (g) ⊕ i f (g), then f extends to an
isomorphism gC −→ h of complex Lie algebras. Now f extends to a Lie algebra
homomorphism fC : gC −→ h by part (ii). To see that this is an isomorphism,
note that it is surjective since f (g) spans h. To prove that it is injective, if
fC (X +iY ) = 0 with X, Y ∈ g, then f (X)+if (Y ) = 0. Now f (X) = f (Y ) = 0
because f (g) ∩ if (g) = 0. Since f is injective, X = Y = 0.
70 11 Extension of Scalars
Let
1 0
x= , y= ,
0 1
be the standard basis of C2 . We have a corresponding basis of k + 1 elements
in ∨k C2 , which we will label by integers k, k − 2, k − 4, . . . , −k for reasons that
will become clear presently. Thus, we let
R R R R
···
v−k v−k+2 v−k+4 vk
L L L L
For example, if k = 3, then with respect to the basis v3 , v1 , v−1 , v−3 , we find
that
⎛ ⎞
0 1 0 0
⎜0 0 2 0⎟
∨3 R = ⎜
⎝ 0 0 0 3 ⎠,
⎟
0 0 0 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 3 0 0 0
⎜3 0 0 0⎟ ⎜0 1 0 0 ⎟
∨3 L = ⎜⎝ 0 2 0 0 ⎠,
⎟ ∨3 H = ⎜⎝ 0 0 −1 0 ⎠ .
⎟
0 0 1 0 0 0 0 −3
It may be checked directly that these matrices satisfy the commutation rela-
tions (12.1).
proving (10.3). Dual to the basis H, R, L of sl(2, R) is the basis H, 2L, 2R,
and it follows from Theorem 10.2 that the Casimir element
Δ = H 2 + 2RL + 2LR
is an element of the center of U sl(2, R) .
Δ = H 2 + 2H + 4LR. (12.5)
Proposition 12.8. Let g = sl(2, R), su(2), or sl(2, C). Let (π, V ) be a finite-
dimensional complex representation of g.
(i) If v ∈ V and Δ2 v = 0, then Δv = 0.
(ii) We have V = V0 ⊕ V1 , where V0 is the kernel of Δ and V1 is the image of
Δ. Both are invariant subspaces. If X ∈ g and v ∈ V0 , then π(X)v = 0.
(iii) The subspace V0 = {v ∈ V | π(X) = 0 for all X ∈ g}.
(iv) If 0 −→ V −→ W −→ Q −→ 0 is an exact sequence of g-modules, then
there is an exact sequence 0 −→ V0 −→ W0 −→ Q0 −→ 0.
0 V0 W0 Q0
0 V W Q 0
Δ Δ Δ
0 V W Q 0
V/V1 W/W1
Exactness of the two middle rows implies exactness of the top row. We must
show that W0 −→ Q0 is surjective. We will deduce this from the Snake Lemma.
The cokernel of Δ : V −→ V is V /V1 ∼ = V0 , and similarly the cokernel of
Δ : W −→ W is W/W1 ∼ = W0 , so the Snake lemma gives us a long exact
sequence:
0 −→ V0 −→ W0 −→ Q0 −→ V0 −→ W0 .
Since the last map is injective, the map Q0 −→ V0 is zero, and hence W0 −→
Q0 is surjective.
Exercises
Exercise 12.1. If (π, V ) is a representation of SL(2, R), SU(2) or SL(2, C), then we
may restrict the character of π to the diagonal subgroup. This gives
t
ξπ (t) = tr π −1 ,
t
(i) Compute ξπ (t) for the symmetric power representations. Show that the polyno-
mials ξπ (t) are linearly independent and determine the representation π.
(ii) Show that if Π = π ⊗ π , then ξΠ = ξπ ξπ . Use this observation to compute the
decomposition of π ⊗ π into irreducibles when π = ∨n C2 and π = ∨m C2 .
Exercise 12.2. Show that each representation of sl(2, R) comes from a representa-
tion of SL(2, R).
Exercise 12.3. Let g = sl(3, R). Let Δ be the Casimir element with respect to the
invariant bilinear form 12 tr(xy) on g. Show that if (π, V ) is an irreducible represen-
tation with Δ · V = 0, then V is trivial.
[Hint: Here are some suggestions for a direct approach. Let
⎛ ⎞ ⎛ ⎞
1 0
H1 = ⎝ −1 ⎠ , H2 = ⎝ 1 ⎠,
0 −1
and (denoting by Eij the matrix with 1 in the i, j position, 0 elsewhere) let R1 = E12 ,
R2 = E23 , R3 = E13 , L1 = E21 , L2 = E32 , L3 = E31 . These eight elements are a
basis. Since [H1 , H2 ] = 0 there exists a vector vλ that is a simultaneous eigenvector,
so that H1 vλ = (λ1 − λ2 )vλ and H2 vλ = (λ2 − λ3 )vλ for some triple (λ1 , λ2 , λ3 ) of
real numbers. (We may normalize them so λ1 + λ2 + λ3 = 0, and it may then be
shown that λi ∈ 13 Z, though you may not need that fact.) Let Vλ be the space of
such vectors. Show that R1 maps Vλ into Vλ+α1 and R2 maps Vλ into Vλ+α2 where
α1 = (1, −1, 0) and α2 = (0, 1, −1). (What does R3 do to Vλ , and what do the Li
do?) Conclude that there is a nonzero vector vλ in some Vλ that is annihilated by
R1 , R2 and R3 . Show that λ1 λ2 λ3 . For this, it may be useful to observe that
there are two copies of sl2 (R) in sl3 (R) spanned by Hi , Ri , Li with i = 1, 2, so you
may restrict the representation to these and make use of the theory in the text.
Compute the eigenvector of Δ and show that Δvλ = 0 implies λ = (0, 0, 0).]
Exercise 12.4. Show that complex representations of su(3), sl(3, R), and sl(3, C)
are completely reducible.
of sl(2, R),
Exercise 12.5. Show that if (π, V ) is a faithful representation then the
trace bilinear form BV : g×g → C defined by BV (X, Y ) = tr π(X) π(Y ) is nonzero.
Exercise 12.6. Let g be a simple Lie algebra. Assume that g contains a subalgebra
isomorphic to sl(2, R). Let π : g → End(V ) be an irreducible representation. Assume
that π is not the trivial representation.
(ii) Show that the trace bilinear form BV on g defined by B(X, Y ) = tr(π(X)π(Y ))
is nondegenerate. (Hint: First show that it is nonzero.)
(iii) By (ii) there exists a Casimir element Δ in the center of U (g) as in Theorem 10.2.
Show that the eigenvalue of Δ on V is 1/dim(V ). (Hint: Take traces.)
(iv) Show that representations of g are completely reducible. (Hint: Use Proposi-
tion 10.5.)
Exercise 12.7. Show that complex representations of su(n), sl(n, R), and sl(n, C)
are completely reducible.
13
The Universal Cover
the endpoints. This means that ht (0) = p(0) = p (0) and ht (1) = p(1) = p (1)
for all t. We call h a path-homotopy, and we write p ≈ p if a path-homotopy
exists.
Suppose there exists a continuous function f : [0, 1] −→ [0, 1] such
that f (0) = 0 and f (1) = 1 and that p = p ◦ f . Then we say that p
is a reparametrization of p. The paths
are path-homotopic since we can
consider pt (u) = p (1 − t)u + tf (u) . Because the interval [0, 1] is convex,
(1 − t)u + tf (u) ∈ [0, 1] and pt : p p .
Let us say that a map of topological spaces is trivial if it is constant,
mapping the entire domain to a single point. A topological space U is
contractible if the identity map U −→ U is homotopic to a trivial map. A space
U is path-connected if for all x, y ∈ U there exists a path p : [0, 1] −→ U such
that p(0) = x and p(1) = y.
Suppose that p : [0, 1] −→ U and q : [0, 1] −→ U are two paths in the
space U such that the right endpoint of p coincides with the left endpoint of
q; that is, p(1) = q(0). Then we can concatenate the paths to form the path
p q:
We may also reverse a path: −p is the path (−p)(t) = p(1 − t). These
operations are compatible with path-homotopy, and the path p (−p) is
homotopic to the trivial path p0 (t) = p(0). To see this, define
pt (u) = p(2tu)
if 0 u 1/2,
p 2t(1 − u) if 1/2 u 1.
π −1 (U ) ∼
= U × π −1 (x) −→ U,
where the second map is the projection, coincides with the given map π.
We say that the cover is trivial if N is homeomorphic to M × F , where
the space F is discrete, in such a way that π is the composition N ∼
=M×
F −→ M (where the second map is the projection). Thus, each m ∈ M has
a neighborhood U such that the restricted covering map π −1 (U ) −→ U is
trivial, a property we will cite as local triviality of the cover.
Proposition 13.2. Let π : N −→ M be a covering map.
(i) If p : [0, 1] −→ M is a path, and if y ∈ π −1 p(0) , then there exists a
unique path p̃ : [0, 1] −→ N such that π ◦ p̃ = p and p̃(0) = y.
(ii) If p̃, p̃ : [0, 1] −→ N are paths with p̃(0) = p̃ (0), and if the paths π ◦ p̃ and
π ◦ p̃ are path-homotopic, then the paths p̃ and p̃ are path-homotopic.
We refer to (i) as the path lifting property of the covering space. We refer to
(ii) as the homotopy lifting property.
Proof. If the cover is trivial, then we may assume that N = M × F where
F is discrete, and if y = (x, f ), where
x
= p(0) and f ∈ F , then the unique
solution to this problem is p̃(t) = p(t), f .
Since p([0, 1]) is compact, and since the cover is locally trivial, there are a
finite number of open sets U1 , U2, . . . , Un and points x0 = 0 < x1 < · · · < xn =
1 such that p([xi−1 , xi ]) ⊂ Ui and such that the restriction of the cover to Ui
is trivial. On each interval [xi−1 , x], there is a unique solution, and patching
these together gives the unique general solution. This proves (i).
For (ii), since p = π ◦ p̃ and p = π ◦ p̃ are path-homotopic, there exists a
continuous map (u, t) → pt (u) from [0, 1]×[0, 1] −→ M such that p0 (u) = p(u)
and p1 (u) = p (u). For each t, using (i) there is a unique path p,t : [0, 1] −→ M̃
such that pt = π ◦ p̃t and p̃t (0) = p(0). One may check that (u, t) → p̃t (u) is
continuous, and p,0 = p̃ and p,1 = p̃ , so p̃ and p̃ are path-homotopic.
Covering spaces of a fixed space M form a category: if π : N −→ M and
π : N −→ M are covering maps, a morphism is a covering map f : N −→ N
such that π = π ◦ f . If M is a pointed space, we are actually interested in the
subcategory of pointed covering maps: if x0 is the base point of M , the base
point of N must lie in the fiber π −1 (x0 ), and in this category the morphism
f must preserve base points. We call this category the category of pointed
covering maps or pointed covers of M .
84 13 The Universal Cover
q(0) = y0 , and Proposition 13.2 (ii) shows that the path-homotopy class of q
depends only on the path-homotopy class of p. Then mapping p → q(1) is the
unique morphism M̃ −→ N of pointed covers of M .
If N is simply connected, any covering map M −→ N is an isomorphism
by Proposition 13.3.
This result shows that the universal cover is a functor: if M̃ and Ñ are the
universal covers of M and N , then this proposition implies that a continuous
map φ : M → N induces a map φ̃ : M̃ → Ñ .
Proof. Let x0 be a base point for M , and let y0 be an element of N such that
π(y0 ) = y0 where y0 = f (x0 ). If x ∈ M , we may find a path p : [0, 1] −→ M
such that p(0) = x0 and p(1) = x. By Proposition 13.2 (i) we may then find
a path p̃ : [0, 1] −→ N such that π ◦ p̃ = f ◦ p and p̃(0) = y0 . We will define
f (x) = p̃(1), but first we must check that this is well-defined. If q is another
path with q(0) = x0 and q(1) = x, and if q̃ : [0, 1] −→ N is the corresponding
lift of f ◦p with q̃(0) = y0 , then we must show q̃(1) = p̃(1). The paths p and p
are homotopic because M is simply connected. That is, the concatenation of p
with the inverse path to p is a loop, hence contractible, and this implies that
p and p are homotopic. It follows that the paths q̃ and p̃ are path-homotopic,
and in particular they have the same right endpoint q̃(1) = p̃(1). Hence we
may define f (x) = p̃(1) and this is the required map.
If M is a pointed space and x0 is its base point, then the fiber π̃ −1 (x0 )
coincides with its fundamental group π1 (M ). We are interested in the case
where M = G is a Lie group. We take the base point to be the origin.
by (t, u) −→ p(t)q(u). Taking different routes from (0, 0) to (1, 1) will give
path-homotopic paths. Going directly via t → f (t, t) = p(t)q(t) gives p · q,
while going indirectly via
f (2t, 0) = p(2t) if 0 t 12 ,
t →
f (1, 2t − 1) = q(2t − 1) if 12 t 1,
Proof. We may identify the circle S 1 with the unit circle in C. Then x →
e2πix is a covering map R −→ S 1 . The space R is contractible and hence
simply-connected, so it is the universal covering space. If we give S 1 ⊂ C×
the group structure it inherits from C× , then this map R −→ S 1 is a group
homomorphism, so by Theorem 13.2 we may identify the kernel Z with π1 (S 1 ).
To see that S r is simply connected for r 2, let p : [0, 1] −→ S r be a path.
Since it is a mapping from a lower-dimensional manifold, perturbing the path
slightly if necessary, we may assume that p is not surjective. If it omits one
point P ∈ S r , its image is contained in S r − {P }, which is homeomorphic to
Rr and hence contractible. Therefore p, is path-homotopic to a trivial path.
It follows from the fact that φ is a local homomorphism that the right-hand
side is
φ p(xi ) p(xi−1 )−1 .
Replacing i by i − 1 in (13.2) and taking v = xi , this equals q(xi )q(xi−1 )−1 .
Now (13.1) follows for this path by noting that if = 12 min |xi+1 − xi |, then
when |u − v| < , u, v ∈ [0, 1], there exists an i such that u, v ∈ [xi−1 , xi+1 ],
and (13.1) follows from (13.2) and the fact that φ is a local homomorphism.
This proves that the path q exists. To show that it is unique, assume
that (13.1) is valid for |u−v| < , and choose the xi so that |xi −xi+1 | < ; then
for v ∈ [xi , xi+1 ], (13.2) is true, and the values of q are determined by this
property.
Next we indicate how one can show that if p and p are path-homotopic,
and if q and q are the corresponding paths in H, then q(1) = q (1). It is
sufficient to prove this in the special case of a path-homotopy t → pt , where
p0 = p and p1 = p , such that there exists a sequence 0 = x1 · · · xn = 1
with pt (u)pt (v)−1 ∈ U when u, v ∈ [xi−1 , xi+1 ] and t and t ∈ [0, 1]. For
although a general path-homotopy may not satisfy this assumption, it can be
broken into steps, each of which does. In this case, we define
88 13 The Universal Cover
qt (v) = φ pt (v) p(xi )−1 q(xi )
Let
q (2t) if 0 t 1/2 ,
q (t) =
q(2t − 1)q (1) if 1/2 t 1.
Proposition 13.7.
(i) If g1 and g2 are positive definite Hermitian matrices, and if g12 = g22 , then
g1 = g2 .
(ii) If g1 and g2 are Hermitian matrices and eg1 = eg2 , then g1 = g2 .
Proof. To prove (i), assume that the gi are positive definite and that g12 = g22 .
We may write gi = ki ai ki−1 , where ai is diagonal with positive entries, and we
may arrange it so the entries in ai are in descending order. Since a21 and a22 are
similar diagonal matrices with their entries in descending order, they are equal,
and since the squaring map on the positive reals is injective, a1 = a2 . Denote
13 The Universal Cover 89
and
π1 SL(n, R) ∼= π1 SO(n) .
We have omitted GL(n, R) from this list because it is not connected. There is
a general principle here: the fundamental group of a connected Lie group is
often the same as the fundamental group of a maximal compact subgroup.
Proof. First, let G = GL(n, C), K = U(n), and P be the space of positive
definite Hermitian matrices. By the Cartan decomposition, multiplication K ×
P −→ G is a bijection, and in fact, a homeomorphism, so it will follow that
π1 (K) ∼
= π1 (G) if we can show that P is contractible. However, the exponential
map from the space p of Hermitian matrices to P is bijective (in fact, a
homeomorphism) by Proposition 13.7, and the space p is a real vector space
and hence contractible.
For G = SL(n, C), one argues similarly, with K = SU(n) and P the space of
positive definite Hermitian matrices of determinant one. The exponential map
from the space p of Hermitian matrices of trace zero is again a homeomorphism
of a real vector space onto P .
Finally, for G = SL(n, R), one takes K = SO(n), P to be the space of
positive definite real matrices of determinant one, and p to be the space of
real symmetric matrices of trace zero.
The remainder of this chapter will be less self-contained, but can be skipped
with no loss of continuity. We will calculate the fundamental groups of SO(n)
and SU(n), making use of some facts from algebraic topology that we do not
prove. (These fundamental groups can alternatively be computed using the
method of Chap. 23. See Exercise 23.4.)
If G is a Hausdorff topological group and H is a closed subgroup, then
the coset space G/H is a Hausdorff space with the quotient topology. Such a
quotient is called a homogeneous space.
Theorem 13.6. The groups SU(n) are simply connected for all n. On the
other hand,
Z if n = 2 ,
π1 SO(n) ∼ =
Z/2Z if n > 2.
Proof.
Since its
fundamental group is Z. By Proposition 13.6
SO(2) is a circle,
π1 SO(3) ∼ = Z/2Z and π1 SU(2) is trivial. The group SO(n) acts transitively
on the unit sphere S n−1 in Rn , and the isotropy subgroup is SO(n − 1), so
SO(n)/SO(n
− 1) is homeomorphic
to S n−1 . By Proposition 13.8, we see that
∼
π1 SO(n) = π1 SO(n − 1) if n 4. Similarly, SU(n) acts on the unit sphere
S 2n−1 in Cn , and so SU(n)/SU(n − 1) ∼= S 2n−1 , whence SU(n) ∼ = SU(n − 1)
for n 2.
If n , the universal covering group of SO(n) is called the spin group and is
denoted Spin(n). We will take a closer look at it in Chap. 31.
Exercises
R) be the universal covering group of SL(2, R). Let π :
Exercise 13.1. Let SL(2,
R) −→ GL(V ) be any finite-dimensional irreducible representation. Show that
SL(2,
π factors through SL(2, R) and is hence not a faithful representation. (Hint: Use
Exercise 12.2.)
14
The Local Frobenius Theorem
Proof. Any vector field subordinate to D has the form (locally near x)
i fi Xi , where fi are smooth functions. To check that the commutator of
two such vector fields is also of the same form amounts to using the formula
which follows easily on applying both sides to a function h and using the fact
that X and Y are derivations of C ∞ (M ).
Proof. Since this is a strictly local statement, it is sufficient to prove this when
M is an open set in Rn and x is the origin.
We show first that if X is a vector field that does not vanish at x, then
we may find a system y1 , . . . , yn of coordinates in which X = ∂/∂yn . Let
x1 , . . . , xn be the standard Cartesian functions. Since X does not vanish at
the origin, the function X(xi ) does not vanish at the origin for some i, so after
permuting the variables if necessary, we may assume that X(xn )
= 0. Write
∂ ∂
X = a1 + · · · + an
∂x1 ∂xn
in terms of smooth functions ai = ai (x1 , . . . , xn ). Then an (0, . . . , 0)
= 0.
The new coordinate system y1 , . . . , yn will have the property that
This path is to be an integral curve for the vector field through the point
(u1 , . . . , un−1 , 0). By Proposition 8.1 a unique such path exists (for t small).
14 The Local Frobenius Theorem 95
Thus, we have a path that is (in the x coordinates) t −→ (x1 (t), . . . , xn (t)) ,
satisfying the first-order system
xi (t) = ai x1 (t), . . . , xn (t) , (14.1)
(xi (0), . . . , xn (0)) = (u1 , . . . , un−1 , 0).
Now ∂xi /∂yn = ai because the partial derivative is the derivative along one
of the paths (14.1). Thus
∂ ∂xi ∂ ∂
= = ai = X.
∂yn i
∂y n ∂xi i
∂x i
d−1
[Xd , Xi ] = gij Xj , (i < d). (14.2)
j=1
Indeed, writing
n
∂
Xi = hik , (i = 1, . . . , d − 1), (14.3)
∂yk
k=1
n−1
∂
Xi = hik , (i = 1, . . . , d − 1). (14.4)
∂yk
k=1
In other words, we may assume that Xi does not involve ∂/∂yn for i < d.
Now
∂hik ∂
n−1
[Xd , Xi ] = . (14.5)
∂yn ∂yj
k=1
96 14 The Local Frobenius Theorem
d−1 n−1
d−1 ∂ ∂
[Xd , Xi ] = gij Xj + gid Xd = gij hjk + gid .
j=1 j=1 k=1
∂yk ∂yn
and - .
d−1
Xd , fi Xi = 0.
i=1
Indeed,
- .
d−1
d=1
∂fi
d−1
Xd , f i Xi = Xi + fi gij Xj .
i=1 i=1
∂yn i,j=1
∂fj
d−1
+ gij fi = 0, j = 1, . . . , d − 1.
∂yn i=1
This first-order system has a solution locally with the prescribed initial con-
dition.
Since the ci can be arbitrary, we may choose
1 if i = 1,
ci =
0 otherwise.
Then the vector field fi Xi agrees with X1 on the hyperplane yn = 0.
Replacing X1 by fi Xi , we may therefore assume that [Xd , X1 ] = 0.
Repeating this process, we may similarly assume that [Xd , Xi ] = 0 for all
i < d. Now with the hij as in (14.3), this means that ∂hij /∂yn = 0, so the hij
are independent of yn .
Since the hij are independent of yn , we may interpret (14.4) as defining
d − 1 vector fields on Rn−1 . They span a (d − 1)-dimensional involutory family
of tangent vectors in Rn−1 and by induction there exists an integral manifold
for this vector field. If this manifold is N0 ⊂ Rn−1 , then it is clear that
Theorem 14.2. Let G and H be Lie groups with Lie algebras g and h, respec-
tively, and let π : g −→ h be a Lie algebra homomorphism. Assume that G is
simply connected. Then there exists a Lie group homomorphism π : G −→ H
with differential π.
We can now give another proof of Theorem 12.2. The basic idea here is to
use a compact subgroup to prove the complete reducibility of some class of
representations of a noncompact group. This idea was called the “Unitarian
Trick” by Hermann Weyl. We will extend the validity of Theorem 12.2, though
the algebraic method would work as well for this.
Theorem 14.3. Let G and K be Lie groups with Lie algebras g and k. Assume
K is compact and simply connected. Suppose that g and k have isomorphic
complexifications. Then every finite-dimensional irreducible complex represen-
tation of g is completely reducible. If G is connected, then every irreducible
complex representation of G is completely reducible.
Proof. We will prove this for sl(n, R) and su(n). By Theorem 13.6, K is
simply-connected and the hypotheses of Theorem 14.3 are satisfied. For
sl(n, R), we can take G = SL(n, R), K = SU(n). For su(n), we can take
G = K = SU(n).
The case of sl(n, C) requires a minor modification to Theorem 14.3 and is
left to the reader.
Exercises
Exercise 14.1. Let G be a connected complex analytic Lie group, and let K ⊂ G be
a compact Lie subgroup. Let g and k ⊂ g be the Lie algebras of G and K, respectively.
Assume that g is the complexification of k and that K is simply-connected. Prove
that every finite-dimensional irreducible complex representation of g is completely
reducible. If G is connected, then every irreducible complex analytic representation
of G is completely reducible.
15
Tori
d
f p(t) = (Y f ) p(t) .
dt
Let q(t, u) = etX euY . The vector field Y is invariant under left translation, in
particular left translation by etX , so
∂
f q(t, u) = (Y f )(etX euY ).
∂u
Similarly (making use of etX euY = euY etX ),
∂
f q(t, u) = (Xf )(etX euY ).
∂t
Now, by the chain rule,
d
∂
∂
f q(v, v) = f q(t, u) + f q(t, u)
dv ∂t t=u=v
∂u t=u=v
= (Y f + Xf ) q(v, v) .
This means that the path v −→ r(v) = q(v, v) satisfies r∗ (d/dv) = (X +Y )r(v)
whence ev(X+Y ) = evX evY . Taking v = 1, the result is proved.
A compact torus is a compact connected Lie group that is Abelian. In the
context of Lie group theory a compact torus is usually just called a torus,
though in the context of algebraic groups the term “torus” is used slightly
differently.
For example, T = {z ∈ C× |z| = 1} is a torus. This group is isomorphic to
R/Z. Even though R and Z are additive groups, we may, during the following
discussion, sometimes write the group law in R/Z multiplicatively.
Proposition 15.3. Let T be a torus, and let t be its Lie algebra. Then exp :
= (R/Z)r ∼
t −→ T is a homomorphism, and its kernel is a lattice. We have T ∼ =
T , where r is the dimension of T .
r
Proposition 15.5. Let T = (Z/R)r and let (π, V ) be an irreducible real rep-
resentation. Then either π is trivial or π is two-dimensional and is one of
the irreducible representations (15.2) with ki ∈ Z not all zero. In the two-
dimensional case the complexified module VC = C ⊗ V decomposes into two
one-dimensional representations corresponding to a character and its inverse.
Proof. Without loss of generality, we may assume that T = (R/Z)r and that
TC = (C× )r , where the embedding T −→ TC is the map (x1 , . . . , xr ) −→
(e2πix1 , . . . , e2πixr ). Every linear character of T is given by (15.1) /
for suitable
ki ∈ Z, and this extends to the rational character (t1 , . . . , tr ) −→ tki i of TC .
Since a rational character is holomorphic, it is determined by its values on the
image Tr of T .
We can express (ii) by saying that Aut(T ) is discrete since if it is given the
discrete topology, then (h, t) −→ f (h)t is continuous if and only if f is locally
constant.
In the remainder of this chapter, we will consider tori embedded in Lie groups.
First, we prove a general statement that implies the existence of tori.
Theorem 15.2. Let G be a Lie group and H a closed Abelian subgroup. Then
H is a Lie subgroup of G. If G is compact, then the connected component of
the identity in H is a torus.
The assumption that H is Abelian is unnecessary. See Remark 7.2 for refer-
ences to a result without this assumption.
and hn by eZn , and we still have hn −→ 1, but log(hn ) ∈ / h, and after this
substitution we have Xn ∈ p.
Let us put an inner product on g. We choose it so that the unit ball
is contained in U . The vectors Xn /|Xn | lie on the unit ball in p, which is
compact, so they have an accumulation point. Passing to a subsequence, we
may assume that Xn /|Xn | −→ X∞ , where X∞ lies in the unit ball in p. We
will show that X∞ ∈ h, which is a contradiction since h ∩ p = {0}.
To show that X∞ ∈ h, we must show that etX∞ ∈ H. It is sufficient
to show this for t < 1. With t fixed, let rn be the smallest integer greater
than t/|Xn |. Since Xn → 0 we have rn |Xn | → t. Thus, rn Xn −→ tX∞ and
ern Xn = (eXn )rn ∈ H since eXn ∈ H. Since H is closed, etX∞ ∈ H and the
proof that H is a Lie group is complete.
If G is compact, then so is H. The connected component of the identity
in H is a connected compact Abelian Lie group and hence a torus.
If G is a group and H a subgroup, we will denote by NG (H) and CG (H) the
normalizer and centralizers of H. If no confusion is possible, we will denote
them as simply N (H) and C(H).
Let G be a compact, connected Lie group. It contains tori, for example
{1}, and an ascending chain T1 T2 T3 · · · has length bounded by
the dimension of G. Therefore, G contains maximal tori. Let T be a maximal
torus.
The normalizer N (T ) = {g ∈ G | gT g −1 = T }. It is a closed subgroup since
if t ∈ T is a generator, N (T ) is the inverse image of T under the continuous
map g −→ gtg −1 .
Proposition 15.8. Let G be a compact Lie group and T a maximal torus.
Then N(T) is a closed subgroup of G. The connected component N (T )◦ of the
identity in N (T ) is T itself. The quotient N (T )/T is a finite group.
Proof. We have a homomorphism N (T ) −→ Aut(T ) in which the action is
by conjugation. By Proposition 15.7, Aut(T ) ∼ = GL(r, Z) is discrete, so any
connected group of automorphisms must act trivially. Thus, if n ∈ N (T )◦ , n
commutes with T . If N (T )◦
= T , then it contains a one-parameter subgroup
R t −→ n(t), and the closure of the group generated by T and n(t) is a
closed commutative subgroup strictly larger than T . By Theorem 15.2, it is a
torus, contradicting the maximality of T . It follows that T = N (T )◦ .
The quotient group N (T )◦ /T is both discrete and compact and hence
finite.
The quotient N (T )/T is called the Weyl group of G with respect to T .
Example 15.1. Suppose that G = U(n). A maximal torus is
⎧⎛ ⎞ ⎫
⎪
⎨ t1 ⎪
⎬
⎜ . ⎟
T = ⎝ .. ⎠ |t1 | = · · · = |tn | = 1 .
⎪
⎩ ⎪
⎭
tn
15 Tori 107
Exercises
Exercise 15.1. Compute the dimensions of the flag manifolds for su(n), sp(2n) and
so(n).
16
Geodesics and Maximal Tori
Thus, the matrix (gij ) representing the inner product is positive definite sym-
metric. Smoothness of the inner product means that the gij are smooth func-
tions of x ∈ U .
We also define (g ij ) to be the inverse matrix to (gij ). Thus, the functions
ij
g satisfy
1 if i = k ,
jk k k
gij g = δi , where δi = (16.2)
0 otherwise ,
j
and of course
gij = gji , g ij = g ji .
Suppose that p : [0, 1] −→ M is a path in the Riemannian manifold M . We say
p is admissible if it is smooth, and moreover the movement along the path
never “stops,” that is, the tangent vector p∗ (d/dt), where t is the coordinate
function on [0, 1], is never zero. The length or arclength of p is
1
|p| = p∗ d dt. (16.3)
dt
0
In terms of local coordinates, if we write xi (t) = xi p(t) the integrand is
∂xi ∂xj
p∗ d = gij .
dt ∂t ∂t
i,j
for all 0 a 1. Intuitively, this means that the point p(t) moves along the
path at a constant “velocity.”
It is an easy application of the chain rule that the arclength of p is
unchanged under reparametrization. Moreover, each path has a unique rep-
arametrization that is well-paced.
A Riemannian manifold becomes a complete metric space by defining the
distance between two points a and b as the infimum of the lengths of the paths
connecting them. It is not immediately obvious that there will be a shortest
path, and indeed there may not be for some Riemannian manifolds, but it is
easy to check that this definition satisfies the triangle inequality and induces
the standard topology.
We will encounter various quantities indexed by 1 i, j, k, · · · n, where
n is the dimension of the manifold M under consideration. We will make use
of Einstein’s summation convention (in this chapter only). According to this
convention, if any index is repeated in a term, it is summed. For example,
suppose that p : [0, 1] −→ M is a path lying entirely in a single chart U ⊂
M with coordinate functions x1 , . . . , xn . Then
we may regard x1 , . . . , xn as
functions of t ∈ [0, 1], namely xi (t) = xi p(t) . If f : U −→ C is a smooth
function, then according to the chain rule
df
n
dxi ∂f
x1 (t), . . . , xn (t) = x1 (t), . . . , xn (t) .
dt i=1
dt ∂xi
According to the summation convention, we can write this as simply
df dxi ∂f
= ,
dt dt ∂xi
and the summation over i is understood because it is a repeated index.
16 Geodesics and Maximal Tori 111
b
a
c
d2 xk dxi dxj
2
= −{ij, k} . (16.4)
dt dt dt
Proof. Let us consider the effect of deforming the path. We consider a family
pu of paths parametrized by u ∈ (−, ), where > 0 is a small real number.
It is assumed that the family of paths varies smoothly, so (t, u) −→ pu (t) is a
smooth map (−, ) × [0, 1] −→ M .
We regard the coordinate functions xi of the point x = pu (t) to be func-
tions of u and t.
It is assumed that p0 (t) = p(t) and that the endpoints are fixed, so that
pu (0) = p(u) and pu (1) = p(1) for all u ∈ (−, ). Therefore,
∂xi
=0 when t = 0 or 1. (16.5)
∂u
In local coordinates, the arclength (16.3) becomes
15
∂xi ∂xj
|pu | = gij dt. (16.6)
0 ∂t ∂t
Because the path p(t) = p0 (t) is well-paced, the integrand is constant (inde-
pendent of t) when u = 0, so
5
∂ ∂xi ∂xj
gij = 0 when u = 0. (16.7)
∂t ∂t ∂t
We do not need to assume that the deformed path p(t, u) is well-paced for
any u
= 0.
Let f (u) = |pu |. We have
15
∂ ∂xi ∂xj
f (u) = gij dt .
∂u 0 ∂t ∂t
This equals
1 − 12 !
∂xi ∂xj 1 ∂gij ∂xi ∂xj 1 ∂ 2 xi ∂xj 1 ∂xi ∂ 2 xj
gij + gij + gij dt
0 ∂t ∂t 2 ∂u ∂t ∂t 2 ∂u∂t ∂t 2 ∂t ∂u∂t
1 − 12 !
∂xi ∂xj 1 ∂gij ∂xl ∂xi ∂xj ∂ 2 xi ∂xj
= gij + gij dt,
0 ∂t ∂t 2 ∂xl ∂u ∂t ∂t ∂u∂t ∂t
where we have used the chain rule, and combined two terms that are equal.
(The variables i and j are summed by the summation convention, so we may
16 Geodesics and Maximal Tori 113
interchange them, and using gij = gji , the last two terms on the left-hand side
are equal.) We integrate the second term by parts with respect to t, making
use of (16.5) and (16.7) to obtain
1 − 12 !
∂xi ∂xj 1 ∂gij ∂xl ∂xi ∂xj ∂xi ∂ ∂xj
f (0) = gij − gij dt
0 ∂t ∂t 2 ∂xl ∂u ∂t ∂t ∂u ∂t ∂t
1 − 12 !
∂xi ∂xj 1 ∂gij ∂xi ∂xj ∂ ∂xj ∂xl
= gij − glj dt.
0 ∂t ∂t 2 ∂xl ∂t ∂t ∂t ∂t ∂u
Now all the partial derivatives are evaluated when u = 0. The last step is just
a relabeling of a summed index.
We observe that the displacements ∂xl /∂u are arbitrary except that they
must vanish when t = 0 and t = 1. (We did not assume the deformed path to
be well-paced except when u = 0.) Thus, the path is of stationary length if
and only if
1 ∂gij ∂xi ∂xj ∂ ∂xj
0= − glj ,
2 ∂xl ∂t ∂t ∂t ∂t
so the condition is
∂ 2 xj 1 ∂gij ∂xi ∂xj ∂glj ∂xj
glj 2
= − .
∂t 2 ∂xl ∂t ∂t ∂t ∂t
Now !
∂glj ∂xj ∂glj ∂xi ∂xj 1 ∂glj ∂gli ∂xi ∂xj
= = + .
∂t ∂t ∂xi dt ∂t 2 ∂xi ∂xj dt ∂t
The two terms on the right-hand side are of course equal since both i and j
are summed indices. We obtain in terms of the Christoffel symbols
∂ 2 xj ∂xi ∂xj
glj 2
= −[ij, l] .
∂t ∂t ∂t
Multiplying by g kl , summing the repeated index l, and using (16.2), we obtain
(16.4).
dxi
= yi ,
dt
dyk
= −{ij, k} yiyj .
dt
The conditions p(0) = x and p∗ (d/dt) = X amount to initial conditions for
this first-order system, and the existence and uniqueness of the solution follow
from the general theory of first-order systems.
In view of (16.1), this amounts to saying that the geodesic curves (having
tangent vector ∂/∂x1 ) are orthogonal to the level hypersurfaces x1 = constant
(having tangent spaces spanned by ∂/∂x2 , . . . , ∂/∂xn ), such as U and U in
Fig. 16.2.
Proof. Having chosen coordinates so that the path t −→ (t, x2, . . . , xn ) is a
geodesic, we see that if all dxi /dt = 0 in (16.4), for i
= 1, then d2 xk /dt2 = 0
for all k. This means that {11, k} = 0. Since the matrix (gkl ) is invertible, it
follows that [11, k] = 0, so
∂g1k 1 ∂g11
= . (16.8)
∂x1 2 ∂xk
First, take k = 1 in (16.8). We see that ∂g11 /∂x1 = 0, so if x2 , . . . , xn are held
constant, g11 is constant. When x1 = 0, the initial condition of the geodesic
curve px through (0, x2 , . . . , xn ) is that it is tangent to the unit normal to
16 Geodesics and Maximal Tori 115
the surface, that is, its tangent vector ∂/∂x1 has length one, and by (16.1)
it follows that g11 = 1 when x1 = 0, so g11 = 1 throughout the geodesic
coordinate neighborhood.
Now let 2 k n in (16.8). Since g11 is constant, ∂g1k /∂x1 = 0, and
so g1k is also constant when x2 , . . . , xn are held constant. When x1 = 0, our
assumption that the geodesic curve px is normal to the surface means that
∂/∂x1 and ∂/∂xk are orthogonal, so by (16.1), g1k vanishes when x1 = 0 and
so it vanishes for all x1 .
With these preparations, we may now prove that short geodesics are paths
of shortest length.
Proposition 16.4.
(i) Let p : [0, 1] −→ M be a geodesic. Then there exists an > 0 such that
the restriction of p to [0, ] is the unique path of shortest length from p(0)
to p().
(ii) Let x ∈ M . There exists a neighborhood N of x such that for all y ∈ N
there exists a unique path of shortest distance from x to y, and that path
is a geodesic.
Proof. We choose a hypersurface U orthogonal to p at t = 0 and construct
geodesic coordinates as explained before Proposition 16.3. We choose and
B so small that the set N of points with coordinates {x1 ∈ [0, ], 0
|x2 |, . . . , |xn | B} is contained within the interior of this geodesic coordi-
nate neighborhood. We can assume that the coordinates of p(0) are (0, . . . , 0),
so by construction p(t) = (t, 0, . . . , 0). Then |p| = , where now |p| denotes
the length of the restriction of the path to the interval from 0 to .
We will show that if q : [0, ] −→ M is any path with q(0) = p(0) and
q() = p(), then |q| |p|.
First, we consider paths q : [0, ] −→ M that lie entirely within the
set N and such that the x1 -coordinate of q(t) is monotonically increasing.
Reparametrizing q, we may arrange that q(t) and p(t) have the
same x1 -
coordinate, which equals t. Let us write q(t) = t, x2 (t), . . . , xn (t) . We also
denote x1 (t) = t. Since g1k = gk1 = 0 when k 2 and g11 = 1, we have
116 16 Geodesics and Maximal Tori
dxi dxj
|q| = gij dt
0 i,j
dt dt
dxi dxj
= 1+ gij dt.
0 dt dt
2i,jn
Now since the matrix (gij )1i,jn is positive definite, its principal minor
(gij )2i,jn is also positive definite, so
dxi dxj
gij 0
dt dt
2i,jn
and √
|q| 1 dt = = |p|.
0
This argument is easily extended to include all paths such that the values
of x1 for those t such that q(t) ∈ N cover the entire interval [0, ]. Paths for
which this is not true must be long enough to reach the edges of the box
xi > B, and after reducing if necessary, they must be longer than . This
completes our discussion of (i).
For (ii), given each unit tangent vector X ∈ Tx (M ), there is a unique
geodesic pX : [0, X ] −→ M through x tangent to X, and X > 0 may
be chosen so that this geodesic is a path of shortest length. We assert that
X may be chosen so that the same value X is valid for nearby unit tangent
vectors Y . We leave this point to the reader except to remark that it is perhaps
easiest to see this by applying a diffeomorphism of M that moves X to Y and
regarding X as fixed while the metric gij varies; if Y is sufficiently near X, the
variation of gij will be small and the in part (i) can be chosen to work for
small variations of the gij . So for each unit tangent vector X ∈ Tx (M ) there
exists an X > 0 and a neighborhood NX of X in the unit ball of Tx (M ) such
that pY : [0, X ] −→ M is a path of shortest length for all Y ∈ NX . Since the
unit tangent ball in Tx (M ) is compact, a finite number of NX suffice to cover
it, and if is the minimum of the corresponding X , then we can take N to
be the set of all points connected to x by a geodesic of length < .
If M is a connected Riemannian manifold, we make M into a metric space
by defining d(x, y) to be the infimum of |p|, where p is a smooth path from x
to y.
Theorem 16.1. Let M be a compact connected Riemannian manifold, and
let x and y be points of M . Then there is a geodesic p : [0, 1] −→ M with
p(0) = x and p(1) = y.
A more precise statement may be found in Kobayashi and Nomizu [110], The-
orem 4.2 on p. 172. It is proved there that if M is connected and geodesically
complete, meaning that any well-paced geodesic can be extended to (−∞, ∞),
then the conclusion of the theorem is true. (It is not hard to see that a compact
manifold is geodesically complete.)
16 Geodesics and Maximal Tori 117
itself. By a linear change of variables, we may assume that the inner product
on Rn = Te (G) corresponding to the Riemannian structure is the standard
Euclidean inner product. Since the Riemannian structure is invariant under
translation it follows that G ∼= Rn /Λ is a Riemannian manifold as well as a
group. Geodesics are straight lines and so are translates of the exponential
map.
We turn now to the general case. If X ∈ g, let EX : (−, ) −→ G denote
the geodesic through the origin tangent to X ∈ g. It is defined for sufficiently
small (depending on X). If λ ∈ R, then t −→ EX (λt) is the geodesic through
the origin tangent to λX, so EX (λt) = EλX (t). Thus, there is a neighborhood
U of the origin in g and a map E : U −→ G such that EX (t) = E(tX) for
X, tX ∈ U . We must show that E coincides with the exponential map.
118 16 Geodesics and Maximal Tori
Theorem 16.3. Let G be a compact Lie group and g its Lie algebra. Then
the exponential map g −→ G is surjective.
[H, X], Y = 0
for all Y . Since an inner product is by definition positive definite, the bilinear
form , is nondegenerate, which implies that [H, X] = 0. Now, by Proposi-
tion 15.2, eH commutes with etX for all t ∈ R. Since eH generates the maximal
16 Geodesics and Maximal Tori 119
Theorem 16.5 (E. Cartan). Let G be a compact connected Lie group, and
let T be a maximal torus. Then every maximal torus is conjugate to T , and
every element of G is contained in a conjugate of T .
Proof. The second statement is contained in Theorem 16.4. As for the first
statement, let T be another maximal torus, and let t be a generator. Then
t is contained in kT k −1 for some k, so T ⊆ kT k −1. Since both are maximal
tori, they are equal.
Exercises
Exercise 16.1. Give an example of a connected Riemannian manifold with two
points P and Q such that no geodesic connects P and Q.
Exercise 16.2. Let G be a compact connected Lie group and let g ∈ G. Show that
the centralizer CG (g) of g is connected.
Exercise 16.3. Show that the conclusion of Exercise 16.2 fails for the connected
noncompact Lie group SL(2, R) by exhibiting an element with a centralizer that is
not connected.
a b
a, b ∈ C, |a| + |b| = 1 ,
2 2
SU(2) =
−b̄ ā
ab
SU(1, 1) = a, b ∈ C, |a|2
− |b|2
= 1 ,
b̄ ā
and
a0 2
K= |a| = 1 ∼= U(1).
0 ā
It will be shown in Chap. 28 that the group SU(1, 1) is conjugate in SL(2, C) to
SL(2, R). Let G be one of the groups SU(2), A, or SU(1, 1). The stabilizer of 0 ∈ R
is the group K, so we may identify the orbit of 0 ∈ R with the homogeneous space
G/H by the bijection g(0) ←→ gH. The orbit of 0 is given in the following table.
G K orbit of 0 ∈ R
SU(1, 1) U(1) D
A U(1) C
SU(2) U(1) H
Exercise 16.5. Show that if G is one of the groups SU(1, 1), A, or SU(2), then the
quotient G/K, which we may identify with D, C, or H, has a unique G-invariant
Riemannian structure.
Exercise 16.6. Show that the inclusions D −→ C −→ R are conformal maps but
are not isometric.
Exercise 16.7.
(i) Show that the group SL(n, C) preserves the set of circles. Show, however, that
a linear fractional transformation g ∈ SL(n, C) may take a circle with center α
to a circle with center different from g(α).
(ii) Show that if M = D, C or R, then each geodesic is a circle, but not each circle
is a geodesic.
(iii) Show that the geodesics in C are the straight lines and that the geodesics in D
are the curves C ∩ D, where C is a circle in C perpendicular to ∂D.
(iv) Show that ∂D is a geodesic in R.
17
The Weyl Integration Formula
Proof. We may choose the normalization of dμH so that H has total volume 1.
We define a map λ : Cc (G) −→ Cc (G/H) by
(λf )(g) = f (gh) dμH (h).
H
provided we check that this is well defined. We must show that if λf = 0 then
f (g) dμG (g) = 0. (17.1)
G
We note that the function (g, h) −→ f (gh) is compactly supported and con-
tinuous on G × H, so if λf = 0 we may use Fubini’s theorem to write
0= (λf )(g) dμG (g) = f (gh) dμG (g) dμH (h).
G H G
In the inner integral on the right-hand side we make the variable change
g −→ gh−1 . Recalling that dμG (g) is left Haar measure, this produces a
factor of δG (h), where δG is the modular quasicharacter on G. Thus,
0= δG (h) f (g) dμG (g) dμH (h).
H G
Theorem 17.1.
(i) Two elements of T are conjugate in G if and only if they are conjugate
in N (T ).
(ii) The inclusion T −→ G induces a bijection between the orbits of W on T
and the conjugacy classes of G.
Proposition 17.3. There exists a dense open set Ω of T such that the |W |
elements wtw−1 (w ∈ W ) are all distinct for t ∈ Ω.
Theorem 17.2 (Weyl). Let G be a compact connected Lie group, and let p
be as in Proposition 15.9. If f is a class function, and if dg and dt are Haar
measures on G and T (normalized so that G and T have volume 1), then
1
f (g) dg = f (t) det [Ad(t−1 ) − Ip ] p dt.
G |W | T
Proof. Let X = G/T . We give X the measure dX invariant under left trans-
lation by G such that X has volume 1. Consider the map
φ : X × T −→ G, φ(xT, t) = xtx−1 .
126 17 The Weyl Integration Formula
To compute the Jacobian of this map, we translate on the left by t−1 x−1 and
on the right by x. There is no harm in this because these maps are Haar
isometries. We are reduced to computing the Jacobian of the map
−1
(U, V ) → t−1 eU teV e−U = eAd(t )U V −U
e e .
Identifying the tangent space of the real vector space p × t with itself (that is,
with g = p ⊕ t), the differential of this map is
U + V → Ad(t−1 ) − Ip U + V.
Proposition 17.4. Let G = U(n), and let T be the diagonal torus. Writing
⎛ ⎞
t1
⎜ .. ⎟
t=⎝ . ⎠ ∈ T,
tn
and letting T dt be the Haar measure on T normalized so that its volume is
1, we have
⎛ ⎞
t1
1 ⎜ .. ⎟7
f (g) dg = f⎝ . ⎠ |ti − tj |2 dt. (17.3)
G n! T i<j
tn
Proof. This will follow from Theorem 17.2 once we check that
7
det [Ad(t−1 ) − Ip ] | p = |ti − tj |2 .
i<j
Exercises
Exercise 17.1. Let G = SO(2n + 1). Choose the realization of Exercise 5.3.
Show that
⎛ ⎞
t1
⎜ .. ⎟
⎜ ⎟
⎜ . ⎟
⎜ ⎟
⎜ tn ⎟
1 ⎜ ⎟
f (g) dg = n f⎜ 1 ⎟
2 n! Tn ⎜ −1 ⎟
SO(2n+1)
⎜ tn ⎟
⎜ ⎟
⎜ .. ⎟
⎝ . ⎠
t−1
1
2
× |ti − tj |2 |ti − t−1
j | |ti − 1|2 dt1 · · · dtn .
i<j i
128 17 The Weyl Integration Formula
Exercise 17.2. Let G = SO(2n). Choose the realization of Exercise 5.3. Show that
⎛ ⎞
t1
⎜ .. ⎟
⎜ ⎟
⎜ . ⎟
⎜ ⎟
1 ⎜ tn ⎟
f (g) dg = n−1 f⎜ −1 ⎟
2 n! Tn ⎜ tn ⎟
SO(2n)
⎜ ⎟
⎜ .. ⎟
⎝ . ⎠
−1
t1
2
× |ti − tj |2 |ti − t−1
j | dt1 · · · dtn .
i<j
Exercise 17.3. Describe the Haar measure on Sp(2n) as an integral over the diag-
onal maximal torus.
For the first proof, check that this is true for every irreducible character. For the
second proof, show that a(n) = a(−n). Then use the Weyl integration formula and
make use of the fact that a(2) = a(−2).
A Euclidean space is a real vector space V endowed with an inner product, that
is, a positive definite symmetric bilinear form. We denote this inner product
by , . If 0
= α ∈ V, consider the transformation sα : V −→ V given by
2 x, α
sα (x) = x − α. (18.1)
α, α
This is the reflection attached to α. Geometrically, it is the reflection in the
plane perpendicular to α. We have sα (α) = −α, while any element of that
plane (with x, α = 0) is unchanged by sα .
Definition 18.1. Let V be a finite-dimensional real Euclidean space, Φ ⊂ V
a finite subset of nonzero vectors. Then Φ is called a root system if for all
α ∈ Φ, sα (Φ) = Φ, and if α, β ∈ Φ then 2 α, β / α, α ∈ Z. The root system
is called reduced if α, λα ∈ Φ, λ ∈ R implies that λ = ±1.
There is another, more modern notion which was introduced in Demazure [10]
(Exposé XXI). This notion is known as a root datum. We will give the
definition, then discuss the relationship between the two notions. We will
find both structures in a compact Lie group.
A root datum consists of a quadruple (Λ, Φ, Λ∨ , Φ∨ ) of data which are to
be as follows. First, Λ is a lattice, that is, a free Z-module, and let Λ∨ =
Hom(Λ, Z) is the dual lattice. Inside each lattice there is given a finite set
of nonzero vectors, denoted Φ ⊂ Λ and Φ̂ ⊂ Λ∨ , together with a bijection
α → α∨ from Φ to Φ̂. It is required that α∨ (α) = 2 and that α∨ (Φ) ⊂ Z.
Using these we may define, for each α ∈ Φ, linear maps sα : Λ → Λ and
sα∨ : Λ∨ → Λ∨ of order 2. These are defined by the formulas
Let us now explain the relationship between the root system and the root
datum. We will always obtain the root system with another piece of data: a
lattice Λ that spans V such that Φ ⊂ Λ. It will have the property of being
invariant under the sα . Let V ∗ be the real dual space of V. The dual lattice
Λ∨ is the set of linear functionals v ∗ : V → R such that v ∗ (L) ⊂ Z. It can be
identified with Hom(Λ, Z). If α ∈ Φ the linear functional
2 x, α
α∨ (x) = (18.2)
α, α
(See Proposition 18.13.) In view of this close relationship both the α∨ and the
Hα may be referred to as coroots.
Define
1 01 00
Hα = , Xα = , X−α = . (18.6)
−1 00 10
We will see that Hα ∈ it is the coroot, and that Xα and X−α span the
one-dimensional weight spaces Xα and X−α . Thus, the root system Φ =
{α, −α}.
Let us say that λk is the highest weight in an irreducible representation
V if k is maximal such that λk occurs in the restriction of the representation
to V . This ad hoc definition is a special case of a partial order on the weights
for general compact Lie groups in Chap. 21.
This is not the same as the group introduced in Example 5.5, but it is
conjugate to that group in GL(4, C). The subgroup Sp(4) is the intersection
of Sp(4, C) with U(4). A maximal torus T can be taken to be the group of
diagonal elements, and the we will show that the roots are the eight characters
18 The Root System 133
⎧
⎪
⎪ α1 (t) = t1 t−1 2 ,
⎪
⎪ 2
⎛ ⎞ ⎪
⎪ α2 (t) = t 2 ,
⎪
⎪
t1 ⎪
⎪ (α1 + α2 )(t) = t1 t2 ,
⎜ ⎟ ⎨ (2α + α )(t) = t2 ,
t2
T t=⎜ ⎟ −→ 1 2 1
⎝ t−1 ⎠ ⎪ −α (t) = t −1
1 t2 ,
2 ⎪
⎪
1
−1 ⎪
⎪ −α (t) = t −2
t1 ⎪ 2 2 ,
⎪
⎪ −1 −1
⎪
⎪ −(α + α )(t) = t 1 t2 ,
⎩ 1 2
−(2α1 + α2 )(t) = t−21 .
They form a configuration in V that can be seen in Fig. 19.4 of the next
chapter. The reader can check that this forms a root system.
The complexified Lie algebra gC consists of matrices of the form
⎛ ⎞
t1 x12 x13 x14
⎜ x21 t2 x23 x13 ⎟
⎜ ⎟
⎝ x31 x32 −t2 −x12 ⎠ . (18.7)
x41 x31 −x21 −t1
The spaces Xα1 and X−α1 are spanned by the vectors
⎛ ⎞ ⎛ ⎞
0 1 0 0 0 0 0 0
⎜0 0 0 0 ⎟ ⎜1 0 0 0⎟
Xα1 = ⎜
⎝0 0
⎟, X−α1 = ⎜ ⎟.
0 −1 ⎠ ⎝0 0 0 0⎠
0 0 0 0 0 0 −1 0
Note that Hα ∈/ t, but −iHα ∈ t, since the elements of t are diagonal and
purely imaginary. The Hα satisfy
134 18 The Root System
and they are elements of the intersection of it with the complex Lie algebra
generated by Xα and X−α . We note that Xα and X−α are only determined
up to constant multiples by the description we have given, but Hα is fully
characterized. The Hα will be constructed in Proposition 18.8 below. They
form a root system that is dual to the one we want to construct—if α is a
long root, then Hα is short, and conversely, in root systems where not all the
roots have the same length. (See Exercise 18.1.)
A key step will be to construct an element wα of the Weyl group W =
N (T )/T corresponding to the reflection sα in (18.1). In order to produce this,
we will construct a homomorphism
iα : SU(2) −→ G. Then wα ∈ N (T ) will
0 −1
then be the image of under iα .
1 0
Let us offer a word about how one can get a grip on iα . The centralizer
C(Tα ) of the kernel Tα of the homomorphism α : T −→ C× is a close relative of
this group iα (SU(2)). In fact, C(Tα ) = iα (SU(2))·T . Later, in Proposition 18.6
we will use this circumstance to show that Xα is one-dimensional, after which
the structure of C(Tα ) will become clear: since this group has only two roots
α and −α, it is itself a close relative of SU (2). Its Lie algebra contains a copy
of su(2) (Proposition 18.8) and using this fact we will be able to construct the
homomorphism iα in Theorem 18.1.
Let us take a look at the groups C(Tα ) and the homomorphisms iα in the
example of Sp(4). The subgroup Tα1 of T is characterized by t1 = t2 , so its
centralizer consists of elements of the form
⎛ ⎞
a b
⎜c d ⎟ a b
⎜ ⎟, ∈ U(2),
⎝ ∗ ∗⎠ c d
∗ ∗
where the elements marked ∗ are determined by the requirement that the
matrix be in Sp(4). The homomorphism iα1 is given by
⎛ ⎞
a b
a b ⎜c d ⎟ ab
iα1 =⎝⎜ ⎟ , ∈ SU(2).
c d a −b ⎠ cd
−c d
Lemma 18.1. Suppose that G is a compact Lie group with Lie algebra g,
π : G −→ GL(V ) a representation, and dπ : g −→ End(V ) the differential. If
v ∈ V and X ∈ g such that dπ(X)n v = 0 for any n > 1, then dπ(X)v = 0.
(ii) We have
Taking the derivative and setting t = 0, using (18.3) we obtain (18.8). When
V = gC and π = Ad, we have Xα = V (λ). Remembering that the differential
of Ad is ad (Theorem 8.2), we see that (18.9) is a special case of (18.8), and
(ii) follows.
For (iii), we have, by (18.9),
so w ∈ V (λ + α).
We may write gC = g + ig. Let c : gC −→ gC be the conjugation with respect
to g, that is, the real linear transformation X + iY −→ X − iY (X, Y ∈ g).
Although c is not complex linear, it is an automorphism of gC as a real Lie
algebra. We have c(aZ) = a · c(Z), so c is complex antilinear.
Proposition 18.4.
(i) We have c(Xα ) = X−α .
(ii) If Xα ∈ Xα , Xβ ∈ Xβ , α, β ∈ Φ, then
tC if β = −α ,
[Xα , Xβ ] ∈
Xα+β if α + β ∈ Φ .
so [Xα , c(Xα )] ∈ it. We show that [Xα , c(Xα )]
= 0. Let tα ⊂ t be the kernel
of dα. It is of course a subspace of codimension 1. Let H1 , . . . , Hr−1 be a basis.
If [Xα , c(Xα )] = 0, then denoting
Yα = 12 Xα + c(Xα ) , 1
Zα = 2i Xα − c(Xα ) , (18.10)
H1 , . . . , Hr−1 , Yα , Zα
[Yα , Zα ] = 12 H0 , [Yα , H0 ] = 0.
Proposition 18.6.
(i) If α ∈ Φ, then dim(Xα ) = 1.
(ii) If α, β ∈ Φ and α = λβ, λ ∈ R, then λ = ±1.
Thus H/Tα is a rank 1 group with maximal torus T /Tα . Its complexified
Lie
algebra is therefore three-dimensional by Proposition 18.5. However, Xλα
is embedded injectively in this complexified Lie algebra, so λ = ±1 are the
only λ, and X±α are one-dimensional.
Proposition 18.7.
(i) Let g be the Lie algebra of a compact Lie group. If X, Y ∈ g such that
[X, Y ] = cY with c a nonzero real constant, then Y = 0.
(ii) There does not exist any embedding of sl(2, R) into the Lie algebra of a
compact Lie group.
Proof. Let G be a compact Lie group with Lie algebra g. Then given any
finite-dimensional representation π : G −→ GL(V ) on a real vector space
V , there exists a positive-definite symmetric bilinear form B on V such that
B(π(g)v, π(g)w) = B(v, w) for g ∈ G, v, w ∈ V . By Proposition 10.3, we have
B(dπ(X)v, w) = −B(v, dπ(X)w) for X ∈ g. Now let us apply this with V = g
and π = Ad, so by Theorem 8.2 we have B([X, Y ], Z) = −B(Y, [X, Z]). If X
and Y such that [X, Y ] = cY with c a nonzero constant, then
cB(Y, Y ) = B([X, Y ], Y ) = −B(Y, [X, Y ]) = −cB(Y, Y ).
Since c
= 0 and B is positive definite, it follows that Y = 0. This proves (i).
As for (ii), if g contains a subalgebra
isomorphic
to sl(2, R) then we may
1 01
take X and Y to be the images of and and obtain a contra-
−1 00
diction to (i).
We remind the reader that the Lie algebra su(2) of SU(2) consists of the
trace zero skew-Hermitian matrices in Mat(2, C). The Lie algebra sl(2, C) of
SL(2, C) consists of all trace zero matrices. Any trace zero matrix may X may
be uniquely written as X1 + iX2 where X1 and X2 are in su(2), so sl(2, C) is
the complexification of su(2).
Proposition 18.8. Let α ∈ Φ and let 0
= Xα ∈ Xα . Let X−α = −c(Xα ) ∈
X−α . Then Xα and X−α generate a complex Lie subalgebra gα,C of gC isomor-
phic to sl(2, C). Its intersection gα = g ∩ gα,C is isomorphic to su(2). We may
choose Xα and the isomorphism iα : sl(2, C) −→ gα,C so that
1 0 1 0 0
iα = Hα , iα = Xα , iα = X−α , (18.11)
−1 0 0 1 0
and these generate the Lie algebra su(1, 1), which is isomorphic to sl(2, R) by
Exercise 5.9. This is a contradiction because sl(2, R) cannot be embedded in
the Lie algebra of a compact Lie group by Proposition 18.7.
Proof. H centralizes Xα and X−α by (18.9); that is, [H, Xα ] = [H, X−α ] = 0,
and it follows that [H, X] = 0 for all X ∈ gα .
We gave the ambient vector space V of the set Φ of roots an inner product
(Euclidean structure) invariant under W . The Weyl group acts on T by conju-
gation and hence it acts on X ∗ (T ). It acts on p by the adjoint representation
140 18 The Root System
(induced from conjugation) so it permutes the roots. The Weyl group elements
are realized as orthogonal motions with respect to this metric.
We may now give a method of constructing Weyl group elements. Let
α ∈ Φ. Let Tα = {t ∈ T | α(t) = 1}.
Proof. Since SU(2) is simply connected, it follows from Theorem 14.2 that the
Lie algebra homomorphism su(2) −→ g of Proposition 18.8 is the differential
of a homomorphism iα : SU(2) −→ G. By Proposition
18.9, gα centralizes tα ,
and since SU(2) is connected, it follows that iα SU(2) ⊆ C(Tα )◦ .
By Proposition 18.4, −iH
α ∈/ tα , so t is generated by its codimension-
one subspace tα and iα su(2) ∩ t.
Since Lie(Tα ) = tα , it follows that T is
generated by Tα and T ∩ iα SU(2) . By construction, wα normalizes
y
T ∩ iα SU(2) = iα y ∈ C, |y| = 1 ,
y −1
Proof. Let
&
W = V (λ + kα).
k∈Z
Theorem 18.2. If Φ is the set of roots associated with a compact Lie group
and its maximal torus T , then Φ is a reduced root system.
18 The Root System 141
Proof. First let us show that λ and α are orthogonal if and only if dλ(Hα ) = 0
with Hα as in Proposition 18.4. It is sufficient to show that the orthogonal
complement of α is contained in the kernel of this functional since both are
subspaces of codimension 1. Assuming therefore that α and λ are orthogonal,
sα (λ) = λ, and since the action of W on X ∗ (T ) and V = R⊗X ∗ (T ) is induced
by the action of W on T by conjugation, whose differential is the action of W
on t via Ad, we have
142 18 The Root System
dλ(Hα ) = dλ Ad(wα )Hα = −dλ(Hα )
by Proposition 18.12.
The result is now proved in the case where λ and α are orthogonal.
Therefore dλ(Hα ) = cα∨ (λ) for some constant c. To show that c = 1, we
take λ = α and (remembering that iHα ∈ t) check that dα(iHα ) = 2i. Indeed
we have
it
i d e d
dα iα = α iα = e2it |t=0 = 2i.
−i dt e −it dt
t=0
Proposition 18.14.
6
(i) 4α∈Φ Tα is the center Z(G).
(ii) α∈Φ Tα is the set of singular elements of T .
6
Of course, Tα = T−α , so we could equally well write Z(G) = α∈Φ+ Tα .
linear functionals dα. Let us partition t into open chambers which are the
complements of the tα . Let C be one of these. Choose the counterexample w
to minimize the number of hyperplanes tα separating the chambers C and wC.
Since wα reflects in the hyperplane tα , we must have w(C) = C. We will argue
that w = 1, which will be a contradiction. Let n ∈ N (T ) represent w. What
we need to show is that n ∈ T .
Since w has finite order and maps C to itself, and since C is convex, we may
find an element H of C such that w(H) = H; simply averaging any element
over its orbit under powers of w will produce such an H. Since H does not lie in
any of the tα , the one-parameter subgroup S = {exp(tH) | t ∈ R} ⊂ T contains
regular elements. Since Ad(n) fixes H, n is in the centralizer CG (S). We claim
that CG (S) = T . First note that if g ∈ CG (S) then gT g −1 contains regular
elements of T , so gT g −1 = T . Thus CG (S) ⊂ NG (T ). But CG (S) is connected
by Theorem 16.6, so n ∈ CG (S) ⊆ NG (T )◦ = T by Proposition 15.8. Therefore
n ∈ T , as required.
Proof. We may identify the decomposition (18.14) with the irreducible module
described in (12.1). Now Xα is iα (R) in the notation of that Proposition. Since
α + β is a root, Xβ is vk−2l with l > 0, and the nonvanishing of [Xα , Xβ ] =
ad(Xα )Xβ follows from (12.3).
Exercises
Exercise 18.1.
(i) Let Φ be a root system in a Euclidean space V and for α ∈ Φ let
2α
α∨ = .
α, α
Show that the α∨ also form a root system. Note that long roots in Φ correspond
to short vectors in Φ∨ . (Hint: Prove this first for rank two root systems, then note
that if α, β ∈ Φ are linearly independent roots the intersection of Φ with their span
is a rank two root system.)
(ii) Explain why this implies that the Hα form a root system in it.
Exercise 18.2. Analyze the root system of SO(5) similarly to the case of Sp(4) in
the text. It may be helpful to use Exercise 7.3.
19
Examples of Root Systems
It may be easiest to read the next chapter with examples in mind. In this
chapter we will describe various root systems and in particular we will illus-
trate the rank 2 root systems. Since the purpose of this chapter is to give
examples, we will state various facts here without proof. The proofs will come
in later chapters.
A root system Φ in a Euclidean space V is called reducible if we can
decompose V = V1 ⊕ V2 into orthogonal subspaces with Φ = Φ1 ∪ Φ2 , with
both Φi = Vi ∩ Φ nonempty. Then the Φi are themselves smaller root systems.
In classifying root systems, one may clearly restrict to irreducible root sys-
tems, and these were classified by Killing and Cartan. The irreducible root
systems are classified by a Cartan type which can be one of the classical Cartan
types Ar (r 1), Br (r 2), Cr (r 2) Dr (r 4), or one of the five excep-
tional types G2 , F4 , E6 , E7 and E8 . The subscript is the (semisimple) rank of
the corresponding Lie groups. We have an accidental isomorphism B2 ∼ = C2 .
The Cartan types D2 and D3 are usually excluded, but it may be helpful to
consider D2 as a synonym for the reducible Cartan type A1 × A1 (that is, A1
is the Cartan type of both Φ1 and Φ2 in the orthogonal decomposition); and
D3 as a synonym for A3 .
In the last chapter we saw how to associate a root system Φ with a compact
Lie group G. The Euclidean V containing Φ is R⊗X ∗ (T ) where T is a maximal
torus. The group G is called semisimple if the root system Φ spans V = R ⊗ Λ
where Λ = X ∗ (T ) is the group of rational characters of a maximal torus T .
We will denote by g the Lie algebra of G and other notations will be as in
Chap. 18.
Within each Cartan type there may be several Lie groups to consider,
but in each case there is a unique semisimple simply connected group. There
is also a unique simple semisimple group, which is isomorphic to the simply
connected group modulo its finite center. This is called the adjoint group since
it is isomorphic to its image in GL(g) under the adjoint representation. Here is
a table giving the simply connected and adjoint groups for each of the classical
Cartan types.
Let us consider first the Cartan type Ar . We will describe three distinct
groups, U(r + 1), SU(r + 1) and PU(r + 1), which is U (r + 1) modulo its
one-dimensional center. These have the same root system, but the ambient
vector space V is different in each case.
The group U(r + 1) is not semisimple. Its rank is r + 1 but its semisimple
rank is r. The maximal torus T consists of diagonal matrices t with eigen-
values t1 , . . . , tr+1 in T, the group of complex numbers of absolute value 1.
We may identify Λ = X ∗ (T ) ∼ / Z , in which λ = (λ1 , . . . , λr+1 ) with λi ∈ Z
= r+1
ei − ej , i = j, (19.1)
having exactly two nonzero entries, one being 1 and the other −1. To see
that this is the root system, we recall that the complexified Lie algebra is
C⊗g ∼ = glr+1 (C) = Matr+1 (C), since Matr+1 (C) = g ⊕ ig. (Every complex
matrix can be written uniquely as X + iY with X and Y skew-Hermitian.)
If α = ei − ej , the one-dimensional vector space Xα of Matr+1 (C) spanned
by the matrix Eij with a 1 in the i, j-position and 0’s everywhere else is an
eigenspace for T affording the character α, and these eigenspaces, together
with the Lie algebra of T , span V . So the ei − ej are precisely the roots of
U(r+1). The group U(r+1) has semisimple rank r, since that is the dimension
of the space spanned by these vectors.
Next consider the group SU(r + 1). This is the semisimple and simply
connected group with the same root system. The ambient space V is one
dimension
/ smaller than for U(r + 1), because the ti are subject to the equa-
tion ti = det(t) = 1. Therefore, the character represented by λ ∈ Zr+1 is
trivial if λ is in the diagonal lattice Δ = Z(1, . . . , 1). Thus, for this group, the
weight lattice, which we will denote ΛSU(r+1) , is Zr+1 /Δ, and the space V is
r-dimensional. It is spanned by the roots, so this group is semisimple.
The group PU(r + 1) is U(r + 1) modulo its one-dimensional central torus.
It is the adjoint group for the Cartan type Ar . It is isomorphic to SU(r + 1)
modulo its finite center of order r + 1. A character of U(r + 1) parametrized
by λ ∈ Zr+1 is well-defined
if and only if it is trivial on the center of U(r + 1),
which requires λi = 0. So the lattice ΛPU(r+1) is isomorphic to the sublat-
tice of Zr+1 determined by this condition. The composition
where the first map is the inclusion and the second the projection is injective,
so we may regard ΛPU(r+1) as a sublattice of ΛSU(r+1) . Its index is r + 1.
Turning now to the general case, the set Φ of roots will be partitioned into
two parts, called Φ+ and Φ− . Exactly half the roots will be in Φ+ and the other
half in Φ− . This is accomplished by choosing a hyperplane through the origin
in V that does not pass through any root, and taking Φ+ to be the roots on one
side of the hyperplane, Φ− the roots on the other side. Although the choice
−
of the hyperplane is arbitrary, if another such decomposition Φ = Φ+ 1 ∪ Φ1
is found by choosing a different hyperplane, a Weyl group element w can be
− −
found such that w(Φ+ ) = Φ+ 1 and w(Φ ) = Φ1 , so the procedure is not as
+
arbitrary as one might think. The roots in Φ will be called positive. In the
figures of this chapter, the positive roots are labeled •, and the negative roots
are labeled ◦.
If G is a semisimple compact connected Lie group, then its universal cover
G̃ is a cover of finite degree, and as in the last example (where G = PU(r + 1)
and G̃ = SU(r + 1)) the weight lattice of G is a sublattice of G̃. Moreover,
if π : G −→ GL(V ) is an irreducible representation, then we may compose
it with the canonical map G̃ −→ G and get a representation of G̃. So if we
understand the representation theory of G̃ we understand the representation
theory of G. For this reason, we will consider mainly the case where G is
simply connected and semisimple in the remaining examples of this chapter.
Assuming G is semisimple, so the αi span V, we will define certain special
elements of V as follows. If Σ = {α1 , . . . , αr } are the simple positive roots
then let {α∨ ∨
1 . . . , αr } be the corresponding coroots. In the semisimple case,
the coroots span V ∗ , and the fundamental dominant weights i are the dual
basis of V. Thus,
α∨
j (i ) = δij (Kronecker δ).
We will show later that if G is simply connected, then the i are in the weight
lattice Λ = X ∗ (T ), though if G is not simply connected, they may not all be
in Λ.
Another important particular vector is ρ, sometimes called the Weyl
vector. It may be characterized as half the sum of the positive roots, and
in the semisimple case it may also be characterized as the sum of the funda-
mental weights. (See Proposition 20.17.)
For example, the root system of type A2 , pictured in Fig. 19.1, consists of
α1 = (1, −1, 0), α2 = (0, 1, −1), (1, 0, −1),
(−1, 1, 0), (0, −1, 1), (−1, 0, 1).
With G = SU(3), we really mean the images of these vectors in Z3 /Δ, as
explained above. Taking T to be the diagonal torus of SU(3), α1 and α2 ∈
X ∗ (T ) are the roots
⎛ ⎞
t1
α1 (t) = t1 t−1
2 , a2 (t) = t2 t−1
3 , t = ⎝ t2 ⎠ ∈ T.
t3
148 19 Examples of Root Systems
α2 C+
ρ
1
α1
Figure 19.1 shows the root system of type A2 associated with the Lie group
SU(3). The shaded region in Fig. 19.1 is the positive Weyl chamber C+ , which
consists of {x ∈ V | x, α 0 for all α ∈ Φ+ }. It is a fundamental domain for
the Weyl group.
A role will also be played by a partial order on V. We define x y
if x − y 0, where x 0 if x is a linear combination, with nonnegative
coefficients, of the elements of Σ. The shaded region in Fig. 19.2 is the set of
x such that x 0 for the root system of type A2 .
Next we turn to the remaining classical root systems. The root system of
type Bn is associated with the odd orthogonal group SO(2n + 1) or with its
double cover spin(2n + 1). The root system of type Cn is associated with the
symplectic group Sp(2n). Finally, the root system of type Dn is associated
with the even orthogonal group SO(2n) or its double cover spin(2n). We will
now describe these root systems. Let ei = (0, . . . , 0, 1, 0, . . . , 0) be the standard
basis of Rn .
19 Examples of Root Systems 149
{x 0}
ρ
α2
2
α1
The root system of type Bn can be embedded in Rn . The roots are not all
of the same length. There are 2n short roots
±ei (1 i n)
±ei ± ej (i = j).
To see that this is the root system of SO(2n + 1), it is most convenient to use
the representation of SO(2n + 1) in Exercise 5.3. Thus, we replace the usual
realization of SO(2n + 1) as a group of real matrices by the subgroup of all
g ∈ U(2n + 1) that satisfy g J t g = J, where
⎛ ⎞
1
.
J = ⎝ .. ⎠ .
1
A maximal torus consists of all diagonal elements, which have the form (when
n = 4, for example)
⎛ ⎞
t1
⎜ t2 ⎟
⎜ ⎟
⎜ t ⎟
⎜ 3 ⎟
⎜ t ⎟
⎜ 4 ⎟
t=⎜⎜ 1 ⎟.
⎟
⎜ t −1 ⎟
⎜ 4 ⎟
⎜ t −1 ⎟
⎜ 3 ⎟
⎝ t−1 ⎠
2
−1
t1
150 19 Examples of Root Systems
Fig. 19.3. The Lie algebra so(9). The Dynkin diagram, which will be explained in
Chap. 25, has been superimposed on top of the Lie algebra
We order the roots so the root spaces Xα with α ∈ Φ+ are upper triangular.
In particular, the simple roots are α1 (t) = t1 t−1
2 , acting on Xα1 , the space of
matrices in which all entries are zero except x12 ; α2 (t) = t2 t−1
3 , with root space
corresponding to x23 ; α3 (t) = t3 t−1
4 corresponding to x34 ; and α4 (t) = t4 ,
corresponding to x45 . We have circled these positions. Note, however that (for
example) x12 appears in a second place which has not been circled. The lines
connecting the circles, one of them double, map out the Dynkin diagram,
which will explained in greater detail Chap. 25. Briefly, the Dynkin diagram
is a graph whose vertices correspond to the simple roots; simple roots are
connected in the Dynkin diagram if they are not perpendicular. We have
drawn the nodes corresponding to each simple root on top of the variable xij
for the corresponding eigenspace.
We have drawn a double bond with an arrow pointing from a long root
to a short root, which is the convention when two nonadjacent roots have
different lengths.
19 Examples of Root Systems 151
±2ei (1 i n)
±ei ± ej (i = j).
±ei ± ej (i = j).
To see that Dn is the root system of SO(2n), one may again use the realiza-
tion of Exercise 5.3. We leave this verification to the reader in Exercise 19.2.
(Fig. 30.1 may help with this.)
α2
2
α1
Fig. 19.4. The root system of type C2 , which coincides with type B2
Weyl chamber. (We have labeled the roots so that the order coincides with
the root system C2 in the notations of Bourbaki [23], in the appendix at the
back of the book. For type B2 , the roots α1 and α2 would be switched.)
There is a nonreduced root system whose type is called BCn . The root
system of type BCn can be realized as all elements of the form
α2
1
α1
α2
2 ρ
1 α1
Exercises
Exercise 19.1. Show that any irreducible rank 2 root system is isomorphic to one
of those described in this chapter, of type A2 , B2 , G2 or BC2 .
Exercise 19.2. Verify, as we did for type SO(2n + 1), that the root system of the
Lie group SO(2n) is of type Dn and that the root system of Sp(2n) is of type Cn .
154 19 Examples of Root Systems
Exercise 19.3. Show that the root systems of types B2 and C2 are isomorphic.
Exercise 19.4. Show that the root system of SO(6) is isomorphic to that of SU(4).
What can you say about the root system of SO(4)?
Exercise 19.5. Suppose that G is a compact Lie group with root system Φ, and
that H is a Lie subgroup of G having the same maximal torus. Show that every root
of H is a root of G, and that if Φ ⊆ Φ is the root system of H, then
If α, β ∈ Φ and α + β ∈ Φ then α + β ∈ Φ . (19.2)
Exercise 19.6. Conversely, let G be a compact Lie group with root system Φ. Let
Φ ⊆ Φ be a root system such that (19.2) is satisfied. Show that in the notation of
Chap. 18, tC and Xα (α ∈ Φ ) form a complex Lie algebra hC , and that h = hC is a
Lie subalgebra of g.
Exercise 19.7. Let Φ be the root system of type G2 .
(i) Show that the long roots form a root system Φ satisfying (19.2).
(ii) Assume the following fact: there exists a simply connected compact Lie group G
whose root system is Φ. This Lie group G2 may be constructed as the group of
automorphisms of the octonions (Jacobson [87]). Prove that there exists a non-
trivial homomorphism SU(3) → G (known to be injective). (Hint: Use Exercise
19.6 and Theorem 14.2.)
(ii) Exhibit another root system in Φ of rank two satisfying (19.2). Note that you
cannot use the short roots for this.
It may be shown that the root systems Φ in (i) and (ii) of the last exercise
correspond to Lie groups [su(3) for part (i)] that may be embedded in the exceptional
group G2 .
Exercise 19.8. Let ei (i = 1, 2, 3, 4) be the standard basis elements of R4 . Show
that the 48 vectors
1
±ei (1 i 4), ±ei ± ej (1 i < j 4), (±e1 ± e2 ± e3 ± e4 ) ,
2
form a root system. This is the root system of Cartan type F4 . Compute the order of
the Weyl group. Show that this root system contains smaller root systems of types
B3 and C3 .
Exercise 19.9. Let Φ8 consist of the following vectors in R8 . First, the 112 vectors
± e i ± ej 1 i < j 8. (19.3)
Second, the 128 vectors
1
(±e1 ± e2 ± e3 ± e4 ± e5 ± e6 ± e7 ± e8 ) (19.4)
2
where the number of − signs is even. We will refer to the vectors (19.3) as integral
roots and the vectors (19.4) as half-integral roots. Prove that Φ8 is a root system.
This is the exceptional root system of type E8 . Note that the integral roots form a
root system of type D8 .
Hint: To show that if α and β are roots then sα (β) ∈ Φ8 , observe that the D8 Weyl
group permutes
the roots, and using this action we may assume that α = e1 + e2
or α = 12 ei . The first case is easy so assume α = 12 ei . We may then use the
action of the symmetric group on β and there are only a few cases to check.
19 Examples of Root Systems 155
In this chapter, we will associate a Weyl group with an abstract root system,
and develop some of its properties.
Let V be a Euclidean space and Φ ⊂ V a reduced root system. (At the end
of the chapter we will remove the assumption that Φ is reduced, but many of
the results of this chapter are false without it.)
Since Φ is a finite set of nonzero vectors, we may choose ρ0 ∈ V such that
α, ρ0
= 0 for all α ∈ Φ. Let Φ+ be the set of roots α such that α, ρ0 > 0.
This consists of exactly half the roots since evidently a root α ∈ Φ+ if and
only if −α ∈ / Φ+ . Elements of Φ+ are called positive roots. Elements of set
Φ− = Φ − Φ+ are called negative roots.
If α, β ∈ Φ+ and α + β ∈ Φ, then evidently α + β ∈ Φ+ . Let Σ be the set
of elements in Φ+ that cannot be expressed as a sum of other elements of Φ+ .
If α ∈ Σ, then we call α a simple positive root, or sometimes just a simple
root and we call sα defined by (18.1) a simple reflection.
Proposition 20.1.
(i) The elements of Σ are linearly independent.
(ii) If α ∈ Σ and β ∈ Φ+ , then either β = α or sα (β) ∈ Φ+ .
(iii) If α and β are distinct elements of Σ, then α, β 0.
(iv) Each element α ∈ Φ can be expressed uniquely as a linear combination
α= nβ · β
β∈Σ
where nγ = nγ + nγ . There exists some γ ∈ Σ such that nγ
= 0, because β is
not α, and (the root system being reduced), it follows that β is not a multiple
of α. Therefore,
(α∨ (β) − nα ) α = nγ · γ,
γ ∈ Σ
γ = α
and the right-hand side is not zero. Taking the inner product with ρ0 shows
that the coefficient on the left-hand side is strictly positive; dividing by
this positive constant, we see that α may be expressed as a linear combi-
nation of the elements γ ∈ Σ distinct from α, and so α may be omitted
from Σ , contradicting its assumed minimality. This contradiction shows that
sα (β) ∈ Φ+ .
Next we show that if α and β are distinct elements of Σ , then α, β 0.
We have already shown that sα (β) ∈ Φ+ . If α, β > 0, then by (18.2) we have
α∨ (β) > 0. Write
β = sα (β) + α∨ (β) α. (20.1)
We have already shown that sα (β) ∈ Φ+ . Writing sα (β) as a linear combina-
tion with nonnegative coefficients of the elements of Σ , and noting that the
coefficient of α on the right-hand side of (20.1) is strictly positive, we may
write
β= nγ · γ,
γ∈Σ
At least one coefficient nα > 0 on the right, so taking the inner product with
ρ0 we see that 1 − nβ > 0. Thus, β is a linear combination with nonnegative
20 Abstract Weyl Groups 159
where Σ1 and Σ2 are disjoint subsets of Σ and the coefficients cα , dβ are all
positive. Call this vector v. We have
v, v = cα dβ α, β 0
α ∈ Σ1
β ∈ Σ2
Since α, ρ0 > 0 and cα > 0, this is a contradiction. This proves the linear
independence of the elements of Σ .
Next let us show that every element of Φ+ may be expressed as a linear
combination of elements of Σ with integer coefficients. We define a function
h from Φ+ to the positive real numbers as follows. If α ∈ Φ+ we may write
α= nβ · β, nβ 0.
β∈Σ
The coefficients nβ are uniquely determined since the elements of Σ are lin-
early independent. We define
h(α) = nβ . (20.3)
Evidently h(α) > 0. We want to show that the coefficients nβ are integers.
Assume a counterexample with h(α) minimal. Evidently, α ∈ / Σ since if
α ∈ Σ , then nα = 1 while all other nβ = 0, so such an α has all nβ ∈ Z.
Since
0 < α, α = nβ α, β , (20.4)
β∈Σ
where
nβ if β
= γ ,
nβ =
nγ − γ ∨ (α) if β = γ .
Since γ, α > 0, we have
h(α ) < h(α) ,
so by induction we have nβ ∈ Z. Since Φ is a root system, γ ∨ (α) ∈ Z, so
nβ ∈ Z for all β ∈ Σ . This is a contradiction.
Finally, let us show that Σ = Σ .
If α ∈ Σ, then by definition of Σ, α cannot be expressed as a linear
combination with integer coefficients of other elements of Φ+ . Hence α cannot
be omitted from Σ . Thus, Σ ⊂ Σ .
On the other hand, if α ∈ Σ , then we claim that α ∈ Σ. Otherwise, we
may write α = β + γ with β, γ ∈ Φ+ , and β and γ may both be written as
linear combinations of elements of Σ with positive integer coefficients, and
thus h(β), h(γ) 1, so h(α) = h(β) + h(γ) > 1. But evidently h(α) = 1 since
α ∈ Σ . This contradiction shows that Σ ⊂ Σ.
s1 s2 · · · sk = s1 s2 · · · ŝj · · · sk sα , (20.8)
where the “hat” on the right signifies the omission of the single element sj .
or
sj+1 · · · sk sα = sj sj+1 · · · sk .
This implies (20.8).
where the “hats” on the right signify omission of the elements si and sj .
Proof. Evidently there is a first j such that l (s1 s2 · · · sj ) < j, and [since
l (s1 ) = 1] we have j > 1. Then l (s1 s2 · · · sj−1 ) = j − 1, and by Propo-
sition 20.2 we have s1 s2 · · · sj−1 (αj ) ∈ Φ− . The existence of i satisfying
s1 · · · sj−1 = s1 · · · ŝi · · · sj−1 sj now follows from Proposition 20.3, which im-
plies (20.9).
w = s1 s2 · · · ŝi · · · ŝj · · · sk .
The element w0 is often called the long element of the Weyl group.
C+ = {x|{α, x} for α ∈ Φ+ },
the element w0 such that wC+ = −C+ sends positive roots to negative roots.
Thus, its length equals the number of positive roots, and is maximal.
sα (ρ) = ρ − α, α ∈ Σ. (20.11)
Proof. This follows since sα changes the sign of α and permutes the remaining
positive roots.
Now let us assume that Φ spans V. This will be true if G is semisimple. Let Λ̃
to be the set of vectors v such that α∨ (v) ∈ Z for α∨ ∈ Φ∨ . In the semisimple
case the α∨ span V ∗ , Λ̃ is a lattice. We have Λ̃ ⊇ Λ ⊇ Λroot , and all three
lattices span V, so [Λ̃ : Λroot ] < ∞. The α∨i are linearly independent, and in
the semisimple case they are a basis of V, so let i be the dual basis of V. In
other words, these vectors are defined by α∨ i (j ) = δij (Kronecker delta). The
i are called the fundamental dominant weights. Strictly speaking, because
in our usage only elements of Λ will be called weights, the i might not
be weights by our conventions. However, we will call them the fundamental
weights because this terminology is standard. Clearly the i span Λ̃ as a
Z-module.
Up until now we have assumed that Φ is a reduced root system, and much
of the foregoing is false without this assumption. In Chap. 18, and indeed
most of the book, the root systems are reduced, so this is enough for now. In
Chap. 29, however, we will encounter relative root systems, which may not be
reduced, so let us say a few words about them. If Φ ⊂ V is not reduced, then
we may still choose v0 and partition Φ into positive and negative roots. We
call a positive root simple if it cannot be expressed as a linear combination
(with nonnegative coefficients) of other positive roots.
Proposition 20.18. Let (Φ, V) be a root system that is not necessarily re-
duced. If α and λα ∈ Φ with λ > 0, then λ = 1, 2 or 12 . Partition Φ into
positive and negative roots, and let Σ be the set of simple roots. The elements
of Σ are linearly independent. Any positive root may be expressed as a linear
combination of elements of Σ with nonnegative integer coefficients.
Exercises
Exercise 20.1. Suppose that S is any subset of Φ such that if α ∈ Φ, then either
α ∈ S or −α ∈ S. Assume further more that if α, β ∈ S and if α + β ∈ Φ then
α + β ∈ S. Show that there exists w ∈ W such that w(S) ⊇ Φ+ . If either for every
α ∈ Φ either α ∈ S or −α ∈ W but never both, then w is unique.
If G is a compact connected Lie group, we will show in Chap. 22 that its ir-
reducible representations are parametrized uniquely by their highest weight
vectors. In this chapter, we will explain what this means and give some illus-
trative examples. This chapter is to some extent a continuation of the example
Chap. 19. As in that chapter, we will make many assertions that will only be
proved in later chapters, mostly Chap. 22.
We return to the figures in Chap. 19 (which the reader should review). Let
T be a maximal torus in G, with Λ = X ∗ (T ) embedded as a lattice in the
Euclidean space V = R ⊗ X ∗ (T ). Let Λroot ⊆ Λ be the lattice spanned by the
roots.
If G is semisimple, then Λroot spans V and has finite codimension in Λ. In
this case, the coroots also span V ∗ , so we may ask for the dual basis of V. These
are elements called i such that α∨ i (j ) = δij . These are the fundamental
dominant weights. They are not necessarily in Λ, however: they are in Λ if
G is simply connected as well as semisimple. We only will call elements of
V weights if they are in Λ, so if G is not connected, the term “fundamental
dominant weight” is a misnomer. But if G is semisimple and simply connected,
the i are uniquely defined and span the weight lattice Λ. The fundamental
dominant weights do not play a major role in the general theory but they give
a convenient parametrization of Λ when G is semisimple, since then every
element of Λ is of the form ni i with ni nonnegative integers. (This is true
even if G is not simply connected.) Since our examples will be semisimple, we
will make use of the fundamental dominant weights.
Our first example is G = SU(3). The lattices Λ and its sublattice Λroot
(of index 3) are marked in Fig. 21.1. The positive Weyl chamber C + is the
shaded cone. It is a fundamental group for the Weyl group W , acting by
simple reflections, which are the reflections in the two walls of C + . The weight
lattice Λ is marked with light dots and the root sublattice with darker ones. In
this case G is semisimple and simply connected, so the fundamental dominant
weights 1 and 2 are defined and span the weight lattice. The root lattice
is of codimension 3 in Λ.
α2
C+
1
α1
1
1 1
1 1
1
1 1 1 1
1 2 2 2 1
1 2 3 3 2 1
1 2 3 4 3 2 1
1 2 3 4 4 3 2 1
1 2 3 4 4 4 3 2 1
1 2 3 4 4 4 4 3 2 1
1 2 3 3 3 3 3 2 1
1 2 2 2 2 2 2 1
1 1 1 1 1 1 1
Fig. 21.3. The irreducible representation π31 +62 of SU(2). The shaded region is
the positive Weyl chamber, and λ = 31 + 6ω2 is circled
The highest weight can be any element of Λ∩C + . In fact, there is a bijection
between Λ ∩ C + and the isomorphism classes of irreducible representations of
G. Since there is a unique irreducible representation with the given highest
weight λ, we will denote it by πλ . For example, if λ = 31 + 62 , the weight
diagram of πλ is shown in Fig. 21.3. Note that λ = 31 + 62 is marked with
a circle.
From this we can see several features of the general situation. The set of
weights of πλ can be characterized as follows. First, if μ is a weight of πλ
then λ μ in the partial order. This puts μ in the translate by λ of the
172 21 Highest Weight Vectors
cone {μ 0}. This is the shaded region in Fig. 21.4. Moreover since the set
of weights is invariant under the Weyl group W , we can actually say that
λ w(μ) for all w ∈ W . In Fig. 21.3, this puts μ in the hexagonal region
that is the convex hull of the W -orbit W λ = {w(λ) | w ∈ W }. This region is
marked with dashed lines.
1 1 1 1
1 2 2 2 1
1 2 3 3 2 1
1 2 3 4 3 2 1
1 2 3 4 4 3 2 1
1 2 3 4 4 4 3 2 1
1 2 3 4 4 4 4 3 2 1
1 2 3 3 3 3 3 2 1
1 2 2 2 2 2 2 1
1 1 1 1 1 1 1
Fig. 21.4. With λ the highest weight vector (circled ) the shaded region is {μ|μ λ}
It will be noted that not every element of Λ inside the hexagon is a weight
of πλ . Indeed, if μ is a weight of Λ then λ − μ ∈ Λroot . In the particular
example of Fig. 21.3, λ is itself in Λroot , so the weights of πλ are elements of
the root lattice. What is true in general is that the weights of πλ are the μ
inside the convex hull of W λ such that λ − μ ∈ Λroot .
Next let G = Sp(4). The root system is of type C2 . This group is also
simply connected, so again the fundamental dominant weights 1 and 2
are in the weight lattice. The weight lattice and root lattice are illustrated in
Fig. 21.5.
As in Fig. 21.1, the weight lattice Λ is marked with light dots and the
root sublattice with darker ones. We have also marked the positive Weyl
chamber, which is a fundamental group for the Weyl group W , acting by
simple reflections.
The group Sp(4) admits a homomorphism Sp(4) −→ SO(5), so it has both
a four-dimensional and a five-dimensional irreducible representation. These
are π1 and π2 , respectively. Their root diagrams may be found in Fig. 21.6.
21 Highest Weight Vectors 173
α2 2
α1
Fig. 21.5. The root and weight lattices of the C2 root system
1
1 1
1 1 1
1 1
1
1 1 1
1 2 3 2 1
1 2 4 4 4 2 1
1 2 4 5 6 5 4 2 1
1 3 4 6 6 6 4 3 1
1 2 4 5 6 5 4 2 1
1 2 4 4 4 2 1
1 2 3 2 1
1 1 1
Exercises
Exercise 21.1. Consider the adjoint representation of SU(3) acting on the eight-
dimensional Lie algebra g of SU(3). (It may be shown to be irreducible.) Show that
the highest weight vector is 1 + 2 , and construct a weight diagram.
21 Highest Weight Vectors 175
Exercise 21.2. Construct a weight diagram for the adjoint representation of Sp(4)
or, equivalently, SO(5).
Exercise 21.4. Consider the tensor product of the contragredient of the standard
representation of SU(3), having highest weight vector 2 , with the adjoint represen-
tation, having highest weight vector 1 + 2 . We will see later in Exercise 22.4 that
this tensor product has three irreducible constituents. They are the contragredient
of the standard representation, the symmetric square of the standard representa-
tion, and another piece, which we will call π1 +22 . The first two pieces are known,
and the third can be obtained by subtracting the two others. Accepting for now
the validity of this decomposition, construct the weight diagram for the irreducible
representation π1 +22 .
The character formula of Weyl [174] is the gem of the representation theory
of compact Lie groups.
Let G be a compact connected Lie group and T a maximal torus. Let
Λ = X ∗ (T ), and let Λroot be the lattice spanned by the roots. Then Λ ⊇ Λroot .
The index [Λ : Λroot ] may be finite (e.g. if G = SU(n)) or infinite (e.g. if
G = U(n)). If the index is finite, then we say G is semisimple, and this
corresponds to the semisimple case in Chap. 20. Elements of Λ will be called
weights.
We have written the characters of T additively. Sometimes we want to
write them multiplicatively, however, so we introduce symbols eλ for λ ∈ V
subject to the rule eλ eμ = eλ+μ . More formally, let E(R) denote the free
R-module
on the set of symbols {eλ |λ ∈ Λ}. It consists of all formal sums
λ∈Λ nλ e with nλ ∈ R such that nλ = 0 for all but finitely many λ. It is a
λ
This makes sense because only finitely many nλ and only finitely many mμ
are nonzero. Of course, E(R) is just the group algebra over R of Λ. The Weyl
group acts on E(R), and we will denote by E(R)W the subring of W -invariant
elements. Usually, we are interested in the case R = Z, and we will denote
E = E(Z), E W = E(Z)W . We will find it sometimes convenient to work in the
larger ringE2 , which is the free Abelian group on 12 Λ.
If ξ = λ nλ · eλ , we will sometimes
denote m(ξ, λ) = nλ , the multiplicity
of λ in ξ. We will denote by ξ = λ nλ · e−λ the conjugate of ξ.
By Theorem 17.1, class functions on G are the same thing as W -invariant
functions on T . In particular, if χ is the character of a representation of G,
then its restriction to T is a sum of characters of T and is invariant under
the action of W . Thus, if λ ∈ Λ, let nλ (χ) denote the multiplicity of λ in this
restriction. We associate with χ the element
The equivalence of the two expressions follows easily from the fact that 2ρ is
the sum of the positive roots.
Proposition 22.1. We have w(Δ) = (−1)l(w) Δ for all w ∈ W .
Proof. It is sufficient to check that sβ (Δ) = −Δ for every simple root β. We
recall that sβ changes the sign of β and permutes the remaining simple roots.
Of the factors in the first expression for Δ in (22.2), only two are changed:
e−ρ and (eβ − 1). These become [see (20.11)] e−ρ+β and (e−β − 1). The net
effect is that Δ changes sign.
An alternative way of explaining the same proof begins with the equation
7
Δ= (eα/2 − e−α/2 ). (22.3)
α∈Φ+
Here α/2 may not be an element of Λ, so each individual factor on the right
is not really an element of E but of the larger ring E2 . Proposition 22.1 follows
by noting that by Proposition 20.1(ii) each simple reflection alters the sign of
exactly one term in (22.3), and the result follows.
Proposition 22.2. If ξ ∈ E satisfies w(ξ) = (−1)l(w) ξ for all w ∈ W , then ξ
is divisible by Δ in E.
Proof. In the ring E, by Proposition 22.1, Δ is a product of distinct irreducible
elements 1 − eα , where α runs through Φ+ , times a unit e−ρ . It is therefore
by each 1 − e . By Proposition 20.12,
α
sufficient to show that ξ is divisible
we have sα (ξ) = −ξ. Write ξ = λ∈Λ nλ · l. Since sα (ξ) = −ξ, we have
nsα (λ) = −nλ . Noting that sα (λ) = λ − kα where k = α∨ (λ) ∈ Z, we see that
22 The Weyl Character Formula 179
ξ= nλ (eλ − eλ−kα ).
λ∈Λ
λ mod
sα
The notation means that we choose only one representative for each sα orbit
of Λ. (If sα (λ) = λ, then nλ = 0.) Since
this is divisible by Δ.
If λ ∈ Λ ∩ C+ , let
χ(λ) = Δ−1 (−1)l(w) ew(λ+ρ) . (22.4)
w∈W
By
Proposition 22.2, χ(λ) ∈ E. Moreover, applying w ∈ W multiplies both
w∈W (−1)w w(λ+ρ)
e and Δ by (−1)w , so χ(λ) is actually in E W .
We will eventually prove that if λ ∈ Λ ∩ C+ this is an irreducible character
of G. Then (22.4) is called the Weyl character formula.
If ξ = nλ eλ ∈ E, we define the support of ξ to be the finite set supp(ξ) =
{λ ∈ L | nλ
= 0}. We define a partial order on V by λ μ if λ = μ+ α∈Σ cα α,
where cα 0.
◦
Proposition 22.3. If λ ∈ C+ , then λ w(λ) for w ∈ W . If λ ∈ C+ and
w
= 1, then w(λ) λ.
Proposition
22.4. Let λ ∈ C+ . Then λ ∈ supp χ(λ). Indeed, writing χ(λ) =
n
μ μ · e μ
, we have nλ = 1. Moreover, if μ ∈ supp χ(λ), then λ μ, and
λ − μ ∈ Λroot . In particular, λ is the largest weight in the support of χ(λ).
that λ v. This means that, in the product (22.1), only finitely many terms
will be nonzero, so Ê is a ring. We can write
7
Δ = eρ (1 − e−α ),
α∈Φ+
so in Ê we have
7
Δ−1 = e−ρ (1 + e−α + e−2α + · · · ).
α∈Φ+
Therefore,
7
χ(λ) = eλ (1 + e−α + e−2α + · · · ) (−1)l(w) ew(λ+ρ)−(λ+ρ) . (22.5)
α∈Φ+ w∈W
Each factor in the product is ≺ 0 except 1, and by Proposition 22.3 each term
in the sum is ≺ 0 except that corresponding to w = 1. Hence, each term in
the expansion is λ, and exactly one term contributes λ itself.
It remains to be seen that if eμ appears in the expansion of the right-hand
side of (22.5), then λ−μ is an element of Λroot . We note that w(λ+ρ)−(λ+ρ) ∈
Λroot by Proposition 20.16, and of course all the contributions coming from
the product over α ∈ Φ+ are roots, and the result follows.
Now let us write the Weyl integration formula in terms of Δ.
Theorem 22.1. If f is a class function on G, we have
1
f (g) dg = f (t) |Δ(t)|2 dt. (22.6)
G |W | T
Here there is an abuse of notation since Δ is itself only an element of E, not
even W -invariant, so it is not identifiable as a function on the group. However,
it will follow from the proof that ΔΔ̄ is always a function on the group, and
we will naturally denote ΔΔ̄ as |Δ|2 .
Proof. We will show that, in the notation of Theorem 17.2,
det [Ad(t−1 ) − Ip ] | p = ΔΔ̄. (22.7)
Indeed, since the complexification of p is the direct sum of the spaces Xα on
each of which t ∈ T acts by α(t) in the adjoint representation,
7
7
det [Ad(t−1 ) − Ip ] | p = α(t)−1 − 1 = | α(t) − 1 |2 .
α∈Φ α∈Φ+
α∈Φ+ α∈Φ+
Proof. The L2 inner product of ξ and η is just the integral of ξ·η over the group
and, using (22.6), this is just W −1 times the multiplicity of 0 in (ξΔ)(ηΔ).
Proof. The linear independence of the χ(λ) follows from their orthogonality.
We must show that they span. Clearly, E W is spanned by elements of the form
B(λ) = eμ , l ∈ Λ ∩ C+ ,
μ∈W ·λ
We have
1 = χ, χ = n2λ .
λ
Therefore, exactly one nλ is nonzero, and that has value ±1. Thus, either χ(λ)
or its negative is an irreducible character of G. To see that −χ(λ) is not a
character, consider its restriction to T . By Proposition 22.4, the multiplicity
of the character λ in −χ(λ) is −1, which is impossible if −χ(λ) is a character.
Hence, χ(λ) = χ is an irreducible character of G.
We have shown that every irreducible character of G is a χ(λ). It remains
to be shown that every χ(λ) is a character. Since the class functions on G are
identical to the W -invariant functions on T , the closure in L2 (G) of E(C)W is
identified with the space of all class functions on G. By Proposition 22.6, the
χ(λ) form an L2 -basis of E(C)W . Since by the Peter–Weyl theorem the set of
irreducible characters of G are an L2 basis of the space of class functions, the
characters of G cannot be a proper subset of the set of χ(λ).
Now let us step back and see what we have established. We know that in group
representation theory there is a duality between the irreducible characters of
a group and its conjugacy classes. We can study both the conjugacy classes
and the irreducible representations of a compact Lie group by restricting them
to T . We find that the conjugacy classes of G are in one-to-one correspon-
dence with the W -orbits of T . Dually, the irreducible representations of G are
parametrized by the orbits of W on Λ = X ∗ (T ).
We study these orbits by embedding X ∗ (T ) in a Euclidean space V. The
positive Weyl chamber C+ is a fundamental domain for the action of W on V,
and so the dominant weights—those in C+ —are thus used to parametrize the
irreducible representations. Of the weights that appear in the parametrized
representation χ(λ), the parametrizing weight λ ∈ C+ ∩ X ∗ (T ) is maximal
with respect to the partial order. We therefore call it the highest weight vector
of the representation.
Proposition 22.7. We have
Δ= (−1)l(w) ew(ρ) . (22.9)
w∈W
22 The Weyl Character Formula 183
Weyl gave a formula for the dimension of the irreducible representation with
character χλ . Of course, this is the value χλ at the identity element of G,
but we cannot simply plug the identity into the Weyl character formula since
the numerator and denominator both vanish there. Naturally, the solution is
to use L’Hôpital’s rule, which can be formulated purely algebraically in this
context.
But by Proposition 20.1(ii), the set of w(α) consists of Φ+ with just one
element, namely β, replaced by its negative. So (22.11) is proved.
184 22 The Weyl Character Formula
where 8 9
7 > ?
−α/2
θ =Ω◦∂ e α/2
−e
α∈Φ+
Suppose that for each ν with m(ν)
= 0 the weight λ + ν is dominant. Then
χλ χμ = m(ν)χλ+ν . (22.14)
ν
Since χλ χμ is the character of the tensor product representation, this gives the
decomposition of this tensor product into irreducibles. The method of proof
can be extended to the case where λ + ν is not dominant for all ν, though the
answer is a bit more complicated to state (Exercise 22.5).
Interchange the order of summation, so that the sum over ν is the inner sum,
and make the variable change ν −→ w(ν). Since m(ν) = m(wν), we get
Δ−1 m(ν) (−1)l(w) ew(λ+ν+ρ) .
w ν
Now we may interchange the order of summation again and apply the Weyl
character formula to obtain (22.14).
δ = (n − 1, n − 2, . . . , 1, 0) ∈ X ∗ (T ). (22.16)
where
Δ0 = (−1)l(w) ew(δ) .
w∈W
We have simply multiplied the numerator and the denominator by the same
W -invariant element so that both the numerator and the denominator are in
X ∗ (T ).
In (22.7), we write the factor |Δ|2 = |Δ0 |2 since (Δ0 /Δ)2 = e2(δ−ρ) . As a
function on the group, this is just det(g)n−1 , which has absolute value 1.
Therefore, we may write the Weyl integration formula in the form
1
f (g) dg = f (t) |Δ0 (t)|2 dt. (22.18)
G |W | T
Exercises
In the first batch of exercises, G = SU(3) and, as usual, 1 and 2 are the funda-
mental dominant weights.
Exercise 22.1. By Proposition 22.4, all the weights in χλ lie in the set
Confirm by examining the weights that this is true for all the examples in Chap. 21—
in fact, for all these examples, S(μ) is exactly the set of weights.
Exercise 22.2. Use the Weyl dimension formula to compute the dimension of χ21 .
Deduce from this that the symmetric square of the standard representation is irre-
ducible.
Exercise 22.3. Use the Weyl dimension formula to compute the dimension of
χ1 +22 . Deduce from this that the symmetric square of the standard represen-
tation is irreducible.
Exercise 22.4. Use the Brauer–Klimyk method (Proposition 22.9) to compute the
tensor product of the contragredient of the standard representation (with character
χ2 ) and the adjoint representation (with character χ1 +2 ).
22 The Weyl Character Formula 187
Exercise 22.5. Prove the following extension of Proposition 22.9. Suppose that λ
is dominant and that ν is any weight. By Proposition 20.1, there exists a Weyl group
element such that w(ν + λ + ρ) ∈ C+ . The point w(ν + λ + ρ) is uniquely determined,
even though w may not be. If w(ν+λ+ρ) is on the boundary of C+ , define ξ(ν, λ) = 0.
If w(ν + λ + ρ) is not on the boundary of C+ , explain why w(ν + λ + ρ) − ρ ∈ C+ and
w is uniquely determined. In this case, define
ξ(ν, λ) = (−1)l(w) χw(ν+λ+ρ)−ρ . Prove
ν
that if μ is a dominant weight, and χμ = m(ν)e , then
χμ χλ = m(ν)ξ(ν, λ).
ν
Exercise 22.6. Use the last exercise to compute the decomposition of χ21 into
irreducibles, and obtain another proof that the symmetric square of the standard
representation is irreducible.
The Kostant partition function P(μ) is defined to be the number of vector partitions
of μ. Note that this is zero unless μ 0. Let Ê be the completion of E defined in
the proof of Proposition 22.4. Show that in Ê
(1 − e−α )−1 = P(μ)e−μ ,
α∈Φ μ ∈ Λroot
μ0
and from (22.5) deduce the Kostant multiplicity formula, for λ a dominant weight:
the multiplicity of μ in χ(λ) is
(−1)l(w) P w(λ + ρ) − ρ − μ .
w μ
Exercise 22.9. Show that if −w0 λ = λ for all weights in G, then every element of
G is conjugate to its inverse.
Exercise 22.12. Generalize the last exercise as follows. Let V be the adjoint rep-
resentation of Sp(2n). Its degree is n + 2n2 . What is the decomposition of the
symmetric and exterior squares of this representation into irreducibles? You might
want to use a computer program such as Sage to collect some data but prove your
answer.
Exercise 22.13. Let the weight lattices of Sp(2n) and SO(2n + 1) be identified
with Zn as in Chap. 19. Denote by ρC and ρB the Weyl vectors of these two groups.
Show that
1 3 1
ρC = (n, n − 1, . . . , 1), ρB = n − , n − , . . . , .
2 2 2
Recall that Spin(2n + 1) is the double cover of SO(2n + 1). The root lattice of
Spin(2n + 1) is naturally embedded in that of SO(2n + 1). Show that the root lattice
of Spin(2n+1) consists of tuples (μ1 , . . . , μn ) such that 2μi ∈ Z, and the μi are either
all integers or all half integers, that is, the 2μi are either all even or odd. (Hint:
Use Proposition 18.10 and look ahead to Proposition 31.2 if you need help.) Now
let λ = (λ1 , . . . , λn ) ∈ Zn such that λ1 · · · λn , and let μ = (μ1 , . . . , μn ) where
μi = λi + 12 . Show that λ and μ are dominant weights for Sp(2n) and SO(2n+1). Let
Vλ and Wμ be irreducible representations of Sp(2n) and SO(2n + 1), respectively.
Show that
dim(Wμ ) = 2n · dim(Vλ ).
[Hint: It may be easiest to show that dim(Wμ )/ dim(Vλ ) is constant, then take l = 0
to determine the constant.]
The generalized χ(g 2 ) is (in the language of lambda rings) the second Adams
operation applied to χ. More generally, for k 0, the Adams operation ψ k χ(g) =
χ(g k ). We will return to the Adams operations in Chap. 33, and in particular we
will see that ψ k χ is a generalized character for all k; for k = 2 this follows from
Exercise 22.14. Suppose that for μ in the weight lattice Λ of G, m(μ) is the weight
multiplicity for χ, so that in the notation of Chap. 22, we have
χ= m(μ) eμ .
μ∈Λ
Then clearly
ψk χ = m(μ) ekμ .
μ∈Λ
1 r
k
∨k χ = (ψ χ)(∨k χ),
k r=1
1
k
∧k χ = (−1)r (ψ r χ)(∧k χ).
k r=1
Hint: The symmetric polynomial identity (37.3) below may be of use.
Exercise 22.19. Let 1 and 2 be the fundamental dominant weights for SU(3).
Show that the irreducible representation with highest weight k1 + l2 is self-
contragredient if and only if k = l, and in this case, it is orthogonal.
Exercise 22.20. Let G be a semisimple Lie group. Show that the adjoint repre-
sentation is orthogonal. (Hint: You may assume that G is simple. Use the Killing
form.)
23
The Fundamental Group
Proposition 23.4.
(i) The map φ is a covering map of degree |W |.
(ii) If t ∈ Treg , then the |W | elements wtw−1 , w ∈ W are all distinct.
Proof. For t ∈ Treg , the Jacobian of this map, computed in (17.2), is nonzero.
Thus the map φ is a local homeomorphism.
We define an action of W = N (T )/T on G/T × Treg by
W acts freely on G/T , so the quotient map G/T × Treg −→ W \(G/T × Treg )
is a covering map of degree |W |. The map φ factors through W \(G/T ×
Treg ). Consider the induced map ψ : W \(G/T × Treg ) −→ Greg . We have a
commutative diagram:
23 The Fundamental Group 193
ψ
φ
Greg
Proof. Suppose that p0 and p1 are loops in X with the same endpoints whose
images in Y are path-homotopic. It is an immediate consequence of Proposi-
tion 13.2 that p0 and p1 are themselves path-homotopic.
In Exercise 27.4 we will see that the Bruhat decomposition gives an alternative
proof of this fact.
Proof. Let t0 ∈ Treg and consider the map f0 : G/T −→ G, f0 (gT ) = gt0 g −1 .
We will show that the map π1 (G/T ) −→ π1 (G) induced by f0 is injective. We
may factor f0 as
υ φ
G/T −→ G/T × Treg −→ Greg −→ G,
194 23 The Fundamental Group
where the first map υ sends gT −→ (gT, t0 ). We will show that each
induced map
υ φ
π1 (G/T ) −→ π1 (G/T × Treg ) −→ π1 (Greg ) −→ π1 (G) (23.1)
is injective. It should be noted that Treg might not be connected, so G/T ×Treg
might not be connected, and π1 (G/T × Treg ) depends on the choice of a
connected component for its base point. We choose the base point to be (T, t0 ).
υ
We can factor the identity map G/T as G/T −→ G/T × Treg −→ G/T ,
where the second map is the projection. Applying the functor π1 , we see that
π1 (υ) has a left inverse and is therefore injective. Also π1 (φ) is injective by
Propositions 23.4 and 23.5, and the third map is injective by Proposition 23.6.
This proves that the map induced by f0 injects π1 (G/T ) −→ π1 (G).
However, the map f0 : G/T → G is homotopic in G to the trivial map
mapping G/t to a single point, as we can see by moving t0 to 1 ∈ G. Thus f0
induces the trivial map π1 (G/T ) −→ π1 (G) and so π1 (G/T ) = 1.
Proof. One way to see this is to use have the exact sequence
π1 (T ) −→ π1 (G) −→ π1 (G/T )
Proof. Since t → λ etH is a character of R we have λ etH = e2πiθt for some
1 d
θ = θ(λ, H). Then θ = 2πi dt λ(e
tH
) |t=0 = 2πi
1
dλ(H). On the other hand
1
d 1 2πiθ
λ(eH ) = λ(etH ) dt = e −1 ,
0 dt 2πiθ
α∨ = τ (2πiHα ). (23.2)
α2∨
0 α1∨
α0∨
Fig. 23.1. The Cartan subalgebra t, partitioned into alcoves, when G = SU(3) or
PU(3). We are identifying t = V ∗ via the isomorphism τ , so the coroots are in t
Proposition 23.9.
(i) The affine Weyl group acts transitively on the alcoves.
(ii) The group Waff is generated by s0 , s1 , . . . , sr .
(iii) The group Waff contains the group of translations by elements of Λ∨
coroot as
a normal subgroup and is the semidirect product of W and the translation
group of Λ∨
coroot .
Proof. Let Waff be the subgroup s0 , s1 , . . . , sr . Consider the orbit Waff F of
alcoves. If F1 is in this orbit, and F2 is another alcove adjacent to F1 , then
F = wF1 for some w ∈ WF and so wF2 is adjacent to F , i.e. wF2 = si F for
some si . It follows that F2 is in Waff F also. Since every alcove adjacent to
an alcove in Waff F is also in it, it is clear that Waff F consists of all alcoves,
proving (i).
We may now show that Waff = Waff . It is enough to show that the reflection
r = sα,k in the hyperplane Hα,k is in Waff . Let A be an alcove that is adjacent
to Hα,k , and find w ∈ Waff such that w(A) = F . Then w maps Hα,k to one of
the walls of F , so wrw−1 is the reflection in that wall. This means wrw−1 = si
for some i and so r = w−1 si w ∈ Waff
. This completes the proof that Waff is
generated by the si .
Now we recall that the group G of all affine linear maps of the real vector
space t is the semidirect product of the group of linear transformations (which
fix the origin) by the group of translations, a normal subgroup. If v ∈ t we
will denote by T (v) the translation x → x + v. So T (t) is normal in G and
G = GL(t) · T (t). This means that G/T (t) ∼ = GL(t). The homomorphism
G −→ GL(t) maps the reflection in Hα,k to the reflection in the parallel
hyperplane Hα,0 through the origin. This induces a homomorphism from Waff
to W , and we wish to determine the kernel K, which is the group of translations
in Waff .
First observe T (α∨ ) is in Waff , since it is the reflection in Hα,0 followed by
the reflection in Hα,1 . Let us check that Λ∨ coroot is normalized by W . Indeed we
check easily that sα,0 T (β ∨ )sα0 is translation by sα,0 (β ∨ ) = β ∨ − α(β ∨ )α∨ ∈
Λ∨ ∨
coroot . Therefore W · T (Λcoroot ) is a subgroup of Waff . Finally, we note that
∨ ∨
sα,k = T (kα )sα,0 , so W · T (Λcoroot) contains generators for Waff , and the
proof is complete.
The group Wextended generated by Waff and translations by Λ may equal
Waff or it may be larger. This often larger group is called the extended affine
Weyl group. We will not need it, but it is important and we mention it for
completeness.
We have constructed a surjective homomorphism Λ∨ −→ π1 (G). To re-
capitulate, the exponential map t −→ T is a covering by a simply-connected
group, so its kernel Λ∨ may be identified with the fundamental group π1 (T ),
and we have seen that the map π1 (T ) −→ π1 (G) induced by inclusion is
surjective.
We recall that if X is a topological space then a path p : [0, 1] −→ X is
called a loop if p(0) = p(1).
198 23 The Fundamental Group
Proof. First let us show that a coroot α∨ is in the kernel of this homomor-
phism. In view of (23.2) this means, concretely that if we take a path from
0 to 2πiHα in t then the exponential of this map (which has both left and
right endpoint at the identity) is contractible in G. We may modify the path
so that it is the straight-line path, passing through πiHα . Then we write it
as the concatenation of two paths, p q in the notation of Chap. 13, where
p(0) = 0 and p(1) = πiHα while q(0) = πiHα and q(1) = 2πiHα .
The exponential of this path epq = ep eq is a loop. In fact, we have
p(0) q(1) 1 p(1) q(0) −1
e =e = iα , e =e = iα .
1 −1
show that we may assume that K lies in the closure of one of these alcoves.
Indeed F + K is an alcove, so there is some w ∈ Waff such that F + K = w F .
Moreover, w may be represented as T (K )w with K ∈ Λ∨ coroot and w ∈ W .
Thus F + K − K = wF and since we have already shown that K maps to
the identity in π1 (G), we may replace K by K − K and assume that K is in
the closure of wF .
Our goal is to prove that K = 0, which will of course establish that estab-
lish K ∈ Λ∨ coroot , as required. Since K and the origin are in the same alcove,
we may find a path p : [0, 1] −→ t from 0 to K such that p(u) is in the interior
of the alcove for p(u)
= 0, 1, while p(0) = 0 and p(1) = K are vertices of
the alcove.
Let treg be the preimage of Treg in t under the exponential map. It is the
complement of the hyperplanes Hα,k , or equivalently, the union of the alcove
interiors. We will make use of the map ψ : G/T × treg −→ Greg defined by
φ(gT, H) = geH g −1 . It is the composition of the covering map φ in Proposi-
tion 23.4 with the exponential map treg −→ Treg , which is also a covering, so
ψ is a covering map.
Let N be a connected neighborhood of the identity in G that is closed
under conjugation such that the preimage under exp of N in t consists of
disjoint open sets, each containing an a single element of the kernel Λ∨ of the
exponential map on t. Let Nt = t ∩ exp−1 (N ) be this reimage. Each connected
component of Nt contains a unique element of Λ∨ .
We will modify p so that it is outside treg only near t = 1. Let H ∈ t be
a small vector in the interior of wF ∩ N and let p : [0, 1] −→ t be the path
shifted by H. The vector H can be chosen in the connected component of Nt
that contains 0 but no other element of Λ∨ . So p (0) = H and p (1) = K + H.
When t is near 1 the path p may cross some of the Hα,k but only inside N .
The exponentiated path ep (t) will be a loop close to ep(t) hence con-
p (t)
tractible. And e will be near the identity 1 = eK in G at the end of
the path where p (t) may not be in the interior of wF . Because, by Proposi-
tion 23.3, Greg is a union of codimension 3 submanifolds of G, we may find
a loop q : [0, 1] −→ Greg that coincides with ep until near the end where
p (t) is near v + H and ep (t) reenters N . At the end of the path, q dodges
out of T to avoid the singular subset of G, but stays near the identity. More
precisely, if q is close enough to ep the paths will agree until ep (t) and q
are both inside N .
The loop q will have the same endpoints as ep . It is still contractible by
Proposition 23.6. Therefore we may use the path lifting property of covering
maps [Proposition 13.2(i)] to lift the path q to a path p : [0, 1] −→ G/T ×treg
by means of ψ, with p (0) = (1·T, H), the lifted path is a loop by Lemma 23.3.
Thus p (1) = p (0) = (1 · T, H). Now consider the projection p of p onto
treg . At the end of the path, the paths ep and ψ ◦ p are both within N , so
p and p can only vary within Nt . In particular their endpoints, which are
H + K and H respectively, must be the same, so K = 0 as required.
200 23 The Fundamental Group
Regarding these as lattices in the dual real vector spaces V and V ∗ , which we
have identified with R ⊗ X ∗ (T ) and with t, respectively, Λ̃ is the dual lattice
of Λ∨ ∨ ∨
coroot , Λ is the dual lattice of Λ and Λroot is the dual lattice of Λ̃ . Both
π1 (G) and Z(G) are finite Abelian groups and
π1 (G) ∼
= Λ̃/Λ ∼
= Λ∨ /Λ∨
coroot , = Λ̃∨ /Λ∨ ∼
Z(G) ∼ = Λ/Λroot. (23.3)
Proof. By Proposition 18.10 we have Λ̃ ⊇ Λ and Λ ⊇ Λroot is clear since roots
are characters of X ∗ (T ). That Λ and Λ∨ are dual lattices is Lemma 23.2. That
Λroot and Λ̃∨ are dual lattices and that Λ̃ and Λ∨ coroot are dual lattices are both
by definition. The inclusions Λ̃ ⊇ Λ ⊇ Λroot then imply Λ̃∨ ⊇ Λ∨ ⊇ Λ∨ root .
Moreover, Λ̃/Λ ∼ = Λ∨ /Λ∨ root follows from the fact that two Abelian groups in
duality are isomorphic, and the vector space pairing V × V ∗ −→ R induces a
∨ ∼
perfect pairing Λ̃/Λ × Λ∨ /Λ∨ ∨
root −→ R/Z. Similarly Λ̃ /Λ = Λ/Λroot . The
fact that π1 (G) ∼ ∨ ∨
= Λ /Λcoroot follows from Proposition 23.10. It remains to
be shown that Z(G) ∼ = Λ̃∨ /Λ∨ . We know from Proposition 18.14 that Z(G) is
the intersection of the Tα . Thus H ∈ t exponentiates into Z(G) if it is in the
kernel of all the root groups, that is, if α(H) ∈ Z for all α ∈ Φ. This means
that the exponential induces a surjection Λ̃∨ −→ Z(G). Since Λ∨ is the kernel
of the exponential on t, the statement follows.
23 The Fundamental Group 201
Exercises
Exercise 23.1. If g is a Lie algebra let [g, g] be the vector space spanned by [X, Y ]
with X, Y ∈ g. Show that [g, g] is an ideal of g.
Exercise 23.2. Suppose that g is a real or complex Lie algebra. Assume that there
exists an invariant inner product B : g × g −→ C. Thus B is positive definite
symmetric or Hermitian and satisfies the ad-invariance property (10.3). Let z be the
center of g. Show that the orthogonal complement of g is [g, g].
Exercise 23.3. Let G be a semisimple group of adjoint type, and let G be its
universal cover. Show that the fundamental group of G is isomorphic to the center
of G . (Both are finite Abelian groups.)
Exercise 23.4.
(i) Consider a simply-connected semisimple group G. Explain why Λ = Λ in the
notation of Theorem 23.2.
(ii) Using the description of the root lattices for each of the four classical Cartan
types in Chap. 19, consider a simply-connected semisimple group G and compute
the weight lattice Λ using Theorem 23.2. Confirm the following table:
Exercise 23.5. If g is a Lie algebra, the center of g is the set of all Z ∈ g such that
[Z, X] = 0 for all X ∈ G. Show that if G is a connected Lie group with Lie algebra
g then the center of g is the Lie algebra of Z(G).
Exercise 23.6. (i) Let g be the Lie algebra of a compact Lie group G. If a is an
Abelian ideal, show that a is contained in the center of g.
(ii) Show by example that this may fail without the assumption that g is the Lie
algebra of a compact Lie group. Thus give a Lie algebra and an Abelian ideal
that is not central.
Exercise 23.7. Let G be a compact Lie group and g its Lie algebra. Let T , t,
and other notations be as in this chapter. Let t be the linear span of the coroots
α∨ = 2πiHα . Let z be the center of g. Show that t = t z.
Part III
Proof. Given any complex analytic group H and any Lie group homomorphism
f : SL(n, R) −→ H, the differential is a Lie algebra homomorphism sl(n, R)−→
Lie(H). Since Lie(H) is a complex Lie algebra, this homomorphism extends
uniquely to a complex Lie algebra homomorphism sl(n, C) −→ Lie(H) by
Proposition 11.3. By Theorems 13.5 and 13.6, SL(n, C) is simply connected,
so by Theorem 14.2 this map is the differential of a Lie group homomorphism
F : SL(n, C) −→ H. We need to show that F is analytic. Consider the com-
mutative diagram
The top, left, and right arrows are all holomorphic maps, and exp : sl(n, C) −→
SL(n, C) is a local homeomorphism in a neighborhood of the identity. Hence
F is holomorphic
near 1. If g ∈ SL(n, C) and if l(g) : SL(n, C) −→ SL(n, C)
and l F (g) : H −→
H denote left translation with
respect to g and F (g),
then l(g) and l F (g) are analytic, and F = l F (g) ◦ F ◦ l(g)−1 . Since F is
analytic at 1, it follows that it is analytic at g.
We recall from Chap. 14, particularly the proof of Proposition 14.1, that if G is
a Lie group and h a Lie subalgebra of Lie(G), then there is an involutory family
of tangent vectors spanned by the left-invariant vector fields corresponding to
the elements of h. Since these vector fields are left-invariant, this involutory
family is invariant under left translation.
One must not conclude from this that every Lie subalgebra of Lie(G) is the
Lie algebra of a closed Lie subgroup. For example, if G = (R/Z)2 , then the
one-dimensional subalgebra spanned by a vector (x1 , x2 ) ∈ Lie(G) = R2 is
the Lie algebra of a closed subgroup only if x1 /x2 is rational or x2 = 0.
G = {(x, y) | x2 + y 2 = ±1},
which is an algebraic group with group law (x, y)(z, w) = (xz − yw, xw + yz),
but which has one Zariski-connected component with no real points.
We see that the algebraic complexification is a functor not from the
category of Lie groups but rather from the category of algebraic groups G
defined over R. So the algebraic complexification of a Lie group K depends
on more than just the isomorphism class of K as a Lie group—it also depends
on its realization as the group of real points of an algebraic group. We illustrate
this point with an example.
24 Complexification 209
Proof. This is clear since a polynomial function extends uniquely from G(R)
to G(C).
Compare this with Proposition 11.4, which is the Lie algebra analog of this
statement.
A = 12 (g + t g −1 ), B= 1
2i (g − t g −1 ). (24.1)
It can be seen similarly that SU(n) and SL(n, R) are C/R Galois forms of each
other. One has only to impose in the definition of the second group G2 an
additional polynomial relation corresponding to the condition det(A+Bi) = 1.
(This condition, written out in terms of matrix entries, will not involve i, so
the resulting algebraic group is defined over R.)
Galois forms are important because if G1 and G2 are Galois forms of each
other, then we expect the representation theories of G1 and G2 to be related.
We have already seen this principle applied (for example) in Theorem 14.3.
Our next proposition gives a typical application.
Exercises
Exercise 24.1. If F is a field, let
⎛ ⎞
1
SOJ (n, F ) = g ∈ SL(n, F ) | g J t g = J , J =⎝ . .. ⎠.
1
Show that SOJ (C) is the complexification of SO(n). (Use Exercise 5.3.)
25
Coxeter Groups
As we will see in this chapter, Weyl groups and affine Weyl groups are
examples of Coxeter groups, an important family of groups generated by
“reflections.”
Let G be a group, and let I be a set of generators of G, each of which has
order 2. In practice, we will usually denote the elements of I by {s1 , s2 , . . . , sr }
or {s0 , . . . , sr } with some definite indexing by integers. If si , sj ∈ I, let
n(i, j) = n(si , sj ) be the order of si sj . [Strictly speaking we should write
n(si , sj ) but prefer less uncluttered notation.] We assume n(i, j) to be finite
for all si , sj . The pair (G, I) is called a Coxeter group if the relations
negative roots and denote by Σ the simple positive roots. Let I = {s1 , . . . , sr }
be of simple reflections. By definition, W is generated by the set I. Let n(i, j)
denote the order of si sj . We will show that (W, I) is a Coxeter group. It is
evident that the relations (25.1) are satisfied, but we need to see that they
give a presentation of W .
Theorem 25.1. Let W be the Weyl group of the root system Φ, and let I =
{s1 , . . . , sr } be the simple reflections. Then (W, I) is a Coxeter group.
We will give a geometric proof of this fact, making use of the system of Weyl
chambers. As it turns out, every Coxeter group has a geometric action on a
simplicial complex, a Coxeter complex, which for Weyl groups is closely related
to the action on Weyl chambers. This point of view leads to the theory of build-
ings. See Tits [163] and Abramenko and Brown [1] as well as Bourbaki [23].
Proof. Let (W , I ) be the Coxeter group with generators s1 , . . . , sj and
the relations (25.1) where n(i, j) is the order of si sj . Since the relations
(25.1) are true in W we have a surjective homomorphism W −→ W send-
ing si → si . We must show that it is injective. Let si1 · · · sin = sj1 · · · sjm
be two decompositions of the same element w into products of simple reflec-
tions. We will show that we may go from one word w = (si1 , . . . , sin ) to the
other w = (sj1 , . . . , sjm ) , only making changes corresponding to relations in
the Coxeter group presentation. That is, we may insert or remove a pair of
adjacent equal si , or we may replace a segment (si , sj , si , . . .) by (sj , si , sj , . . .)
where the total number of si and sj is each the order of 2n(si , sj ). This will
show that si1 · · · sin = sj1 · · · sjm so the homomorphism W −→ W is indeed
injective.
Let C be the positive Weyl chamber. We have si1 · · · sin C = sj1 · · · sjm C.
Let
w1 = si1 , w2 = si1 si2 , · · · wn = si1 · · · sin .
The sequence of chambers
C, w1 C, w2 C, . . . , wn C = wC (25.2)
are adjacent. We will say that (C, w1 C, w2 , . . . , wC) is the gallery associated
with the word w = (si1 , . . . , sik ) representing w. We find a path p from a point
in the interior of C to wC passing exactly through this sequence of chambers.
Similarly we have a gallery associated with the word (sj1 , . . . , sjm ). We may
similarly consider a path q from C to wC having the same endpoints as p
passing through the chambers of this gallery. We will consider what happens
when we deform p to q.
If α ∈ Φ let Hα be the set of v ∈ V such that α∨ (v) = 0. This is the hyper-
plane perpendicular to the root α, and these hyperplanes are the walls of Weyl
chambers. Let K2 be the closed subset of V where two or more hyperplanes
Hα intersect. It is a subset of codimension 2, that is, it is a (locally) finite
25 Coxeter Groups 215
wk−1C
wk−1C = wk+1C
wkC
There are two ways the word can change. If ik = ik+1 then wk−1 = wk+1
and we have a crossing as in Fig. 25.1. The path may move to eliminate
(or create) the crossing. This corresponds to eliminating (or inserting) a rep-
eated sik = sik+1 from the word, and since sik has order 2 in the Coxeter
group, the corresponding elements of the Coxeter group will be the same.
Since the deformation avoids K3 , the only other way that the word can
change is if the deformation causes the path to cross K2 , that is, some point
where two or more hyperplanes Hα1 , Hα2 , . . . , Hαn intersect, with the roots
α1 , . . . , αn in a two-dimensional subspace. In this case the transition looks
like Fig. 25.2.
This can happen if ik = ik+2 = · · · = i and ik+1 = ik+3 = · · · = j, and
the effect of the crossing is to replace a subword of the form (si , sj , si , . . .) by
an equivalent (sj , si , sj , . . .), where the total number of si and sj is 2n(si , sj ).
We have si sj si · · · = sj si sj · · · in W , so this type of transition also does not
change the element of W . We see that si1 · · · sin = sj1 · · · sjm , proving that
W ∼ = W . This concludes the proof that W is a Coxeter group.
The Coxeter group (W, I) has a close relative called the associated braid group.
We note that in the Coxeter group (W, I) with generators si satisfying s2i = 1,
the relation (si sj )n(i,j) = 1 (true when i
= j) can be written
si sj si · · · = sj si sj · · · , (25.3)
216 25 Coxeter Groups
wkC
wk+3C
wk+1C wk−1C wk−1C
wkC
wk+2C
wk+1C
where the number of factors on both sides is n(i, j). Written this way, we call
equation (25.3) the braid relation.
Now let us consider a group B with generators ui in bijection with the si
that are assumed to satisfy the braid relations but are not assumed to be of
order two. Thus,
ui uj ui uj · · · = uj ui uj ui · · · , (25.4)
where there are n(i, j) terms on both sides. Note that since the relation s2i = 1
is not true for the ui , it is not true that n(i, j) is the order of ui uj , and in
fact ui uj has infinite order. The group B is called the braid group.
The term braid group is used due to the fact that the braid group of type
An is Artin’s original braid group, which is a fundamental object in knot
theory. Although Artin’s braid group will not play any role in this book,
abstract braid groups will play a role in our discussion of Hecke algebras in
Chap. 46, and the relationship between Weyl groups and braid groups und-
erlies many unexpected developments beginning with the use by Jones [91]
of Hecke algebras in defining new knot invariants and continuing with the
work of Reshetikhin and Turaev [135] based on the Yang–Baxter equation,
with connections to quantum groups and applications to knot and ribbon
invariants.
Consider a set of paths represented by a set of n+1 nonintersecting strings
connected to two (infinite) parallel posts in R3 to be a braid . Braids are
equivalent if they are homotopic. The “multiplication” in the braid group
is concatenation: to multiply two braids, the endpoints of the first braid on
the right post are tied to the endpoints of the second braid on the left post.
In Fig. 25.3, we give generators u1 and u2 for the braid group of type A2 and
calculate their product. In Fig. 25.4, we consider u1 u2 u1 and u2 u1 u2 ; clearly
these two braids are homotopic, so the braid relation u1 u2 u1 = u2 u1 u2 is
satisfied.
We did not have to make the map n part of the defining data in the Coxeter
group since n(i, j) is just the order of si sj . This is no longer true in the braid
group. Coxeter groups are often finite, but the braid group (B, I) is infinite if
|I| > 1.
25 Coxeter Groups 217
u1 × u2 = u 1u 2
As a typical example of how the theorem of Matsumoto and Tits is used, let
us define the divided difference operators Di on E. They were introduced by
Lascoux and Schutzenberger, and independently by Bernstein, Gelfand, and
Gelfand, in the cohomology of flag varieties. The divided difference operators
are sometimes denoted ∂i , but we will reserve that notation for the Demazure
operators we will introduce below. Di acts on the group algebra of the weight
lattice Λ; this algebra was denoted E in Chap. 22. It has a basis eλ indexed
by weights λ ∈ Λ. We define
−1
Di f = (eαi − 1) (f − si f ) .
Di Dj Di · · · = Dj Di Dj · · · (25.5)
where the number of factors on both sides is n(i, j). Moreover, this equals
- .
7 l(w)
H(α) (−1) w
α w
where the product is over roots α in the rank two root system spanned by αi
and αj , and the sum is over the rank two Weyl group generated by si , sj .
Proof. This calculation can be done separately for the four possible cases
n(i, j) = 2, 3, 4 or 6. The case n(i, j) = 2 is trivial so let us assume n(i, j) = 3.
We will show that
l(w)
Di Dj Di = H(αi )H(αj )H(αi + αj ) (−1) w, (25.6)
w∈
si ,sj
∂i2 = ∂i , si ∂i = ∂i ,
−1
Proof. We have si ∂i = (1 − eαi ) (s − eαi ) since si eλ s−1
i = esi (λ) and in par-
ticular si e−αi s−1
i = e αi
. Multiplying both the numerator and the denomina-
tor by −e−αi then shows that si ∂i = ∂i . This identity shows that for any f ∈ E
the element ∂i f is si invariant. Moreover, if f is si -invariant, then ∂i f = f
−1
because ∂i f = (1 − e−αi ) (1 − e−αi ) f = f . Since ∂i f is si invariant, we have
∂i2 f = ∂i f . The action of D on E is easily seen to be faithful so this proves
∂i2 = ∂i (or check this by direct computation. The last identity follows from the
−1
formula for a finite geometric series, (1 − x) 1 − xN +1 = 1 + x + · · · + xN
together with si λ = λ − α∨ i (λ) αi .
It is easy to check that the Demazure and divided difference operators are
related by Di = ∂i − 1.
Proposition 25.3. The Demazure operators also satisfy the braid relations
Di Dj Di · · · = Dj Di Dj · · · (25.7)
Now, by the theorem of Matsumoto and Tits, we may define ∂w = ∂i1 · · · ∂ik
where w = si1 · · · sik is any reduced expression, and this is well defined.
We return to the setting of Chap. 22. Thus, let λ be a dominant weight in
Λ = X ∗ (T ), where T is a maximal torus in the compact Lie group G. Let χλ
be the character of the corresponding highest weight module, which may be
regarded as an element of E.
Theorem 25.3 (Demazure). Let w0 be the long element in the Weyl group,
and let λ be a dominant weight. Then
χλ = ∂w0 eλ .
25 Coxeter Groups 221
where Δ is the Weyl denominator as in Chap. 22. This is sufficient, for then
∂w0 eλ = χλ when λ is dominant by the Weyl character formula.
Let N = l(w0 ). For each i, l(si w0 ) = N − 1, so we may find a reduced word
si w0 = si2 · · · siN . Then w0 = si si2 · · · siN in which i1 = i, so ∂w0 = ∂i ∂si w0 .
Since ∂i2 = ∂i and si ∂i = ∂i this means that ∂i ∂w0 = ∂w0 and si ∂w0 = ∂w0 .
A consequence is that ∂w0 eλ is W -invariant
for every weight λ (dominant or
not). Therefore, if we write ∂w0 = fw · w with fw ∈ M we have w(fw ) =
fww . Since w(Δ−1 ) = (−1)
l(w)
, we now have only to check (25.8) for one
particular w. Fortunately when w = w0 it is possible to do this without too
much work. Choosing a reduced word, we have
−1
−1
∂w0 = 1 − e−αi1 1 − e−αi1 si1 · · · 1 − e−αiN 1 − e−αiN siN .
Expanding out thefactors 1 − sik e−αk there is only one way to get w0 in
the decomposition fw · w, namely we must take −sik e−αik in every factor.
Therefore,
−1 −α1
−1 −αN
fw0 · w0 = 1 − e−αi1 −e si1 · · · 1 − e−αiN −e siN
= (−1)l(w0 ) H(α1 ) si1 · · · H(αiN ) siN .
Theorem 25.4. The affine Weyl group is also a Coxeter group (generated by
s0 , . . . , sr ). Moreover, the analog of the Matsumoto-Tits theorem is true for
the affine Weyl group: if w of length k has two reduced decompositions w =
si1 · · · sik = sj1 · · · sjk , then the word (si1 , . . . , sik ) may be transformed into
(sj1 , . . . , sjk ) by a series of substitutions, in which a subword (si , sj , si , . . .) is
changed to (sj , si , sj , . . .), both subwords having n(i, j) elements.
222 25 Coxeter Groups
Proof. This may be proved by the same method as Theorem 25.1 and 25.2
(Exercise 25.3).
As a last application of the theorem of Matsumoto and Tits, we discuss the
Bruhat order on the Weyl group, which we will meet again in Chap. 27. This
is a partial order, with the long Weyl group element maximal and the identity
element minimal. If v and u are elements of the Weyl group W , then we write
u v if, given a reduced decomposition v = si1 · · · sik then there exists a
subsequence (j1 , . . . , jl ) of (i1 , . . . , ik ) such that u = sj1 · · · sjl . By Proposi-
tion 20.4 we may assume that u = sj1 · · · sjl is a reduced decomposition.
Proposition 25.4. This definition does not depend on the reduced decompo-
sition v = si1 · · · sik .
Proof. By Theorem 25.2 it is sufficient to check that if (i1 , . . . , ik ) is changed
by a braid relation, then we can still find a subsequence (j1 , . . . , jl ) repre-
senting u. We therefore find a subsequence of the form (t, u, t, . . .) where the
number of elements is the order of st su , and we replace this by (u, t, u, . . .).
We divide the subsequence (j1 , . . . , jl ) into three parts: the portion extracted
from that part of (i1 , . . . , ik ) before the changed subsequence, the portion ex-
tracted from the changed subsequence, and the portion extracted from after
the changed subsequence. The first and last part do not need to be altered.
A subsequence can be extracted from the portion in the middle to repre-
sent any element of the dihedral group generated by st and su whether it is
(t, u, t, . . .) or (u, t, u, . . .), so changing this portion has no effect.
We now describe (without proof) the classification of the possible reduced
root systems and their associated finite Coxeter groups. See Bourbaki [23] for
proofs. If Φ1 and Φ2 are root systems in vector spaces V1 , V2 , then Φ1 ∪ Φ2 is
a root system in V1 ⊕ V2 . Such a root system is called reducible. Naturally, it
is enough to classify the irreducible root systems.
The Dynkin diagram represents the Coxeter group in compact form. It is
a graph whose vertices are in bijection with Σ. Let us label Σ = {α1 , . . . , αr },
and let si = sαi . Let θ(αi , αj ) be the angle between the roots αi and αj . Then
⎧
⎪
⎪ 2 if θ(αi , αj ) = π2 ,
⎪
⎪
⎪
⎪
⎪
⎪
⎨ 3 if θ(αi , αj ) = 2π
3 ,
n(si , sj ) =
⎪
⎪
⎪
⎪ 4 if θ(αi , αj ) = 3π
4 ,
⎪
⎪
⎪
⎪
⎩
6 if θ(αi , αj ) = 5π
6 .
connect them with a single bond; if they make an angle of 6π/4, we connect
them with a double bond; and if they make an angle of 5π/6, we connect them
with a triple bond. The latter case only arises with the exceptional group G2 .
If αi and αj make an angle of 3π/4 or 5π/6, then these two roots have
different lengths; see Figs. 19.4 and 19.6. In the Dynkin diagram, there will
be a double or triple bond in these examples, and we draw an arrow from
the long root to the short root. The triple bond (corresponding to an angle
of 5π/6) is rare—it is only found in the Dynkin diagram of a single group,
the exceptional group G2 . If there are no double or triple bonds, the Dynkin
diagram is called simply laced.
α1 α2 α3 α4 α5
Fig. 25.5. The Dynkin diagram for the type A5 root system
The root system of type An is associated with the Lie group SU(n + 1).
The corresponding abstract root system is described in Chap. 19. All roots
have the same length, so the Dynkin diagram is simply laced. In Fig. 25.5
we illustrate the Dynkin diagram when n = 5. The case of general n is the
same—exactly n nodes strung together in a line (•—•— · · · —•).
α1 α2 α3 α4 α5
Fig. 25.6. The Dynkin diagram for the type B5 root system
The root system of type Bn is associated with the odd orthogonal group
SO(2n + 1). The corresponding abstract root system is described in Chap. 19.
There are both long and short roots, so the Dynkin diagram is not simply
laced. See Fig. 25.6 for the Dynkin diagram of type B5 . The general case is
the same (•—•— · · · —•= >
=•), with the arrow pointing towards the αn node
corresponding to the unique short simple root.
α1 α2 α3 α4 α5
Fig. 25.7. The Dynkin diagram for the type C5 root system
The root system of type Cn is associated with the symplectic group Sp(2n).
The corresponding abstract root system is described in Chap. 19. There are
both long and short roots, so the Dynkin diagram is not simply laced. See
Fig. 25.7 for the Dynkin diagram of type C5 . The general case is the same
(•—•— · · · —•=<=•), with the arrow pointing from the αn node corresponding
to the unique long simple root, towards αn−1 .
224 25 Coxeter Groups
α5
α1 α2 α3 α4
α6
Fig. 25.8. The Dynkin diagram for the type D6 root system
The root system of type Dn is associated with the even orthogonal group
O(2n). All roots have the same length, so the Dynkin diagram is simply-laced.
See Fig. 25.8 for the Dynkin diagram of type D6 . The general case is similar,
but the cases n = 2 or n = 3 are degenerate, and coincide with the root
systems A1 × A1 and A3 . For this reason, the family Dn is usually considered
to begin with n = 4. See Fig. 30.2 and the discussion in Chap. 30 for further
information about these degenerate cases.
These are the “classical” root systems, which come in infinite families.
There are also five exceptional root systems, denoted E6 , E7 , E8 , F4 and G2 .
Their Dynkin diagrams are illustrated in Figs. 25.9–25.12.
α2
α1 α3 α4 α5 α6
Fig. 25.9. The Dynkin diagram for the type E6 root system
α2
α1 α3 α4 α5 α6 α7
Fig. 25.10. The Dynkin diagram for the type E7 root system
α2
α1 α3 α4 α5 α6 α7 α8
Fig. 25.11. The Dynkin diagram for the type E8 root system
25 Coxeter Groups 225
α1 α2 α3 α4 α1 α2
Exercises
Exercise 25.1. For the root systems of types An , Bn , Cn , Dn and G2 described in
Chap. 19, identify the simple roots and the angles between them. Confirm that their
Dynkin diagrams are as described in this chapter.
Exercise 25.2. Let Φ be a root system in a Euclidean space V. Let W be the Weyl
group, and let W be the group of all linear transformations of V that preserve Φ.
Show that W is a normal subgroup of W and that W /W is isomorphic to the
group of all symmetries of the Dynkin diagram of the associated Coxeter group.
(Use Proposition 20.13.)
Exercise 25.3. Prove Theorem 25.4 by imitating the proof of Theorem 25.1
Exercise 25.4. How many reduced expressions are there in the A3 Weyl group
representing the long Weyl group element?
Exercise 25.6. Let W be a Weyl group. Let w = si1 si2 · · · siN be a decompo-
sition of w into a product of simple reflections. Construct a path through the
sequence (25.2) of chambers as in the proof of Theorem 25.1. Observe that the
word (i1 , i2 , . . . , iN ) representing w is reduced if and only if this path does not cross
any of the hyperplanes H orthogonal to the roots twice. Suppose that the word is
not reduced, and that it meets some hyperplane H in two points, P and Q. Then for
some k, with notation as in (25.2), P lies between wk−1 C and wk−1 sik C. Similarly
Q lies between wl−1 C and wl−1 sil C. Show that ik = il , and that
where the “hat” means that the two entries are omitted. (Hint: Reflect the segment
of the path between P and Q in the hyperplane H.)
Exercise 25.7. Prove that the Bruhat order has the following properties.
[Hint: Any one of these four properties implies the others. For example, to deduce
(ii) from (i), replace u by su].
226 25 Coxeter Groups
Observe that su < u if and only if l(su) < l(u), a condition that is easy to check.
Therefore, (i) and (ii) give a convenient method of checking (recursively) whether
u v.
Exercise 25.8. Let w0 be the long element in a Weyl group W . Show that if u, v ∈
W then u v if and only if uw0 vw0 .
26
The Borel Subgroup
For example, if n = 3,
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎨ 0 ∗ ∗ ⎬ ⎨ 0 0 ∗ ⎬
n = n1 = ⎝ 0 0 ∗ ⎠ , n2 = ⎝ 0 0 0 ⎠ , n3 = {0}.
⎩ ⎭ ⎩ ⎭
0 0 0 0 0 0
We also say that a Lie algebra b is solvable if there exists a finite chain of
Lie subalgebras
b = b1 ⊃ b2 ⊃ · · · ⊃ bN = {0} (26.1)
such that [bi , bi ] ⊆ bi+1 . It is not necessarily true that bi is an ideal in b.
However, the assumption that [bi , bi ] ⊆ bi+1 obviously implies that [bi , bi+1 ] ⊆
bi+1 , so bi+1 is an ideal in bi .
Clearly, a nilpotent Lie algebra is solvable. The converse is not true, as
the next example shows.
Example 26.2. Let F be a field, and let b be the Lie algebra over F consisting
of all upper triangular matrices in GL(n, F ). Let
Thus, if n = 3,
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎨ ∗ ∗ ∗ ⎬ ⎨ 0 ∗ ∗ ⎬
b = b1 = ⎝ 0 ∗ ∗ ⎠ , b2 = ⎝ 0 0 ∗ ⎠ ,
⎩ ⎭ ⎩ ⎭
0 0 ∗ 0 0 0
⎧⎛ ⎞⎫
⎨ 0 0 ∗ ⎬
b3 = ⎝ 0 0 0⎠ , b4 = {0}.
⎩ ⎭
0 0 0
as required.
Proof. To prove (i), we may clearly assume that b
= 0. Let us first observe
that b has an ideal h of codimension 1. Indeed, since b is solvable, [b, b] is
a proper ideal, and the quotient Lie algebra b/[b, b] is Abelian; hence any
26 The Borel Subgroup 231
Proof. Let Φ+k be the set of positive roots α such that α is expressible as the
sum of at least k simple positive roots. Thus, Φ+1 = Φ, Φ1 ⊃ Φ2 ⊃ Φ3 ⊃ · · · ,
+ + +
+
and eventually Φk is empty. Define
&
nk = Xα .
α∈Φ+
k
It follows from Proposition 18.4 (ii) that [n, nk ] ⊆ nk+1 , and eventually nk is
zero, so n is nilpotent.
Now let t be the Lie algebra of T , and let b = tC ⊕n. Since [tC , Xα ] ⊆ Xα , it
is clear that b, like n, is closed under the Lie bracket and forms a complex Lie
algebra. Moreover, since tC is Abelian and normalizes n, we have [b, b] ⊂ n,
and since n is nilpotent and hence solvable, it follows that b is solvable.
232 26 The Borel Subgroup
We aim to show that both n and b are the Lie algebras of closed complex
Lie subgroups of G.
Proof. By Theorem 26.1, we may choose a basis of V such that all π(X) are
upper triangular for X ∈ b, where we are identifying π(X) with its matrix
with respect to the chosen basis. What we must show is that if X ∈ n, then
the diagonal entries of this matrix are zero. It is sufficient to show this if
X ∈ Xα , where α is a positive root.
By the definition of a root, the character α of T is nonzero, and so its
differential dα is nonzero. This means that there exists H ∈ t such that
dα(H)
= 0, and by (18.9) the commutator [π(H), π(Xα )] is a nonzero multiple
of π(Xα ). Because it is a nonzero multiple of the commutator of two upper
triangular matrices, it follows that π(Xα ) is an upper triangular matrix with
zeros on the diagonal. Thus, it is nilpotent.
Y = I + X + 12 X 2 + . . . + 1 n
n! X .
This is now a series with only finitely many terms since X is nilpotent by
Proposition 26.5. Moreover, Y − I is a finite sum of upper triangular nilpotent
matrices and hence is itself nilpotent, and reverting the exponential series, we
have X = log(Y ), where we define
A×N×K A × N × K
exp exp
G G
where the vertical arrows are multiplications and the horizontal arrows are
inclusions. By Proposition 26.1, the composition
A × N × K −→ A × N × K −→ G (26.7)
Proof. Let n be the algebra generated by the Xα with α simple. Let us define
the height of a positive root to be the number of simple roots into which it
may be decomposed, counted with multiplicities. If α ∈ Φ+ is not simple, we
may write α = β + γ where β and γ are in Φ+ , and by induction on the height
of α, we may assume that Xβ and Xγ are in n . By Corollary 18.1, [Xβ , Xγ ]
is a nonzero multiple of Xα and so Xα is in n . Thus, n = n.
Xk = det(Xij )1i,jk .
and since the value of xμ on X is X1μ1 −μ2 X2μ2 −μ3 · · · if this is a polynomial of
X we must have a(μ) = 0 unless μ1 μ2 · · · . Thus the xμ with μ dominant
form a basis of the N -invariants. We have seen that there is one irreducible
representation for each basis vector. Under the action in Proposition 26.8, the
μ
@ 2x nis 2μ. Hence the highest weights of the irreducible rep-
weight of the vector
resentations in (∨ C ) are the even dominant weights.
Theorem 26.7. Assume that G is an affine algebraic group over the complex
numbers. Assume that it is also the complexification of a compact Lie group K.
Let X be a complex affine algebraic variety on which the group G acts
algebraically. Assume that the Borel subgroup B has a dense open orbit in
X. Let W be the space of algebraic functions on X. Then W is a multiplicity-
free G-module.
The open orbit, if it exists, is always unique and dense, so the word “dense”
could be eliminated from the statement. The theory of algebraic group actions
is an important topic, and the standard monograph is Mumford, Fulton, and
Kirwan [132]. Varieties (whether affine or not) with an open B-orbit are called
spherical .
240 26 The Borel Subgroup
Exercises
Exercise 26.1. Let Ω be the vector space of n × n skew-symmetric matrices. G =
GL(n, C) acts on Ω by the space of polynomial functions on Ω by (gf )(X) = f (t g ·
X · g).
(i) Show that the symmetric algebra on the exterior square of the standard module
of GL(n, C), that is, (∧2 Cn ), is isomorphic as a G-module to the ring of
polynomialfunctions on Ω.
(ii) Show that (∧2 Cn ) decomposes as a direct sum of irreducible representations,
each with multiplicity one, and that if λ = (λ1 , λ2 , . . . , λn ) is a dominant weight,
then λ occurs in this decomposition if and only if λ1 = λ2 , λ3 = λ4 , . . ., and if
n is odd, then λn = 0.
Exercise 26.2. Let G = GL(n, C), and let H = O(n, C). As in Proposition 26.8, let
Ω be space of symmetric n × n complex matrices, and let Ω ◦ be the open subset of
invertible n × n matrices. Let P(Ω) be the ring of polynomials on Ω, and let P(Ω ◦ )
be the space of polynomial functions on Ω ◦ ; it is generated by P(Ω) together with
g −→ det(g)−1 . The group G acts on both P(Ω) and P(Ω ◦ ) as in Proposition 26.8.
The stabilizer of I ∈ P(Ω ◦ ) is the group H, and the action on P(Ω ◦ ) is transitive,
so P(Ω ◦ ) is in bijection with G/H. Let (π, V ) be an irreducible representation of G.
(i) Show that (π, V ) has a nonzero H-fixed vector if and only if its contragredient
(π̂, V ∗ ) does. [Hint: Show that π̂ is equivalent to the representation π : G −→ V
defined by π (g) = t g −1 by comparing their characters.]
26 The Borel Subgroup 241
(ii) Assume that π has a nonzero H-fixed vector. By (i) there is an H-invariant
linear functional φ : V −→ C. Define Φ : V −→ P(Ω ◦ ) by letting Φ(v) be the
function Φv defined by
Show that this is well defined and that Φgv (X) = Φv (t gXg). Deduce that v −→
Φv is an embedding of V into P(Ω ◦ ).
(iii) Show that π has a nonzero H-fixed vector if and only if π can be embedded in
P(Ω ◦ ). [Hint: One direction is (ii). For the other, prove instead that π has an
H-invariant linear functional.]
Remark: The argument in (ii) and (iii) is formally very similar to the proof
of Frobenius reciprocity (Proposition 32.2) with P(Ω ◦ ) playing the role of the
induced representation.)
(iv) Show that an irreducible representation of P(Ω ◦ ) can be extended to P(Ω) if
and only if its highest weight λ is effective.
(v) Let πλ be an irreducible representation of G with highest weight λ. Assume that
λ is effective, so that λ is a partition. Show that πλ has an O(n)-fixed vector if
and only if λ is even.
(vi) Assume again that λ is effective, but only assume that πλ has a fixed vector for
SO(n). What λ are possible?
Exercise 26.3. The last exercise shows that if (π, V ) is an irreducible representation
of GL(n, C), then the multiplicity of the trivial representation of O(n, C) in its
restriction to this subgroup is at most one. Show by example that there are other
representations that can occur with higher multiplicity, for example when n = 5.
The next exercise is essentially a proof of the Cauchy identity, which is the
subject of Chap. 38.
Exercise 26.5. Let G be a complex analytic Lie group and let H1 , H2 be closed
analytic subgroups. Then G acts on the homogeneous space G/H1 , as does its sub-
group H2 . The quotient is the space of double cosets, H2 \G/H1 , which might also
be obtained by letting H1 act on the right on H2 \G.
242 26 The Borel Subgroup
(iii) Show that H2 has an open orbit on G/H1 if and only if H1 has an open orbit
on H2 \G.
27
The Bruhat Decomposition
The Bruhat decomposition was discovered quite late in the history of Lie
groups, which is surprising in view of its fundamental importance. It was
preceded by Ehresmann’s discovery of a closely related cell decomposition for
flag manifolds. The Bruhat decomposition was axiomatized by Tits in the
notion of a Group with (B, N ) pair or Tits’ system. This is a generalization
of the notion of a Coxeter group, and indeed every (B, N ) gives rise to a
Coxeter group. We have remarked after Theorem 25.1 that Coxeter groups
always act on simplicial complexes whose geometry is closely connected with
their properties. As it turns out a group with (B, N ) pair also acts on a
simplicial complex, the Tits’ building. We will not have space to discuss this
important concept but see Tits [163] and Abramenko and Brown [1].
In this chapter, in order to be consistent with the notation in the litera-
ture on Tits’ systems, particularly Bourbaki [23], we will modify our notation
slightly. In other chapters such as the previous one, N denotes the subgroup
(26.6) of the Borel subgroup. That group will appear in this Chapter also,
but we will denote it as U , reserving the letter N for the normalizer of T .
Similarly, in this chapter U will be the subgroup formerly denoted N .
Let G = GL(n, F ), where F is a field, and let B be the Borel subgroup of
upper triangular matrices in G. Taking T ⊂ B to be the subgroup of diagonal
matrices in G, the normalizer N (T ) consists of all monomial matrices. The
Weyl group W = N (T )/T ∼ = Sn . If w ∈ W is represented by ω ∈ N (T )
then since T ⊂ B the double coset BωB is independent of the choice of
representative ω, so by abuse of notation we write BwB for BωB. It is a
remarkable and extremely important fact that w −→ BwB is a bijection
between the elements of W and the double cosets B\G/B. We will prove the
following Bruhat decomposition:
A
G= BwB (disjoint). (27.1)
ab
The example of GL(2, F ) is worth writing out explicitly. If g = ,
cd
then g ∈ B if c = 0. Therefore to prove the Bruhat decomposition, then for a
representative
ωof the long Weyl group element it will be convenient to take
0 −Δc−1
ω= where Δ = ad − bc. Then this follows from the identity
c 0
ab 1 a/c 1 d/c
= ω
cd 1 1
We will prove this and also obtain a similar statement in complex Lie
groups. Specifically, if G is a complex Lie group obtained by complexification
of a compact connected Lie group, we will prove a “Bruhat decomposition”
analogous to (27.1) in G. A more general Bruhat decomposition will be found
in Theorem 29.5.
We will prove the Bruhat decomposition for a group with a Tits’ system,
which consists of a pair of subgroups B and N satisfying certain axioms. The
use of the notation N differs from that of Chap. 26, though the results of that
chapter are very relevant here.
Let G be a group, and let B and N be subgroups such that T = B ∩ N is
normal in N . Let W be the quotient group N/T . As with GL(n, F ), we write
wB instead of wB when ω ∈ N represents the Weyl group element w, and
similarly we will denote Bw = Bω and BwB = BωB.
Let G be a group with subgroups B and N satisfying the following condi-
tions.
We will prove in Theorem 27.1 below that this (B, N, I) is a Tits’ system.
The proof will require introducing a root system into GL(n, F ). Of course, we
have already done this if F = C, but let us revisit the definitions in this new
context.
Let X ∗ (T ) be the group of rational characters of T . In case F is a finite
field, we don’t want any torsion in X ∗ (T ); that is, we want χ ∈ X ∗ (T ) to have
infinite order so that R ⊗ X ∗ (T ) will be nonzero. So we define an element of
X ∗ (T ) to be a character of T (F ), the group of diagonal matrices in GL(n, F ),
where F is the algebraic closure of F , of the form
⎛ ⎞
t1
⎜ .. ⎟
⎠ −→ t11 . . . tnn ,
k k
⎝ . (27.5)
tn
where ki ∈ Z. Then X ∗ (T ) ∼ = Zn , so V = R ⊗ X ∗ (T ) ∼
= Rn .
∗
As usual, we write the group law in X (T ) additively.
In this context, by a root of T in G we mean an element α ∈ X ∗ (T ) such
that there exists a group isomorphism xα of F onto a subgroup Xα of G
consisting of unipotent matrices such that
t xα (λ) t−1 = xα α(t) λ , t ∈ T, λ ∈ F. (27.6)
xα (λ) = I + λEij ,
246 27 The Bruhat Decomposition
then (27.6) is clearly valid. The set Φ consisting of αij is a root system; we
leave the reader to check this but in fact it is identical to the root system
of GL(n, C) or its maximal compact subgroup U(n) already introduced in
Chap. 18 when n = C. Let Φ+ consist of the “positive roots” αij with i < j,
and let Σ consist of the “simple roots” αi,i+1 . We will sometimes denote the
simple reflections si = sα , where α = αi,i+1 .
Suppose that α is a simple root. Let Tα ⊂ T be the kernel of α. Let Mα be
the centralizer of Tα , and let Pα be the subgroup generated by B and Mα . By
abuse of language, Pα is called a minimal parabolic subgroup. Observe that it is
a parabolic subgroup since it contains the Borel subgroup. Strictly speaking
it is not minimal amoung the parabolic subgroups, since the Borel itself is
smaller. However it is minimal among non-Borel parabolic subgroups, and it
is commonly called a minimal parabolic. in Chap. 30.) We have a semidirect
product decomposition Pα = Mα Uα , where Uα is the group generated by the
xβ (λ) with β ∈ Φ+ − {α}. For example, if n = 4 and α = α23 , then
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎪
⎪ t1 ⎪
⎪ ⎪
⎪ ∗ ⎪
⎪
⎨⎜ ⎟⎬ ⎨⎜ ⎟⎬
t2 ∗ ∗
Tα = ⎜⎝
⎟ ,
⎠⎪ Mα = ⎜
⎝ ∗ ∗
⎟ ,
⎠⎪
⎪
⎪ t2 ⎪ ⎪
⎪ ⎪
⎩ ⎭ ⎩ ⎭
t4 ∗
⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎪
⎪ ∗ ∗ ∗ ∗ ⎪⎪ ⎪
⎪ 1 ∗ ∗ ∗ ⎪⎪
⎨⎜ ⎬ ⎨⎜ ⎬
∗ ∗ ∗⎟ 1 ∗⎟
Pα = ⎜ ⎟ ,
⎝ ∗ ∗ ∗ ⎠⎪ Uα = ⎜⎝
⎟ ,
⎪
⎪ ⎪ ⎪
⎪ 1 ∗ ⎠⎪
⎪
⎩ ⎭ ⎩ ⎭
∗ 1
where ∗ indicates an arbitrary value.
Lemma 27.1. Let G = GL(n, F ) for any field F , and let other notations be
as above. If s is a simple reflection, then B ∪ C(s) is a subgroup of G.
Proof. First, let us check this when n = 2. In this case, there is only one
simple root sα where α = α12 . We check easily that
ab
C(sα ) = Bsα B = ∈ GL(2, F ) c
= 0 ,
cd
so C(sα ) ∪ B = G.
In the general case, both C(sα ) and B are subsets of Pα . We claim that
their union is all of Pα . Both double cosets are right-invariant by Uα since
Uα ⊂ B, so it is sufficient to show that C(sα ) ∪ B ⊃ Mα . Passing to the
quotient in Pα /Uα ∼= Mα ∼ = GL(2) × (F × )n−2 , this reduces to the case n = 2
just considered.
Lemma 27.2. Let G = GL(n, F ) for any field F , and let other notations
be as above. If α is a simple root and w ∈ W such that w(α) ∈ Φ+ , then
C(w) C(s) = C(ws).
Proposition 27.1. Let G = GL(n, F ) for any field F , and let other notations
be as above. If w, w ∈ W are such that l(ww ) = l(w) + l(w ), then
Indeed, assuming we know this fact, let w = s1 . . . sr be a decomposition into
simple reflections with r = l(r ). Then s1 . . . sr s1 . . . sr is a decomposition of
ww into simple reflections with l(ww ) = r + r , so
Proof. Only Axiom TS3 requires proof; the others can be safely left to the
reader. Let α ∈ Σ such that s = sα .
First, suppose that w(α) ∈ Φ+ . In this case, it follows from Lemma 27.2
that wBs ⊂ BwsB.
Next suppose that w(α)
∈ Φ+ . Then wsα (α) = w(−α) = −w(α) ∈ Φ+ , so
we may apply the case just considered, with wsα replacing w, to see that
Bs ⊂ sB ∪ sBsB.
Using (27.10),
by Axiom TS3. Since two double cosets are either disjoint or equal, this means
that either
C(w ) = C(w ) or C(w ) = C(w s).
Our induction hypothesis implies that either w = w or w = w s. The first
case is impossible since l(w ) < l(w) l(w ). Therefore w = w s. Hence
w = w s = w , as required.
Exercises
Exercise 27.1. Explain why Yw has complex dimension l(w), or real dimension
2l(w). Also explain why Yw is open in Xw . Since Xw is a union of Yw and subsets
of lower dimension, we may say that l(w) is the dimension of the Schubert variety
Xw .
If W is a finite Weyl group then W has a longest element w0 . By the last exercise,
the Bruhat cell Bw0 B is the largest in the sense of dimension. It is therefore called
the big Bruhat cell.
Exercise 27.2. Let G = GL(n, C). Show that g = (gij ) ∈ G is in the big Bruhat
cell if and only if all the bottom left minors
gn−2,1 gn−2,2 gn−2,3
gn−1,1 gn−1,2
gn,1 , gn−1,1 gn−1,2 gn−1,3 ,
gn,1 gn,2 , ...
gn,1 gn,2 gn,3
Xw −
−−−−
→ Xw
where the horizontal maps are the bijections described above, φ is the canonical
map Zw → Xw , and Φ is the map that sends the configuration (U0 , . . . , Uk ) to its
last flag Uk .
V3
V2 U2
V1 W1 U1
V0
Exercise 27.7. (i) Show (for GL(3)) that if w = (1, 2) or (2, 1) φ is an isomorphism.
(ii) Give a similar analysis when w = (2, 1, 2).
V4 V4
V3 U3 V3 W3 U3
V2 W2 U2 V2 U2
V1 U1 V1 W1 U1
V0 V0
K = {g ∈ G | θ(g) = g} (28.1)
Remark 28.1. We will see in Theorem 28.3 that every noncompact semisimple
Lie group admits a Cartan involution θ such that this hypothesis is satisfied.
Our proof of Theorem 28.3 will not be self-contained, but we do not really
need to rely on it as motivation because we will give numerous examples in
this chapter and the next where Hypothesis 28.1 is satisfied.
g : x −→ g x t g. (28.3)
On the other hand, the group U(n) acts on the space En (R) of unitary sym-
metric matrices by the same formula (28.3). [The notation En (R) does not
imply that the elements of this space are real matrices.]
Proof. If x ∈ Pn (R), then (i) is, of course, just the spectral theorem. However,
if x ∈ En (R), this statement may be less familiar. It is instructive to give a
unified
proof of the two cases. Give Cn its usual inner product, so u, v =
i ui vi .
Let λ be an eigenvalue of x. We will show that the eigenspace Vλ = {v ∈
Cn | xv = λv} is stable under complex conjugation. Suppose that v ∈ Vλ .
If x ∈ Pn (R), then both x and λ are real, and simply conjugating the identity
−1
xv = λv gives xv = λv. On the other hand, if x ∈ En (R), then x = t x = x−1
and |λ| = 1 so λ = λ−1 . Thus, conjugating xv = λv gives x−1 v = λ−1 v, which
implies that xv = λv.
Now we can show that Cn has an orthonormal basis consisting of eigen-
vectors v1 , . . . , vn such that vi ∈ Rn . The adjoint of x with respect to the
standard inner product is x or x−1 depending on whether x ∈ Pn (R) or En (R).
In either case, x is the matrix of a normal operator—one that commutes with
its adjoint—and Cn is the orthogonal direct sum of the eigenspaces of x.
Each eigenspace has an orthonormal basis consisting of real vectors. Indeed,
if v1 , . . . , vk is a basis of Vλ , then since we have proved that vi ∈ Vλ , the space
is spanned by 12 (vi + vi ) and 2i 1
(vi − vi ); selecting a basis from this spanning
set and applying the usual Gram–Schmidt orthogonalization process gives an
orthonormal basis of real vectors.
In either case, we see that Cn has an orthonormal basis consisting of
eigenvectors v1 , . . . , vn such that vi ∈ Rn . Let xvi = λi vi . Then, if k ∈ O(n) is
the matrix with columns xi and d is the diagonal matrix with diagonal entries
λi , we have xk = kd so k −1 xk = δ. As k −1 = t k we may take the matrix
g = k −1 . If the determinant of k is −1, we can switch the sign of the first
entry without harm, so we may assume k ∈ SO(n) and (i) is proved.
For (i), we have shown that each orbit in Pn (R) or En (R) contains a
diagonal matrix. The eigenvalues are positive real if x ∈ Pn (R) or of absolute
value 1 if x ∈ En (R). In either case, applying the action (28.3) with g ∈
GL(n, R) or U(n) diagonal will reduce to the identity, proving (ii). For (iii),
we use (ii) to write an arbitrary element x of Pn (R) or En (R) as kdk −1 , where
k is orthogonal and d diagonal. The eigenvalues of d are either positive real
if x ∈ Pn (R) or of absolute value 1 if x ∈ En (R). Thus, d = eY , where Y is
real or purely imaginary, and x = eX or eiX , where X = kY k −1 or −ikY k −1
is real.
We will show that the Type II and Type IV symmetric spaces are in
duality. For this, we need a couple of lemmas. If R is a ring and e, f ∈ R we
call e and f orthogonal central idempotents if ex = xe and f x = xf for all
x ∈ R, e2 = e, f 2 = f , and ef = f e.
Proof. Let
e = 12 (1 ⊗ 1 + i ⊗ i), f = 12 (1 ⊗ 1 − i ⊗ i). (28.4)
It is easily checked that e and f are orthogonal central idempotents whose
sum is the identity element 1 ⊗ 1, and so we obtain a Peirce decomposition by
Lemma 28.1. The ideals generated by e and f are both isomorphic to C.
Theorem 28.1. Let K0 be a compact connected Lie group. Then the compact
and noncompact symmetric spaces of Examples 28.2 and 28.3 are in duality.
262 28 Symmetric Spaces
θ : a ⊗ b ⊗ X −→ a ⊗ b ⊗ X, a, b ∈ C, X ∈ k0 .
Table 28.1. Real forms and Type I and Type III symmetric spaces
where Inn(Gc ) is the group of inner automorphisms, then the group is an inner
form. Looking ahead to the next chapter, where we introduce the Satake dia-
grams, G is an inner form if and only if the symmetry of the Satake diagram,
corresponding to the permutation α −→ −θ(α) of the relative root system, is
trivial. Thus, from Fig. 29.3, we see that SO(6, 6) is an inner form, but the
quasisplit group SO(7, 5) is an outer form. For the exceptional groups, only
E6 admits an outer automorphism (corresponding to the nontrivial automor-
phism of its Dynkin diagram). Thus, for the other exceptional groups, this
parameter is omitted from the notation.
The number r is the (real) rank—the dimension of the group A = exp(a),
where a is a maximal Abelian subspace of p. The number d is the dimension
of the anisotropic kernel , which is the maximal compact subgroup of the
centralizer of A. Both A and M will play an extensive role in the next chapter.
We have listed the rank for the groups of classical type but not the excep-
tional ones since for those the rank is contained in the Tits’ classification.
For classification matters we recommend Tits [162] supplemented by Borel
[20]. The definitive classification in this paper, from the point of view of alge-
braic groups, includes not only real groups but also groups over p-adic fields,
number fields, and finite fields. Knapp [106], Helgason [66], Onishchik and
Vinberg [166], and Satake [144] are also very useful.
Example 28.4. Consider SL(2, R)/SO(2) and SU(2)/SO(2). Unlike the general
case of SL(n, R)/SO(n) and SU(n)/SO(n), these two symmetric spaces have
complex structures. Specifically, SL(2, R) acts transitively on the Poincaré
upper half-plane H = {z = x + iy | x, y ∈ R, y > 0} by linear fractional
transformations:
ab az + b
SL(2, R) : z −→ .
cd cz + d
The stabilizer of the point i ∈ H is SO(2), so we may identify H with
SL(2, R)/SO(2). Equally, let R be the Riemann sphere C ∪ {∞}, which is
the same as the complex projective line P1 (C). The group SU(2) acts transi-
tively on R, also by linear fractional transformations:
a b az + b
SU(2) : z −→ , |a|2 + |b|2 = 1.
−b̄ ā −b̄z + ā
Both H and R are naturally complex manifolds, and the action of SL(2, R)
or SU(2) consists of holomorphic mappings. They are examples of Hermitian
symmetric spaces, which we now define. A Hermitian manifold is the complex
analog of a Riemannian manifold. A Hermitian manifold is a complex manifold
on which there is given a (smoothly varying) positive definite Hermitian inner
product on each tangent space (which has a complex structure because the
266 28 Symmetric Spaces
space is a complex manifold). The real part of this positive definite Hermitian
inner product is a positive definite symmetric bilinear form, so a Hermitian
manifold becomes a Riemannian manifold. A real-valued symmetric bilinear
form B on a complex vector space V is the real part of a positive definite
Hermitian form H if and only if it satisfies
B(iX, iY ) = B(X, Y ),
is the unique positive definite Hermitian form with real part H. From this
remark, a complex manifold is Hermitian by our definition if and only if it is
Hermitian by the definition in Helgason [66].
A symmetric space X is called Hermitian if it is a Hermitian manifold that
is homogeneous with respect to a group of holomorphic Hermitian isometries
that is connected and contains the geodesic-reversing reflection around each
point. Thus, if X = G/K, the group G consists of holomorphic mappings,
and if g(x) = y for x, y ∈ X, g ∈ X, then g induces an isometry between the
tangent spaces at x and y.
The irreducible Hermitian symmetric spaces can easily be recognized by
the following criterion.
Proposition 28.3. Let X = G/K and Xc = Gc /K be a pair of irreducible
symmetric spaces in duality. If one is a Hermitian symmetric space, then they
both are. This will be true if and only if the center of K is a one-dimensional
central torus Z. In this case, the rank of Gc equals the rank of K.
In a nutshell, if K has a one-dimensional central torus, then there exists a
homomorphism of T into the center of K. The image of T induces a group
of isometries of X fixing the base point x0 ∈ X corresponding to the coset
of K. The content of the proposition is that X may be given the structure of
a complex manifold in such a way that the maps on the tangent space at x0
induced by this family of isometries correspond to multiplication by T, which
is regarded as a subgroup of C× .
Proof. See Helgason [66], Theorem 6.1 and Proposition 6.2, or Wolf [176],
Corollary 8.7.10, for the first statement. The latter reference has two other
very interesting conditions for the space to be symmetric. The fact that Gc
and K are of equal rank is contained in Helgason [66] in the first paragraph
of “Proof of Theorem 7.1 (ii), algebraic part” on p. 383.
A particularly important family of Hermitian symmetric spaces are the
Siegel upper half-spaces Hn , also known as Siegel spaces which generalize the
Poincaré upper half-plane H = H1 . We will discuss this family of examples in
considerable detail since many features of the general case are already present
in this example and are perhaps best understood with an example in mind.
28 Symmetric Spaces 267
A · t B = B · t A, C · t D = D · t C,
(28.5)
A · t D − B · t C = I, D · t A − C · t B = I.
Simplifying this gives (28.9). From this it is clear that W is Hermitian and that
W > 0. It is real, of course, though that is less clear from this expression. Since
it is real and positive definite Hermitian, it is a positive definite symmetric
matrix.
It is easy to check that g g (Z) = (gg )(Z), so this is a group action.
To show that this action is transitive, we note that if Z = X + iY ∈ Hn , then
I −X
∈ Sp(2n, R) ,
I
identify Xc with the Riemann sphere R, then the action of SU(2) was by
linear fractional transformations and so is the extension to SL(2, C).
As a consequence, we have an action of G = SL(2, R) on Xc since G ⊂ GC
and GC acts on Xc . This is just the action by linear fractional transformations
on R = C ∪ {∞}. There are three orbits: H, the projective real line P1 (R) =
R ∪ {∞}, and the lower half-plane H.
The Cayley transform is the element c ∈ SU(2) given by
1 −i i i
c = √12i , so c−1 = √12i . (28.11)
1 i −1 1
gave unbounded realizations. Korányi and Wolf [111, 112] gave a completely
general theory relating bounded symmetric domains to unbounded ones by
means of the Cayley transform.
Now let us consider the Cayley transform for Hn . Let G = Sp(2n, R),
K = U(n), Gc = Sp(2n), and GC = Sp(2n, C). Let
1 In −iIn −1 1 iIn iIn
c= √ , c = √ .
2i In iIn 2i −In In
They are elements of Sp(2n). We will see that the scenario uncovered for
SL(2, R) extends to the symplectic group.
Our first goal is to show that Hn can be embedded in its compact dual,
a fact already noted when n = 1. The first step is to interpret Gc /K as an
analog of the Riemann sphere R, a space on which the actions of both groups
G and Gc may be realized as linear fractional transformations. Specifically,
we will construct a space Rn that contains a dense open subspace R◦n that
can be naturally identified with the vector space of all complex symmetric
matrices. What we want is for GC to act on Rn , and if g ∈ GC , with both
Z, g(Z) ∈ R◦n , then g(Z) is expressed in terms of Z by (28.8).
Toward the goal of constructing Rn , let
h I X
P = t −1 h ∈ GL(n, C), X ∈ Mat n (C), X = t
X .
h I
(28.12)
This group is called the Siegel parabolic subgroup of Sp(2n, C). (The term
parabolic subgroup will be formally defined in Chap. 30.) We will define Rn
to be the quotient GC /P . Let us consider how an element of this space can
(usually) be regarded as a complex symmetric matrix, and the action of GC
is by linear fractional transformations as in (28.8).
Proposition 28.5. We have P Gc = Sp(2n, C) and P ∩ Gc = cKc−1 .
Proof. Indeed, P contains a Borel subgroup, the group B of matrices (28.12)
with g upper triangular, so P Gc = Sp(2n, C) follows from the Iwasawa
decomposition (Theorem 26.3). The group K is U(n) embedded via (28.7),
and it is easy to check that
g
cKc−1 = t −1 g ∈ U(n) . (28.13)
g
Proof. Most of this is safely left to the reader. We only point out the reason
that AC −1 is symmetric. By (28.6), the matrix t CA is symmetric, so t C −1 ·
t
CA · C −1 = AC −1 is also.
if and only if AC −1 = Z.
AB
Proposition 28.6. If σ(Z) and g σ(Z) are both in R◦n , where g =
CD
is an element of Sp(2n, C), then CZ + D is invertible and
g σ(Z) = σ (AZ + B)(CZ + D)−1 .
Proof. We have
A B Z −I AZ + B −A
g σ(Z) = P = P .
C D I CZ + D −C
In view of Proposition 28.6 we will identify Rn with its image in R◦n . Thus,
elements of R◦n become for us complex symmetric matrices, and the action of
Sp(2n, C) is by linear fractional transformations.
We can also identify Rn with the compact symmetric space Gc /K by
means of the composition of bijections
Gc /K −→ Gc /cKc−1 −→ GC /P = Rn .
272 28 Symmetric Spaces
where we have used the fact that both W and (W −I)(W +I)−1 are symmetric
to rewrite the second term. This will be positive definite if and only if (W +
I)Y (W + I) is positive definite. This equals
− 21 (W + I)(W − I) + (W − I)(W + I) = I − W W.
Proof. The diagonal entries in W W are the squares of the lengths of the rows
of the symmetric matrix W . If I − W W is positive definite, these must be less
than 1. So Dn is a bounded domain within the set R◦n of symmetric complex
matrices. The rest of (i) is also clear.
For (ii), if g ∈ c G c−1 , then by Proposition 28.7 the matrix g has the form
(28.14). Using the fact that both W and W are symmetric, we have
Like D, the matrix W is diagonal, and its diagonal entries equal to 1 corre-
spond to the diagonal entries of D equal to 0. These correspond to diagonal
entries of I − W W equal to 0, so the diagonal matrices D and I − W W
have the same rank. But by (ii), the ranks of I − W W and I − W W are
equal, so the rank of D is r. Clearly, W has the special form (28.15).
Now let us fix r < n and consider
Wr
Br = r
W ∈ D n−r .
In−r
We now see that both the dual symmetric spaces Pn (R) and En (R) appear
in connection with Hn . The construction of Hn involved building a tube
domain over the cone Pn (R), while the dual En (R) appeared as the Bergman–
Shilov boundary. (Since Pn◦ (R) and En (R)◦ are in duality, it is natural to
extend the notion of duality to the reducible symmetric spaces Pn (R) and
En (R) and to say that these are in duality.)
This scenario repeats itself: there are four infinite families and one isolated
example of Hermitian symmetric spaces that appear as tube domains over
cones. In each case, the space can be mapped to a bounded symmetric domain
by a Cayley transform, and the compact dual of the cone appears as the
Bergman–Shilov boundary of the cone. These statements follow from the work
of Koecher, Vinberg, and Piatetski-Shapiro [133], culminating in Korányi and
Wolf [111, 112].
Let us describe this setup in a bit more detail. Let V be a real vector space
with an inner product , . A cone C ⊂ V is a convex open set consisting of a
union of rays through the origin but not containing any line. The dual cone to
C is {x ∈ V |x, y > 0 for all y ∈ C}. If C is its own dual, it is naturally called
self-dual . It is called homogeneous if it admits a transitive automorphism
group.
A homogeneous self-dual cone is a symmetric space. It is not irreducible
since it is invariant under similitudes (i.e, transformations x −→ λx where
λ ∈ R× ). The orbit of a typical point under the commutator subgroup of
the group of automorphisms of the cone sits inside the cone, inscribed like a
hyperboloid, though this description is a little misleading since it may be the
constant locus of an invariant of degree > 2. For example, Pn◦ (R) is the locus
of det(x) = 1, and det is a homogeneous polynomial of degree n.
Homogeneous self-dual cones were investigated and applied to symmetric
domains by Koecher, Vinberg, and others. A Jordan algebra over a field F is
a nonassociative algebra over F with multiplication that is commutative and
satisfies the weakened associative law (ab)a2 = a(ba2 ). The basic principle is
that if C ⊂ V is a self-dual convex cone, then V can be given the structure of
a Jordan algebra in such a way that C becomes the set of squares in V .
276 28 Symmetric Spaces
D R C H R C H
Gn(D) SO(n) U(n) Sp(2n) − − −
Gn (D) U(n) U(n) × U(n) U(2n) GL(n, R) GL(n, C) GL(n, H)
Gn(D) Sp(2n) U(2n) SO(4n) Sp(2n, R) GU(n, n) SO(4n)∗
Fig. 28.2. The 3 × 3 square. Left: compact forms. Right: noncompact forms
Indeed, dim Gn (D) − dim Gn (D) is the dimension of the tube domain, and
this is twice the dimension dim G (D) − dim Gn (D) of the cone.
Although in presenting the 3 × 3 square—valid for all n—in Fig. 28.2 it
seems best to take the full unitary groups in the second rows and columns,
this does not work so well for the 4×4 magic square. Let us therefore note that
we can also use modified groups that we call Hn (D) ⊂ Hn (D) ⊂ Hn , which
are the derived groups of the Gn (D). We must modify (28.16) accordingly:
D R C H R C H
1
Hn(D) SO(n) SU(n) Sp(2n) 2n(n − 1) n2 − 1 n(2n + 1)
Hn (D) SU(n) SU(n) × SU(n) SU(2n) n2 − 1 2n2 − 2 4n2 − 1
Hn(D) Sp(2n) SU(2n) SO(4n) n(2n + 1) 4n2 − 1 2n(4n − 1)
Exercises
In the exercises, we look at the complex unit ball, which is a Hermitian symmetric
space that is not a tube domain. For these spaces, Piatetski-Shapiro [133] gave
unbounded realizations that are called Siegel domains of Type II . (Siegel domains
of Type I are tube domains over self-dual cones.)
D R C H O R C H O
H3(D) SO(3) SU(3) Sp(6) F4 3 8 21 52
H3 (D) SU(3) SU(3) × SU(3) SU(6) E6 8 16 35 78
H3(D) Sp(6) SU(6) SO(12) E7 21 35 66 133
H3(D) F4 E6 E 7 E8 52 78 133 248
Let
⎧ ⎛ ⎞ ⎫
⎪
⎨ w1 ⎪
⎬
⎜ ⎟
Bn = w = ⎝ ... ⎠ |w1 |2 + · · · + |wn |2 < 1
⎪
⎩ ⎪
⎭
wn
be the complex unit ball. Write
Ab
g= , A ∈ Matn (C), b ∈ Matn×1 (C), c ∈ Mat1,n (C), d ∈ C.
c d
Show that there are holomorphic maps c : Hn −→ Bn and c−1 : Bn −→ Hn that are
inverses of each other and are given by
⎛ −1 ⎞ ⎛ ⎞
√1 − i)(z1 + i)−1
(z √ + w1 )(1 − w1−1
i(1 )−1
⎜ 2iz2 (z1 + i) ⎟ ⎜ 2iw2 (1 − w1 ) ⎟
⎜ ⎟ ⎜ ⎟
c(z) = ⎜ .. ⎟, c−1 (w) = ⎜ .. ⎟.
⎝ . ⎠ ⎝ . ⎠
√ −1
√ −1
2izn (z1 + i) 2iwn (1 − w1 )
Note: If we extend the action (28.19) to allow g ∈ GL(n + 1, C), these “Cayley
transforms” are represented by the matrices
⎛ √ √ ⎞ ⎛ √ √ ⎞
1/ 2i −i/ 2i i/ 2i i/ 2i
c = ⎝ √ In−1 √
⎠, c−1 = ⎝ √ In−1 √
⎠.
1/ 2i i/ 2i −1/ 2i 1/ 2i
Exercise 28.3. Show that c−1 SU(n, 1)c = SUξ , where SUξ is the group of all g ∈
GL(n, C) satisfying g ξ t ḡ −1 = ξ, where
⎛ ⎞
−i
ξ = ⎝ In−1 ⎠.
i
280 28 Symmetric Spaces
Let us write
⎛ ⎞
z1 ⎛ ⎞
⎜ z2 ⎟ z2
⎜ ⎟ z1 ⎜ ⎟
z=⎜ . ⎟= , ζ = ⎝ ... ⎠ .
⎝ .. ⎠ ζ
zn
zn
According to (28.19), a typical element of H should act by
i 2
z1 −→ z1 + ibζ + |b| + ia,
2
ζ −→ ζ + b.
Check directly that H is invariant under such a transformation. Also show that SUξ
contains the group
⎧⎛ ⎞ ⎫
⎨ u ⎬
M= ⎝ h ⎠ u, v ∈ C× , h ∈ U(n − 1) .
⎩ ⎭
ū−1
Describe the action of this group. Show that the subgroup of SUξ generated by H
and M is transitive on Hn , and deduce that the action of SU(n, 1) on Bn is also
transitive.
Exercise 28.4. Observe that the subgroup K = S U(n) × U(1) of SU(n, 1) acts
transitively on the topological boundary of Bn , and explain why this shows that the
Bergman–Shilov boundary is the whole topological boundary. Contrast this with the
case of Dn .
Exercise 28.5. Emulate the construction of Rn and R◦n to show that the compact
dual of Bn has a dense open subset that can be identified with Cn in such a way that
GC = GL(n + 1, C) acts by (28.19). Show that Bn can be embedded in its compact
dual, just as Dn is in the case of the symplectic group.
29
Relative Root Systems
In this chapter, we will consider root systems and Weyl groups associated
with a Lie group G. We will assume that G satisfies the assumptions in Hy-
pothesis 28.1 of the last chapter. Thus, G is semisimple and comes with a
compact dual Gc . In Chap. 18, we associated with Gc a root system and Weyl
group. That root system and Weyl group we will call the absolute root system
Φ and Weyl group W . We will introduce another root system Φrel , called the
relative or restricted root system, and a Weyl group Wrel called the relative
Weyl group. The relation between the two root systems will be discussed. The
structures that we will find give Iwasawa and Bruhat decompositions in this
context. This chapter may be skipped with no loss of continuity.
As we saw in Theorem 28.3, every semisimple Lie group admits a Car-
tan decomposition, and Hypothesis 28.1 will be satisfied. The assumption of
semisimplicity can be relaxed—it is sufficient for G to be reductive, though in
this book we only define the term “reductive” when G is a complex analytic
group. A more significant generalization of the results of this chapter is that
relative and absolute root systems and Weyl groups can be defined whenever
G is a reductive algebraic group defined over a field F . If F is algebraically
closed, these coincide. If F = R, they coincide with the structures defined in
this chapter. But reductive groups over p-adic fields, number fields, or finite
fields have many applications, and this reason alone is enough to prefer an ap-
proach based on algebraic groups. For this, see Borel [20] as well as Borel and
Tits [21], Tits [162] (and other papers in the same volume), and Satake [144].
Consider, for example, the group G = SL(r, H), the construction of which
we recall. The group GL(r, H) is the group of units of the central simple algebra
Matr (H) over R. We have C ⊗ H ∼ = Mat2 (C) as C-algebras. Consequently,
C ⊗ Matr (H) ∼ = Mat2r (C). The reduced norm ν : Matr (H) −→ R is a map
determined by the commutativity of the diagram
Matr(H) Mat2r(C)
ν det
R C
Proposition 29.1. Assume that the assumptions of Hypothesis 28.1 are sat-
isfied. Then the map
(Z, k) −→ exp(Z)k (29.1)
is a diffeomorphism p × K −→ G.
Proposition 29.2. Assume that the assumptions of Hypothesis 28.1 are sat-
isfied. Let a be a maximal Abelian subspace of p. Then A = exp(a) is a closed
Lie subgroup of G, and a is its Lie algebra. There exists a θ-stable maximal
torus T of Gc such that A is contained in the complexification TC regarded as
a subgroup of GC . If r = dim(a), then A ∼
= (R× r
+ ) . Moreover, Ac = exp(ia) is
a compact torus contained in T . We have T = Ac TM , where TM = (T ∩ K)◦ .
is Abelian. Thus, both H and θ H are in the centralizer of t. Now we can write
H = H1 + H2 , where H1 = 12 (H + θ H) and H2 = 12 (H − θ H). Note that the
torus Si , which is the closure of {exp(tHi ) | t ∈ R}, is θ stable – indeed θ is
trivial on S1 and induces the automorphism x −→ x−1 on S2 . Also Si ⊆ T
centralizes T . Consequently, T Si is a θ-stable torus and, by maximality of T ,
Si ⊆ T . It follows that Hi ∈ t, and so H ∈ t. We have proven that t = t and
so T = T is a maximal torus.
It remains to be shown that T = Ac TM . It is sufficient to show that the Lie
algebra of T decomposes as ia ⊕ tM , where tM is the Lie algebra of TM . Since
θ stabilizes T , it induces an endomorphism of order 2 of t = Lie(T ). The +1
eigenspace is tM = t ∩ k since the +1 eigenspace of θ on gc is k. On the other
hand, the −1 eigenspace of θ on t contains ia and is contained in ip, which is
the −1 eigenspace of θ on gc . Since a is a maximal Abelian subspace of p, it
follows that the −1 eigenspace of θ on t is exactly ia, so t = ia ⊕ tM .
Then exp(Z) has the same form with λi replaced by exp(λi ). Since the λi are
distinct real numbers, the exp(λi ) are also distinct, and it follows that g has
the form ⎛ ⎞
g1
⎜ .. ⎟
⎝ . ⎠,
gh
where gi is an ri × ri block. Thus g commutes with Z.
The groups X ∗ (A) and X ∗ (Ac ) are isomorphic: extending a rational character
of A to a complex analytic character of AC and then restricting it to Ac gives
every character of Ac exactly once.
Proof. Obviously (29.2) is a rational character. Extending any rational char-
acter of A to an analytic character of AC and then restricting it to Ac
gives a homomorphism X ∗ (A) −→ X ∗ (Ac ), and since the characters of
X ∗ (Ac ) are classified by Proposition 15.4, we see that every rational char-
acter has the form (29.2) and that the homomorphism X ∗ (A) −→ X ∗ (Ac ) is
an isomorphism.
Since the compact tori T and Ac satisfy T ⊃ Ac , we may restrict characters
of T to Ac . Some characters may restrict trivially. In any case, if α ∈ X ∗ (T )
restricts to β ∈ X ∗ (A) = X ∗ (Ac ), we write α|β. Assuming that α and hence
β are not the trivial character, as in Chap. 18 we will denote by Xβ the
β-eigenspace of T on gC . We will also denote by Xrelα the α-eigenspace of Ac
on gC . Since X ∗ (Ac ) = X ∗ (A), we may write
286 29 Relative Root Systems
Xrel
α = {X ∈ gC | Ad(a)X = α(a)X for all a ∈ A}.
Let Φ be the set of β ∈ X ∗ (T ) such that Xβ
= 0, and let Φrel be the set of
α ∈ X ∗ (A) such that Xrelα
= 0.
If β ∈ X ∗ (T ), let dβ : t −→ C be the differential of β. Thus
d
dβ(H) = β(etH ) , H ∈ t.
dt t=0
As in Chap. 18, the linear form dβ is pure imaginary on the Lie algebra tM ⊕ia
of the compact torus T . This means that dβ is real on a and purely imaginary
on tM .
If α ∈ Φrel , the α-eigenspace Xrel
α can be characterized by either the con-
dition (for X ∈ Xrelα )
Ad(a)X = α(a)X, a ∈ A,
or
[H, X] = dα(H) X, H ∈ a. (29.3)
Let c : gC −→ gC denote the conjugation with respect to g. Thus, if
Z ∈ gC is written as X + iY , where X, Y ∈ g, then c(Z) = X − iY so
g = {Z ∈ gC | c(Z) = Z}. Let m be the Lie algebra of M . Thus, the Lie
algebra of CG (A) = M A is m ⊕ a. It is the 0-eigenspace of A on g, so
&
gC = C(m ⊕ a) ⊕ Xα (29.4)
α∈Φrel
applying c to (29.3) gives the same condition, with Z replaced by c(Z). Now
2i Z − c(Z) ,
1 1
2 Z + c(Z) ,
are in g, and at least one of them is not in the span of X1 , . . . , Xi since Z is not.
We may add this to the linearly independent set X1 , . . . , Xh , contradicting the
assumed maximality. This proves (i).
α to X−α . Indeed, if X ∈ Xα , then for
For (ii), let us show that θ maps Xrel rel rel
−1
a ∈ A we have Ad(a)X = α(a)Xα . Since θ(a) = a , replacing a by its inverse
and applying θ, it follows that Ad(a)θ(X) = α(a−1 ) θ(X). Since the group law
in X ∗ (A) is written additively, (−α)(a) = α(a−1 ). Therefore θ(X) ∈ X−α .
Since θ and c commute, if X ∈ g, then θ(X) ∈ g.
The last point we need to check for (ii) is that if 0
= X ∈ Xrel α ∩ g, then
[X, θ(X)]
= 0. Since Ad : Gc −→ GL(gc ) is a real representation of a compact
group, there exists a positive definite symmetric bilinear form B on gc that is
Gc -invariant. We extend B to a symmetric C-bilinear form B : gC × gC −→ C
by linearity. We note that Z = X + θ(X) ∈ k since θ(Z) = Z and Z ∈ g.
In particular Z ∈ gc . It cannot vanish since X and θ(X) lie in Xα and X−α ,
which have a trivial intersection. Therefore, B(Z, Z)
= 0. Choose H ∈ a such
that dα(H)
= 0. We have
B X + θ(X), [H, X − θ(X)] = B Z, dα(H)Z
= 0.
On the other hand, by (10.3) this equals
−B [X + θ(X), X − θ(X)], H = 2B [X, θ(X)], H .
Therefore, [X, θ(X)]
= 0.
For (ii), we will prove that Xrel
α is invariant under CG (A), which con-
tains M . Since g is obviously an Ad-invariant real subspace of gC it will follow
that Xrel
α ∩ g is Ad(M )-invariant. Since CG (A) is connected by Theorem 16.6,
it is sufficient to show that Xrel
α is invariant under ad(Z) when Z is in the Lie
algebra centralizer of a. Thus, if H ∈ a we have [H, Z] = 0. Now if X ∈ Xrel α
we have
[H, [Z, X]] = [[H, Z], X] + [Z, [H, X]] = [Z, dα(H)X] = dα(H)[Z, X].
Therefore, Ad(Z)X = [Z, X] ∈ Xrel α .
Part (iv) is entirely similar to Proposition 18.4 (ii), and we leave it to the
reader.
The roots in Φ can now be divided into two classes. First, there are those
that restrict nontrivially to A and hence correspond to roots in Φrel . On the
other hand, some roots do restrict trivially, and we will show that these cor-
respond to roots of the compact Lie group M . Let m = Lie(M ).
288 29 Relative Root Systems
and
[Y, Z] = 2iHα .
Now dα(Hα )
= 0. Indeed, if dα(Hα ) = 0, then ad(Z)2 Y = 0 while
ad(Z)Y
= 0, contradicting Lemma 18.1, since Z ∈ k. After replacing Xα by a
positive multiple, we may assume that dα(H) = 2.
Now at least we have a Lie algebra homomorphism sl(2, R) −→ g with the
effect (29.8), and we have to show that it is the differential of a Lie group
homomorphism SL(2, R) −→ G. We begin by constructing the corresponding
map SU(2) −→ Gc . Note that iHα , Y , and Z are all elements of gc , and so
we have a homomorphism su(2) −→ k that maps
i i 1
−→ iHα , −→ Y, −→ Z.
−i i −1
Theorem 29.1. In the context of Proposition 29.7, the set Φrel of restricted
roots is a root system. If α ∈ Φrel , there exists wα ∈ K that normalizes A and
that induces on X ∗ (A) the reflection sα .
Proof. Let
1
wα = iα .
−1
We note wα ∈ K. Indeed, it is the exponential of
290 29 Relative Root Systems
π 1 π
diα = Xα − X−α ∈ k
2 −1 2
since
1 cos(t) sin(t)
exp t = .
−1 − sin(t) cos(t)
in SL(2, R), and applying iα gives Ad(wα )Hα = −Hα . Since a is spanned by
the codimension 1 subspace aα and the vector Hα , it follows that (in its action
on Vrel ) wα has order 2 and eigenvalue −1 with multiplicity 1. It therefore
induces the reflection sα in its action on Vrel .
Now the proof that Φrel is a root system follows the structure of the proof
of Theorem 18.2. The existence of the simple reflection wα in the Weyl group
implies that sα preserves the set Φ.
For the proof that if α and β are in Φ then 2 α, β / α, α ∈ Z, we adapt
the proof of Proposition 18.10. If λ ∈ X ∗ (Ac ), we will denote (in this proof
only) by Xλ the λ-eigenspace of Ac in the adjoint representation. We normally
use this notation only if λ
= 0 is a root. If λ = 0, then Xλ is the complexified
Lie algebra of CG (A); that is, C(m ⊕ a). Let
&
W = Xβ+kα ⊆ XC .
k∈Z
We claim that W is invariant under iα SL(2, C)
. To prove this, it is sufficient
to show that it is invariant under diα sl(2, C) , which is generated by Xα and
X−α , since these are the images under iα of a pair of generators of sl(2, C)
by (29.8). These are the images of diα and iα , respectively. From (29.7), we see
that ad(Xα )Xγ ∈ Xγ+2α and ad(X−α )Xγ ∈ Xγ−2α , proving that iα (SL(2, C))
is invariant. In particular, W is invariant under wα ∈ SL(2, C). Since ad(wα )
induces sα on Vrel , it follows that the set {β + kα|k ∈ Z} is invariant under
sα and, by (18.1), this implies that 2 α, β / α, α ∈ Z.
it was assumed that the root system was reduced. Proposition 20.18 contains
all we need about nonreduced root systems.
The relationship between the three root systems Φ, ΦM , and Φrel can be
expressed in a “short exact sequence of root systems,”
0 −→ ΦM −→ Φ −→ Φrel −→ 0, (29.9)
We see that the restriction map induces a surjective mapping from the set
of simple roots in Φ that have nonzero restrictions to the simple roots in Φrel .
The last question that needs to be answered is when two simple roots of Φ
can have the same nonzero restriction to Φrel .
Proof. The fact that β and −θ(β) have the same restriction follows from
Proposition 29.5 (ii). It follows immediately that −θ(β) is a positive root
in Φ. The map β −→ −θ(β) permutes the positive roots, is additive, and
therefore preserves the simple positive roots.
Suppose that α is a simple root of Φrel and β, β are simple roots of Φrel
both restricting to α. Since β −β has trivial
restriction
to
Ac , it is θ-invariant.
Rewrite β − β = θ(β − β ) as β + − θ(β) = β + θ(−β ) . This expresses the
sum of two simple positive roots as the sum of another two simple positive
roots. Since the simple positive roots are linearly independent by Proposi-
tion 20.18, it follows that either β = β or β = −θ(β).
292 29 Relative Root Systems
0 ΦM Φ Φrel 0
A1 × A1 × A1 A5 A2
Fig. 29.1. The “short exact sequence of root systems” for SL(3, H)
since this symmetric matrix also has eigenvalues 1 with multiplicity n and −1
with multiplicity −1. Thus, if
⎛ √ √ ⎞
1/ 2 −1/ 2
u = ⎝ √ In−1 √
⎠,
1/ 2 1/ 2
Fig. 29.2. Satake diagrams for the rank 1 groups SO(n, 1) (a) SO(11, 1) (Type
DII) (b) SO(10, 1) (Type BII)
Fig. 29.3. Split and quasisplit even orthogonal groups (a) SO(6, 6) (Type DI, split)
(b) SO (7, 5) (Type DI, quasisplit)
For n = 5, the Lie algebra of SO(6, 4) is shown in Fig. 29.4. For n = 6, the
Satake diagram of SO(7, 5) is shown in Fig. 29.3.
The circling of the x45 and x46 positions in Fig. 29.4 is slightly misleading
because, as we will now explain, these do not correspond exactly to roots.
296 29 Relative Root Systems
Indeed, each of the circled coordinates x12 , x23 , and x34 corresponds to a
one-dimensional subspace of g spanning a space Xαi , where i = 1, 2, 3 are
the first three simple roots in Φ. In contrast, the root spaces Xα4 and Xα5
are divided between the x45 and x46 positions. To see this, the torus T in
Gc ⊂ GC consists of matrices
⎛ it1 ⎞
e
⎜ eit2 ⎟
⎜ eit3 ⎟
⎜ ⎟
⎜ it4 ⎟
⎜ e ⎟
⎜ cos(t5 ) sin(t5 ) ⎟
t=⎜ ⎜
⎟
⎟
⎜ − sin(t5 ) cos(t5 ) ⎟
⎜ e−it4 ⎟
⎜ −it3 ⎟
⎜ e ⎟
⎝ e −it2 ⎠
e−it1
and
α4 (t) = ei(t4 −t5 ) , α5 (t) = ei(t4 +t5 ) .
The eigenspaces Xα4 and Xα5 are spanned by Xα4 and Xα5 , where
⎛ ⎞
0 0 0 0 0 0 0 0
⎜ 0 0 0 0 0 0 0 0 ⎟
⎜ 0 0 0 1 i 0 0 0 ⎟
⎜ ⎟
⎜ 0 0 0 0 0 −1 0 0 ⎟
Xα4 = ⎜⎜ 0 0 0 0 0 −i 0
⎟,
⎜ 0 ⎟
⎟
⎜ 0 0 0 0 0 0 0 0 ⎟
⎝ ⎠
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
and its conjugate is Xα5 .
The involution θ is transpose-inverse. In its effect on the torus T , θ(t−1 )
does not change t1 , t2 , t3 , or t4 but sends t5 −→ −t5 . Therefore, −θ inter-
changes the simple roots α4 and α5 , as indicated in the Satake diagram in
Figs. 29.3 and 29.4.
As a last example, we look next at the Lie group SU(p, q), where p > q.
We will see that this has type BCq . Recall from Chap. 19 that the root system
of type BCq can be realized as all elements of the form
where ei are standard basis vectors of Rn . See Fig. 19.5 for the case q = 2.
We defined U(p, q) to be
g ∈ GL(p + q, C) | g J t ḡ = J ,
29 Relative Root Systems 297
where J = J1 , but (as with the group O(n, 1) discussed above) we could just
as well take J = J2 , where now
⎛ ⎞
Iq
Ip
J1 = , J2 = ⎝ Ip−q ⎠ .
−Iq
Iq
This has the advantage of making the group A diagonal. We can take A to be
the group of matrices of the form
⎛ ⎞
t1
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
⎜ tq ⎟
⎜ ⎟
⎜ Ip−q ⎟.
⎜ ⎟
⎜ −1
t1 ⎟
⎜ ⎟
⎜ .. ⎟
⎝ . ⎠
t−1
q
in such a way that the rational character ti corresponds to the standard basis
vector ei , we see that Φrel is a root system of type BCq . The Satake diagram
is illustrated in Fig. 29.5.
We turn now to the Iwasawa decomposition for G admitting a Cartan
decomposition as in Hypothesis 28.1. The construction is rather similar to
what we have already done in Chap. 26.
q nodes
q nodes
We have
0 = [H, Z] = [H, Z0 ] + [H, Zα ] = dα(H)Zα .
α∈Φrel α∈Φrel
Thus, the relative root system does not dependent in any essential way on the
choice of a. The argument is similar to the proof of Theorem 16.4.
Exercises
∼ Mat2n (C) as C-algebras and that the
Exercise 29.1. Show that C ⊗ Matn (H) =
composition
Matn (H) −→ C ⊗ Matn (H) ∼
det
= Mat2n (C) −→ C
takes values in R.
Exercise 29.2. Compute the Satake diagrams for SO(p, q) with p q for all p
and q.
Exercise 29.3. Prove an analog of Theorem 18.3 showing that Wrel is generated
by the reflections constructed in Theorem 29.1.
30
Embeddings of Lie Groups
In this chapter, we will contemplate how Lie groups embed in one another.
Our aim is not to be systematic or even completely precise but to give the
reader some tools for thinking about the relationships between different Lie
groups.
If G is a Lie group and H a subgroup, then there exists a chain of Lie
subgroups of G,
G = G0 ⊃ G1 ⊃ · · · ⊃ Gn = H
such that each Gi is maximal in Gi−1 . Dynkin [45–47] classified the maxi-
mal subgroups of semisimple complex analytic groups. Thus, the lattice of
semisimple complex analytic subgroups of such a group is known.
Let K1 and K2 be compact connected Lie groups, and let G1 and G2 be
their complexifications. Given an embedding K1 −→ K2 , there is a unique
analytic embedding G1 −→ G2 . The converse is also true: given an analytic
embedding G1 −→ G2 , then K1 embeds as a compact subgroup of G2 . How-
ever, any compact subgroup of G2 is conjugate to a subgroup of K2 (Theorem
28.2), so K1 is conjugate to a subgroup of K2 . Thus, embeddings of compact
connected Lie groups and analytic embeddings of their complexifications are
essentially the same thing. To be definite, let us specify that in this chapter
we are talking about analytic embeddings of complex analytic groups, with
the understanding that the ideas will be applicable in other contexts. By a
“torus,” we therefore mean a group analytically isomorphic to (C)n for some
n. We will allow ourselves to be a bit sloppy in this chapter, and we will
sometimes write O(n) when we should really write O(n, C).
So let us start with embeddings of complex analytic Lie groups. A useful
class of complex analytic groups that is slightly larger than the semisimple
ones is the class of reductive complex analytic groups. A complex analytic
group G (connected, let us assume) is called reductive if its linear analytic
representations are completely reducible. For example, GL(n, C) is reductive,
though it is not semisimple.
Examples of groups that are not reductive are parabolic subgroups. Let G
be the complexification of the compact connected Lie group K, and let B be
the Borel subgroup described in Theorem 26.2. A subgroup of G containing B
is called a standard parabolic subgroup. (Any conjugate of a standard parabolic
subgroup is called parabolic.)
As an example of a group that is not reductive, let P ⊂ GL(n, C) be the
maximal parabolic subgroup consisting of matrices
g1 ∗
, g1 ∈ GL(r, C), g2 ∈ GL(s, C), r + s = n.
g2
In the standard representation corresponding to the inclusion P −→ GL(n, C),
the set of matrices which have last s entries that are zero is a P -invariant sub-
space of Cn that has no invariant complement. Therefore, this representation
is not completely reducible, and so P is not reductive.
If G is the complexification of a connected compact group, then analytic
representations of G are completely reducible by Theorem 24.1. It turns out
that the converse is true—a connected complex analytic reductive group is the
complexification of a compact Lie group. We will not prove this, but it is useful
to bear in mind that whatever we prove for complexifications of connected
compact groups is applicable to the class of reductive complex analytic Lie
groups.
Even if we restrict ourselves to finding reductive subgroups of reductive
Lie groups, the problem is very difficult. After all, any faithful representation
gives an embedding of a Lie group in another. There is an important class of
embeddings for which it is possible to give a systematic discussion. Following
Dynkin, we call an embedding of Lie groups or Lie algebras regular if it takes
a maximal torus into a maximal torus and roots into roots. Our first aim is
to show how regular embeddings can be recognized using extended Dynkin
diagrams.
We will use orthogonal groups to illustrate some points. It is convenient
to take the orthogonal group in the form
⎛ ⎞
1
.
OJ (n, F ) = g ∈ GL(n, F ) | g J t g = J , J = ⎝ .. ⎠ .
1
We will take the realization OJ (n, C) ∩ U(n) = ∼ O(n) of the usual orthogonal
group in Exercise 5.3 with the maximal torus T consisting of diagonal ele-
ments of OJ (n, C) ∩ U(n). Then, as in Exercise 24.1, OJ (n, C) is the analytic
complexification of the usual orthogonal group O(n). We can take the ordering
of the roots so that the root eigenspaces Xα with α ∈ Φ+ are upper triangular.
We recall that the root system of type Dn is the root system for SO(2n).
Normally, one only considers Dn when n 4. The reason for this is that the
Lie groups SO(4) and SO(6) have root systems of types A1 × A1 and A3 ,
respectively. To see this, consider the Lie algebra of type SO(8). This consists
of the set of all matrices of the form in Fig. 30.1.
30 Embeddings of Lie Groups 305
D2 = A1 × A1 D3 = A3 D4
Remark 30.1. Although we could have worked with SL(2, C) × SL(2, C) at the
outset, over a field F that was not algebraically closed, it is often better to
use the realization G/Z Δ ∼ = SO(4). The reason is that if F is not algebraically
closed, the image of the homomorphism SL(2, F )×SL(2, F ) −→ SO(4, F ) may
not be all of SO(4). Identifying SL(2)×SL(2) with the algebraic group Spin(4),
this is a special instance of the fact that the covering map Spin(n) −→ SO(n),
though surjective over an algebraically closed field, is not generally surjective
on rational points over a field that is not algebraically closed. A surjective map
may instead be obtained by working with the group of similitudes GSpin(n),
which when n = 4 is the group G. This is analogous to the fact that the
homomorphism SL(2, F ) −→ PGL(2, F ) is not surjective if F is algebraically
closed, which is why the adjoint group PGL(2, F ) of SL(2) is constructed as
GL(2, F ) modulo the center, not SL(2) modulo the center.
∧2
GL(∧2 W ) −→ GL(∧4 W ) ∼
= C× .
(Each vi has to move past each wj producing rs sign changes.) Hence we may
regard ∧2 as a quadratic form on GL(∧2 W ). The subspace preserving the
determinant is therefore isomorphic to SO(6). The composite
2 2
∧ ∧
GL(W ) −→ GL(∧2 W ) −→ GL(∧4 W ) ∼
= C×
Proposition 30.1. Suppose in this setting that S is any set of roots such that
if α, β ∈ S and if α + β ⊂ Φ, then α + β ∈ S. Then
&
h = tC ⊕ Xα
α∈S
Proof. It is immediate from Proposition 18.4 (ii) and Proposition 18.3 (ii)
that this vector space is closed under the bracket.
We will not worry too much about verifying that h is the Lie algebra of a
closed Lie subgroup of G except to remark that we have some tools for this,
such as Theorem 14.3.
We have already introduced the Dynkin diagram in Chap. 25. We recall
that the Dynkin diagram is obtained as a graph whose vertices are in bijection
with Σ. Let us label Σ = {α1 , . . . , αr }, and let si = sαi . Let θ(αi , αj ) be the
angle between the roots αi and αj . Then
⎧
⎪
⎪ 2 if θ(αi , αj ) = π2 ,
⎪
⎪
⎪
⎪
⎪
⎪
⎨ 3 if θ(αi , αj ) = 2π 3 ,
n(si , sj ) =
⎪
⎪
⎪
⎪ 4 if θ(αi , αj ) = 3π
4 ,
⎪
⎪
⎪
⎪
⎩
6 if θ(αi , αj ) = 5π
6 .
The extended Dynkin diagram adjoins to the graph of the Dynkin diagram
one more node, which corresponds to the negative root α0 such that −α0
308 30 Embeddings of Lie Groups
is the highest weight vector in the adjoint representation. The negative root
α0 is sometimes called the affine root, because of its role in the affine root
system (Chap. 23). As in the usual Dynkin diagram, we connect the vertices
corresponding to αi and αj only if the roots are not orthogonal. If they make
an angle of 2π/3, we connect them with a single bond; if they make an angle
of 6π/4, we connect them with a double bond; and if they make an angle of
5π/6, we connect them with a triple bond.
The basic paradigm is that if we remove a node from the extended Dynkin
diagram, what remains will be the Dynkin diagram of a subgroup of G. To get
some feeling for why this is true, let us consider an example in the exceptional
group G2 . We may take S in Proposition 30.1 to be the set of six long roots.
These form a root system of type A2 , and h is the Lie algebra of a Lie subgroup
isomorphic to SL(3, C). Since SL(3, C) is the complexification of the simply-
connected compact Lie group SU(2), it follows from Theorem 14.3 that there
is a homomorphism SL(3, C) −→ G.
α2
α1
α0
The ordinary Dynkin diagram of G2 does not reflect the existence of this
embedding. However, from Fig. 30.3, we see that the roots α2 and α0 can be
taken as the simple roots of SL(3, C). The embedding SL(3, C) can be under-
stood as an inclusion of the A2 (ordinary) Dynkin diagram in the extended
G2 Dynkin diagram (Fig. 30.4).
α0
α1 α2 α3 αn−2 αn−1 αn
α0 αn−1
α1 α2 α3 αn−3 αn−2
αn
which is in O(2n + 2) but not SO(2n + 2). The fixed subgroup of this outer
automorphism stabilizes the vector v0 = t (0, . . . , 0, 1, −1, 0, . . . , 0). This vector
is not isotropic (that is, it does not have length zero) so the stabilizer is the
group SO(2n + 1) fixing the 2n + 1-dimensional orthogonal complement of
v0 . In this embedding SO(2n + 1) −→ SO(2n + 1), the short simple root of
SO(2n + 1) is embedded into the direct sum of Xαn and Xαn+1 . We invite the
reader to confirm this for the embedding of SO(9) −→ SO(10) with the above
matrices. We envision the Dn+1 Dynkin diagram being folded into the Bn
diagram, as in Fig. 30.12.
The Dynkin diagram of type D4 admits a rare symmetry of order 3
(Fig. 30.13). This is associated with a phenomenon known as triality, which
we now discuss.
Referring to Fig. 30.1, the groups Xαi (i = 1, 2, 3, 4) correspond to x12 , x23 ,
x34 and x35 , respectively. The Lie algebra will thus have an automorphism
τ that sends x12 −→ x34 −→ x35 −→ x12 and fixes x23 . Let us consider the
effect on tC , which is the subalgebra of elements t with all xij = 0. Noting
that dα1 (t) = t1 − t2 , dα2 (t) = t2 − t3 , dα3 (t) = t3 − t4 , and dα4 (t) = t3 + t4 ,
we must have ⎧
⎪ t1 − t2 −→ t3 − t4
⎪
⎨
t2 − t3 −→ t2 − t3
τ: ,
⎪
⎪ t3 − t4 −→ t3 + t4
⎩
t3 + t4 −→ t1 − t2
from which we deduce that
30 Embeddings of Lie Groups 311
α0 α1 α2 α3 αn−2 αn−1 αn
α0
α1 α2 α3 αn−2 αn−1
αn
α1 α2
α0 Left: G2, F4, E6.
Right: E7, E8. α2
α1 α2 α3 α4 α0
α0 α1 α3 α4 α5 α6 α7
α0
α2 α2
α0
α1 α3 α4 α5 α6 α1 α3 α4 α5 α6 α7 α8
τ (t1 ) = 12 (t1 + t2 + t3 − t4 ) ,
τ (t2 ) = 12 (t1 + t2 − t3 + t4 ) ,
τ (t3 ) = 12 (t1 − t2 + t3 + t4 ) ,
τ (t4 ) = 12 (t1 − t2 − t3 − t4 ) .
312 30 Embeddings of Lie Groups
αn
α1 α2 α3 αn−2 αn−1
αn+1
α1 α2 α3 αn−2 αn−1 αn
α4
α2
α1
α3
where B B
t1 = t1 t2 t3 t−1
4 , t2 = t1 t2 t−1
3 t4 ,
B B
t3 = t1 t−1
2 t3 t4 , t4 = t1 t−1 −1 −1
2 t3 t4 .
Due to the ambiguity of the square roots, this is not a univalent map.
The explanation is that since SO(8) is not simply-connected, a Lie alge-
bra automorphism cannot necessarily be lifted to the group. However, there
is automatically induced an automorphism τ of the simply-connected double
cover Spin(8). The center of Spin(8) is (Z/2Z) × (Z/2Z), which has an auto-
morphism of order 3 that does not preserve the kernel (of order 2) of τ . If we
divide Spin(8) by its entire center (Z/2Z) × (Z/2Z), we obtain the adjoint
group PGO(8), and the triality automorphism of Spin(8) induces an auto-
morphism of order 3 of PGO(8). To summarize, triality is an automorphism
of order 3 of either Spin(8) or PGO(8) but not of SO(8).
30 Embeddings of Lie Groups 313
The linear transformations f2 and f3 are only determined up to sign. The maps
f1 −→ f2 and f1 −→ f3 , though thus not well-defined as an automorphisms
of SO(8), do lift to well-defined automorphisms of Spin(8), and the resulting
automorphism f1 −→ f2 is the triality automorphism. Triality permutes the
three orthogonal maps f1 , f2 , and f3 cyclicly. Note that if f1 = f2 = f3 ,
then f1 is an automorphism of the octonion ring, so the fixed group G2 is
the automorphism group of the octonions. See Chevalley [36], p.188. As an
alternative to Chevalley’s approach, one may first prove a local form of triality
as in Jacobson [88] and then deduce the global form. See also Schafer [146].
Over an algebraically closed field, the octonion algebra is unique. Over the
real numbers there are two forms, which correspond to the compact group
O(8) and the split form O(4, 4).
So far, the examples we have given of folding correspond to automorphisms
of the group G. For an example that does not, consider the embedding of G2
into Spin(7) (Fig. 30.14, right).
α1 α1
α2 α2
α3
α4 α3
α2 α1 α2 α1
can be found in Table 28.1, for in this list, the compact subgroup K is the
fixed point of an involution in the compact group Gc , and this relationship is
also true for the complexifications. For example, the first entry, corresponding
to Cartan’s classification AI, is the symmetric space with Gc = SU(n) and the
subgroup K = SO(n). Assuming that we use the version of the orthogonal
group in Exercise 5.3, the involution θ is g → J t g −1 J, where J is given
by (5.3). This involution extends to the complexification SL(n, C), and the
fixed point set is the subgroup SO(n, C). If n is odd, then every simple root
eigenspace of SO(n, C) embeds in the direct sum of one or two simple root
eigenspaces of SL(2, C), and the embedding may be understood as an example
of the root folding paradigm. But if n = 2r is even, then one of the roots of
SO(2r), namely the simple root er−1 + er , involves non-simple roots of SL(n).
Suppose that V1 and V2 are quadratic spaces (that is, vector spaces
equipped with nondegenerate symmetric bilinear forms). Then V1 ⊕ V2 is
naturally a quadratic space, so we have an embedding O(V1 ) × O(V2 ) −→
O(V1 ⊕ V2 ). The same is true if V1 and V2 are symplectic (that is, equipped
with nondegenerate skew-symmetric bilinear forms). It follows that we have
embeddings
O(r) × O(s) −→ O(r + s), Sp(2r) × Sp(2s) −→ Sp 2(r + s) .
including a local theory. This is a topic that transcends Lie theory since in
much of the literature one will consider O(s) or Sp(2r) as algebraic groups
defined over a p-adic field or a number field (and its adele ring). Expositions of
pure Lie group applications may be found in Howe and Tan [78] and Goodman
and Wallach [56].
The classification of dual reductive pairs in Sp(2n), described in Weil [173]
and Howe [73], has its origins in the theory of algebras with involutions, due
to Albert [5]. The connection between algebras with involutions and the the-
ory of algebraic groups was emphasized earlier by Weil [171]. A modern and
immensely valuable treatise on algebras with involutions and their relations
with the theory of algebraic groups may be found in Knus, Merkurjev, Rost,
and Tignol [107].
A classification of dual reductive pairs in exceptional groups is in Ruben-
thaler [138]. These examples have proved interesting in the theory of auto-
morphic forms since an analog of the Weil representation is available.
So far, our point of view has been to start with a group G and understand
its large subgroups H, and we have a set of examples sufficient for understand-
ing most, but not all such pairs. Let us consider the alternative question: given
H, how can we embed it in a larger group G?
Suppose, therefore that π : H → GL(V ) is a representation. We assume
that it is faithful and irreducible. Then we get an embedding of H into GL(V ).
However sometimes there is a smaller subgroup G ⊂ GL(V ) such that the
image of π is contained in G. A frequent case is that G is an orthogonal or
symplectic group. These cases may be classified by considering the theory
of the Frobenius-Schur indicator, which is discussed in the exercises to this
chapter and again in Chap. 43. The Frobenius-Schur indicator (π) is the
multiplicity of the trivial character in the generalized character g → χ(g 2 ),
where χ is the character of π. It equals 0 unless π = π̂ is self-contragredient,
in which case either it equals 1 and π is orthogonal, or −1 and π is symplectic.
This means that if (π) = 1, then we may take G = O(n) where n = dim(V ),
while if (π) = −1, then dim(V ) is even, and we may take G = Sp(n).
The examples (30.1) can be understood this way. Here’s a couple more.
Let H = SL(2), and let π be the symmetric k-th power representation. The
vector space V is k + 1-dimensional. Exercise 22.15 computes the Frobenius-
Schur indicator, and we see that H embeds in SO(k + 1) if k is even, and
Sp(k + 1) if k is odd. For another example, if H is any simple Lie group,
then the adjoint representation is orthogonal since the Killing form on the Lie
algebra is a nondegenerate symmetric bilinear form. Thus for example we get
an embedding of SL(3) into SO(8).
As a final topic, we discuss parabolic subgroups. Just as regular subgroups
of G can be read off from the extended Dynkin diagram, the parabolic sub-
groups can be read off from the regular Dynkin diagram. Let Σ ⊂ Σ be any
proper subset of the set of simple roots. Then Σ is the set of vertices of a
(possibly disconnected) Dynkin diagram D contained in that of G. There will
316 30 Embeddings of Lie Groups
x61 x62 x63 x64 x65 −t5 −x45 −x35 −x25 −x15
x71 x72 x73 x74 x64 −x54 −t4 −x34 −x24 −x14
x81 x82 x83 x73 x63 −x53 −x43 −t3 −x23 −x13
x91 x92 x82 x72 x62 −x52 −x42 −x32 −t2 −x12
x01 x91 x81 x71 x61 −x51 −x41 −x31 −x21 −t1
x61 x62 x63 x64 x65 −t5 −x45 −x35 −x25 −x15
x71 x72 x73 x74 x64 −x54 −t4 −x34 −x24 −x14
x81 x82 x83 x73 x63 −x53 −x43 −t3 −x23 −x13
x91 x92 x82 x72 x62 −x52 −x42 −x32 −t2 −x12
x01 x91 x81 x71 x61 −x51 −x41 −x31 −x21 −t1
Exercises
Exercise 30.1. Discuss as many as possible of the embeddings K−→Gc in Table
28.1 of Chap. 28 using the extended Dynkin diagram of Gc .
Exercise 30.2. In doing the last exercise, one case you may have trouble with is the
embedding of S(O(p) × O(q)) into SO(p + q) when p and q are both odd. To get some
318 30 Embeddings of Lie Groups
insight, consider the embedding of SO(5)×SO(5) into SO(10). (Note: S O(p)×O(q)
is the group of elements of determinant 1 in O(p) × O(q) and contains SO(p) × SO(q)
as a subgroup of index 2. For this exercise,
it does not matter whether you work
with SO(5) × SO(5) or S O(5) × O(5) .) Take the form of SO(10) in Fig. 30.8.
This stabilizes the quadratic form x1 x10 + x2 x9 + x3 x8 + x4 x7 + x5 x6 . Consider the
subspaces ⎧⎛ ⎞⎫ ⎧⎛ ⎞⎫
⎪ a ⎪
⎪ ⎪ ⎪ 0 ⎪
⎪ ⎪
⎪
⎪ ⎜ ⎟ ⎪
⎪ ⎪
⎪ ⎜ 0 ⎟⎪ ⎪
⎪
⎪ ⎜ b ⎟⎪⎪ ⎪
⎪ ⎜ ⎟⎪ ⎪
⎪
⎪ ⎜ 0 ⎟⎪ ⎪ ⎪
⎪ ⎜ t ⎟⎪ ⎪
⎪
⎪ ⎜ ⎟ ⎪
⎪ ⎪
⎪ ⎪
⎪⎜
⎪ ⎟⎪⎪ ⎪
⎜
⎪⎜ ⎟⎪ ⎟ ⎪
⎪
⎪
⎪ 0 ⎪ ⎪ u ⎟⎪
⎨⎜ ⎟⎪
⎜ c ⎟⎬ ⎨⎜
⎪
⎜ v ⎟⎬
⎪
V1 = ⎜ ⎟ ,
⎜ −c ⎟⎪ V 2 = ⎜ ⎟ .
⎜ v ⎟⎪
⎪
⎪ ⎜ ⎟⎪ ⎪
⎪ ⎜ ⎟⎪
⎪
⎪ ⎜ 0 ⎟⎪ ⎪ ⎪
⎪ ⎜ ⎟⎪ ⎪
⎪
⎪⎜ ⎟⎪⎪
⎪ ⎪⎜ w ⎟⎪
⎪ ⎪
⎪
⎪ ⎜ ⎟ ⎪ ⎪
⎪ ⎜ ⎟ ⎪
⎪
⎪
⎪ ⎜ 0 ⎟⎪ ⎪ ⎪
⎪ ⎜ x ⎟⎪ ⎪
⎪
⎪ ⎝ d ⎠⎪ ⎪ ⎪
⎪ ⎝ 0 ⎠⎪ ⎪
⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪
⎩ ⎭ ⎩ ⎭
e 0
Observe that these five-dimensional spaces are mutually orthogonal and that the
restriction of the quadratic form is nondegenerate, so the stabilizers of these two
spaces are mutually centralizing copies of SO(5). Compute the Lie algebras of these
two subgroups, and describe how the roots of SO(10) restrict to SO(5) × SO(5).
Exercise 30.3. Let G be a semisimple Lie group. Assume that the Dynkin diagram
of G has no automorphisms. Show that every representation is self-contragredient.
Exercise 30.4. Let 1 and 2 be the fundamental dominant weights for Spin(5),
so that 2 is the highest weight of the spin representation. Show that the irreducible
representation with highest weight k1 +l2 is orthogonal if l is even, and symplectic
of l is odd.
Exercise 30.5. The group Spin(8) has three distinct irreducible eight-dimensional
representations, namely the standard representation of SO(8) and the two spin rep-
resentations. Show that these are permuted cyclicly by the triality automorphism.
Exercise 30.6. Prove that if G is semisimple and its Dynkin diagram has no aut-
omorphisms, then every element in G is conjugate to its inverse. Is the converse
true?
31
Spin
This chapter does not depend on the last few chapters, and may be read at
any point after Chap. 23, or even earlier. The results of Chap. 23 are not used
here, but are illustrated by the results of this chapter.
We will take a closer look at the groups SO(N ) and their double cov-
ers, Spin(N ). We assume that N 3 and that N = 2n + 1 or 2n. In this
Chapter, we will take a closer look at the groups SO(N ) and their double
covers, Spin(N ). These groups have remarkable “spin” representations of dim-
ension 2n , where N = 2n or 2n + 1. We will first show that this follows from
the Weyl theorem of Chap. 22. We will then take a different point of view
and give a different construction, using Clifford algebras and a uniqueness
principle.
The group Spin(N ) was constructed at the end
of Chap. 13 as the universal
cover of SO(N ). Since we proved that π1 SO(N ) ∼ = Z/2Z, it is a double cover.
In this chapter, we will construct and study the interesting and important spin
representations of the group Spin(N ). We will also show how to compute the
center of Spin(N ).
Let G = SO(N ) and let G̃ = Spin(N ). We will take G in the realization
of Exercise 5.3; that is, as the group of unitary matrices satisfying g J t g = J,
where J is (5.3). Let p : G̃ −→ G be the covering map.
Let
T be the diagonal
torus in G, and let T̃ = p−1 (T ). Thus ker(p) ∼= π1 SO(N ) ∼ = Z/2Z.
ti −→ t±1
i or ei −→ ±ei .
Under this embedding, the Weyl group Sn of U(n) gets embedded in the Weyl
group of O(N ). In its action on the torus, the ti are simply permuted, and in
the action on X ∗ (T ), the ei are permuted. The simple i-th simple reflection
in Sn has the sends αi to its negative (1 i n − 1) while permuting the
other positive roots, so it coincides with the i-th simple reflection in SO(N ).
Now let us consider the simple reflection with respect to αn . If N = 2n+ 1,
then since αn = en this just has the effect en −→ −en , and all other ei −→ ei .
A representative in N (T ) can be taken to be
⎛ ⎞
In−1
⎜ 0 0 1 ⎟
⎜ ⎟
wn = ⎜⎜ 0 −1 0 ⎟.
⎟
⎝ 1 0 0 ⎠
In−1
It is clear that all elements of the group H described in the statement of the
theorem that change the sign of exactly one ei can be generated by conjugating
322 31 Spin
1 = e1 ,
2 = e1 + e2 ,
..
.
n−1 = e1 + e2 + . . . + en−1 ,
n = 12 (e1 + e2 + . . . + en−1 + en ).
On the other hand, if N = 2n is even, the last two are a little changed. In this
case, the fundamental weights are
1 = e1 ,
2 = e1 + e2 ,
..
.
n−1 = 12 (e1 + e2 + . . . + en−1 − en ),
n = 12 (e1 + e2 + . . . + en−1 + en ).
Of course, to check the correctness of these weights, what one must check is
that 2 i , αj / αj , αj = 1 if i = j, and 0 if i
= j, and this is easily done.
We say that a weight is integral if it is in X ∗ (T ) and half-integral if it
is not. Thus a weight is integral if it is of the form (31.4) with the ci even,
and half-integral if they are odd. Dominant integral weights, of course, are
highest weight vectors of representations of SO(N ). By Proposition 31.2, the
dominant half-integral weights are highest weight vectors of representations
of Spin(N ). They are not highest weight vectors of representations of SO(N ).
If N = 2n+ 1, we see that just the last fundamental weight is half-integral,
but if N = 2n, the last two fundamental weights are half-integral. The repre-
sentations with highest weight vectors n (when N = 2n + 1) or n−1 and
n (when N = 2n) are called the spin representations.
(ii) If N = 2n, the dimensions of the spin representations π(n−1 ) and π(n )
are each 2n−1 . The weights that occur with nonzero multiplicity in this
representation all occur with multiplicity one; they are
1
2 (±e1 ± e2 ± . . . ± en ),
where the number of minus signs is odd for π(n−1 ) and even for π(n ).
Now let us consider the spin representations from a different point of view.
If V is a complex vector space of dimension N with a nondegenerate quadratic
form q, and if W ⊂ V is a maximal subspace on which q restricts to zero, we
will call W Lagrangian. We will see that the dimension of such a Lagrangian
(
subspace W is n where N = 2n or 2n + 1, so the exterior algebra W has
dimension 2n , and we will construct the spin representation on this vector
space.
To construct the spin representation, we will make use of the properties
of Clifford algebras. We digress to develop what we need. For more about
Clifford algebras, see Artin [9], Chevalley [36], Goodman and Wallach [56],
and Lawson and Michelsohn [118].
By a Z/2Z-graded algebra we mean an F -algebra A that decomposes into
a direct sum A0 ⊕ A1 , where A0 is a subalgbra, with Ai · Aj ⊆ Ai+j where i + j
is modulo 2. We require that F be contained in A0 and that F is central in A,
but A0 may be strictly larger than F . An element a of A is called homogeneous
if it is in Ai with i ∈ {0, 1} and then we call deg(a) = i the degree of a.
If A and B are Z/2Z-graded algebras, then we may define a Z/2Z-graded
algebra A ⊗ B. As a vector space, this is the usual tensor product of A and
B, with the following Z/2Z grading:
(A ⊗ B)0 = A0 ⊗ B0 ⊕ A1 ⊗ B1 , (A ⊗ B)1 = A1 ⊗ B0 ⊕ A0 ⊗ B1 .
Proof. We may identify D(A) with A⊕A as a vector space in which a⊗1+b⊗ζ
is identified with the ordered pair (a, b). In view of (31.5) the multiplication
is
(a, b)(c, d) = (ac − bd, ad + bc).
326 31 Spin
Now D(A)0 consists of pairs (a, b) with a of degree zero and b of degree 1,
and for this subring, the multiplication is
In order to construct the Clifford algebra, we may take the tensor algebra
T (V ) modulo the ideal I = IV generated by elements of the form x ⊗ y + y ⊗
x − 2B(x, y) with x, y ∈ V .
Proof. Let v be any vector with q(v)
= 0, and let w be any nonzero vector
in the orthogonal complement of v. Then q(w)
= 0 also since V otherwise
it is in the kernel of the associated symmetric bilinear form B, but B is
nondegenerate. Let q(v) = a2 and q(w) = b2 . Then x = bv − aw and y =
bv + aw are linearly independent isotropic vectors since a, b
= 0. Clearly the
space V0 spanned by x and y is a hyperbolic plane, and we may take V to be
its orthogonal complement.
ω(x)ξ = x ∧ ξ, x ∈ W, (31.6)
k
ω(y)ξ = 2B(y, wi )(−1)i+1 w1 ∧ . . . ∧ w
Ci ∧ . . . ∧ wk , y ∈ W . (31.7)
i=1
Proof. We can define ω by the (31.6) and (31.7) and (if N is odd) the req-
uirement that ω(z)ξ = (−1)i ξ. Regarding (31.7) this is well-defined by the
universal property of the exterior power because it is easy to check that the
right-hand side is multiplied by −1 if wi and wi+1 are interchanged, so it is
alternating. We have to check that ω is an algebra homomorphism.
We will show that if x, y ∈ V then
Exercises
Exercise 31.1. Check the details in the proof of Theorem 31.2. That is, verify that
S(λ) is exactly the set of characters stated in the theorem and that it consists of
just the W orbit of λ.
Exercise 31.2. Prove that the restriction of the spin representation of Spin(2n + 1)
to Spin(2n) is the sum of the two spin representations of Spin(2n).
Exercise 31.3. Prove that the restriction of either spin representation of Spin(2n)
to Spin(2n − 1) is the spin representation of Spin(2n).
Exercise 31.4. Show that one of the spin representations of Spin(6) gives an iso-
morphism Spin(6) ∼ = SU(4). What is the significance of the fact that there are two
spin representations?
Exercise 31.6. Let G be a compact connected Lie group whose root system is of
type G2 . (See Fig. 19.6.) Prove that G is simply-connected.
Part IV
Before showing how Mackey theory can be used to prove (32.2), let us observe
that this implies the proposition. First, if χ1
= χ2 , then it shows that
HomG (Vχ , Vχ ) is one-dimensional, so Vχ is irreducible. Moreover, (32.2) shows
exactly when there is a nonzero intertwining operator Vχ −→ Vμ , and the
second statement is easily deduced.
To prove (32.2), we make use of Mackey’s theorem, which we will prove
later in the chapter. We recall that if f1 and f2 are functions on G, their
convolution is the function
(f1 ∗ f2 )(g) = f1 (gh) f2 (h−1 ) = f1 (h) f2 (h−1 g).
h∈G h∈G
Mackey’s theorem (Theorem 32.1 below) asserts that any intertwining operator
T : Vχ −→ Vμ is of the form T f = Δ ∗ f where Δ : G −→ C is a function
satisfying
Δ(b2 g b1 ) = μ(b2 )Δ(g)χ(b1 ).
Such a function is determined by its values on a set of representatives for
the double cosets B\G/B. By the Bruhat decomposition, there are just two
double cosets:
1
G = B 1B ∪ Bw0 B, w0 = ,
1
where 1 is, of course, the identity matrix. [A quick proof is given below (27.1).]
32 Mackey Theory 339
that is,
Δ(1) = μ1 (t1 )μ2 (t2 )χ1 (t1 )−1 χ2 (t2 )−1 Δ(1).
Unless χ1 = μ1 and χ2 = μ2 , we may choose t1 and t2 so that
HomG (U, V G ) ∼
= HomH (UH , V ). (32.3)
HomG (V G , U ) ∼
= HomH (V, U ). (32.5)
This is slightly more difficult than Proposition 32.2, and it also involves ideas
that we will need in our discussion of Mackey theory. We will approach this
by means of a universal property.
Thus is H-equivariant.
32 Mackey Theory 341
Using (32.7), each term on the right-hand side of (32.8) is independent of the
choice of representatives γ of the cosets in G/H. Let us apply the right-hand
side to g ∈ G. We get
f (γ −1 ) (gγ).
G/H
The variable change γ −→ gγ permutes the cosets in G/H and shows that
J π G (g)f = σ(gγ)j f (γ −1 ) = σ(g)J(f ),
γ∈G/H
as required.
Proof. Let Δ satisfying (32.9) be given. It is easy to check, using (32.9) and
the fact that f ∈ V1G , that (32.10) is independent of the choice of coset
representatives γ for G/H1 . Moreover, if h2 ∈ H2 , then the variable change
γ −→ h2 γ permutes the cosets of G/H1 , and again using (32.9), this variable
change shows that Δ∗f ∈ V2G . Thus f −→ Δ∗f is a well-defined map V1G −→
V2G , and using the fact that G acts on both these spaces by right translation,
it is straightforward to see that Λ(f ) = Δ ∗ f defines an intertwining operator
V1G −→ V2G .
To show that this map Δ → Λ is an isomorphism of the space of Δ satisfy-
ing (32.9) to HomG (V1G , V2G ), we make use of Corollary 32.1. We must relate
the space of Δ satisfying (32.9) to HomH1 (V1 , V2G ). Given λ ∈ HomH1 (V1 , V2G )
corresponding to Λ ∈ HomG (V1G , V2G ) as in that corollary, define Δ : G −→
Hom(V1 , V2 ) by Δ(g)v1 = λ(v1 )(g). The condition that λ(v1 ) ∈ V2G for all
v1 ∈ V1 is equivalent to
Δ(h2 g) = π2 (h2 ) ◦ Δ(g), h 2 ∈ H2 ,
and the condition that λ : V1 −→ V2G is H1 -equivariant is equivalent to
Δ(gh1 ) = Δ(g) ◦ π1 (h1 ), h 1 ∈ H1 .
Of course, these two properties together are equivalent to (32.9). We see that
Corollary 32.1 implies a linear isomorphism between the space of functions Δ
satisfying (32.9) and the elements of HomG (V1G , V2G ). We have only to show
that this correspondence is given by (32.10). In (32.6), we take H = H1 ,
(σ, U ) = (π2G , V2G ), and j = λ. Then J = Λ and (32.6) gives us, for f ∈ V1G ,
Λ(f ) = π2G (γ)λ f (γ −1 ) .
γ∈G/H1
Applying this to g ∈ G,
Λ(f )(g) = λ f (γ −1 ) (gγ) = Δ(gγ)f (γ −1 ).
γ∈G/H1 γ∈G/H1
Remark 32.1. Although we are working here with finite groups, Mackey’s
theorem is (since Bruhat [26]) a standard tool in representation theory of
Lie groups also. The function Δ becomes a distribution.
Remark 32.2. Suppose that H1 , H2 , and (πi , Vi ) are as in Theorem 32.1. The
function Δ : G −→ Hom(V1 , V2 ) associated with an intertwining operator
Λ : V1G −→ V2G is clearly determined by its values on a set of representatives
for the double cosets in H2 \G/H1 . The simplest case is when Δ is supported
on a single double coset H2 γH1 . In this case, we say that the intertwining
operator Λ is supported on H2 γH1 .
h
dim HomG (V1G , V2G ) = dim HomHγi (π1γi , π2γi ). (32.12)
i=1
Proof. If Δ is as in Theorem 32.1, write Δ = i Δi , where
Δ(g) if g ∈ H2 γi H1 ,
Δi (g) =
0 otherwise.
Then Δi satisfy (32.9). Let Λi be the intertwining operator. Then Λi is sup-
ported on a single double coset, and the dimension of the space of such inter-
twining operators is computed in Proposition 32.4.
344 32 Mackey Theory
Corollary 32.2. Assume that the ground field F is of characteristic zero. Let
H1 and H2 be subgroups of G and let (π, V ) be an irreducible representation
of H1 . Let γ1 , . . . , γh be a complete set of representatives of the double cosets
in H2 \G/H1 . If γ ∈ G, let Hγ = H2 ∩ γH1 γ −1 , and let π γ : Hγ −→ GL(V )
be the representation π γ (g) = π(γ −1 gγ). Then the restriction of π G to H2 is
isomorphic to
&h
IndH 2 γi
Hγ (π ). (32.13)
i
i=1
In a word, first inducing and then restricting gives the same result as restrict-
ing, then inducing. This way of explaining the result is a pithy oversimplifi-
cation that has to be correctly understood. More precisely, there are different
ways we can restrict, namely given γ we may restrict to Hγ , then induce; we
have to sum over all these different ways. And the different ways depend only
on the double coset H2 γH1 .
h
dim HomH2 (V G , V2 ) = dim HomG (V G , V2G ) = dim HomHγi (π γi , π2γi )
i=1
h
dim HomH2 IndH2 γi
Hγ (π ), π2 .
i
i=1
a(w ⊗ v) = aw ⊗ v, a ∈ A.
HomR (V, U ) ∼
= HomA (A ⊗R V, U ). (32.14)
Finally, if F = C, let us recall the formula for the character of the induced
representation. If χ is a class function of the subgroup H of G, let χ̇ : G −→ C
be the function
χ(g) if g ∈ H ,
χ̇(g) =
0 otherwise,
and let χG : G −→ C be the function
χG (g) = χ̇(xgx−1 ). (32.15)
x∈H\G
1 1
χ(h) η(g).
|G| |H|
g∈G h∈H x∈G
xgx−1 = h
Given h ∈ H, we can enumerate the pairs (g, x) ∈ G×G that satisfy xgx−1 = h
by noting that they are the pairs (x−1 hx, x) with x ∈ G. So the sum equals
32 Mackey Theory 347
1 1 1
χ(h) η(x−1 hx) = χ(h) η(h) = χ, ηH
|G| |H| |H|
h∈H x∈G h∈H
−1
since η(x hx) = η(h).
Exercises
Exercise 32.1. Some points in the proof of Proposition 32.2 were left to the reader.
Write out a complete proof.
Exercise 32.2. Let H1 , H2 , and H3 be subgroups of G, with (πi , Vi ) a repre-
sentation of Hi . Let there be given intertwining operators Λ1 : V1G → V2G and
Λ2 : V2G → V3G . Let Δ1 : G → Hom(V1 , V2 ) and Δ2 : G → Hom(V2 , V3 ) being
the corresponding functions as in Theorem 32.1. Express the Δ : G → Hom(V1 , V3 )
corresponding to the composition Λ2 ◦ Λ1 in terms of Δ1 and Δ2 .
Exercise 32.3. Let H be a subgroup of G, and ψ : H → C× a linear character.
Prove that the ring of G-module endomorphisms of the induced representation ψ G
is isomorphic to the convolution ring of functions Δ : G → C× such that
Δ(h2 gh1 ) = ψ(h2 ) Δ(g) ψ(h1 ), h1 , h2 ∈ H.
G
What can you say about ψ if this ring is commutative?
Exercise 32.4. Let G = GL(2, F ), where F is a finite field. Let B be the Borel
subgroup of upper triangular matrices, and let N be its subgroup of unipotent
matrices. Let ψF : F → C be any nontrivial character. Define a character of N as
follows:
1x
ψ = ψF (x).
1
Let χ be a linear character of B as in (32.1). Show that up to scalar multiple there
is a unique intertwining operator IndGB (χ) → IndN (ψ).
G
If k > n, then ek = 0, although this is not true for hk . Our convention is that
e0 = h0 = 1.
Let E(t) be the generating function for the elementary symmetric
polynomials:
n
E(t) = e k tk .
k=0
Then
E(t) = (1 + x1 t)(1 + x2 t) · · · (1 + xn t) (33.1)
since expanding the right-hand side and collecting the coefficients of tk will
give each monomial in the definition of ek exactly once. Similarly, if
∞
H(t) = h k tk ,
k=0
then
7
n
H(t) = (1 + xi t + x2i t2 + · · · ) = (1 − x1 t)−1 · · · (1 − xn t)−1 . (33.2)
i=0
We see that
H(t)E(−t) = 1.
Equating the coefficients in this identity gives us recursive relations
hk − e1 hk−1 + e2 hk−2 − · · · + (−1)k ek = 0, k > 0. (33.3)
These can be used to express the h’s in terms of the e’s or vice versa.
Proposition 33.1. The ring Zsym [x1 , . . . , xn ] is generated as a Z-algebra by
e1 , . . . , en , and they are algebraically independent. Thus, Zsym [x1 , . . . , xn ] =
Z[e1 , . . . , en ] is a polynomial ring. It is also generated by h1 , . . . , hn , which are
algebraically independent, and Zsym [x1 , . . . , xn ] = Z[h1 , . . . , hn ].
Proof. The fact that the ei generate Zsym [x1 , . . . , xn ] is Theorem 6.1 on p. 191
of Lang [116], and their algebraic independence is proved on p. 192 of that
reference. The fact that h1 , . . . , hn also generate follows since (33.3) can be
solved recursively to express the ei in terms of the hi . The hi must be alge-
braically independent since if they were dependent the transcendence degree
of the field of fractions of Zsym [x1 , . . . , xn ] would be less than n, so the ei
would also be algebraically dependent, which is a contradiction.
If V is a vector space, let ∧k V and ∨k V denote the kth exterior and
symmetric powers. If T : V −→ W is a linear transformation, then there are
induced linear transformations ∧k T : ∧k V −→ ∧k W and ∨k T : ∨k V −→
∨k W .
33 Characters of GL(n, C) 351
Proof. First, assume that T is diagonalizable and that v1 , . . . , vn are its eigen-
vectors, so T vi = ti vi . Then a basis of ∧k V consists of the vectors
it follows that the constructible polynomials also form a ring. The ek are
constructible by Proposition 33.2 and generate Zsym [x1 , . . . , xn ] by Proposi-
tion 33.1. Thus, the ring of constructible polynomials is all of Zsym [x1 , . . . , xn ].
Replacing t by txi in this identity, summing over the xi , and using (33.1), we
see that
∞
(−1)k−1
log E(t) = p k tk .
k
k=1
Let us return to the context of Theorem 33.2. Let G be a group and χ the
character of a representation π : G −→ GL(n, C). As we saw in that theorem,
the functions g −→ χk (g) = χ(g k ) are generalized characters; indeed they are
the functions ψpk ◦ π. They are conveniently computable and therefore useful.
The operations χ −→ χk on the ring of generalized characters of G are called
the Adams operations. See also the exercises in Chap. 22 for more about the
Adams operations.
Let us consider an example. Consider the polynomial
s(x1 , . . . , xn ) = x2i xj + 2 xi xj xk . (33.9)
i=j i<j<k
We find that
p31 = x3i + 3 x2i xj + 6 xi xj xk ,
i i=j i<j<k
so
s = 13 (p31 − p3 ). (33.10)
Hence, if π : G −→ GL(n, C) is a representation affording the character χ,
then we have
(ψs ◦ π)(g) = 13 χ(g)3 − χ(g 3 ) . (33.11)
Such a composition of a representation with a ψf is called a plethysm. The
expression on the right-hand side is useful for calculating the values of this
function, which we have proved is a virtual character of GL(n, C), provided we
know the values of the character χ. We will show in the next chapter that (for
this particular s) this plethysm is actually a proper character. Indeed, we will
actually prove that ψs is a character of GL(n, C), not just a virtual character.
This will require ideas different from those than used in this chapter.
EXERCISES
Exercise 33.1. Express each of the sets of polynomials {ek | k 5} and {pk | k 5}
in terms of the other.
Exercise 33.2. Here is the character table of S4 .
1 (123) (12)(34) (12) (1234)
χ1 1 1 1 1 1
χ2 1 1 1 −1 −1
χ3 3 0 −1 1 −1
χ4 3 0 −1 −1 1
χ5 2 −1 2 0 0
Let s be as in (33.9). Using (33.11), compute ψs ◦π when (π, V ) is an irreducible rep-
resentation with character χi for each i, and decompose the resulting class function
into irreducible characters, confirming that it is a generalized character.
34
Duality Between Sk and GL(n, C)
$
Let V be a complex vector space, and let k V = V ⊗ · · · ⊗ V be the k-fold
tensor of V . (Unadorned ⊗ means ⊗C .) We consider this to be a right module
over the group ring C[Sk ], where σ ∈ Sk acts by permuting the factors:
(v1 ⊗ · · · ⊗ vk )σ = vσ(1) ⊗ · · · ⊗ vσ(k) . (34.1)
It may be checked that with this definition
((v1 ⊗ · · · ⊗ vk )σ) τ = (v1 ⊗ · · · ⊗ vk )(στ ).
$k
If A is C-algebra and V is an A-module, then V has an A-module struc-
ture; namely, a ∈ A acts diagonally:
a(v1 ⊗ · · · ⊗ vk ) = av1 ⊗ · · · ⊗ avk .
This action commutes with the action (34.1) of the symmetric group, so it
$k
makes V an (A, C[Sk ])-bimodule. Suppose that ρ : Sk −→ GL(Nρ ) is a
representation. Then Nρ is an Sk -module, so by Remark 32.3
%
k
Vρ = V ⊗C[Sk ] Nρ (34.2)
is a left A-module.
We can take A = End(V ). Embedding GL(V ) −→ A, we obtain a rep-
resentation of GL(V ) parametrized by a module Nρ of Sk . Thus, Vρ is a
GL(V )-module. This is the basic construction of Frobenius–Schur duality.
We now give a reinterpretation of the symmetric and exterior powers,
which were used in the proof of Theorem 33.1. Let Csym be a left C[Sk ]-
module for the trivial representation, and let Calt be a C[Sk ]-module for the
alternating character. Thus, Calt is C with the Sk -module structure
σ x = ε(σ) x,
for σ ∈ Sk , x ∈ Calt , where ε : Sk → {±1} is the alternating character.
The universal property is that any such alternating map factors uniquely
through ∧k V . That is, the map (v1 , . . . , vk ) → v1 ∧· · ·∧vk is itself alternating,
and given any alternating map f : V × · · · × V −→ W there exists a unique
linear map>F : ∧k V? −→ W such that f (v1 , . . . , vk ) = F (v1 ∧ · · · ∧ vk ). We will
$k
show that V ⊗C[Sk ] Calt has the same universal property.
We are identifying the underlying
>$ space ? of Calt with C, so 1 ∈ Calt . There
k
exists a map i : V × · · · × V → V ⊗C[Sk ] Calt given by
such
>$ that ? f = F ◦ i. Uniqueness is clear since the image of i spans the space
k
V ⊗C[Sk ] Calt . To prove existence, we observe first that by the universal
$k
property of the tensor product there exists a linear map f : V → W such
that f (v1 , . . . , vk ) = f (v1 ⊗ · · · ⊗ vk ). Now consider the map
%
k
V × Calt → W
34 Duality Between Sk and GL(n, C) 357
defined by (ξ, t) −→ t f (ξ). It follows from the fact that f is alternating that
this map is C[Sk ]-balanced and consequently induces a map
%
k
F : V ⊗C[Sk ] Calt → W.
>$ ?
k
This is the map we are seeking. We see that V ⊗C[Sk ] Calt satisfies the
same universal property as the exterior power, so it is naturally isomorphic
to ∧k V .
Proof. First let us prove this for g restricted to the subgroup of diagonal
matrices. Let ξ1 , . . . , ξn be the standard basis of V . In other words, identifying
V with Cn , let ξi = (0, . . . , 1, . . . , 0), where the 1 is in the ith position. The
vectors (ξi1 ⊗ · · · ⊗ ξik ) ⊗ ν, where ν runs through a basis of Nρ , and 1
i1 · · · ik n span Vρ . They will generally not be linearly independent,
but there will be a linearly independent subset that forms a basis of Vρ . For g
diagonal, if g(ξi ) = ti ξi , then (ξi1 ⊗ · · · ⊗ ξik ) ⊗ ν will be an eigenvector for g
in Vρ with eigenvalue ti1 · · · tik . Thus, we see that there exists a homogeneous
polynomial sρ of degree k such that (34.3) is true for diagonal matrices g.
To see that sρ is symmetric, we have pointed out that the action of Sk on
⊗k V commutes with the action of GL(n, C). In particular, it commutes with
the action of the permutation matrices in GL(n, C), which form a subgroup
isomorphic to Sn . These permute the eigenvectors (ξi1 ⊗ · · · ⊗ ξik ) ⊗ ν of g
and hence their eigenvalues. Thus, the polynomial sρ must be symmetric.
Since the eigenvalues of a matrix are equal to the eigenvalues of any con-
jugate, we see that (34.3) must be true for any matrix that is conjugate to a
diagonal matrix. Since these are dense in GL(n, C), (34.3) follows for all g by
continuity.
∼
multiplicity of ρi in the regular representation is di , that is, C[Sk ] =
The
di Nρi , and hence
%k & %k &
V ∼= di V ⊗C[Sk ] Nρi = di Vρi . (34.6)
i i
p31 = h3 + e3 + 2sρ3 ,
5, 4 + 1, 3 + 2, 3 + 1 + 1, 2 + 2 + 1, 2 + 1 + 1 + 1, 1 + 1 + 1 + 1 + 1.
Note that the partitions 3 + 2 and 2 + 3 are considered equal. We may arrange
the terms in a partition into descending order. Hence, a partition λ of k may
be more formally defined to be a sequence of nonnegative integers (λ1 , . . . , λl )
such that λ1 λ2 · · · λl 0 and i λi = k. It is sometimes convenient to
allow some of the parts λi to be zero, in which case we identify two sequences
if they differ only by trailing zeros. Thus, (3, 2, 0, 0) is considered to be the
same partition as (3, 2). The length or number of parts l(λ) of the partition λ
is the largest i such that λi
= 0, so the length of the partition (3, 2) is two.
We will denote by p(k) the number of partitions of k, so that p(5) = 7.
If λ is a partition of k, there is another partition, called the conjugate
partition and denoted λt , which may be constructed as follows. We construct
from λ a diagram in which the ith row is a series of λi boxes. Thus, the
diagram corresponding to the partition λ = (3, 2) is
Λ(n) ∼
= Z[h1 , . . . , hn ].
(n) (n)
Λ(n) is a graded ring. We have Λ(n) = Λk , where Λk consists of all
(n)
homogeneous polynomials of degree k in Λ .
(n)
Proposition 34.3. The homogeneous part Λk is a free Abelian group of rank
equal to the number of partitions of k into no more than n parts.
there exists a unique element whose image under the projection Λ → Λ(n) is
ek for all n k; we naturally denote this element ek , and (34.8) implies that
Λ∼
= Z[e1 , e2 , e3 , . . .]
Λ∼
= Z[h1 , h2 , h3 , . . .].
Proof. Of course, it does not matter which ring we work in. Since Λ ∼ =
Z[e1 , e2 , e3 , . . .], and since the ei are algebraically independent, if u1 , u2 , . . . are
arbitrarily elements of Λ, there exists a unique ring homomorphism Λ −→ Λ
such that ei −→ ui . What we must show is that if we take the ui = hi , then
this same homomorphism maps hi −→ ui . This follows from the fact that the
34 Duality Between Sk and GL(n, C) 363
recursive identity (33.3), from which we may solve for the e’s in terms of the
h’s or conversely, is unchanged if we interchange ei ←→ hi .
EXERCISES
Exercise 34.1. Let s = h1 h2 − h3 . Show that ι s = s.
35
The Jacobi–Trudi Identity
Proof. Let us show how to prove this fact using exterior algebra. Suppose that
A is an N × N matrix. Let V = CN . Then ∧N V is one-dimensional, and we
(∧kA, ∧N−kA)
(∧kV ) × (∧N−kV ) (∧kV ) × (∧N−kV )
∧ ∧
∧NA
∧NV ∧NV
η η
det A
C C
η(v1 ∧ · · · ∧ vN ) = 1.
Then a pair of dual bases of ∧k V and ∧N −k V with respect to this pairing are
{i1 , . . . , ik }, {j1 , . . . , jN −k } ,
of {1, . . . , N } are complementary. [The sign of the second basis vector will
be (−1)d , where d = (i1 − 1) + (i2 − 2) + · · · + (ik − k).] If det(A) = 1,
then the bottom arrow is the identity map, and therefore we see that the map
∧N −k A : ∧N −k V → ∧N −k V is the inverse of the adjoint of ∧k A : ∧k V → ∧k V
with respect to this dual pairing. Hence, if we use the above dual bases to
compute matrices for these two maps, the matrix of ∧N −k A is the transpose
of the inverse of the matrix of ∧k A. Thus, if B is the inverse of the adjoint
of A with respect to the inner product on V for which v1 , . . . , vN are an
orthonormal basis, then the matrix of ∧N −k B is the same as the matrix of
∧k A. Now, with respect to the chosen dual bases, the coefficients in the matrix
of ∧k A are the k × k minors of A, while the matrix coefficients of ∧N −k B are
(up to sign) the complementary (N − k) × (N − k) minors of B. Hence, these
are equal.
s + i − λi , (i = 1, . . . , r),
s − j + μj + 1, (j = 1, . . . , s),
are 1, 2, 3, . . . , r + s rearranged.
35 The Jacobi–Trudi Identity 367
Proof. First note that the r + s integers all lie between 0 and r + s. Indeed,
if 1 i r, then
0 s + i − λi s + r
because s is greater than or equal to the length l(μ) = λ1 λi , so s + i − λi
s − λi 0, and s + i − λi s + i s + r; and if 1 j s, then
0 s − j + μj + 1 s + r
if k 1.
(n)
Proof. Choose n k so that the characteristic map ch(n) : Rk → Λk is
injective. It is then sufficient to prove that ch(n) annihilates the left-hand
side. Since ch(n) (ei ) = ei and ch(n) (hi ) = hi , this follows from (33.3).
Later, in Proposition 35.1 we will see that the sign in (35.3) is always +. This
could be proved now by carefully keeping track of the sign, but this is more
trouble than it is worth because we will determine the sign in a different way.
conjugated by ⎛ ⎞
1
⎜ −1 ⎟
⎜ ⎟
⎜ .. ⎟.
⎝ . ⎠
(−1)r+s−1
We only need to compute the minors up to sign, and conjugation by the
latter matrix only changes the signs of these minors. Hence, it follows from
Proposition 35.1 that each minor of (35.4) is, up to sign, the same as the
complementary minor of (35.5). Let us choose the minor of (35.4) with columns
s+1, . . . , s+r and rows s+i−λi (i = 1, . . . , r). This minor is the left-hand side
of (35.3). By Proposition 35.2, the complementary minor of (35.5) is formed
with columns 1, . . . , s and rows s − j + μj + 1 (j = 1, . . . , s). After conjugating
this matrix by ⎛ ⎞
1
.
⎝ .. ⎠ ,
1
we obtain the right-hand side of (35.3).
λ1 + · · · + λi μ1 + · · · + μi , (i = 1, 2, 3, . . .).
where ε is the alternating character of Sλ , and IndSSkμ (ε) denotes the cor-
responding induced representation of Sk . This is because eμi ∈ Rμi is the
alternating character of Sλi , and the multiplication in R is defined so that
the product eμ = eμ1 · · · eμs is obtained by induction from Sμ .
By Mackey’s theorem, we must count the number of double cosets in
Sμ \Sk /Sλ that support intertwining operators. (See Remark 32.2.) Simply
counting these double cosets is sufficient because the representations that we
are inducing are both one-dimensional, so each space on the right-hand side
of (32.12) is either one-dimensional (if the coset supports an intertwining op-
erator) or zero-dimensional (if it doesn’t).
First, we will show that the double cosets in Sμ \Sk /Sλ may be parametrized
by s × r matrices with nonnegative integer coefficients such that the sum of
the ith row is equal to μi and the sum of the jth column is equal to λj . Then
we will show that the double cosets that support intertwining operators are
precisely those that have no entry > 1. This will prove the first assertion.
We will identify Sk with the group of k × k permutation matrices. (A
permutation matrix is one that has only zeros and ones as entries, with ex-
actly one nonzero entry in each row and column.) Then Sλ is the subgroup
consisting of elements of the form
370 35 The Jacobi–Trudi Identity
⎛ ⎞
D1 0 ··· 0
⎜ 0 D2 · · · 0 ⎟
⎜ ⎟
⎜ .. .. . . .. ⎟,
⎝ . . . . ⎠
0 0 ··· Dr
where Gij is a μi × λj block. Let γij be the rank of Gij , which is the number
of nonzero entries. Then the matrix r × s matrix (γij ) is independent of the
choice of representative of the double coset. It has the property that the sum
of the ith row is equal to μi and the sum of the jth column is equal to λj .
Moreover, it is easy to see that any such matrix arises from a double coset in
this manner and determines the double coset uniquely. This establishes the
correspondence between the matrices (γij ) and the double cosets.
Next we show that a double coset supports an intertwining operator if and
only if each γij 1. A double coset Sμ gSλ supports an intertwining operator
if and only if there exists a nonzero function Δ : Sk → C with support in
Sμ gSλ such that
Δ(τ hσ) = ε(τ )Δ(h) (35.8)
for τ ∈ Sμ , σ ∈ Sλ .
First, suppose the matrix (γij ) is given such that for some particular i, j,
we have γ = γij > 1. Then we may take as our representative of the double
coset a matrix g such that
Iγ 0
Gij = .
0 0
so Δ(g) = 0 and therefore Δ is identically zero. We see that if any γij > 1, then
the corresponding double coset does not support an intertwining operator.
On the other hand, if each γij 1, then we will show that for g a repre-
sentative of the corresponding double coset, g −1 Sμ g ∩ Sλ = {1}, or
with τμi ∈ Sμi and σλi ∈ Sλi and letting g be as in (35.7), we have τμi Gij =
Gij σλj . If τμi
= I, then
τμi Gi1 · · · Gir
= Gi1 · · · Gir
since the rows of the second matrix are distinct. Thus τμi Gij
= Gij for some i.
Since Gij has at most one nonzero entry, it is impossible that after reordering
the rows (which is the effect of left multiplication by τμi ) this nonzero entry
could be restored to its original position by reordering the columns (which
is the effect of right multiplication by σλ−1 j
). Thus, τμi Gij
= Gij implies that
τμi Gij
= Gij σλj . This contradiction proves (35.9).
Now (35.9) shows that each element of the double coset has a unique
representation as τ gσ with τ ∈ Sμ and σ ∈ Sλ . Hence, we may define
ε(τ ) if h = τ gσ with τ ∈ Sμ and σ ∈ Sλ ,
Δ(h) =
0 otherwise,
and this is well-defined. Hence, such a double coset does support an intertwin-
ing operator.
Now we have asserted further that (35.6) is nonzero if and only if μt λ
and that if μt = λ, then the inner product is 1. Let us ask, therefore, for given
λ and μ, whether we can construct a matrix (γij ) with each γij = 0 or 1 such
that the sum of the ith row is μi and the sum of the jth column is λj . Let
ν = μt . Then
νi = card {j | μj i}.
That is, νi is the number of rows that will accommodate up to i 1’s. Now
ν1 + ν2 + · · · + νt is equal to the number of rows that will take a 1, plus the
number of rows that will take two 1’s, and so forth. Let us ask how many 1’s
we may put in the first t columns. Each nonzero entry must lie in a different
row , so to put as many 1’s as possible in the first t columns, we should put νt
of them in those rows that will accommodate t nonzero entries, νt−1 of them in
those rows that will accommodate t−1 entries, and so forth. Thus, ν1 +· · ·+νt
is the maximum number of 1’s we can put in the first t columns. We need to
place λ1 + · · · + λt ones in these rows, so in order for the construction to be
possible, what we need is
λ1 + · · · + λt ν1 + · · · + νt
for each t, that is, for ν λ. It is easy to see that if ν = λ, then the location of
the ones in the matrix (γij ) is forced so that in this case there exists a unique
intertwining operator.
372 35 The Jacobi–Trudi Identity
λ1 λ1 − 1 + j1 λ1 ,
and similarly j1 + j2 3 so
and so forth.
Similarly, expanding the right-hand side gives a sum of terms of the form
±eμ , where μ μ, and the term eμ also occurs exactly once.
35 The Jacobi–Trudi Identity 373
EXERCISES
Exercise 35.1. Let λ and μ be partitions of k. Show that
hλ , hμ = eλ , eμ
and that this inner product is equal to the number of r × s matrices with each
coefficient a nonnegative integer such that the sum of the ith row is equal to λi , and
the sum of the jth column is equal to μj .
Similarly we will define Txy when x, y are s, e or h to denote the transition matrices
between the bases sλ , eλ and hλ of Λk .
(i) Show that Tsh (λ, μ) = 0 unless μ λ.
(ii) Show that Ths (λ, μ) = 0 unless μ λ.
(iii) Show that Tse (λ, μ) = 0 unless μ λt .
(iv) Show that Tes (λ, μ) = 0 unless μt λ.
(v) Show that The (λ, μ) = 0 unless μt λ.
(v) Show that Teh (λ, μ) = 0 unless μt λ.
Zelevinsky [178] shows how the ring R may be given the structure of a
graded Hopf algebra. This extra algebraic structure (actually introduced earlier by
Geissinger) encodes all the information about the representations of Sk that comes
from Mackey theory. Moreover, a similar structure exists in a ring R(q) analogous
to R, constructed from the representations of GL(k, Fq ), which we will consider in
Chap. 47. Thus, Zelevinsky is able to give a unified discussion of important aspects
of the two theories. In the next exercises, we will establish the basic fact that R is
a Hopf algebra.
We begin by reviewing the notion of a Hopf algebra. We recommend Majid [125]
for further insight. (Apart from its use as an introduction to quantum groups, this is
good for gaining facility with Hopf algebra methods such as the Sweedler notation.)
Let A be a commutative ring. An A-algebra is normally defined to be a ring R with
a homomorphism u into the center of R. The homomorphism u (called the unit)
then makes R into an A-module. The multiplication map R × R → R is A-bilinear
hence induces a linear map m : R ⊗ R → R. The associative law for multiplication
may be interpreted as the commutativity of the diagram:
m⊗1
R⊗R⊗R R⊗R
1⊗m m
m
R⊗R R
R⊗R R⊗A R
u⊗1 m
1⊗u m
A⊗R R R⊗R
The dual notion to an algebra is that of a coalgebra. The definition and axioms
are obtained by reversing all the arrows. That is, we require an A-module R together
with linear maps Δ : R → R ⊗ R and : R → A such that we have commutative
diagrams
Δ
R R⊗R
Δ 1⊗Δ
Δ⊗1
R⊗R R⊗R⊗R
R A⊗R R⊗R
Δ 1⊗
Δ ⊗1
R⊗R R R⊗A
Exercise 35.4. Let R be a algebra that is also a coalgebra. Show that the three
statements are equivalent.
(i) The comultiplication Δ : R −→ R ⊗ R and counit R → A are homomorphisms
of algebras.
(ii) The multiplication m : R ⊗ R −→ R and unit A → R are homomorphisms of
coalgebras.
(iii) The following diagram is commutative:
Here τ is the “transposition” map R ⊗ R → R ⊗ R that sends x ⊗ y to y ⊗ x.
We will refer to this property as the Hopf axiom.
Δ⊗Δ 1⊗τ⊗1
R⊗R R⊗R⊗R⊗R R⊗R⊗R⊗R
m m⊗m
Δ
R R⊗R
Exercise 35.5. Let G be a finite group, A = C and let R be the group algebra.
Define a map Δ : R → R ⊗ R by extending the diagonal map G → G × G to a linear
map R → R ⊗ R, and let : R → A be the augmentation map that sends every
element of G to 1. Show that R is a bialgebra.
Exercise 35.6. Suppose that k+l = m. Let ⊗ denote ⊗Z . The group Rk ⊗Rl can be
identified with the free Abelian group identified with the irreducible representations
of Sk × Sl . (Explain.) So restriction of a representation from Sm to Sk × Sl gives a
group homomorphism Rm −→ Rk ⊗ Rl . Combining these maps gives a map
'
Δ : Rm −→ Rk ⊗ Rl = (R ⊗ R)m .
k+l=m
Show that this homomorphism of graded Z-algebras makes R into a graded coalge-
bra.
where a + b = k, c + d = l, a + c = p, and b + d = q.
(ii) Use (i) and Mackey theory to prove that R is a graded bialgebra over Z.
Hint: Both parts are similar to parts of the proof of Proposition 35.5.
1⊗S A S⊗1
R⊗R m R m R⊗R
35 The Jacobi–Trudi Identity 377
Exercise 35.8. Show that a group algebra is a Hopf algebra. The antipode is the
map S(g) = g −1 .
Exercise 35.9. Show that R is a Hopf algebra. We have S(hk ) = (−1)k ek and
S(ek ) = (−1)k hk .
The next exercise is from the 2013 senior thesis of Seth Shelley-Abrahamson.
Similar statements relate the higher Hopf power maps to other wreath products.
Interest in the Hopf square map and higher-power maps has been stimulated by
recent investigations of Diaconis, Pang, and Ram.
Now let sμ (x1 , . . . , xn ) be the symmetric polynomial ch(n) (sμ ); we will use the
same notation sμ for the element ch(sμ ) of the inverse limit ring Λ defined by
(34.10). These are the Schur polynomials.
Theorem 36.1. Assume that n l(λ). We have
λ1 +n−1 λ1 +n−1
x1 λ1 +n−1
λ +n−2 xλ2 +n−2 · · · xλn +n−2
x 2 x2 2 · · · xn2
1
. .
.. ..
xλn x2λn
· · · xn λn
sλ (x1 , . . . , xn ) = 1
n−1 n−1 , (36.1)
x1 x2 · · · xn−1
n−2 n−2 n
x n−2
1 x2 · · · xn
. ..
.. .
x1 x2 · · · xn
1 1 ··· 1
provided that n is greater than or equal to the length of the partition k, so that
we may denote λ = (λ1 , . . . , λn ) (possibly with trailing zeros). In this case
sλ
= 0.
It is worth recalling that the Vandermonde determinant in the denominator
can be factored:
n−1 n−1
x1 x2 · · · xn−1
n−2 n−2 n
x n−2
1 x2 · · · xn 7
. .. =
.. . (xi − xj ).
x1 x2 · · · xn i<j
1 1 ··· 1
It is also worth noting, since it is not immediately obvious from the expression
(36.1), that the Schur polynomial sλ in n + 1 variables restricts to the Schur
polynomial also denoted sλ under the map (34.9). This is of course clear from
Proposition 34.6 and the fact that ch(sλ ) = sλ .
(i)
Proof. Let ek be the kth elementary symmetric matrix in n − 1 variables
x1 , . . . , xi−i , xi+1 , . . . , xn ,
omitting xi . We have, using (33.1) and (33.2) and omitting one variable in
(33.1),
∞ 7n
(i)
(−1)k ek tk = (1 − xj t),
k=0 j=i
∞ 7 n
hk t k
= (1 − xi t)−1 ,
k=0 i=1
and therefore
-∞ .- ∞ .
k (i) k
(−1) ek t hk t = (1 − txi )−1 = 1 + txi + t2 x2i + · · · .
k
k=0 k=0
Suppose that V and W are vector spaces over a field of characteristic zero
and B : V × · · · × V −→ W is a symmetric k-linear map. Let Q : V −→ W be
the function Q(v) = B(v, . . . , v). The function B can be reconstructed from
Q, and this process is called polarization. For example, if k = 2 we have
1
B(v, w) = (Q(v + w) − Q(v) − Q(w)) ,
2
as we may see by expanding the right-hand side and using B(v, w) = B(w, v).
Therefore,
(−1)k−|S| Q(uS ) = B(ui1 , . . . , uik ) (−1)k−|S| .
S⊆I 1 i1 k S⊇{i1 ,...,ik }
.
.
.
1 ik k
Suppose that there are repetitions among the list i1 , . . . , ik . Then there will be
some j ∈ I such that j ∈ / {i1 , . . . , ik }, and pairing
those subsets containing j
with those not containing j, we see that the sum S⊇{i1 ,...,ik } (−1)k−|S| = 0.
Hence, we need only consider those terms where {i1 , . . . , ik } is a permutation
of {1, . . . , k}. Remembering that B is symmetric, these terms all contribute
equally and the result follows.
Theorem 36.2. Let λ be a partition of k, and let n l(λ). Then there exists
GL(n)
an irreducible representation πλ = πλ of GL(n, C) with character χλ such
that if g ∈ GL(n, C) has eigenvalues t1 , . . . , tn , then
f1 ⊗ · · · ⊗ fk −→ fσ(1) ⊗ · · · ⊗ fσ(k)
$k $k
of End(V ). If ξ ∈ End(V ) commutes with this action, then ξ is a
linear combination of elements of the form B(f1 , . . . , fk ), where B : End(V ) ×
$
· · · × End(V ) −→ k End(V ) is the symmetric k-linear map
B(f1 , . . . , fk ) = fσ(1) ⊗ · · · ⊗ fσ(k) .
σ∈Sk
It $
follows from Proposition 36.1 that the vector space of such elements
k
of End(V ) is spanned by those of the form Q(f ) = B(f, . . . , f ) with
f ∈ End(V ). Since GL(n, C) is dense in End(V ), the elements Q(f ) with f
invertible span the same vector space. This proves that$the transformations of
k
the form (36.4) span the space of transformations of V commuting with
the action of Sk . $k
We temporarily restrict the action of GL(n, C) × Sk on V to the com-
pact subgroup U(n) × Sk . Representations of a compact group are completely
reducible, and the irreducible representations of U(n) × Sk are of the form
π ⊗ ρ, where π is an irreducible representation of U(n) and ρ is an irreducible
representation of Sk . Thus, we write
%k
V ∼
= πi ⊗ ρi , (36.5)
i
Lie algebra action of u(n) and hence is invariant under the action of the
complexified Lie algebra u(n) + iu(n) = gl(n, C) and therefore under its exp-
onential, GL(n, C). So we may regard the decomposition (36.5) as a decom-
position with respect to GL(n, C) × Sk .
We claim that there are no repetitions among the isomorphism classes of
the representations ρi of Sk that occur. This is because if ρi ∼= ρj , then if we
denote by f an intertwining map ρi −→ ρj and by τ an arbitrary nonzero
linear transformation from the space of πi to the space of πj , then τ ⊗ f is
a map from the space of πi ⊗ ρi to the space of πj ⊗ ρj that commutes with
the action of Sk . Extending it by zero on direct summands in (36.5) beside
$k
πi ⊗ ρi gives an endomorphism of V that commutes with the action of Sk .
It therefore is in the span of the endomorphisms (36.4). But this is impossible
because those endomorphisms leave πi ⊗ ρi invariant and this one does not.
This contradiction shows that the ρi all have distinct isomorphism classes.
It follows from this that at most one ρi>can be?isomorphic to the contragre-
$k
dient representation of ρλ . Thus, in Vρ = V ⊗C[Sk ] Nρ at most one term
can survive, and that term will be isomorphic to πi as a GL(n, C) module for
this unique i. We know that Vρ is nonzero since by Theorem 36.1 the polyno-
mial sλ
= 0 under our hypothesis that l(λ) n. Thus, such a πi does exist,
and it is irreducible as a U(n)-module a fortiori as a GL(n, C)-module.
It remains to be shown that if μ
= λ, then χμ
= χλ . Indeed, the Schur
polynomials sμ and sλ are distinct since the partition λ can be read off from
the numerator in (36.1).
We have constructed an irreducible representation of GL(n, C) for every
partition λ = (λ1 , . . . , λn ) of length n.
Proposition 36.2. Suppose that n l(λ). Let
λ = (λ1 − λn , λ2 − λn , . . . , λn−1 − λn , 0).
In the ring Λ(n) of symmetric polynomials in n variables, we have
sλ (x1 , . . . , xn ) = en (x1 , . . . , xn )λn sλ (x1 , . . . , xn ). (36.6)
In terms of the characters of GL(n, C), we have
χλ (g) = det(g)λn χλ (g). (36.7)
Note that en (x1 , . . . , xn ) = x1 · · · xn . Caution: This identity is special to Λ(n) .
The corresponding statement is not true in Λ.
Proof. It follows from (36.1) that sλ (x1 , . . . , xn ) is divisible by (x1 · · · xn )λn .
Indeed, each entry of the first column of the matrix in the numerator is divis-
ible by xλ1 n , so we may pull xλ1 n out of the first column, xλ2 n out of the second
column, and so forth, obtaining (36.6).
If the eigenvalues of g are t1 , . . . , tn , then en (t1 , . . . , tn ) = t1 · · · tn = det(g)
and (36.7) follows from (36.6) and (36.3).
384 36 Schur Polynomials and GL(n, C)
Proof. We may assume that π1 and π2 act on the same complex vector space
V , and that π1 (g) = π2 (g) when g ∈ U(n). Applying Lemma 36.1 to the matrix
coefficients of π1 and π2 it follows that π1 (g) = π2 (g) for all g ∈ GL(n, C).
36 Schur Polynomials and GL(n, C) 385
Proposition 36.5. Suppose that λ is a partition and l(λ) > n. Then we have
sλ (x1 , . . . , xn ) = 0 in the ring Λ(n) .
Proof. Most of this was proved in the proof of Theorem 36.2. Particularly, we
saw there that each irreducible representation of Sk occurring in (36.5) occurs
at most once and is paired with an irreducible representation of GL(n, C).
If l(λ) n, we saw in the proof of Theorem 36.2 that ρλ does occur and
is paired with πλ . The one fact that was not proved there is that ρλ with
l(λ) > n do not occur, and this follows from Proposition 36.5.
37
Schur Polynomials and Sk
Frobenius [51] discovered that the characters of the symmetric group can be
computed using symmetric functions. We will explain this from our point of
view. We highly recommend Curtis [39] as an account, both historical and
mathematical, of the work of Frobenius and Schur on representation theory.
We remind the reader that the elements of Rk , as generalized characters,
are class functions on Sk . The conjugacy classes of Sk are parametrized by
the partitions as follows. Let λ = (λ1 , . . . , λr ) be a partition of k. Let Cλ
be the conjugacy class consisting of products of disjoint cycles of lengths
λ1 , λ2 , . . . . Thus, if k = 7 and λ = (3, 3, 1), then Cλ consists of the conjugates
of (123) (456) (7) = (123) (456). We say that the partition λ is the cycle type
of the permutations in the conjugacy class Cλ . Let zλ = |Sk |/|Cλ |.
The support of σ ∈ Sk is the set of x ∈ {1, 2, 3, . . . , k} such that σ(x)
= x.
Proof. We have
1
sμ , pλ = zλ sμ (x).
|Sk |
x ∈ Cλ
The summand is constant on Cλ and equals zλ sμ (g) for any fixed representa-
tive g. The cardinality of Cλ is |Sk |/zλ and the result follows.
Proof. From the definitions, pλ1 · · · pλh is induced from the class function f
on the subgroup Sλ of Sk which has a value on (σ1 , . . . , σh ) that is
λ1 · · · λh if each σi is a λi -cycle ,
0 otherwise.
The formula (32.15) may be used to compute this induced class function. It
is clear that pλ1 · · · pλh is supported on the conjugacy class of cycle type λ,
and so it is a constant multiple of pλ . We write pλ1 · · · pλh = cpλ and use a
trick to show that c = 1. By Proposition 37.3, since hk = s(k) is the trivial
character of Sk , we have hk , pλ Sk = 1. On the other hand, by Frobenius
reciprocity, hk , pλ1 · · · pλh Sk = hk , f Sλ . As a class function, hk is just the
constant function on Sk equal to 1, so this inner product is
7
hλi , pλi Sλ = 1.
i
i
Therefore, c = 1.
Proposition 37.6. We have
k
khk = pr hk−r . (37.3)
r=1
Since
∞
d
log(1 − xi t)−1 = xri tr−1 ,
dt r=1
we obtain -∞ .
∞
∞
khk tk−1 = h k tk pr tr−1 .
k=1 k=0 r=1
Then the coefficient cλμ is the value of the irreducible character sμ on elements
of the conjugacy class Cλ .
37 Schur Polynomials and Sk 391
cλμ = pλ , sμ .
in the polynomal pλ Δ.
Proof. By Theorem 37.2, we have pλ = μ cλμ sμ , and by (36.1) this means
that
pλ Δ = cλμ det(xjμi +n−i ),
μ
the determinant being the determinant in the numerator in (36.1). The mono-
mial (37.6) appears only in the μ term and the statement follows.
and
p(3) = x3i ,
p(21) = x3i + i=j x2i xj ,
p(111) = x3i + 3 i=j x2i xj + 6 i<j<k xi xj xk ,
so
p(111) = s(3) + s(111) + 2s(21) ,
p(3) = s(3) + s(111) − s(21) ,
p(21 ) = s(3) − s(111) .
392 37 Schur Polynomials and Sk
1 (123) (12)
s(3) 1 1 1
s(111) 1 1 −1
s(21) 2 −1 0
Proof. This may be proved either directly from the definition of the induced
representation or by using (32.15).
Proof. Let us denote by τ : Rk −→ Rk the linear map that takes a class func-
tion f on Sk and multiplies it by ε, and assemble the τ in different degrees to a
linear map of R to itself. We want to prove that τ and ι are the same. By the
definition of the ek and hk , they are interchanged by τ , and by Theorem 35.2
they are interchanged by ι. Since the ek generate R as a ring, the result will
follow if we check that τ is a ring homomorphism.
Applying Lemma 37.1 with G = Sk+l , H = Sk × Sl , and ρ = ε shows
that multiplying the characters χ and η of Sk and Sl each by ε to obtain the
characters τ χ and τ η and then inducing the character τ χ ⊗ τ η of Sk × Sl to
Sk+l gives the same result as inducing χ ⊗ η and multiplying it by ε. This
shows that τ is a ring homomorphism.
EXERCISES
Exercise 37.1. Compute the character table of S4 using symmetric polynomials by
the method of this chapter.
37 Schur Polynomials and Sk 393
k
kek = (−1)r pr ek−r .
r=1
Let us say that a partition λ is a ribbon partition if its Young diagram only
has entries in the first row and column. The ribbon partitions of k are of the form
(k − r, 1r ) with 0 r k, where the notation means the partition with one part of
length k − r and r parts of length 1.
where the sum is over all partitions λ (of all k). The series is absolutely
convergent if all |αi |, |βi | < 1. It can also be regarded as an equality of formal
power series.
The general context for our discussion of the Cauchy identity will be the
Frobenius–Schur duality. For other approaches, see Exercises 26.4 and 38.4.
We recall from Chap. 34 that the characteristic map ch : R −→ Λ(N )
allows us to interpret a character (or class function) on the symmetric group
Sk as a symmetric polynomial in N variables that is homogeneous of degree k.
Here is a simple fact we will need. Notations are as in Chap. 37.
Proof. Both sides polynomials in the αi and βi that are symmetric and homo-
geneous of degree k in either set of variables. Use the Frobenius–Schur duality
to transfer the function on the right-hand side to a function on Sk ×Sk . In view
of Theorem 37.1 this is the function
Δ(σ, τ ) = zλ−1 pλ (σ) pλ (τ )
λ a partition of k
that maps (σ, τ ) ∈ Sk × Sk to the function that has the value zλ if σ and τ
are in the conjugacy class Cλ , and is zero if σ and τ are not conjugate. This
function may be characterized as follows: if f is a class function, then
1
Δ(σ, τ ) f (τ ) = f (σ).
k!
τ ∈Sk
is also a class function in σ and τ separately, and it has the same reproducing
property, as a consequence of Schur orthogonality. Hence these are equal.
We see that
sλ (σ) sλ (τ ) = zλ−1 pλ (σ) pλ (τ ),
λ a partition of k λ a partition of k
7
n 7
m
(1 − αi βj )−1 = sλ (α1 , . . . , αn ) sλ (β1 , . . . , βm ). (38.3)
i=1 j=1 λ
so it is sufficient to show
sλ (α1 , . . . , αn ) sλ (β1 , . . . , βm ) = hk (αi βj ). (38.4)
λ a partition of k
We now make the observation that pk (αi βj ), which is the kth power sum sym-
metric polynomial in nm variables αi βj , equals pk (α)pk (β), and so pλ (αi βj ) =
pλ (α)pλ (β). The statement now follows from Proposition 38.1.
by analytic continuation in t.
398 38 The Cauchy Identity
Proposition 38.3. Let G = GLn (C) × GLm (C) acting on the tensor product
Ω = Cn ⊗ Cm of the standard modules of GLn (C) and GLm (C). Then the
symmetric algebra
' &
Ω∼
= πλGLn ⊗ πλGLm (38.5)
λ
Proof. If g has eigenvalues αi and h has eigenvalues βj , then (g, h) has eigen-
hk (αi βj ) on ∨ Ω. By the Cauchy identity
k
values αi βj on Ω, hence has trace
in the form (38.4), this equals sλ (α) sλ (β) where the sum is over partitions
of k. [If the length of the partition λ is > n, we interpret sλ (α) as zero.] Com-
bining the contributions over all k, the statement follows.
Note that now each partition λ is paired with its conjugate partition λt . This
may be regarded as a decomposition of the exterior algebra on Matn (C)∗ .
Proof. Let α1 , . . . , αn be fixed complex numbers, and let Λ(m) be the ring
of symmetric polynomials in β1 , . . . , βm with integer coefficients. We recall
from Theorems 34.3 and 35.2 that Λ has an involution ι that interchanges sλ
and sλ . We have to be careful how we use ι because it does not induce an
involution of Λ(m) . Indeed, it is possible that in Λ(m) one of sλ and sλ is zero
and the other is not, so no involution exists that simply interchanges them.
We write the Cauchy identity in the form
-∞ .
7n
αki hk (β1 , . . . , βm ) = sλ (α1 , . . . , αn ) sλ (β1 , . . . , bm ).
i=1 k=0 λ
where the hk on the left and the second occurrence of sλ on the right are
regarded as elements of the ring Λ, which is the inverse limit (34.10) of the
rings Λ(m) , while αi and sλ (α1 , . . . , αn ) are regarded as complex numbers.
To this identity we may apply ι and obtain
38 The Cauchy Identity 399
∞
- .
7
n
αki ek = sλ (α1 , . . . , αn ) sλ ,
i=1 k=0 λ
since taking the inner product of the left-hand side with sλ and using Frobe-
nius reciprocity gives the coefficient of ρμ ⊗ ρν in the restriction of ρλ from
Sk+l to Sk × Sl . On the other hand (ii) can be expressed as the identity
sμ (x1 , . . . , xn ) sν (x1 , . . . , xn ) = cλμν sλ (x1 , . . . , xn )
λ
Since the functions sλ (γ) are linearly independent as λ varies, we may compare
the coefficients of sλ (γ) and obtain (38.8).
It is worth pondering the mechanism behind the proof that (ii) is equivalent
to (iii). We will reconsider it after some preliminaries.
We begin with the notion of a correspondence in the sense of Howe, who
wrote many papers on this subject: see Howe [75, 77]. A correspondence is a
bijection between a set of irreducible representations of a group G and another
group H. The relevant examples arise in the following manner.
Let G be a group with a representation Θ, and let G and H be subgroups
that centralize each other. Thus, we have a homomorphism G × H −→ G.
(Often this homomorphism is injective so G × H is a subgroup of G, but
we do not require this.) We assume given a representation Θ of G with the
following property: when Θ is restricted to G × H, it becomes a direct sum
πi ⊗ πi , where πi are irreducible representations of G, and πi are irreducible
representations of H. We assume that each πi ⊗ πi occurs with multiplicity at
most one, and moreover, there are no repetitions between the representations
πi and no repetitions among the πi . (This definition is adequate if G and H
are compact but might need to be generalized slightly if they are not.) If this
condition is satisfied, we say the representation Θ induces a correspondence
for G and H. The correspondence is the bijection πi ←→ πi . Here are some
examples.
• Let G = Sk , H = GL(n, C), and G = G × H. The representation Θ
is the action on ⊗k Cn in Theorem 36.4. That theorem implies that Θ
induces a correspondence. Indeed, by Theorem 36.4 the correspondence
GL(n)
is the bijection ρλ ←→ πλ , as λ runs through partitions of k that
have length n. Thus, the Frobenius–Schur duality is an example of a
correspondence.
• Consider G = GL(n, C), H = GL(m, C) acting on Ω = Cn ⊗ Cm as above.
Since (g ⊗ Im )(In ⊗ h) = (g ⊗ h) = (In ⊗ h)(g ⊗ Im ), the actions of G and H
commute. Let Θ be the action on the symmetric algebra of Ω. It is actually
a representation of G = GL(Ω). As we have already explained, when
restricted to G × H, the Cauchy identity implies the decomposition (38.5),
GL(n) GL(m)
so Θ induces a correspondence. This is the bijection πλ ←→ πλ
as λ runs through all partitions of length min(m, n). This equivalence
is sometimes referred to as GL(n) × GL(m)-duality.
• Howe conjectured [73], and it was eventually proved, that if G and H are
reductive subgroups of Sp(2N, F ), where F is a local field (including R
or C), then the Weil (oscillator) representation induces a correspondence.
In some cases one of the groups of the correspondence must be replaced
by a covering group. In one most important case, G = Sp(2n) and
H = O(m), where nm = N , so the correspondence relates representa-
tions of symplectic groups (or their double covers) to representations of
402 38 The Cauchy Identity
G G
H H
(38.9)
The vertical lines are inclusions, and the diagonal lines are correspondences.
Now we can show that the pairs G, H and G , H have the same branching
rule (except inverted with respect to inclusion).
Proposition 38.4. Let there be given a see-saw (38.9). Let πiG and πiH be cor-
responding representations of G and H , and let πjG and πjH be corresponding
representations of G and H. Then the multiplicity of πjH in πiG equals the
multiplicity of πiH in πjG .
and similarly
&
πjG |H = d(i, j)πiH .
i∈I
What we must prove is that c(i, j) = d(i, j). Now combining the first equation
in (38.10) with (38.11) we get
&
ΘH×H = c(i, j)πjH ⊗ πiH
i,j
and similarly
&
ΘH×H = d(i, j)πjH ⊗ πiH .
i,j
Now let us reconsider the equivalence of (ii) and (iii) in Theorem 38.3.
This may be understood as a refllection of the following see-saw:
The left vertical line is the diagonal embedding GL(n) −→ GL(n) × GL(n),
and the right vertical line is the Levi embedding GL(p)×GL(q) −→ GL(p+q).
The ambient
@ group is GL(Ω) where Ω = Cn ⊗ Cp+q acting on the symmetric
algebra Ω. More specifically, with H = GL(n) and G = GL(p + q), H × G
acts on Cn ⊗ Cp+q in the obvious way; with G = GL(n) × GL(n) and H =
GL(p) × GL(q) we use the isomorphism
Cn ⊗ Cp+q ∼
= (Cn ⊗ Cp ) ⊕ (Cn ⊗ Cq )
with the first GL(n) and GL(p) acting on the first component, and the second
GL(n) and GL(q) on the second component. Proposition 38.4 asserts that the
two branching rules are the same, which is the equivalence of (ii) and (iii) in
Theorem 38.3.
The paper of Howe, Tan, and Willenbring [79] gives many more examples
of see-saws applied to branching rules. Kudla [113] showed that many con-
structions in the theory of automorphic forms could be explained by see-saws.
Branching rules are important for many problems and are the subject
of considerable literature. Branching rules for the orthogonal and symplectic
404 38 The Cauchy Identity
groups are discussed in Goodman and Wallach [56], Chap. 8. King [101] is a
useful survey of branching rules for classical groups. Many branching rules are
programmed into Sage.
Exercises
Exercise 38.1. Let n and m be integers. Define a bijection between partitions λ =
(λ1 , . . . , λn ) with λ1 m and partitions μ = (μ1 , . . . , μm ) with μ1 n as follows.
The shapes of the partitions λ and μ must sit as complementary pieces in an n × m
box, with the λi being the lengths of the rows of one piece, and the μj being the
lengths of the columns of the other. For example, suppose n = 3 and μ = 5 we could
have λ = (4, 2, 1) and μ = (3, 2, 2, 1). As usual, a partition may be padded with
zeros, so we identify this μ with (3, 2, 2, 1, 0), and the diagram is as follows:
λ
μ
n
m
(xi + yj ) = sλ (x1 , . . . , xn ) sμ (y1 , . . . , ym ),
i=1 j=1
Exercise 38.2. Give another proof of Proposition 38.1 as follows. Show that
(1 − αi t)−1 = zλ pλ (α1 , . . . , αn )t|λ| (38.12)
i λ
The next two exercises lead to another proof of the Cauchy identity.
Exercise 38.4. (i) Show that every matrix coefficient of U(n) extends uniquely to
a regular function on GL(n, C),
so the ring of matrix coefficients on U(n) may
be identified with O GL(n, C . Deduce
that the ring of matrix coefficients of
U(n) may be identified with the O GL(n, C) . Let GL(n, C) × GL(n, C) act on
functions on either GL(n, C) or Matn (C) by
Exercise 38.5. (i) Let α = (α1 , α2 , . . .), β = (β1 , β2 , . . .), γ = (γ1 , γ2 , . . .), δ =
(δ1 , δ2 , . . .) be three sets of variables. Using (38.8), evaluate ν sν (α, β) sν (γ, δ)
in two different ways and obtain the identity
ν ν λ μ θ τ
cλμ cθτ = cφψ cξη cφξ cψη . (38.13)
ν φ,ψ,ξ,η
406 38 The Cauchy Identity
(ii) Show that (38.13) implies the Hopf axiom, that is, the commutativity of the
diagram in Exercise 35.4.
Let α = (α1 , α2 , . . .), β = (β1 , β2 , . . .) be two sets of variables. Define the sup-
ersymmetric Schur polynomial (Littlewood [120] (pages 66–70), Berele and Remmel
[15], Macdonald [123], Bump and Gamburd [29]) by the formula
λ
sλ (α/β) = cμν sμ (α)sν t (β)
μ,ν
In this chapter, we will work not with GL(n, C) but with its compact subgroup
U(n). As in the previous chapters, we will consider elements of Rk as gener-
(n)
alized characters on Sk . If f ∈ Rk , then f = ch(n) (f ) ∈ Λk is a symmetric
polynomial in n variables, homogeneous of weight k. Then ψf : U(n) −→ C,
defined by (33.6), is the function on U(n) obtained by applying f to the
eigenvalues of g ∈ U(n). We will denote ψf = Ch(n) (f ). Thus, Ch(n) maps the
additive group of generalized characters on Sk to the additive group of gener-
alized characters on U(n). It extends by linearity to a map from the Hilbert
space of class functions on Sk to the Hilbert space of class functions on U (n).
Proposition 39.1. Let f be a class function on Sk . Write f = λ cλ sλ , where
the sum is over the partitions of k. Then
|f |2 = |cλ |2 , |Ch(n) (f )|2 = |cλ |2 .
λ l(λ)n
Proof. The sλ are orthonormal by Schur orthogonality, so |f |2 = |cλ |2 .
(n)
By Theorem 36.2, Ch (sλ ) are distinct irreducible characters when λ runs
through the partitions of k with length n, while, by Proposition 36.5,
Ch(n) (sλ ) = 0 if l(λ) > n. Therefore, we may write
Ch(n) (f ) = cλ Ch(n) (sλ ),
l(λ)n
and the Ch(n) (sλ ) in this decomposition are orthonormal by Schur orthogo-
nality on U(n). Thus, |Ch(n) (f )|2 = l(λ)n |cλ |2 .
1 −(x2 +y2 )
dμ(z) = e dx ∧ dy, z = x + iy. (39.2)
π
Let us consider how surprising this is! As n varies, the number of eigen-
values increases and one might expect the standard deviation of the traces to
increase with n. This is what would happen were the eigenvalues of a random
symmetric matrix uncorrelated. That it converges to a fixed Gaussian mea-
sure means that the eigenvalues of a random unitary matrix are quite evenly
distributed around the circle.
Intuitively, the eigenvalues “repel” and tend not to lie too close together.
This is reflected in the property of the trace—that its distribution does not
39 Random Matrix Theory 409
100
R2(1, θ)
θ π
or in other words
lim φ tr(g) dg = φ(x + iy) dμ(z). (39.4)
n−→∞ U(n) C
while
|tr(g)|2k dg k! ,
U(n)
Of course, the sum of the squares of the degrees of the irreducible representa-
tions of Sk is |Sk | = k!, and (39.5) is proved. If k >
n, then the same method
can be used to evaluate the trace, and we obtain λ d2λ , where now the sum
is restricted to partitions of length n. This is < k! .
Theorem 39.2. Suppose that φ(z) is a polynomial in z and z of degree 2n.
Then
φ tr(g) dg = φ(z) dμ(z), (39.6)
U(n) C
Proof omitted. A proof may be found in Billingsley [18], Theorem 26.3 (when
N = 1) and Sect. 28 (for general N ). The precise statement we need is on
p. 383 before Theorem 29.4.
We have
- ∞ k+l k k l l
∞
.
i z w z w
μ̂n (w) = dμn (z)
C k! l!
k=0 l=0
∞
∞ k+l !
i
= z z dμn (z) wk w l .
k l
k! l! C
k=0 l=0
The interchange of the summation and the integration is justified since the
measure dμn is compactly supported, and the series is uniformly convergent
when z is restricted to a compact set. By Proposition 39.2 and the definition
(39.1) of μn , the integral inside brackets vanishes unless k = l, so
∞
(−1)k 2k
μ̂n (w) = Fn (|w|), Fn (r) = ak,n r ,
k!
k=0
1
ak,n = |z| dμn (z).
2k
k! C
by the Plancherel formula and, since we have proved that μ̂n −→ μ uniformly
on compact sets (39.4) is clear for such φ.
Diaconis and Shahshahani [42] proved a much stronger statement to the
effect that the quantities
tr(g), tr(g 2 ), . . . , tr(g r ),
where g is a Haar random element of U(n), are distributed like the moments
of r independent Gaussian random variables. Strikingly, what the proof req-
uires is the full representation theory of the symmetric group in the form of
Theorem 37.1!
39 Random Matrix Theory 413
since pλ = pλ1 · · · pλr , and applying pλi to the eigenvalues of g gives tr(g λi ).
The left-hand side of (39.7) is thus the L2 norm of Ch(n) , and if k n,
then by Theorem 39.1 we may compute this L2 norm in Sk . It equals
1
|pλ (σ)|2 = zλ
|Sk |
σ∈Sk
by (37.2). This is the right-hand side of (39.7). If k > n, the proof is identical
except that Theorem 39.1 only gives an inequality in (39.7).
Proof. Indeed, this follows along the lines of Corollary 39.1 using the fact that
the moments of the measure (39.8)
7r 7r
1 −π|zj |2 /j
|z1 |2k1 |z2 |2k2 · · · |zr |2kr πe dxj ∧ dyj = j kj kj ! ,
C j=1
j j=1
dual of GL(n, C)/U(n) is just U(n). Haar measure makes this symmetric space
into the circular unitary ensemble (CUE). The ensemble is called circular
because the eigenvalues of a unitary matrix lie on the unit circle instead of
the real line. It is the CUE that we have studied in this chapter. Note that
in the GUE, we cannot use Haar measure to make GL(n, C)/U(n) into a
measure space, since we want a probability measure on each ensemble, but
the noncompact group GL(n, C) has infinite volume. This is an important
advantage of the CUE over the GUE. And as Dyson observed, as far as the
local statistics of random matrices are concerned—for examples, with matters
of spacing of eigenvalues—the circular ensembles are faithful mirrors of the
Gaussian ones. The circular orthogonal and symplectic ensembles (COE and
CSE) are similarly the measure spaces U(n)/O(n) and U(2n)/Sp(2n) with
their unique invariant probability measures.
In recent years, random matrix theory has found a new application in the
study of the zeros of the Riemann zeta function and similar arithmetic data.
The observation that the distribution of the zeros of the Riemann zeta function
should have a local distribution similar to that of the eigenvalues of a random
Hermitian matrix in the GUE originated in a conversation between Dyson
and Montgomery, and was confirmed numerically by Odlyzko and others; see
Rubinstein [139]. See Katz and Sarnak [94] and Conrey [38] for surveys of this
field, Keating and Snaith [99] for a typical paper from the extensive literature.
The paper of Keating and Snaith is important because it marked a paradigm
shift away from the study of the spacing of the zeros of ζ(s) to the distribution
of the values of ζ( 12 +it), which are, in the new paradigm, related to the values
of the characteristic polynomial of a random matrix.
This was proved by Keating and Snaith using the Selberg integral. How-
ever an alternative proof was found by Alex Gamburd (see Bump and Gam-
burd [29]) which we will give here. This proof is similar to that of Theorem 39.4
in that we will transfer the computation from U(n) to another group. Whereas
in Theorem 39.4 we used the Frobenius–Schur duality to transfer the com-
putation to the symmetric group Sk , here we will use the Cauchy identity
to transfer the computation from U(n) to U(2k). The two procedures are
extremely analogous and closely related to each other.
7
k 7
n
(1 + ti αj )(1 + ui αj ) det(g)−k
i=1 j=1
k
= sλ (α1 , . . . , αn ) sλt (t1 , . . . , tn , u1 , . . . , un ) det(g) .
λ
Exercises
Let m n. The m-level correlation function of Dyson [48] for unitary statistics
is a function Rm on Tm defined by the requirement that if f is a test function on
Tm (piecewise continuous, let us say) then
∗
Rm (t1 , . . . , tm ) f (t1 , . . . , tm ) dt1 · · · dtm = f (ti1 , . . . , tim ) dg, (39.11)
Tm U(n)
where the sum is over all distinct m-tuples (i1 , . . . , im ) of distinct integers between
1 and n, and t1 , . . . , tn are the eigenvalues of g. Intuitively, this function gives the
probability density that t1 , . . . , tn are the eigenvalues of g ∈ U(n).
The purpose of the exercises is to prove (and generalize) Dyson’s formula
Rm (t1 , . . . , tm ) = det sn (θj − θk ) j,k , ti = eiθj , (39.12)
where *
sin(nθ/2)
sin(θ/2)
if θ
= 0,
sn (θ) =
n if θ = 0.
As a special case, when m = 2, the graph of the “pair correlation” R2 (1, θ) may be
found in Fig. 39.1. This shows graphically the repulsion of the zeros – as we can see,
the probability of two zeros being close together is small, but for moderate distances
there is no correlation.
39 Random Matrix Theory 417
[Since n = m, the matrix A is square and we have det(A · t Ā) = | det(A)|2 . Reduce
to the case where the test function f is symmetric. Then use the Weyl integration
formula.]
Observe that if m < n, then A is not square, so we may no longer factor the
determinant. Deduce Dyson’s formula (39.12).
Exercise 39.4. (Bump, Diaconis and Keller [30]) Generalize Dyson’s formula
as follows. Let λ be a partition of length n. The measure |χλ (g)|2 dg is a probability
measure, and we may define an m-level correlation function for it exactly as in
(39.11). Denote this as Rm,λ . Prove that
⎛ −λ1 1−λ2 ⎞
t1 t1 · · · t1−λn +n−1
⎜ t2 1 t2 2 · · · t2 n
−λ 1−λ −λ +n−1 ⎟
⎜ ⎟
Rm,λ (t1 , . . . , tn ) = det(A · t Ā), A=⎜ . . ⎟.
⎝ .. .. ⎠
−λ1 1−λ2 −λn +n−1
tm tm · · · tm
Exercise 39.5. Let us consider the distribution of the traces of g ∈ SU(2). In this
cases the traces are real valued so we must modify (39.1) to read
φ tr(g) dg = φ(x) dμ(x).
SU(2) R
Since |tr(g)| 2, and since the map g → −g takes SU(2) to itself, the measure dμ
will be even and supported between −2 and 2. Show that
∞ +
1 1 2k
4 − x2 x2k dx =
2π −∞ k+1 k
and deduce that
1 +
dμ(x) = 4 − x2 dx.
2π
40
Symmetric Group Branching Rules
and Tableaux
Proof. Statements (i) and (ii) are equivalent by Frobenius reciprocity. Noting
that S1 is the trivial group, we have Sk−1 = Sk−1 × S1 . By definition, sμ e1
is the character of Sk induced from the character sμ ⊗ e1 of Sk−1 × S1 . With
this in mind, this theorem is just a paraphrase of Proposition 40.1.
1 2 3 1 2 4 1 2 5 1 3 5 1 3 4
4 5 3 5 3 4 2 4 2 5
Sk ⊃ Sk−1 ⊃ · · · ⊃ S1 .
These have the remarkable property that the restriction of each irreducible
representation of Si to Si−1 is multiplicity-free and the branching rule is
explicitly known. Although this is a rare phenomenon, there are a couple
of other important cases:
and
O(n) ⊃ O(n − 1) ⊃ · · · ⊃ O(2).
Proof. Removing the top box (labeled k) from a tableau of shape λ results
in another tableau, of shape μ (say), where μ ⊂ λ. Thus, the set of tableaux
of shape λ is in bijection with the set of tableaux of shape μ, where μ runs
through the partitions of k − 1 contained in λ.
The restriction of ρλ to Sk−1 is the direct sum of the irreducible repre-
sentations ρμ , where μ runs through the partitions of k − 1 contained in λ,
and by induction the degree of each such ρμ equals the number of tableaux of
shape μ. The result follows.
422 40 Symmetric Group Branching Rules and Tableaux
9 8 7 4 2
8 7 6 3 1
B 6 5 4 1
4 3 2
3 2 1
where the sum is over partitions ν of r + k. The coefficients cνλμ are called
the Littlewood–Richardson coefficients. They are integers since the sν are a
Z-basis of the free Abelian group Rr+k .
Applying ch(n) , we may also write
sλ sμ = cνλμ sν
λ
as a decomposition of Schur polynomials, or χλ χμ = cνλμ χν in terms of the
irreducible characters of U (n) parametrized by λ, μ, and ν. Using the fact
that the sλ are orthonormal, we have also
cνλμ = sλ sμ , sν .
Proof. Since by Theorems 34.3 and 35.2 applying the involution ι interchanges
er and hr and also interchanges sμ and sλ , the second statement follows from
the first, which we prove.
424 40 Symmetric Group Branching Rules and Tableaux
The proof that sμ er is the sum of the sλ as λ runs through the partitions
of k + r containing μ such that λ\μ is a vertical strip is actually identical
to the proof of Proposition 40.1. Choose n k + r and, applying ch, it is
sufficient to prove the corresponding result for Schur polynomials.
With notations as in that proof, we see that Δer sμ equals the sum of
k
r terms, each of which is obtained by multiplying M , defined by (40.1),
by a monomial xi1 · · · xir , where i1 < · · · < ir . Multiplying M by xi1 · · · xir
amounts to increasing the exponent of xir in the ir th column by one. Thus,
we get Δer sμ if we take M , increase the exponents in r columns each by one,
and then add the resulting kr determinants.
We claim that this gives the same result as taking M , increasing
the expo-
nents in r rows each by one, and then adding the resulting kr determinants.
Indeed, either way, we get the result of taking each monomial occurring in the
expansion of the determinant M , increasing the exponents of exactly r of the
xi each by one, and adding all resulting terms.
Thus, er sμ equals the sum of all terms (36.1) where (λ1 , . . . , λn ) is obtained
from (μ1 , . . . , μn ) by increasing exactly r of the μi by one. Some of these
terms may not be partitions, in which case the determinant in the numerator
of (36.1) will be zero since it will have repeated rows. The terms that remain
will be the partitions of k + r of length such that λ\μ is a vertical strip. These
partitions all have length n because we chose n large enough. Thus, er sμ is
the sum of sλ for these λ, as required.
Exercises
The next problem generalizes Exercise 37.3. If λ and μ are partitions such that
the Young diagram of λ contains that of μ, then the pair (λ, μ) is called a skew shape
and is denoted λ\μ. Its Young diagram is the set-theoretic difference between the
Young diagrams of λ and μ. The skew shape is called a ribbon shape if the diagram
is connected and contains no 2 × 2 squares. For example, if λ = (5, 4, 4, 3, 2) and
μ = (5, 3, 2, 2, 2) then the skew shape λ\μ is a ribbon shape. Its diagram is the
shaded region in the following figure.
If λ\μ is a ribbon shape, we call its height, denoted ht(λ\μ) one less than the
number of rows involved in its Young diagram. In the example, the height is 2.
40 Symmetric Group Branching Rules and Tableaux 425
where the sum is over all partitions λ of k + r such that λ\μ is a ribbon shape.
[Hint: If λ ∈ Zn , let
λ
F (λ) = det(xi j )/Δ,
where Δ is the denominator in (36.1). Thus if ρ = (n − 1, n − 2, . . . , 0), then (36.1)
can be written F (λ + ρ) = sλ . Show that
n
p r sλ = F (λ + ρ + rek ),
k=1
where ek = (0, . . . , 1, . . . , 0) is the kth standard basis vector of Zn . Show that each
term in this sum is either zero or ±sμ where λ\μ is a ribbon shape.]
Exercise 40.2. Since the hk generate the ring R, knowing how to multiply them
gives complete information about the multiplication in R. Thus, Pieri’s formula
contains full information about the Littlewood–Richardson coefficients. This exercise
gives a concrete illustration. Using Pieri’s formula (or the Jacobi–Trudi identity),
check that
h2 h1 − h3 = s(21) .
Use this to show that
Exercise 40.3. Let λ be a partition of k into at most n parts. Prove that the
number of standard tableaux of shape λ is
tr(g)k χλ (g) dg.
U(n)
(i) Show that thisrmultinomial coefficient is the coefficient of tk1 1 · · · tkr r in the
k
expansion of ( i=1 ti ) .
(ii) Prove that if λ is a partition of k into at most n parts, then the number of
standard tableaux of shape λ is
( )
l(w) k
(−1) .
w∈S
λ1 − 1 + w(1), λ2 − 2 + w(2), . . . , λn − n + w(n)
n
the number of standard tableaux with shape λ. (Hint: Use Theorems 37.3
and 40.2.)
(iii) Let λ be a partition of k into at most n parts. Let μ = λ + δ, where δ =
(n − 1, n − 2, . . . , 1, 0). Show that the number of standard tableaux of shape λ is
( )
k!
, (μi − μj ) .
i μi ! i<j
is a polynomial of degree 1
2
n(n − 1) in μ1 , . . . , μn , and that it vanishes when
μi = μj .]
Exercise 40.5. (i) Show that the product of the hooks in the ith row is
μi !
, .
j>i μi − μj
parametrized by μ.
We embed GL(n − 1, C) −→ GL(n, C) by
g
g −→ . (41.1)
1
Proposition 41.1. Suppose that λn and μn−1 are nonnegative, so the integer
sequences λ and μ are partitions. Then λ and μ interlace if and only if λ ⊃ μ
and the skew partition λ\μ is a horizontal strip.
This is obvious if one draws a diagram.
implies that the diagram of λ\μ contains two entries in the μj + 1 column,
namely in the j and j +1 rows, which is a contradiction since λ\μ was assumed
to be a horizontal strip. We have proved that λ and μ interlace. The converse
is similar.
We can now give a combinatorial formula for the degree of the irreducible
representation πλ of GL(n, C), where λ = (λ1 , . . . , λn ) and λ1 · · · λn . A
Gelfand–Tsetlin pattern of degree n consists of n decreasing integer sequences
of lengths n, n − 1, . . . , 1 such that each adjacent pair interlaces. For example,
if the top row is 3, 2, 1, there are eight possible Gelfand–Tsetlin patterns:
3 2 1 3 2 1 3 2 1
3 2 3 2 3 1
3 2 3
3 2 1 3 2 1 3 2 1
3 1 3 1 2 2
2 1 2
3 2 1 3 2 1
2 1 and 2 1
2 1
We continue removing the boxes labeled n − 1, and the resulting shape is the
third row: ⎧ ⎫
⎨5 2 1⎬
1 1 1 4 3
⎩ ⎭
3
Continuing in this way we obtain a Gelfand–Tsetlin pattern. We leave it to
the reader to convince themselves that this is a bijection.
The second version of RSK (Schensted) gives a bijection between the set
of sequences {m1 , . . . , mk } with mi ∈ {1, . . . , n} (called words) and
F
SYT(λ) × SSYT(λ, n). (41.2)
λ a partition of k
l(λ) n
Let us again observe that these sets have the same cardinality, which we may
prove using Frobenius–Schur duality in the form
&
⊗ k Cn ∼
GL(n)
= ρλ ⊗ πλ .
λ a partition of k
l(λ) n
1 1 1 2 2 3
2 3 3
3
We’ve shaded the box where the inserted 1 will go. The 1 bumps the 2 which
is then inserted into the second row:
1 1 1 2 3
2 2 3 3
3
again we’ve shaded the location where the inserted 2 will go. The 3 that is
bumped will then go in the third row. This time it will go at the end:
1 1 1 2 3
2 2 3
3 3
1 1 1 2 3
2 2 3
3 3
Now let us explain the second RSK algorithm mentioned above, which is
the bijection of {1, . . . , n}k with (41.2). We begin with a sequence (m1 , . . . , mk )
and the empty tableau, which we will denote by ∅. We insert the mi one by
one, finally ending up with a tableau
Q = m1 → m2 → · · · → mk → ∅
1 1
1 2 2 1 1 2 1 1 1
2 2 2
2 2
2
1
1 2 3 1 1 3
2 2
1 1 1
2 2 2 1 2 2 1 1 2 1 1 1
3 3 2 3 3
2
2 2
1 3 3
2
1 1
2 2 3 1 2 3 1 1 3 1 1 1
3 3 3
2 1 2 2 1
2 1 2 1
1 1
1
2 3 3 1 3 3
3 3
Fig. 41.1. The crystal with highest weight λ = (3, 1, 0). The weight diagram (see
Chap. 21) is supplied to the right, to orient the reader
sλ (z1 , . . . , zn ) = zwt(P ) (41.4)
P
where zwt(P ) is the product of zi as i runs through the entries in the tableau P .
We will not prove (41.4), which is due to Littlewood, but proofs may be found
in Fulton [53] or Stanley [153].
Crystals have a purely combinatorial tensor product rule that exactly par-
allels the decomposition rule for tensor products of Lie group representations.
That is, if C and C are crystals, a crystal C ⊗ C is defined which is the disjoint
union of “irreducible” crystals, each isomorphic to a crystal of the type Bλ .
If λ, μ and ν are dominant weights, then the number of copies of Bλ in the
decomposition of Bμ ⊗ Bν equals the multiplicity of πλG in πμG ⊗ πνG .
Crystals give an explanation of the RSK algorithm. The point is that
the tensor product operation is closely related to Schensted insertion. Let
B = B(1) . This is Kashiwara’s standard crystal . It looks like this:
1 2 n−1
1 −→ 2 −→ · · · −→ n .
as μ runs through all the partitions with Young diagrams that are obtained
from λ by adding one box. In this isomorphism, if P is a tableau of shape λ,
the element i ⊗ P in B ⊗ Bλ corresponds to the tableaux i → P obtained
41 Unitary Branching Rules and Tableaux 435
Exercises
Exercise 41.1. Illustrate the bijection described in the proof of Proposition 41.2 by
translating the eight Gelfand–Tsetlin patterns with top row {3, 2, 1} into tableaux.
Exercise 41.2. Illustrate the second RSK bijection by showing that the word
(1, 2, 3, 2, 3, 1) corresponds to the tableau Q with recording tableau P where
1 1 2 1 2 4
Q= 2 3 P= 3 5
3 3
Exercise 41.3. Let g be a Lie algebra. Let U = U (g) be its universal enveloping
algebra.
This chapter can be read immediately after Chap. 39. It may also be skipped
without loss of continuity. It gives further examples of how Frobenius–Schur
duality can be used to give information about problems related to random
matrix theory.
∞
Let f (t) = n=−∞ dn tn be a Laurent series representing a function f :
T −→ C on the unit circle. We consider the Toeplitz matrix
⎛ ⎞
d0 d1 · · · dn−1
⎜ d−1 d0 · · · dn−2 ⎟
⎜ ⎟
Tn−1 (f ) = ⎜ .. .. .. ⎟ .
⎝ . . . ⎠
d1−n d2−n · · · d0
Szegö [157] considered the asymptotics of Dn−1 (f ) = det Tn−1 (f ) as
n −→ ∞. He proved, under certain assumptions, that if
8 ∞
9
n
f (t) = exp cn t ,
−∞
then 8 9
∞
Dn−1 (f ) ∼ exp nc0 + kck c−k . (42.1)
k=1
One may form a minor of a Toeplitz matrix by either striking some rows
and columns or by shifting some rows and columns. For example, if we strike
the second row and first column of T4 (f ), we get
⎛ ⎞
d1 d2 d3 d4
⎜ d−1 d0 d1 d2 ⎟
⎜ ⎟
⎝ d−2 d−1 d0 d1 ⎠ .
d−3 d−2 d−1 d0
This is the same result as we would get by simply shifting the indices in T3 (f );
that is, it is the determinant det(dλi −i+j )1i,j4 where λ is the partition (1).
The most general Toeplitz minor has the form det(dλi −μj −i+j ), where λ and
μ are partitions. The asymptotic formula of Bump and Diaconis holds λ and
μ fixed and lets n −→ ∞.
The formula with μ omitted [i.e., for det(dλi −i+j )] is somewhat simpler to
state than the formula, involving Laguerre polynomials, with both λ and μ.
We will content ourselves with the special case where μ is trivial.
We will take the opportunity in the proof of Theorem 42.1 to correct a
minor error in [28]. The statement before (3.4) of [28] that “. . . the only terms
that survive have αk = βk ” is only correct for terms of degree n. (We thank
Barry Simon for pointing this out.)
If λ is a partition, let χλ denote the character of U(n) defined in Chap. 36.
We will use the notation like that at the end of Chap. 22, which we review
next. Although we hark back to Chap. 22 for our notation, the only “deep”
fact that we need from Part II of this book is the Weyl integration formula. For
example, the Weyl character formula in the form that we need it is identical
to the combination of (36.1) and (36.3). The proof of Theorem 42.1 in [28],
based on the Jacobi–Trudi and Cauchy identities, did not make use of the Weyl
integration formula, so even this aspect of the proof can be made independent
of Part II.
Let T be the diagonal torus in U(n). We will identify X ∗ (T ) ∼ = Zn by
∗
mapping the character (22.15) to (k1 , . . . , kn ). If χ ∈ X (T ) we will use the
“multiplicative” notation eχ for χ so as to be able to form linear combinations
of characters yet still write X ∗ (T ) additively. The Weyl group W can be
identified with the symmetric group Sn acting on X ∗ (T ) = Zn by permuting
the characters. Let E be the free Abelian group on X ∗ (T ). (This differs slightly
from the use of E at the end of Chap. 22.)
Elements of E are naturally functions on T . Since each conjugacy class of
U(n) has a representative in T , and two elements of T are conjugate in G if
and only if they are equivalent by W , class functions on G are the same as
W -invariant functions on W . In particular, a W -invariant element of E may
be regarded as a function on the group. We write the Weyl character formula
in the form (22.17) with δ = (n − 1, n − 2, . . . , 1, 0) as in (22.16).
If λ and μ are partitions of length n, let
λ,μ
Dn−1 (f ) = det(dλi −μj −i+j ).
It is easy to see that this is a minor in a larger Toeplitz matrix.
42 Minors of Toeplitz Matrices 439
Proof. By the Weyl integration formula in the form (22.18), and the Weyl
character formula in the form (22.17), we have
Φn,f (g)χλ (g)χμ (g) dg
U(n)
8 98 9
1
l(w) w(μ+δ) l(w ) −w (λ+δ)
= Φn,f (t) (−1) e (−1) e dt
n! T w∈W w ∈W
⎛ ⎞
1 l(w)+l(w ) w(μ+δ)−w (λ+δ) ⎠
= Φn,f (t) ⎝ (−1) e dt.
n! T w,w ∈W
Each w contributes equally, and we may simply drop the summation over w
and the 1/n! to get
> ?
Φn,f (t) (−1)l(w) ew(μ+δ)−λ−δ dt.
w∈W T
Since the Weyl group is Sn and (−1)l(w) is the sign character, by the definition
λ,μ
of the determinant, this is the determinant Dn−1 (f ).
As we already mentioned, we will only consider here the special case where
μ is (0, . . . , 0). We refer to [28] for the general case. If μ is trivial, then
Theorem 42.1 reduces to the formula
λ
Dn−1 (f ) = Φn,f (g) χλ (g) dg, (42.2)
U(n)
where
λ
Dn−1 (f ) = det(dλi −i+j ).
which is enough tojustify2 all of the following manipulations. (We will use the
assumption that |kck | < ∞ later.)
First, take λ to be trivial, so that m = 0. This special case is Szegö’s
original theorem. By (42.2),
8 9
7
k
Dn−1 (f ) = exp ck tr(g ) dg = exp ck tr(g k ) dg .
U(n) k U(n) k
We can pull out the factor exp(nc0 ) since tr(1) = n, substitute the series
expansion for the exponential function, and group together the contributions
for k and −k. We get
- ∞ .- ∞ β .
7 cα k c−k k
βk
nc0 k k αk k
e tr(g ) tr(g ) dg
U(n) k α !
αk =0 k αk =0 k
β !
8 98 9
7 cαk 7 cβ−kk
β k
= enc0 k
tr(g k )αk tr(g k ) dg,
U(n) αk ! βk !
(αk ) (βk ) k k
This means that the integral is zero unless kαk = kβk . Assuming this, we
look more closely at these terms.
/ By Theorem 37.1, in notation introduced in
Chap. 39,
the function
g −
→ k tr(g k αk
) is Ch (n)
(p ν ), where ν is a partition
of r = kαk = kβk with αk = αk (ν) parts of size k, and similarly we will
denote by σ the partition of r with βk parts of size k. This point was discussed
in the last chapter in connection with (39.7). We therefore obtain
∞
λ
Dn−1 (f ) = enc0 C(r, n),
r=0
where
8 9
7 cαk βk G H
k c−k
C(r, n) = Ch(n) (pν ), Ch(n) (pσ ) .
αk !βk !
ν, σ partitions of r k
(This is the same fact we used in the proof of Proposition 39.4.) Thus, when
r n, we have C(r, n) = C(r) where, using the explicit form (37.1) of zν ,
we have
8 9
7 (ck c−k )αk
nc0
C(r) = e zν
(αk !)2
ν a partition of r k
8 9
7 (kck c−k )αk
nc0
=e .
αk !
ν a partition of r k
Now
∞
7 7
nc0 (kck c−k )αk
C(r) = e = enc0 exp(kck c−k ),
r α =0
αk !
k k k
so as n −→ ∞, the series r C(r, n) stabilizes to the series r C(r) that
converges to the right-hand side of (42.1).
To prove (42.1), we must bound the tails of the series r C(r, n). It is
enough to show that there exists an absolutely convergent series r |D(r)| <
∞ such that |C(r, n)| |D(r)|. First, let us consider the case where ck = c−k .
In thiscase, we may take D(r) = C(r). The absolute convergence of the
series |D(r)| follows from our assumption that |k| |ck |2 < ∞ and the
Cauchy–Schwarz inequality. In this case,
I 8 9 I2
I I
I 7 cαk I
I
C(r, n) = I k
Ch (pν )I
(n)
α ! I ,
Iν a partition of r k k I
where now the inner product is taken in Sr , and of course this is C(r). If we do
not assume ck = c−k , we may use the Cauchy–Schwarz inequality and bound
C(r, n) by
I 8 9 II 8 9 I
I 7 cαk II 7 cβ−k I
I I I k
I
I k I I
Ch (pν )I·I
(n)
Ch (pσ )I
(n)
I I.
Iν a partition of r k αk ! I Iσ a partition of r k βk ! I
Now (42.1) is proved, which is the special case with λ trivial. We turn now
to the general case.
We will make use of the identity
sλ = zμ−1 sλ (ξμ ) pμ
μ a partition of m
For each μ, let γk (ξμ ) be the number of cycles of length k in the decomposition
of ξμ into a product of disjoint cycles. By Theorem 37.1, we may write this
identity
7
χλ = zμ−1 sλ (ξμ ) tr(g k )γk (ξμ ) .
μ a partition of m k
λ
Now, proceeding as before from (42.2), we see that Dn−1 (f ) equals
nc0 −1
e zμ sλ (ξμ )
μ a partition of m
8 ∞
9⎛ ∞
⎞
7 cα k cβ−k
k
βk +γk (ξμ )
× k
tr(g k )αk ⎝ tr(g k ) ⎠ dg.
U(n) k α =0
α k ! βk !
k βk =0
Since Sm contains m!/zλ elements of cycle type μ and sλ has the same value
sλ (ξμ ) on all of them, we may write this as
1
enc0 sλ (ξ)
m!
ξ∈Sm
8 98 9
7 cαk 7 cβ−kk
βk +γk (ξ)
× k k αk
tr(g ) k
tr(g ) dg.
αk ! βk !
(αk ) (βk ) U(n) k k
As in the previous case, the contribution vanishes unless kαk = kβk +m,
and we assume this. We get
∞
λ 1
Dn−1 (f ) = enc0 sλ (ξ) C(r, n, ξ),
m! r=0
ξ∈Sm
where now
C(r, n, ξ)
( )( )
cαk cβ−k
k
βk +γk (ξ)
k k αk
= tr(g ) tr(g k ) dg.
U(n) αk ! βk !
(αk ) k k
(βk )
kαk = r + m kβk = r
444 42 Minors of Toeplitz Matrices
and using this value, we see that when r n we have C(r, n, ξ) = C(r, ξ),
where
(kck c−k )βk
C(r, ξ) = Δ(f, ξ) .
βk !
(βk )
kβk = r
The series is
8 ∞
9
C(r, ξ) = Δ(f, ξ) exp nc0 + kck c−k ,
r k=1
result will follow as before if we can show that |C(r, n, ξ)| < |D(r, ξ)|
so the
where |D(r, ξ)| < ∞. The method is the same as before, based on the fact
that the characteristic map is a contraction, and we leave it to the reader.
Exercises
Exercise 42.1 (Bump et al. [30]).
(i) If f is a continuous function on T, show that there is a well-defined continuous
function uf : U(n) −→ U(n) such that if ti ∈ T and h ∈ U(n), we have
⎛ ⎛ ⎞ ⎞ ⎛ ⎞
t1 f (t1 )
⎜ ⎜ ⎟ ⎟ ⎜ ⎟ −1
uf ⎝h ⎝ . . . ⎠ h−1 ⎠ = h ⎝ ..
. ⎠h .
tn f (tn )
n
(ii) If g is an n × n matrix, with n m, let Em (g) denote the sum of the m
principal m × m minors of g. Thus, if n = 4, then E2 (g) is
g11 g12 g11 g13 g11 g14 g22 g23 g22 g24 g33 g34
+ + + + +
g21 g22 g31 g33 g41 g44 g32 g33 g42 g44 g43 g44 .
Prove that if f (t) = dk tk , then
μ,λ
Em (uf (g)) χλ (g) χμ (g) dg = Em (Tn−1 ),
U(n)
μ,λ
where Tn−1 is the n × n matrix whose i, jth entry is dλi −μj −i+j . (Hint: Deduce
this from the special case m = n.)
43
The Involution Model for Sk
Now (v, w) → B(w, v) has the same property as B, and so B(w, v) = B(v, w)
for some constant . Applying this identity twice, 2 B(v, w) = B(v, w) so
= ±1.
If (π, V ) is self-contragredient, let π be the constant in Proposition 43.1;
otherwise let π = 0. If π = 1, then we say that π is orthogonal or real ; if
π = −1, we say that π is symplectic or quaternionic. We call π the Frobenius–
Schur indicator of π.
Theorem 43.1 (Frobenius and Schur). Let (π, V ) be an irreducible rep-
resentation of the compact group G. Then
π = χ(g 2 ) dg.
G
= h2 (x1 , . . . , xn ) − e2 (x1 , . . . , xn ) .
By (33.8) and Proposition 33.2, this means that
χ(g 2 ) = tr ∨2 π(g) − tr ∧2 π(g) .
We see that π is
tr ∨2 π(g) dg − tr ∧2 π(g) dg.
G G
Thus, what we need to know is that ∨2 π(g) contains the trivial representation
if and only if π = 1, while ∧2 π(g) contains the trivial representation if and
only if π = −1.
If ∨2 π(g) contains the trivial representation, let ξ ∈ ∨2 V be a ∨2 π(g)-
fixed vector. Let , be a G-invariant inner product on V . There is induced
a G-invariant Hermitian inner product on ∨2 V such that v1 ∨ v2 , w1 ∨ w2 =
v1 , v2 w1 , w2 , and we may define a symmetric bilinear form on V by
B(v, w) = v ∨ w, ξ. Thus, π = 1.
Conversely, if π = 1, let B be a symmetric invariant bilinear form. By the
universal property of the symmetric square, there exists a linear form L :
∨2 V −→ C such that B(v, w) = L(v ∨ w), and hence a vector ξ ∈ ∨2 V such
that B(v, w) = v ∨ w, ξ, which is a ∨2 π(g)-fixed vector.
The case where π = −1 is identical using the exterior square.
Proposition 43.2. Let (π, V ) be an irreducible complex representation of the
compact group G. Then π is the complexification of a real representation if
and only if π = 1. If this is true, π(G) is conjugate to a subgroup of the
orthogonal group O(n).
43 The Involution Model for Sk 447
Exercises
The first exercise generalizes Theorem 43.1 of Frobenius and Schur. Suppose that
G is a compact group and θ : G −→ G is an involution (i.e,, an automorphism
satisfying θ2 = 1). Let (π, V ) be an irreducible representation of G. If π ∼
= θ π, where
θ
π : V −→ V is the “twisted” representation π(g) = π( g), then by an obvious
θ θ
Exercise 43.1. Assuming the hypotheses of the stated theorem, define a group H
that is the semidirect product of G by a cyclic group t generated by an element t
of order 2 such that tgt−1 = θ g for g ∈ G. Thus, the index [G : H] = 2. The idea
is to use Theorem 43.1 for the group H to obtain the theorem of Kawanaka and
Matsuyama for G. Proceed as follows.
Case 1: Assume that π ∼ = θ π. In this case, show that there exists an endomor-
phism T : V −→ V such that T ◦ π(g) = π(θ g) ◦ T and T 2 = 1V . Extend π to a
representation πH of H such that πH (t) = T . Let Bθ : V × V −→ C satisfy (43.3).
Then B(v, w) = Bθ (v, T w) satisfies (43.1), as does B(T v, T w) = Bθ (T v, w). Thus,
there exists a constant δ such that B(T v, T w) = δB(v, w). Show that δ 2 = 1 and
that
θ (π) = δ(π) . (43.5)
Apply Theorem 43.1 to the representation πH , bearing in mind that the Haar mea-
sure on H restricted to G is only half the Haar measure on G because both measures
are normalized to have total volume 1. This gives
1
(πH ) = (π) + χ(g · θ g) dg . (43.6)
2 G
Show using direct constructions with bilinear forms on V and V H that if either (π)
or θ (π) is nonzero, then πH is self-contragredient, while if πH is self-contragredient,
then exactly one of (π) or θ (π) is nonzero, and whichever one is nonzero equals
(πH ).
43 The Involution Model for Sk 453
For example, if G = GL(n, Fq ), it was shown independently by Gow [57] and Kly-
achko [103] that the conclusions to Exercise 43.2 are satisfied when G = GL(n, Fq )
and θ is the automorphism g −→ t g −1 .
For the next group of exercises, the group B2k is a Coxeter group with generators
and (2k − 1, 2k). It is thus a Weyl group of Cartan type Bk with order k!2k . It has
a linear character ξ2k having value −1 on these “simple reflections.” This is the
S
character (−1)l(w) of Proposition 20.12. Let η2k = IndB2k2k
(ξ2k ) be the character of
S2k induced from this linear character of B2k . The goal of this exercise will be to
prove analogs of Theorem 43.4 and the other results of this chapter for η2k .
Exercise 43.3. Prove the analog of Proposition 43.3. That is, show that inducing
the restriction of η2r to S2r−1 is isomorphic to the character of S2r−1 induced from
the character η2r−2 to S2r−1 .
Let S2k be the set of characters sλ of S2k where λ is a partition of 2k such that if
μ is the conjugate partition, then μi = λi + 1 for all i such that λi i. For example,
the partition λ = (5, 5, 4, 3, 3, 2) has conjugate (6, 6, 5, 3, 2), and the hypothesis is
satisfied. Visually, this assumption means that the diagram of λ can be assembled
from two congruent pieces, as in Fig. 43.1. We will describe these as the “top piece”
and the “bottom piece,” respectively.
Top Piece
Bottom
Piece
Let T2k+1 be the set of partitions of 2k + 1 with a diagram that contains an element
of S2k .
Exercise 43.4. Prove that if λ ∈ T2k+1 , then there are unique partitions μ ∈ S2k
and ν ∈ S2k+2 such that the diagram of λ contains the diagram of μ and is contained
in the diagram of ν. (Hint: The diagrams of the skew partitions λ − μ and ν − λ,
each consisting of a single node, must be corresponding nodes of the top piece and
bottom piece.)
454 43 The Involution Model for Sk
Exercise 43.6. Show that η2k is multiplicity-free and that the representations
occurring in it are precisely the sλ with λ ∈ S2k .
44
Some Symmetric Algebras
The results of the last chapter can be translated into statements about the
representation theory of U(n). For example, we will see that each irreducible
representation of U(n) occurs exactly once in the decomposition of the sym-
metric algebra of V ⊕ ∧2 V , where V = Cn is the standard module of U(n).
The results of this chapter are also proved in Goodman and Wallach [56],
Howe [77], Littlewood [120], and Macdonald [124]. See Theorem 26.6 and the
exercises to Chap. 26 for alternative proofs of some of these results.
Let us recall some ideas that already
@ appeared
( in Chap. 38. If ρ : G −→
GL(V ) is a representation, then V and V become modules for G and
we may ask for their decomposition into irreducible representations of V . For
some representations ρ, this question will have a simple answer, and for others
the answer will be complex. The very simplest case is where G = GL(V ).
In this case, each ∨k V is itself irreducible, and each ∧k V is either irreducible
(if k < dim(V )) or zero.
We can encode the solution to this question with generating functions
∞
∞
Pρ∨ (g; t) = tr g| ∨k V tk , Pρ∧ (g; t) = tr g ∧k V tk .
k=0 k=0
We see that for all g, Pρ∨ (g, t) is convergent if t < max(|γi |−1 ) and has
meromorphic continuation in t, while Pρ∧ (g, t) is a polynomial in t of degree
equal to the dimension of V . We will denote Pρ∨ (g) = Pρ∨ (g, 1) and Pρ∧ (g) =
Pρ∧ (g, 1). Then we specialize t = 1 in (44.1) and write
7 7
Pρ∨ (g) = (1 − γi )−1 , Pρ∧ (g) = (1 + γi ). (44.2)
i i
as GL(n, C)-modules. It is the direct sum of the πλ as λ runs through all even
partitions of k.
$
Here B2k ⊂ S2k acts on k V on the right by the action (34.1).
First, we note that the map
commutes with the right action of B2k . It is 2k-linear and hence induces a
map
%2k
α: V −→ ∨k (∨2 V ), α(v1 ⊗ · · · ⊗ v2k ) = (v1 ∨ v2 ) ∨ · · · ∨ (v2k−1 ∨ v2k ),
44 Some Symmetric Algebras 457
As for the other direction, we first note that for v3 , v4 , . . . , v2k fixed, using
the fact that ⊗C[B2k ] is B2k -balanced, the map
%
2k
(v1 , v2 ) −→ (v1 ⊗ v2 ⊗ v3 ⊗ · · · ⊗ v2k ) ⊗ 1 ∈ V ⊗C[B2k ] Ctrivial
Using the fact that ⊗C[B2k ] is B2k -balanced, the map μ is symmetric and hence
>$ ?
2k
induces a map ∨k (∨2 V ) −→ V ⊗C[B2k ] Ctrivial that is the inverse of the
map previously constructed. We have now proved (44.3).
458 44 Some Symmetric Algebras
Proof. This follows from Proposition 44.2, Theorem 36.4, and the explicit
decomposition of Theorem 43.4.
- .⎡ ⎤
= e k tk ⎣ sλ t2r ⎦
k λ an even partition of 2r
is
ek−2r (α1 , . . . , αn ) sλ .
2rk λ an even partition of 2r
This is the image of ek−2r ω2r under the characteristic map, and it equals the
sum of the sλ for all partitions of k by Theorem 43.5. Taking t = 1, the result
follows.
44 Some Symmetric Algebras 459
Proposition 44.3. The character ω̃2k is the sum of the sλ , where λ runs
through all the partitions of k such that the conjugate partition λt is even.
Proof. This may be deduced from Theorem 43.4 as follows. Applying this with
G = S2k , H = B2k , and ρ = ε, we see that ω̃2k is the same as ω2k multiplied
by the character ε. By Theorem 37.4, this is ι ω2k , and by Theorems 43.4, and
35.2, this is the sum of the sλ with λt even.
Exercises
Exercise 44.1. Let η2k be the character of S2k from the exercises of the last chapter,
and let S2k be the set of partitions of 2k defined there. Show that
-2k
∧k (∧2 V ) ∼
= V ⊗C[S2k ] η2k ,
(1 + αi αj ) = sλ (α1 , . . . , αn ).
1ijn k t λ∈S
2k
Explain why, in contrast with (44.4) and (44.5), there are only finitely many nonzero
terms on the right-hand side in these identities.
45
Gelfand Pairs
Proof. In the decomposition (45.1), we have EndG (θ) = Matdi (C). This is
commutative if and only if all di 1.
Let G be a group, finite for the time being, and H a subgroup. Then (G, H)
is called a Gelfand pair if the representation of G induced by the trivial repre-
sentation of H is multiplicity-free. We also refer to H as a Gelfand subgroup.
More generally, if π is an irreducible representation of H, then (G, H, π) is
called a Gelfand triple if π G is multiplicity-free. See Gross [59] for a lively
discussion of Gelfand pairs.
From Proposition 45.1, Gelfand pairs are characterized by the commuta-
tivity of the endomorphism ring of an induced representation. To study it, we
make use of Mackey theory.
Proof. Note that, using (32.9), the summand Δ2 (gγ −1 )Δ1 (γ) does not depend
on the choice of representative γ ∈ H2 \G. The result is easily checked.
Theorem 45.1. Let H be a subgroup of the finite group G, and let (π, V ) be
a representation of H. Then (G, H, π) is a Gelfand triple if and only if the
convolution algebra H of functions Δ : G −→ EndC (V ) satisfying
is commutative.
Theorem 45.2. Let H be a subgroup of the finite group G, and suppose that
G admits an involution fixing H such that every double coset of H is invariant:
HgH = H ι g H. Then H is a Gelfand subgroup.
Proof. The ring H of Theorem 45.1 is just the convolution ring of H-bi-
invariant functions on G. We have an involution on this ring:
ι
Δ(g) = Δ(ι g).
On the other hand, each Δ is constant on each double coset, and these are
invariant under ι by hypothesis, so ι is the identity map. This proves that H
is commutative, so (G, H) is a Gelfand pair.
We already know this: the representation of Sn+m induced from the trivial
character of Sn × Sm is the product in the ring R of hn by hm . By Pieri’s
formula, one computes, assuming without loss of generality that n > m,
m
hn h m = s(n+m−k,k) .
k=0
Here In and 0n are the n × n identity and zero matrices, and the remaining
0 matrices are rectangular blocks.
We start with g in block form,
AB
,
CD
where the sizes of the square blocks are indicated by subscripts. The matrices
T , U , V , and W are permutation matrices (invertible). Left multiplication by
element of Sr × Sn−r × Sm−n+r × Sn−r can now replace these four matrices
by identity matrices. This proves that (45.2) is a complete set of double coset
representatives.
Since these double coset representatives are all invariant under the invo-
lution, by Theorem 45.2 it follows that Sn × Sm is a Gelfand subgroup.
Proposition 45.4. Suppose that (G, H, ψ) is a Gelfand triple, and let (π, V )
be an irreducible representation of G. Then there exists at most one space M
of functions on G satisfying
such that M is closed under right translation and such that the representation
of G on M by right translation is isomorphic to π.
Proof. This is just the Frobenius reciprocity. The space of functions satis-
fying (45.3)
H (ψ), so M, if it exists, is the image of an element of
is Ind
G
HomG V, IndG H (ψ) . This is one-dimensional since the induced representation
is assumed to be multiplicity-free.
45 Gelfand Pairs 465
π(φξ,η ) v = 1
dim(V ) v, ξ η. (45.4)
Indeed, taking the inner product of the left-hand side with an arbitrary vector
θ ∈ V , Schur orthogonality (Theorem 2.4) gives
π(φξ,η )v, θ = G π(g) v, θ π(g)ξ, ηdg = dim(V ) v, ξ η, θ ,
1
The image of this is contained in the linear span of η, and taking v = ξ shows
that the map is nonzero. Since H is assumed commutative, this also equals
π(φξ,η ∗ φη,ξ ). Hence, its image is also equal to C ξ, and so we see that ξ and
η both belong to the same one-dimensional subspace of V .
466 45 Gelfand Pairs
This implies that k −1 h−1 g ∈ O(n), so g and k lie in the same double coset.
Now let us think a bit about what this means in concrete terms. The
quotient G/H may be identified with the sphere S n . Indeed, thinking of S n
as the unit sphere in Rn+1 , G acts transitively and H is the stabilizer of a
point in S n .
Consequently, we have an action of G on L2 (S n ), and this may be thought
of as the representation induced from the trivial representation of O(n).
This gives us a concrete model for at least some representations of O(n + 1).
φ(v)(ξ0 ) = v, η .
By (45.7), we have
φ(v) π(g) ξ0 = π(g −1 )v, η = v, π(g)η .
This makes it clear that φ is determined by η, and it also shows that η is O(n)-
invariant since ξ0 ∈ S n is O(n)-fixed. Since the space of O(n)-fixed vectors is
at most one-dimensional, the theorem is proved.
Proposition 45.6. If g ∈ U(n), then there exist k1 and k2 ∈ O(n) such that
k1 gk2 is diagonal.
468 45 Gelfand Pairs
Exercises
Exercise 45.1. Let G be any compact group. Let H = G × G, and embed G into
H diagonally, that is, by the map g −→ (g, g). Use the involution method to prove
that G is a Gelfand subgroup of H.
Exercise 45.2. Use the involution method to show that O(n) is a Gelfand subgroup
of U (n).
Exercise 45.3. Show that each irreducible representation of O(3) has an O(2)-fixed
vector, and deduce that L2 (S 2 ) is the (Hilbert space) direct sum of all irreducible
representations of O(3), each with multiplicity one.
Exercise 45.4 (Gelfand and Graev). Let G = GL(n, Fq ) and let N be the sub-
group of upper triangular unipotent matrices. Let ψ : Fq −→ C× be a nontrivial
additive character. Define a character ψN of N by
⎛ ⎞
1 x12 x13 · · · x1n
⎜ 1 x23 · · · x2n ⎟
⎜ ⎟
⎜ 1 ⎟
ψN ⎜ ⎟ = ψ(x12 + x23 + · · · + xn−1,n ).
⎜ . . ⎟
⎝ . . .. ⎠
1
The object of this exercise is to show that IndN G (ψN ) is multiplicity-free. This
Gelfand–Graev representation is important because it contains most irreducible rep-
resentations of the group; those it contains are therefore called generic. We will
denote by Φ the root system of GL(n, Fq ) and by Φ+ the positive roots αij such
that i < j. Let Σ be the simple positive roots αi,i+1 .
(i) Show that each double coset in N \G/N has a representative m that is a mono-
mial matrix. In the notation of Chap. 27, this means that m ∈ N (T ), where T
is the group of diagonal matrices. (Make use of the Bruhat decomposition.) Let
w ∈ W = N (T )/T be the corresponding Weyl group element.
45 Gelfand Pairs 469
Note that N and its character ψN are invariant under ι. Interpret (iv) as show-
ing that every double coset that supports an intertwining operator Ind(ψN ) −→
Ind(ψN ) has a representative that is invariant under ι, and deduce that
EndG Ind(ψN ) is commutative and that Ind(ψN ) is multiplicity-free.
46
Hecke Algebras
A Coxeter group (Chap. 25) is a group W which may be given the following
description. The group W has generators si (i = 1, 2, . . . , r) with relations
s2i = 1 and for each pair of indices i and j the “braid relations”
si sj si · · · = sj si sj · · ·
where the number of terms on both sides is the same integer n(i, j). An
example is the symmetric group Sk , where si is the transposition (i, i + 1).
In this case r = k − 1.
Given a Coxeter group W , we may deform its group algebra as follows.
Let H(W ) be the ring with generators ti satisfying the same braid relations
ti tj ti · · · = tj ti tj · · · ,
but we replace the relation s2i = 1 by a more general relation
t2i = (q − 1)ti + q.
The parameter q may be a complex number or an indeterminate. If q = 1, we
recover the group algebra of W .
Hecke algebras are ubiquitous. They arise in various seemingly different
ways: as endomorphism rings of induced representations for the groups of
Lie type such as GL(k, Fq ) (Iwahori [84], Howlett and Lehrer [80]); as
convolution rings of functions on p-adic groups (Iwahori and Matsumoto [86]);
as rings of operators acting on the equivariant K-theory of flag varieties
(Lusztig [122], Kazhdan and Lusztig [98]); as rings of transfer matrices in
statistical mechanics and quantum mechanics (Temperley and Lieb [160],
Jimbo [90]), in knot theory (Jones [91]), and other areas. It is the con-
text for defining the Kazhdan–Lusztig polynomials, which occur in seemingly
unrelated questions in representation theory, geometry and combinatorics [97].
Some of these different occurrences of Hecke algebras may seem unrelated to
each other, but this can be an illusion when in fact deep and surprising con-
nections exist.
Following Iwahori [84], we will study a certain “Hecke algebra” Hk (q) that,
as we will see, is isomorphic to the Hecke algebra of the symmetric group Sk .
The ring Hk (q) can actually be defined if q is any complex number, but if q is a
prime power, it has a representation-theoretic interpretation. We will see that
it is the endomorphism ring of the representation of G = GL(k, Fq ), where
Fq is the finite field with q elements, induced from the trivial representation
of the Borel subgroup B of upper triangular matrices in G. The fact that it
is a deformation of C[Sk ] amounts to a parametrization of a certain set of
irreducible representations of G—the so-called unipotent ones—by partitions.
If instead of G = GL(k, Fq ) we take G = GL(k, Qp ), where Qp is the
p-adic field, and we take B to be the Iwahori subgroup consisting of elements
g of K = GL(k, Zp ) that are upper triangular modulo p, then one obtains the
affine Hecke algebra, which is similar to Hk (q) but infinite-dimensional. It was
introduced by Iwahori and Matsumoto [86]. The role of the Bruhat decompo-
sition in the proofs requires a generalization of the Tits’ system described in
Iwahori [85]. This Hecke algebra contains a copy of Hk (p). On the other hand,
it also contains the ring of K-bi-invariant functions, the so-called spherical
Hecke algebra (Satake [143], Tamagawa [158]). The spherical Hecke algebra is
commutative since K is a Gelfand subgroup of G. The spherical Hecke algebra
is (when k = 2) essentially the portion corresponding to the prime p of the
original Hecke algebra introduced by Hecke [65] to explain the appearance
of Euler products as the L-series of automorphic forms. See Howe [76] and
Rogawski [137] for the representation theory of the affine Hecke algebra.
Let F be a field. Let G = GL(k, F ) and, as in Chap. 27, let B be the
Borel subgroup of upper triangular matrices in G. A subgroup P containing
B is called a standard parabolic subgroup. (More generally, any conjugate of a
standard parabolic subgroup is called parabolic.)
Let k1 , . . . , kr be positive integers such that i ki = k. Then Sk has a
subgroup isomorphic to Sk1 ×· · ·×Skr in which the first Sk1 acts on {1, . . . , k1 },
the second Sk2 acts on {k1 + 1, . . . , k1 + k2 }, and so forth. Let Σ denote the
set of k − 1 transpositions {(1, 2), (2, 3), . . . , (k − 1, k)}.
Proof. If J contains (1, 2), (2, 3), . . . , (k1 − 1, k1 ), then the subgroup they
generate is the symmetric group Sk1 acting on {1, . . . , k1 }. Taking k1 as large
as possible, assume that J omits (k1 , k1 + 1). Taking k2 as large as possible
such that J contains (k1 + 1, k1 + 2), . . . , (k1 + k2 − 1, k1 + k2 ), the subgroup
they generate is the symmetric group Sk2 acting on {k1 + 1, . . . , k1 + k2 }, and
so forth. Thus J contains generators of each factor in Sk1 × · · · × Skr and does
not contain any element that is not in this product, so this is the group it
generates.
The notations from Chap. 27 will also be followed. Let T be the maximal
torus of diagonal elements in G, N the normalizer of T , and W = N/T the
46 Hecke Algebras 473
Weyl group. Moreover, Φ will be the set of all roots, Φ+ the positive roots, and
Σ the simple positive roots. Concretely, elements of Φ are the k 2 − k rational
characters of T of the form
⎛ ⎞
t1
⎜ .. ⎟ −1
αij ⎝ . ⎠ = ti tj ,
tn
where 1 i, j n, Φ+ consists of {αij i < j}, and Σ = {αi,i+1 }. Identifying
W with Sk , the set Σ in Lemma 46.1 is then the set of simple reflections.
Let J be any subset of Σ. Let WJ be the subgroup of W generated by the
sα with α ∈ Σ. Then, by Lemma 46.1, we have (for suitable ki )
WJ ∼
= Sk1 × · · · × Skr . (46.1)
Let NJ be the preimage of WJ in N under the canonical projection to W . Let
PJ be the group generated by B and NJ . Then
⎧⎛ ⎞⎫
⎪
⎪ G11 G12 · · · G1r ⎪⎪
⎪
⎨⎜ 0 G22 · · · G2r ⎟⎪ ⎬
⎜ ⎟
PJ = ⎜ . . . ⎟ , (46.2)
⎪
⎪ ⎝ .. . . .. ⎠⎪
⎪
⎪
⎩ ⎪
⎭
0 0 · · · Grr
where each Gij is a ki × kj block. The group PJ is a semidirect product
PJ = MJ UJ = UJ MJ , where MJ is characterized by the condition that
Gij = 0 unless i = j, and the normal subgroup UJ is characterized by
the condition that each Gii is a scalar multiple of the identity matrix in
GL(ki ). The groups PJ with J a proper subset of Σ are called the standard
parabolic subgroups, and more generally any subgroup conjugate to a PJ is
called parabolic. The subgroup UJ is the unipotent radical of PJ (that is, its
maximal normal unipotent subgroup), and MJ is called the standard Levi
subgroup of PJ . Evidently,
MJ ∼
= GL(k1 , F ) × · · · × GL(kr , F ). (46.3)
Any subgroup conjugate in PJ to MJ (which is not normal) would also be
called a Levi subgroup.
As in Chap. 27, we note that a double coset BωB, or more generally PI ωPJ
with I, J ⊂ Σ, does not depend on the choice ω ∈ N of representative for an
element w ∈ W , and we will use the notation BwB = C(w) or PI wPJ for
this double coset. Let BJ = MJ ∩ B. This is the standard “Borel subgroup”
of MJ .
Proposition 46.1.
(i) Let J ⊆ Σ. Then
A
MJ = BJ wBJ (disjoint).
w∈WJ
474 46 Hecke Algebras
Proof. For (i), we have (46.3) for suitable ki . Now BJ is the direct product of
the Borel subgroups of these GL(ki , F ), and WJ is the direct product (46.1).
Part (i) follows directly from the Bruhat decomposition for GL(k, F ) as proved
in Chap. 27.
As for (ii), since BWI ⊂ PI and WJ B ⊂ PJ , we have BWI wWJ B ⊆
PI wPJ . To prove the opposite inclusion, we first note that
wBWJ ⊆ BwWJ B. (46.5)
Indeed, any element of WJ can be written as s1 · · · sr , where si = sαi , with
αi ∈ J. Using Axiom TS3 from Chap. 27, we have
wBs1 · · · sr ⊂ BwBs2 · · · sr B ∪ Bws1 Bs2 · · · sr B
and, by induction on r, both sets on the right are contained in BwWJ B. This
proves (46.5). A similar argument shows that
WI BwWJ ⊆ BWI wWJ B. (46.6)
Now, using (i),
PI wPJ = UI MI wMJ UJ ⊂ UI BI WI BI wBJ WJ BJ UJ ⊂ BWI BwBWJ B.
Applying (46.5) and (46.6), we obtain BWI wWJ B ⊇ PI wPJ , whence (46.4).
As for (iii), since by the Bruhat decomposition w −→ BwB is a bijec-
tion W −→ B\G/B, (46.4) implies that w −→ PI wPJ induces a bijection
WI \W/WJ −→ PI \G/PJ .
To proceed further, we will assume that F = Fq is a finite field. We
recall from Chap. 34 that Rk denotes the free Abelian group generated by the
isomorphism classes of irreducible representations of the symmetric group Sk ,
or, as we sometimes prefer, the additive group of generalized characters. It can
be identified with the character ring of Sk . However, we do not need its ring
structure, only its additive structure and its inner product, in which the dis-
tinct isomorphism classes of irreducible representations form an orthonormal
basis.
Similarly, let Rk (q) be the free Abelian group generated by the isomor-
phism classes of irreducible representations of GL(n, Fq ) or equivalently the
additive group of generalized characters. Like Rk , we can make Rk (q) into
the k-homogeneous part of a graded ring, a point we will take up in the next
chapter.
46 Hecke Algebras 475
Proof. By the geometric form of Mackey’s theorem (Theorem 32.1), the space
of intertwining maps from IndH H
M1 (1) to IndM2 (1) is isomorphic to the space
of functions Δ : H −→ Hom(C, C) ∼ = C that satisfy Δ(m2 hm1 ) = Δ(h) for
mi ∈ Mi . Of course, a function has this property if and only if it is constant
on double cosets, so the dimension of the space of such functions is equal to
the number of double cosets. On the other hand, the dimension of the space
of intertwining operators equals the inner product in the character ring by
(2.7).
where the sum is over subsets of Σ. We need to verify that this is well-defined
and an isometry.
By Proposition 46.1, if I, J ⊆ Σ, the cardinality of WI \W/WJ equals the
cardinality of PI \G/PJ . By Proposition 46.2, it follows that
Expanding this gives a sum of exactly 2k−1 monomials in the hi , which are
in one-to-one correspondence with the subsets J of Σ. Indeed, let J be given,
and let k1 , k2 , k3 , . . . be as in Lemma 46.1. Then there is a monomial that
has |J| 1’s taken from below the diagonal; namely, if αi,i+1 ∈ Σ, then there
is a 1 taken from the i + 1, i position, and there is an hk1 taken from the
1, k1 position, an hk2 taken from the k1 + 1, k1 + k2 position, and so forth.
This monomial equals (−1)|J| hk1 hk2 · · · , which is (−1)|J| times the character
induced from the trivial representation of WJ = Sk1 × Sk2 × · · · .
Theorem 46.2. As a virtual representation, the Steinberg representation
ek (q) of GL(k, Fq ) admits the following expression:
ek (q) = (−1)|J| IndP
PJ (1).
J⊆Σ
Proof. This follows immediately from Proposition 46.3 on applying the map-
ping of Theorem 46.1.
For our next considerations, there is no reason that F needs to be finite,
so we return to the case where G = GL(k, F ) of a general field F . We will
denote by U the group of upper triangular unipotent matrices in GL(k, F ).
Proposition 46.4. Suppose that S is any subset of Φ such that if α ∈ S, then
−α ∈/ S, and if α, β ∈ S and α + β ∈ Φ, then α + β ∈ S. Let US be the set
of g = (gij ) in GL(k, F ) such that gii = 1, and if i
= j, then gij = 0 unless
αij ∈ S. Then US is a group.
Proof. Let S̃ be the set of (i, j) such that the root αij ∈ S. Translating the
hypothesis on S into a statement about S̃, if (i, j) ∈ S̃ we have i < j, and
if both (i, j) and (j, k) are in S̃, then i = k and (i, k) ∈ S̃. (46.9)
From this it is easy to see that if g and h are in US , then so are g −1 and
gh.
As a particular case, if w ∈ W , then S = Φ+ ∩wΦ− satisfies the hypothesis
of Proposition 46.4, and we denote
|Uw− | = q l(w) .
Proof. We will prove this if F is finite, the only case we need. In this case
Uw+ ∩ Uw− = {1} by definition since the sets Φ+ ∩ wΦ− and Φ+ ∩ wΦ+ are
− + − ± ±
1 u1 = u2 u2 with ui ∈ Uw , then (u2 )
disjoint. Thus, if u+ u1 = u−
+ −1 + − −1
2 (u1 ) ∈
− ± ± −
Uw ∩ Uw so u1 = u2 . Therefore, the multiplication map Uw × Uw −→ U is
+ +
We are interested in the size of the double coset BwB. In geometric terms,
G/B can be identified with the space of F -rational points of a projective
algebraic variety, and the closure of BwB/B is an algebraic subvariety in
which BwB/B is an open subset; the dimension of this “Schubert cell” turns
out to be l(w).
If F = Fq , an equally good measure of the size of BwB is its cardinality.
It can of course be decomposed into right cosets of B, and its cardinality will
be the order of B times the cardinality of the quotient BwB/B.
where
1
|f | = |f (x)|.
|B|
x∈G
φww = φw φw .
φs ∗ φs = qφ1 + (q − 1)φs .
Proof. By (27.2), we have C(s) C(s) ⊆ C(1) ∪ C(s). Therefore, there exist
constants λ and μ such that φs ∗ φs = λφ1 + μφs . Evaluating both sides
at the identity gives λ = q. Now applying the augmentation and using the
special cases (φs ) = q, (f1 ) = 1 of (46.10), we have q 2 = λ · 1 + μ · q = q + μq,
so μ = q − 1.
480 46 Hecke Algebras
Proof. Let B be the braid group generated by uαi parametrized by the simple
roots αi , with n(uαi , uαj ) equal to the order (2 or 3) of sαi sαj . Let si = sβi
and si = sγi with βi , γi ∈ Σ, and let ui = uαi , ui = uβi be the corresponding
elements of B. By Theorem 25.2, we have
Since the fαi satisfy the braid relations, there is a homomorphism of B into
the group of invertible elements of Hk (q) such that uαi −→ fαi . Applying this
homomorphism to (46.15), we obtain (46.14).
There exists a rational prime p such that cw (p) are not all zero. Let H be
the convolution ring of B-bi-invariant functions on GL(k, Fp ). It follows from
Propositions 46.8 and 46.9 that (46.11)–(46.13) are all satisfied by the stan-
dard generators of H, so we have a homomorphism Hk (q) −→ H mapping
each fw to the corresponding generator φw of H and mapping q −→ p. The
images of the fw are linearly independent in H, yet since the cw (p) are not
all zero, we obtain a relation of linear independence. This is a contradiction.
The result is proved if q is transcendental. If 0
= q0 ∈ C, then there
is a homomorphism R −→ C, and a compatible homomorphism Hk (q) −→
482 46 Hecke Algebras
Hk (q0 ), in which q −→ q0 . What we must show is that the R-basis elements
fw remain linearly independent when projected to Hk (q0 ). To prove this, we
note that in Hk (q) we have
fw fw = aw,w ,w (q, q −1 )fw ,
w ∈W
where aw,w ,w is a polynomial in q and q −1 . We may construct ring H̃k (q0 )
over C with basis elements f˜w indexed by W and specialized ring structure
constants aw,w ,w (q0 , q0 −1 ). The associative law in Hk (q) boils down to a
polynomial identity that remains true in this new ring, so this ring exists.
Clearly, the identities (46.11)–(46.13) are true in the new ring, so there exists
a homomorphism Hk (q0 ) −→ H̃k (q0 ) mapping the fw to the f˜w . Since the f˜w
are linearly independent, so are the fw in Hk (q0 ).
Let us return to the case where q is a prime power.
Theorem 46.4. Let q be a prime power. Then the Hecke algebra Hk (q) is
isomorphic to the convolution ring of B-bi-invariant functions on GL(k, Fq ),
where B is the Borel subgroup of upper triangular matrices in GL(n, Fq ). In
this isomorphism, the standard basis element fw (w ∈ W ) corresponds to the
characteristic function of the double coset BwB.
Proof. It follows from Propositions 46.8 and 46.9 that (46.11)–(46.13) are
all satisfied by the elements φw in the ring H of B-bi-invariant functions on
GL(n, Fq ), so there exists a homomorphism Hk (q) −→ H such that fw −→ φw .
Since the {fw } are a basis of Hk (q) and the φw are a basis of H, this ring
homomorphism is an isomorphism.
Exercises
Exercise 46.1. Show that any subgroup of GL(n, F ) containing B is of the form
(46.2).
Exercise 46.2. For G = GL(3), describe Uw+ and Uw− explicitly for each of the six
Weyl group elements.
Exercise 46.3. Let G be a finite group and H a subgroup. Let H be the “Hecke
algebra” of H bi-invariant functions, with multiplication being the convolution
product normalized by
1
(f1 ∗ f2 )(g) = f1 (x)f2 (x−1 g).
|H| x∈G
Exercise 46.4. In the setting of Exercise 46.3, show that (π, V ) −→ V H is a
bijection between the isomorphism classes of irreducible representations of G with
V H
= 0 and isomorphism classes of irreducible H-modules.
Exercise 46.7. Prove that the degree of the irreducible character sλ (q) of GL(k, Fq )
is a polynomial in q whose value when q = 1 is the degree dλ of the irreducible
character sλ of Sk .
There are four theories that deserve to be studied in parallel. These are:
• The representation theory of symmetric groups Sk ;
• The representation theory of GL(k, Fq );
• The representation theory of GL(k, F ) where F is a local field;
• The theory of automorphic forms on GL(k).
In this description, a local field is R, C, or a field such as the p-adic field
Qp that is complete with respect to a non-Archimedean valuation. Roughly
speaking, each successive theory can be thought of as an elaboration of its
predecessor. Both similarities and differences are important. We list some
parallels between the four theories in Table 47.1.
The plan of this chapter is to discuss all four theories in general terms,
giving proofs only for the second stage in this tower of theories, the represen-
tation theory of GL(n, Fq ). (The first stage is already adequately covered.)
Although the third and fourth stages are outside the scope of this book, our
goal is to prepare the reader for their study by exposing the parallels with the
finite field case.
There is one important way in which these four theories are similar: there
are certain representations that are the “atoms” from which all other repre-
sentations are built and a “constructive process” from which the other repre-
sentations are built. Depending on the context, the “atomic” representations
are called cuspidal or discrete series representations. The constructive process
is parabolic induction or Eisenstein series. The constructive process usually
(but not always) produces an irreducible representation.
Harish-Chandra [62] used the term “philosophy of cusp forms” to describe
this parallel, which will be the subject of this chapter. One may substitute any
reductive group for GL(k) and most of what we have to say will be applicable.
But GL(k) is enough to fix the ideas.
In order to explain the philosophy of cusp forms, we will briefly summarize
the theory of Eisenstein series before discussing (in a more serious way) a
part of the representation theory of GL(k) over a finite field. The reader
only interested in the latter may skip the paragraphs on automorphic forms.
When we discuss automorphic forms, we will prove nothing and state exactly
what seems relevant in order to see the parallel. For GL(k, Fq ), we prove
more, but mainly what we think is essential to see the parallel. Our treatment
is greatly influenced by Howe [74] and Zelevinsky [178]. To go deeper into
the representation theory of the finite groups of Lie type, Carter [32] is an
exceedingly useful reference.
For the symmetric groups, there is only one “atom”—the trivial represen-
tation of S1 . The constructive process is ordinary induction from Sk × Sl to
Sk+l , which was the multiplication ◦ in the ring R introduced in Chap. 34.
The element that we have identified as atomic was called h1 there. It does not
generate the ring R. However, hk1 is the regular representation (or character)
of Sk , and it contains every irreducible representation. To construct every
irreducible representation of Sk from this single irreducible representation of
S1 , the constructive process embodied in the multiplicative structure of the
ring R must be supplemented by a further procedure. This is the extraction
of an irreducible from a bigger representation hk1 that includes it. This ex-
traction amounts to finding a description for the “Hecke algebra” that is the
endomorphism ring of hk1 . This “Hecke algebra” is isomorphic to the group
ring of Sk .
For the groups GL(k, Fq ), let us construct a graded ring R(q) analogous
to the ring R in Chap. 34. The homogeneous part Rk (q) will be the free
Abelian group on the set of isomorphism classes of irreducible representations
of GL(k, Fq ), which may be identified with the character ring of this group;
the multiplicative structure of the character ring is not used. Instead, there
is a multiplication Rk (q) × Rl (q) −→ Rk+l (q), called parabolic induction.
Consider the maximal parabolic subgroup P = M U of GL(k + l, Fq ), where
∼ g1
M = GL(k, Fq ) × GL(l, Fq ) = g1 ∈ GL(k, Fq ), g2 ∈ GL(l, Fq )
g2
and
Ik X
U= X ∈ Matk×l (Fq ) .
Il
The group P is a semidirect product, since U is normal, and the composition
M −→ P −→ P/U
for all z ∈ R×
+ . The character ω is the central character . It is fixed throughout
the discussion and is assumed unitary; that is, |ω(z)| = 1.
Let V be a vector space on which K and g both act. The actions are as-
sumed to be compatible in the sense that both induce the same representation
of Lie(K). We ask that V decomposes into a direct sum of finite-dimensional
irreducible subspaces under K. Then V is called a (g, K)-module. If every
irreducible representation of K appears with only finite multiplicity, then we
say that V is admissible. For example, let (π, H) be an irreducible unitary
representation of G on a Hilbert space H, and let V be the space of K-finite
vectors in H. It is a dense subspace and is closed under actions of both g
and K, so it is a (g, K)-module. The (g, K)-modules form a category that can
be studied by purely algebraic methods, which captures the essence of the
representations.
The space A(Γ \G) of automorphic forms is not closed under ρ because
K-finiteness is not preserved by ρ(g) unless g ∈ K. Still, both K and g
preserve the space A(Γ \G). A subspace that is invariant under these actions
and irreducible in the obvious sense is called an automorphic representation.
It is a (g, K)-module.
Given an automorphic form f on G = GL(k, R) with respect to Γ =
GL(k, Z), if k = r + t we can consider the constant term along the parabolic
subgroup P with Levi factor GL(r) × GL(t). This is the function
IX g1
f dX
Matr×t (Z)\Matr×t (R) I g2
for (g1 , g2 ) ∈ GL(r, R) × GL(t, R). If the constant term of f along every maxi-
mal parabolic subgroup vanishes then f is called a cusp form. An automorphic
representation is called automorphic cuspidal if its elements are cusp forms.
Let L2 (Γ \G, ω) be the space of measurable functions on g that are auto-
morphic and have central character ω and such that
|f (g)|2 dg < ∞.
Γ Z\G
is true.
47 The Philosophy of Cusp Forms 491
This integral may be shown to be convergent if re(s) > 12 . For other values
of s, it has analytic continuation. This integral emerges when one looks at
the constant term of the Eisenstein series with respect to Q. We will not
explain this further but mention it because these intertwining integrals are
extremely important and will reappear in the finite field case in the proof of
Proposition 47.3.
The two constructions—constant term and Eisenstein series—have paral-
lels in the representation theory of GL(k, F ), where F is a local field including
F = R, C, or a p-adic field. These constructions are functors between repre-
sentations of GL(k, F ) and those of the Levi factor of any parabolic subgroup.
They are the Jacquet functors in one direction and parabolic induction in the
other. (We will not define the Jacquet Functors, but they are the functors rU,1
in Bernstein and Zelevinsky [17].) Moreover, these constructions also descend
to the case of representation theory of GL(n, Fq ), which we look at next.
V U −→ ρ ⊗ τ −→ (π1 ◦ · · · ◦ πh ) ⊗ (πh+1 ◦ · · · ◦ πm ).
V −→ (π1 ◦ · · · ◦ πh ) ◦ (πh+1 ◦ · · · ◦ πm ) = π1 ◦ · · · ◦ πm .
Mλ = GL(λ1 , Fq ) × · · · × GL(λr , Fq )
Mλ −→ Pλ −→ Pλ /Uλ
is an isomorphism, where the first map is inclusion and the second projection.
This means that the representation πλ of Mλ may be regarded as a represen-
tation of Pλ in which Uλ acts trivially. Then π1 ◦ · · · ◦ πr is the representation
induced from Pλ .
Proof. We will frame our proof in terms of characters rather than repre-
sentations, so in this proof elements of Rk (q) are generalized characters of
GL(k, Fq ).
We make use of the involution ι : GL(k, Fq ) −→ GL(k, Fq ) defined by
⎛ ⎞
1
.
ι
g = wk · t g −1 · wk , wk = ⎝ . . ⎠ .
1
they are closely connected with the constant terms of the Eisenstein series
and with the functional equations. It is for this reason that we give a second,
longer proof of a weaker statement.
−1
Making the
variable change
X −→ g2 Xg1 and
then using (47.4) and the fact
that τ ◦ π1 (g1 ) ⊗ π2 (g2 ) = π2 (g2 ) ⊗ π1 (g1 ) ◦ τ shows that
g2
Mf h = π2 (g2 ) ⊗ π1 (g1 ) M f (h).
g1
Thus M f ∈ V2 ◦ V1 .
The map M is an intertwining operator since G acts on both the spaces of
π1 ◦ π2 and π2 ◦ π1 by right translation, and f −→ M f obviously commutes
with right translation. We must show that it is nonzero. Choose a nonzero
vector ξ ∈ V1 ⊗ V2 . Define
AB π1 (A) ⊗ π2 (D) ξ if C = 0,
f =
CD 0 otherwise,
and the term is zero unless X = 0, so this equals τ (ξ)
= 0. This proves that
the intertwining operator M is nonzero.
Returning momentarily to automorphic forms, the functional equation
(47.1) extends to Eisenstein series in several complex variables attached to
cusp forms for general parabolic subgroups. We will not try to formulate a
precise theorem, but suffice it to say that if πi are automorphic cuspidal rep-
resentations of GL(λi , R) and s = (s1 , . . . , sr ) ∈ Cr , and if ds : Pλ (R) −→ C
is the quasicharacter
⎛ ⎞
g1
⎜ .. ⎟
⎠ = | det(g1 )| 1 · · · | det(gr )| r ,
s s
ds ⎝ .
gr
then there is a representation Ind(π1 ⊗ · · · ⊗ πr ⊗ ds ) of GL(k, R) induced
parabolically from the representation π1 ⊗ · · · ⊗ πr ⊗ ds of Mλ . One may form
an Eisenstein series by a series that is absolutely convergent if re(si − sj ) are
sufficiently large and that has meromorphic continuation to all si . There are
functional equations that permute the constituents | det |si ⊗ πi .
If some of the πi are equal, the Eisenstein series will have poles. The polar
divisor maps out the places where the representations Ind(π1 ⊗ · · · ⊗ πr ⊗ ds )
are reducible. Restricting ourselves to the subspace of Cr where si = 0,
the following picture emerges. If all of the πi are equal, then the polar divisor
will consist of r(r − 1) hyperplanes in parallel pairs. There will be r! points
where r − 1 hyperplanes meet in pairs. These are the points where the induced
representation Ind(π1 ⊗ · · · ⊗ πr ⊗ ds ) is maximally reducible. Regarding the
reducibility of representations, we will see that there are both similarities and
dissimilarities with the finite field case.
Returning to the case of a finite field, we will denote by T the subgroup
of diagonal matrices in GL(k, Fq ). If α is a root, we will denote by Uα the
one-dimensional unipotent of GL(k, Fq ) corresponding to α. Thus, if α = αij
in the notation (27.7), then Xα consists of the matrices of the form I + tEij ,
where Eij has a 1 in the i, jth position and zeros elsewhere.
If λ = (λ1 , . . . , λr ) is an ordered partition of k, πi are representations of
GL(λi , Fq ), and πλ = π1 ⊗ · · · ⊗ πr is the corresponding representation of Mλ ,
we will use Ind(πλ ) as an alternative notation for π1 ◦ · · · ◦ πr .
Theorem 47.3 (Harish-Chandra). Suppose that λ = (λ1 , . . . , λr ) and μ =
(μ1 , . . . , μt ) are ordered partitions of k, and let πλ = ⊗πi and πμ = ⊗πj be
cuspidal representations of Mλ and Mμ , respectively. Then
dim HomGL(k,Fq ) Ind(πλ ), Ind(πμ )
is zero unless r = t. If r = t, it is the number of permutations σ of {1, 2, . . . , r}
such that λσ(i) = μi and πσ(i) ∼ = πi .
496 47 The Philosophy of Cusp Forms
Now, since the representative w is only determined modulo left and right
multiplication by Mμ and Mλ , respectively, we may assume that w takes
positive roots of Mλ to positive roots of Mμ . Thus, a representative of w is a
“block permutation matrix” of the form
⎛ ⎞
w11 · · · w1r
⎜ .. ⎟ ,
w = ⎝ ... . ⎠
wt1 · · · wtr
Thus
⎛ ⎞ ⎛ ⎞
g1 gσ(1)
⎜ .. ⎟ ⎜ .. ⎟
Δ(w) ◦ πλ ⎝ . ⎠ = πμ ⎝ . ⎠ ◦ Δ(w),
gr gσ(r)
so
Δ(w) ◦ π1 (g1 ) ⊗ · · · ⊗ πr (gr ) = π1 (gσ(1) ) ⊗ · · · ⊗ πr (gσ(r) ) ◦ Δ(w).
Remark 47.1. This is the usual case. If q is large, the probability that there is
a repetition among a list of randomly chosen cuspidal representations is small.
Remark 47.2. The statement that the isomorphism class is unchanged if the λi
and πi are permuted is the analog of the functional equations of the Eisenstein
series.
Proof. Proofs may be found in Howlett and Lehrer [80] and Howe [74].
Corollary 47.1. There exists a natural bijection between the set of partitions
λ of t and the irreducible constituents σλ(π) of π0◦t . The multiplicity of σλ(π)
in π0◦t equals the degree of the irreducible character sλ of the symmetric group
St parametrized by λ.
Proof. The multiplicity of σλ(π) in π0◦t equals the multiplicity of the corre-
sponding module of Ht (q l ). By Exercise 46.5, this is the degree of sλ .
Although we will not make use of the multiplicative structure that is con-
tained in this theorem of Howlett and Lehrer, we may at least see immediately
that
dim End(π0◦t ) = t!, (47.6)
by Theorem 47.3, taking μ = λ and all πi , πi to be π0 . This is enough for the
following result.
Proof. We
note the following general principle: χ is a character of any group,
and if χ = di χi is a decomposition into subrepresentations such that
χ, χ = d2i ,
also. By the “general principle” stated at the beginning of this proof, it fol-
lows that the representations θ1j1 ◦ · · · ◦ θrjr are irreducible and mutually
nonisomorphic.
Next we show that every irreducible representation π is of the form Ind(θλ ).
If π is cuspidal, then π is monatomic, and so we can just take r = t1 = 1,
θ1 = π1 . We assume that π is not cuspidal. Then by Proposition 47.2 we
may embed π into π1 ◦ · · · ◦ πm for some cuspidal representations πi . By
Proposition 47.4, we may order these so that isomorphic πi are adjacent, so
π is embedded in a representation of the form π1◦t1 ◦ · · · ◦ πr◦tr , where πi are
nonisomorphic cuspidal representations. We have determined the irreducible
constituents of such a representation, and they are of the form Ind(θλ ), where
θi is πi -monatomic. Hence π is of this form.
We leave the final uniqueness assertion for the reader to deduce from
Theorem 47.3.
500 47 The Philosophy of Cusp Forms
The great paper of Green [58] constructs all the irreducible representations
of GL(k, Fq ). Systematic use is made of the ring R(q). However, Green does
not start with the cuspidal representations. Instead, Green takes as his ba-
sic building blocks certain generalized characters that are “lifts” of modular
characters, described in the following theorem.
Theorem 47.7 (Green). Let G be a finite group, and let ρ : G −→ GL(k, Fq )
be a representation. Let f ∈ Z[X1 , . . . , Xk ] be a symmetric polynomial with
integer coefficients. Let θ : F̄×
q −→ C
×
be any character. Let χ : G −→ C be
the function
χ(g) = f θ(α1 ), . . . , θ(αk ) .
Then θ is a generalized character.
Proof. First, we reduceto the following case: θ : F̄× q −→ C
×
is injective
and f (X1 , . . . , Xk ) = Xi . If this case is known, then by replacing ρ by
its exterior powers we get the same result for the elementary symmetric
polynomials, and hence for all symmetric polynomials. Then we can take
f (X1 , . . . , Xk ) = Xir , effectively replacing θ by θr . We may choose r to
match any given character on a finite field containing all eigenvalues of all g,
obtaining the result in full generality.
We recall that if l is a prime, a group is l-elementary if it is the direct
product of a cyclic group and an l-group. According to Brauer’s characteriza-
tion of characters (Theorem 8.4(a) on p. 127 of Isaacs [83]), a class function
is a generalized character if and only if its restriction to every l-elementary
subgroup H (for all l) is a generalized character. Thus, we may assume that
G is l-elementary. If p is the characteristic of Fq , whether l = p or not, we may
write G = P × Q where P is a p-group and p |Q|. The restriction of χ to Q
is a character by Isaacs, [83], Theorem 15.13 on p. 268. The result will follow
if we show that χ(gp q) = χ(q) for gp ∈ P , q ∈ Q. Since gp and q commute,
using the Jordan canonical form, we may find a basis for the representation
space of ρ over F̄q such that ρ(q) is diagonal and ρ(gp ) is upper triangular.
Because the order of gp is a power of p, its diagonal entries are 1’s, so q and
gp q have the same eigenvalues, whence χ(gp q) = χ(q).
Since the proof of this theorem of Green is purely character-theoretic, it
does not directly produce irreducible representations. And the characters that
it produces are not irreducible. (We will look more closely at them later.) How-
ever, Green’s generalized characters have two important advantages. First,
their values are easily described. By contrast, the values of cuspidal repre-
sentations are easily described on the semisimple conjugacy classes, but at
other classes require knowledge of “degeneracy rules” which we will not de-
scribe. Second, Green’s generalized character can be extended to a generalized
character of GL(n, Fqr ) for any r, a property that ordinary characters do not
have.
Still, the cuspidal characters have a satisfactory direct description, which
we turn to next. Choosing a basis for Fqk as a k-dimensional vector space
47 The Philosophy of Cusp Forms 501
where the last consists of the conjugacy classes of matrices whose eigenvalues
are ν and ν q , where ν ∈ Fq2 − Fq . In the duality, these four types of conjugacy
classes correspond to the four types of irreducible representations: the q + 1-
dimensional principal series, induced from a pair of distinct characters of
GL(1); the one-dimensional representations χ ◦ det, where χ is a character
of F×
q ; the q-dimensional representations obtained by tensoring the Steinberg
representation with a one-dimensional character; and the q − 1-dimensional
cuspidal representations.
Let f (X) = X d + ad−1 X d−1 + · · · + a0 be a monic irreducible polynomial
over Fq of degree d. Let
⎛ ⎞
0 1 0 ··· 0
⎜ 0 0 1 0 ⎟
⎜ ⎟
⎜ .. . .. ⎟
U (f ) = ⎜ . .. . ⎟
⎜ ⎟
⎝ 0 0 0 ··· 1 ⎠
−a0 −a1 −a2 · · · −ad−1
be the rational canonical form. Let
⎛ ⎞
U (f ) Id 0 ··· 0
⎜ 0 U (f ) Id ⎟
⎜ ⎟
⎜ .. ⎟
⎜
Ur (f ) = ⎜ 0 0 U (f ) . ⎟ ⎟,
⎜ . . ⎟
⎝ . . . . ⎠
0 ··· U (f )
({ri }, {di }, {(λ )i }, {fi }) parametrize the same conjugacy class if and only if
they both have the same length m and there exists a permutation σ ∈ Sm
such that ri = rσ(i) , di = dσ(i) , (λ )i = λσ(i) and fi = fσ(i) .
We say two conjugacy classes are of the same type if the parametrizing
data have the same length m and there exists a permutation σ ∈ Sm such
that ri = rσ(i) , di = dσ(i) , (λ )i = λσ(i) . (The fi and fi are allowed to differ.)
The set of types of conjugacy classes depends on k, but is independent of q
(though if q is too small, some types might be empty).
Lemma 47.1. Let {N1 , N2 , . . .} be a sequence of numbers, and for each Nk
let Xk be a set of cardinality Nk (Xk disjoint). Let Σk be the following set.
An element of Σk consists of a 4-tuple ({ri }, {di }, {λi }, {xi }), where {ri } =
{r1 , .
. . , rm } and {di } = {d1 , . . . , dm } are sequences of positive integers, such
that ri di = k, together with a sequence {λi } of partitions of ri and an
element xi ∈ Xdi , such that no xi are equal. Define an equivalence relation ∼
on Σk in which two elements are considered equivalent if they can be obtained
by permuting the data, that is, if σ ∈ Sm then
k−1
j
σk (g) = (−1)k−1 θ(αq ).
j=0
By Theorem 47.8, the number of σk,θ is the total number of cuspidal repre-
sentations, so this is a complete list.
We will first construct σk under the assumption that θ, regarded as a
character of F× qk
, can be extended to a character θ : F̄× q −→ C
×
that is
injective. This is assumption is too restrictive, and we will later relax it. We
will also postpone showing that that σk is independent of the extension of θ
to F̄×
q . Eventually we will settle these points completely in the special case
where k is a prime.
Let
k
χk (g) = θ(αi ), (47.9)
i=1
47 The Philosophy of Cusp Forms 505
the last step using the fact that the complement of the Tλreg in Tλ is of codi-
mension one. We note that the restriction of χk to Tλ is the sum of k distinct
characters, so
|χk (g)|2 = k|Tλ |.
g∈Tλ
We have
1 1
[G : NG (Tλ )]|Tλ | = [G : NG (Tλ )]|Tλreg | + O(q −1 )
|G| |G|
λ λ
1
= |Greg | + O(q −1 )
|G|
= 1 + O(q −1 ).
The result is now proved for q sufficiently large.
506 47 The Philosophy of Cusp Forms
To prove the result for all q, we will show that the inner product χk , χk
is a polynomial in q. This will follow if we can show that if S is the subset of G
consisting of the union of conjugacy classes of a single type, then [G : CG (g)]
is constant for g ∈ S and
|χk (g)|2 (47.10)
g∈S
is a polynomial in q. We note that for each type, the index of the centralizer
of (47.7) is the same for all such matrices, and that this index is polynomial
in q. Thus it is sufficient to show that the sum over the representatives (47.7)
is a polynomial in q. Moreover, the value of χk is unchanged if every instance
of a Ur (f ) is replaced with r blocks of U (f ), so we may restrictourselves
to semisimple conjugacy classes in confirming this. Thus if k = di ri , we
consider the sum (47.10), where the sum is over all matrices
⎛ ⎞
U(r1 ) (f1 )
⎜ .. ⎟
⎝ . ⎠,
U(rm ) (fm )
k−1
i
σk,θ (g) = (−1)k−1 θ(ν q ). (47.11)
i=0
χn , σk ◦ 1n−k G = χn , σk ⊗ 1n−k M .
Let
m1
m= ∈ M, m1 ∈ GL(k, Fq ), m2 ∈ GL(n − k, Fq ).
m2
Clearly, χn (m) = χk (m1 ) + χn−k (m2 ). Now using the induction hypothe-
sis, χn−k does not contain the trivial character of GL(n − k, Fq ), hence it is
orthogonal to 1n−k on GL(n − k, Fq ); so we can ignore χn−k (m2 ). Thus,
χn , σk ◦ 1n−k G = χk , σk GL(k,Fq ) .
since by Exercise 47.2(ii), this will show that the representation affording the
character σn has no U -invariants, the definition of cuspidality. The summand
on the left-hand side is independent of u, and by the definition of χn the
left-hand side is just χk (m1 ) + χn−k (m2 ). By Exercise 47.4, the right-hand
side can also be evaluated. Using (47.11), which we have assumed inductively
for σr with r < n, the terms r = k and r = n − k contribute χk (m1 ) and
χn−k (m2 ) and all other terms are zero.
To evaluate the sign in (47.13), we compare the values at the identity to
get the relation
n−1 k−17 7
n−1
n
n= (−1)k−1 (q j − 1) ± (q j − 1),
k (q) j=1 j=1
k=1
where
/n
n j=1 (q
j
− 1)
= >/ ? >/ ?
k (q) k
(q j − 1) n−k j
(q − 1)
j=1 j=1
is the Gaussian binomial coefficient, which is the index of the parabolic sub-
group with Levi factor GL(k) × GL(n − k). Substituting q = 0 in this identity
shows that the missing sign must be (−1)n−1 .
If g is a regular element of T(k) , then the value of σk on a regular element
of T(k ) is now given by (47.11) since if k < n then σk ◦ 1n−k vanishes on g,
which is not conjugate to any element of the parabolic subgroup from which
σk ◦ 1n−k is induced.
See Exercise 47.9 for an example showing that the cuspidal characters
that we have constructed are not enough because of our assumption that
θ is injective. Without attempting a completely general result, we will now
give a variation of Theorem 47.9 that is sufficient to construct all cuspidal
representations of GL(k, Fq ) when k is prime.
Proof. The proof is similar to Proposition 47.4. It is sufficient to show this for
sufficiently large q. As in that proposition, the sum is
47 The Philosophy of Cusp Forms 509
1
[G : NG (Tλ )] |χk (g)|2 + O(q −1 ).
|G|
λ a partition of k g∈Tλ
We can interpret this as a sum over the symmetric group. If σ ∈ Sk , let r(σ)
be the number of fixed points of σ. In the conjugacy class of shape λ, there
are k!/zλ elements, and so
1 1
(k + r2 − r) = (k + r(σ)2 − r(σ)).
zλ k!
λ σ∈Sk
where the
inner product
is now over the symmetric group. Clearly hk , hk =
1 and s(k−1,1) , hk = 0. Since the character s(k−1,1) is real and hk is the
constant function equal to 1,
G H
s2(k−1,1) , hk = s(k−1,1) , s(k−1,1) = 1,
k
χk = (−1)r−1 σr ◦ 1k−r .
r=1
If θ is trivial onF×
q , it is still true that the restriction of θ to Fqd does
not factor through the norm map to Fqr for any proper divisor of d whenever
k n. So χk , χk = k + 1 by Theorem 47.5. Now, we can proceed as before,
except that σ1 = 11 , so σ1 ◦ 1k−1 is not irreducible—it is the sum of two
irreducible representations s(k−1,1) (q) and s(k) (q) of GL(k, Fq ), in the notation
of Chap. 46. Of course, s(k) (q) is the same as 1k in the notation we have been
using. The rest of the argument goes through as in Theorem 47.9. In particular
the inner product formula χk , χk = k + 1 together with fact that 11 ◦ 1k−1
accounts for two representations
in the decomposition of χk guarantees that
σk , defined to be χk − r<k (−1)r σr ◦ 1k−r is irreducible.
The cuspidal characters we have constructed are linearly independent
by (47.14). They are equal in number to the total number of cuspidal rep-
resentations, and so we have constructed all of them.
Exercises
Exercise 47.1 (Transitivity of parabolic induction).
(i) Let P be a parabolic subgroup of GL(k) with Levi factor M and unipotent
radical U , so P = M U . Suppose that Q is a parabolic subgroup of M with
Levi factor MQ and unipotent radical UQ . Show that MQ is the Levi factor of
a parabolic subgroup R of GL(k) with unipotent radical UQ U .
(ii) In the setting of (i), show that parabolic induction from MQ directly to GL(k)
gives the same result as parabolically inducing first to M , and then from M to
GL(k).
(iii) Show that the multiplication ◦ is associative and that R(q) is a ring.
HomP (V, W ) ∼ U
= HomM (V , W ).
Let V0 be the span of elements of the form w − π(u)w with u ∈ U . Show that
V = V U ⊕ V0 , as M -modules, and that any P -equivariant map V −→ W factors
through V /V0 ∼= V U .)
(ii) Let χ be a character of G, and let σ be a character of M . Let Ind(σ) be the
character of the representation of G parabolically induced from σ, and let χU
be the function on M defined by
1
χU (m) = χ(mu).
|U | u∈U
The next exercise is very analogous to the computation of the constant terms of
Eisenstein series. For example, the computation around pages 39–40 of Langlands
[117] is a near exact analog.
514 47 The Philosophy of Cusp Forms
[Hint: Both sides are class functions, so it is sufficient to compare the inner products
with ρ1 ⊗ ρ2 where ρ1 and ρ2 are irreducible representations of GL(k, Fq )andGL(n −
k, Fq ), respectively. Using Exercise 47.2 this amounts to comparing σ1 ◦σ2 and ρ1 ◦ρ2 .
To do this, explain why in the last statement in Theorem 47.6 the assumption that
the θi are monatomic with respect to distinct cuspidals may be omitted provided
this assumption is made for the θi .]
Let P be the “mirabolic” subgroup of g ∈ G where the bottom row is (0, . . . , 0, 1).
(Note that P is not a parabolic subgroup.) Call an irreducible representation of
P cuspidal if it has no U -fixed vector for the unipotent radical U of any standard
parabolic subgroup of G. Note that U is contained in P for each such U . If 1 r < k
let Gr be GL(r, Fq ) embedded in G in the upper left-hand corner, and let N r be the
subgroup of x ∈ N in which xij = 0 if i < j r.
(i) Show that the representation κ = IndP N (ψ) is irreducible. [Hint: Use Mackey
theory to compute HomP (κ, κ).]
(ii) Let (π, V ) be a cuspidal representation of P . Let Lr be the set of all linear
functionals λ on V such that λ(π(x)v) = ψN (x)v for v ∈ V and x ∈ Lr . Show
r > 1 then there exists γ ∈ Gr−1 such that λ ∈ Lr−1 , where
that if λ ∈ Lr and
λ (v) = λ π(γ)v .
(iii) Show that the restriction of an cuspidal representation π of GL(k, Fq ) to P is
a direct sum of copies of κ. Then use Exercise 45.4 to show that at most one
copy can occur, so π|P = κ.
47 The Philosophy of Cusp Forms 515
(iv) Show that each irreducible cuspidal representation of GL(k, Fq ) has dimension
(q − 1)(q 2 − 1) · · · (q k−1 − 1).
Exercise 47.9. Obtain a character table of GL(2, F3 ), a group of order 48. Show
that there are three distinct characters θ of F× 9 such that θ does not factor through
the norm map Fqk −→ Fqd for any proper divisor of d. Of these, two (of order
eight) can be extended to an injective homomorphism F̄× ×
3 −→ C , but the third (of
order four) cannot. If θ is this third character, then χ2 defined by (47.9) defines a
character that splits as χtriv +χsteinberg −σ2 , where χtriv and χsteinberg are the trivial
and Steinberg characters, and σ2 is the character of a cuspidal representation. Show
also that σ2 differs from the sum of the two one-dimensional characters of GL(2, F3 )
only on the two non-semisimple conjugacy classes, of elements of orders 3 and 6.
Exercise 47.12. Let θ be an injective character of F̄q . Prove the following result.
Theorem. Let λ be a partition of n and let t ∈ Tλ . Then σk,θ (t) = 0 unless λ = (n).
Hint: Assume by induction that the statement is true for all k < n. Write
t = (t1 , . . . , tr ) where ti ∈ GL(λi , Fq ) has distinct eigenvalues in Fqλi . Show that
(σk ◦ 1n−k )(t) = σk (ti ).
λi
48
Cohomology of Grassmannians
In this chapter, we will deviate from our usual policy of giving complete proofs
in order to explain some important matters. Among other things, we will see
that the ring R introduced in Chap. 34 has yet another interpretation in terms
of the cohomology of Grassmannians.
References for this chapter are Fulton [53], Hiller [70], Hodge and Pedoe
[71], Kleiman [102], and Manivel [126].
We recall the notion of a CW-complex . Intuitively, this is just a space
decomposed into open cells, the closure of each cell being contained in
the union of cells of lower dimension—for example, a simplicial complex. (See
Dold [44], Chap. 5, and the appendix in Milnor and Stasheff [129].) Let Bn
be the closed unit ball in Euclidean n-space. Let B◦n be its interior, the unit
disk, and let Sn−1 be its boundary, the n − 1 sphere. We are given a Hausdorff
topological space X together with set S of subspaces of X. It is assumed that
X is the disjoint union of the Ci ∈ S, which are called cells. Each space Ci ∈ S
is homeomorphic to B◦d(i) for some d(i) by a homeomorphism εi : B◦d(i) −→ Ci
that extends to a continuous map εi : Bd(i) −→ X. The image of Sd(i)−1 under
εi lies in the union of cells Ci of strictly lower dimension. Thus, if we define
the n-skeleton
A
Xn = Ci ,
d(i)n
A curve of degree 2
(hyperbola) deformed
into a pair of lines.
Bezout’s theorem can be used to illustrate Chow’s lemma. First, note that
a curve of degree d is rationally equivalent to a sum of d lines (Fig. 48.1), so
Y is linearly equivalent to a sum of d(Y ) lines, and Z is linearly equivalent
to a sum of d(Z) lines. Since two lines have a unique point of intersection,
the first set of d(Y ) lines will intersect the second set of d(Z) lines in exactly
d(Y ) d(Z) points, which is Bezout’s theorem for P2 (Fig. 48.2).
What is the
multiplicity of one
circle intersecting? Four!
... deform
To compute one copy
Y · Y ... of the circle
B\G/P ∼
= W/WP ,
with Levi factor M = GL(r, C) × GL(s, C). The quotient Xr,s = G/P is then
the Grassmannian, a compact complex manifold of dimension rs. In this case,
the cohomology ring H ∗ (Xr,s ) is closely related to the ring R introduced in
Chap. 34.
To explain this point, let us explain how to “truncate” the ring R and
obtain a finite-dimensional algebra that will be isomorphic to H ∗ (Xr,s ).
Suppose that Jr is the linear span of all sλ such that the length of λ is
> r. Then Jr is an ideal, and the quotient R/Jr ∼ = Λ(r) by the characteristic
map. Indeed, it follows from Proposition 36.5 that Jr is the kernel of the
homomorphism ch(n) : R −→ Λ(n) .
We can also consider the ideal ι Js , where ι is the involution of Theorem 34.3.
By Proposition 35.2, this is the span of the sλ in which the length of λt is
greater than s—in other words, in which λ1 > s. So Jr + ι Js is the span of all
sλ such that the diagram of λ does not fit in an r × s box. Therefore, the ring
Rr,s = R/(Jr + ι Js ) is spanned by the images of sλ where the diagram of λ
does fit in an r × s box. For example, R3,2 is spanned by s() , s(1) , s(2) , s(11) ,
522 48 Cohomology of Grassmannians
s(21) , s(22) , s(111) , s(211) , s(221) , and s(222) . It is a free Z-module of rank 10.
In general the rank of the ring Rr,s is r+s r , which is the number of partitions
of r + s of length r into parts not exceeding s—that is, partitions with
diagrams that fit into a box of dimensions r × s.
We will not prove this. Proofs (all rather similar and based on a method
of Hodge) may be found in Fulton [53], Hiller [70], Hodge and Pedoe [71],
and Manivel [126]. We will instead give an informal discussion of the result,
including a precise description of the isomorphism and an example.
Let us explain how to associate a partition λ with a diagram that is con-
tained in the r×s box with a Schubert cell of codimension equal to |λ|. In fact,
to every coset wWP in W/WP we will associate such a partition.
Right multiplication by an element of WP ∼ = Sr × Ss consists of reordering
the first r columns and the last s columns. Hence, the representative w of the
given coset in W/WP may be chosen to be a permutation matrix such that
the entries in the first r columns are in ascending order, and so that the
entries in the last s columns are in ascending order. In other words, if σ is the
permutation such that wσ(j),j
= 0, then
σ(1) < σ(2) < · · · < σ(r), σ(r + 1) < σ(r + 2) < · · · < σ(r + s). (48.2)
Now, we collect the marked columns and read off the permutation. For each
row containing marks, there will be a part of the permutation equal to the
number of marks in that row. In the three examples above, the respective
permutations λ are:
Their diagrams fit into a 2 × 3 box. We will write Cλ for the closed Schubert
cell C(w) when λ and w are related this way.
48 Cohomology of Grassmannians 523
0 d0 d1 · · · dr+s = r, 0 di 1, (48.4)
dim(V ∩ Fi ) = di . (48.5)
dim(V ∩ Fi ) di . (48.6)
The function V −→ dim(V ∩ Fi ) is upper semicontinuous on Gr,s , that is, for
any integer n, {V | dim(V ∩ Fi ) n} is closed. Therefore, C(d) is closed, and
in fact it is the closure of C◦(d) .
Proof. If di+1 > di and dim(V ∩ Fi+1 ) di+1 , then since V ∩ Fi has
codimension at most 1 in V ∩Fi+1 we do not need to assume dim(V ∩Fi ) di .
If di = di−1 and dim(V ∩ Fi−1 ) di−1 then dim(V ∩ Fi−1 ) di−1 .
We will show C◦(d) is the image in Gr,s of an open Schubert cell. For example,
with r = 3 and s = 2, taking w to be the first matrix in (48.3), we consider the
Schubert cell BwP/P , which has the image in G3,2 that consists of all bwF3 ,
where b ∈ B. A one-dimensional unipotent subspace of B is sufficient to
produce all of these elements, and a typical such space consists of all matrices
of the form
⎛ ⎞⎛ ⎞⎛ ⎞ ⎛ ⎞
1 1 x1 x1
⎜ 1 ⎟⎜ 1 ⎟ ⎜ x2 ⎟ ⎜ x2 ⎟
⎜ ⎟⎜ ⎟⎜ ⎟ ⎜ ⎟
⎜ 1 α ⎟⎜ 1 ⎟ ⎜ x3 ⎟ = ⎜ αx3 ⎟
⎜ ⎟⎜ ⎟⎜ ⎟ ⎜ ⎟
⎝ 1 ⎠⎝ 1 ⎠ ⎝ 0 ⎠ ⎝ x3 ⎠
1 1 0 0
Proposition 48.1. The image in Gr,s of the Schubert cell C(w) corresponding
to the partition λ (the diagram of which, we have noted, must fit in an r × s
box) is C(d) , where the integer sequence (d0 , d1 , . . . , dr+s ) where
dk = i ⇐⇒ s + i − λi k s + i − λi+1 . (48.7)
We note that, by Lemma 48.1, if (d) is the sequence in (48.7), the closed
Schubert cell C(d) is characterized by the conditions
Also, by Lemma 48.1, this only needs to be checked when λi > λi+1 . [The
characterization of the open Schubert cell still requires dim(V ∩ Fk ) to be
specified for all k, not just those of the form s + i − λi .]
Proof. We will prove this for the open cell. The image of C(w)◦ in Gr,s consists
of all spaces bwFr with b ∈ B, so we must show that, with di as in (48.7), we
have
dim(bwFr ∩ Fi ) = di .
Since b stabilizes Fi , we may apply b−1 , and we are reduced to showing that
dim(wFr ∩ Fi ) = di .
When (d) and λ are related as in (48.7), we will also denote the Schubert
cell C(d) by Cλ .
As we asserted earlier, the cohomology ring Xr,s is isomorphic to the
quotient Rr,s of the ring R, which has played such a role in this last part
of the book. To get some intuition for this, let us consider the identity in R
So our expectation is that if we deform C(1) into two copies C(1) and C(1) that
intersect properly, the intersection will be rationally equivalent to the sum of
C(2) and C(11) . We may choose spaces Gs and Hs of codimension s such that
Gs ∩ Hs = Fs−1 and Gs + Hs = Fs+1 . Now let us consider the intersection of
If V lies in both C(1) and C(1) , then let v and v be nonzero vectors in V ∩ Gs
and V ∩ Hs , respectively. There are two possibilities. Either v and v are
proportional, in which case they lie in V ∩ Fs−1 , so V ∈ C(2) , or they are
linearly independent. In the second case, both lie in Fs+1 , so V ∈ C(11) .
The intersection theory of flag manifolds is very similar to that of Grass-
mannians. The difference is that while the cohomology of Grassmannians for
GL(r) is modeled on the ring R, which can be identified as in Chap. 34 with
the ring Λ of symmetric polynomials, the cohomology of flag manifolds is
modeled on a polynomial ring. Specifically, if B is the Borel subgroup of
G = GL(r, C), then the cohomology ring of G/B is a quotient of the poly-
nomial ring Z[x1 , . . . , xr ], where each xi is homogeneous of degree 2. Las-
coux and Schützenberger defined elements of the polynomial ring Z[x1 , . . . , xr ]
called Schubert polynomials which play a role analogous to that of the Schur
polynomials (See Fulton [53] and Manivel [126]).
A minor problem is that H ∗ (G/B) is not precisely the polynomial ring
Z[x1 , . . . , xr ] but a quotient, just as H ∗ (Gr,s ) is not precisely R or even its
quotient R/Jr , which is isomorphic to the ring of symmetric polynomials in
Z[x1 , . . . , xr ].
The ring Z[x1 , . . . , xr ] should be more properly regarded as the cohomology
ring of an infinite CW-complex, which is the cohomology ring of the space Fr
of r-flags in C∞ . That is, let Fr,s be the space of r-flags in Cr+s :
We may embed Fr,s −→Fr,s+1 , and the union of the Fr,s (topologized as
the direct limit) is Fr . The open Schubert cells in Fr,s correspond to double
526 48 Cohomology of Grassmannians
2n
|X(Fq )| = (−1)k tr(φ|H k ) .
k=0
The dimensions of the l-adic cohomology groups are the same as the complex
cohomology, and in these examples (since all the cohomology comes from
algebraic cycles) the odd-dimensional cohomology vanishes while on H 2i (X)
the Frobenius endomorphism acts by the scalar q i . Thus,
n
|X(Fq )| = dim H 2k (X) q k .
k=0
The l-adic cohomology groups have the same dimensions as the complex ones.
Hence, the Grothendieck–Lefschetz fixed-point formula explains the extraor-
dinary fact that the number of points over a finite field of the Grassmannian
or flag varieties is a generating function for the complex cohomology.
48 Cohomology of Grassmannians 527
Exercises
Exercise 48.1. Consider the space Fr,s (Fq ) of r-flags in Fr+s . Compute the car-
dinality by representing it as GL(n, Fq )/P (Fiq ), where P is the parabolic subgroup
(48.11).
r+i−1 Show that |F r,s (F q )| = i di (r, s) q , where for fixed s, we have di (r, s) =
i
.
http://www.sagemath.org/doc/thematic tutorials/lie.html
sage: B3=WeylCharacterRing("B3"); B3
The Weyl Character Ring of Type [’B’, 3] with Integer Ring
coefficients
sage: B3.fundamental_weights()
Finite family {1: (1,0,0), 2: (1,1,0), 3: (1/2,1/2,1/2)}
sage: [B3(f) for f in B3.fundamental_weights()]
[B3(1,0,0), B3(1,1,0), B3(1/2,1/2,1/2)]
sage: [B3(f).degree() for f in B3.fundamental_weights()]
[7, 21, 8]
sage: B3(1,1,0).symmetric_power(3)
B3(1,0,0) + 2*B3(1,1,0) + B3(2,1,1) + B3(2,2,1) + B3(3,1,0)
+ B3(3,3,0)
sage: [f1,f2,f3]=B3.fundamental_weights()
sage: B3(f3)
B3(1/2,1/2,1/2)
sage: B3(f3)^3
4*B3(1/2,1/2,1/2) + 3*B3(3/2,1/2,1/2) + 2*B3(3/2,3/2,1/2)
+ B3(3/2,3/2,3/2)
This illustrates different ways of interacting with Sage as a command-line
interpreter. I prefer to run Sage from within an Emacs buffer; others prefer
the notebook. For complicated tasks, such as loading some Python code, you
may write your commands in a file and load or attach it. Whatever your
method of interacting with the program, you can have a dialog with this one.
Sage provides a prompt (“sage:”) after which you type a command. Sage will
sometimes produce some output, sometimes not; in any case, when it is done
it will give you another prompt.
The first line contains two commands, separated by a semicolon. The first
command creates the WeylCharacterRing B3 but produces no output. The
second command “B3” prints the name of the ring you have just created.
Elements of the ring are virtual representations of the Lie group Spin (7)
having Cartan type B3 . Addition corresponds to direct sum, multiplication to
tensor product.
The ring B3 is a Python class, and like every Python class it has methods
and attributes which you can use to perform various tasks. If at the sage:
prompt you type B3 then hit the tab key, you will get a list of Python
methods and attributes that B3 has. For example, you will notice methods
dynkin diagram and extended dynkin diagram. If you want more informa-
tion about one of them, you may access the on-line documentation, with
examples of how to use it, by typing B3.dynkin diagram?
Turning to the next command, the Python class B3 has a method called
fundamental weights. This returns a Python dictionary with elements that
are the fundamental weights. The third command gives the irreducible repre-
sentations with these highest weights, as a Python list. After that, we compute
the degrees of these, the symmetric cube of a representation, the spin repre-
sentation B3(1/2,1/2,1/2) and its square.
Appendix: Sage 531
To get documentation about Sage’s branching rules, either see the thematic
tutorial or enter the command:
sage: get_branching_rule?
As another example, let us compute
|tr(g)|20 dg.
SU(2)
There are different ways of doing this computation. An efficient way is just
to compute decompose the tenth power of the standard character
2 into irre-
ducibles: if tr(g)10 = aλ χλ then its modulus squared is aλ .
sage: A1=WeylCharacterRing("A1")
sage: A1(1)
A1(0,0)
sage: A1([1])
A1(1,0)
sage: A1([1])^10
42*A1(5,5) + 90*A1(6,4) + 75*A1(7,3) + 35*A1(8,2) + 9*A1(9,1)
+ A1(10,0)
sage: (A1([1])^10).monomial_coefficients()
{(8, 2): 35, (10, 0): 1, (9, 1): 9, (5, 5): 42,
(7, 3): 75, (6, 4): 90}
sage: sum(v^2 for v in (A1([1])^10).monomial_coefficients().values())
16796
You may also use the method weight multiplicities of a Weyl character
to get a dictionary of weight multiplicities indexed by weight.
sage: A2=WeylCharacterRing("A2")
sage: d=A2(6,2,0).weight_multiplicities(); d
{(0, 6, 2): 1, (5, 0, 3): 1, (3, 5, 0): 1, ...
Appendix: Sage 533