Sie sind auf Seite 1von 290

Dualities and Representations of Lie

Superalgebras

Shun-Jen Cheng
Weiqiang Wang

I NSTITUTE OF M ATHEMATICS , ACADEMIA S INICA , TAIPEI , TAIWAN 10617


E-mail address: chengsj@math.sinica.edu.tw

D EPARTMENT OF M ATHEMATICS , U NIVERSITY OF V IRGINIA , C HARLOTTESVILLE ,


VA 22904
E-mail address: ww9c@virginia.edu
2000 Mathematics Subject Classification. Primary

NSC.
NSF.

A BSTRACT. Abstract here.


Contents

Preface ix

Chapter 1. Lie superalgebra ABC 1


§1.1. Lie superalgebras: Definition and Examples 1
1.1.1. Basic definitions 2
1.1.2. The general and special linear Lie superalgebras 4
1.1.3. The ortho-symplectic Lie superalgebras 6
1.1.4. The queer Lie superalgebras 8
1.1.5. The periplectic and exceptional Lie superalgebras 9
1.1.6. The Cartan series 10
1.1.7. The classification theorem 12
§1.2. Structures of classical Lie superalgebras 13
1.2.1. A basic structure theorem 13
1.2.2. Invariant bilinear forms for gl and osp 15
1.2.3. Root system and Weyl group for gl(m|n) 16
1.2.4. Root system and Weyl group for spo(2m|2n + 1) 16
1.2.5. Root system and Weyl group for spo(2m|2n) 17
1.2.6. Root system and odd invariant form for q(n) 17
§1.3. Non-conjugate positive systems and odd reflections 19
1.3.1. Positive systems and fundamental systems 19
1.3.2. Positive and fundamental systems for gl(m|n) 21
1.3.3. Positive and fundamental systems for spo(2m|2n + 1) 22
1.3.4. Positive and fundamental systems for spo(2m|2n) 23
1.3.5. Conjugacy classes of fundamental systems 25
1.3.6. Odd reflections 26
§1.4. Highest weight theory 30

iii
iv Contents

1.4.1. The Poincaré-Birkhoff-Witt (PBW) Theorem 30


1.4.2. Representations of solvable Lie superalgebras 31
1.4.3. Highest weight theory for basic Lie superalgebras 32
1.4.4. Highest weight theory for q(n) 34
§1.5. Notes 36

Chapter 2. Finite-dimensional modules 39


§2.1. Classification of finite-dimensional simple modules 39
2.1.1. Finite-dimensional simple modules of gl(m|n) 39
2.1.2. Finite-dimensional simple modules of spo(2m|2) 41
2.1.3. A virtual character formula 41
2.1.4. Finite-dimensional simple modules of spo(2m|2n + 1) 43
2.1.5. Finite-dimensional simple modules of spo(2m|2n) 46
2.1.6. Finite-dimensional simple modules of q(n) 48
§2.2. Harish-Chandra homomorphism and linkage 50
2.2.1. Supersymmetrization 50
2.2.2. The central characters 51
2.2.3. Harish-Chandra homomorphism for basic Lie superalgebras 52
2.2.4. Invariant polynomials for gl and osp 54
2.2.5. Image of Harish-Chandra homomorphism for gl and osp 56
2.2.6. Linkage for gl and osp 59
2.2.7. Typical finite-dimensional irreducible characters 63
§2.3. Harish-Chandra homomorphism and linkage for q(n) 64
2.3.1. The central characters for q(n) 64
2.3.2. Harish-Chandra homomorphism for q(n) 65
2.3.3. Linkage for q(n) 68
2.3.4. Typical finite-dimensional characters of q(n) 70
§2.4. Extremal weights of finite-dimensional simple modules 71
2.4.1. Extremal weights for gl(m|n) 72
2.4.2. Extremal weights for spo(2m|2n + 1) 75
2.4.3. Extremal weights for spo(2m|2n) 77
§2.5. Notes 80

Chapter 3. Schur duality 83


§3.1. Generalities for associative superalgebras 83
3.1.1. Classification of simple superalgebras 84
3.1.2. Wedderburn theorem and Schur’s Lemma 86
3.1.3. Double centralizer property for superalgebras 87
3.1.4. Split conjugacy classes in a finite supergroup 88
§3.2. Schur-Sergeev duality of type A 90
3.2.1. Schur-Sergeev duality, I 90
Contents v

3.2.2. Schur-Sergeev duality, II 92


3.2.3. The character formula 95
3.2.4. The classical Schur duality 97
3.2.5. Degree of atypicality of λ♮ 98
3.2.6. Category of polynomial modules 100
§3.3. Representation theory of the algebra Hn 101
3.3.1. A double cover 102
3.3.2. Split conjugacy classes in Ben 103
3.3.3. A ring structure on R− 106
3.3.4. The characteristic map 108
3.3.5. The basic spin module 109
3.3.6. The irreducible characters 111
§3.4. Schur-Sergeev duality for q(n) 113
3.4.1. A double centralizer property 113
3.4.2. The Sergeev duality 114
3.4.3. The irreducible character formula 116
§3.5. Notes 117

Chapter 4. Classical invariant theory 119


§4.1. FFT for the general linear Lie group 119
4.1.1. General invariant theory 119
4.1.2. Tensor and multilinear FFT for GL(V ) 120
4.1.3. Formulation of the polynomial FFT for GL(V ) 122
4.1.4. Polarization and restitution 123
§4.2. Polynomial FFT for classical groups 124
4.2.1. A reduction theorem of Weyl 125
4.2.2. The symplectic and orthogonal groups 126
4.2.3. Formulation of the polynomial FFT 128
4.2.4. From basic to general polynomial FFT 129
4.2.5. The basic case 129
§4.3. Tensor and supersymmetric FFT for classical groups 132
4.3.1. Tensor FFT for classical groups 133
4.3.2. From tensor FFT to supersymmetric FFT 135
§4.4. Notes 138

Chapter 5. Howe duality 139


§5.1. Weyl-Clifford algebra and classical Lie superalgebras 140
5.1.1. The Weyl-Clifford algebra 140
5.1.2. A filtration on Weyl-Clifford algebra 142
5.1.3. Relation to classical Lie superalgebras 143
5.1.4. A general duality theorem 145
vi Contents

5.1.5. A duality for Weyl-Clifford algebras 147


§5.2. Howe duality for type A and type Q 148
5.2.1. Howe dual pair (GL(k), gl(m|n)) 148
5.2.2. (GL(k), gl(m|n))-Howe duality 150
5.2.3. Formulas for highest weight vectors 152
5.2.4. (q(m), q(n))-Howe duality 154
§5.3. Howe duality for symplectic and orthogonal groups 157
5.3.1. Howe dual pair (Sp(V ), osp(2m|2n)) 157
5.3.2. (Sp(V ), osp(2m|2n))-Howe duality 160
5.3.3. Irreducible modules of O(V ) 163
5.3.4. Howe dual pair (O(k), spo(2m|2n)) 164
5.3.5. (O(V ), spo(2m|2n))-Howe duality 166
§5.4. Howe duality for infinite-dimensional Lie algebras 168
5.4.1. Lie algebras a∞ , c∞ and d∞ 168
5.4.2. The fermionic Fock space 171
5.4.3. (GL(ℓ), a∞ )-Howe duality 172
5.4.4. (Sp(k), c∞ )-Howe duality 175
5.4.5. (O(k), d∞ )-Howe duality 177
§5.5. Character formula for Lie superalgebras 179
5.5.1. Characters for modules of Lie algebras c∞ and d∞ 179
5.5.2. Characters for oscillator osp(2m|2n)-modules 181
5.5.3. Characters for oscillator spo(2m|2n)-modules 182
§5.6. Notes 184
Chapter 6. Super duality 187
§6.1. Lie superalgebras of classical types 187
6.1.1. Head, tail, and master diagrams 188
6.1.2. The index sets 189
6.1.3. Infinite-rank Lie superalgebras 190
6.1.4. The case of m = 0 193
6.1.5. Finite-dimensional Lie superalgebras 194
6.1.6. Central extensions 194
§6.2. The module categories 196
6.2.1. Category of polynomial modules revisited 196
6.2.2. Parabolic subalgebras and dominant weights 199
6.2.3. The categories O, O and Oe 200
6.2.4. en
The categories On , On and O 202
6.2.5. Truncation functors 202
§6.3. The irreducible character formulas 203
6.3.1. Two sequences of Borel subalgebras of e
g 204
6.3.2. Odd reflections and highest weight modules 206
Contents vii

6.3.3. The functors T and T 209


6.3.4. Character formulas 212
§6.4. Kostant homology and KLV polynomials 213
6.4.1. Homology and cohomology of Lie superalgebras 214
6.4.2. Kostant u− -homology and u-cohomology 216
6.4.3. Comparison of Kostant homology groups 218
6.4.4. Kazhdan-Lusztig-Vogan (KLV) polynomials 220
6.4.5. Stability of KLV polynomials 221
§6.5. Super duality as an equivalence of categories 222
6.5.1. Extensions à la Baer-Yoneda 223
6.5.2. e
Relating extensions in O, O, and O 225
6.5.3. f f
Categories O , O and O e f 228
6.5.4. Lifting highest weight modules 228
6.5.5. Super duality and strategy of proof 229
6.5.6. The proof of super duality 231
§6.6. Notes 236

Appendix A. Symmetric functions 239


§A.1. The ring Λ and Schur functions 239
A.1.1. The ring Λ 239
A.1.2. Schur functions 243
A.1.3. Skew Schur functions 246
A.1.4. The Frobenius characteristic map 248
§A.2. Supersymmetric Polynomials 249
A.2.1. The ring of supersymmetric polynomials 249
A.2.2. Super Schur functions 251
§A.3. The ring Γ and Schur Q-functions 253
A.3.1. The ring Γ 253
A.3.2. Schur Q-functions 255
A.3.3. The inner product on Γ 256
A.3.4. A characterization of Γ 258
§A.4. Boson-Fermion Correspondence 259
A.4.1. The Maya diagrams 259
A.4.2. Partitions 259
A.4.3. Femions and fermionic Fock space 261
A.4.4. Charge and energy 263
A.4.5. From Boson to Fermion 264
A.4.6. Jacobi triple product identity 266
§A.5. Notes 266

Bibliography 267
viii Contents

Index 273
Preface

Lie algebras, Lie groups, and their representation theories are parts of the math-
ematical language to describe symmetries, and they have played a central role in
modern mathematics. An early motivation of studying Lie superalgebras as a gen-
eralization of Lie algebras came from supersymmetry in mathematical physics.
Ever since a Cartan-Killing type classification of finite dimensional complex Lie
superalgebra was obtained by Kac [K1] in 1977, the theory of Lie superalgebras
has established itself as a prominent subject in modern mathematics. An inde-
pendent classification of the finite-dimensional complex simple Lie superalgebras
whose even subalgebras are reductive (called the basic Lie superalgebras) was
given by Scheunert, Nahm, and Rittenberg in [SNR].
The goal of this book is a systematic account of the structure and representation
theory of finite dimensional complex Lie superalgebras of classical type. The book
intends to serve as a rigorous introduction to representation theory of Lie super-
algebras on one hand, and on the other hand, it covers a new approach developed
in the past few years toward understanding the Bernstein-Gelfand-Gelfand cate-
gory for classical Lie superalgebras. In spite of much interest in representations
of Lie superalgebras stimulated by mathematical physics, these basic topics have
not been treated in depth in a book form before. The reason seems to be that the
representation theory of Lie superalgebras is dramatically different from complex
semisimple Lie algebras, and a systematic yet accessible approach toward the basic
problem of finding irreducible characters for Lie superalgebras was not available
in a great generality until very recently.
We are consciously aware that there is an enormous literature and numerous
partial results for Lie superalgebras, and it is not our intention to make this book
an encyclopedia. Rather, we treat in depth the representation theory of the three

ix
x Preface

most important classes of Lie superalgebras, namely, the general linear Lie super-
algebras gl(m|n), the ortho-symplectic Lie superalgebras osp(m|2n), and the queer
Lie superalgebras q(n). A prototype of this book is the lecture notes [CW5] by the
authors. The presentation in this book is organized around three dualities with a
unifying theme of determining irreducible characters:

Schur duality, Howe duality, Super duality.

There are two superalgebra generalizations of Schur duality. The first one,
due to Sergeev [Sv1] and independently Berele-Regev [BeR], is an interplay be-
tween the general linear Lie superalgebra gl(m|n) and the symmetric group which
incorporates the trivial and sign modules in a unified framework. The irreducible
polynomial characters of gl(m|n) arising this way are given by the super Schur
polynomials. The second one, called Sergeev duality, is an interplay between the
queer Lie superalgebra q(n) and a twisted hyperoctahedral group algebra. The
Schur Q-functions and related combinatorics of shifted tableaux appear naturally
in the description of the irreducible polynomial characters of q(n).
It has been observed that much of the classical invariant theory for the polyno-
mial algebra has a parallel theory for the exterior algebra as well. The First Funda-
mental Theorem (FFT) for both polynomial invariants and skew invariants for clas-
sical groups admits natural reformulation and extension in the theory of Howe’s
reductive dual pairs [H1, H2]. Lie superalgebras allow an elegant and uniform
treatment of Howe duality on the polynomial and exterior algebras (cf. Cheng-
Wang [CW1]). For the general linear Lie groups, Schur duality, Howe duality
and FFT are equivalent. Unlike Schur duality, Howe duality treats classical Lie
groups and (super)algebras beyond type A equally well. The Howe dualities allow
us to determine the character formulas for the irreducible modules appearing in the
dualities.
The third duality, super duality, has a completely different flavor. It views
representation theories of Lie superalgebras and Lie algebras as two sides of the
same coin, and it is an unexpected and rather powerful approach developed in the
past few years by the authors and their collaborators, culminating in Cheng-Lam-
Wang [CLW]. The super duality approach allows one to overcome in a conceptual
way various major obstacles in the super representation theory, via an equivalence
of module categories of Lie algebras and Lie superalgebras.
Schur, Howe, and super dualities provide approaches to the irreducible charac-
ter problem in increasing generality and sophistication. Schur and Howe dualities
only offer a solution to the irreducible character problem for modules in some
semisimple subcategories. On the other hand, super duality provides a concep-
tual solution to the long-standing irreducible character problem in fairly general
Preface xi

BGG categories (including all finite-dimensional modules) over classical Lie su-
peralgebras in terms of the usual Kazhdan-Lusztig polynomials of classical Lie al-
gebras. Totally different approaches to the irreducible character problem of finite-
dimensional gl(m|n)-modules have been developed Serganova [Sva] and Brundan
[Br1].
The book is largely self-contained, and should be accessible to graduate stu-
dents and non-experts as well. Besides assuming basic knowledge of entry-level
graduate algebra, the only other prerequisite is a one-semester course in the the-
ory of finite-dimensional semisimple Lie algebras. For example, either the book
by Humphreys or the first half of the book by Carter on semisimple Lie algebras is
more than sufficient. Some familiarity with symmetric functions can be useful, and
Appendix A provides a quick summary for our purpose. It is possible that super
experts may also benefit from the book, as several “folklore” results are clarified
and occasionally corrected in great detail here, sometimes with new proofs. The
proofs of some of these results can be at times rather difficult to trace or read in the
literature (not merely because they might be in a different language).
Here is a broad outline of the book chapter by chapter. Each chapter ends with
some historical notes and references. Though we have tried to attribute the main
results accurately and fairly, we apologize beforehand for any unintended omission
and mistakes.
Chapter 1 starts with defining various classes of Lie superalgebras. For the
basic Lie superalgebras, we introduce the invariant bilinear forms, root systems,
fundamental systems, and Weyl groups. Positive systems and fundamental systems
for basic Lie superalgebras are not conjugate under the Weyl group, and the notion
of odd reflections is introduced to relate non-conjugate positive systems. The PBW
theorem for the universal enveloping algebra of a Lie superalgebra is formulated,
and highest weight theory for basic Lie superalgebras and q(n) is developed.
In Chapter 2, we focus on Lie superalgebras of types gl, osp and q. We classify
their finite-dimensional simple modules using odd reflection techniques. We then
formulate and establish precisely the image of Harish-Chandra homomorphism and
linkage principle. We end with a Young diagramatic description of the extremal
weights in the simple polynomial gl(m|n)-modules and finite-dimensional simple
osp(m|2n)-modules. It takes considerably more effort to formulate and prove these
results for Lie superalgebras in contrast to semisimple Lie algebras, because of the
existence of non-conjugate Borel subalgebras and the limited role of Weyl groups
for Lie superalgebras.
Schur duality for Lie superalgebras is developed in Chapter 3. We start with
some results on the structure of associative superalgebras including the super vari-
ants of Wedderburn theorem, Schur’s lemma, and double centralizer property. The
Schur-Sergeev duality for gl(m|n) is proved, and it provides a classification of ir-
reducible polynomial gl(m|n)-modules. As a consequence, the characters of the
xii Preface

simple polynomial gl(m|n)-modules are given by the super Schur polynomials. On


the algebraic combinatorial level, there is a natural super generalization of the no-
tion of semistandard tableaux which is a hybrid of the traditional version and its
conjugate counterpart. The Schur-Sergeev duality for q(n) requires understanding
the representation theory of a twisted hyperoctahedral group algebra, which we
develop from scratch. The characters of the simple polynomial q(n)-modules are
given by the Schur Q-polynomials up to some 2-powers.
In Chapter 4, we give a quick introduction to classical invariant theory, which
serves as a preparation for Howe duality in the next chapter. We describe sev-
eral versions of the FFT for the classical groups, i.e., a tensor algebra version, a
polynomial algebra version, and a supersymmetric algebra version.
Howe duality is the main topic of Chapter 5. Like Schur duality, Howe duality
involves commuting actions of a classical Lie group G and a classical superalgebra
g′ on a supersymmetric algebra. The precise relation between the classical Lie su-
peralgebras and Weyl-Clifford algebras WC is established. According to the FFT
for classical invariant theory in Chapter 4, when applied to the G-action on the
associated graded algebra gr WC, the basic invariants generating (gr WC)G turn
out to form the associated graded space for a Lie superalgebra g′ . From this it
follows that the algebra of G-invariants WCG is generated by g′ . Multiplicity-free
decompositions for various (G, g′ )-Howe dualities are obtained explicitly. Charac-
ter formulas for the irreducible g′ -modules appearing in (G, g′ )-Howe duality are
then obtained, via a comparison with Howe duality involving classical groups G
and infinite-dimensional Lie algebras, which we develop in detail.
Finally in Chapter 6, we develop a super duality approach to obtain a complete
and conceptual solution of the irreducible character problem in certain parabolic
Bernstein-Gelfand-Gelfand categories for general linear and ortho-symplectic Lie
superalgebras. Super duality is an equivalence of categories between parabolic
categories for Lie superalgebras and their Lie algebra counterparts at an infinite-
rank limit, and it matches the corresponding Verma modules, irreducible modules,
Kostant u-homology groups, and Kazhdan-Lusztig-Vogan polynomials. Trunca-
tion functors are introduced to relate the BGG categories for infinite-rank and
finite-rank Lie superalgebras. In this way, we obtain a solution à la Kazhdan-
Lusztig of the irreducible character problem in the corresponding parabolic BGG
categories for finite-dimensional basic Lie superalgebras.
There is an appendix in the book. In Appendix A, we have included a fairly
self-contained treatment of some elementary aspects of symmetric function the-
ory, including Schur functions, supersymmetric functions and Schur Q-functions.
The celebrated boson-fermion correspondence serves as a prominent example re-
lating superalgebras to mathematical physics and algebraic combinatorics. The
Fock space therein is used in setting up the Howe duality for infinite-dimensional
Lie algebras in Chapter 5.
Preface xiii

Acknowledgment. The book project started with the lecture notes [CW5]
of the authors, which was an expanded written account for a series of lectures
delivered by the second-named author in the summer school at East China Normal
University, Shanghai, in July 2009. In a graduate course at University of Virginia in
Spring 2010, the second-named author lectured on what became a large portion of
Chapters 3, 4 and 5 of the book. The materials of Chapter 3 on Schur duality have
been used by the second-named author in the winter school in Taipei in December
2010. Part of the materials of the first three chapters have also been used by the
first-named author in a lecture series in Shanghai in March 2011, and then in a
lecture series by both authors in a workshop in Tehran in May 2011. We thank the
participants in all these occasions for their helpful suggestions and feedbacks.

Shun-Jen Cheng
Weiqiang Wang
Chapter 1

Lie superalgebra ABC

We start by introducing the basic notions and definitions in the theory of Lie super-
algebras, such as basic Lie superalgebras and queer Lie superalgebras, Cartan sub-
algebras, Borel subalgebras, root systems, positive and fundamental systems. We
formulate the main structure results for the basic Lie superalgebras and the queer
Lie superalgebras. We describe in detail the structures of Lie superalgebras of type
gl, osp and q. A distinguished feature for Lie superalgebras is that Borel subal-
gebras, positive systems, or fundamental systems of a simple finite-dimensional
Lie superalgebra may not be conjugate under the action of the corresponding Weyl
group; instead, they are shown to be related to each other by real and odd reflec-
tions. A highest weight theory is developed for Lie superalgebras. We describe
how fundamental systems are related and how highest weights are transformed by
an odd reflection.

1.1. Lie superalgebras: Definition and Examples


Throughout this book we will assume that our field is C, the field of complex
numbers. The symbol Z2 = {0̄, 1̄} will stand for the group of two elements and Sn
stands for the symmetric group in n letters.
In this section, we introduce many examples of Lie superalgebras. The ex-
amples, most relevant to this book, of the general linear and ortho-symplectic Lie
superalgebras are introduced first. Other series of finite-dimensional simple Lie

1
2 1. Lie superalgebra ABC

superalgebras of classical type, namely, the queer and the periplectic Lie super-
algebras, along with the three exceptional ones, are then described. The finite-
dimensional Cartan type Lie superalgebras are then realized explicitly as subal-
gebras of the Lie superalgebra of polynomial vector fields on a purely odd di-
mensional superspace. The section ends with Kac’s classification theorem of the
finite-dimensional simple Lie superalgebras over C, which we state without proof.

1.1.1. Basic definitions. A vector superspace V is a vector space endowed with


a Z2 -gradation: V = V0̄ ⊕V1̄ . The dimension of the vector superspace V is the tu-
ple dimV = (dimV0̄ | dimV1̄ ). The superdimension of V is defined to be dimV0̄ −
dimV1̄ and is denoted by sdimV . We denote the superspace with even subspace
Cm and odd subspace Cn by Cm|n . It has dimension (m|n). The parity of a ho-
mogeneous element a ∈ Vi is denoted by |a| = i, i ∈ Z2 . An element in V0̄ is
called even, while an element in V1̄ is called odd. A subspace of a vector su-
perspace V = V0̄ ⊕ V1̄ is a vector superspace W = W0̄ ⊕ W1̄ ⊆ V with compatible
Z2 -gradation, i.e., Wi ⊆ Vi , for i ∈ Z2 .
Let V be a superspace. Throughout the book, when we write |v| for an element
v ∈ V , we will always implicitly assume that v is a homogeneous element, and au-
tomatically extend the relevant formulas by linearity (whenever applicable). Also,
note that if V and W are superspaces, then the space of linear transformations from
V to W is naturally a vector superspace. In particular, the space of endomorphisms
of V , denoted by End(V ), is a vector superspace. When V = Cm|n we write I = Im|n
for the identity matrix on V .
There is a parity reversing functor Π on the category of vector superspaces: for
a vector superspace V = V0̄ ⊕V1̄ , we let
Π(V ) = Π(V )0̄ ⊕ Π(V )1̄ , Π(V )i = Vi+1̄ , ∀i ∈ Z2 .
Clearly, Π2 = I.

Definition 1.1. A superalgebra A, sometimes also called a Z2 -graded algebra,


is a vector superspace A = A0̄ ⊕ A1̄ , equipped with a multiplication satisfying
Ai A j ⊆ Ai+ j , for i, j ∈ Z2 . A module M over a superalgebra A is always under-
stood in the Z2 -graded sense, that is M = M0̄ ⊕ M1̄ such that Ai M j ⊆ Mi+ j , for
i, j ∈ Z2 . Subalgebras and ideals of superalgebras are also understood in the Z2 -
graded sense. A superalgebra that has no non-trivial ideal is called simple. A
homomorphism between A-modules M and N is a linear map f : M → N satisfy-
ing that f (am) = a f (m), for all a ∈ A, m ∈ M.

A homomorphism between modules M and N of a superalgebra A is sometimes


understood in the literature as a linear map f : M → N of parity | f | ∈ Z2 which
satisfies (⋆) f (am) = (−1)|a|·| f | a f (m), for homogeneous a ∈ A, m ∈ M. Let us call
such a map a ⋆-homomorphism. These two definitions can be converted to each
1.1. Lie superalgebras: Definition and Examples 3

other as follows. Given a homomorphism f : M → N of degree | f | in the sense of


Definition 1.1, we define f † : M → N by the formula
(1.1) f † (x) := (−1)| f |·|x| f (x).
Then f † is a ⋆-homomorphism. Conversely, (1.1) also converts a ⋆-homomorphism
into a homomorphism.
Now we come to the definition of the main object of our study.
Definition 1.2. A Lie superalgebra is a superalgebra g = g0̄ ⊕ g1̄ with multiplica-
tion [., .] satisfying the following two axioms: (for homogeneous a, b, c ∈ g)
(1) Skew-supersymmetry: [a, b] = −(−1)|a|·|b| [b, a].
(2) Super Jacobi identity: [a, [b, c]] = [[a, b], c] + (−1)|a|·|b| [b, [a, c]].

A bilinear form (·, ·) : g × g → C on a Lie superalgebra g is called invariant if


([a, b], c) = (a, [b, c]), for all a, b, c ∈ g.
For a Lie superalgebra g = g0̄ ⊕ g1̄ , the even part g0̄ is a Lie algebra. Hence,
if g1̄ = 0, then g is just a usual Lie algebra. A Lie superalgebra g with purely odd
part, i.e., g0̄ = 0, has to be abelian, i.e., [g, g] = 0.
Definition 1.3. Let g and g′ be Lie superalgebras. A homomorphism of Lie su-
peralgebras is an even linear map f : g → g′ satisfying
f ([a, b]) = [ f (a), f (b)], a, b ∈ g.
Example 1.4. (1) Let A = A0̄ ⊕ A1̄ be an associative superalgebra. We can make A
into a Lie superalgebra by letting
[a, b] := ab − (−1)|a|·|b| ba,
for homogeneous a, b ∈ A, and extending [., .] by bi-linearity.
(2) Let g be a Lie superalgebra. Then End(g) is an associative superalgebra,
and hence it carries a structure of a Lie superalgebra by (1). We define the adjoint
map ad : g → End(g) by
ad (a)(b) := [a, b], a, b ∈ g.
Then ad is a homomorphism of Lie superalgebras due to the super Jacobi identity.
The resulting action of g on itself is called the adjoint action.
(3) Let A = A0̄ ⊕ A1̄ be an associative superalgebra. An endomorphism D ∈
End(A)s is called a derivation of degree s, if it satisfies the identity
D(ab) = D(a)b + (−1)s|a| aD(b), a, b ∈ A.
Denote by Der(A)s the space of derivations on A of degree s. One verifies that the
superspace of derivations of A, Der(A) = Der(A)0̄ ⊕ Der(A)1̄ , is a subalgebra of the
Lie superalgebra (End(A), [·, ·]).
4 1. Lie superalgebra ABC

(4) Let g = g0̄ ⊕ g1̄ be a superspace such that g0̄ = Cz is one-dimensional.


Suppose that we have a symmetric bilinear form B(·, ·) on g1̄ . We can make g into
a Lie superalgebra by letting z commute with g and declaring
[v, w] := B(v, w)z, v, w ∈ g1̄ .
The special cases when B(·, ·) is zero, or when B(·, ·) is non-degenerate respec-
tively, are of particular interest, and they roughly correspond to the exterior and
Clifford superalgebras of Section 1.1.6 and (3.19), respectively.

For a Lie superalgebra g = g0̄ ⊕ g1̄ , the restriction of the adjoint homomor-
phism ad |g0̄ : g0̄ → End(g1̄ ) is a homomorphism of Lie algebras, namely, g1̄ is a
g0̄ -module under the adjoint action.
Remark 1.5. To a Lie superalgebra g = g0̄ ⊕ g1̄ we associate the following data:
(1) A Lie algebra g0̄ .
(2) A g0̄ -module g1̄ induced by the adjoint action.
(3) A g0̄ -homomorphism S2 (g1̄ ) → g0̄ induced by the Lie bracket.
(4) The condition coming from Definition 1.2(2) with a, b, c ∈ g1̄ .
Conversely, the above data determine a Lie superalgebra structure on g0̄ ⊕ g1̄ .

1.1.2. The general and special linear Lie superalgebras. Let V = V0̄ ⊕ V1̄ be
a vector superspace so that End(V ) is an associative superalgebra. As in Ex-
ample 1.4, End(V ), equipped with the supercommutator, forms a Lie superalge-
bra, called the general linear Lie superalgebra and is denoted by gl(V ). When
V = Cm|n we also write gl(m|n) for gl(V ).
Choose bases for V0̄ and V1̄ that combine to a homogeneous basis for V . We
will make it a convention to parameterize such a basis by the set
(1.2) I(m|n) = {1, . . . , m; 1, . . . , n}
with total order
(1.3) 1 < . . . < m < 0 < 1 < . . . < n.
Here 0 is inserted for notational convenience later on. The elementary matrices
are accordingly denoted by Ei j , with i, j ∈ I(m|n). With respect to such an ordered
basis, End(V ) and gl(V ) can be realized as (m + n) × (m + n) complex matrices of
the block form
( )
a b
(1.4) g= ,
c d
where a, b, c, and d are respectively m × m, m × n, n × m, and n × n matrices. The
even subalgebra gl(V )0̄ consists of matrices of the form (1.4) with b = c = 0, while
the odd subspace gl(V )1̄ consists of those with a = d = 0. In particular, gl(V )0̄ ∼
=
1.1. Lie superalgebras: Definition and Examples 5

gl(m) ⊕ gl(n), and as a gl(V )0̄ -module, gl(V )1̄ is self-dual and is isomorphic to
(Cm ⊗ Cn∗ ) ⊕ (Cm∗ ⊗ Cn ). Here and below, Cn∗ denotes the dual space of Cn .
For each element g ∈ gl(m|n) of the form (1.4) we define the supertrace
str(g) := tr(a) − tr(d),
where tr(x) denotes the trace of the square matrix x. One checks that
str([g, g′ ]) = 0, for g, g′ ∈ gl(m|n).
Thus, the subspace
sl(m|n) := {g ∈ gl(m|n) | str(g) = 0}
is a subalgebra of gl(m|n), called the special linear Lie superalgebra. One verifies
directly that [gl(m|n), gl(m|n)] = sl(m|n). Furthermore, sl(m|n) ∼ = sl(n|m), and
when m ̸= n, sl(m|n) is simple. When m = n, sl(m|m) contains a non-trivial center
generated by the identity matrix Im|m . For m ≥ 2, sl(m|m)/CIm|m is simple.
Example 1.6. Let g = gl(1|1) and consider the following basis for g:
( ) ( ) ( ) ( )
0 1 0 0 1 0 0 0
e= , f= , E1̄.1̄ = , E11 = .
0 0 1 0 0 0 0 1
Set h := E1̄,1̄ + E11 = I1|1 . Then h is central, [e, f ] = h, and sl(1|1) has a basis
{e, h, f }.

Let I = I0̄ ⊔ I1̄ be a parametrization of a homogeneous basis of the superspace


V = V0̄ ⊕ V1̄ , where I0̄ and I1̄ parameterize the corresponding bases of V0̄ and V1̄ ,
respectively. For an element i in I we define
{
0, if i ∈ I0̄ ,
|i| :=
1, if i ∈ I1̄ .
For example, for the parametrization I(m|n) with ordering (1.3), we have |i| = 0
for i < 0, and |i| = 1 for i > 0. Choosing a total ordering of the homogeneous
basis we may identify gl(V ) with the space of |I| × |I| matrices. For such a matrix
A = ∑i, j∈I ai j Ei j , ai j ∈ C, we define the supertranspose of A to be

(1.5) Ast := ∑ (−1)| j|(|i|+| j|) ai j E ji .


i, j∈I

We define Chevalley automorphism τ : gl(V ) → gl(V ) by the formula


(1.6) τ(A) := −Ast .
It is straightforward to check that τ is an automorphism of Lie superalgebras. We
note that τ restricts to an automorphism of sl(V ). Also, for m, n ̸= 0, τ has order 4,
and hence is in general not an involution.
6 1. Lie superalgebra ABC

1.1.3. The ortho-symplectic Lie superalgebras.

Definition 1.7. Let V = V0̄ ⊕V1̄ be a vector superspace. A bilinear form

B(·, ·) : V ×V −→ V

is called even (respectively, odd), if B(Vi ,V j ) = 0 unless i + j = 0̄ (respectively,


i + j = 1̄). An even bilinear form B is said to be supersymmetric, if B|V0̄ ×V0̄ is
symmetric and B|V1̄ ×V1̄ is skew-symmetric, and it is called skew-supersymmetric,
if B|V0̄ ×V0̄ is skew-symmetric and B|V1̄ ×V1̄ is symmetric.

Let B be a non-degenerate even supersymmetric bilinear form on a vector su-


perspace V = V0̄ ⊕V1̄ . It follows that dimV1̄ is necessarily even. For s ∈ Z2 , let

osp(V )s := {g ∈ gl(V )s | B(g(x), y) = −(−1)s·|x| B(x, g(y)), ∀x, y ∈ V },


osp(V ) := osp(V )0̄ ⊕ osp(V )1̄ .

One checks that osp(V ) is a Lie superalgebra, called the ortho-symplectic Lie
superalgebra, that is, osp(V ) is the subalgebra of gl(V ) that preserves a non-
degenerate supersymmetric bilinear form. Its even subalgebra is isomorphic to
so(V0̄ ) ⊕ sp(V1̄ ), a direct sum of the orthogonal Lie algebra on V0̄ and the symplec-
tic Lie algebra on V1̄ . When V = Cℓ|2m , we write osp(ℓ|2m) for osp(V ). Note that
when ℓ or m is zero, the ortho-symplectic Lie superalgebra reduces to classical Lie
algebra sp(2m) or so(ℓ), respectively.
Similarly, we define the Lie superalgebra spo(V ) as the subalgebra of gl(V )
that preserves a non-degenerate skew-supersymmetric bilinear form on V (here
dimV0̄ has to be even). When V = C2m|ℓ , we write spo(2m|ℓ) for spo(V ).
A supersymmetric bilinear form B on V induces a skew-supersymmetric bilin-
ear form C(·, ·) on Π(V ) defined by C(Π(v), Π(v′ )) := B(v, v′ ), for v, v′ ∈ V , where
Π is the parity reversing functor defined in Section 1.1.1. Define an isomorphism of
Lie superalgebras from End(V ) to End(ΠV ) by sending T to ΠT Π−1 . One checks
that the restriction of the map to osp(V ) sends osp(V ) to spo(ΠV ). It follows that
osp(ℓ|2m) ∼ = spo(2m|ℓ) as Lie superalgebras.
We now give an explicit matrix realization of the ortho-symplectic Lie super-
algebra. To this end, we first observe that the supertranspose (1.5) of a matrix in
the block form (1.4), is equal to
( )st ( )
a b at ct
= ,
c d −bt dt

where xt denotes the usual transpose of a matrix x.


1.1. Lie superalgebras: Definition and Examples 7

Define the (2m + 2n + 1) × (2m + 2n + 1) matrix in the (m|m|n|n|1)-block form


 
0 Im 0 0 0
−Im 0 0 0 0
 
(1.7) J2m|2n+1 :=  0 0 0 In 0 .
 0 0 In 0 0
0 0 0 0 1
Let J2m|2n denote the (2m + 2n) × (2m + 2n) matrix obtained from J2m|2n+1 by
deleting the last row and column. For ℓ = 2n or 2n + 1, by definition spo(2m|ℓ)
is the subalgebra of gl(2m|ℓ) that preserves the bilinear form on C2m|ℓ with matrix
J2m|ℓ relative to the standard basis of C2m|ℓ , and hence
spo(2m|ℓ) = {g ∈ gl(2m|ℓ) | gst J2m|ℓ + J2m|ℓ g = 0}.
By a direct computation, spo(2m|2n + 1) consists of the (2m + 2n + 1) × (2m +
2n + 1) matrices of the following (m|m|n|n|1)-block form
 
d e yt1 xt1 zt1
 f −d t −yt −xt −zt 
 
(1.8)  x x1 a b −vt   , b, c skew-symmetric, e, f symmetric.
 y y1 c −at −ut 
z z1 u v 0
∼ C2m ⊗ C2n+1 (which is self-dual) as a module over
Note that spo(2m|2n + 1)1̄ =
spo(2m|2n + 1)0̄ ∼
= sp(2m) ⊕ so(2n + 1).
The Lie superalgebra spo(2m|2n) consists of matrices (1.8) with the last row
and column removed. Note that spo(2m|2n)1̄ ∼
= C2m ⊗ C2n (which is self-dual) as
a module over spo(2m|2n)0̄ ∼
= sp(2m) ⊕ so(2n).
Here and below, the rows and columns of the matrices J2m|ℓ and (1.8) (or its
modification) are indexed by the finite set I(2m|ℓ).

Proposition 1.8. The automorphism τ in (1.6) restricts to an automorphism of


spo(2m|ℓ).

Proof. Take an element g ∈ gl(2m|ℓ) that lies in spo(2m|ℓ). Thus, J2m|ℓ g+gst J2m|ℓ =
0 and hence
(1.9) gst Jst2m|ℓ + Jst2m|ℓ (gst )st = 0.

Observe that Jst2m|ℓ J2m|ℓ = I2m|ℓ so that if we multiply (1.9) on the left and on the
right by J2m|ℓ , we get the identity

J2m|ℓ gst + (gst )st J2m|ℓ = 0.

This gives the identity J2m|ℓ τ(g) + τ(g)st J2m|ℓ = 0, and so τ(g) ∈ spo(2m|ℓ). 
8 1. Lie superalgebra ABC

1.1.4. The queer Lie superalgebras. Let V = V0̄ ⊕V1̄ be a vector superspace with
dimV0̄ = dimV1̄ . Choose P ∈ End(V )1̄ such that P2 = In|n . The subspace
q(V ) = {T ∈ End(V ) | [T, P] = 0}
is a subalgebra of gl(V ), called the queer Lie superalgebra. (Different choices of
P give rise to isomorphic queer Lie superalgebras.) If V = Cn|n , then q(V ) is also
denoted by q(n).
In order to give an explicit realization of q(n) in terms of matrices, let us take
P to be the 2n × 2n matrix
( )
√ 0 In
(1.10) P := −1 .
−In 0

(The factor −1 makes no difference here, and it is introduced for later conve-
nience.) Then, for g ∈ gl(n|n) of the form (1.4) we have g ∈ q(n) if and only if
gP − (−1)|g| Pg = 0, and in turn, if and only if
( )
a b
(1.11) g= ,
b a
where a, b are arbitrary complex n × n matrices. Thus we have q(n)0̄ ∼ = gl(n) as Lie
algebras, and q(n) ∼
1̄ = gl(n) as the adjoint q(n) -module. A linear basis for q(n)

consists of the following elements
Eei j := Eī j¯ + Ei j , E i j := Ei j¯ + Eī j , 1 ≤ i, j ≤ n.

The derived superalgebra [q(n), q(n)] consists of matrices of the form (1.11),
with a ∈ gl(n) and b ∈ sl(n), and so it contains a one-dimensional center generated
by the identity matrix In|n . The quotient superalgebra [q(n), q(n)]/CIn|n has even
part isomorphic to sl(n) and odd part isomorphic to the adjoint module, and one
can show that it is simple for n ≥ 3. For n = 2, the odd part of the quotient Lie
superalgebra is an abelian ideal, since the adjoint module of sl(2) does not appear
in the symmetric square of the adjoint module.
Remark 1.9. Consider the subspace e
q(n) of gl(n|n) consisting of elements that
commute with P. That is,
e
q(n) := {g ∈ gl(n|n)|gP − Pg = 0}.
In matrix form e
q(n) consists of the following n|n-block matrices:
( )
a b
(1.12) ,
−b a
where a and b are arbitrary n × n matrices. One checks that e q(n) is closed under
the Lie bracket and hence is a subalgebra of gl(n|n). Indeed e q(n) is isomorphic to
q(n), since the map τ in (1.6) sends q(n) to e
q(n), and vice versa. Thus, (1.12) gives
another realization of the queer Lie superalgebra.
1.1. Lie superalgebras: Definition and Examples 9

1.1.5. The periplectic and exceptional Lie superalgebras. Let us describe more
examples of Lie superalgebras.
The periplectic Lie superalgebras. Let V = V0̄ ⊕ V1̄ be a vector superspace
with dimV0̄ = dimV1̄ . Let C(·, ·) be a non-degenerate odd symmetric bilinear form
on V . One checks that the subspace of gl(V ) preserving C is closed under the Lie
bracket and hence is a Lie subalgebra of gl(V ). This superalgebra is called the
periplectic Lie superalgebra and will be denoted by p(V ). (Different choices of
C give rise to isomorphic periplectic Lie superalgebras.) In the case V = Cn|n , p(V )
is also denoted by p(n).
To write down an explicit matrix realization of p(n) as subalgebra of gl(n|n),
let us take the 2n × 2n matrix
( )
0 In
(1.13) P := ,
In 0
which determines an odd symmetric bilinear form C on Cn|n . Then, g ∈ p(n) if and
only if gst P + Pg = 0. It follows that
{( ) }
a b
(1.14) p(n) = , where b is symmetric and c is skew-symmetric .
c −at
We have
(1.15) p(n)0̄ ∼
= gl(n), p(n)1̄ ∼
= S2 (Cn ) ⊕ ∧2 (Cn∗ ).
For n ≥ 3, the
( derived)superalgebra [p(n), p(n)] is simple, and it consists of matrices
a b
of the form , with tr(a) = 0, bt = b, and ct = −c.
c −at
Remark 1.10. One checks that the Lie subalgebra e p(n) of gl(n|n) preserving the
non-degenerate odd skew-symmetric( bilinear
) form corresponding to P in (1.10)
a b
consists of matrices of the form with bt = −b and ct = c. The superal-
c −at
gebra e
p(n) is isomorphic to p(n).

The exceptional Lie superalgebra D(2, 1, α). We take three copies of the Lie
algebra sl(2) denoted by gi (i = 1, 2, 3), and we associate to each gi a copy of the
standard sl(2)-module Vi .
Clearly as gi -modules we have an isomorphism S2 (Vi ) ∼
= gi . By Schur’s Lemma
we may associate a non-zero scalar αi ∈ C to any such an isomorphism. Now con-
sider g = g0̄ ⊕ g1̄ , where g0̄ ∼
= g1 ⊕ g2 ⊕ g3 , and g1̄ is the irreducible g0̄ -module
V1 ⊗ V2 ⊗ V3 . We can associate three non-zero complex numbers αi , i = 1, 2, 3,
to any surjective g0̄ -homomorphism from S2 (g1̄ ) to g0̄ . Thus our vector super-
space g satisfies Conditions (1)–(3) of Remark 1.5. It is easy to see that Con-
dition (4) of Remark 1.5 is equivalent to ∑3i=1 αi = 0. Thus we obtain a Lie su-
peralgebra g(α1 , α2 , α3 ) depending on three non-zero parameters αi , i = 1, 2, 3.
10 1. Lie superalgebra ABC

For σ ∈ S3 , we have g(α1 , α2 , α3 ) ∼


= g(ασ(1) , ασ(2) , ασ(3) ). Also, g(α1 , α2 , α3 ) ∼
=
g(λα1 , λα2 , λα3 ), for any non-zero λ ∈ C. Thus we have a one-parameter family
of Lie superalgebras D(2, 1, α) := g(α, 1, −1 − α), which are simple for α ̸= 0, −1.
We have g(α, 1, −1−α) ∼ = g(1, α, −1−α) ∼ = g(1, α1 , − α1 −1), and g(α, 1, −1−

α) = g(−1 − α, 1, α), which imply
1
D(2, 1, α) ∼
= D(2, 1, ) ∼
= D(2, 1, −1 − α).
α
The maps α 7→ α1 and α 7→ (−1 − α) generate an action of S3 on C \ {0, −1}.
We have D(2, 1, α) ∼
= D(2, 1, β) if and only if β ∈ S3 · α. This gives additional
isomorphisms:

D(2, 1, α) ∼
= D(2, 1, −(1 + α)−1 α) ∼
= D(2, 1, −1 − α−1 ) ∼
= D(2, 1, −(1 + α)−1 ).

Thus, any orbit of (C \ {0, −1}) /S3 consists of six points, except for the orbit
corresponding to the three points α = 1, −2, − 12 , and the orbit corresponding to the

two points α = − 12 ± − 34 . Finally, note that D(2, 1, 1) ∼= osp(4|2).
The exceptional Lie superalgebra F(4). There is a simple Lie superalgebra
F(4) with F(4)0̄ ∼
= sl(2) ⊕ so(7). The odd part, as an F(4)0̄ -module, is isomorphic
the tensor product of the standard sl(2)-module and the simple so(7)-spin module.
Hence, dim F(4) = (24|16).
The exceptional Lie superalgebra G(3). There is a simple Lie superalgebra
G(3) with G(3)0̄ ∼= sl(2) ⊕ G2 . The odd part as a G(3)0̄ -module is isomorphic to
the tensor product of the standard sl(2)-module and the fundamental 7-dimensional
G2 -module. Hence, dim G(3) = (17|14).

1.1.6. The Cartan series. In this subsection, we describe the Cartan series of
finite-dimensional simple Lie superalgebras without proof. This subsection is in-
cluded for the sake of completeness of classification of finite-dimensional simple
Lie superalgebras, and will not be used in the rest of the book.
Lie superalgebra W (n). Let ∧(n) be the exterior algebra in n indeterminates
ξ1 , ξ2 , . . . , ξn , with a Z2 -gradation given by setting |ξi | = 1̄, for all i. We have
ξi ξ j = −ξ j ξi , for all i, j, and in particular, ξ2i = 0, for all i.
By general construction in Example 1.4, we have a finite-dimensional Lie su-
peralgebra of derivations on ∧(n), which will be denoted by W (n).

For i = 1, . . . , n, a derivation ∂ξi
: ∧(n) → ∧(n) of degree 1̄ is uniquely deter-
mined by

(ξ j ) = δi j , j = 1, . . . , n.
∂ξi
1.1. Lie superalgebras: Definition and Examples 11

Given an element f = ( f1 , f2 , . . . , fn ) ∈ ∧(n)n , with | fi | = | f j |, for all i, j, the linear


map D f : ∧(n) → ∧(n) of the form
n

D f = ∑ fi
i=1 ∂ξi
is a derivation of ∧(n). Furthermore, all homogeneous derivations of ∧(n) are of
this form, since a derivation is determined by its values at ξi for all i. Therefore,
sending f 7→ D f defines a linear isomorphism from ∧(n)n to W (n), and so W (n)
has dimension 2n n.
Setting degξi = 1 and deg ∂ξ∂ i = −1, for all i, gives rise to a Z-gradation on
W (n), called the principal gradation. We have

n−1
W (n) = W (n) j .
j=−1

The Z-gradation is compatible with the super structure on W (n), that is, W (n)s =
⊕ j≡s mod 2W (n) j , for s ∈ Z2 . The 0th degree component W (n)0 is a Lie algebra iso-
morphic to gl(n), and each W (n) j is a gl(n)-module isomorphic to ∧ j+1 (Cn )⊗Cn∗ .
In particular, when n = 2, we have W (2)0 ∼ = gl(2), W (2)−1 ∼
= C2∗ and W (2)1 ∼= C2
as gl(2)-modules. Hence, we have isomorphisms of Lie superalgebras W (2) ∼ =

osp(2|2) = sl(2|1).
The Lie superalgebra W (n) is simple, for n ≥ 2. Moreover, W (n) contains
the following three series of simple Lie superalgebras as subalgebras that we shall
describe.
Lie superalgebra S(n). The first series is the super-analogue of the Lie algebra
of divergence vector fields given by
{ n ∂ n
∂ }
S(n) := ∑ f j ∈ W (n) | ∑ ( f j) = 0 .
j=1 ∂ξ j j=1 ∂ξ j

The Lie superalgebra S(n) is a Z-graded subalgebra of W (n) and we have S(n) =
⊕n−2
j=−1 S(n) j . The Lie algebra S(n)0 is isomorphic to sl(n), and the jth degree
component S(n) j is isomorphic to the top irreducible summand of the sl(n)-module
∧ j+1 (Cn ) ⊗ Cn∗ . The Lie superalgebra S(n) is simple, for n ≥ 3.
e
Lie superalgebra S(n). Let n be even so that ω = 1 + ξ1 ξ2 · · · ξn is a Z2 -
homogeneous invertible element in ∧(n). Consider the subspace of W (n) given
by
{ n }
e := ∑ f j ∂ | ∑ ∂ (ω f j ) = 0 .
n
S(n)
j=1 ∂ξ j j=1 ∂ξ j

e is a subalgebra of Lie superalgebra W (n). The Lie super-


It can be shown that S(n)
e is no longer Z-graded, as the defining condition is not homogeneous.
algebra S(n)
12 1. Lie superalgebra ABC

e inherits a natural filtration from the filtration on W (n) induced by


However, S(n)
e is isomor-
the principal gradation. The associated graded Lie superalgebra of S(n)
e is the following direct sum of vector spaces inside
phic to S(n). Explicitly, S(n)
W (n):

n−2
(1.16) e =
S(n) e j,
S(n)
j=−1

e −1 is spanned by {(1 − ξ1 ξ2 · · · ξn ) ∂ |i = 1, . . . , n}, and S(n)


where S(n) e j = S(n) j ,
∂ξi
e is simple.
for j = 0, . . . , n − 2. For n ≥ 4 and n even, S(n)
Lie superalgebra H(n). As in the classical setting, W (n) contains a subalgebra
H(n) as defined below, which is a super analogue of the Lie algebra of Hamiltonian
vector fields. For f , g ∈ ∧(n) we define a Poisson bracket by
n
∂ f ∂g
{ f , g} := (−1)| f | ∑ .
j=1 ∂ξ j ∂ξ j

The Poisson bracket makes ∧(n) into a Lie superalgebra, which we will denote by
e
H(n). e
Now putting deg f := k − 2, for f ∈ ∧(n)k , H(n) becomes a Z-graded Lie
e
superalgebra. The superalgebra H(n) is not simple, as it has center C1. However,
e
the derived superalgebra of H(n)/C1, which we denote by H(n), is simple, for
⊕n−3
n ≥ 4. Moreover, H(n) = j=−1 H(n) j is a graded Lie superalgebra. The 0th
degree component is a Lie algebra isomorphic to so(n). As an so(n)-module we
have H(n) j ∼= ∧ j+2 (Cn ), for −1 ≤ j ≤ n − 3. Finally, H(n) can be viewed as a
∂f ∂
subalgebra of W (n), since the assignment f 7→ (−1)| f | ∑nj=1 ∂ξ j ∂ξ j
for f ∈ ∧(n),
gives rise to an embedding of Lie superalgebras from H(n) into W (n).

1.1.7. The classification theorem. The following theorem of Kac [K1] gives a
classification of finite-dimensional complex simple Lie superalgebras. Note the
following isomorphisms of simple Lie superalgebras of low ranks:
osp(2|2) ∼
= sl(2|1) ∼
= W (2), sl(2|2)/CI2|2 ∼
= H(4), [p(3), p(3)] ∼
= S(3).
Theorem 1.11. The following is a complete list of pairwise non-isomorphic finite-
dimensional simple Lie superalgebras over C.
(1) A finite-dimensional simple Lie algebra in the Killing-Cartan list.
(2) sl(m|n), for m > n ≥ 1 (excluding (m, n) = (2, 1)); sl(m|m)/CIm|m , for
m ≥ 3; spo(2m|n), for m, n ≥ 1.
( )
(3) D(2, 1, α), for α ∈ C \ {0, ±1, −2, − 12 } /S3 ; F(4); and G(3).
(4) [p(n), p(n)] and [q(n), q(n)]/CIn|n , for n ≥ 3.
e
(5) W (n), for n ≥ 3; S(n), for n ≥ 3; S(2n), for n ≥ 2; and H(n), for n ≥ 4.
1.2. Structures of classical Lie superalgebras 13

The Lie superalgebras g = g0̄ ⊕ g1̄ in Theorem 1.11(1)-(4) have the property
that g0̄ is a reductive Lie algebra and that the adjoint g0̄ -module g1̄ is semisimple.
To distinguish them from the Cartan series of Theorem 1.11(5), we introduce the
following terminology.

Definition 1.12. The Lie superalgebras g = g0̄ ⊕ g1̄ in Theorem 1.11(2)-(4) are
called classical. Classical Lie superalgebras in Theorem 1.11(2)-(3) are called
basic. The Lie superalgebra gl(m|n) for m, n ≥ 1 is also declared to be basic.

Remark 1.13. Let us comment on the simplicity of the Lie superalgebras in The-
orem 1.11. It is not difficult to check directly the simplicity of the sl series,
spo(2m|2), and those in (4). A Lie superalgebra g in the remaining cases in (2)
and (3), with the exception of spo(2m|2), satisfies the properties that the adjoint
g0̄ -module g1̄ is irreducible and faithful, and [g1̄ , g1̄ ] = g0̄ . The simplicity of such
a g follows easily from these properties. For the Z-graded Cartan type Lie superal-

gebras g = j≥−1 gi in (5), we first note that they are all transitive (i.e., [g−1 , x] = 0
implies that x = 0, for x ∈ g j and j ≥ 0,) and irreducible (i.e., the g0 -module g−1 is
irreducible). Furthermore, we have [g−1 , g1 ] = g0 , [g0 , g1 ] = g1 , and g j is generated
by g1 , for j ≥ 1. From these properties simplicity follows. Finally, simplicity of
e can be proved with a bit of extra work using the explicit realization given in
S(n)
(1.16).

1.2. Structures of classical Lie superalgebras


In this section, Cartan subalgebras, root systems, Weyl groups, and invariant bi-
linear form for basic Lie superalgebras are introduced, and they are described in
detail for type gl, osp and q. A structure theorem is formulated for the basic and
type q Lie superalgebras.

1.2.1. A basic structure theorem. In this subsection, we shall assume that g =


g0̄ ⊕ g1̄ is a basic Lie superalgebra.
A Cartan subalgebra h of g is defined to be a Cartan subalgebra of the even
subalgebra g0̄ . Since every inner automorphism of g0̄ extends to one of Lie super-
algebra g and Cartan subalgebras of g0̄ are conjugate under inner automorphisms,
we conclude that the Cartan subalgebras of g are conjugate under inner automor-
phisms.
Let h be a Cartan subalgebra of g. For α ∈ h∗ , let
gα = {g ∈ g | [h, g] = α(h)g, ∀h ∈ h}.
The root system for g is defined to be
Φ = {α ∈ h∗ | gα ̸= 0, α ̸= 0}.
14 1. Lie superalgebra ABC

Define the even and odd roots to be


Φ0̄ := {α ∈ Φ|gα ∩ g0̄ ̸= 0}, Φ1̄ := {α ∈ Φ|gα ∩ g1̄ ̸= 0}.
Definition 1.14. For a basic or type q Lie superalgebra g = g0̄ ⊕ g1̄ , the Weyl
group W of g is defined to be the Weyl group of the reductive Lie algebra g0̄ .

As we shall see, the Weyl groups play a less vital though still important role
in determining central characters and the linkage principle for Lie superalgebras,
which is somewhat different from the theory of semisimple Lie algebras.
The following structure theorem shows that the basic Lie superalgebras behave
similarly to semisimple Lie algebras.
Theorem 1.15. Let g be a basic Lie superalgebra with a Cartan subalgebra h.
(1) We have a root space decomposition of g with respect to h:

g = h⊕ gα , and g0 = h.
α∈Φ
(2) dim gα = 1, for α ∈ Φ. (Fix now some nonzero eα ∈ gα .)
(3) [gα , gβ ] ⊆ gα+β , for α, β, α + β ∈ Φ.
(4) Φ, Φ0̄ and Φ1̄ are invariant under the action of the Weyl group W on h∗ .
(5) There exists a non-degenerate even invariant supersymmetric bilinear
form (·, ·) on g.
(6) (gα , gβ ) = 0 unless α = −β ∈ Φ.
(7) The restriction of the bilinear form (·, ·) on h × h is non-degenerate and
W -invariant.
(8) [eα , e−α ] = (eα , e−α )hα , where hα is the coroot determined by (hα , h) =
α(h) for h ∈ h.
(9) Φ = −Φ, Φ0̄ = −Φ0̄ , and Φ1̄ = −Φ1̄ .
(10) Let α ∈ Φ. Then, kα ∈ Φ for k ̸= ±1 if and only if α ∈ Φ1̄ and (α, α) ̸= 0;
in this case, k = ±2.

Proof. For g = gl(m|n) or osp(2m|ℓ), we will describe explicitly the root systems
and trace forms below in this section, and so the theorem in these cases follows
by inspection. The theorem for the other infinite series basic Lie superalgebras
sl(m|n) follows by some easy modification from the gl(m|n) case.
Part (4) follows by the invariance, under the action of the Weyl group W , of
the sets of weights for the adjoint g0̄ -modules g0̄ and g1̄ , respectively.
Since the remaining three exceptional Lie superalgebras D(2, 1, α), G(3) and
F(4) will not be studied in detail in the book, we will be sketchy. Most of the
items of the theorem, with an exception of (5), again follow by inspection from
the constructions of these superalgebras and standard arguments as for simple Lie
1.2. Structures of classical Lie superalgebras 15

algebras (with the help of (5)). As the exceptional Lie superalgebras are simple,
any nonzero invariant supersymmetric bilinear form must be non-degenerate. For
G(3) and F(4), the Killing form is such a nonzero even form. A direct and ad hoc
construction of a nonzero invariant bilinear form on D(2, 1, α) is possible. 

It follows by Theorem 1.15(1) that h is self-normalizing in g (and h is abelian),


justifying the terminology of a Cartan subalgebra. Since h ⊆ g0̄ and dim gα = 1 for
each α ∈ Φ by Theorem 1.15(2), there exists i ∈ Z2 such that gα ⊆ gi . Hence Φ is
a disjoint union of Φ0̄ and Φ1̄ , and we have

Φi = {α ∈ Φ | gα ⊆ gi }, i ∈ Z2 .

Remark 1.16. One uniform approach to establish Theorem 1.15 is as follows (see
[K1, Proposition 2.5.3]). One follows the standard construction of contragredi-
ent (i.e., Kac-Moody) (super)algebras (see Carter [Car, Chapter 14]) to show that
any Lie superalgebra in Theorem 1.15 is a Kac-Moody superalgebra (or rather
the quotient by its possibly nontrivial center) associated to some generalized Car-
tan matrix with Z2 -grading. These (quotients of) Kac-Moody superalgebras carry
non-degenerate even invariant supersymmetric bilinear forms, by a standard Kac-
Moody type argument (see Carter [Car, Chapter 16]).

A root α ∈ Φ is called isotropic if (α, α) = 0. An isotropic root is necessarily


an odd root. Denote the set of isotropic odd roots by
Φ̄1̄ := {α ∈ Φ1̄ | (α, α) = 0}
(1.17)
= {α ∈ Φ1̄ | 2α ̸∈ Φ}.
The second equation above follows by Theorem 1.15(10). Moreover, we have
1
e2α = [eα , eα ] = 0, for α ∈ Φ̄1̄ .
2

1.2.2. Invariant bilinear forms for gl and osp. In contrast to the semisimple Lie
algebras, the Killing form for a basic Lie superalgebra may be zero, and even when
it is nonzero, it may not be positive definite on the real vector space spanned by
Φ. In this subsection, we give a down-to-earth description of an invariant non-
degenerate even supersymmetric bilinear form for Lie superalgebras of type gl and
osp.
The supertrace str on the general linear Lie superalgebra gives rise to a non-
degenerate supersymmetric bilinear form

(·, ·) : gl(m|n) × gl(m|n) → C, (a, b) = str(ab).


It is straightforward to check that this form is invariant. Restricting to the Cartan
subalgebra h of diagonal matrices, we obtain a non-degenerate symmetric bilinear
16 1. Lie superalgebra ABC

form on h:

 1 if 1 ≤ i = j ≤ m,
(Eii , E j j ) = −1 if 1 ≤ i = j ≤ n,

0 if i ̸= j,
where i, j ∈ I(m|n). We recall here that I(m|n) is defined in (1.2). Denote by
{δi , ε j }i, j the basis of h∗ dual to {Ei i , E j j }i, j , where 1 ≤ i ≤ m and 1 ≤ j ≤ n.
Using the bilinear form (·, ·) we can identify δi with (Ei,i , ·) and ε j with −(E j j , ·).
Whenever it is convenient we also use the notation
(1.18) εi := δi , for 1 ≤ i ≤ m.

The form (·, ·) on h induces a non-degenerate bilinear from on h∗ , which will


be denoted by (·, ·) as well. Then, for i, j ∈ I(m|n), we have

 1 if 1 ≤ i = j ≤ m,
(1.19) (εi , ε j ) = −1 if 1 ≤ i = j ≤ n,

0 if i ̸= j.

Such a bilinear form on gl(2n|ℓ) restricts to a non-degenerate invariant super-


symmetric bilinear form on the subalgebra spo(2n|ℓ), which will also denoted by
(·, ·). The further restriction to a Cartan subalgebra of spo(2n|ℓ) remains to be non-
degenerate. This allows us to identify a Cartan subalgebra h with its dual h∗ , and
also induces a bilinear form on the dual h∗ .

1.2.3. Root system and Weyl group for gl(m|n). Let g = gl(m|n) and h be the
Cartan subalgebra of diagonal matrices. Its root system Φ = Φ0̄ ∪ Φ1̄ is given by
specifying
Φ0̄ = {εi − ε j | i ̸= j ∈ I(m|n), i, j > 0 or i, j < 0},
Φ1̄ = {±(εi − ε j ) | i, j ∈ I(m|n), i < 0 < j}.
Observe that Ei j is a root vector corresponding to the root εi −ε j , for i ̸= j ∈ I(m|n).
The Weyl group of gl(m|n), which is by definition the Weyl group of g0̄ =
gl(m) ⊕ gl(n), is isomorphic to Sm × Sn , where we recall that Sn denotes the
symmetric group of n letters.

1.2.4. Root system and Weyl group for spo(2m|2n + 1). Now we describe the
root system for the ortho-symplectic Lie superalgebra spo(2m|2n + 1), which is
defined in matrix form (1.8) in Section 1.1.3. Recall that the rows and columns
of the matrices are indexed by I(2m|2n + 1). The subalgebra h of g of diagonal
matrices has a basis given by
Hi := Ei,i − Em+i,m+i , 1 ≤ i ≤ m,
H j := E j j − En+ j,n+ j , 1 ≤ j ≤ n,
1.2. Structures of classical Lie superalgebras 17

and it is a Cartan subalgebra for spo(2m|2n + 1). With respect to h, the root system
Φ = Φ0̄ ∪ Φ1̄ for spo(2m|2n + 1) is
{±δi ± δ j , ±2δ p , ±εk ± εl , ±εq } ∪ {±δ p ± εq , ±δ p },
where 1 ≤ i < j ≤ m, 1 ≤ k < l ≤ n, 1 ≤ p ≤ m, 1 ≤ q ≤ n.
A root vector is a nonzero vector in gα for α ∈ Φ, and will be denoted by
eα . The root vectors for spo(2m|2n + 1) can be computed explicitly as follows
(1 ≤ i ̸= j ≤ m, 1 ≤ k ̸= l ≤ n):
eεk = E2n+1,k+n − Ek,2n+1 , e−εk = E2n+1,k − Ek+n,2n+1 ,
(1.20) e2δi = Ei,i+m , e−2δi = Ei+m,i ,
(1.21) eδi +δ j = Ei, j+m + E j,i+m , e−δi −δ j = E j+m,i + Ei+m, j ,
(1.22) eδi −δ j = Ei, j − E j+m,i+m , eεk −εl = Ekl − El+n,k+n ,
(1.23) eεk +εl = Ek,l+n − El,k+n , e−εk −εl = Ek+n,l − El+n,k ,

(1.24) eδi +εk = Ek,i+m + Ei,k+n , e−δi −εk = Ek+n,i − Ei+m,k ,


(1.25) eδi −εk = Ek+n,i+m + Ei,k , e−δi +εk = Ek,i − Ei+m,k+n ,
eδi = E2n+1,i+m + Ei,2n+1 , e−δi = E2n+1,i − Ei+m,2n+1 .

The Weyl group of spo(2m|2n + 1), which is by definition the Weyl group of
g0̄ = sp(2m) ⊕ so(2n + 1), is isomorphic to (Zm
2 o Sm ) × (Z2 o Sn ).
n

1.2.5. Root system and Weyl group for spo(2m|2n). Let g = spo(2m|2n). The
abelian Lie subalgebra h spanned by {Hi , H j |1 ≤ i ≤ m, 1 ≤ j ≤ n} is a Cartan
subalgebra for spo(2m|2n). Let {δi , ε j | 1 ≤ i ≤ m, 1 ≤ j ≤ n} be the corresponding
dual basis. Again, when it is convenient, we will use the notation εī = δi , for
1 ≤ i ≤ m. With respect to h, the root system Φ = Φ0̄ ∪ Φ1̄ is given by
{±δi ± δ j , ±2δ p , ±εk ± εl } ∪ {±δ p ± εq },
where 1 ≤ i < j ≤ m, 1 ≤ k < l ≤ n, 1 ≤ p ≤ m, 1 ≤ q ≤ n.
The root vectors for spo(2m|2n) are given by (1.20)–(1.25).
The Weyl group of spo(2m|2n), which is by definition the Weyl group of g0̄ =
sp(2m) ⊕ so(2n), is an index 2 subgroup of (Zm2 o Sm ) × (Z2 o Sn ), with only
n

even number of signs in Zn2 permitted.

1.2.6. Root system and odd invariant form for q(n). In this subsection let g be
the queer Lie superalgebra q(n), see Section 1.1.4. As the case of q(n) is different
from the basic Lie superalgebra case, we will describe altogether its Cartan sub-
algebras, root systems, positive systems, Borel subalgebras, and invariant bilinear
form.
18 1. Lie superalgebra ABC

(
)
a b
Recall that q(n) can be realized as matrices in the n|n block form
b a
indexed by I(n|n). The subalgebra consisting of matrices with a, b being diago-
nal (which we refer to as “block diagonal matrices”) will be called the standard
Cartan subalgebra. We define a Cartan subalgebra h to be any subalgebra con-
jugate by some element in GL(n) to the standard Cartan subalgebra. Note that h
is self-normalizing in g and h is nilpotent, justifying the terminology of a Cartain
subalgebra. However, h = h0̄ ⊕h1̄ is not abelian, since [h0̄ , h] = 0, and [h1̄ , h1̄ ] = h0̄ .
Now fix h = h0̄ ⊕ h1̄ to be the standard Cartan subalgebra. The vectors
Hi := Eī ī + Eii , i = 1, . . . , n,
form a basis for h0̄ , while the vectors
H i := Eī i + Ei ī , i = 1, . . . , n,
form a basis for h1̄ . We let {εi | i = 1, . . . , n} denote the corresponding dual basis

in h∗0̄ . With respect to h0̄ , we have the root space decomposition g = h ⊕ α∈Φ gα
with root system Φ = Φ0̄ ∪Φ1̄ , where both Φ0̄ and Φ1̄ coincide with the root system
for gl(n):
Φ0̄ = Φ1̄ = {εi − ε j | 1 ≤ i ̸= j ≤ n}.
We have dimC gα = 1, for each α ∈ Φ, and gα ⊆ gi , for α ∈ Φi and i ∈ Z2 . The
Weyl group of q(n) is identified with the symmetric group Sn .
( )
a b
The matrices with a, b being upper triangular (which we refer to as
b a
“block upper triangular matrices”) form a solvable subalgebra b, which will be
called the standard Borel subalgebra of g. The positive system corresponding to
the Borel subalgebra b is Φ+ = Φ+ 0̄
∪ Φ+

, where Φ+

= Φ+

= {εi − ε j | 1 ≤ i < j ≤
n}. Let Φ− = −Φ+ and so Φ = Φ+ ∪ Φ− . Let
⊕ ⊕
n+ = gα , n− = gα .
α∈Φ+ α∈Φ−

Then, we have b = h ⊕ n+ , and we have a triangular decomposition


g = n− ⊕ h ⊕ n+ .
To be compatible with the definition of Weyl vector ρ in (1.28) for basic Lie super-
algebras later on, it makes sense to regard ρ = 0 for q(n).
Any subalgebra of g that is conjugate to b by GL(n) will be referred to as a
Borel subalgebra of g.
( )
a b
For g = in q(n), the odd trace is defined to be
b a
(1.26) otr(g) := tr(b).
1.3. Non-conjugate positive systems and odd reflections 19

Using this, we obtain an odd non-degenerate invariant symmetric bilinear form


(·, ·) on q(n) defined by
(g, g′ ) = otr(gg′ ), g, g′ ∈ q(n).
Here “odd” is understood in the sense of Definition 1.7.

1.3. Non-conjugate positive systems and odd reflections


In this section, positive systems, fundamental systems, and Dynkin diagrams for
basic Lie superalgebras are defined and classified, along with Borel subalgebras.
In contrast to semisimple Lie algebras, the fundamental systems for a Lie super-
algebra may not be conjugate under the Weyl group action. We introduce odd
reflections which permute the fundamental systems.

1.3.1. Positive systems and fundamental systems. Let Φ be a root system for a
basic Lie superalgebra and let E be the real vector space spanned by Φ. We have
E ⊗R C = h∗ , for g ̸= gl(m|n). For g = gl(m|n) the space E ⊗R C is a subspace of
h∗ of codimension one.
A total ordering ≥ on E below is always assumed to be compatible with the
real vector space structure, that is, v ≥ w and v′ ≥ w′ imply that v + v′ ≥ w + w′ ,
−w ≥ −v, and cv ≥ cw for c ∈ R>0 .
A positive system Φ+ is a subset of Φ consisting precisely of all those roots
α ∈ Φ satisfying α > 0 for some total ordering of E. Given a positive system
Φ+ , we define the fundamental system Π ⊂ Φ+ to be the set of α ∈ Φ+ which
cannot be written as a sum of two roots in Φ+ . We refer to elements in Φ+ as
positive roots and elements in Π as simple roots. Similarly, we denote by Φ− the
− −
corresponding set of negative roots. Set Φ+i = Φ ∩ Φi and Φi = Φ ∩ Φi , for
+

i ∈ Z2 . By Theorem 1.15(9), we have Φ− = −Φ+ and Φi = −Φi , for i ∈ Z2 .
+

Then, we have
Φ+ = Φ+0̄
∪ Φ+

.
Recall Φ̄1̄ from (1.17). Associated to a positive system Φ+ , we let
(1.27) Φ̄+

:= Φ̄1̄ ∩ Φ+ .
Proposition 1.17. Let g be a basic Lie superalgebra. There is a one-to-one corre-
spondence between the set of positive systems and the set of fundamental systems.
The Weyl group of g acts naturally on the set of the positive systems (respectively,
fundamental systems).

Proof. It follows by definition that the fundamental system exists and is unique for
a given positive system. A positive root, if not simple, can be written as a sum of
two positive roots. Continuing this way, any positive root is a Z+ -linear combina-
tion of simple roots, and hence a positive system is determined by its fundamental
system.
20 1. Lie superalgebra ABC

By Theorem 1.15, Φ = −Φ and Φ is W -invariant. Then W acts naturally on the


set of positive systems, and then on the set of fundamental systems by the above
correspondence. 

We define
⊕ ⊕
n+ = gα , n− = gα .
α∈Φ+ α∈Φ−

Then n± are ad h-invariant nilpotent subalgebras of g and we obtain a triangular


decomposition
g = n− ⊕ h ⊕ n+ .
The solvable subalgebra b = h ⊕ n+ is called a Borel subalgebra of g (correspond-
ing to Φ+ ). We have b = b0̄ ⊕ b1̄ , where bi = b ∩ gi for i ∈ Z2 .

Remark 1.18. The rank one subalgebra corresponding to an isotropic simple root
is isomorphic to sl(1|1), which is solvable. Therefore, if we enlarge a Borel sub-
algebra by adding the root space corresponding to a negative isotropic simple root,
then the resulting subalgebra is still solvable. Thus, a Borel subalgebra is not a
maximal solvable subalgebra for Lie superalgebras in general.

Given a positive system Φ+ = Φ+



∪ Φ+

, the Weyl vector ρ is defined by

(1.28) ρ = ρ0̄ − ρ1̄ ,

where
1 1
ρ0̄ = ∑+ α,
2 α∈Φ
ρ1̄ = ∑+ β.
2 β∈Φ
0̄ 1̄

Denote

(1.29) 1m|n = (δ1 + . . . + δm ) − (ε1 + . . . + εn ).

The following lemma is proved by a direct computation.

Lemma 1.19. We have the following formulas for the Weyl vector ρ for the stan-
dard positive system Φ+ :

i=1 (m − i + 1)δi − ∑ j=1 jε j − 2 (m + n + 1)1m|n , for g = gl(m|n).


(1) ρ = ∑m n 1

i=1 (m − n − i + 2 )δi − ∑ j=1 (n − j + 2 )ε j , for g = spo(2m|2n + 1).


(2) ρ = ∑m 1 n 1

i=1 (m − n − i + 1)δi − ∑ j=1 (n − j)ε j , for g = spo(2m|2n).


(3) ρ = ∑m n

We shall next describe completely the positive and fundamental systems for
Lie superalgebras of type gl and osp case-by-case.
1.3. Non-conjugate positive systems and odd reflections 21

1.3.2. Positive and fundamental systems for gl(m|n). Recall the root system Φ
for gl(m|n) and the standard Cartan subalgebra h described in Section 1.2.3. The
subalgebra of upper triangular matrices is the standard Borel subalgebra of g
which contains h, and the corresponding standard positive system of Φ is given
by {εi − ε j | i, j ∈ I(m|n), i < j}. Bearing in mind εi = δi , its standard fundamental
system is
{δi − δi+1 , ε j − ε j+1 , δm − ε1 | 1 ≤ i ≤ m − 1, 1 ≤ j ≤ n − 1},
with the corresponding simple root vectors ei := Ei,i+1 , for i ∈ I(m − 1|n − 1), and
em := Em,1 . The simple coroots are h j := E j j − E j+1, j+1 , for j ∈ I(m − 1|n − 1),
and hm := Em,m + E11 . Denote fi := Ei+1,i , for i ∈ I(m − 1|n − 1), and fm := E1,m .
Then {ei , hi , fi | i ∈ I(m|n − 1)} is a set of Chevalley generators for sl(m|n).
Note that
(δi − δi+1 , δi − δi+1 ) = 2, 1 ≤ i ≤ m − 1,
(δm − ε1 , δm − ε1 ) = 0,
(ε j − ε j+1 , ε j − ε j+1 ) = −2, 1 ≤ j ≤ n − 1.
Thus δm − ε1 is an isotropic simple root. Following the usual convention for Lie
algebras, we draw the corresponding standard Dynkin diagram:

⃝ ⃝ ··· ⃝ ··· ⃝ ⃝
δ1 − δ2 δ2 − δ3 δm − ε1 ε1 − ε2 εn−2 − εn−1 εn−1 − εn

where we denote by ⃝ an even simple root α such that 12 α is not a root, and denote

by an odd isotropic simple root.
Let us classify all possible positive systems for gl(m|n), keeping in mind εī =
δi . If we ignore the parity of roots for a moment, the root system of gl(m|n) is the
same as the root system for gl(m+n), and hence by definition their positive systems
(respectively, fundamental systems) are exactly described in the same way, and
there are (m + n)! of them in total. It follows from the well-known classification
for gl(m + n) that a fundamental system for gl(m|n) consists of (m + n − 1) roots
εi1 − εi2 , εi2 − εi3 , . . . , εim+n−1 − εim+n , where {i1 , i2 , . . . , im+n } = I(m|n). Then we
restore the parity of the simple roots in a fundamental system for gl(m|n). The
corresponding Dynkin diagram is of the form
⊙ ⊙ ⊙ ⊙ ⊙ ⊙
··· ···
εi1 − εi2 εi2 − εi3 εik − εik+1 εim+n−1 − εim+n

⊙ ⊗
where is ⃝ or , depending on the corresponding simple root is even or odd.
Example 1.20. For n = m, there exists a fundamental system which consists of
only odd roots {δ1 −ε1 , ε1 −δ2 , δ2 −ε2 , . . . , δm −εm }, whose corresponding Dynkin
diagram is
22 1. Lie superalgebra ABC

⊗ ⊗ ⊗ ⊗ ⊗ ⊗
··· ···
δ1 − ε1 ε1 − δ2 δk − εk εk − δk+1 εm−1 − δm δm − εm

The εδ-sequence for a fundamental system Π is obtained by switching the


ordered sequence εi1 εi2 . . . εim+n for Π to the εδ-notation via the identification εī = δi
and then dropping the indices. Clearly, an εδ-sequence has m δ’s and n ε’s. In
general, there exist positive systems for Φ which are not conjugate to each other
under the action of the Weyl group, in contrast to the semisimple Lie algebra case.
Example 1.21. The three W -conjugacy classes of fundamental systems for gl(1|2)
correspond to the three sequences δεε, εδε, εεδ, respectively. Also the standard
Borel of gl(m|n) corresponds to the sequence δ| ·{z
· · δ} ε| ·{z
· · }ε while the opposite Borel
m n
to the standard one corresponds to ε| ·{z
· · }ε |δ ·{z
· · δ}.
n m

1.3.3. Positive and fundamental systems for spo(2m|2n + 1). Now we describe
the positive/fundamental systems and Dynkin diagrams for spo(2m|2n + 1) whose
root system is described in Section 1.2.4. The standard positive system Φ+ =
Φ+0̄
∪ Φ+

corresponding to the standard Borel subalgebra for spo(2m|2n + 1) is
{δi ± δ j , 2δ p , εk ± εl , εq } ∪ {δ p ± εq , δ p },
where 1 ≤ i < j ≤ m, 1 ≤ k < l ≤ n, 1 ≤ p ≤ m, 1 ≤ q ≤ n. The fundamental system
Π of Φ+ contains one odd simple root δm − ε1 , and it is given by
Π = {δi − δi+1 , δm − ε1 , εk − εk+1 , εn | 1 ≤ i ≤ m − 1, 1 ≤ k ≤ n − 1}.
The corresponding standard Dynkin diagram for spo(2m|2n + 1) is

⃝ ⃝ ··· ⃝ ··· ⃝=⇒⃝
δ1 − δ2 δ2 − δ3 δm − ε1 ε1 − ε2 εn−1 − εn εn

Another often used positive system in Φ for spo(2m|2n + 1) is given by


{δi ± δ j , 2δ p , εk ± εl , εq } ∪ {εq ± δ p , δ p },
where 1 ≤ i < j ≤ m, 1 ≤ k < l ≤ n, 1 ≤ p ≤ m, 1 ≤ q ≤ n, with the following
fundamental system
{εk − εk+1 , εn − δ1 , δi − δi+1 , δm |1 ≤ i ≤ m − 1, 1 ≤ k ≤ n − 1}.
The corresponding Dynkin diagram is

⃝ ⃝ ··· ⃝ ··· ⃝=⇒a
ε1 − ε2 ε2 − ε3 εn − δ1 δ1 − δ2 δm−1 − δm δm
1.3. Non-conjugate positive systems and odd reflections 23

where we use a to denote a non-isotropic odd simple root, as (δm , δm ) = 1.


Now let us classify all the possible fundamental systems for spo(2m|2n + 1),
keeping in mind εī = δi . Note that 2ε p ∈ Φ+ if and only if ε p ∈ Φ+ , and that ±2ε p
are never in any fundamental system by definition. Hence, for the sake of classi-
fication of positive systems and fundamental systems in Φ, it suffices to consider
the subset Φ̃ := Φ − {±2ε p | 1 ≤ p ≤ m}. Ignoring the parity of the roots, Φ̃ may
be identified with the root system of the classical Lie algebra so(2m + 2n + 1),
whose fundamental systems are completely known and they are acted upon sim-
ply transitively by the Weyl group of so(2m + 2n + 1) (which is ∼ = Zm+n
2 o Sm+n ).
The number of W -congugacy classes( ) of fundamental systems for spo(2m|2n + 1)
is |W (so(2m + 2n + 1))|/|W | = m+n m .
The εδ-sequence associated to a fundamental system (or a Dynkin diagram)
for spo(2m|2n + 1) is defined as for gl(m|n), starting from the type A end of the
Dynkin diagram (to fix the ambiguity). For example, the εδ-sequence associated
to the standard Dynkin diagram is m δ’s followed by n ε’s.

Example 1.22. Let g = osp(1|2). Then its even subalgebra g0̄ is isomorphic to
sl(2) = C⟨e, h, f ⟩, and as a g0̄ -module g1̄ is isomorphic to the 2-dimensional natural
sl(2)-module CE + CF. The simple root consists of a (unique) odd non-isotropic
root δ so that 2δ is an even root. The Dynkin diagram is a . The root vectors
E and F associated to the odd roots δ and −δ can be chosen such that [E, E] =
2e, [F, F] = −2 f , [E, F] = h.

1.3.4. Positive and fundamental systems for spo(2m|2n). Now we consider g =


spo(2m|2n), whose root system is described in Section 1.2.5. The standard pos-
itive system Φ+ = Φ+ 0̄
∪ Φ+

in Φ corresponding to the standard Borel subalgebra
is
{δi ± δ j , 2δ p , εk ± εl } ∪ {δ p ± εq },
where 1 ≤ i < j ≤ m, 1 ≤ k < l ≤ n, 1 ≤ p ≤ m, 1 ≤ q ≤ n, with its fundamental
system being
Π = {δi − δi+1 , δm − ε1 , εk − εk+1 , εn−1 + εn | 1 ≤ i ≤ m − 1, 1 ≤ k ≤ n − 1}.
The corresponding standard Dynkin diagram is
⃝ εn−1 − εn

⃝ ⃝ ··· ⃝ ··· ⃝
δ1 − δ2 δ2 − δ3 δm − ε1 ε1 − ε2 εn−2 − εn−1@
@⃝ εn−1 + εn

Another often used positive system in Φ is given by


{δi ± δ j , 2δ p , εk ± εl } ∪ {εq ± δ p },
24 1. Lie superalgebra ABC

where 1 ≤ i < j ≤ m, 1 ≤ k < l ≤ n, 1 ≤ p ≤ m, 1 ≤ q ≤ n, with its fundamental


system being
{εk − εk+1 , εn − δ1 , δi − δi+1 , 2δm | 1 ≤ i ≤ m − 1, 1 ≤ k ≤ n − 1}.
The corresponding Dynkin diagram is

⃝ ⃝ ··· ⃝ ··· ⃝⇐=⃝
ε1 − ε2 ε2 − ε3 εn − δ1 δ1 − δ2 δm−1 − δm 2δm

As we already observed above, there are (at least) two Dynkin diagrams for
spo(2m|2n) of different shapes. The classification of all fundamental systems for
the root system Φ of spo(2m|2n) is divided into 2 cases below. As usual we keep
in mind εī = δi .
(1) First we classify the fundamental systems Π in Φ that do not contain
any long root (i.e., a root of the form ±2ε p ). With parity ignored, the subset
Φ̃ := Φ − {±2ε p | 1 ≤ p ≤ m} may be identified with the root system of the clas-
sical Lie algebra so(2m + 2n), whose fundamental systems are completely known
and they are acted upon simply transitively by the Weyl group of so(2m + 2n)
(which is an index 2 subgroup of Zm+n 2 o Sm+n ). Observe that the positive sys-
tem Φ+ for Φ corresponding to Π is completely determined by the positive system
Φ̃ ∩ Φ+ for Φ̃ (and vice versa). We conclude that the fundamental systems for
Φ that do not contain any long root are exactly the fundamental systems for Φ̃.
However, not every fundamental system of Φ̃ gives rise to a fundamental system
of Φ. To be precise, a fundamental system of Φ cannot contain a pair of roots
of the form {±εi ± ε p , ±εi ∓ ε p }, i ̸= p and 1 ≤ p ≤ m. It follows that there are
m+n · |W (so(2m + 2n))| such fundamental systems of Φ, and hence the number
n
of
(m+n−1)
W -conjugacy classes for g in this case is m+n · |W (so(2m + 2n))|/|W | =
n
m .
(2) Next, we classify the fundamental systems Π that contain some long root.
In this case, we consider Φ b := Φ ∪ {±2ε j , 1 ≤ j ≤ n}. With the parity ignored,
b may be identified with the root system of sp(2m + 2n), whose fundamental sys-
Φ
tems are completely described. Observe that the positive system Φ+ for Φ which
corresponds to Π is completely determined by the positive system Φ b + of Φb which
contains Φ . We conclude that the fundamental systems for Φ that contain some
+

long root are exactly the fundamental systems for Φ b whose long root is of the form
±2ε p . It follows that the number of W -conjugacy classes of such fundamental
( )
m
systems for g is m+n · |W (sp(2m + 2n))|/|W | = 2 m+n−1
n .
The εδ-sequence for spo(2m|2n) associated to a positive system is now de-
fined as follows. We can first obtain an ordered sequence of ε’s and δ’s just as
for spo(2m|2n + 1). If this sequence has an ε as its last member, then it is the
εδ-sequence. If the last member in this sequence is a δ, then the εδ-sequence is
obtained from this sequence by attaching a sign to the last ε. For example, for
spo(4|4) the εδ-sequences εδδε, εδεδ, and εδ(−ε)δ are distinct.
1.3. Non-conjugate positive systems and odd reflections 25

1.3.5. Conjugacy classes of fundamental systems. We now describe the classi-


fication of the W -conjugacy classes of fundamental systems for these Lie superal-
gebras via the εδ-sequences.
Proposition 1.23. Let Φ be the root system and W be the Weyl group for a Lie
superalgebra g of type gl or osp. Then the W -conjugacy classes of fundamental
systems in Φ are in one-to-one correspondence with the associated εδ-sequences.
( )
More precisely, there are m+n conjugacy classes of Borel subalgebras for
m ( ) (m+n−1)
gl(m|n) and spo(2m|2n + 1), while there are m+n
m + n conjugacy classes of
Borel subalgebras for spo(2m|2n).

Proof. By the case-by-case classification of fundamental systems in Φ and defini-


tion of εδ-sequences, we clearly have a well-defined map
Θ : {fundamental systems in Φ}/W −→ {εδ-sequences for g}.
As we can easily construct a fundamental system Π for a given εδ-sequence, Θ is
surjective. To show Θ is a bijection, it remains to show that the two finite sets have
the same cardinalities.
The number of W -conjugacy classes of fundamental systems equals |W ′ |/|W |,
where W ′ denotes the Weyl group of gl(m + n) when g = gl(m|n) and W ′ denotes
the Weyl group
( of ) so(2m + 2n + 1) when g = spo(2m|2n + 1). In either case,
|W ′ |/|W | = m+n
m . On the other hand, in either case, an εδ-sequence (
is simply an
)
ordered arrangement of m δ’s and n ε’s. Thus the total number is also m+n
m .
In the case of spo(2m|2n), when the last slot of an εδ-sequence is an ε, the
(m+n−1) (m + n − 1) slots can be filled with m δ’s and (n − 1) ε’s. Thus we obtain
previous
m different εδ-sequences this way. When the last slot in an εδ-sequence is
a δ, then the previous (m + n − 1) slots are filled with (m − 1) δ’s and n ε’s, with
the last ε) having either a positive or negative
(m+n−1 (m+n−1sign.
) (This way
) we obtain additional
2 n distinct εδ-sequences. Now m +2 n m+n−1
is exactly the number
of W -conjugacy classes of fundamental systems for spo(2m|2n) that we have com-
puted earlier. 

The fundamental and positive systems for the exceptional Lie superalgebras
G(3), F(4), D(2, 1, α) are also listed completely by Kac [K1] (with one missing
for F(4); see [WZ, 5.1]).
¿From the classification of fundamental and positive systems we can read off
the following.
Proposition 1.24. Let g be a basic Lie superalgebra, excluding sl(n|n)/CIn|n and
gl(m|n). Let h be a Cartan subalgebra of g. Then any fundamental system forms
a basis in h∗ . For g = gl(m|n), any fundamental system is linearly independent in
h∗ .
26 1. Lie superalgebra ABC

Lemma 1.25. Let g be a basic Lie superalgebra. Let Π be the fundamental system
in the positive system Φ+ , and let α be an isotropic odd root in Φ+ . Then there
exists w ∈ W such that w(α) ∈ Π.

1.3.6. Odd reflections. We have seen that the fundamental systems of a root sys-
tem Φ are not W -conjugate due to the existence of odd (simple) roots. The follow-
ing fundamental lemma will play an important role in representation theory of Lie
superalgebras.

Lemma 1.26. Let g be a basic Lie superalgebra. Let α be an isotropic odd simple
root in a positive system Φ+ . Then,
Φ+
α := {−α} ∪ Φ \{α}
+

is a new positive system, whose corresponding fundamental system Πα is given by


(1.30) Πα = {β ∈ Π | (β, α) = 0, β ̸= α} ∪ {β + α | β ∈ Π, (β, α) ̸= 0} ∪ {−α}.

Let b = h + n+ , where n+ corresponds to the positive system Φ+ . Then the


new Borel subalgebra corresponding to Φ+
α is given by

bα := h ⊕ gβ .
β∈Φ+
α

The process of obtaining Πα (respectively, Φ+ α


α or b ) from Π (respectively, Φ or
+

b) will be referred to as an odd reflection (with respect to α), and will be denoted
by rα . We shall write
rα (Π) = Πα , rα (Φ+ ) = Φ+
α, rα (b) = bα .
Note that r−α rα = 1.

Proof. For g = gl(m|n), spo(2m|2n) or spo(2m|2n + 1), which we will be mostly


concerned about in this book, we have already obtained complete descriptions of all
fundamental systems. For most Π’s the set {β ∈ Π | (β, α) ̸= 0} has cardinality at
most 2 and consists of roots of the form ±εi ± ε j for i ̸= j ∈ I(m|n). By inspection
we see that Πα is a fundamental system. There are a few extra cases when the
root α corresponds to a node in the corresponding Dynkin diagram which is either
connected to a long simple root, or connected to a short non-isotropic root, or a
branching node, or belongs to one of the (short) end nodes of a branching node, as
follows.
⊙ ⊙
⊗ ⊗ ⊕ ⊗ ⊙
··· ⇐=⃝ ··· =⇒ ··· ···
α α α @ ⊙ @ ⊗
@ @ α
1.3. Non-conjugate positive systems and odd reflections 27


Here the vertical dashed lines indicate an edge when it connects two ’s and no
⊕ ⊙ ⊗
edge otherwise, means either ⃝ or a , and means either or ⃝. It can be
checked directly in these cases that Πα is also a fundamental system.
Take β ∈ Φ+ α ∩ Φ . Since Π is a fundamental system for Φ , by Proposi-
+ +

tion 1.24 we have a unique expression β = ∑γ∈Π\{α} mγ γ + mα α for mγ ∈ Z+ and


mα ∈ Z+ . It follows by definition of Πα that β can be expressed as a linear combi-
nation of Πα of the form β = ∑γ∈Πα \{−α} mγ γ + m′α (−α) for some suitable integer
m′α . By choice of β we have mγ > 0 for some γ, and hence m′α ∈ Z+ , since Πα is
a fundamental system. This shows that Φ+ α is the positive system corresponding to
Πα .
For an exceptional Lie superalgebra g = D(2, 1, α), G(3) or F(4), which will
not be studied in any detail in this book, our proof shall be rather sketchy. A
conceptual approach is to follow Remark 1.16 to regard g as a Kac-Moody super-
algebra with Chevalley generators ei , fi , hi for i ∈ I associated to a Cartan matrix A
(and a fundamental system Π = {αi | i ∈ I}), where I is Z2 -graded. For an isotropic
odd root α ∈ Π, one can construct a new set of Chevalley generators e′i , fi′ , h′i asso-
ciated to Πα , which gives rise to a new Cartan matrix A′ . By standard machinery of
Kac-Moody theory, the Kac-Moody superalgebra associated to A and A′ coincide
with g. From this it follows that Πα is a fundamental system for g. The same argu-
ment as above shows that Φα is the positive system associated to the fundamental
system Πα . This uniform approach is applicable to all basic Lie superalgebras. 
Corollary 1.27. For two fundamental systems Π and ′ Π of a basic Lie superal-
gebra g, there exists a sequence consisting of real and odd reflections r1 , r2 , . . . , rk
such that rk . . . r2 r1 (Π) = ′ Π.

Proof. Denote by Φ+ and ′ Φ+ the positive systems associated to Π and ′ Π respec-


tively. We prove the corollary by induction on |Φ+ ∩ ′ Φ− |. If |Φ+ ∩ ′ Φ− | = 0,
then Π = ′ Π. Assume now |Φ+ ∩ ′ Φ− | > 0. Then Π+ ̸= ′ Π+ , and we pick
α ∈ Π+ ∩ ′ Φ− ̸= 0,/ and apply the real or odd simple reflection rα . Observe that
Φα = {−α} ∪ Φ \{α} is the positive system with fundamental system Πα =
+ +

rα (Π), regardless of the parity of the simple root α. Note that


′ − + ′ −
|Φ+
α ∩ Φ | < |Φ ∩ Φ |.
By the inductive assumption, there exists a sequence of real and odd reflections
r2 , . . . , rk such that rk . . . r2 (Πα ) = ′ Π. Hence rk . . . r2 rα (Π) = ′ Π. 
Proposition 1.28. Let g be a basic Lie superalgebra. Let Φ+ be a positive system
with Π as its fundamental system and ρ as its associated Weyl vector. Then, (ρ, β) =
2 (β, β) for every simple root β ∈ Π.
1

Proof. Let α ∈ Π. Let Πα and Φ+ α be the fundamental and positive systems


obtained by a reflection rα (see Lemma 1.26 for α isotropic odd; for α non-
isotropic, rα = r2α ), respectively. Let ρα denote the Weyl vector for Φ+
α . Since
28 1. Lie superalgebra ABC

Φ+
α = {−α} ∪ Φ \{α}, we compute that
+
{
ρ − α, for α even or non-isotropic odd,
ρα =
ρ + α, for α isotropic odd.
Claim. Assume that the proposition holds for a fundamental system Π, then it
holds for the fundamental system Πα for α ∈ Π.
Let us prove the claim. First assume α ∈ Π is even or non-isotropic odd. Since
ρα = rα (ρ) and Πα = rα (Π), we have for each β ∈ Π that
1 1
(ρα , rα (β)) = (ρ, β) = (β, β) = (rα (β), rα (β)).
2 2
Now assume α ∈ Π is isotropic odd. We will check case-by-case, recalling the def-
inition of Πα from (1.30). If (β, α) = 0 for β ∈ Π, then β ∈ Πα . By the assumption
of the claim, we have
1
(ρα , β) = (ρ + α, β) = (ρ, β) = (β, β).
2
If (β, α) ̸= 0 for β ∈ Π, then β + α ∈ Πα . By the assumption of the claim, (ρ, α) =
2 (α, α) = 0, and hence
1

1 1
(ρα , β + α) = (ρ + α, β + α) = (ρ + α, β) = (β, β) + (α, β) = (β + α, β + α).
2 2
This completes the proof of the claim.
By the claim and Corollary 1.27, it suffices to prove the proposition for just
one fundamental system, e.g., the standard fundamental system for type gl and
osp. Assume that β is an even simple root or an non-isotropic odd simple root. In
either case one checks that rβ (ρ) = ρ − β. Since (·, ·) is W -invariant, we have
( )
(ρ, β) = rβ (ρ), rβ (β) = (ρ − β, −β) = −(ρ, β) + (β, β).
It follows that (ρ, β) = 12 (β, β) just as for semisimple Lie algebras.
For isotropic odd simple root β, we do a case-by-case inspection. For type gl
or osp, the equation (ρ, β) = 12 (β, β) follows directly from Lemma 1.19. For the
three exceptional cases, it can be checked directly by using a fundamental system
with exactly one isotropic simple root and the detailed description of root systems
(cf. Kac [K1]). We skip the details. 
Remark 1.29. In contrast to a real reflection, an odd reflection rα with respect to
an isotropic odd root α may not be extended to a linear transformation on h∗ which
sends a simple root in Π to a simple root in Πα . For example, for spo(2m|2n)
with a fundamental system Π corresponding to the first Dynkin diagram below,
and take α = δm − εn . The odd reflection transforms Π to Πα corresponding to
the second Dynkin diagram below (with the · · · portion unchanged). A plausible
transformation would have to fix all possible δi , ε j and interchange εn and δm . But
this transformation would not send δm + εn to 2δm .
1.3. Non-conjugate positive systems and odd reflections 29


δm + εn
⊗ ⊗
··· ··· ⃝ ⇐=⃝
εn−1 − δm@
@⊗ δm − εn
εn−1 − εn εn − δm 2δm

Example 1.30. Associated with gl(1|2), we have Φ0̄ = {±(ε1 − ε2 )}, and Φ1̄ =
{±(δ1 − ε1 ), ±(δ1 − ε2 )}. There are 6 sets of simple roots, that are related by real
and odd reflections as follows. There are three conjugacy classes of Borel subalge-
bras corresponding to the three columns below, and each vertical pair corresponds
to such a conjugacy class.

rδ1 −ε1 rδ1 −ε2


⊗ ⊗ ⊗ ⊗
⃝   ⃝
δ1 − ε1 ε1 − ε2 ε1 − δ1 δ1 − ε2 ε1 − ε2 ε2 − δ1

↕ rε1 −ε2
rδ1 −ε2
↕ rε1 −ε2
rδ1 −ε1
↕ rε1 −ε2
⊗ ⊗ ⊗ ⊗
⃝   ⃝
δ1 − ε2 ε2 − ε1 ε2 − δ1 δ1 − ε1 ε2 − ε1 ε1 − δ1

Example 1.31. Let g = gl(4|2). The following sequence of fundamental systems


Π0 = Πst , Π1 , Π2 , Π3 , Π4 (in lowering order) are obtained from the standard funda-
mental system Πst by applying consecutively the sequences of odd reflections with
respect to the odd roots δ4 − ε1 , δ3 − ε1 , δ2 − ε1 and δ1 − ε1 .


⃝ ⃝ ⃝ ⃝
XX
rδ4 −ε1
δ1 − δ2 δ2 − δ3 δ3 − δ4 δ4 − ε1 ε1 − ε2 X X ⊗ ⊗ ⊗
rδ3 −ε1 ⃝ ⃝

⊗ ⊗ ⊗  δ1 − δ2 δ2 − δ3 δ3 − ε1 ε1 − δ4 δ4 − ε2
⃝ ⃝
δ1 − δ2 δ2 − ε1 ε1 − δ3 δ3 − δ4 δ4 − ε2 XX
rδ2 −ε1
X X ⊗ ⊗ ⊗
rδ1 −ε1 ⃝ ⃝

⊗ ⊗  δ1 − ε1 ε1 − δ2 δ2 − δ3 δ3 − δ4 δ4 − ε2
⃝ ⃝ ⃝
ε1 − δ1 δ1 − δ2 δ2 − δ3 δ3 − δ4 δ4 − ε2

If we continue to apply to Π4 consecutively the sequence of odd reflections with


respect to the odd roots δ4 −ε2 , δ3 −ε2 , δ2 −ε2 and δ1 −ε2 , we obtain the following
fundamental systems Π4 , Π5 , Π6 , Π7 , Π8 :
30 1. Lie superalgebra ABC

⊗ ⊗
⃝ ⃝ ⃝
XX
rδ4 −ε2
X
ε1 − δ1 δ1 − δ2 δ2 − δ3 δ3 − δ4 δ4 − ε2 X ⊗ ⊗ ⊗
rδ3 −ε2 ⃝ ⃝

⊗ ⊗ ⊗  ε1 − δ1 δ1 − δ2 δ2 − δ3 δ3 − ε2 ε2 − δ4
⃝ ⃝
ε1 − δ1 δ1 − δ2 δ2 − ε2 ε2 − δ3 δ3 − δ4 XX
rδ2 −ε2
X X ⊗ ⊗ ⊗
rδ1 −ε2 ⃝ ⃝

⊗  ε1 − δ1 δ1 − ε2 ε2 − δ2 δ2 − δ3 δ3 − δ4
⃝ ⃝ ⃝ ⃝
ε1 − ε2 ε2 − δ1 δ1 − δ2 δ2 − δ3 δ3 − δ4

1.4. Highest weight theory


In this section, we formulate the Poincaré-Birkhoff-Witt Theorem for Lie superal-
gebras. Finite-dimensional irreducible modules over certain solvable Lie superal-
gebras, including all Borel subalgebras, are classified. This is then used to develop
a highest weight theory of the basic and queer Lie superalgebras.

1.4.1. The Poincaré-Birkhoff-Witt (PBW) Theorem. Let g = g0̄ ⊕ g1̄ be a Lie


superalgebra. A universal enveloping algebra of g is an associative superalgebra
U(g) together with a homomorphism of Lie superalgebras ι : g → U(g) character-
ized by the following universal property: given an associative superalgebra A and
a homomorphism of Lie superalgebras φ : g → A, there exists a unique homomor-
phism of associative superalgebras ψ : U(g) → A such that the following diagram
commutes:
ι / U(g)
gB
B BB
BB ∃ψ
φ BBB
! 
A
In particular, this implies that representations of g are equivalent to representa-
tions of U(g). It follows by a standard tensor algebra construction that a universal
enveloping algebra exists, and it is unique up to isomorphism by the defining uni-
versal property.
Theorem 1.32 (Poincaré-Birkhoff-Witt Theorem). Let {x1 , x2 , . . . , x p } be a basis
for g0̄ and {y1 , y2 , . . . , yq } be a basis for g1̄ . The set
{ r1 r2 r s }
x1 x2 . . . x pp ys11 ys22 . . . yqq | r1 , r2 , . . . , r p ∈ Z+ , s1 , s2 , . . . , sq ∈ {0, 1}
is a basis for U(g).

The proof for Theorem 1.32 is a straightforward super generalization of the


Lie algebra case, and will be omitted (see [Ro, Theorem 2.1] for detail).
1.4. Highest weight theory 31

The superalgebra U(g) carries a filtered algebra structure by letting


deg(x1r1 x2r2 . . . x pp ys11 ys22 . . . yqq ) = ∑ ri + ∑ s j ,
r s

i j

and its associated graded algebra is isomorphic to S(g0̄ ) ⊗ ∧(g1̄ ).


Given a Lie subalgebra l of a finite-dimensional Lie superalgebra g and an
l-module V , we define the induced module
IndglV = U(g) ⊗U(l) V.
By the PBW Theorem, if V is finite-dimensional, then so is Indgg0̄ V .

1.4.2. Representations of solvable Lie superalgebras. Just as for Lie algebras,


a finite-dimensional Lie superalgebra g = g0̄ ⊕ g1̄ is called solvable if g(n) = 0 for
some n ≥ 1, where we define inductively g(n) = [g(n−1) , g(n−1) ] and g(0) = g.
Let g = g0̄ ⊕ g1̄ be a finite-dimensional solvable Lie superalgebra such that
[g1̄ , g1̄ ] ⊆ [g0̄ , g0̄ ]. Given λ ∈ g∗0̄ with λ([g0̄ , g0̄ ]) = 0, we define a one-dimensional
g-module Cλ = Cvλ , where
xvλ = λ(x)vλ , for x ∈ g0̄ ,
yvλ = 0, for y ∈ g1̄ .
There is a canonical linear isomorphism (g0̄ /[g0̄ , g0̄ ])∗ ∼
= {λ ∈ g∗0̄ | λ([g0̄ , g0̄ ]) = 0}.
Lemma 1.33. Let g = g0̄ ⊕ g1̄ be a finite-dimensional solvable Lie superalgebra
such that [g1̄ , g1̄ ] ⊆ [g0̄ , g0̄ ]. Then every finite-dimensional irreducible g-module
is one-dimensional. A complete list of finite-dimensional irreducible g-modules is
given by Cλ , for λ ∈ (g0̄ /[g0̄ , g0̄ ])∗ .

Proof. Clearly every one-dimensional g-module is isomorphic to Cλ , for some


λ ∈ (g0̄ /[g0̄ , g0̄ ])∗ . So it suffices to show that every finite-dimensional irreducible
g-module V is one dimensional.
(n)
The even subalgebra g0̄ is solvable since g is solvable and g0̄ ⊆ g(n) for every
n. By applying Lie’s theorem to g0̄ , V contains a non-zero g0̄ -invariant vector vλ ,
where λ ∈ g∗0̄ with λ([g0̄ , g0̄ ]) = 0. Here, by a g0̄ -invariant vector we mean that
xvλ = λ(x)vλ , ∀x ∈ g0̄ . Now Frobenius reciprocity implies that there exists a g-
epimorphism from Indgg0̄ Cvλ ∼ = ∧(g1̄ ) ⊗ Cvλ onto V . Thus, to prove that V is one
dimensional, it suffices to prove that Indgg0̄ Cvλ has a composition series with one-
dimensional composition factors. We shall construct such a composition series
explicitly.
Set dim g1̄ = n. By Lie’s theorem, the g0̄ -module g1̄ has an ordered basis
{y j | j = 1, . . . , n} such that for all 1 ≤ i ≤ n we have
n n
(1.31) ad g0̄ (yi ) ⊆ ∑ Cy j , ad [g0̄ , g0̄ ](yi ) ⊆ ∑ Cy j .
j=i j=i+1
32 1. Lie superalgebra ABC

Let B1 = {1, y1 }, and define inductively Bk = {Bk−1 , Bk−1 yk }, for 2 ≤ k ≤ n,


where the set Bk−1 yk denotes the ordered set of elements obtained by multiplying
all elements in Bk−1 on the right by yk . Then Bn vλ is an ordered basis for Indgg0̄ Cvλ
n
of cardinality 2n . Denote this ordered basis by {v1 , · · · , v2n }. Set Vi := ⊕2j=i Cv j .
Then we have a filtration of g0̄ -modules
V = V1 ⊇ V2 ⊇ V3 ⊇ · · · ⊇ · · ·V2n ⊇ 0.
Clearly we have, for 1 ≤ k, ℓ ≤ n and i1 < . . . < iℓ ≤ n,
(1.32) yk yi1 yi2 . . . yiℓ vλ = [yk , yi1 ] yi2 . . . yiℓ vλ − yi1 yk yi2 . . . yiℓ vλ .

Using the assumption that [g1̄ , g1̄ ] ⊆ [g0̄ , g0̄ ] , it follows from (1.31) and (1.32)
that every yk leaves Vi invariant, and hence each Vi is a g-module. Thus, the above
filtration is a composition series with one-dimensional composition factors. 
Example 1.34. The Lie superalgebra g = Cz ⊕ g1̄ in Example 1.4(4) associated to
a non-degenerated symmetric bilinear form B on a nonzero space g1̄ is solvable.
But an irreducible module of g with the central element z acting as a nonzero
scalar has dimension more than one, as they are irreducible modules of the Clifford
superalgebra on (g1̄ , B). So the condition [g1̄ , g1̄ ] ⊆ [g0̄ , g0̄ ] in Lemma 1.33 cannot
be dropped.

1.4.3. Highest weight theory for basic Lie superalgebras. Let g be a basic Lie
superalgebra. Let h be the standard Cartan subalgebra and Φ be the root system.
Let b = h ⊕ n+ be a Borel subalgebra of g = b ⊕ n− and let Φ+ be the associated
positive system. The condition of Lemma 1.33 is satisfied for the solvable Lie
superalgebra b, since we have b1̄ = n1̄ , and
[b1̄ , b1̄ ] = [n1̄ , n1̄ ] ⊆ n0̄ = [h, n0̄ ] ⊆ [b0̄ , b0̄ ].
Let V be a finite-dimensional irreducible representation of g. Then by Lemma 1.33
V contains a one-dimensional b-module, which is of the form Cλ = Cvλ , for λ ∈
h∗ ∼
= (b/[b, b])∗ . That is,
hvλ = λ(h)vλ (h ∈ h), xvλ = 0 (x ∈ n+ ).
By the PBW theorem and the irreducibility of V , we obtain that V = U(n− )vλ , and
thus a weight space decomposition

(1.33) V= Vµ ,
µ∈h∗

where the µ-weight space Vµ is given by


Vµ := {v ∈ V | hv = µ(h)v, ∀h ∈ h}.
By (1.33), Vµ = 0 unless λ − µ is a Z+ -linear combination of positive roots. The
weight λ is called the b-highest weight (and sometime called an extremal weight)
of V , the space Cvλ is called the b-highest weight space, and the vector vλ is called
1.4. Highest weight theory 33

a b-highest weight vector for V . When no confusion arises, we will simply say
highest weight by dropping b. Hence, we have established the following.

Proposition 1.35. Let g be a basic Lie superalgebra with a Borel subalgebra b.


Any finite-dimensional irreducible g-module is a b-highest weight module.

We shall denote the highest weight irreducible module of highest weight λ


by L(λ), L(g, λ), or L(g, b, λ), depending on whether b and g are clear from the
context.
Recall from 1.3.6 the notations Πα and bα associated to an isotropic odd simple
root α. Denote by ⟨·, ·⟩ : h∗ × h → C the standard bilinear pairing. Denote by hα
the corresponding coroot for α, and denote by eα and fα the root vectors of roots
α and −α, respectively.

Lemma 1.36. Let L be a simple g-module and let v be a b-highest weight vector
of L of b-highest weight λ. Let α be an isotropic odd simple root.

(1) If ⟨λ, hα ⟩ = 0, then L is a g-module of bα -highest weight λ and v is a


bα -highest weight vector.
(2) If ⟨λ, hα ⟩ ̸= 0, then L is a g-module of bα -highest weight (λ − α) and fα v
is a bα -highest weight vector.

Proof. We first observe three simple identities:


(i) eα fα v = [eα , fα ]v = hα v = ⟨λ, hα ⟩v.
(ii) eβ fα v = [eβ , fα ]v = 0 for any β ∈ Φ+ ∩ Φ+
α , since either β − α is not a root
or it belongs to Φ+ ∩ Φ+ α .
(iii) fα2 v = 0, since α is an isotropic odd root.
Now, we consider the two cases separately.
(1) Assume that ⟨λ, hα ⟩ = 0. Then we must have fα v = 0, for otherwise fα v
would be a b-singular vector in the simple g-module V by (i) and (ii). This together
with Lemma 1.26 implies that v is a bα -highest weight vector of weight λ in the
g-module V .
(2) Assume that ⟨λ, hα ⟩ ≠ 0. Then (i), (ii), (iii) and Lemma 1.26 imply that fα v
is nonzero and it is a bα -highest weight vector of weight λ − α in V . 

Example 1.37. Let g = gl(4|2). Denote the sequence of Borel subalgebras cor-
responding to the fundamental systems Πi in Example 1.31 as bi , for 0 ≤ i ≤ 8,
with bst = b0 . Consider the finite-dimensional irreducible gl(4|2)-module L(λ),
where λ = a1 δ1 + a2 δ2 + a3 δ3 + a4 δ4 + b1 ε1 + b2 ε2 with a1 ≥ a2 ≥ a3 ≥ a4 ≥ 2
and b1 ≥ b2 ≥ 0. We identify λ = (a1 , a2 , a3 , a4 |b1 , b2 ). The bi -extremal weights,
34 1. Lie superalgebra ABC

denoted by λi , for 0 ≤ i ≤ 4 are computed as follows:


λ0 = λ = (a1 , a2 , a3 , a4 |b1 , b2 ),
λ1 = (a1 , a2 , a3 , a4 − 1|b1 + 1, b2 ),
λ2 = (a1 , a2 , a3 − 1, a4 − 1|b1 + 2, b2 ),
λ3 = (a1 , a2 − 1, a3 − 1, a4 − 1|b1 + 3, b2 ),
λ4 = (a1 − 1, a2 − 1, a3 − 1, a4 − 1|b1 + 4, b2 ).
If we continue to applying λ4 consecutively the sequence of odd reflections with
respect to the odd roots δ4 −ε2 , δ3 −ε2 , δ2 −ε2 and δ1 −ε2 , we obtain the following
bi -extremal weights λi , for 5 ≤ i ≤ 8:
λ5 = (a1 − 1, a2 − 1, a3 − 1, a4 − 2|b1 + 4, b2 + 1),
λ6 = (a1 − 1, a2 − 1, a3 − 2, a4 − 2|b1 + 4, b2 + 2),
λ7 = (a1 − 1, a2 − 2, a3 − 2, a4 − 2|b1 + 4, b2 + 3),
λ8 = (a1 − 2, a2 − 2, a3 − 2, a4 − 2|b1 + 4, b2 + 4).
We shall see in Chapter 2, Section 2.4 that all these weights λi afford very simple
visualization in terms of Young diagrams.

1.4.4. Highest weight theory for q(n). Let g = q(n) be the queer Lie superalge-
bra. We recall several subalgebras of g from 1.2.6. Let h be the standard Car-
tan subalgebra consisting of block diagonal matrices and let b be the standard
Borel subalgebra of block upper triangular matrices of g. We have b = h ⊕ n+
and g = b ⊕ n− . The Cartan subalgebra h = h0̄ ⊕ h1̄ is a solvable but nonabelian
Lie superalgebra, since [h1̄ , h1̄ ] = h0̄ and [h0̄ , h0̄ + h1̄ ] = 0.
For λ ∈ h∗0̄ , define a symmetric bilinear form ⟨·, ·⟩λ on h1̄ by
⟨v, w⟩λ := λ([v, w]), for v, w ∈ h1̄ .
Denote by Rad⟨·, ·⟩λ the radical of the form ⟨·, ·⟩λ . Then the form ⟨·, ·⟩λ descends
to a nondegenerate symmetric bilinear form on h1̄ /Rad⟨·, ·⟩λ , and it gives rise to a
Clifford superalgebra C = Cλ := C(h1̄ /Rad⟨·, ·⟩λ ). By definition we have an iso-
morphism of superalgebras
(1.34) Cλ ∼
= U(h)/Iλ ,
where Iλ denotes the ideal of U(h) generated by Rad⟨·, ·⟩λ and a − λ(a) for a ∈ h0̄ .
Let h′1̄ ⊆ h1̄ be a maximal isotropic subspace with respect to ⟨·, ·⟩λ , and define
the Lie subalgebra h′ := h0̄ ⊕ h′1̄ . The one-dimensional h0̄ -module Cvλ , defined by
hvλ = λ(h)vλ , extends to an h′ -module by letting h′1̄ .vλ = 0. Define the induced
h-module
Wλ := Indhh′ Cvλ .
1.4. Highest weight theory 35

Recall the well-known fact that the Clifford superalgebra C admits a unique
eλ (See ?? in Chapter 3).
irreducible (Z2 -graded) module W
Lemma 1.38. For λ ∈ h∗0̄ , the h-module Wλ is isomorphic to W eλ (viewed as an
h-module via the pullback through (1.34)), and is irreducible. Furthermore, every
finite-dimensional irreducible h-module is isomorphic to Wλ , for some λ ∈ h∗0̄ .

Lemma 1.38 shows that the h-module Wλ is independent of a choice of a max-


imal isotropic subspace h′1̄ .

Proof. The action of U(h) on Wλ descends to an action of U(h)/Iλ , and via (1.34)
eλ of the Clifford superalgebra
we identify Wλ with the unique irreducible module W
Cλ . Hence, the h-module Wλ is irreducible.
Suppose we are given an irreducible h-module U. Then it contains an h0̄ -
weight vector v′λ of weight λ ∈ h∗0̄ . Recall that h′ = h0̄ ⊕ h′1̄ , and consider the
( )
h′ -submodule π,U(h′ )v′λ of U such that π([h′1̄ , h′1̄ ]) = π([h′0̄ , h′0̄ ]) = 0. By ap-
plying Lemma 1.33 to Lie superalgebra π(h′ ), there exists a one-dimensional h′ -
submodule Cvλ of U(h′ )v′λ ⊆ U. By Frobenious reciprocity we have a surjective
h-homomorphism from Wλ onto U. Since Wλ is irreducible, we have U ∼ = Wλ . 

Let V be a finite-dimensional irreducible g-module. Pick an irreducible h-


module Wλ in V , where λ ∈ h∗0̄ can be taken to be maximal in the partial order
induced by the positive system Φ+ by the finite dimensionality of V . By defini-
tion, Wλ is h0̄ -semisimple of weight λ. For any α ∈ Φ+ with associated even root
vector eα and odd root vector eα in n+ , the space CeαWλ + CeαWλ is an h-module
which is h0̄ -semisimple of weight λ + α. If CeαWλ + CeαWλ ̸= 0 for some α ∈ Φ+ ,
then it contains an isomorphic copy of Wλ+α as an h-submodule, contradicting the
maximal weight assumption of λ. Hence, we have n+Wλ = 0. By irreducibility of
V we must have U(n− )Wλ = V , which gives rise to a weight space decomposition

of V = µ∈h∗ Vµ . The space Wλ = Vλ is the highest weight space of V , and it

completely determines the irreducible module V . We denote V by L(g, λ), or L(λ),
if g is evident from the context. Summarizing, we have proved the following.
Proposition 1.39. Let g = q(n). Any finite-dimensional irreducible g-module is a
highest weight module.

Let ℓ(λ) be the number of nonzero parts in a composition λ (the notation here
is consistent with the notation for length ℓ(λ) of a partition λ). We set
{
0, if ℓ(λ) is even,
(1.35) δ(λ) =
1, if ℓ(λ) is odd.
Now for λ ∈ h∗ , recalling the notation Hi from 1.2.6, we identify λ with the
composition λ = (λ(H1 ), . . . , λ(Hn )), and hence, ℓ(λ) is equal to the dimension
of the space h1̄ /Rad⟨·|·⟩λ . We remark that the highest weight space Wλ of L(λ)
36 1. Lie superalgebra ABC

has dimension 2(ℓ(λ)+δ(λ))/2 . Note that the Clifford algebra Cλ admits an odd au-
tomorphism if and only if ℓ(λ) is odd. Hence, the h-module Wλ , or equivalently
the irreducible C-module Wλ , has an odd automorphism if and only if ℓ(λ) is an
odd integer. An automorphism of the irreducible g-module L(λ) clearly induces
an h-module automorphism of its highest weight space. Conversely, any h-module
automorphism on Wλ induces an automorphism of the g-module IndgbWλ . Since an
automorphism preserves the maximal submodule, it induces an automorphism of
the unique irreducible quotient g-module. We have proved the following.
Lemma 1.40. Let g = q(n) be the queer Lie superalgebra with a Cartan subalge-
bra h. Let L(λ) be the irreducible g-module of highest weight λ ∈ h∗0̄ . Then,
dim(Endg (L(λ)) = 2δ(λ) .

This suggests that Schur’s Lemma for superalgebras requires modification.


This will be discussed in depth in Chapter 3, Lemma 3.4.

1.5. Notes
Section 1.1. The classification of finite-dimensional simple complex Lie superal-
gebras was first announced by Kac in [K4]. The detailed proof of the classifica-
tion, along with many other fundamental results on Lie superalgebras, appeared in
Kac [K1] two years later. An independent proof of the classification of the finite-
dimensional complex simple Lie superalgebras whose even subalgebras are reduc-
tive was given in the two papers by Scheunert, Nahm, and Rittenberg in [SNR]
around the same time.
Section 1.2. The structure theory of root systems, root space decompositions
and invariant bilinear forms of basic Lie superalgebras presented here is fairly stan-
dard. More details can be found in the standard references (see Kac [K1, K2] and
Scheunert [Sch]).
Section 1.3. Odd reflections was introduced by Leites, Saveliev, and Serganova
[LSS] to relate non-conjugate Borel subalgebras, fundamental and positive sys-
tems. The list of conjugacy classes of fundamental systems under the Weyl group
action for the basic Lie algebras was given by Kac [K1]. We introduce a no-
tion of εδ-sequences (which appeared in Cheng-Wang [CW5]) to facilitate the
parametrization of the conjugacy classes of fundamental systems for Lie super-
algebras of type gl and osp. The fundamental Lemma 1.26 on odd reflections is
sometimes attributed to Serganova’s 1988 thesis (cf., e.g., [Sva2]), and more gen-
erally for Kac-Moody Lie superalgebras in Kac-Wakimoto [KW, Lemma 1.2].
Section 1.4. A detailed proof of the PBW Theorem for Lie superalgebras, The-
orem 1.32, can be found in Milnor-Moore [MM, Theorem 6.20] and Ross [Ro,
Theorem 2.1]. As in the classical theory, the first step to develop a highest weight
theory for the basic and queer Lie superalgebras is to study representations of Borel
1.5. Notes 37

subalgebras. Lemma 1.33 on solvable Lie superalgebras appeared in Kac [K1,


Proposition 5.2.4]. The proof given here is different. Lemma 1.36 on the change
of the extremal weights of an irreducible module under an odd reflection (which
appeared in Penkov-Serganova [PS, Lemma 1]) plays a fundamental role in repre-
sentation theory of Lie superalgebras developed in the book.
Chapter 2

Finite-dimensional
modules

In this chapter, we mainly work on Lie superalgebras of type gl, osp or q. The
finite-dimensional irreducible modules for these Lie superalgebras are classified in
terms of highest weights. In contrast to semisimple Lie algebras, there is more
work to do to achieve such classifications for Lie superalgebras. We describe the
images of the Harish-Chandra homomorphisms in terms of (super)symmetric func-
tions and formulate the linkage principle on composition factors of a Verma mod-
ule. As an application, we present a Weyl-type character formula for the so-called
typical finite dimensional irreducible modules. The chapter concludes with a study
of extremal weights, i.e., highest weights with respect to (not necessarily conju-
gate) Borel subalgebras, of various finite dimensional irreducible modules.

2.1. Classification of finite-dimensional simple modules


In this section, the highest weights of finite-dimensional irreducible modules over
Lie superalgebras of type gl, osp and q are determined.

2.1.1. Finite-dimensional simple modules of gl(m|n). Let g be the Lie superal-


gebra gl(m|n) and let h be the Cartan subalgebra of diagonal matrices spanned by
the basis elements {Eii |i ∈ I(m|n)}. Let n+ (respectively, n− ) be the subalgebra of
strictly upper (respectively, strictly lower) triangular matrices of gl(m|n). Then we
have the standard triangular decomposition

g = n− ⊕ h ⊕ n+ .

39
40 2. Finite-dimensional modules

The even subalgebra admits a compatible triangular decomposition


g0̄ = n−

⊕ h ⊕ n+

,
where n±

= g0̄ ∩ n± . Let b = h ⊕ n+ and b0̄ = h ⊕ n+

.
Moreover, the Lie superalgebra g admits a Z-gradation
(2.1) g = g−1 ⊕ g0 ⊕ g1 ,
where g−1 (respectively, g1 ) is spanned by all Ei j with i, j ∈ I(m|n) such that i >
0 > j (respectively, i < 0 < j). Note that g−1 and g1 are abelian Lie superalgebras,
and that the Z-degree zero subspace g0 coincides with the Z2 -degree zero subspace
g0̄ .
For λ ∈ h∗ , let L0 (λ) be the simple g0̄ -module of highest weight λ ∈ h∗ (relative
to the Borel b0̄ ). Then L0 (λ) may be extended trivially to a g0̄ ⊕ g1 -module due to
(2.1). Define the Kac module over g by
K(λ) = Indgg0 ⊕g1 L0 (λ),
which, as a vector space, can be identified by the PBW Theorem 1.32 with
(2.2) K(λ) = ∧(g−1 ) ⊗ L0 (λ).
Proposition 2.1. Let g = gl(m|n). There exists a surjective g-module homomor-
phism (unique up to a scalar multiple) K(λ)  L(λ). Moreover, the following are
equivalent:
(1) L(λ) is finite dimensional.
(2) L0 (λ) is finite dimensional.
(3) K(λ) is finite dimensional.

Proof. The existence of a surjective homomorphism K(λ)  L(λ) follows by the


embedding of g0̄ -modules L0 (λ) ,→ L(λ) and the Frobenius reciprocity.
(1) ⇒ (2). Note that L0 (λ) is an irreducible summand of L(λ) regarded as
g0̄ -module.
(2) ⇒ (3). Follows from (2.2).
(3) ⇒ (1). Follows from the surjectivity of the map K(λ)  L(λ). 

According to Proposition 1.35, every finite-dimensional simple g-module is a


highest weight module L(λ), for some λ. Moreover, L(λ) ∼ ̸ L(µ), if λ ̸= µ. By
=
Proposition 2.1, the classification of finite-dimensional simple g-modules is the
same as the classification of finite-dimensional simple modules for g0̄ = gl(m) ⊕
gl(n), which is well known. Hence, we have established the following.
Proposition 2.2. A complete list of pairwise non-isomorphic finite-dimensional
simple gl(m|n)-modules are L(λ), for λ = ∑m ∗
i=1 λi δi + ∑ j=1 υ j ε j ∈ h satisfying
n

λi − λi+1 ∈ Z+ and υ j − υ j+1 ∈ Z+ for all possible i, j.


2.1. Classification of finite-dimensional simple modules 41

Remark 2.3. Note that g0̄ is a Levi subalgebra of g = gl(m|n) (corresponding to


the removal of the odd simple root from the standard Dynkin diagram of g), and
hence the Kac module K(λ) is a distinguished parabolic Verma module relative
to the standard Borel subalgebra. Since K(λ) is a highest weight module, it is
indecomposable. However, for m ≥ 1 and n ≥ 1, K(λ) is reducible for suitable
λ, e.g., when λ(hα ) = 0 for the odd simple root α by Lemma 1.36. Hence, the
category of finite-dimensional gl(m|n)-modules is not semisimple when m ≥ 1 and
n ≥ 1.

2.1.2. Finite-dimensional simple modules of spo(2m|2). The classification of


finite-dimensional simple modules for g = spo(2m|2) is carried out just like the
gl(m|n) case. Let Φ+ be the standard positive system of the root system Φ of g.
Associated to Φ+ we have the triangular decomposition g = n− ⊕ h ⊕ n+ , compat-
ible with that of its even subalgebra g0̄ = n−0̄
⊕ h ⊕ n+ 0̄
, where n±

= g0̄ ∩ n± . In
addition, g = spo(2m|2) admits the following two favorable properties which are
shared by gl(m|n) but not by other spo superalgebras: first, g admits a Z-graded Lie
superalgebra structure of the form g = g−1 ⊕ g0 ⊕ g1 , where g0 = g0̄ , g1̄ = g−1 ⊕ g1 ,
and the set of roots for g1 is given by Φ+

= {ε1 ± δi , 1 ≤ i ≤ m}. Secondly, g0̄ is a
Levi subalgebra of g corresponding to the removal of the odd simple root from the
standard Dynkin diagram of g.
Hence we define the Kac module K(λ) of g as before: K(λ) = Indgg0 ⊕g1 L0 (λ),
where L0 (λ) denotes the simple g0̄ -module of highest weight λ relative to Φ+ 0̄
.
Proposition 2.1 remains valid in the current setting, and this implies the classifi-
cation of finite-dimensional simple g-modules is the same as the classification of
finite-dimensional simple modules for g0̄ ∼
= sp(2m) ⊕ C.
Proposition 2.4. A complete list of pairwise non-isomorphic finite-dimensional
simple spo(2m|2)-modules are L(λ), for λ = ∑m ∗
i=1 λi δi +υ1 ε1 ∈ h such that υ1 ∈ C
and (λ1 , . . . , λm ) is a partition.

2.1.3. A virtual character formula. Let g be a Lie superalgebra of type gl or osp.


With respect to the standard positive system Φ+ = Φ+ 0̄
∪ Φ+

(see Sections 1.3.2,

1.3.3 and 1.3.4), we have the triangular decomposition g = n ⊕ h ⊕ n+ . Letting
b = n+ ⊕ h, we define the Verma module
∆(λ) = Indgb Cvλ
associated to a weight λ ∈ h∗ . Here, as usual, Cvλ stands for the one-dimensional
b-module on which h acts by the character λ, and n+ acts trivially. Then ∆(λ)
admits a unique simple quotient g-module, which is the irreducible g-module L(λ)
of highest weight λ.
A weight λ ∈ h∗ is called dominant integral with respect to Φ+ 0̄
or sim-
ply Φ+

-dominant integral, if ⟨λ, h α ⟩ ∈ Z+ , for all α ∈ Φ+

. A difference of two
42 2. Finite-dimensional modules

characters of finite dimensional g-modules is called a virtual character of finite-


dimensional g-modules. Also recall the Weyl vector ρ = ρ0 − ρ1 from (1.28). We
shall need the following proposition (see Santos [San, Proposition 5.7]) to deter-
mine if certain weights could be highest weights of finite dimensional g-modules.
Proposition 2.5. Let g be of type gl or osp. For any Φ+

-dominant integral weight

λ ∈ h , the expression
∏β∈Φ+ (eβ/2 + e−β/2 )
(2.3)


∑ (−1)ℓ(w) ew(λ+ρ)
(eα/2 − e−α/2 ) w∈W
α∈Φ+

is a virtual character of finite-dimensional g-modules.

Proof. Following the proof of [San, Proposition 5.7], we shall make use of the
so-called Bernstein functor L0 [San, (12)] from the category of h-semisimple g-
modules to the category of g0̄ -semisimple g-modules. Actually, when restricting to
the category of h-semisimple g0̄ -module, the functor L0 is just the classical Bern-
stein functor, which is a special case of the functor P defined in Knapp and Vogan
[KV, (2.8)]. As we shall only need some standard properties of the Bernstein func-
tor, which can be found in [KV, San], we will not recall here the precise definition
which is quite involved.
An immediate consequence of the definition of the functor L0 is that every g0̄ -
composition factor of L0 (V ) is also a g0̄ -composition factor of V , counting multi-
plicity. Hence L0 takes a finitely generated h-semisimple g0̄ -module in the BGG
category to a finite-dimensional g0̄ -module. Denote by Li the ith derived functor of
L0 . Then, each Li (V ) is a finite-dimensional g-module, for any Verma g-module
V (which lies in the BGG category of g0̄ -modules). Now it is a classical result
(cf. [KV, Corollary 4.160]) that the resulting Euler characteristic when applying
L0 to a g0̄ -Verma module is given by the Weyl character formula, i.e.,

∑w∈W (−1)ℓ(w) ew(λ+ρ0̄ )
∑ (−1)i ch Li (Indb Cλ ) =
g0̄
,
i=0

∏α∈Φ+ (eα/2 − e−α/2 )

g
and Li (Indb0̄ Cλ ) = 0 for i ≫ 0.

Given a short exact sequence of finite-dimensional h-semisimple g0̄ -modules
0 −→ M −→ E −→ N −→ 0.
It follows by the Euler-Poincaré principle that
∞ ∞ ∞
∑ (−1)i ch Li (M) + ∑ (−1)i ch Li (N) = ∑ (−1)i ch Li (E).
i=0 i=0 i=0

As a g0̄ -module, ∆(λ) and Indb0̄ Cλ ⊗ Λ(n∗1̄ ) have Verma flags, where a Verma
g

flag is a filtration whose sections are Verma modules. Since these two g0̄ -modules
2.1. Classification of finite-dimensional simple modules 43

also have identical characters, their Verma flags have the same Verma module mul-
tiplicities. Now the character of the b0̄ -module Cλ ⊗ Λ(n∗1̄ ) equals to eλ ∏β∈Φ+ (1 +

e−β ). Noting the W -invariance of ∏β∈Φ+ (eβ/2 + e−β/2 ), we have

∞ ∞ ( )
∑ (−1)i ch Li (∆(λ)) = ∑ (−1)i ch Li Indb0̄ Cλ ⊗ Λ(n∗1̄ )
g

i=0 i=0
( )
∑w∈W (−1)ℓ(w) w eλ+ρ0̄ ∏β∈Φ+ (1 + e−β )

=
∏α∈Φ+ (eα/2 − e−α/2 )
( )

∑w∈W (−1) w eλ+ρ0̄ −ρ1̄ ∏β∈Φ+ (eβ/2 + e−β/2 )


ℓ(w)

=
∏α∈Φ+ (eα/2 − e−α/2 )

β/2 + e−β/2 )
∏β∈Φ + (e
=

(eα/2 − e−α/2 ) w∈W


∑ (−1)ℓ(w) ew(λ+ρ) .
α∈Φ+

Since the left hand side is a virtual character of finite-dimensional g-modules, so is


the right hand side. This proves the proposition. 
Remark 2.6. Upon changing the sign + in the numerator of (2.3) to −, the result-
ing expression is a virtual supercharacter (see Section 2.2.4) of finite-dimensional
g-modules.

2.1.4. Finite-dimensional simple modules of spo(2m|2n + 1). We start with a


general remark for Lie superalgebra g of type osp. By Proposition 1.35, every
finite-dimensional irreducible g-module is necessarily of the form L(λ), for some
λ ∈ h∗ . Moreover, we have L(λ) ∼ = L(µ) if and only if λ = µ. However, the clas-
sification of finite-dimensional simple g-modules is non-trivial, partly because the
even subalgebra g0̄ of g is not a Levi subalgebra.
Lemma 2.7. Let g be a basic Lie superalgebra. A necessary condition for the finite
dimensionality of the g-module L(λ) is that λ is Φ+

-dominant integral.

Proof. Follows by noting that λ is a highest weight with respect to Φ+



for L(λ),
regarded as a g0̄ -module. 

In this subsection we consider g = spo(2m|2n + 1) and let Φ+ be the stan-


dard positive system for g. The next lemma is a reformulation of the well-known
dominance integral (D.I.) conditions for Lie algebras sp(2m) and so(2n + 1).
Lemma 2.8. Let g = spo(2m|2n + 1) with g0̄ = sp(2m) ⊕ so(2n + 1). A weight
λ = ∑mi=1 λi δi + ∑ j=1 υ j ε j is Φ0̄ -dominant integral if and only if the following two
n +

conditions are satisfied:


(DI.c) λ1 ≥ . . . ≥ λm with all λi ∈ Z+ .
(DI.b) υ1 ≥ . . . ≥ υn , with either (i) all υ j ∈ Z+ , or (ii) all υ j ∈ 12 + Z+ .
44 2. Finite-dimensional modules

For g = spo(2m|2n + 1), a Φ+ 0̄


-dominant integral weight is called an integer
weight if it satisfies (DI.b-i), and it is called a half-integer weight if it satisfies
(DI.b-ii). The L(λ)’s for integer and half-integer weights λ turn out to be quite
different, and they are analyzed separately.

Proposition 2.9. Let g = spo(2m|2n + 1) and let λ = ∑m ∗


i=1 λi δi + ∑ j=1 υ j ε j ∈ h
n

be a half-integer Φ0̄ -dominant integral weight. Then the highest weight g-module
+

L(λ) is finite dimensional if and only if λm ≥ n.

Proof. Take a half-integer Φ+0̄


-dominant integral weight λ such that λm ≥ n. We
will look for the highest weight appearing in the virtual character (2.3) of finite-
dimensional g-modules in Proposition 2.5. Recalling the Weyl vector ρ = ρ0 − ρ1
from (1.28), we compute that
2n + 1 m
ρ1 =
2 i=1∑ δi ,
ρ0 = mδ1 + (m − 1)δ2 + . . . + δm + (n − 1/2)ε1 + (n − 3/2)ε2 + . . . + (1/2)εn .

Since the Weyl group of sp(2m) is ∼ = Zn2 o Sn and it consists of the signed permu-
tations among δ1 , . . . , δm , we conclude by the assumption λm ≥ n that the weight
λ + ρ remains Φ+ 0̄
-dominant and regular. Thus the highest weight occurring in the
virtual character of finite-dimensional g-modules (2.3) is (λ + ρ) + ρ1 − ρ0 = λ,
and so L(λ) must be finite dimensional.
Conversely, suppose that L(λ) is finite dimensional, for a half-integer Φ+ 0̄
-
dominant integral weight λ. We apply the following sequence of mn odd reflections
to the standard fundamental and positive systems of g (see 1.3.3). We first apply
the n odd reflections with respect to δm − ε1 , δm − ε2 , . . . , δm − εn , then the n odd
reflections with respect to δm−1 − ε1 , . . . , δm−1 − εn etc., until we finally apply the
n odd reflections with respect to δ1 − ε1 , δ1 − ε2 , . . . , δ1 − εn . After a computation
similar to Example 1.31, the resulting fundamental system has the following form


⃝ ⃝ ··· ⃝ ··· ⃝=⇒a
ε1 − ε2 ε2 − ε3 εn − δ1 δ1 − δ2 δm−1 − δm δm

By the half-integer weight assumption on λ, the value of λ at any odd coroot be-
longs to 12 + Z (and hence is nonzero). By Lemma 1.36, after an odd reflection with
respect to an odd root α the new highest weight is λ − α, which again takes value
in 12 + Z at any odd coroot. Continuing this way, by a computation very similar
to Example 1.37 we see that the new highest weight corresponding to the above
Dynkin diagram is equal to λ − n ∑m i=1 δi + m ∑ j=1 ε j . Now this weight must take
n

non-negative integer value at h2δm by the finite dimensionality of L(λ), from which
we conclude that λm ≥ n. 
2.1. Classification of finite-dimensional simple modules 45

We next determine for which integer Φ+ 0̄


-dominant integral weight λ the g-
module L(λ) is finite dimensional. Recall µ′ denotes the conjugate partition of a
partition µ. The following definition plays a fundamental role in the book.
Definition 2.10. A partition µ = (µ1 , µ2 , . . .) is called an (m|n)-hook partition
(or simply a hook partition when m, n are implicitly understood), if µm+1 ≤ n.
Equivalently, µ is an (m|n)-hook partition if µ′n+1 ≤ m.

Diagrammatically, the hook condition means that the (m + 1, n + 1)-box in the


Young diagram is missing, and it can be visualized as follows:
n

µ
m

Given an (m|n)-hook partition µ, we denote by µ+ = (µm+1 , µm+2 , . . .) and write


its conjugate, which is necessarily of length ≤ n, as ν = (µ+ )′ = (ν1 , . . . , νn ). We
define the weights
(2.4) µ♮ = µ1 δ1 + . . . + µm δm + ν1 ε1 + . . . + νn−1 εn−1 + νn εn ,
µ♮− = µ1 δ1 + . . . + µm δm + ν1 ε1 + . . . + νn−1 εn−1 − νn εn .
(µ♮− is only used for spo(2m|2n) below.) It is sometimes convenient to identify µ♮
and µ♮− with the following tuples of integers:
µ♮ = (µ1 , . . . , µm ; ν1 , . . . , νn ),
µ♮− = (µ1 , . . . , µm ; ν1 , . . . , −νn ).
Theorem 2.11. Let g = spo(2m|2n + 1), and let λ = ∑m ∗
i=1 λi δi + ∑ j=1 υ j ε j ∈ h be
n

a Φ+

-dominant integral weight.
(1) Suppose that λ is a half-integer weight. Then L(λ) is finite dimensional if
and only if λm ≥ n.
(2) Suppose that λ is an integer weight. Then L(λ) is finite dimensional if
and only if λ is of the form µ♮ , where µ is an (m|n)-hook partition.

Proof. Part (1) is Proposition 2.9. So it remains to prove (2).


Let µ be an (m|n)-hook partition. For M ≥ ℓ(µ) consider the Lie superalge-
bra spo(2M|2n + 1) acting on its standard representation C2M|2n+1 . It has highest
weight δ1 . The module ∧k (C2M|2n+1 ) contains a highest weight vector of highest
46 2. Finite-dimensional modules

weight ∑kj=1 δ j , for k ≤ M. Write the conjugate partition µ′ = (ν1 , ν2 , . . .). Then
the spo(2M|2n + 1)-module ∧ν1 (C2M|2n+1 ) ⊗ ∧ν1 (C2M|2n+1 ) ⊗ · · · contains the ir-
reducible spo(2M|2n + 1)-module V of highest weight µ as a quotient, and hence
V is finite dimensional. Now, we use a sequence of odd reflections to change
the standard Borel of spo(2M|2n + 1) into a Borel subalgebra that contains as a
subalgebra the standard Borel of spo(2m|2n + 1). (Here spo(2m|2n + 1) is re-
garded as a subalgebra of spo(2M|2n + 1)). A sequence of odd reflections that
we may use for this purpose is the following: We first apply the n odd reflections
with respect to δM − ε1 , δM − ε2 , . . . , δM − εn , then the n odd reflections with re-
spect to δM−1 − ε1 , . . . , δM−1 − εn etc., until we finally apply the n odd reflections
with respect to δm+1 − ε1 , δm+1 − ε2 , . . . , δm+1 − εn . Thus, we have applied a total
of (M − m)n odd reflections. The subdiagram consisting of the first (m + n − 1)
vertices of the new Dynkin diagram is of type gl(m|n). In a similar way as in Ex-
ample 1.37, one shows that the highest weight of V with respect to the new Borel
is µ♮ . Hence by restriction we conclude that the spo(2m|2n + 1)-module L(µ♮ ) is
finite dimensional, proving the “if” direction of (2).
By Lemma 2.8, an integer Φ+ 0̄
-dominant integral highest weight for a simple
finite-dimensional g-module is necessarily of the form µ = µ1 δ1 + . . . + µm δm +
ν1 ε1 + . . . + νn εn , where (µ1 , . . . , µm ) and (ν1 , . . . , νn ) are partitions. To prove the
remaining condition ν′1 ≤ µm , it suffices to prove it in the case of m = 1 by noting
that spo(2|2n + 1) is a subalgebra of spo(2m|2n + 1). Let V be a finite-dimensional
simple spo(2|2n + 1)-module. Via the sequence of odd reflections with respect to
δ1 − ε1 , δ1 − ε2 , . . . , δ1 − εn we change the standard Borel to a Borel with an odd
non-isotropic simple root δ1 . Now if µ1 ≥ n, then there is no condition on µ and
we are done. Suppose now that µ1 = k < n, and we shall prove by contradiction
by assuming that νk+1 ≥ 1 (and so ν′1 > µ1 ). This implies that νi ≥ 1, for all
1 ≤ i ≤ k, and hence, when applying the odd reflection with respect to to δ1 − εi ,
the new highest weight is obtained from the old one by subtracting δ1 − εi , for
1 ≤ i ≤ k. This remains true also when applying the odd reflection with respect
to δ1 − εk+1 . Hence the resulting new highest weight has negative δ1 -coefficient.
But this implies that the new highest weight evaluated at the coroot h2δ1 must be
negative, contradicting the finite dimensionality of the module V . Hence µ must
satisfy the hook condition, as claimed. 

2.1.5. Finite-dimensional simple modules of spo(2m|2n). Let n ≥ 2. Let g =


spo(2m|2n) with g0̄ = sp(2m) ⊕ so(2n). The next lemma is the well-known domi-
nance integral conditions for classical Lie algebras sp(2m) and so(2n).
Lemma 2.12. Let g = spo(2m|2n). Assume n ≥ 2. A weight λ = ∑m i=1 λi δi +
∑nj=1 υ j ε j ∈ h∗ is Φ+

-dominant integral if and only if the following two conditions
are satisfied:
(DI.c) λ1 ≥ . . . ≥ λm with all λi ∈ Z+ .
2.1. Classification of finite-dimensional simple modules 47

(DI.d) υ1 ≥ . . . ≥ υn−1 ≥ |υn |, with either (i) all υ j ∈ Z, or (ii) all υ j ∈ 12 + Z.

For g = spo(2m|2n), a Φ+ 0̄
-dominant integral weight λ ∈ h∗ is called an integer
weight if it satisfies (DI.d-i), and it is called a half-integer weight if it satisfies
(DI.d-ii). The L(λ)’s for integer and half-integer weights λ again are quite different.
Proposition 2.13. Let g = spo(2m|2n) with n ≥ 2. Let λ = ∑m i=1 λi δi + ∑ j=1 υ j ε j
n

be a half-integer Φ0̄ -dominant integral weight. Then the highest weight g-module
+

L(λ) is finite dimensional if and only if λm ≥ n.

Proof. Take a half-integer Φ+ 0̄


-dominant integral weight λ with λm ≥ n. We com-
pute that ρ1 = n ∑i=1 δi , and so µ := λ − ρ1 is a Φ+
m

-dominant integral weight,
by Lemma 2.12. By virtue of the classical Weyl character formula, the virtual
character of finite-dimensional g-modules (2.3) in Proposition 2.5 can be rewritten
as chL0 (µ) ∏β∈Φ+ (eβ/2 + e−β/2 ). The highest weight in this expression is clearly

µ + ρ1 = λ, and we conclude that L(λ) is finite dimensional.
Now let λ be a half-integer Φ+

-dominant integral weight such that L(λ) is finite
dimensional. We apply exactly the same sequence of mn odd reflections as in the
proof of Proposition 2.9 now to the standard fundamental system for spo(2m|2n)
in 1.3.4. The resulting fundamental system is given by

⃝ ⃝ ··· ⃝ ··· ⃝⇐=⃝
ε1 − ε2 ε2 − ε3 εn − δ1 δ1 − δ2 δm−1 − δm 2δm

Bearing in mind that λ is half-integer, as in the proof of Proposition 2.9 we can


show that the highest weight with respect to this new Borel subalgebra is λ −
i=1 δi + m ∑ j=1 ε j . Since this weight must take non-negative integer value at
n ∑m n

h2δm to ensure that L(λ) is finite dimensional, we conclude that λm ≥ n. 


Theorem 2.14. Let g = spo(2m|2n) with n ≥ 2. Let λ = ∑m
i=1 λi δi + ∑ j=1 υ j ε j be
n

a Φ+

-dominant integral weight.
(1) Suppose that λ is a half-integer weight. Then L(λ) is finite dimensional if
and only if λm ≥ n.
(2) Suppose that λ is an integer weight. Then L(λ) is finite dimensional if
and only if λ is of the form µ♮ or µ♮− , where µ is an (m|n)-hook partition.

Proof. Part (1) is Proposition 2.13. Since the proof for (2) is very similar to the
proof for Theorem 2.11(2), we shall only give a sketch below.
We recall from 1.3.4 that the standard Dynkin diagram of g = spo(2m|2n)
possesses a non-trivial diagram involution, which induces an involution of the Lie
superalgebra g. Twisting the module structure with this involution interchanges the
irreducible modules L(µ♮ ) and L(µ♮− ).
48 2. Finite-dimensional modules

Given an (m|n)-hook partition µ, we choose M ≥ ℓ(µ) and consider the Lie


superalgebra spo(2M|2n). Then, arguing as in the proof of Theorem 2.11, the
irreducible spo(2M|2n)-module of highest weight µ is finite dimensional. We apply
the same sequence of odd reflections to change the standard Borel of spo(2M|2n)
to a Borel that is compatible with the standard Borel of spo(2m|2n). Again, the
subdiagram consisting of the first (m + n − 1) vertices of the new Dynkin diagram
is of type gl(m|n), and it is easy to see that the new highest weight equals µ♮ . By
restriction to spo(2m|2n), we see that L(µ♮ ) is finite dimensional, and so L(µ♮− ) is
also finite dimensional by applying the Dynkin diagram involution.
Conversely, let µ = µ1 δ1 + . . . + µm δm + ν1 ε1 + . . . + νn εn be an integer Φ+ 0̄
-
dominant integral weight. By Lemma 2.12, (µ1 , . . . , µm ) and (ν1 , . . . , |νn |) are par-
titions. To prove that ν′1 ≤ µm it suffices to prove it for m = 1. Using the sequence
of odd reflections with respect to δ1 − ε1 , δ1 − ε2 , . . . , δ1 − εn , we change the stan-
dard Borel to a Borel with an even simple root 2δ1 . Now the same argument as in
the proof of Theorem 2.11(2) shows that if µ does not satisfy the hook condition,
then µ must take a negative value at h2δ1 , contradicting the finite dimensionality of
L(µ). 

2.1.6. Finite-dimensional simple modules of q(n). Now suppose that g is the


queer Lie superalgebra q(n). Let h be the standard Cartan subalgebra of block
diagonal matrices and let b = h ⊕ n+ be the standard Borel subalgebra of block
upper triangular matrices, cf. Section 1.2.6. Recall that {Hi = Ei,i + Eii | 1 ≤ i ≤ n}
forms the standard basis for h0̄ , with dual basis {εi | 1 ≤ i ≤ n} for h∗0̄ . We have a
triangular decomposition g = n− ⊕ h ⊕ n+ .
Recall from Lemma 1.38 that Wλ is a finite dimensional irreducible h-module,
for λ ∈ h∗0 . Regarding Wλ as a b-module by letting n+ act trivially, we define the
Verma module of g to be

∆(λ) = IndgbWλ .

The highest weight space of ∆(λ) can be naturally identified with Wλ of dimension
2(ℓ(λ)+δ(λ))/2 , and any nonzero vector in the highest weight space Wλ generates the
whole g-module ∆(λ). Following standard arguments for semisimple Lie algebras,
any g-submodule of ∆(λ) has a weight space decomposition, and then the Verma
module ∆(λ) contains a unique maximal proper submodule J(λ) (which is the sum
of all proper submodules). Hence, ∆(λ) has a unique irreducible quotient g-module
L(λ) = ∆(λ)/J(λ).
Assume that the g-module L(λ) is finite dimensional, and write λ = ∑ni=1 λi εi .
Since g0̄ ∼
= gl(n), the g0̄ -dominance integral condition imposes a necessary condi-
tion on λ: λi − λi+1 ∈ Z+ , for all 1 ≤ i ≤ n − 1. However, this condition on λ is not
sufficient for L(λ) to be finite dimensional, as we see in the following lemma.
2.1. Classification of finite-dimensional simple modules 49

Lemma 2.15. Let g = q(2) and λ = mε1 + mε2 ∈ h∗ with m ̸= 0. Then L(λ) is
infinite dimensional.

Proof. Recall the notations Eei j , E i j etc. from 1.1.4 and 1.2.6. Then Ee12 , Ee21 and
H1 − H2 = Ee11 − Ee22 form a standard sl(2)-triple s. Take a highest weight vector
vλ in L(λ). For λ = mε1 + mε2 , we have

Ee12 vλ = 0 = (H1 − H2 )vλ .

To prove the lemma, it suffices to show that Ee21 vλ ̸= 0 in L(λ). Indeed, by the
sl(2) representation theory, Ee21 vλ ̸= 0 implies that the s-submodule generated by
vλ must be the infinite-dimensional Verma module of zero highest weight.
To that end, recalling H i = Eii + Eii , we compute that

E 12 Ee21 vλ = Ee21 E 12 vλ + (H 1 − H 2 )vλ


= (H 1 − H 2 )vλ ,

and (H 1 − H 2 )2 vλ = (Ee11 + Ee22 )vλ = 2mvλ ̸= 0. Hence, we have (H 1 − H 2 )vλ ̸= 0


and Ee21 vλ ̸= 0. This proves the lemma. 

Theorem 2.16. The highest weight irreducible q(n)-module L(λ) is finite dimen-
sional if and only if λ = ∑ni=1 λi εi satisfies the conditions (i)-(ii):
(i) λi − λi+1 ∈ Z+ ; (ii) λi = λi+1 implies that λi = 0, for 1 ≤ i ≤ n − 1.

Proof. For L(λ) to be finite dimensional, λ = ∑ni=1 λi εi must satisfy the usual dom-
inance integral condition for q(n)0̄ ∼
= gl(n), which is Condition (i); λ also must
satisfy Condition (ii) by applying Lemma 2.15 to various standard subalgebras of
q(n) which are isomorphic to q(2). This proves the “only if” part.
Assume that λ satisfies the conditions (i)-(ii), and assume in addition that λi ̸∈
Z. This implies that λi ̸= 0, for all i, and λ j > λ j+1 , for j = 1, . . . , n − 1. We want
to prove that L(λ) is finite dimensional. Write gε = n− +
ε + hε + nε for the standard
triangular decomposition of gε , for ε ∈ Z2 . The Verma g-module ∆(λ), regarded as
a g0̄ -module, has a g0̄ -Verma flag, and we have
( )
ch∆(λ) = 2⌊(n+1)/2⌋ ch Indh0̄ +n+ Cλ ⊗ ∧(n−
g

) ,
0̄ 0̄

where ⌊r⌋ for a real number r denotes the largest integer no greater than r. As in the
proof of Proposition 2.5 we apply the Bernstein functor L0 and its derived functor
Li (with g0̄ = gl(n)) to ∆(λ) and compute the corresponding Euler characteristic
50 2. Finite-dimensional modules

(recall Φ1̄ and Φ0̄ are identical copies of the root system for gl(n)):
( )
∑ ⌊(n+1)/2⌋
∑ −
g0̄
(−1) i
chL i (∆(λ)) =2 (−1)i
chL i Ind C
h +n+ λ
⊗ ∧(n1̄
)
0̄ 0̄
i≥0 i≥0
( )
∑w∈W (−1)ℓ(w) w eλ+ρ0̄ ∏α∈Φ+ (1 + e−α )
=2⌊(n+1)/2⌋ 1̄

∏α∈Φ+ (eα/2 − e−α/2 )


∏α∈Φ+ (1 + e−α )
=2 ⌊(n+1)/2⌋ 1̄

∏α∈Φ+ (1 − e−α ) w∈W


∑ (−1)ℓ(w) ew(λ) .

Since the highest weight appearing in this last expression is clearly λ and the Euler
characteristic is a virtual character of finite-dimensional g-modules (see the proof
of Proposition 2.5), L(λ) is a finite-dimensional module.
Now suppose that λ satisfies the conditions (i)-(ii) of the theorem and in ad-
dition that λi ∈ Z for each i. We write λ = (λ1 , . . . , λk , 0, . . . , 0, λl , . . . , λn ) such
that λ1 > · · · > λk > 0 > λl > · · · > λn . Denote λ+ = (λ1 , . . . , λk , 0, . . . , 0) and
λ− = (0, . . . , 0, λl , . . . , λn ). By Theorem 3.46 on Sergeev duality in Chapter 3, L(µ)
is finite dimensional for µ = ∑i µi εi corresponding to a strict partition (µ1 , . . . , µn ).
Then, the dual module L(µ)∗ ∼ = L(µ∗ ) is finite dimensional, where we have denoted

µ = − ∑i=1 µn+1−i εi . It follows that both L(λ+ ) and L(λ− ) are finite dimensional.
n

Hence, as a quotient of the finite-dimensional g-module L(λ+ ) ⊗ L(λ− ), L(λ) must


be finite dimensional. 

2.2. Harish-Chandra homomorphism and linkage


In this section, we study the center of the universal enveloping algebra U(g), when
g is a basic Lie superalgebra, in particular of type gl and osp. The image of
the Harish-Chandra homomorphism is described with the help of supersymmet-
ric functions (see Appendix A). Then, we provide a characterization on when two
central characters are equal and hence obtain a linkage principle on composition
factors of a Verma module. The finite-dimensional typical irreducible representa-
tions of g are classified, and a Weyl-type character formula for these modules is
obtained.

2.2.1. Supersymmetrization. Let g be a finite-dimensional Lie superalgebra. De-


note by Z(g) = Z(g)0̄ ⊕ Z(g)1̄ the center of the enveloping superalgebra U(g),
where
Z(g)i = {z ∈ U(g)i | za = (−1)ik za, ∀a ∈ gk for k ∈ Z2 }, i ∈ Z2 .
Recall S(g) denotes the symmetric superalgebra of g. As a straightforward super
generalization of a standard construction for Lie algebras (cf. Carter [Car, Propo-
sition 11.4]), we may define a supersymmetrization map
γ : S(g)−→U(g)
2.2. Harish-Chandra homomorphism and linkage 51

and show that it is a g-module isomorphism. When restricting to the g-invariants,


we have the following.

Proposition 2.17. The supersymmetrization map γ : S(g)→U(g) induces a linear


isomorphism γ : S(g)g → Z(g).

Given a finite-dimensional representation (π,V ) of a Lie superalgebra g we


denote its supercharacter by schV. That is,
schV = ∑ sdim(Vµ )eµ ,
µ∈h∗

where we recall from Section 1.1.1 that sdimW = dimW0̄ − dimW1̄ stands for the
superdimension of a vector superspace W = W0̄ ⊕ W1̄ . If we regard e as the Euler
number, a formal expansion of eµ gives us

1 [ ]
[schV ](h) = ∑ str π(h)i , ∀h ∈ g.
i=0 i!

A straightforward super generalization


[ of the
] arguments for Lie algebras (see Carter
[Car, pp.212-3]) shows that x 7→ str π(x)i defines a g-invariant polynomial func-
tion on g. From this we conclude the following.

Proposition 2.18. Let g be a basic Lie superalgebra with Cartan subalgebra[ h. Let]
(π,V ) be a finite-dimensional g-module. Then the polynomial functions str π(h)i
on h, for i > 0, can be identified with the homogeneous components of the super-
character of π, and they arise as restrictions of g-invariant polynomial functions
on g.

2.2.2. The central characters. In this subsection, we let g be a basic Lie superal-
gebra, and h be its Cartan subalgebra.
Recall that g is equipped with an even non-degenerate invariant supersymmet-
ric bilinear form (see Section 1.2.2). Furthermore, the bilinear form (·, ·) on g
allows us to identify g with its dual g∗ , and h with h∗ so that we have correspond-
ing isomorphisms S(g)g ∼ = S(g∗ )g and S(h) ∼= S(h∗ ). Via such identifications, the
∗ ∗
restriction map S(g ) → S(h ) and the induced map η : S(g) → S(h) are homomor-
phisms of algebras.
Let g = n− ⊕h⊕n+ be a triangular decomposition. By the PBW Theorem 1.32,
we have U(g) = U(h) ⊕ (n−U(g) +U(g)n+ ), and we denote by ϕ : U(g) → U(h)
the projection associated to this direct sum decomposition. Since [h, n± ] ⊆ n± ,
by considering ad h(z) = 0 for any element z ∈ Z(g), we conclude that z affords a
unique expression of the form
(2.5) z = hz + ∑ n− +
i hi ni , for hz , hi ∈ U(h), n± ± ±
i ∈ n U(n ).
i
52 2. Finite-dimensional modules

Hence Z(g) consists of only even elements. The restriction of ϕ to Z(g), denoted
again by ϕ, defines an algebra homomorphism from Z(g) to U(h), which sends z
to hz .
As usual, we freely identify S(h) with the algebra of polynomial functions on
h∗ . Extending the standard pairing ⟨·, ·⟩ : h∗ × h → C, we can make sense of ⟨λ, p⟩
for λ ∈ h∗ and p ∈ S(h) by regarding it as the value at λ of the polynomial function
p on h∗ . For λ ∈ h∗ , define a linear map χλ : Z(g) → C by
χλ (z) = ⟨λ, ϕ(z)⟩.
Lemma 2.19. Let g be a basic Lie superalgebra. An element z ∈ Z(g) acts as the
scalar χλ (z) on any highest weight g-module V (λ) of highest weight λ.

Proof. Our proof here is phrased so that it makes sense for q(n) later on as well.
Let z ∈ Z(g). The highest weight space V (λ)λ is an irreducible h-module,
which is one-dimensional for basic g (but may fail to be one-dimensional in the
q(n) case). The even element z commutes with h, and hence by Schur’s Lemma, z
acts as a scalar ξ(z) on V (λ)λ . Now let vλ be a highest weight vector of V (λ), and
write an arbitrary vector u ∈ V (λ) as u = xvλ for x ∈ U(h + n− ). Since z is central,
we have zu = xzvλ = ξ(z)u. It follows by (2.5) that
( )
zvλ = hz + ∑ n− i hi ni vλ = ⟨λ, hz ⟩vλ = χλ (z)vλ .
+
i
Hence ξ(z) = χλ (z). The lemma is proved. 

Therefore χλ : Z(g) → C defines a one-dimensional representation of Z(g),


which will be called the central character associated to a highest weight λ.

2.2.3. Harish-Chandra homomorphism for basic Lie superalgebras. In this


subsection, we assume that g is a basic Lie superalgebra.
Let Φ+ be a positive system in the root system Φ for g. Recall the Weyl vector
ρ from (1.28). Regarding S(h) as the algebra of polynomial functions on h∗ , we
define an automorphism τ of the algebra S(h) by
τ( f )(λ) := f (λ − ρ), ∀ f ∈ S(h).
We then define the Harish-Chandra homomorphism to be the composition
hc = τϕ : Z(g) −→ S(h).
The non-degenerate invariant bilinear form (·, ·) on h induces a W -equivariant alge-
bra isomorphism θ : S(h) → S(h∗ ), and hence an isomorphism θ : S(h)W → S(h∗ )W .
By composition we obtain an algebra homomorphism hc∗ = θhc : Z(g) → S(h∗ ).
Proposition 2.20. Let g be a basic Lie superalgebra. We have
(1) χλ = χw(λ+ρ)−ρ , for all w ∈ W and λ ∈ h∗ .
(2) hc(Z(g)) ⊆ S(h)W , and hc∗ (Z(g)) ⊆ S(h∗ )W .
2.2. Harish-Chandra homomorphism and linkage 53

Proof. The character of the Verma module ∆(µ) of highest weight µ ∈ h∗ is given
by ch∆(µ) = eµ+ρ D−1 , where
∏β∈Φ+ (eβ/2 − e−β/2 )

D := .
∏ α∈Φ+ (eα/2 + e−α/2 )

Since the character of a module equals the sum of the characters of its compo-
sition factors, we have
(2.6) ch∆(λ) = ∑ bµλ chL(µ),
µ

where bµλ ∈ Z+ and bλλ = 1. Since ∆(λ) is a highest weight module, the µ’s in the
summation need to satisfy λ − µ ∈ ∑α∈Φ+ Z+ α and also χλ = χµ by Lemma 2.19.
We choose a total order ≥ on the set λ − ∑α Z+ α with the property that ν ≥ µ
if ν − µ ∈ ∑α∈Φ+ Z+ α. Then the ordered sets {ch∆(µ)}µ,≥ and {chL(µ)}µ,≥ are
two ordered bases for the same space. Now (2.6) says that the matrix expressing
{ch∆(µ)}µ,≥ in terms of {chL(µ)}µ,≥ is upper triangular with 1 along the diagonal,
and hence we have
chL(λ) = ∑ aµλ ch∆(µ),
µ

where aµλ ∈ Z with aλλ = 1, and


(2.7) aµλ = 0 unless λ − µ ∈ ∑ Z+ α and χλ = χµ .
α∈Φ+
Thus
(2.8) DchL(λ) = ∑ aµλ eµ+ρ .
µ

Assume for now that λ ∈ h∗ is chosen so that the irreducible g-module L(λ)
is finite dimensional. Thus L(λ) is a semisimple g0̄ -module, and so chL(λ) is W -
invariant. On the other hand, D is W -anti-invariant by Theorem 1.15(4). Thus the
right-hand side of (2.8) is W -anti-invariant, and hence can be written as
(2.9) ∑ aµλ ∑ (−1)ℓ(w) ew(µ+ρ) ,
µ∈X w∈W

where X consists of Φ+

-dominant integral weights such that aµλ ̸= 0. We compute
that aw(λ+ρ)−ρ,λ = ±aλλ = ±1, and hence by (2.7) we have χλ = χw(λ+ρ)−ρ , for all
w ∈ W.
Now (1) as a polynomial identity on λ ∈ h∗ follows from the following claim.
Claim. S := {λ ∈ h∗ | L(λ) is finite dimensional} is Zariski dense in h∗ .
Indeed, given a Φ+ 0̄
-dominant integral weight µ ∈ h∗ , the induced module
g 0
Indg0̄ L (µ) is finite dimensional and has a highest weight µ+2ρ1 . It is clear that the
set of such weights µ + 2ρ1 associated to all Φ+ 0̄
-dominant integral µ ∈ h∗ (which

is a subset of S) is Zariski dense in h . Hence the claim follows.
54 2. Finite-dimensional modules

By (1), we have that ⟨λ + ρ, hc(z)⟩ = ⟨w(λ + ρ), hc(z)⟩, for all w ∈ W and z ∈
Z(g). Thus hc(z) ∈ S(h)W , and equivalently, hc∗ (Z(g)) ⊆ S(h∗ )W by definition of
hc∗ . 

By Proposition 2.20, we have


hc : Z(g) −→ S(h)W , hc∗ = θhc : Z(g) −→ S(h∗ )W ,
and we further define a homomorphism
η∗ = θη : S(g)g −→ S(h∗ ).

In the remainder of this section, we shall strengthen Proposition 2.20 for g of


types gl and osp by describing hc∗ (Z(g)) precisely.

2.2.4. Invariant polynomials for gl and osp. Suppose that g is of type gl or osp.
The goal of this subsection is to establish Proposition 2.21 which partially describes
the image Im(η∗ ) via supersymmetric polynomials, see Appendix A.2.1 (The word
“partially” can and will be removed subsequently). The description of this image
plays an important role in the study of the Harsih-Chandra map later on.
Associated to a positive system Φ+ , we have the subset Φ̄+

of positive isotropic
odd roots defined in (1.27). Introduce the following element
(2.10) P := ∏ α ∈ S(h∗ ).
α∈Φ̄+

Invariant polynomials for g = gl(m|n). Recall the Weyl group W for g =


gl(m|n), which is by definition the same as for its even subalgebra gl(m) ⊕ gl(n),
is W = Sm × Sn . Applying Proposition 2.18 to the natural representation (ρ,V ) of
g = gl(m|n), we see that
m n
(2.11) σkm,n := ∑ δki − ∑ εkj ∈ Im(η∗ ), ∀k ∈ Z+ .
i=1 j=1

Here and further, we shall identify S(h∗ ) = C[δ, ε], where we set
δ = {δ1 , . . . , δm }, ε = {ε, . . . , εn }.
Hence an element of S(h∗ )W is precisely a polynomial which is symmetric in δ and
symmetric in ε. Recalling from A.2.1 the definition of supersymmetric polynomi-
als, we introduce the following subalgebra of S(h∗ )W :
(2.12) S(h∗ )W
sup = {supersymmetric polynomials in variables δ, ε}.

By (A.31), the polynomials {σkm,n | k ∈ Z+ } generate the algebra S(h∗ )W


sup . Thus
we have S(h∗ )W
sup ⊆ Im(η ∗ ).
2.2. Harish-Chandra homomorphism and linkage 55

The element P defined in (2.10), up to a possible sign which depends on the


choice of positive systems for gl(m|n), can be computed to be
P= ∏ (δi − ε j ).
1≤i≤m,1≤ j≤n

It follows by (A.28) that f P ∈ S(h∗ )W ∗ W


sup for any f ∈ S(h ) .

Invariant polynomials for g = spo(2m|2n + 1). Recall the Weyl group W for
g = spo(2m|2n + 1), which is by definition the same as for sp(2m) ⊕ so(2n + 1), is

2 o Sm ) × (Z2 o Sn ),
WB := (Zm n
(2.13)
and hence an element f ∈ S(h∗ )W = C[δ, ε]W is exactly a polynomial which is
symmetric among δ2i ’s and symmetric among ε2j ’s. We introduce the following
subalgebra of S(h∗ )W :
(2.14) S(h∗ )W
sup = {supersymmetric polynomials in δi , ε j , 1 ≤ i ≤ m, 1 ≤ j ≤ n}.
2 2

Applying Proposition 2.18 to the natural representation of spo(2m|2n + 1), we


have that
m n
m,n = ∑ δi − ∑ ε j ∈ Im(η∗ ),
σ2k ∀k ∈ Z+ .
2k 2k
i=1 j=1

Hence we have S(h∗ )W ∗ W


sup ⊆ Im(η∗ ), since σm,n for k ≥ 0 generate S(h )sup by (A.31).
2k

In this case, we calculate that Φ̄+



= {δi ± ε j | 1 ≤ i ≤ m, 1 ≤ j ≤ n} for the
standard positive system Φ+ , and hence, P in (2.10) is given by
P= ∏ (δi − ε j )(δi + ε j ) = ∏ (δ2i − ε2j ).
1≤i≤m,1≤ j≤n 1≤i≤m,1≤ j≤n

(For other positive systems, P might differ from the above formula by an irrelevant
sign.) By (A.28), we have that f P ∈ S(h∗ )W ∗ W
sup for any f ∈ S(h ) .

Invariant polynomials for g = spo(2m|2n). We regard g = spo(2m|2n) as a


subalgebra of spo(2m|2n + 1) with the same Cartan subalgebra h. We continue
to denote by WB the Weyl group of spo(2m|2n + 1), and then identify the Weyl
group W of g = spo(2m|2n) with a subgroup of WB of index 2. By the same con-
sideration as for g = spo(2m|2n + 1) above using Proposition 2.18, we conclude
that the image Im(η∗ ) contains S(h∗ )W sup , which consists of all the supersymmetric
B

polynomials in δi , ε j , 1 ≤ i ≤ m, 1 ≤ j ≤ n.
2 2

Noting now that Φ+1̄


= {δi ±ε j | 1 ≤ i ≤ m, 1 ≤ j ≤ n} associated to the standard
positive system Φ+ , we set
D1̄ : = ∏ (eα/2 − e−α/2 )
α∈Φ+

= ∏(eδi /2−ε j /2 − e−δi /2+ε j /2 )(eδi /2+ε j /2 − e−δi /2−ε j /2 ).


i, j
56 2. Finite-dimensional modules

For a homogeneous polynomial g ∈ Z[δ, ε]WB , the expression


n
Q := ∏ (eε j − e−ε j ) · g(eδ1 − e−δ1 , . . . , eδm − e−δm ; eε1 − e−ε1 , . . . , eεn − e−εn )
j=1

has ε1 ε2 · · · εn g(δ1 , . . . , δm ; ε1 , . . . , εn ) as its homogeneous term of lowest degree.


Also observe that D1̄ has P as its lowest degree term, where the element P in (2.10)
associated to the standard positive system Φ+ of spo(2m|2n) is given by
P= ∏ (δ2i − ε2j ).
1≤i≤m,1≤ j≤n

(For other positive systems, P might differ by a sign from the above formula.)
Since Q is an integral polynomial in e±δi and e±ε j that is W -invariant, Q is actually
a virtual g0̄ -character. By Remark 2.6, D1̄ Q is a virtual supercharacter of finite-
dimensional g-modules. Hence every homogeneous term of D1̄ Q lies in Im(η∗ ) by
Proposition 2.18, and in particular so does the lowest degree term of D1̄ Q, i.e.,
ε1 ε2 · · · εn Pg ∈ Im(η∗ ).

Introduce the following subspace S(h∗ )W


sup (which can be easily shown to be a
subalgebra) of S(h∗ )W = C[δ, ε]W :

(2.15) S(h∗ )W ∗ WB
sup = S(h )sup C[δ, ε]WB ε1 ε2 · · · εn P.
Summarizing the above discussions, we have shown that S(h∗ )W
sup ⊆ Im(η∗ ).
It is a standard fact about the classical Weyl group W that any element f ∈
S(h∗ )W = C[δ, ε]W can be written as
f = h + ε1 ε2 · · · εn g,
where h, g ∈ C[δ, ε]WB . Thus, we have hP ∈ S(h∗ )W ∗ W
sup and ε1 ε2 · · · εn gP ∈ S(h )sup ,
and hence, f P ∈ S(h∗ )Wsup .
Summarizing all the three cases, we have established the following.
Proposition 2.21. Let g be either gl(m|n), spo(2m|2n + 1), or spo(2m|2n), and let
W be its Weyl group. Then, we have
(1) P ∈ S(h∗ )W .
(2) f P ∈ S(h∗ )W ∗ W
sup , for all f ∈ S(h ) .
(3) S(h∗ )W
sup ⊆ Im(η∗ ).

2.2.5. Image of Harish-Chandra homomorphism for gl and osp. Let g be a


basic Lie superalgebra for now. We have a non-degenerate invariant bilinear form
(·, ·) on g, h, and h∗ . We recall here that (α, α) = 0, for any isotropic odd root α.
The non-degenerate form (·, ·) induces an isomorphism h ∼ = h∗ .
Lemma 2.22. Let g be a basic Lie superalgebra. Let α be an isotropic odd root
such that (λ + ρ, α) = 0. Then we have χλ = χλ+tα , for all t ∈ C.
2.2. Harish-Chandra homomorphism and linkage 57

Proof. Fix a positive system Φ+ . We may assume that the isotropic odd root α ∈
Φ+ (otherwise, use −α to replace α).
First suppose that α is an isotropic simple root in Φ+ . Then (ρ, α) = 12 (α, α) =
0, and hence (λ, α) = 0 by assumption. Let vλ be a highest weight vector in the
Verma module ∆(λ). Then,
eα fα vλ = [eα , fα ]vλ = hα vλ = (λ, α)vλ = 0.
Also, we have eγ fα vλ = 0 for other simple roots γ by weight considerations. Hence,
fα vλ is a singular vector in ∆(λ). This gives rise to a non-trivial g-module homo-
morphism ∆(λ − α) → ∆(λ), whence χλ−α = χλ .
Now suppose α is any isotropic root in Φ+ . By Lemma 1.25, there exists
w ∈ W such that β = w(α) is simple. Then (α, α) = (β, β) = 0 by the W -invariance
of (·, ·). By Proposition 2.20, we have
χλ−α = χw(λ−α+ρ)−ρ = χµ ,
where we have denoted µ := w(λ + ρ) − ρ − β and used w(α) = β. Now by the
W -invariance of (·, ·) again, we compute that
(µ + ρ, β) = (w(λ + ρ), w(α)) − (β, β) = (λ + ρ, α) − 0 = 0,
Thus, we can apply the special case established in the preceding paragraph to the
weight µ and simple root β to obtain that
χλ−α = χµ = χµ+β = χw(λ+ρ)−ρ = χλ ,
where we have used Proposition 2.20 again in the last equality. This implies that
χλ = χλ−tα , for any t ∈ Z+ and any isotropic odd root α with (λ + ρ, α) = 0. Since
Z+ is Zariski dense in C, we conclude that the polynomial identity χλ = χλ+tα
holds for all t ∈ C. 

Now we specialize to the type gl and osp.


Proposition 2.23. Let g be either gl(m|n), spo(2m|2n + 1), or spo(2m|2n). Then
we have hc∗ (Z(g)) ⊆ S(h∗ )W ∗ W
sup , where S(h )sup is given in (2.12), (2.14), and (2.15).

Proof. Let us proceed case-by-case.


First let us consider the case of g = gl(m|n). Assume that λ + ρ = ∑m i=1 ai δi +
∑nj=1 b j ε j
satisfies (λ + ρ, α) = 0 for some isotropic root α = δi − ε j , which means
that ai = −b j . Let z ∈ Z(g). Then by Lemma 2.22 we have
⟨λ + ρ, hc(z)⟩ = χλ (z) = χλ−tα (z) = ⟨λ − tα + ρ, hc(z)⟩, ∀t ∈ C.
This implies that the specialization hc∗ (z)|δi =−ε j =t is independent of t, where we
recall hc∗ = θhc. In addition, we have hc∗ (z) ∈ S(h∗ )W by Proposition 2.20, whence
hc∗ (z) ∈ S(h∗ )W
sup by the definition (2.12).
The case for g = spo(2m|2n + 1) is analogous. Let λ ∈ h∗ be such that (λ +
ρ, α) = 0 for some isotropic root α = δi ± ε j , we show by the same argument as
58 2. Finite-dimensional modules

above that the specialization hc∗ (z)|δi =∓ε j =t is independent of t. In addition, it


follows by Proposition 2.20 that hc∗ (z) ∈ S(h∗ )W for W = WB (see (2.13)), which
means that hc∗ (z) is symmetric in δ2i for all i and symmetric in ε2j for all j. Then
we conclude that hc∗ (z) ∈ S(h∗ )W
sup by the definition (2.14).
Now suppose that g = spo(2m|2n). Let W be the Weyl group of g, which is
regarded as a subgroup of WB of index 2 as usual. Given z ∈ Z(g), we can write
hc∗ (z) ∈ S(h∗ )W as
hc∗ (z) = f + ε1 · · · εn g, for f , g ∈ S(h∗ )WB = C[δ2 , ε2 ]Sm ×Sn .
Since χλ = χλ−tα , for any λ and any isotropic root α = δi − ε j with (λ + ρ, α) = 0,
the specialization hc∗ (z)|δi =ε j =t = f1 + g1 is independent of t, where we denote
f1 := f |δi =ε j =t , g1 := tε1 · · · ε j−1 ε j+1 · · · εn g|δi =ε j =t .
As a polynomial of t, f1 has even degree while g1 has odd degree. Thus both f1
and g1 must be independent of t. Hence, recalling (2.14), we have that
f ∈ S(h∗ )W
sup ,
B
g|δi =ε j =t = 0.
Since the polynomial g is symmetric in δ21 , . . . , δ2m and symmetric in ε21 , . . . , ε2n , we
conclude that g is divisible by (δ2i − ε2j ), for all i and j, and hence divisible by
P = ∏i, j (δ2i − ε2j ). Thus we have g = g0 P for some g0 ∈ S(h∗ )WB .
The inclusion hc∗ (Z(g)) ⊆ S(h∗ )W
sup now follows by the definition (2.15). 
Theorem 2.24. Let g = gl(m|n), spo(2m|2n + 1), or spo(2m|2n), and let W be its
Weyl group. Then we have
hc∗ (Z(g)) = Im(η∗ ) = S(h∗ )W
sup ,

where S(h∗ )W
sup is defined in (2.12), (2.14), and (2.15), respectively.

Proof. By Propositions 2.21 and 2.23, we have a map hc∗ : Z(g) → Im(η∗ ), and
that hc∗ (Z(g)) ⊆ S(h∗ )W
sup ⊆ Im(η∗ ). So it suffices to show that hc∗ (Z(g)) = Im(η∗ )
in order to complete the proof. The commutative diagram (where gr denotes the
associated graded)
gr γ−1
(2.16) gr Z(g) / S(g)g
H ∼
HH=
HH
HH η∗
HH
gr hc∗ $ 
S(h∗ )
implies that gr hc∗ : gr Z(g) → Im(η∗ ) is surjective.
Now the surjectivity of the map hc∗ : Z(g) → Im(η∗ ) follows by a standard
filtered algebra argument. Indeed, consider hc∗ (Z(g)) and Im(η∗ ) as filtered spaces:
C= ∼ hc (Z(g))0 ⊆ hc (Z(g))1 ⊆ · · · , C∼= Im(η∗ )0 ⊆ Im(η∗ )1 ⊆ · · · .
∗ ∗
2.2. Harish-Chandra homomorphism and linkage 59

We shall prove by induction on k that hc∗ (Z(g))k = Im(η∗ )k , for all k. Clearly we
have hc∗ (Z(g))0 = Im(η∗ )0 . Now assume that hc∗ (Z(g))k−1 = Im(η∗ )k−1 . Let us
take a = η∗ (x) ∈ Im(η∗ )k for x ∈ S(g)g . Then, a = gr (hc∗ γ)(x), and

hc∗ γ(x) − a ∈ Im(η∗ )k−1 = hc∗ (Z(g))k−1 .

Therefore, we conclude that a ∈ hc∗ (Z(g))k . 

Define the subalgebra S(h)W ∗ W


sup of S(h) to be the preimage of S(h )sup under

the isomorphism θ : S(h) → S(h ). Various maps in this section can now be put
together in the following diagram, which only becomes commutative when we pass
from Z(g) to its associated graded (note that the other algebras are already graded).

γ−1
(2.17) Z(g) / S(g)g

= t
tt
tt η
hc∗
ttt η∗
 zt ∼ 
S(h∗ )W o
=
sup S(h)W sup
θ

Remark 2.25. The surjective homomorphism hc : Z(g) → S(h)W sup is actually injec-
tive as well, and hence all the homomorphisms in the diagram (2.17) are isomor-
phisms. Actually Sergeev [Sv3, Corollary 1.1] showed that η : S(g)g → S(h) is
injective, and the injectivity of hc follows from this by a filtered algebra argument.
Except for its intrinsic value, the injectivity does not seem to play any crucial role
in the representation theory of g.

The following corollary is immediate from Proposition 2.21 and Theorem 2.24,
and it will be used later on.

Corollary 2.26. Let g be one of the Lie superalgebras gl(m|n), spo(2m|2n + 1), or
spo(2m|2n). Then, for any f ∈ S(h)W , we have f P ∈ hc∗ (Z(g)).

2.2.6. Linkage for gl and osp. In this subsection, we let g denote the Lie su-
peralgebra gl(m|n), spo(2m|2n), or spo(2m|2n + 1). Using the description of the
images of Harish-Chandra homomorphisms, we obtain a necessary and sufficient
condition for two central characters of g to be equal.
Let h be its standard Cartan subalgebra of diagonal matrices. Let {δi , ε j |1 ≤
i ≤ m, 1 ≤ j ≤ n} be the standard basis of h∗ . Let Φ+ be the standard positive
system of the root system Φ for g. The following lemma is proved by a direct
computation.
We define a relation ∼ on h∗ by declaring

λ ∼ µ, for λ, µ ∈ h∗ ,
60 2. Finite-dimensional modules

if there exist mutually orthogonal isotropic odd roots α1 , α2 , . . . , αℓ , complex num-


bers c1 , c2 , . . . , cℓ , and an element w ∈ W satisfying that
( ℓ ) ( )
(2.18) µ + ρ = w λ + ρ − ∑ ca αa , λ + ρ, α j = 0, j = 1, . . . , ℓ.
a=1
The weights λ and µ are said to be linked if λ ∼ µ. It follows from Theorem 2.27
below that linkage is an equivalence relation.
Theorem 2.27. Let g be gl(m|n), spo(2m|2n), or spo(2m|2n + 1), and let h be its
Cartan subalgebra. Let λ, µ ∈ h∗ . Then λ is linked to µ if and only if χλ = χµ .

Proof. Assume first that λ ∼ µ. Since the αi ’s are orthogonal to another, the second
condition in (2.18) implies that we can apply repeatedly Lemma 2.22 to conclude
that
χλ = χλ−c1 α1 = χλ−c1 α1 −c2 α2 = . . . = χλ−∑ℓ ca αa .
a=1

Since µ + ρ = w(λ + ρ − ∑ℓa=1 ca αa ) by (2.18) again, it follows by Proposition 2.20


that χλ−∑ℓ = χµ , whence χλ = χµ .
a=1 ca αa

Now assume that χλ = χµ . Let us write λ = ∑m i=1 λi δi + ∑ j=1 ν j ε j and µ =


n

∑mi=1 µi δi + ∑ j=1 η j ε j . Recall from Theorem 2.24 that for g = gl(m|n) the poly-
n

nomials σkm,n given in (2.11) lie in hc∗ (Z(g)), for all k ∈ Z+ , while in the case of
g = spo(2m|2n) or spo(2m|2n + 1), the polynomials σkm,n lie in hc∗ (Z(g)), for all
even k ∈ Z+ . Let zkm,n be an element in Z(g) with hc∗ (zkm,n ) = σkm,n , where k is
assume to be even for type spo. Then,
m n
χλ (zkm,n ) = ∑ (λ + ρ, δi )k − ∑ (λ + ρ, ε j )k .
i=1 j=1

We now proceed case-by-case.


(1) First, suppose that g = gl(m|n). Then, for all k ≥ 0, we have
χλ (zkm,n ) = χµ (zkm,n ).
This implies that, for all k ≥ 0,
m n m n
∑ (λ + ρ, δi )k − ∑ (λ + ρ, ε j )k = ∑ (µ + ρ, δi )k − ∑ (µ + ρ, ε j )k .
i=1 j=1 i=1 j=1
Using exponential generating functions in a formal indeterminate t, this is equiva-
lent to
m n m n
(2.19) ∑ e(λ+ρ,δi )t − ∑ e(λ+ρ,ε j )t = ∑ e(µ+ρ,δi )t − ∑ e(µ+ρ,ε j )t .
i=1 j=1 i=1 j=1

Pick a maximal number of pairs (ia , ja ) such that (λ + ρ, δia ) = (λ + ρ, ε ja ) and


all ia (respectively, all ja ) are distinct, for 1 ≤ a ≤ ℓ. Similarly, pick a maximal
number of pairs (i′b , jb′ ) such that (µ + ρ, δi′b ) = (µ + ρ, ε jb′ ) and all i′b (respec-
tively, all jb′ ) are distinct, for 1 ≤ b ≤ r. Note that the functions eat are linearly
2.2. Harish-Chandra homomorphism and linkage 61

independent for distinct a. In the identity obtained from (2.19) by canceling the
terms corresponding to all the pairs (ia , ja ) and (i′b , jb′ ), the survived (λ + ρ, δi )’s
must match bijectively with the survived (µ + ρ, δi )’s, while the (λ + ρ, ε j )’s and
(µ + ρ, ε j )’s must match bijectively. So ℓ = r. Now we can find a pair of per-
mutations w = (σ1 , σ2 ) ∈ Sm × Sn which extends the above bijections such that
w−1 (µ + ρ) − λ − ρ is a linear combination of these mutually orthogonal isotropic
odd roots δia − ε ja , and hence λ ∼ µ.
(2) Now suppose that g = spo(2m|2n + 1). Then, for all even k ≥ 0, we have
χλ (zkm,n ) = χµ (zkm,n ).
This is equivalent to the following generating function identity:
m n m n
∑ e(λ+ρ,δ ) t − ∑ e(λ+ρ,ε ) t = ∑ e(µ+ρ,δ ) t − ∑ e(µ+ρ,ε ) t .
2 2 2 2
(2.20) i j i j

i=1 j=1 i=1 j=1

Pick a maximal number of pairs (ia , ja ) such that (λ + ρ, δia ) = sa (λ + ρ, ε ja ) for


some sign sa ∈ {±} and all ia (respectively, all ja ) are distinct, for 1 ≤ a ≤ ℓ. Sim-
ilarly, pick a maximal number of pairs (i′b , jb′ ) such that (µ + ρ, δi′b )2 = (µ + ρ, ε jb′ )2
and all i′b (respectively, all jb′ ) are distinct, for 1 ≤ b ≤ r. In the identity ob-
tained from (2.20) by canceling the terms corresponding to all the pairs (ia , ja )
and (i′b , jb′ ), the survived (λ + ρ, δi )’s must match bijectively with the survived
(µ + ρ, δi )’s up to signs, while the (λ + ρ, ε j )’s and (µ + ρ, ε j )’s must match bi-
jectively up to signs. So ℓ = r. Then, there exists a pair of signed permutations
w = (w1 , w2 ) ∈ W = WB = (Zm 2 o Sm ) × (Z2 o Sn ) which extends the above bijec-
n

tions such that



(2.21) w−1 (µ + ρ) − (λ + ρ) = ∑ ca (δi a − sa ε ja ), ca ∈ C.
a=1

This implies that λ ∼ µ.


(3) Finally, consider the case when g = spo(2m|2n). For all even k ≥ 0, we have
χλ (zkm,n ) = χµ (zkm,n ). This is equivalent again to the generating function identity
(2.20). Following the case of spo(2m|2n + 1), we make a similar choice of pairs
(ia , ja ), for 1 ≤ a ≤ ℓ, and (i′b , jb′ ), for 1 ≤ b ≤ ℓ, which leads to a choice of w =
(w1 , w2 ) ∈ WB and a sequence of isotropic odd roots satisfying (2.21). But we are
not done yet, as the Weyl group W of spo(2m|2n) is not WB , but a subgroup of WB
of index 2. We will finish the job by considering two cases separately depending
on whether or not λ is typical.
Assume first that λ is atypical (and hence so is µ), and we have ℓ ≥ 1. If
w = (w1 , w2 ) ∈ W , we are done. Otherwise, let τ1 ∈ WB be the element changing
the sign for δi1 while fixing all ε j and δi (i ̸= i1 ), and let σ1 ∈ WB be the element
changing the sign for ε j1 while fixing all δi and ε j ( j ̸= j1 ). Note that τ1 ∈ W and
σ1 w−1 ∈ W , and τ21 = σ21 = 1. By the definition of the pair (i1 , j1 ) and using (2.21),
62 2. Finite-dimensional modules

we have
τ1 σ1 w−1 (µ + ρ) − (λ + ρ)
( ) ( )
= τ1 σ1 w−1 (µ + ρ) − (λ + ρ) + τ1 σ1 (λ + ρ) − (λ + ρ)
( ) ℓ
= − 2(λ + ρ, δi1 ) + c1 (δi1 − s1 ε j1 ) + ∑ ca (δia − sa ε ja ).
a=2
Hence we have obtained a set of mutually orthogonal of isotropic odd roots, a
corresponding set of complex numbers, and the Weyl group element wσ1 τ1 ∈ W
satisfying (2.18). Thus, λ ∼ µ.
By the description of the center Z(g) in Theorem 2.24 and (2.15), we have
additional elements in Z(g) to use. In particular, we have
⟨ ⟩ ⟨ ⟩
(2.22) ε1 ε2 · · · εn P, θ−1 (λ + ρ) = ε1 ε2 · · · εn P, θ−1 (µ + ρ) .
Assume
( now that λ is typical ) (and so is µ), and(hence ℓ = 0. Then the (m ) + n)-
tuple (λ + ρ, δi )2 , (λ + ρ, ε j )2 i, j coincides with (µ + ρ, δi )2 , (µ + ρ, ε j )2 i, j , up
to a permutation in Sm × Sn . Recall P = ∏i, j (δ2i − ε2j ). Hence, ⟨P, θ−1 (λ + ρ)⟩ =
⟨P, θ−1 (µ + ρ)⟩ =
̸ 0, since λ and µ are typical. By canceling this nonzero factor in
(2.22), we obtain that
⟨ ⟩ ⟨ ⟩
ε1 ε2 · · · εn , θ−1 (λ + ρ) = ε1 ε2 · · · εn , θ−1 (µ + ρ) .
Note that in this case (2.18) reads that µ + ρ = w(λ + ρ). If (εi , λ + ρ) ̸= 0 for all
i, we conclude that w2 changes the signs for an even number of ε j ’s, and hence
w = (w1 , w2 ) lies in the Weyl group W . If (εi , λ + ρ) = 0 for some i and w ̸∈ W ,
then w′ = (w1 , w2 σ) ∈ W satisfies that µ + ρ = w′ (λ + ρ), where σ denotes the sign
change at εi . 

Theorem 2.27 has the following implications. For a composition factor L(µ) in
a Verma module ∆(λ) of g, µ must satisfy the following conditions: λ − µ ∈ Z+ Φ+
and µ ∼ λ. This will be referred to as the linkage principle for Lie superalgebra g
of type gl and osp.
We fix a positive system Φ+ of the root system Φ for g. Recall the subset
+
Φ1̄ ⊆ Φ+ defined in (1.27).
Definition 2.28. The degree of atypicality of an element λ ∈ h∗ , denoted by #λ, is
+
the maximum number of mutually orthogonal roots α ∈ Φ1̄ such that (λ + ρ, α) =
0. An element λ ∈ h∗ is said to be typical (relative to Φ) if #λ = 0, and is atypical
otherwise.
Corollary 2.29. If χλ = χµ , then the degrees of atypicality for λ and µ coincide.

Proof. This can be read off from the proof of Theorem 2.27. 
Corollary 2.30. A finite-dimensional spo(2m|1)-module is completely reducible.
2.2. Harish-Chandra homomorphism and linkage 63

Proof. Recall that the irreducible spo(2m|1)-module L(λ) is finite dimensional if


and only if the sequence (λ1 , . . . , λm ) associated to the highest weight λ = ∑m
i=1 λi δi
is a partition. Let λ and µ be highest weights for two irreducible finite-dimensional
spo(2m|1)-modules. Then λ and µ are Φ+ 0̄
-dominant integral, and hence λ ∼ µ if
and only if λ = µ. By Theorem 2.27 this implies that χλ = χµ if and only if λ =
µ, and hence every finite-dimensional spo(2m|1)-module is completely reducible.


2.2.7. Typical finite-dimensional irreducible characters. In this subsection we


assume that g is of type gl or osp. As an application of the linkage principle, we
obtain a character formula for the typical finite-dimensional irreducible g-modules.
We fix a positive system Φ+ of g, and denote by g = n− ⊕h ⊕n+ the associated
+
triangular decomposition. Recall the subset Φ1̄ ⊆ Φ+ defined in (1.27). Clearly,
+
λ ∈ h∗ is typical (cf. Definition 2.28) if and only if (λ + ρ, α) ̸= 0, for all α ∈ Φ1̄ .
Alternatively, λ is typical if and only if ⟨λ + ρ, P′ ⟩ ̸= 0, where P′ := θ−1 (P) ∈ S(h)
in terms of P in (2.10). The following is a partial converse of Proposition 2.20,
and follows instantly from Theorem 2.27. Below we shall provide a second proof
based on the weaker Corollary 2.26.
Lemma 2.31. Let g be of type gl or osp. Let λ, µ ∈ h∗ with λ typical. Suppose that
χλ = χµ . Then, there exists w ∈ W such that w(λ + ρ) = µ + ρ, and µ is also typical.

Proof. Suppose that there is no w ∈ W such that w(λ + ρ) = µ + ρ. Then we have


W (λ + ρ) ∩ W (µ + ρ) = 0. / We can choose an element f ∈ S(h) which takes the
value 1 on the finite set W (λ + ρ), and 0 on the finite set W (µ + ρ). Averaging
over W if necessary, we may also assume that f ∈ S(h)W . Now by Corollary 2.26
we have θ( f )P ∈ hc∗ (Z(g)). Thus there exists z ∈ Z(g) so that θ( f )P = hc∗ (z) and
equivalently f P′ = hc(z). Hence,
χλ (z) = ⟨λ + ρ, hc(z)⟩ = ⟨λ + ρ, f P′ ⟩ = ⟨λ + ρ, f ⟩⟨λ + ρ, P′ ⟩ = ⟨λ + ρ, P′ ⟩ ̸= 0.
Similarly, we have
χµ (z) = ⟨µ + ρ, hc(z)⟩ = ⟨µ + ρ, f P′ ⟩ = ⟨µ + ρ, f ⟩⟨µ + ρ, P′ ⟩ = 0.
This contradicts χλ = χµ .
+ +
It follows by definition that Φ1̄ = Φ1̄ ∪ −Φ1̄ is W -invariant. Hence, for α ∈
+
Φ1̄ , we have
(µ + ρ, α) = (w(λ + ρ), α) = (λ + ρ, w−1 α) ̸= 0.
Therefore, µ is typical. 

Recall that the finite-dimensional irreducible g-modules have been classified


in terms of highest weights in Section 2.1.
64 2. Finite-dimensional modules

Theorem 2.32. Let g be of type gl or osp. Let λ ∈ h∗ be a typical weight such that
L(λ) is a finite-dimensional irreducible representation of g. Then
∏α∈Φ+ (1 + e−α )
chL(λ) =

(1 − e−β )
∑ (−1)ℓ(w) ew(λ+ρ)−ρ .
β∈Φ+

w∈W

Proof. Recall from the proof of Proposition 2.20 that


(2.23) DchL(λ) = ∑ aµλ ∑ (−1)ℓ(w) ew(µ+ρ) ,
µ∈X w∈W

where X consists of Φ+

-dominant weights such that aµλ ̸= 0.
Since λ is typical and χλ = χµ for all µ ∈ X, it follows by Lemma 2.31 that
X = λ. Recalling that aλλ = 1, we can rewrite (2.23) as
DchL(λ) = ∑ (−1)ℓ(w) ew(λ+ρ) ,
w∈W

from which the theorem follows. 


Example 2.33. Let g = gl(1|1). Then λ = aδ+bε is typical if and only if a+b ̸= 0.
Example 2.34. The character formula in Theorem 2.32 applies to all irreducible
finite-dimensional spo(2m|1)-modules, as their highest weights are always typical.

2.3. Harish-Chandra homomorphism and linkage for q(n)


This section is the counterpart of Section 2.2 for the queer Lie superalgebra q(n).
The image of the Harish-Chandra homomorphism for q(n) is described with the
help of symmetric functions such as Schur Q-functions (see Appendix A). Then,
we obtain a linkage principle on composition factors of a Verma module of q(n).
The finite-dimensional typical irreducible representations of q(n) are classified,
and a Weyl-type character formula for these modules is obtained.

2.3.1. The central characters for q(n). In this section, let g be the queer Lie
superalgebra q(n), and h = h0̄ + h1̄ be its (nonabelian) Cartan subalgebra. As this
subsection is parallel to that of the basic Lie superalgebra case in Section 2.2.2, we
will be brief and only point out the differences from therein.
Recall that g is equipped with a non-degenerate odd invariant supersymmetric
bilinear form (·, ·), which allows us to identify g with its dual g∗ , h with h∗ , and
h0̄ with h∗1̄ . Via such identifications, the restriction map S(g∗ ) → S(h∗ ) and the
induced map η : S(g) → S(h) are homomorphisms of algebras. Let g = n− ⊕ h ⊕
n+ be a triangular decomposition. As before we have a projection ϕ : U(g) →
U(h), and any element z ∈ Z(g) affords a unique expression of the form (2.5).
Moreover, it follows from ad h(z) = 0 that hz ∈ U(h0̄ ). Hence Z(g) consists of only
even elements. The restriction of ϕ to Z(g), also denoted by ϕ, defines an algebra
2.3. Harish-Chandra homomorphism and linkage for q(n) 65

homomorphism from Z(g) to U(h0̄ ), which sends z to hz . For λ ∈ h∗0̄ , define a linear
map χλ : Z(g) → C by χλ (z) = ⟨λ, ϕ(z)⟩.
The proof of Lemma 2.19 also works for the following lemma.
Lemma 2.35. Let g = q(n). An element z ∈ Z(g) acts as the scalar χλ (z) on any
highest weight g-module V (λ) of highest weight λ.

Therefore χλ : Z(g) → C defines a one-dimensional Z(g)-module, which will


be called the central character of g = q(n) associated to the highest weight λ.

2.3.2. Harish-Chandra homomorphism for q(n). Let g = n− ⊕ h ⊕ n+ be the


standard triangular decomposition. Bearing in mind that ρ = 0 for q(n), we call
hc = ϕ : Z(g) → S(h0̄ )
the Harish-Chandra homomorphism for q(n).
Recall the elements Hi , H i , Eei j , E i j in q(n) etc. from Sections 1.1.4 and 1.2.6.
A singular vector in the Verma module ∆(λ) of q(n) is a nonzero vector v such
that Eei j v = E i j v = 0, for 1 ≤ i < j ≤ n.
Lemma 2.36. Let g = q(n). Let λ ∈ h∗0̄ be such that λi − λi+1 = k ∈ N for some
1 ≤ i ≤ n − 1. Then
u := E i,i+1 Eei+1,i
k+1

is a singular vector in the Verma module ∆(λ) of weight si (λ) = λ − k(εi − εi+1 ).

Proof. We have the following identities in ∆(λ), for λ ∈ h∗0̄ and m ≥ 0:


(2.24) (H i − H i+1 )Eei+1,i
m
vλ = Eei+1,i
m
(H i − H i+1 )vλ − 2mE i+1,i Eei+1,i
m−1
vλ ,
(2.25) E i,i+1 Eei+1,i
m+1
vλ = (m + 1)Eei+1,i
m
(H i − H i+1 )vλ − m(m + 1)E i+1,i Eei+1,i
m−1
vλ .
Indeed, (2.24) can be directly verified by induction on m, and then (2.25) follows
by induction on m using (2.24).
It follows by (2.25) for m = k that u ̸= 0. Now note that
E i,i+1 u = (E i,i+1 )2 Eek+1 vλ = 0,
i+1,i

Eei,i+1 u = E i,i+1 Eei,i+1 Eei+1,i


k+1
vλ = 0,
by a standard sl(2)-calculation. By weight considerations in ∆(λ), we must have
E j, j+1 u = Eej, j+1 u = 0, j ̸= i.
Thus, we conclude that u is a singular vector in ∆(λ), and the weight of u is clearly
equal to si (λ) = λ − k(εi − εi+1 ). 

Recall that the Weyl group W for g = q(n) is the symmetric group Sn .
Proposition 2.37. Let g = q(n). We have χλ = χw(λ) , for all w ∈ W and λ ∈ h∗0̄ .
Also, we have hc(Z(g)) ⊆ S(h0̄ )W .
66 2. Finite-dimensional modules

Proof. By Lemma 2.36, we have a nonzero homomorphism ∆(si (λ)) → ∆(λ), for
λ ∈ h∗0̄ with λi − λi+1 ∈ N, which implies by Lemma 2.19 that χλ = χsi (λ) . Hence,
χλ = χw(λ) , for all w ∈ W and for all integral weights λ = ∑ni=1 λi εi satisfying λ1 >
. . . > λn . Since the set of such weights is Zariski dense in h∗0̄ , it follows that χλ =
χw(λ) , for all w ∈ W and λ ∈ h∗0̄ . That is, ⟨λ, hc(z)⟩ = ⟨w(λ), hc(z)⟩, for all z ∈ Z(g),
w ∈ W and λ ∈ h∗0̄ . Hence, we conclude that hc(Z(g)) ⊆ S(h0̄ )W thanks to the W -
invariance of (·, ·). 

Remark 2.38. One can imitate the proof of Proposition 2.20 to give an alternative
proof of Proposition 2.37, bypassing Lemma 2.36. On the other hand, a second
proof of Proposition 2.20 can be given in the spirit of the proof of Proposition 2.37
with some extra work involving odd reflections, as the Weyl group W is not gener-
ated by the even simple reflections.

The odd non-degenerate invariant bilinear form (·, ·) on h induces an (even)


isomorphism h0̄ ∼= Πh∗1̄ , where Π denotes the parity reversing functor from Sec-
tion 1.1.1. This in turn induces (even) isomorphisms θ : S(h0̄ ) → S(Πh∗1̄ ) and
θ : S(h0̄ )W → S(Πh∗1̄ )W . We further define hc∗ = θhc : Z(g) → S(Πh∗1̄ )W and η∗ =
θη : S(g)g → S(Πh∗1̄ ).
Let ε1 , ε2 , . . . , εn ∈ Πh∗1̄ be determined by

εi (H j ) = δi j .

That is, θ(Hi ) = εi , for each i. Let (π,V ) be the natural representation of g =
q(n) on the vector superspace Cn|n . We recall the odd trace operator otr : g →
C from (1.26). A straightforward super generalization of the arguments ( for
) Lie
algebras as given in Carter [Car, pp.212-213] shows that x 7→ otr π(x)2k−1 , for
k ∈ N, defines
( )a g-invariant homogeneous polynomial on g of degree k; and note
that otr π(x)2k = 0 by definition of the odd trace. Restricting to h1̄ we obtain a
homogeneous polynomial on (Πh1̄ of degree ) 2k −1. Then a direct calculation shows
that, for x ∈ h1̄ and k ≥ 1, otr π(x)2k−1 equals the value at x of the polynomial
n
p2k−1 (ε) = ∑ ε2k−1
i , k ∈ N.
i=1

Recall that a ring Γn of symmetric function in n variables ε1 , ε2 , . . . , εn is de-


fined in Appendix A.3. By (A.50), Γn has a basis which consists of Schur Q-
functions Qλ (ε1 , ε2 , . . . , εn ), for λ ∈ SP such that ℓ(λ) ≤ n (where SP denotes the
set of strict partitions). By (A.43), the algebra Γn,C := C ⊗Z Γn is generated by the
odd degree power sums p2k−1 (ε) for k ≥ 1.
Summarizing the above discussions, we have established the following.

Proposition 2.39. Let g = q(n). We have Γn,C ⊆ Im(η∗ ).


2.3. Harish-Chandra homomorphism and linkage for q(n) 67

We define
(2.26) P := ∏ (εi + ε j ) ∈ S(Πh∗1̄ ).
1≤i< j≤n

The following lemma is a reformulation of (A.56) in Appendix A.


Lemma 2.40. Let f ∈ S(Πh∗1̄ )W . Then we have f P ∈ Im(η∗ ).
Lemma 2.41. Assume that a weight λ ∈ h∗0̄ satisfies that λi = −λi+1 = k ∈ C, for
some 1 ≤ i ≤ n − 1.
(1) If k = 0, then u := Eei+1,i vλ for any highest weight vector vλ is a singular
vector in ∆(λ) of weight λ − εi + εi+1 .
(2) If k ̸= 0, then there exists a highest weight vector vλ in the Verma module
∆(λ) such that (H i − H i+1 )vλ ̸= 0 and (H i + H i+1 )vλ = 0. Moreover,
u := Eei+1,i (H i − H i+1 )vλ − 2kE i+1,i vλ
is a singular vector in ∆(λ) of weight λ − εi + εi+1 .

Proof. (1) Assume that k = 0. Note that H i and H i+1 act trivially on the highest
weight space Wλ of ∆(λ), and hence H i vλ = H i+1 vλ = 0 for any highest weight
vector vλ . Then, Eei,i+1 u = (Hi − Hi+1 )vλ = 0, and E i,i+1 u = (H i − H i+1 )vλ = 0.
Also it follows from weight consideration that Eej, j+1 u = E j, j+1 u = 0 for j ̸= i.
Hence, u is a singular vector.
(2) Assume that k ̸= 0. We again identify the highest weight space of the Verma
module ∆(λ) as the h-module Wλ . We compute that
λ([H i ± H i+1 , H i ± H i+1 ]) = 2λ(Hi + Hi+1 ) = λi + λi+1 = 0.
Hence, by Lemma 1.38, there exists a highest weight vector vλ ∈ Wλ such that
wλ := (H i − H i+1 )vλ ̸= 0 and (H i + H i+1 )vλ = 0. It follows that
(H i + H i+1 )wλ = λ(Hi − Hi+1 )vλ = 2kvλ ,
(H i − H i+1 )wλ = 0.
We compute that
Eei,i+1 Eei+1,i wλ = λ(Hi − Hi+1 )wλ = 2kwλ ,
E i,i+1 Eei+1,i wλ = (H i − H i+1 )wλ = 0,
Eei,i+1 E i+1,i vλ = (H i − H i+1 )vλ = wλ ,
E i,i+1 E i+1,i vλ = (H i + H i+1 )vλ = 0.

It follows that Eei,i+1 u = E i,i+1 u = 0. Also we have by weight consideration that


Eej, j+1 u = E j, j+1 u = 0 for j ̸= i. Hence, u is a singular vector in ∆(λ), whose weight
is clearly equal to λ − εi + εi+1 . 
68 2. Finite-dimensional modules

Lemma 2.42. Let λ = ∑ni=1 λi εi ∈ h∗0̄ . Assume that λi = −λ j for some 1 ≤ i ̸= j ≤ n.


Then, χλ = χλ−tα for α = εi − ε j and any t ∈ C.

Proof. We first claim that χλ = χλ−εi +ε j . By Proposition 2.37 we are reduced to


prove the claim when j = i + 1, and the claim in this case follows by Lemmas 2.19
and 2.41. By a repeated application of the claim, we have χλ = χλ−tα for any
t ∈ Z+ . Since Z+ is Zariski dense in C, the lemma follows. 

The following theorem is the q(n)-counterpart of Theorem 2.24.

Theorem 2.43. Let g = q(n). Then hc∗ (Z(g)) = Im(η∗ ) = Γn,C .

Proof. Let λ = ∑ni=1 λi εi be an integral weight. Let z ∈ Z(g). By Proposition 2.37,


we know that hc∗ (z) ∈ S(Πh∗1̄ )W . By unraveling the definition of χλ , Lemma 2.42
implies that hc∗ (z)|εi =−ε j =t is independent of t. Then by the characterization of Γn
given in (A.57), we have hc∗ (z) ∈ Γn,C . Together with Proposition 2.39, we have
shown that hc∗ (Z(g)) ⊆ Γn,C ⊆ Im(η∗ ).
It remains to show that hc∗ (Z(g)) = Im(η∗ ). This follows by the corresponding
identity on the associated graded and then a standard filtered algebra argument,
exactly as in the proof of Theorem 2.24. We leave the details to the interested
reader. 

We summarize various maps for the q(n) case in the following diagram, which
only becomes commutative when we pass from Z(g) to its associated graded (note
that the other algebras are graded).

γ−1
(2.27) Z(g) / S(g)g

=
tt
tt
hc∗ tttη η
 ytt ∼ ∗ 
S(Πh∗ )W o
=
1̄ S(h )W

θ

Remark 2.44. Actually hc and hc∗ are injective (see Sergeev [Sv3]), and so we have
an isomorphism of algebras hc∗ : Z(g) → Γn,C . The injectivity of hc∗ does not play
any role in the representation theory of q(n) in the book.

2.3.3. Linkage for q(n). For a root α of the form εi − ε j , define

α∨ := Hi + H j ∈ h0̄ .

We define a relation on h∗0̄

λ ∼ µ, for λ, µ ∈ h∗0̄ ,
2.3. Harish-Chandra homomorphism and linkage for q(n) 69

if there exist a collection of roots αi , complex numbers ci , for 1 ≤ i ≤ ℓ, and an


element w ∈ W such that
( ℓ ) ⟨ ⟩ ⟨ ⟩
(2.28) µ = w λ − ∑ ca αa , αi , α∨j = 0, λ, α∨j = 0, 1 ≤ i, j ≤ ℓ.
a=1
If λ ∼ µ, we say that λ and µ are linked. The following theorem is a q(n)-
counterpart of Theorem 2.27. It also implies that linkage for q(n) is an equivalence
relation.
Theorem 2.45. Let h = h0̄ + h1̄ be a Cartan subalgebra of q(n), and let λ, µ ∈ h∗0̄ .
Then λ is linked to µ if and only if χλ = χµ .

Proof. Assume first that λ ∼ µ. By a repeated application of Lemma 2.42, the


second condition in (2.28) for λ ∼ µ implies that
χλ = χλ−c1 α1 = χλ−c1 α1 −c2 α2 = . . . = χλ−∑ℓ .
a=1 ca αa

Since µ = w(λ − ∑ℓa=1 ca αa ) by (2.28) again, it follows by Proposition 2.37 that


χλ−∑ℓ ca αa = χµ , whence χλ = χµ .
a=1

Now assume that χλ = χµ . Write λ = ∑ni=1 λi εi and µ = ∑ni=1 µi εi . Let z2k+1 ∈


Z(g) such that hc(z2k+1 ) = p2k+1 , the (2k + 1)st power sum in H1 , . . . , Hn . From
χλ (z2k+1 ) = χµ (z2k+1 ) we obtain that ⟨λ, hc(z2k+1 )⟩ = ⟨µ, hc(z2k+1 )⟩. Hence we have
n n
∑ λ2k+1
i = ∑ µ2k+1
j , ∀k ∈ Z+ ,
i=1 j=1

which is equivalent to the following generating function identity in an indetermi-


nate t:
n n
∑ sinh(λit) = ∑ sinh(µ jt).
i=1 j=1

Here we recall that the function


et − e−t t 2k+1
sinh(t) = =∑
2 k≥0 (2k + 1)!

satisfies sinh(−at) = − sinh(at) for any scalar a ∈ C.


Pick a maximal number of pairs (ia , ja ) such that λia = −λ ja , with 1 ≤ ia <
ja ≤ n and all ia , ja are distinct, for 1 ≤ a ≤ ℓ. Denote
Iλ = {1 ≤ i ≤ n | i ̸= ia , i ̸= ja , ∀a}.
Similarly, pick a maximal number of pairs (i′b , jb′ ) such that µi′b = −µ jb′ , with 1 ≤
i′b < jb′ ≤ n and all i′b , jb′ are distinct, for 1 ≤ b ≤ r, and define Iµ accordingly. It
follows by definition that
(2.29) ∑ sinh(λit) = ∑ sinh(µ jt).
i∈Iλ j∈Iµ
70 2. Finite-dimensional modules

We first observe that |Iλ | = |Iµ | (and hence ℓ = r) by dividing both sides of the
above identity by t and then taking the limit t 7→ 0. By the definition of Iλ and
Iµ , we conclude from (2.29) that there is a bijection between (λi )i∈Iλ and (µ j ) j∈Iµ .
Now we can find a permutation w ∈ Sn which extends the bijection between Iλ and
Iµ above such that w−1 µ − λ is a linear combination of the roots εia − ε ja . Thus,
λ ∼ µ. 

Theorem 2.45 has the following implications. For a composition factor L(µ)
in a Verma module ∆(λ) of q(n), µ must satisfy the following conditions: λ − µ ∈
Z+ Φ+ , and µ ∼ λ. This will be referred to as the linkage principle for q(n).

Definition 2.46. Let g = q(n) and h be its standard Cartan subalgebra. The degree
of atypicality of a weight λ ∈ h∗0̄ , denoted by #λ, is the maximum number of pairs
(ia , ja ) with all ia , ja distinct, such that ⟨λ, Hia + H ja ⟩ = 0. A weight λ ∈ h∗0̄ is called
typical if #λ = 0, and is atypical otherwise.

We have the following corollary from the proof of Theorem 2.45.

Corollary 2.47. If χλ = χµ for λ, µ ∈ h∗0̄ , then the degrees of atypicality for λ and
µ coincide.

2.3.4. Typical finite-dimensional characters of q(n). Note that λ ∈ h∗0̄ is typical


(cf. Definition 2.46) if and only if

∏ ⟨λ, Hi + H j ⟩ ̸= 0.
1≤i< j≤n

Recall that the finite-dimensional irreducible q(n)-modules are classified in


terms of highest weights in Theorem 2.16. The following is the q(n)-counterpart
of Lemma 2.31.

Lemma 2.48. Let g = q(n) and λ, µ ∈ h∗0̄ with λ typical. Suppose that χλ = χµ .
Then, there exists w ∈ W such that µ = w(λ), and µ is typical.

Proof. Recalling P from (2.26), we have εi + ε j = θ(Hi + H j ), and


( )
P=θ ∏ i j .
(H + H )
1≤i< j≤n

Clearly P is W -invariant. By Lemma 2.40 and Theorem 2.43, we have θ( f )P ∈


hc∗ (Z(g)), for f ∈ S(h0̄ )W . Also recall ρ = 0.
With these ingredients in place, the argument in the proof of Lemma 2.31 goes
through in this case as well. We leave the details to the interested reader. 

The following is a q(n)-analogue of Theorem 2.32.


2.4. Extremal weights of finite-dimensional simple modules 71

Theorem 2.49. Let g = q(n) and let λ ∈ h∗0̄ be a typical weight such that the irre-
ducible q(n)-module L(λ) is finite dimensional. Then

ℓ(λ)+δ(λ) ∏β∈Φ+ (1 + e−β )


chL(λ) = 2 2 1̄

∏α∈Φ+ (1 − e−α ) w∈W


∑ (−1)ℓ(w) ew(λ) .

Proof. We follow the strategy of the proof of Theorem 2.32.


The character of the Verma module ∆(λ) is given by

ℓ(λ)+δ(λ)
λ
∏β∈Φ+ (1 + e−β )

ch∆(λ) = 2 2 e .
∏α∈Φ+ (1 − e−α )

Observe that the highest weight spaces of L(µ) and ∆(µ), for every µ ∈ W λ, have
ℓ(λ)+δ(λ)
dimension 2 2 , thanks to ℓ(µ) = ℓ(λ).
Imitating the proof of Proposition 2.20, we obtain that

∏α∈Φ+ (1 − e−α )
(2.30)

(1 + e−β )
chL(λ) = ∑ aµλ ∑ (−1)ℓ(w) ew(µ) ,
β∈Φ+

µ∈X w∈W

where X consists of Φ+

-dominant integral weights such that aµλ ̸= 0.
Since λ is typical and χλ = χµ for all µ ∈ X, it follows by Lemma 2.48 that
ℓ(λ)+δ(λ)
µ = λ. Recalling that aλλ = 2 2 , we can rewrite (2.30) as

∏α∈Φ+ (1 − e−α ) ℓ(λ)+δ(λ)


∏β∈Φ+ (1 + e−β )
chL(λ) = 2 2
∑ (−1)ℓ(w) ew(λ+ρ) ,

w∈W

from which the theorem follows. 

Remark 2.50. The assumption on λ in Theorem 2.49 forces λi > λi+1 for all i, and
so ℓ(λ) = n or ℓ(λ) = n − 1.

2.4. Extremal weights of finite-dimensional simple modules


In this section, we study the extremal weights of a finite-dimensional irreducible
module over a Lie superalgebra of types gl and osp. A main complication arises
from the existence of non-conjugate Borel subalgebras for Lie superalgebras. For
the Lie superalgebra of type gl the extremal weights for irreducible polynomial
representations are determined. For the Lie superalgebras of type osp the extremal
weights of finite-dimensional irreducible representations of integer weights are de-
termined. In all cases, the answers are given in terms of the hook diagrams.
72 2. Finite-dimensional modules

2.4.1. Extremal weights for gl(m|n). We start with some general remarks for any
basic Lie superalgebra g. Let L be a finite-dimensional g-module. Then given any
Borel subalgebra b of g associated to a positive system Φ+ , there exists a unique
weight λb for L such that λb +α is not a weight for L for any α ∈ Φ+ , and the weight
space Lλb is one-dimensional (cf. Proposition 1.35). The weight λb is called the
b-extremal weight for the g-module L. Two Borel subalgebras b and b′ of g are in
general not conjugate. Thus, the b-extremal weight and the b′ -extremal weight of L
may in general not be conjugate by the Weyl group W , in contrast to the semisimple
Lie algebra setting.
In this subsection we let g = gl(m|n). Let h be its standard Cartan subalgebra
with standard Borel subalgebra bst . For an (m|n)-hook partition λ we may regard
λ♮ as an element in h∗ as in (2.4). We shall determine the b-extremal weight of
L(g, bst , λ♮ ) for any Borel subalgebra b. The class of gl(m|n)-modules L(g, bst , λ♮ ),
for all (m|n)-hook partitions λ, are called the polyonomial modules of gl(m|n),
and they will feature significantly in Chapter 3 on Schur duality.
Recall the weights δi and ε j from Section 1.2.3. Let b be a Borel subalgebra
of g and Φ+ be its positive system. Assume that the εδ-sequence from Section 1.3
associated to b is δd1 εe1 δd2 εe2 · · · δdr εer where the exponents denote the correspond-
ing multiplicities (all di , ei are positive except possibly d1 = 0 or er = 0). There
exist a permutation s of {1, . . . , m} and a permutation t of {1, . . . , n} such that the
δ’s and ε’s appearing in the εδ-sequence from left to right are δs(1) , δs(2) , . . . , δs(m) ,
and εt(1) , εt(2) , . . . , εt(n) , respectively.
Define
u u
(2.31) du := ∑ da , and eu := ∑ ea ,
a=1 a=1

for u = 1, . . . , r, and let d0 = e0 = 0. Note dr = m, er = n. Define the b-Frobenius


coordinates (pi |q j ) of an (m|n)-hook partition λ as follows. For 1 ≤ i ≤ m, 1 ≤
j ≤ n, let
pi = max{λi − eu , 0}, if du < i ≤ du+1 for some 0 ≤ u ≤ r − 1,
(2.32)
q j = max{λ′j − du+1 , 0}, if eu < j ≤ eu+1 for some 0 ≤ u ≤ r − 1.
Associated to an (m|n)-hook Young diagram λ and a Borel subalgebra b, we define
a weight λb ∈ h∗ in terms of the b-Frobenius coordinates (pi |qi ) to be
m n
λb := ∑ pi δs(i) + ∑ q j εt( j) .
i=1 j=1

It is elementary to read off the b-Frobenius coordinates of λ from the Young dia-
gram of λ in general, as illustrated by the next example.
Example 2.51. Let b be the Borel subalgebra of gl(5|4) associated to the following
fundamental system:
2.4. Extremal weights of finite-dimensional simple modules 73

δ2 − δ3 ε3 − ε2 δ1 − δ4 ε4 − ε1
⊗ ⊗ ⊗ ⊗
⃝ ⃝ ⃝ ⃝
δ3 − ε3 ε2 − δ1 δ4 − ε4 ε1 − δ5

Consider the (5|4)-hook diagram λ = (14, 11, 8, 8, 7, 4, 3, 2). The b-Frobenius


coordinates for λ is:
p1 = 14, p2 = 11, p3 = p4 = 6, p5 = 3; q1 = q2 = 6, q3 = 3, q4 = 2.
These are read off from the Young diagram of λ by following the εδ sequence
δδεεδδεεδ (with δ for rows and ε for columns of λ) as follows:
n=4
← p1 →
← p2 →
← p3 →
← p4 →
↑ ↑ ↑ ↑ ← p5 →
m = 5 q1 q2 q3 q4
↓ ↓ ↓ ↓

Then we obtain that


λb = p1 δ2 + p2 δ3 + p3 δ1 + p4 δ4 + p5 δ5 + q1 ε3 + q2 ε2 + q3 ε4 + q4 ε1
= 6δ1 + 14δ2 + 11δ3 + 6δ4 + 3δ5 + 2ε1 + 6ε2 + 6ε3 + 3ε4 .
Theorem 2.52. Let λ be an (m|n)-hook partition. Let b be an arbitrary Borel
subalgebra of gl(m|n). Then, the b-extremal weight of the simple gl(m|n)-module
L(g, bst , λ♮ ) is λb .

Proof. Set V = L(g, bst , λ♮ ). One distinguished feature for V is that all weights ν
of V are polynomial in the sense that ν(Eii ) ∈ Z+ for all i ∈ I(m|n); see Chapter 3,
Theorem 3.10.
The theorem holds for the standard Borel subalgebra bst , which corresponds to
st
the sequence of m δ’s followed by n ε’s, thanks to λb = λ♮ . By Corollary 1.27,
bst can be converted to b by a finite number of real and odd reflections. Thus, we
are led to consider two Borel subalgebras b1 and b2 , where b2 is obtained from
b1 by applying a simple reflection corresponding to a simple root α. Let µ be the
b2 -extremal weight for V . To complete the proof of the theorem, we will show that
µ = λb2 , assuming the theorem holds for b1 .
74 2. Finite-dimensional modules

If α is an even root, then rα (λb1 ) = λb2 by definition, and the validity of the
theorem for b1 implies its validity for b2 .
Now assume that α = ±(δi − ε j ) is an odd simple root. The corresponding
coroot is hα = ±(Eī,ī + E j j ).
If λb1 (hα ) = 0, then we must have λb1 (Eī,ī ) = λb1 (E j j ) = 0 (as each has to be
nonnegative), and λb1 = λb2 . Then by Lemma 1.36 we have µ = λb1 = λb2 .
Suppose that λb1 (hα ) ̸= 0 and α = δi − ε j . Diagrammatically we can represent
λb1 (Eī,ī )
= a and λb1 (E j j ) = b by Diagram (ii) below, and this forces that a > 0.
By Lemma 1.36, we have µ = λb1 − δi + ε j , and hence µ(E j j ) = b + 1 and µ(Eī,ī ) =
a − 1, which are represented by Diagram (i) below. The remaining parts of µ are
the same as those for λb1 and hence we conclude that µ = λb2 , as claimed.
Suppose that λb1 (hα ) ̸= 0 and α = ε j − δi . Diagrammatically we can represent
λb1 (Ej j ) = b + 1 and λ (Eī,ī ) = a − 1 by Diagram (i) below, and this forces that
b1

b ≥ 0 and a ≥ 1. Lemma 1.36 implies that µ = λb1 − ε j + δi and hence µ(Eī,ī ) = a


and µ(E j j ) = b can be represented by Diagram (ii) below. Hence µ = λb2 .

a−1 a

b+1
b

(i) (ii)

This completes the proof of the theorem. 


Example 2.53. Let us describe the highest weights in Theorem 2.52 with respect
to the three Borel subalgebras of special interest.
st
(1) As seen above, λb = λ♮ .
(2) If we take the opposite Borel subalgebra bop corresponding to a sequence
of n ε’s followed by m δ’s, then
λb = λ′1 ε1 + . . . + λ′n εn + max{λ1 − n, 0}δ1 + . . . + max{λm − n, 0}δm .
op

(3) In case when |m − n| ≤ 1, we may take a Borel subalgebra bo whose


simple roots are all odd (or equivalently, the corresponding εδ- sequence
is alternating between ε and δ):
⊗ ⊗ ⊗ ⊗ ⊗
···

o
In this case Theorem 2.52 states the coefficients of δ and ε in λb are given
by the modified Frobenius coordinates (pi |qi )i≥1 of the partition λ (re-
spectively, λ′ ), when the first simple root is of the form δ−ε (respectively,
2.4. Extremal weights of finite-dimensional simple modules 75

ε − δ). Here by modified Frobenius coordinates we mean


pi = max{λi − i + 1, 0}, qi = max{λ′i − i, 0}
so that ∑i (pi + qi ) = |λ|; “modified” here refers to a shift by 1 from the
pi coordinates defined in [Mac, Chapter 1, Page 3].
The modified Frobenius coordinates in this case can be read off from
the Young diagram by alternatively reading off (and deleting thereafter)
the number of boxes in rows and columns. As an illustration, if λ =
(7, 5, 4, 3, 1), then we have (p1 , p2 , p3 |q1 , q2 , q3 ) = (7, 4, 2|4, 2, 1).

7
4
2
4 2 1

2.4.2. Extremal weights for spo(2m|2n + 1). Let us denote the weights of the
natural spo(2m|2n + 1)-module C2m|2n+1 by ±δi , 0, ±ε j for 1 ≤ i ≤ m, 1 ≤ j ≤ n.
Recall that the standard Borel subalgebra bst of spo(2m|2n + 1) is associated to
the following fundamental system

δn − ε1

⃝ ⃝ ··· ⃝ ⃝ ··· ⃝=⇒⃝
δ1 − δ2 δ2 − δ3 δn−1 − δn ε1 − ε2 εm−1 − εm εm

Let b be a Borel subalgebra. According to Section 1.3.3 a Dynkin diagram for


spo(2m|2n + 1) always has a type A end while the other end is a short (even or odd)
root. Starting from the type A end, the simple roots for b of spo(2m|2n + 1) give
rise to an εδ-sequence δd1 εe1 δd2 εe2 · · · δdr εer and sequences of ±1’s: (ξi )1≤i≤m ∪
(η j )1≤ j≤n (all the di and e j are positive except possibly d1 = 0 or er = 0). Further-
more, the Dynkin diagram contains a short odd root if and only if er = 0.
Recall the definitions of du and eu from (2.31), and those of pi and q j from
(2.32). Similar to the type A case there exist a permutation s of {1, . . . , m} and a
permutation t of {1, . . . , n}, so that the simple roots for b are given by
ξi δs(i) − ξi+1 δs(i+1) , 1 ≤ i ≤ m, i ̸∈ {du |u = 1, . . . , r};
η j εt( j) − η j+1 εt( j+1) , 1 ≤ j ≤ n, j ̸∈ {eu |u = 1, . . . , r};
ξdu δs(du ) − η1+eu−1 εt(1+eu−1 ) , for 1 ≤ u ≤ r if er > 0 (or 1 ≤ u < r if er = 0);
ηeu εt(eu ) − ξ1+du δs(1+du ) , u = 1, . . . , r − 1;
ηer εt(er ) , if er > 0 (or ξdr δs(dr ) if er = 0).
76 2. Finite-dimensional modules

According to Theorem 2.11, a complete list of finite-dimensional irreducible


modules of spo(2m|2n + 1) with integer weights are the bst -highest weight g-
modules L(g, bst , λ♮ ) of highest weights λ♮ in (2.4), for some (m|n)-hook partition
λ.

Example 2.54. Suppose that the corresponding Dynkin diagram of a Borel subal-
gebra of spo(9|10) is as follows:

δ2 + δ3 −ε3 − ε2 δ1 − δ4 ε4 − ε1
⊗ ⊗ ⊗ ⊗
⃝ ⃝ ⃝ ⃝ =⇒ y
−δ3 + ε3 ε2 − δ1 δ4 − ε4 ε1 + δ5 −δ5

We read off a signed sequence with indices δ2 (−δ1 )(−ε1 )ε2 δ3 δ4 ε4 ε3 (−δ5 ). In par-
ticular, we obtain a sequence δδεεδδεεδ by ignoring the signs and indices. In this
case, d1 = d2 = 2, d3 = 1, and e1 = e2 = 2. Furthermore, the sequences (ξi )1≤i≤5
and (η j )1≤ j≤4 are (1, −1, 1, 1, −1) and (−1, 1, 1, 1), respectively.

Theorem 2.55. Let λ be an (m|n)-hook partition. Let b be a Borel subalgebra of


spo(2m|2n + 1) and retain the above notation. Then, the b-highest weight of the
simple spo(2m|2n + 1)-module L(g, bst , λ♮ ) is
m n
λb := ∑ ξi pi δs(i) + ∑ η j q j εt( j) .
i=1 j=1

Proof. The arguments here are similar to the ones given in the proof of Theorem
2.52 for gl(m|n), and so we will be sketchy.
We first note that the theorem holds for the standard Borel subalgebra bst ,
which corresponds to the sequence of m δ’s followed by n ε’s with all signs ξi
st
and η j being positive, i.e., λb = λ♮ .
All Borels are linked to the standard Borel by a sequence of even and odd
simple reflections. Let b and b′ be two Borel subalgebras be related by a simple
reflection with respect to a simple root α, and assume the theorem holds for b′ . If

α is even, then clearly rα (λb ) = λb , and the theorem holds for b. In the case when
α is odd, we have four cases to consider, by setting α = ±δi ± ε j . Each case is
verified by the same type of argument as in the proof of Theorem 2.52. 

Example 2.56. With respect to the Borel b of spo(10|9) as in Example 2.54, the
b-extremal weight of L(g, bst , λ♮ ) for λ as given in Example 2.51 equals
λb = p1 δ2 − p2 δ3 + p3 δ1 + p4 δ4 − p5 δ5 − q1 ε3 + q2 ε2 + q3 ε4 + q4 ε1
= 6δ1 + 14δ2 − 11δ3 + 6δ4 − 3δ5 + 2ε1 + 6ε2 − 6ε3 + 3ε4 .

Corollary 2.57. Every finite-dimensional irreducible spo(2m|2n + 1)-module of


integer highest weight is self-dual.
2.4. Extremal weights of finite-dimensional simple modules 77

Proof. By a standard fact in highest weight theory, the dual module L(g, bst , λ♮ )∗
has bst -highest weight equal to the opposite of the bst -lowest weight of L(g, bst , λ♮ ).
Denote by bop the opposite Borel to the standard one bst . It follows by Theo-
rem 2.55 that the bop -extremal weight (i.e., the bst -lowest weight) of the module
L(g, bst , λ♮ ) is −λ♮ . This proves the corollary. 

Consider the following Dynkin diagram of spo(2m|2n + 1) with simple roots


attached, and let us denote its associated Borel subalgebra by bsd :


⃝⇐=⃝ ⃝ ··· ⃝ ⃝ ··· ⃝
−ε1 ε1 − ε2 ε2 − ε3 εn − δm δm − δm−1 δ2 − δ1

We have the following immediate corollary of Theorem 2.11 and Theorem 2.55.

Corollary 2.58. An irreducible spo(2m|2n + 1)-module of integer highest weight


with respect to the Borel subalgebra bsd is finite dimensional if and only if the
bsd -highest weight is of the form
m n
λb = − ∑ λi δi − ∑ max{λ′j − m, 0} εn− j+1 ,
sd
(2.33)
i=1 j=1

where λ = (λ1 , λ2 , . . .) is an (m|n)-hook partition.

2.4.3. Extremal weights for spo(2m|2n). Let n ≥ 2. Let us denote the weights of
the natural spo(2m|2n)-module C2m|2n by ±δi , ±ε j for 1 ≤ i ≤ m, 1 ≤ j ≤ n. The
standard Borel subalgebra bst of spo(2m|2n) is the one associated to the following
Dynkin diagram
⃝ εn−1 − εn
δm − ε1

⃝ ⃝ ··· ⃝ ⃝ ··· ⃝
δ1 − δ2 δ2 − δ3 δm−1 − δm ε1 − ε2 εn−2 − εn−1@@
⃝ εn−1 + εn

According to Section 1.3.4 there are two types of Dynkin diagrams and correspond-
ing Borel subalgebras for spo(2m|2n):
(i) Diagrams of |-shape, i.e., Dynkin diagrams with a long simple root ±2δi .
(ii) Diagrams of -shape, i.e., Dynkin diagrams with no long simple root.
We will follow the notation for spo(2m|2n + 1) in Section 2.4.2 for fundamen-
tal systems in terms of signed εδ-sequences, so we have permutations s,t, and signs
ξi , η j . We fix the ambiguity on the choice of the last sign ηn associated to a Borel b
of -shape, by demanding the total number of negative signs among η j (1 ≤ j ≤ n)
to be always even.
78 2. Finite-dimensional modules

Let λ be an (m|n)-hook partition, and let the b-Frobenius coordinates (pi |q j )


be as defined in Section 2.4.1. Introduce the following weights:
m n
λb := ∑ ξi pi δs(i) + ∑ η j q j εt( j) ,
i=1 j=1
m n−1
λb− := ∑ ξi pi δs(i) + ∑ η j q j εt( j) − ηn qn εt(n) .
i=1 j=1

The weight λb− will only be used for Borel b of -shape. Note that λb = λ♮ and
st

st
we shall denote λ♮− := λb− .
Given a Borel b of |-shape, we define s(b) to be the sign of ∏nj=1 η j .
Recall from Theorem 2.14 that a finite-dimensional irreducible spo(2m|2n)-
module of integer highest weight with respect to the standard Borel subalgebra is
of the form L(g, bst , λ♮ ) or L(λ♮− ), where λ is an (m|n)-hook partition.

Theorem 2.59. Let n ≥ 2, and let λ be an (m|n)-hook partition.


(1) Assume b is of -shape. Then,
(i) λb is the b-extremal weight for the module L(g, bst , λ♮ ).
(ii) λb− is the b-extremal weight for the module L(g, bst , λ♮− ).
(2) Assume b is of |-shape. Then,
(i) λb is the b-extremal weight for L(g, bst , λ♮ ) if s(b) = +.
(ii) λb is the b-extremal weight for L(g, bst , λ♮− ) if s(b) = −.

Proof. Let b and b′ be Borel subalgebras. Suppose that they are related by b′ =
w(b), for some Weyl group element w. Note that the set of weights for L(g, bst , λ♮ )
′ ′
is W -invariant. By definition we observe that λb = w(λb ) and λb− = w(λb− ), and
hence the validity of the theorem for b implies its validity for b′ . Thus it suffices to
prove the theorem for a Borel subalgebra b with even roots {δi ± δ j |i ≤ j} ∪ {εk ±
εl |k < l}. We shall make this assumption for the remainder of the proof.
Suppose that the Dynkin diagram of b is of -shape. Thus the corresponding
εδ-sequence ends with an ε, i.e., it is of the form

· · · δ · · · ε · · · ε.

The fundamental system of b can be brought to the fundamental system corre-


sponding to bst by applying a sequence of odd reflections corresponding to odd
isotropic root of the form εi − δ j or δ j − εi . As Part (1) of the theorem is true for
the standard Borel bst , exactly the same argument as in the proof of Theorem 2.52
proves (1).
Now suppose that b is of |-shape. We first consider the Borel subalgebras b1
and b2 corresponding to the two fundamental systems below, respectively:
2.4. Extremal weights of finite-dimensional simple modules 79

εn−1 − εn

⃝ ··· ⃝ ⃝ ··· ⃝⇐=⃝
ε1 − ε2 εn − δ1 δ1 − δ2 δm−1 − δm 2δm

εn−1 + εn

⃝ ··· ⃝ ⃝ ··· ⃝⇐=⃝
ε1 − ε2 −εn − δ1 δ1 − δ2 δm−1 − δm 2δm

We have s(b1 ) = + and s(b2 ) = −.


First, using sequences of odd reflections, we can transform the standard fun-
damental system to the fundamental systems associated with b1 and b2 . We list
below two such sequences consisting of mn odd reflections that will take bst to b1
and b2 , respectively.
{δm − ε1 , . . . , δ1 − ε1 ; δm − ε2 , . . . , δ1 − ε2 ; . . . ; δm − εn , . . . , δ1 − εn },
{δm − ε1 , . . . , δ1 − ε1 ; δm − ε2 , . . . , δ1 − ε2 ; . . . ; δm + εn , . . . , δ1 + εn }.
We note that the above two sequences are identical except for the last n reflections
st
(where + signs replace − signs). Starting from λb± for the standard Borel bst and
repeatedly using Lemma 1.26 at each step, it is straightforward to show that Part
(2) of the theorem is true for b1 and b2 . Now if s(b) = + (respectively, s(b) = −)
then b can be brought to b1 (respectively b2 ) by a sequence of odd reflections
corresponding to odd isotropic roots of the form δ j − εi or εi − δ j . Now the same
type of argument as in the proof of Theorem 2.52 proves (2). 
Corollary 2.60. Let n ≥ 0 be even. Then every finite-dimensional irreducible
spo(2m|2n)-module of integral highest weight is self-dual.

Proof. The argument here is similar to the proof of Corollary 2.57. For n = 0, the
corollary follows by the well-known fact that the longest element in the Weyl group
for sp(2m) is −1. Now assume n ≥ 2 is even. Denote by bop the opposite Borel
to the standard one bst . It follows by Theorem 2.59 that the bop -extremal weight
of the module L(g, bst , λ♮ ) (and respectively, L(g, bst , λ♮− )) is −λ♮ (and respectively,
−λ♮− ). The corollary follows. 
Remark 2.61. The remaining b-extremal weights for the modules L(g, bst , λ♮ )
when s(b) = − or for the modules L(g, bst , λ♮− ) when s(b) = + do not seem to
afford a uniform simple answer.

Consider the following Dynkin diagram of spo(2m|2n) whose Borel subalge-


bra is denoted by bsd .

ε1 − ε2 ⃝
@ εn − δm

@⃝ ··· ⃝ ⃝ ··· ⃝
ε2 − ε3 εn−1 − εn δm − δm−1 δ2 − δ1
−ε1 − ε2 ⃝
80 2. Finite-dimensional modules

We record the following corollary of Theorem 2.14 and Theorem 2.59.

Corollary 2.62. Let n ≥ 2. An irreducible spo(2m|2n)-module of integral highest


weight with respect to the Borel subalgebra bsd is finite dimensional if and only if
the highest weight is of the form
m n−1
(2.34) − ∑ λi δi − ∑ max{λ′j − m, 0} εn− j+1 ± max{λ′n − m, 0} ε1 ,
i=1 j=1

where λ = (λ1 , λ2 , . . .) is an (m|n)-hook partition.

2.5. Notes
Section 2.1. A systematic study of the representation theory of finite-dimensional
(simple) Lie superalgebras was initiated by Kac [K2]. The classification of finite-
dimensional irreducible representations of the Lie superalgebras of types gl and
spo (Propositions 2.2 and 2.2; Theorems 2.11 and 2.14) appeared first in Kac [K1],
where the finite dimensionality condition was formulated in terms of Dynkin labels
instead of hook partitions. The proof using odd reflections in the most challenging
spo type given here follows Shu-Wang [SW] for the cases of integer weights, and
the proof using odd reflections and the Bernstein functor (see Santos [San]) for
the cases of half-integer weights here is new. Yet another closely related approach
to the classification is provided in Azam-Yamane-Yousofzadeh [AYY] by means
of Weyl groupoids. Theorem 2.16 on the classification of finite-dimensional irre-
ducible modules of q(n) was formulated by Penkov [Pe, Theorem 4], who proved
the “only if” direction. We are not aware of a written proof for the “if” direction in
the literature. Our proof uses the Bernstein functor again.
Section 2.2. The Harish-Chandra homomorphisms for Lie superalgebras was
first studied by Kac [K2], who formulated results in terms of fractional fields. The
analogue of Chevalley theorem on invariant polynomials for simple Lie superalge-
bras has been established by Sergeev [Sv3], which are presented here as a part of
Theorem 2.24 for type gl and osp. Our approach of calculating the images of the
Harish-Chandra homomorphisms in this book is different and new, and it uses the
connection to the theory of supersymmetric functions for types gl and spo and the
theory of Schur Q-functions for type q. There is another approach developed by
Gorelik [Go] following Kac [K5] on the centers of the universal enveloping alge-
bras. A variant of Theorem 2.27 on linkage principle for gl and osp was originally
stated as a conjecture in Kac [K2, §6].
The notion of typicality for basic Lie superalgebras is due to Kac [K2]. Corol-
lary 2.26 (which is due to Kac [K2, Theorem 2] by different arguments) was used
by Kac to derive the irreducible character formula for finite-dimensional typical
modules of basic classical Lie superalgebas (including Theorem 2.32).
2.5. Notes 81

Section 2.3. This section is the counterpart of Section 2.2 for the queer Lie
superalgebra q(n). Theorem 2.43 on the image of the Harish-Chandra homomor-
phism for the queer Lie superalgebra was formulated without proof in a short an-
nouncement of Sergeev [Sv4]. We fill in the details of a proof by presenting several
explicit formulas of singular vectors in Verma modules. The notion of typicality
for q(n) was formulated by Penkov [Pe]. Theorem 2.49 on the typical character
formula for q(n) is due to Penkov [Pe, Theorem 2]. Theorem 2.45 on linkage
principle for q(n) or its variant has been regarded as folklore after [Sv4].
Section 2.4. As we shall see in Chapter 3, the irreducible polynomial repre-
sentations of gl(m|n) were classified by Sergeev [Sv1] (also see Berele and Regev
[BeR]). The results on extremal weights of the irreducible polynomial representa-
tions for gl and finite-dimensional irreducible integer weight modules for spo were
due to Ngau Lam and the authors (for the spo case see [CLW] and for the gl case
see [CW5]). The special case for type gl when a fundamental system consists of
only odd simple roots goes back to [CW3].
The irreducible character problem for finite-dimensional modules over Lie su-
peralgebras have been a challenging one. For gl(m|n), there has been complete
solutions in completely different approaches due to Serganova [Sva], Brundan
[Br1], Brundan-Stroppel [BrS], and Cheng-Lam [CL2] (also see Cheng-Wang-
Zhang [CWZ]). There is also a combinatorial dimension formula by Su-Zhang
[SZ]. For q(n), there has been complete solutions due to Penkov-Serganova [PS2]
and Brundan [Br2]. There were numerous partial results in the literature over the
years, see for example [BL, CWZ2, J, JHKT, PS2, San, Sv1, Zou]. Chapter 6
of this book offers a complete solution for a wider class of (including all finite-
dimensional) modules of gl(m|n) and osp(m|2n).
Chapter 3

Schur duality

Schur duality, which was popularized in Weyl’s book The Classical Groups, is
an interplay between representations of the general linear Lie group/algebra and
representations of the symmetric group. On the combinatorial level, it explains
their mutual connections to Schur functions and Young tableaux.
In this chapter, we describe Sergeev’s superalgebra generalization of Schur du-
ality. The first generalization is a duality between the general linear Lie superalge-
bra and the symmetric group. This allows us to classify the irreducible polynomial
representations of the general linear Lie superalgebra, and obtain their character
formula in terms of supertableaux and also in terms of super Schur functions. The
second generalization is a duality between the queer Lie superalgebra and a twisted
(or spin) group algebra Hd of the hyperoctahedral group. The representation theory
of the algebra Hd is systematically developed, and its connection with symmetric
functions such as Schur Q-functions is explained in detail. The irreducible polyno-
mial representations of the queer Lie superalgebra are classified in terms of strict
partitions, and their characters are shown to be the Schur Q-functions up to some
2-powers.

3.1. Generalities for associative superalgebras


In this section, we classify the finite-dimensional simple associative superalgebras
over C. We also formulate the superalgebra generalizations of Schur’s Lemma,
double centralizer property, and the Wedderburn Theorem. By studying the center
of a finite group superalgebra, we obtain a numerical identity relating the number
of simple supermodules to the number of split conjugacy classes.

83
84 3. Schur duality

In this section, a superalgebra is always understood to be a finite-dimensional


associative superalgebra with unity 1, and a module is always understood to be
finite-dimensional.

3.1.1. Classification of simple superalgebras. Let V be a vector superspace with


equal dimensional even and odd subspaces. Given an odd automorphism P of V
of order 2, we define the following subalgebra of the endomorphism superalgebra
End(V ):
Q(V ) = {x ∈ End(V ) | x and P super-commute}.
In case when V = Cn|n and P is the linear transformation in the block matrix form
(1.10), we write Q(V ) as Q(n), which consists of 2n × 2n matrices of the form:
( )
a b
,
b a
where a and b are arbitrary n × n matrices, for n ≥ 1. Note that we have a superal-
gebra isomorphism Q(V ) ∼ = Q(n) by properly choosing coordinates in V , whenever
dimV = n|n.
The matrix superalgebra M(m|n) is the algebra of matrices in the m|n-block
form ( )
a b
,
c d
whose even subspace consists of the matrices with b = 0 and c = 0, and whose odd
subspace consists of the matrices with a = 0 and d = 0. The matrix superalgebra
M(m|n) is isomorphic to the endomorphism superalgebra of Cm|n .
For a superalgebra A, we shall denote by |A| its underlying (i.e., ungraded)
algebra. By an ideal I and a module M of a superalgebra A, we always mean
that I and M are Z2 -graded, i.e., I = (I ∩ A0̄ ) ⊕ (I ∩ A1̄ ), and M = M0̄ ⊕ M1̄ such
that Ai M j ⊆ Mi+ j , for i, j ∈ Z2 . On the other hand, ideals and modules of |A| are
understood in the usual (i.e., ungraded) sense.
Denote by Z(|A|) the center of |A|. Clearly, Z(|A|) = Z(|A|)0 ⊕ Z(|A|)1 ,
where Z(|A|)i = Z(|A|) ∩ Ai for i ∈ Z2 .
Theorem 3.1. Let A be a finite-dimensional simple associative superalgebra.
(1) If Z(|A|)1 = 0, then A is isomorphic to the matrix superalgebra M(m|n)
for some m and n.
(2) If Z(|A|)1 ̸= 0, then A is isomorphic to Q(n) for some n.

We shall need the following lemma. Let pi : A → Ai be the projection for


i ∈ Z2 .
Lemma 3.2. Let A be a simple superalgebra. If J is a proper ideal of |A|, then
the induced maps pi |J : J → Ai are isomorphisms of vector spaces, and moreover,
J ∩ Ai = 0 for i ∈ Z2 .
3.1. Generalities for associative superalgebras 85

Proof. We claim that if I is a nonzero ideal of A0̄ , then


(3.1) I + A1̄ IA1̄ = A0̄ ,
and A1̄ I + IA1̄ = A1̄ . Indeed, the Z2 -graded subspace (I + A1̄ IA1̄ ) + (A1̄ I + IA1̄ )
of A is closed under left and right multiplications by elements of A0̄ and A1̄ , hence
it is an ideal of A. Now the claim follows from the simplicity of A.
For a proper ideal J of |A|, J ∩ A0̄ and p0̄ (J) are ideals of A0̄ , and J ∩ A0̄ ⊆
p0̄ (J). Actually, we must have J ∩ A0̄ ( p0̄ (J), for, otherwise J would be an ideal
of A, contradicting the simplicity of A. By inspection, A1̄ (J ∩ A0̄ )A1̄ ⊆ J ∩ A0̄ and
A1̄ p0̄ (J)A1̄ ⊆ p0̄ (J). It follows by (3.1) that
(3.2) J ∩ A0̄ = 0, p0̄ (J) = A0̄ .

We have A21̄ = A0̄ , for, otherwise A21̄ + A1̄ would be a proper ideal of A, which
contradicts with the simplicity of A. Hence, we have
(3.3) J ∩ A1̄ = 0, p1̄ (J) = A1̄ ,
by the following computations (which use (3.2) and that 1 ∈ A0̄ ):
J ∩ A1̄ = A0̄ (J ∩ A1̄ ) = A21̄ (J ∩ A1̄ ) ⊆ A1̄ (J ∩ A0̄ ) = 0,
p1̄ (J) ⊇ A1̄ p0̄ (J) = A1̄ A0̄ = A1̄ .
Now the lemma follows from (3.2) and (3.3). 

Proof of Theorem 3.1. We claim that Z(|A|)0̄ = C. To see this, let z ∈ Z(|A|)0̄ .
The map ℓz : A → A defined by left multiplication by z has an eigenvalue c ∈ C, and
the kernel of ℓz − c is a nonzero 2-sided ideal of A. The simplicity of A implies
that ℓz − c = 0, and hence z = c · 1.
(1) Assume Z(|A|)1 = 0. We claim that |A| is a simple algebra (and so |A|
is a matrix algebra by a classical structure theorem of simple algebras). Indeed,
assume |A| is a non-simple algebra with a proper ideal J. Consider w = p−10̄
(1) ∈ J
and u = p1̄ (w) ∈ A1̄ , where u ̸= 0 by Lemma 3.2. Then, w = 1 + u. For any
homogeneous element x ∈ Ai for some i ∈ Z2 , we have xw = x+xu and wx = x+ux,
and so pi (xw) = x = pi (wx). By Lemma 3.2, xw = wx and so xu = ux, that is,
u ∈ Z(|A|)1 . It follows by assumption that u = 0, which is a contradiction.
Consider the automorphism σ : |A| → |A| given by σ(ai ) = (−1)i ai for ai ∈ Ai
and i ∈ Z2 . There exists u ∈ |A| such that σ(a) = uau−1 for all a ∈ |A|, since every
automorphism of a matrix algebra is inner. We have u ∈ A0̄ as σ(u) = u. It follows
from σ2 = 1 that u2 ∈ Z(|A|)0̄ = C. We can choose u such that u2 = 1, and then by
a change of basis, take u to be a diagonal matrix with m 1’s followed by n (−1)’s
in the main diagonal. Hence, we conclude that A ∼ = M(m|n).
(2) For a nonzero element w ∈ Z(|A|)1 , we have w2 ∈ Z(|A|)0 = C. We must
have w2 ̸= 0, for, otherwise the annihilator of w is a proper ideal of the super-
algebra A, contrary to the simplicity of A. Then we may assume w2 = 1. We
86 3. Schur duality

have Z(|A|)1 = (Z(|A|)1 w)w ⊆ Z(|A|)0 w = Cw ⊆ Z(|A|)1 , and hence, Z(|A|) =


C + Cw. Moreover, A1̄ = (A1̄ w)w ⊆ A0̄ w ⊆ A1̄ , i.e., A1̄ = A0̄ w. The algebra A0̄
must be simple, for, otherwise a proper ideal I of A0̄ would give rise to a proper
ideal I +Iw of A (which is again a contradiction). Then A0̄ is a matrix algebra M(n)
for some n, and the superalgebra A = M(n) ⊕ M(n)w is isomorphic to Q(n). 

3.1.2. Wedderburn theorem and Schur’s Lemma. The basic results of finite-
dimensional semisimple (unital associative) algebras over C admit natural super
generalizations.
A module M of a superalgebra A is called semisimple, if every A-submodule
of M is a direct summand of M. The following are straightforward generalizations
of the classical structure results on semisimple modules and algebras, and they can
be proved in the same way as in the non-super setting.

Theorem 3.3. (Super Wedderburn Theorem) The following statements are equiva-
lent for a superalgebra A:
(i) Every A-module is completely reducible.
(ii) The left regular module A is a direct sum of minimal left ideals.
(iii) A is a direct sum of simple superalgebras.

A superalgebra A is called semisimple if it satisfies one of the three equivalent


conditions (i)–(iii) Theorem 3.3.
By Theorems 3.1 and 3.3, a finite-dimensional semisimple superalgebra A is
isomorphic to a direct sum of simple superalgebras as follows:

m ⊕
q
(3.4) A∼
= M(ri |si ) ⊕ Q(n j ),
i=1 j=1

where m = m(A) and q = q(A) are invariants of A. Observe that Cr|s is a simple
module (unique up to isomorphism) of the simple superalgebra M(r|s), and Cn|n is
a simple module (unique up to isomorphism) of the simple superalgebra Q(n). A
simple A-module V is annihilated by all but one such summand in (3.4). We say V
is of type M if this summand is of the form M(ri |si ) and of type Q if this summand is
of the form Q(n j ). These two types of simple modules are distinguished by a super
analogue of Schur’s Lemma.

Lemma 3.4. (Super Schur’s Lemma) If M and L are simple modules over a super-
algebra A, then

 1 if M ∼= L is of type M,
dim HomA (M, L) = 2 if M ∼= L is of type Q,

0 if M ̸∼
= L.
3.1. Generalities for associative superalgebras 87

Sketch of a proof. In case M ∼ = L is of type M, we can assume that A ∼


= M(r|s) for
some r, s, and that M = L = Cr|s . Then it is straightforward to check (as in the
non-super case) that HomM(r|s) (Cr|s , Cr|s ) = CI.
In case M ∼ = L is of type Q, we can assume that A ∼= Q(n) for some n, and that
M = L = Cn|n . Then one checks that HomQ(n) (Cn|n , Cn|n ) = CI + CP, where P is
the linear transformation in the block matrix form (1.13). 

3.1.3. Double centralizer property for superalgebras. Given two associative


superalgebras A and B, the tensor product A ⊗ B is naturally a superalgebra, with
multiplication defined by

(a ⊗ b)(a′ ⊗ b′ ) = (−1)|b|·|a | (aa′ ) ⊗ (bb′ ) (a, a′ ∈ A, b, b′ ∈ B).

Note the superalgebra isomorphism


Q(m) ⊗ Q(n) ∼
= M(mn|mn).
Hence, as a Q(m) ⊗ Q(n)-module, the tensor product Cm|m ⊗ Cn|n is a direct sum
of two isomorphic copies of a simple module (which is ∼
= Cmn|mn ), and we have
HomQ(n) (C , C
n|n mn|mn ∼
)=C m|m as a Q(m)-module.
Let A and B be semisimple superalgebras. Let M be a simple A-module of
type Q and let N be a simple B-module of type Q. Then, the A ⊗ B-module M ⊗ N
is a direct sum of two isomorphic copies of a simple module of type M, denoted
by 2−1 M ⊗ N; Moreover, HomB (N, 2−1 M ⊗ N) is naturally an A-module, which is
isomorphic to the A-module M.

Proposition 3.5. Suppose that W is a finite-dimensional vector superspace, and B


is a semisimple subalgebra of End(W ). Let A = EndB (W ). Then, EndA (W ) = B.
As an A ⊗ B-module, W is multiplicity-free, i.e.,

W∼
= ∑ 2−δi Ui ⊗Vi ,
i

where δi ∈ {0, 1}, {Ui } are pairwise non-isomorphic simple A-modules, {Vi } are
pairwise non-isomorphic simple B-modules. Moreover, Ui and Vi are of the same
type, and they are of type M if and only if δi = 0.

Proof. Assume that Va are all the pairwise non-isomorphic simple B-modules of
type M, and Vb are all the pairwise non-isomorphic simple B-modules of type Q.
Then the hom spaces Ua := HomB (Va ,W ) and Ub := HomB (Vb ,W ) are naturally
A-modules. By the super Schur’s Lemma, we know dimUb = kb |kb for some kb .
By the semisimplicity assumption on B, we have an isomorphism of B-modules
⊕ ⊕
W∼
= Ua ⊗Va ⊕ 2−1Ub ⊗Vb .
a b
88 3. Schur duality

By applying the super Schur’s Lemma, we obtain the following superalgebra iso-
morphisms:
⊕ ⊕
A = EndB (W ) ∼
= EndB (Ua ⊗Va ) ⊕ EndB (2−1Ub ⊗Vb )
a b
⊕ ⊕

= End(Ua ) ⊗ IVa ⊕ (End(2−1Ub ) ⊗ Q(1)) ⊗ IVb
a b
⊕ ⊕

= End(Ua ) ⊗ IVa ⊕ Q(Ub ) ⊗ IVb .
a b

Hence, A is a semisimple superalgebra, Ua are all the pairwise non-isomorphic


simple A-modules of type M, and Ub are all the pairwise non-isomorphic simple
A-modules of type Q.
Since A is now a semisimple superalgebra, we can reverse the roles of A and
B in the above computation of EndB (W ), and obtain the following superalgebra
isomorphisms:
⊕ ⊕
EndA (W ) ∼
= IUa ⊗ End(Va ) ⊕ IUb ⊗ Q(Vb ) ∼
= B.
a b

The proposition is proved. 

3.1.4. Split conjugacy classes in a finite supergroup. For a finite group G, the
group algebra CG is a semisimple algebra. It follows by using the Wedderburn
Theorem and comparing two different bases for the center of CG that the number
of simple G-modules coincides with the number of conjugacy classes of G.
Now assume that a finite group G contains a subgroup G0 of index 2. We call
elements in G0 even and elements in G\G0 odd. Then the group algebra of G is
naturally a superalgebra; we shall denote it by C[G, G0 ] = CG0 ⊕ C[G\G0 ] to make
clear its Z2 -grading and its dependence on G0 . Just as for the usual group algebras,
it is standard to show that C[G, G0 ] is a semisimple superalgebra.
Since elements in a given conjugacy class of G share the same parity (i.e.,
Z2 -grading), it makes sense to talk about even and odd conjugacy classes.
Proposition 3.6. (1) The number of simple C[G, G0 ]-modules coincides with
the number of even conjugacy classes of G.
(2) The number of simple C[G, G0 ]-modules of type Q coincides with the num-
ber of odd conjugacy classes of G.

Proof. Set A = C[G, G0 ]. By (3.4) CG can be decomposed as a direct sum of


simple superalgebras:

m ⊕
q
A∼
= M(ri |si ) ⊕ Q(n j ).
i=1 j=1
3.1. Generalities for associative superalgebras 89

The center of the algebra |(M(r|s)| is ∼


( CI, and
= ) the center of the algebra |Q(n)| is
0 In
of dimension 1|1, spanned by I2n and . Hence, dim Z(|A|)0̄ = m + q and
In 0
dim Z(|A|)1̄ = q.
On the other hand, note that |A| is the usual group algebra CG. Hence the
class sums for the even (respectively, odd) conjugacy classes of G form a basis of
Z(|A|)0̄ (respectively, Z(|A|)1̄ ). The proposition is proved. 

e which contains a subgroup G


For a finite group G e0 of index 2, let us further
e
assume that G contains a central element z which is even and of order 2. Setting
e
G = G/{1, z}, we have a short exact sequence of groups
θ
e −→
1 −→ {1, z} −→ G G −→ 1.
Let C be a conjugacy class of G. Depending on whether or not an element xe in
θ−1 (C) is conjugated to ze x, θ−1 (C) is either a single conjugacy class of Ge or it splits
e
into two conjugacy classes of G; in the latter case, C is called a split conjugacy
class, and either conjugacy class in θ−1 (C) will also be called split. An element
x ∈ G is called split if the conjugacy class of x is split. If we denote θ−1 (x) =
{x̃, zx̃}, then x is split if and only if x̃ is not conjugate to zx̃.
Denote by CG− the quotient superalgebra of C[G, eGe0 ] by the ideal generated
e
by z + 1. Now z acts as ±I on any simple G-module by Schur’s Lemma. Thus, we
have an isomorphism of left C[G, eGe0 ]-modules

(3.5) eG
C[G, e0 ] ∼
= C[G, G0 ] CG− ,
according to the eigenvalues of z. Since z is central, (3.5) is an isomorphism of
superalgebras as well. Hence, CG− is naturally a semisimple superalgebra. A
spin C[G,eG e0 ]-module M means that a Z2 -graded module over the superalgebra
eG
C[G, e0 ] on which z acts on by −I, and we will sometimes simply refer to as a spin
e
G-module (with Z2 -grading implicitly assumed). A spin C[G, eG e0 ]-module is then

naturally identified as a CG -module.

Proposition 3.7. eG
(1) The number of simple spin C[G, e0 ]-modules equals the
number of even split conjugacy classes of G.
(2) The number of simple spin C[G, eG e0 ]-modules of type Q equals the number
of odd split conjugacy classes of G.

Proof. Let us denote by A = C[G,eG e0 ], and further denote by m̄ (respectively, q̄)


the number of simple spin A-modules of type M (respectively, type Q). The central
element z acts on Z(|A|) by multiplication with eigenvalues ±1, and the (−1)-
eigenspace Z− (|A|) can be naturally identified with Z(|CG− |), see (3.5). Hence,
using (3.5) and a decomposition of C[G,eG e0 ] as as direct sum of simples, we can
90 3. Schur duality

show by a similar argument as for Proposition 3.6 that

(3.6) dim Z− (|A|)0̄ = m̄ + q̄, dim Z− (|A|)1̄ = q̄.

List all conjugacy classes of G e as follows: D1 , zD1 , . . . , Dℓ , zDℓ , Dℓ+1 , . . . , Dk ,


where Di ∩ zDi = 0/ for 1 ≤ i ≤ ℓ and zDi = Di for ℓ + 1 ≤ i ≤ k. The corresponding
class sums, d1 , zd1 , . . . , dℓ , zdℓ , dℓ+1 , . . . , dk , form a basis for Z(|A|). It follows that
a basis for Z− (|A|) is {di − zdi | 1 ≤ i ≤ ℓ}, and thus, dim Z− (|A|) = ℓ, which is
the number of split conjugacy classes of G. The proposition follows by a division
of the split conjugacy classes by parity and a comparison with (3.6). 

3.2. Schur-Sergeev duality of type A


In this section, we formulate the Schur-Sergeev duality as a double centralizer
property for the commuting actions of the Lie superalgebra gl(V ) and the symmet-
ric group Sd on the tensor space V ⊗d , where V = Cm|n . We obtain a multiplicity-
free decomposition of V ⊗d as a U(gl(V )) ⊗ CSd -module. The character formula
for the irreducible gl(V )-modules arising this way is given in terms of super Schur
polynomials. We then provide a Young diagrammatic interpretation of the degree
of atypicality of the weight λ♮ . We further show that the category of polynomial
gl(m|n)-modules is semisimple.

3.2.1. Schur-Sergeev duality, I. Let g = gl(m|n). Let {ei |i ∈ I(m|n)} be the stan-
dard basis for the natural g-module V = Cm|n . Then V ⊗d is naturally a g-module
by letting

Φd (g).(v1 ⊗ v2 ⊗ . . . ⊗ vd ) =g.v1 ⊗ . . . ⊗ vd + (−1)|g|·|v1 | v1 ⊗ g.v2 ⊗ . . . ⊗ vd


+ . . . + (−1)|g|·(|v1 |+...+|vd−1 |) v1 ⊗ v2 ⊗ . . . ⊗ g.vd ,

where g ∈ g and vi ∈ V are assumed to be Z2 -homogeneous.


On the other hand, let

Ψd ((i, i + 1)).v1 ⊗ . . . vi ⊗ vi+1 ⊗ . . . ⊗ vd


(3.7) = (−1)|vi |·|vi+1 | v1 ⊗ . . . vi+1 ⊗ vi ⊗ . . . ⊗ vd , 1 ≤ i ≤ d − 1,

where (i, j) denotes a transposition in Sd and vi , vi+1 are Z2 -homogeneous.

Lemma 3.8. The formula (3.7) defines a left action of the symmetric group Sd on
V ⊗d . The actions of (gl(m|n), Φd ) and (Sd , Ψd ) on V ⊗d commute with each other.

Proof. It is straightforward to check that the formula (3.7) satisfies the Coxeter
relations for Sd , and hence it defines a left action of Sd .
3.2. Schur-Sergeev duality of type A 91

The verification of the commuting action boils down to the case below when
d = 2 and i = 1 (where we denote s = (1, 2) ∈ S2 , g ∈ gl(m|n), and v1 , v2 ∈ V ):

Ψd (s)Φd (g)(v1 ⊗ v2 )
= Ψd (s)(g.v1 ⊗ v2 + (−1)|g|·|v1 | v1 ⊗ g.v2 )
= (−1)(|g|+|v1 |)|v2 | v2 ⊗ g.v1 + (−1)|g|·|v1 |+|v1 |(|g|+|v2 |) g.v2 ⊗ v1 ,
Φd (g)Ψd (s)(v1 ⊗ v2 ) = (−1)|v1 |·|v2 | (g.v2 ⊗ v1 + (−1)|g|·|v2 | v2 ⊗ g.v1 ).

Clearly, Ψd (s)Φd (g) = Φd (g)Ψd (s). 

We are going to formulate a superalgebra analogue of Schur duality.

Theorem 3.9 (Schur-Sergeev duality, Part I). Let g = gl(m|n). The images of Φd
and Ψd , Φd (U(g)) and Ψd (CSd ), satisfy the double centralizer property, i.e.,

Φd (U(g)) =EndSd (V ⊗d ),
EndU(g) (V ⊗d ) = Ψd (CSd ).

Proof. By Lemma 3.8, we have Φd (U(g)) ⊆ EndSd (V ⊗d ).


We shall proceed to prove that Φd (U(g)) ⊇ EndSd (V ⊗d ).
We have a natural isomorphism of vector superspaces End(V )⊗d ∼ = End(V ⊗d ),
which is then made Sd -equivariant, as both superspaces admit natural Sd -actions.
This isomorphism allows us to identify EndSd (V ⊗d ) ≡ Symd (End(V )), the space
of Sd -invariants in End(V )⊗d .
Denote by Yk , 1 ≤ k ≤ d, the C-span of the super-symmetrization

ϖ(x1 , . . . , xk ) := ∑ σ.(x1 ⊗ . . . ⊗ xk ⊗ 1d−k ),


σ∈Sd

for all xi ∈ End(V ). Note that Yd = Symd (End(V )) ≡ EndSd (V ⊗d ).


Let x̃ = Φd (x) = ∑di=1 1i−1 ⊗ x ⊗ 1d−i , for x ∈ g = End(V ), and denote by Xk ,
1 ≤ k ≤ d, the C-span of x̃1 . . . x̃k for all xi ∈ End(V ).
Claim. We have Yk ⊆ Xk for 1 ≤ k ≤ d.
Assuming the claim is true (in particular for k = d), then the theorem follows
thanks to
EndSd (V ⊗d ) = Yd ⊆ Xd ⊆ Φd (U(g)).

We will prove the claim by induction on k. The case k = 1 holds, thanks to


ϖ(x) = (d − 1)!x̃.
92 3. Schur duality

Assume that Yk−1 ⊆ Xk−1 . Note that ϖ(x1 , . . . , xk−1 ) · x̃k ∈ Xk . On the other
hand, we have
ϖ(x1 , . . . , xk−1 ) · x̃k
= ∑ σ.(x1 ⊗ . . . ⊗ xk−1 ⊗ 1d−k+1 ) · x̃k
σ∈Sd
d ( )
= ∑ ∑ σ. (x1 ⊗ . . . ⊗ xk−1 ⊗ 1d−k+1 ) · (1 j−1 ⊗ xk ⊗ 1d− j ) ,
j=1 σ∈Sd

which can be written as a sum A1 + A2 , where


k−1
A1 = ∑ ϖ(x1 , . . . , x j xk , . . . , xk−1 ) ∈ Yk−1 ,
j=1

and
d
A2 = ∑ ∑ σ.(x1 ⊗ . . . ⊗ xk−1 ⊗ 1 j−k ⊗ xk ⊗ 1d− j )
j=k σ∈Sd
= (d − k + 1)ϖ(x1 , . . . , xk−1 , xk ).

Note that A1 ∈ Xk , since Yk−1 ⊆ Xk−1 ⊆ Xk . Hence, A2 ∈ Xk , and so, Yk ⊆ Xk .


This proves the claim.
Hence, Φd (U(g)) = EndSd (V ⊗d ) = EndB (V ⊗d ), for B := Ψd (CSd ).
Note that B is a semisimple algebra, and so the assumption of Proposition 3.5
is satisfied. Therefore, we have EndU(g) (V ⊗d ) = Ψd (CSd ). 

3.2.2. Schur-Sergeev duality, II. Recall from Definition 2.10 that an (m|n)-hook
partition µ = (µ1 , µ2 , . . .) is a partition with λm+1 ≤ n. Given an (m|n)-hook parti-
tion µ, we denote by µ+ = (µm+1 , µm+2 , . . .) and write its conjugate partition, which
is necessarily of length ≤ n, as ν = (µ+ )′ = (ν1 , . . . , νn ). We recall the weight
defined in (2.4)
µ♮ = µ1 δ1 + . . . + µm δm + ν1 ε1 + . . . + νn εn .

Denote by Pd (m|n) the set of all (m|n)-hook partitions of size d and let
P(m|n) = ∪d≥0 Pd (m|n) = {λ | λ are partitions with λm+1 ≤ n}.
Then, Pd (m) = P(d, m|0) is the set of partitions of d of length ≤ m. Accordingly,
we let
Pd = {λ | λ are partitions of d}.
We denote by L(λ♮ ) for λ ∈ P(m|n) the simple g-module of highest weight λ♮ with
respect to the standard Borel subalgebra. For a partition λ of d, we denote by
Sλ the Specht module of Sd . For example, S(d) is the trivial representation, and
d
S(1 ) = sgnd is the sign representation.
3.2. Schur-Sergeev duality of type A 93

Theorem 3.10 (Schur-Sergeev duality, Part II). As a U(gl(m|n)) ⊗ CSd -module,


we have


(Cm|n )⊗d = L(λ♮ ) ⊗ Sλ .
λ∈Pd (m|n)

We will refer to a U(gl(m|n)) ⊗ CSd -module as a (gl(m|n), Sd )-module. Sim-


ilar conventions also apply to similar setups later on.

Proof. Let V = Cm|n and W = V ⊗d . Note that Ψd (CSd ) is a semisimple algebra


and all its simple modules are of type M. By Proposition 3.5 and Theorem 3.9,
we have a multiplicity-free decomposition of the (gl(m|n), Sd )-module W of the
following form:

W ≡ V ⊗d ∼= L[λ] ⊗ Sλ ,
λ∈Pd (m|n)

where is some simple gl(m|n)-module associated to λ, whose highest weight


L[λ]
(with respect to the standard Borel) is to be determined. Also to be determined is
the index set Pd (m|n) = {λ ⊢ d | L[λ] ̸= 0}.
First we need to prepare some notations.
Let CP(m|n) be the set of pairs ν|µ of compositions ν = (ν1 , . . . , νm ) and µ =
(µ1 , . . . , µn ), and let
CPd (m|n) = {ν|µ ∈ CP(m|n) | ∑ νi + ∑ µ j = d}.
i j

We have the following weight space decomposition with respect to the Cartan sub-
algbra of diagonal matrices h ⊂ gl(m|n):

(3.8) W= Wν|µ ,
ν|µ∈CPd (m|n)

where Wν|µ has a linear basis ei1 ⊗ . . . ⊗ eid , with the indices satisfying the following
equality of sets:
{i1 , . . . , id } = {1, . . . , 1, . . . , m, . . . , m; 1, . . . , 1, . . . , n, . . . , n}.
| {z } | {z } | {z } | {z }
ν1 νm µ1 µn

Let
Sν|µ := Sν1 × . . . × Sνm × Sµ1 × . . . × Sµn .
⊗µ ⊗µ
The span of the vector eν|µ := e⊗ν
1
1
⊗ . . . ⊗ e⊗ν
m ⊗ e1 ⊗ . . . ⊗ en
m 1 n
can be identified
with the Sν|µ -module 1ν ⊗ sgnµ . Since the orbit Sd .eν|µ spans Wν|µ we have a
surjective Sd -homomorphism from IndS Sν|µ (1ν ⊗ sgnµ ) onto Wν|µ by the Frobenius
d

reciprocity. By counting the dimensions we have an Sd -isomorphism:


(3.9) Wν|µ ∼ S
= IndSdν|µ (1ν ⊗ sgnµ ).
94 3. Schur duality

Let us denote the decomposition of Wν|µ into irreducibles by



Wν|µ ∼
= Kλ,ν|µ Sλ , for Kλ,ν|µ ∈ Z+ .
λ

Let λ be a partition which is identified with its Young diagram. Recall I(m|n)
is totally ordered by (1.3). A supertableau T of shape λ, or a super λ-tableau
T , is an assignment of an element in I(m|n) to each box of the Young diagram λ
satisfying the following conditions:
(HT1) The numbers are weakly increasing along each row and column.
(HT2) The numbers from {1, . . . , m} are strictly increasing along each column.
(HT3) The numbers from {1, . . . , n} are strictly increasing along each row.
Such a T is said to have content ν|µ ∈ CP(m|n) if ī (1 ≤ i ≤ m) appears νi times and
j (1 ≤ j ≤ n) appears µ j times. Denote by HT(λ, ν|µ) the set of super λ-tableaux
of content ν|µ. In the case n = 0, a supertableau becomes a usual (semistandard)
tableau.
Lemma 3.11. We have Kλ,ν|µ = #HT(λ, ν|µ).

Proof. Recall that IndS ∼ ⊕ Kλ,ν|µ Sλ .


Sν|µ (1ν ⊗ sgnµ ) =
d
λ
First assume that µ = 0,/ and we prove the formula by induction on the length
r = ℓ(ν). A (semistandard) tableau T of shape λ and content ν gives rise to a
sequence of partitions 0/ = λ0 ⊂ λ1 ⊂ . . . ⊂ λr = λ such that λi has the shape given
by the parts of T with entries ≤ i, and λi /λi−1 has νi boxes for each i. This sets up
a bijection between HT(λ, ν) and the set of such sequences of partitions. Denote

ν = (ν1 , . . . , νr−1 ). We have IndSeν1 1d1 ∼
Sd
d1 = d − νr and e = ρ⊢d1 Kρ,eν Sρ , where
Kρ,eν = #HT(ρ, e ν) by induction hypothesis. Now the induction step is simply the
following representation theoretic version of the Pieri’s formula (A.21) (which can
be seen using the Frobenius characteristic map and (A.24)): for a partition ρ ⊢ d1 ,

IndS
Sd
d
×Sνr (S
ρ
⊗ 1νr ) ∼
= Sλ ,
1
λ

where λ is such that λ/ρ is a horizontal strip of νr boxes (that is, λ/ρ is a skew
diagram of size νr whose columns all have length at most one).
Then using the above special case (for µ = 0) / as the initial step, we complete
the proof in the general case by induction on the length of µ, in which the induction
step is exactly the conjugated Pieri’s formula. 
Lemma 3.12. Let λ ∈ Pd and ν|µ ∈ CPd (m|n). Then Kλ,ν|µ = 0 unless λ ∈ Pd (m|n).

Proof. By the identity Kλ,ν|µ = #HT(λ, ν|µ), it suffices to prove that if a super
λ-tableau T of content ν|µ exists, then λm+1 ≤ n.
3.2. Schur-Sergeev duality of type A 95

By applying the supertableau condition (HT2) to the first column of T , we


see that the first entry k ∈ I(m|n) in row (m + 1) satisfies k > 0. Applying the
supertableau condition (HT3) to the (m + 1)st row, we conclude that λm+1 ≤ n. 

Recall that Pd (m|n) = {λ ⊢ d | L[λ] ̸= 0}. It follows by Lemma 3.12 that


Pd (m|n) ⊂ Pd (m|n). On the other hand, given λ ∈ Pd (m|n), clearly a super λ-
tableau exists, e.g., we can fill in the numbers 1, . . . , m on the first m rows of λ row
by row downward, and then for the (possibly) remaining rows of λ, we fill in the
numbers 1, . . . , n column by column from left to right. This distinguished super
λ-tableau will be denoted by Tλ . Hence, we have proved that
Pd (m|n) = Pd (m|n).
⊕ [λ]
For a given λ ∈ Pd (m|n), we have L[λ] = ν|µ∈CPd (m|n) Lν|µ . Among all the
contents of super λ-tableaux, the one for Tλ corresponds to a highest weight by the
supertableau conditions (HT1-3). Hence, we conclude that L[λ] = L(λ♮ ), the simple
g-module of highest weight λ♮ . This completes the proof of Theorem 3.10. 
Remark 3.13. (1) For n = 0, the Schur-Sergeev duality reduces to the usual
Schur duality. If in addition d = 2, then (Cm )⊗2 = S2 (Cm ) ⊕ ∧2 (Cm ).
This fits well with the well-known fact that as gl(m)-modules S2 (Cm )
and ∧2 (Cm ) are, respectively, irreducible of highest weights 2Λ1 and Λ2 ,
where Λi denotes the ith fundamental weight.

(2) For m = 0, λ♮ = λ′ (the conjugate partition), and Sλ = Sλ ⊗ sgn. In this
case, the Schur-Sergeev duality reduces to a version of Schur duality
twisted by the sign representation of Sd , i.e., as a (gl(m), Sd )-module,

(Cm )⊗d ∼

= L(µ) ⊗ Sµ .
µ∈Pd (m)

(3) If d ≤ mn + m + n, then Pd (m|n) is the set of all partitions of d, and every


simple Sd -module appears in the Schur-Sergeev duality decomposition.
(4) For d = 2, the Schur-Sergeev duality reads that (Cm|n )⊗2 = S2 (Cm|n ) ⊕
∧2 (Cm|n ), where S2 and ∧2 are understood in the super sense. In particu-
lar, as an ordinary vector space,
S2 (Cm|n ) = S2 (Cm ) ⊕ (Cm ⊗ Cn ) ⊕ ∧2 (Cn ).

3.2.3. The character formula. We shall compute the character of the gl(m|n)-
module L(λ♮ )

∑ dim L(λ♮ )ν|µ ∏ xiνi y j j .


E Em̄m̄ E11 µ
chL(λ♮ ) := tr |L(λ♮ ) x1 1̄1̄ . . . xm y1 . . . yEn nn =
ν|µ∈CP(m|n) i, j

Recall that mν denotes the monomial symmetric function associated to a parti-


tion ν (cf. Appendix A), and that hsν denotes the super Schur function (A.35).
96 3. Schur duality

Theorem 3.14. Let λ be an (m|n)-hook partition, i.e., a partition with λm+1 ≤ n.


Let x = (x1 , . . . , xm ) and y = (y1 , . . . , yn ). Then the following character formula
holds:
chL(λ♮ ) = hsλ (x; y).

Proof. It follows from Theorem 3.10 and (3.8) that


⊕ ⊕
Wν|µ ∼
= L(λ♮ ) ⊗ Sλ .
ν|µ∈CPd (m|n) λ∈Pd (m|n)

E
Em̄m̄ yE11 . . . yEnn and
Let us apply simultaneously the trace operator tr |L(λ♮ ) x1 1̄1̄ . . . xm 1 n
the Frobenius characteristic map chF (see Appendix A.1.4) to both sides of the
above isomorphism. Then, summing over all d ≥ 0 and using (3.9) and (A.24), we
have that

∑ mν (x)mµ (y)hν (z)eµ (z) = ∑ chL(λ♮ )sλ (z),


ν|µ∈CP(m|n) λ∈P(m|n)

where z = (z1 , z2 , . . .) is infinite. By the identities in (A.11), we have


∞ m n
1 + y j zk
∑ mν (x)mµ (y)hν (z)eµ (z) = ∏ ∏ ∏ = ∑ hsλ (x; y)sλ (z).
ν|µ∈CP(m|n) k=1 i=1 j=1 1 − xi zk λ∈P(m|n)

Now the theorem follows by comparing the above two equations and noting the
linear independence of sλ (z). 

Given a super λ-tableau T of content ν|µ ∈ CP(m|n), we denote by


(x|y)T := x1ν1 x2ν2 · · · xm
νm 1 µ
y1 · · · yµnn .
We have the following alternative character formula.

Theorem 3.15. Let λ be an (m|n)-hook partition. Then,


(3.10) chL(λ♮ ) = ∑(x|y)T ,
T

where the summation is taken over all super λ-tableaux T . Also, we have
(3.11) chL(λ♮ ) = ∑ #HT(λ, ν|µ) mν (x)mµ (y),
ν,µ

where the summation is over partitions ν and µ of length ≤ m and ≤ n, respectively,


such that |ν| + |µ| = d.

Proof. The formula (3.10) is simply a reformulation of Theorem 3.14 by the defi-
nition of super Schur functions in (A.35) and the combinatorial formula for (skew)
Schur functions (A.18).
3.2. Schur-Sergeev duality of type A 97

Let us collect the same monomials together in the sum (3.10). Regarding L(λ♮ )
as a module over gl(m) × gl(n), we observe that chL(λ♮ ) is symmetric with respect
to x1 , . . . , xm and symmetric with respect to y1 , . . . , yn . Hence, we must have
#HT(λ, e
ν|e
µ) = #HT(λ, ν|µ)
if e
ν and e
µ are compositions obtained by rearranging the parts from ν and µ, respec-
tively. Now (3.11) follows from this observation and (3.10). 
Corollary 3.16. The following weight multiplicity formula for the module L(λ♮ )
holds:
dim L(λ♮ )ν|µ = #HT(λ, ν|µ).

Example 3.17. The standard basis vectors for S2 (Cm|n ) are ei ⊗ e j + (−1)|i|| j| e j ⊗
ei , where i, j ∈ I(m|n) satisfy i ≤ j < 0, i < 0 < j, or 0 < i < j (cf. Remark 3.13 (4)).
They are in bijection with the super tableaux of shape λ = (2):

i j

This is compatible with the identification S2 (Cm|n ) ∼


= L(gl(m|n), 2Λ1 ).

3.2.4. The classical Schur duality. We sketch below a more standard argument
for the standard Schur duality on (Cn )⊗d , which emphasizes the decomposition of
W as a gl(n)-module.
Given a composition (or a partition) µ of d of length ≤ n, we denote by Wµ the
µ-weight space of the gl(n)-module. Clearly Wµ has a basis
(3.12) ei1 ⊗ . . . ⊗ eid , where {i1 , . . . , id } = {1, . . . , 1, . . . , n, . . . , n}.
| {z } | {z }
µ1 µn

As a (gl(n), Sd )-module,

W∼
= L(λ) ⊗U λ ,
λ

where U λ := Homgl(n) (L(λ),W ) ∼


+
= Wλn (the space of highest weight vectors in W
of weight λ). Only λ ∈ Pd (n) can be highest weights of gl(n)-modules that appear
in the decomposition of (Cn )⊗d , and every such λ indeed appears as L(λ) is clearly
′ ′
a summand of the submodule ∧λ1 (Cn ) ⊗ ∧λ2 (Cn ) ⊗ · · · of W .
Note that Pd (n) has two interpretations: one as the polynomial weights for
gl(n) and the other as compositions of d in at most n parts. A remarkable fact is
that the partial order on weights induced by the positive roots of gl(n) coincides
with the dominance partial order ≥ on compositions.
Since W = ⊕µWµ = ⊕µ,λ:λ≥µ L(λ)µ ⊗U λ , we conclude that, as a Sd -module,

(3.13) Wµ ∼
= dim L(λ)µU λ .
λ≥µ
98 3. Schur duality

On the other hand, Sd acts on the basis (3.12) of Wµ transitively, and the stabilizer
⊗µ ⊗µ
of the basis element e1 1 ⊗ . . . ⊗ en n is the Young subgroup Sµ . Therefore we
have

(3.14) Wµ ∼ S
= Ind d 1µ =
Sµ Kλµ Sλ ,
λ≥µ

where Kλµ is the Kostka number which satisfies Kλλ = 1. It is well known that Kλµ
is equal to the number of semistandard λ-tableaux of content µ.
By the double centralizer property (see Proposition 3.5 and Theorem 3.9), U λ
has to be an irreducible Sd -module for each λ. We compare the interpretations
(3.13) and (3.14) of Wµ in the special case when µ is dominant (i.e., a partition).
One by one downward along the dominance order starting with µ = (d), this pro-
vides the identification U µ = Sµ for every µ, and moreover, we obtain the well-
known equality dim L(λ)µ = Kλµ .

3.2.5. Degree of atypicality of λ♮ . In this section, we provide a Young diagram-


matic interpretation of the degree of atypicality of the weight λ♮ , for an (m|n)-hook
partition λ.
Up to a shift by − 12 (m + n + 1)1m|n (which is irrelevant in all applications), the
Weyl vector ρ for the standard positive system of gl(m|n) from Lemma 1.19 can
be written as
m n
ρ = ∑ (m − i + 1)δi − ∑ jε j .
i=1 j=1

We introduce an integer iλ , with 0 ≤ iλ ≤ min{m, n}, to stand for the smallest


integer i such that the (m − i, n − i)-th box belongs to the diagram λ.
Example 3.18. Let λ = (7, 4, 2, 2, 1, 1) with m = 4 and n = 5. Then iλ is the number
of (m − i, n − i) boxes that do not lie in the diagram of λ, for i = 0, 1, . . . , min{m, n}.
Such boxes are marked with crosses in this example. So iλ = 2.
n − iλ n

m − iλ
@
@
m @
@

Lemma 3.19. For 0 ≤ j ≤ iλ − 1, we have λm− j ≤ n − iλ ≤ λm−iλ .

Proof. The lemma is clearly equivalent to the claim that λm−iλ +1 ≤ n − iλ ≤ λm−iλ .
The latter is evident from the diagram in Example 3.18, as iλ is equal to the number
3.2. Schur-Sergeev duality of type A 99

of boxes of coordinates (m − i, n − i) that do not lie in the diagram of λ, for i =


0, 1, . . . , min{m, n}. 

Recall the degree of atypicality #µ for a weight µ ∈ h∗ from Definition 2.28.

Proposition 3.20. Let λ be an (m|n)-hook partition. Then, we have iλ = #λ♮ .

Proof. We compute that


m n
(3.15) λ♮ + ρ = ∑ (λi + m − i + 1)δi − ∑ ( j − ν j )ε j .
i=1 j=1

We observe that the sequence {λ1 +m, λ2 +m−1, . . . , λm +1} is strictly decreasing,
while the sequence {1 − ν1 , 2 − ν2 , . . . , n − νn } is strictly increasing.
Suppose that iλ = 0. This is equivalent to saying that λm ≥ n, by Lemma 3.19.
Thus,
λi + m − i + 1 ≥ λm + 1 > j − ν j , ∀1 ≤ j ≤ n, 1 ≤ i ≤ m.
It follows by Definition 2.28 that #λ♮ = 0.
Now suppose that iλ > 0. Then, λm = n − j0 < n for some 0 < j0 ≤ n. This
implies that νn = · · · = νn− j0 +1 = 0, and thus
{n − j0 + 1 − νn− j0 +1 , . . . , n − νn } = {n − j0 + 1, . . . , n}.
It follows by Lemma 3.19 that, for 0 ≤ j ≤ iλ − 1,
n − j0 + 1 = λm + 1 ≤ λm− j + j + 1 ≤ n − iλ + j + 1 ≤ n.
Thus, in the set {n − j0 + 1 − νn− j0 +1 , . . . , n − νn } = {n − j0 + 1, . . . , n}, there is a
unique element that is equal to λm− j + j + 1, for 0 ≤ j ≤ iλ − 1. Hence, #λ♮ ≥ iλ .
Finally, for i = 1, . . . , m − iλ , Lemma 3.19 implies that
λi + m − i + 1 ≥ λm−iλ + iλ + 1 ≥ (n − iλ ) + iλ + 1 = n + 1.
Thus, any such λi + m − i + 1, for i = 1, . . . , m − iλ , cannot be equal to an element
of the form j − ν j , for j = 1, . . . , n, whence #λ♮ = iλ . 

In light of (3.15), we define


Hλ♮ = {λi + m − i + 1 | 1 ≤ i ≤ m},
Tλ♮ = { j−ν j | 1 ≤ j ≤ n}, Iλ♮ = Hλ♮ ∩ Tλ♮ .

The following corollary can be read off from the proof of Proposition 3.20.

Corollary 3.21. Let λ be an (m|n)-hook partition. Then Iλ♮ consists of precisely


the largest #λ♮ numbers in {1, 2, . . . , n} that are not in Tλ♮ \ Iλ♮ .
100 3. Schur duality

3.2.6. Category of polynomial modules. Recall that, given a Lie (super)algebra


G and G-modules M and N, the vector space Ext1G (M, N) classifies (up to equiva-
lence) the short exact sequences of G-modules of the form
(3.16) 0 −→ N −→ E −→ M −→ 0.
Recall from Section 2.2.2 the notion of central characters χν : Z(g) → C.
Proposition 3.22. Let λ, µ be (m|n)-hook diagrams. Then,
(1) χλ♮ = χµ♮ if and only if λ = µ.
( )
(2) Ext1gl(m|n) L(gl(m|n), λ♮ ), L(gl(m|n), µ♮ ) = 0.

Proof. (1) Suppose that χλ♮ = χµ♮ . Then by Theorem 2.27 we have
Hλ♮ \ Iλ♮ = Hµ♮ \ Iµ♮ , Tλ♮ \ Iλ♮ = Tµ♮ \ Iµ♮ .
Thus, it suffices to show that we can reconstruct λ♮ from the sets Hλ♮ \ Iλ♮ and
Tλ♮ \ Iλ♮ . To that end, note that λ♮ + ρ and λ♮ can be recovered from the sets Hλ♮ and
Tλ♮ , which in turn are determined by the three sets Iλ♮ , Hλ♮ \ Iλ♮ and Tλ♮ \ Iλ♮ . But by
Corollary 3.21 the set Iλ♮ is determined from the set Tλ♮ \ Iλ♮ . This proves (1).
(2) Consider a short exact sequence of the form (3.16) with M = L(gl(m|n), λ♮ )
and N = L(gl(m|n), µ♮ ). First assume λ ̸= µ. It follows by (1) that there exists a
central element z such that χλ♮ (z) ̸= χµ♮ (z). Then, ker(z − χλ♮ (z)) is a nonzero
proper submodule of E, which must be isomorphic to M as it cannot be isomorphic
to N. Hence the short exact sequence splits. Now assume λ = µ. Then by weight
consideration, the two-dimensional µ♮ -weight subspace of E has to be a highest
weight space. Thus, E contains two highest weight submodules of highest weight
µ♮ that must intersect trivially. Hence the short exact sequence splits in this case as
well. 
Definition 3.23. A weight µ = ∑m i=1 ai δi + ∑ j=1 b j ε j is called a polynomial weight
n

for gl(m|n) if all ai , b j are nonnegative integers. A gl(m|n)-module M is called a


polynomial module if the following conditions are satisfied:
(1) M is h-semisimple;
(2) Every weight of M is a polynomial weight;
(3) There exist weights λ1 , λ2 , . . . , λk such that if γ is a weight in M, then
γ ∈ λi − ∑α∈Φ+ Z+ α, for some i (Here Φ+ is the standard positive system).
Proposition 3.24. An irreducible highest weight gl(m|n)-module M is a polyno-
mial module if and only if M ∼
= L(gl(m|n), µ♮ ), for some (m|n)-hook partition µ.

Proof. Assume that M ∼ = L(gl(m|n), µ♮ ), for some (m|n)-hook partition µ. Denote


by d = |µ|. Then by Theorem 3.10, L(gl(m|n), µ♮ ) is a direct summand of (Cm|n )⊗d ,
which is clearly a polynomial module.
3.3. Representation theory of the algebra Hn 101

Now assume that M is an irreducible polynomial gl(m|n)-module, which by


Definition 3.23(3) is isomorphic to L(gl(m|n), λ) for some λ = ∑m i=1 λi δi + ∑ j=1 b j ε j ,
n

with all λi , b j ∈ Z+ . Since L(gl(m) ⊕ gl(n), λ) is a polynomial module, we must


have λi ≥ λi+1 and µ j ≥ µ j+1 , for all possible i, j. Suppose that λ ̸= µ♮ , for
any (m|n)-hook partition µ. Then we have λm = k − 1 and bk > 0, for some
1 ≤ k ≤ n. We apply the sequence of odd reflections corresponding to the odd
roots δm − ε1 , . . . , δm − εk−1 to obtain a new Borel subalgebra e
b from the standard
e
one. By applying Lemma 1.36 repeatedly, we compute the b-highest weight to be
eλ = λ − (k − 1)δm + ε1 + . . . + εk−1 . Observe by a repeated use of Lemma 1.26 that
δm − εk is a simple root of e b. Let weλ be a nonzero e
b-highest weight vector. Then
the vector e −δm +εk λ
we is nonzero in M, and its weight eλ − δm + εk has −1 as the
coefficient for δm , thanks to λm = k − 1. This contradicts the assumption that M is
a polynomial module. 

Theorem 3.25. The category of polynomial gl(m|n)-modules is a semisimple ten-


sor category.

Proof. If M is a polynomial module of gl(m|n), then M is a direct sum of irre-


ducible polynomial modules by Propositions 3.22 and 3.24. Thus, the category of
polynomial modules is semisimple.
It remains to show that the tensor product of any two irreducible polynomial
gl(m|n)-modules is a direct sum of irreducible polynomial gl(m|n)-modules. To
that end, note that by Theorem 3.10 any irreducible polynomial module is a direct
summand of (Cm|n )⊗d , for some d ≥ 0, and furthermore any such tensor power
is a direct sum of irreducible polynomial modules. Now take L(gl(m|n), λ♮ ) ⊆
(Cm|n )⊗d and L(gl(m|n), µ♮ ) ⊆ (Cm|n )⊗l , for (m|n)-hook partitions λ and µ. Then

L(gl(m|n), λ♮ ) ⊗ L(gl(m|n), µ♮ ) ⊆ (Cm|n )⊗k ⊗ (Cm|n )⊗l ∼


= (Cm|n )⊗d+l .

Since (Cm|n )⊗d+l is a direct sum of irreducible polynomial modules, so is the sub-
module L(gl(m|n), λ♮ ) ⊗ L(gl(m|n), µ♮ ). 

3.3. Representation theory of the algebra Hn


In this section, we develop systematically the representation theory of an algebra
Hn , which is equivalent to the spin representation theory of a distinguished double
cover Ben of the hyperoctahedral group Bn . We classify the split conjugacy classes
in Ben . We then define a characteristic map using the character table for the simple
spin modules of Ben , analogous to Frobenius characteristic map for the symmetric
groups. The image of the irreducible spin characters of Ben under the characteristic
map are shown to be Schur Q-functions (up to some 2-powers).
102 3. Schur duality

3.3.1. A double cover. Let Πn be the finite group generated by ai (i = 1, . . . , n)


and the central element z subject to the relations

(3.17) a2i = 1, z2 = 1, ai a j = za j ai (i ̸= j).

The symmetric group Sn acts on Πn by σ(ai ) = aσ(i) , σ ∈ Sn . The semidirect


product Ben := Πn o Sn admits a natural finite group structure and will be called
the twisted hyperoctahedral group. Explicitly, the multiplication in Ben is given by

(a, σ)(a′ , σ′ ) = (aσ(a′ ), σσ′ ), a, a′ ∈ Πn , σ, σ′ ∈ Sn .

Since Πn /{1, z} ≃ Zn2 , the group Ben is a double cover of the hyperoctahedral
group Bn := Zn2 o Sn , and the order |Ben | is 2n+1 n!. That is, we have a short exact
sequence of groups
θn
1 −→ {1, z} −→ Ben −→ Bn −→ 1,

where θn sends each ai to the generator bi of the ith copy of Z2 in Bn .


We define a Z2 -grading on the group Ben by setting the degree of each ai to
be 1̄ and the degree of elements in Sn to be 0̄. Hence the group Ben fits into the
e in Section 3.1.4. The group Bn inherits a Z2 -grading from Ben
general setting of G
via the homomorphism θn . This induces (parity) epimorphisms p : Ben → Z2 and
p : Bn → Z2 .
The conjugacy classes of the group Bn (a special case of a wreath product) can
be described as follows, cf. Macdonald [Mac]. It is convenient to identify Z2 as
{+, −} with + being the identity element. Given a cycle t = (i1 , . . . , im ), we call
the set {i1 , . . . , im } the support of t, denoted by supp(t). The subgroup Zn2 of Bn
consists of elements of the form bI := ∏i∈I bi for I ⊂ {1, . . . , n}. Each element
bI σ ∈ Bn can be written as a product (unique up to reordering)

bI σ = (bI1 σ1 )(bI2 σ2 ) . . . (bIk σk ),

where σ ∈ Sn is a product of disjoint cycles σ = σ1 . . . σk , and Ia ⊂ supp(σa ) for


each 1 ≤ a ≤ k; bIa σa is called a signed cycle of bI σ. The cycle-product of each
signed cycle bIa σa is defined to be the element ∏i∈Ia bi ∈ Z2 (which can be con-

veniently thought as a sign + or −). Let m+ i (respectively, mi ) be the number
of i-cycles of bI σ with associated cycle-product being + (respectively, −). Then

ρ+ = (imi )i≥1 and ρ− = (imi )i≥1 are partitions such that |ρ+ | + |ρ− | = n. The pair
+

of partitions (ρ+ , ρ− ) will be called the type of the element bI σ.


The following is the basic fact on the conjugacy classes of Bn , cf. [Mac, I,
Appendix B].

Lemma 3.26. Two elements of Bn are conjugate if and only if their types are the
same.
3.3. Representation theory of the algebra Hn 103

We shall denote by Cρ+ ,ρ− the conjugacy class of type (ρ+ , ρ− ). Note that if
(bI , σ) ∈ Cρ+ ,ρ− , then Cρ+ ,ρ− is even (respectively, odd) if |I| is even (respectively,
odd).
Example 3.27. Let τ = (1, 2, 3, 4)(5, 6, 7)(8, 9), σ = (1, 3, 8, 6)(2, 7, 9)(4, 5) ∈ S10 .
It is straightforward to check that both x = ((+, +, +, −, +, +, +, −, +, −), τ) and
y = ((+, −, −, −, +, −, −, −, +, −), σ) in B10 have the same type
(ρ+ , ρ− ) = ((3), (4, 2, 1)).
Thus, x is conjugate to y in B10 .

3.3.2. Split conjugacy classes in Ben . A partition λ = (λ1 , . . . , λℓ ) of length ℓ is


called strict if λ1 > λ2 > . . . > λℓ , and it is called odd if each part λi is odd. We
denote by SPn the set of all strict partitions of n, and by OPn the set of all odd
partitions of n. Moreover, we denote
∪ ∪
SP = SPn , OP = OPn .
n≥0 n≥0

Let
P+
n = {λ ∈ Pn | ℓ(λ) is even},
P−
n = {λ ∈ Pn | ℓ(λ) is odd}.
Given σ ∈ Sn of cycle type µ, we denote by
d(σ) = n − ℓ(µ).

For an (ordered) subset I = {i1 , i2 , . . . , im } of {1, . . . , n}, we denote


aI = ai1 i2 ...im = ai1 ai2 . . . aim .
/ then aI aJ = z|I||J| aJ aI . Also we can
It follows that p(aI ) ≡ |I| mod 2. If I ∩ J = 0,
easily show by induction that
(3.18) ai1 i2 ...im = zd(s) as(i1 )s(i2 )...s(im )
for a permutation s such that s fixes the letters other than i1 , i2 , . . . , im .
We can write a general element of Ben as
zk aI s = zk (aI1 s1 ) . . . (aIq sq ),
and s = s1 . . . sq is a cycle decomposition of s and I j ⊂ supp(s j ). We denote by J c
the complement of a subset J ⊆ {1, . . . , n}.

Lemma 3.28. Let aI s = (aI1 s1 ) . . . (aIq sq ) be an element of Ben in its cycle decom-
position. Let J ⊆ supp(s1 ) ∩ I1c . Then
(aJ s1 )(aI s)(aJ s1 )−1 = zd(s1 )+|J||I| aI s.
104 3. Schur duality

Proof. Observe that a2I = z(|I|−1)|I|/2 for any subset I. For k > 1 we have

(aJ s1 )(aIk sk )(aJ s1 )−1 = z|J||Ik | aIk sk .


Therefore it remains to see that
(aJ s1 )(aI1 s1 )(aJ s1 )−1 = z(|J|−1)|J|/2 aJ as1 (I1 ∪J) s1
= z(|J|−1)|J|/2+d(s1 ) aJ a(I1 ∪J) s1
= z(|J|−1)|J|/2+d(s1 )+|J||I1 | a2J aI1 s1
= z|J||I1 |+d(s1 ) aI1 s1 ,
where we have used the fact that supp(s1 ) ⊇ I1 ∪ J and (3.18). 
Theorem 3.29. The conjugacy class Cρ+ ,ρ− in Bn splits if and only if
(1) For even Cρ+ ,ρ− , we have ρ+ ∈ OPn and ρ− = 0;
/
(2) For odd Cρ+ ,ρ− , we have ρ+ = 0/ and ρ− ∈ SP−
n.

Proof. Assume that Cρ+ ,ρ− is an even conjugacy class such that ρ+ ̸∈ OP. Then
θ−1
n (Cρ+ ,ρ− ) contains an element aI s with a signed cycle decomposition of the form

aI s = s1 (aI2 s2 ) . . . (aIp s p ),
where s1 = (1, 2, . . . , r) for r = 2k even, I1 = 0/ and |I| is even. Consider the element
x = a12...r (1, 2, . . . , r) ∈ Ben . By Lemma 3.28 we have
x(aI s)x−1 = z(2k−1)+2k·|I| aI s = zaI s.
Therefore, if an even conjugacy class Cρ+ ,ρ− splits then ρ+ ∈ OP.
Assume that Cρ+ ,ρ− is an even conjugacy class such that ρ− ̸= 0. / Then ρ−
−1
contains at least two parts, and θn (Cρ+ ,ρ− ) contains an element of the form
aI s = (ai1 s1 )(ai2 s2 )(aI3 s3 ) . . . (aIp s p ),
where i1 ∈ supp(s1 ), i2 ∈ supp(s2 ). If both s1 and s2 are of cycle length 1, then
(i1 , i2 )(aI s)(i1 , i2 )−1 = zaI s. Assume that the order ord(s1 ) is ≥ 2, so s−1
1 (i1 ) =

i1 ̸= i1 . Then
(ai1 s1 )−1 aI s(ai1 s1 )
= (ai1 s1 )−1 (ai1 s1 ai2 s2 )(ai1 s1 )(aI3 s3 ) . . . (aIp s p )
= ai2 s2 (ai1 s1 )(aI3 s3 ) . . . (aIp s p )
= z(ai1 s1 )(ai2 s2 )(aI3 s3 ) . . . (aIp s p )
= zaI s.
Hence, if an even conjugacy class Cρ+ ,ρ− splits then ρ− = 0.
/ Together with the
above, we have shown that an even split conjugacy class should satisfy (1).
3.3. Representation theory of the algebra Hn 105

/ Then,
Now assume that Cρ+ ,ρ− is an odd conjugacy class such that ρ+ ̸= 0.
θ−1
n (Cρ+ ,ρ− )
contains an element aI s with the signed cycle decomposition of the
form
aI s = (s1 )(aI2 s2 ) . . . (aIq sq ),
where I1 = 0/ and |I| is odd. Let J = supp(s1 ). Then,
(aJ s1 )(aI s)(aJ s1 )−1 = z(|J|−1)+|I||J| aI s = zaI s,
/
since |I| is odd. Hence, if Cρ+ ,ρ− is an odd split conjugacy class then ρ+ = 0.
Next, assume that Cρ+ ,ρ− is an odd conjugacy class such that ρ− contains two
identical parts. Then θ−1
n (Cρ+ ,ρ− ) contains an element of the form

aI s = (ai1 (i1 , i2 , . . . , ik ))(a j1 ( j1 , j2 , . . . , jk )) . . . (aIq sq ).


Consider the element t = (i1 , j1 ) . . . (ik , jk ). We have
t(aI s)t −1 = at(I) s = a j1 ( j1 , . . . , jk )ai1 (i1 , . . . , ik ) . . . = zaI s.
Hence, if an odd conjugacy class Cρ+ ,ρ− splits then ρ− ∈ SP. Together with the
above, we have shown that an odd split conjugacy class should satisfy (2).
Assume that (1) holds. Then θ−1 n (Cρ+ ,0/ ) contains an element s = s1 . . . sq ∈ Sn
with each si being an odd cycle. Suppose on the contrary the conjugacy class Cρ+ ,0/
does not split, that is, (aJ t)s(aJ t)−1 = zs for some element aJ t. Then (aJ t)s =
zs(aJ t), and so zas(J) = aJ , which implies that supp(s) ⊆ J. On the other hand,
as(J) = zd(s) aJ = aJ by (3.18), since s is a product of disjoint odd cycles. This
contradiction implies that the conjugacy class Cρ+ ,0/ splits.

Now assume that we are given an odd conjugacy class C0,ρ / − with ρ strict
as specified in (2). Thus, θ−1 (C0,ρ/ − ) contains an element aI s = (ai1 s1 ) . . . (aiq sq ),
where q is odd and ik ∈ supp(sk ). Suppose on the contrary that the conjugacy class
−1 = z(a s) for some element a t. It
/ − does not split, that is, (aJ t)(aI s)(aJ t)
C0,ρ I J
r
follows that t commutes with s and hence t = sr11 . . . sqq for 0 ≤ ri < ord(si ), since
the cycle type of s is a strict partition. Write aJ t = (aJ1 t1 ) . . . (aJq tq ) with tm = srmm .
As in the proof of Lemma 3.28 we have
(aJ t)(ai1 s1 )(aJ t)−1
= (aJ1 t1 )(aJ2 t2 ) · · · (aJq tq )(ai1 s1 )((aJ2 t2 ) · · · (aJq tq ))−1 (aJ1 t1 )−1
= z|J|−|J1 | (aJ1 t1 )(ai1 s1 )(aJ1 t1 )−1 ,

which must equal ai1 s1 up to a power of z. Set (aJ1 t1 )(ai1 s1 )(aJ1 t1 )−1 = z∗ ai1 s1
where ∗ is 0 or 1. We claim that ∗ is always 0. Note that aJ1 at1 (i1 ) = z∗ ai1 as1 (J1 ) ,
and so J1 differs from s1 (J1 ) by one element. Without loss of generality, we let
i1 = 1, s1 = (1, 2, . . . , k), J1 = {1, 2, . . . , r} with 0 ≤ r < k, then we have
a1 . . . ar · ar1 +1 = aJ1 at1 (i1 ) = z∗ ai1 as1 (J1 ) = z∗ a1 · a2 . . . ar+1 ,
106 3. Schur duality

which implies that r1 = r and ∗ = 0. Therefore, (aJ t)(ai1 s1 )(aJ t)−1 = z|J|−|J1 | ai1 s1 ,
and similarly we have
(aJ t)(aI s)(aJ t)−1 = (aJ t)(aI1 s1 ) . . . (aIq sq )(aJ t)−1
= zq|J|−(|J1 |+...+|Jq |) aI s = z(q−1)|J| aI s = aI s,
since q is odd. This is a contradiction. 

For α ∈ OPn we let C+ e −1


α be the split conjugacy class in Bn which lies in θn (Cα,0/ )
and contains a permutation in Sn of cycle type α. Then zC+ α is the other conjugacy
class in θ−1
n (Cα,0/ ), which will be denoted by C − . Recall from Appendix A that z
α α
denotes the order of the centralizer of an element of conjugacy type α in Sn . The
order of the centralizer of an element of a given cycle type is known explicitly for
Bn (and actually for any wreath product, cf. Macdonald [Mac, I, Appendix B]).
The next lemma follows from this classical fact.
Lemma 3.30. Let α ∈ OPn . The order of the centralizer of an element in the
conjugacy class C+ e
α of Bn is 2
1+ℓ(α) z . Thus, the order of the conjugacy class C+ is
α α
|C+ n−ℓ(α) z−1 .
α | = n!2 α

3.3.3. A ring structure on R− . The Clifford algebra Cn is the C-algebra generated


by ci (1 ≤ i ≤ n), subject to relations
(3.19) c2i = 1, ci c j = −c j ci (i ̸= j).
The symmetric group Sn acts as automorphisms on the algebra Cn naturally. We
will refer to the semi-direct product Hn := Cn o CSn as the Hecke-Clifford alge-
bra, where
(3.20) σci = cσ(i) σ, ∀σ ∈ Sn .
Note that the algebra Hn is naturally a superalgebra by letting each σ ∈ Sn be even
and each ci be odd.
The quotient algebra CΠn /⟨z + 1⟩ of the group algebra CΠn by the ideal gen-
erated by z + 1 is isomorphic to the Clifford algebra Cn with the identification
āi = ci , 1 ≤ i ≤ n. Hence, we have an isomorphism of superalgebras
(3.21) CBen /⟨z + 1⟩ ∼
= Hn .

Recall our convention (cf. Section 3.1.1) that a module of a superalgebra is


always understood to be Z2 -graded. We shall denote by Hn -mod the category of
modules of the superalgebra Hn (with morphisms of degree one allowed). Thanks
to the superalgebra isomorphism Hn ∼ = CBen /⟨z + 1⟩, Hn -mod is equivalent to the
e
category of spin Bn -modules. (We recall from Section 3.1.4 our convention that a
spin Ben -module means that a Z2 -graded Ben -module M on which z ∈ Ben acts as −I.)
We shall not distinguish these two isomorphic categories below, and the latter one
3.3. Representation theory of the algebra Hn 107

has the advantage that one can apply the standard arguments from the theory of
finite groups directly.
Denote by R−n the Grothendieck group of Hn -mod. As in the usual (ungraded)
case, we may replace the isomorphism classes of modules by their characters, and
then regard R−n as the free abelian group with a basis consisting of the characters
of the simple spin Ben -modules. Let

⊕ ∞

R− := R−
n, R−
Q := Q ⊗Z R−
n,
n=0 n=0

where it is understood that R− 0 = Z. We shall define a ring structure on R as


follows.
Denote by Bem,n the subgroup of Bem+n generated by Sm × Sn and Πm+n . Then
Bm,n can be identified with the quotient group Bem × Ben /{(1, 1), (z, z)}, where Bem ×
e
Ben denotes the product group in the super sense, i.e., elements from Bem and Ben
supercommute with each other.
Given φ ∈ R− − e
m and ψ ∈ Rn , we define a spin character φ×ψ of Bm,n by letting
ˆ

φ×ψ(x,
ˆ y) = φ(x)ψ(y),
where (x, y) is the image of (x, y) in Bem,n ∼
= (Bem × Ben )/{(1, 1), (z, z)}. We define a

product on R by
Be
φ · ψ = IndBem+n (φ×ψ),
ˆ
m,n

where φ ∈ R−
m, ψ∈ R−
nfor all m, n, and Ind denotes the induced character. It
follows from the properties of the induced characters that the multiplication on R−
is commutative and associative.
Remark 3.31. Equivalently, the multiplication in R− can be described as follows.
Let Hm,n be the subalgebra of Hm+n generated by Cm+n and Sm × Sn . For M ∈
Hm -mod and N ∈ Hn -mod, M ⊗ N is naturally an Hm,n -module, and we define the
product
[M] · [N] = [Hm+n ⊗Hm,n (M ⊗ N)],
and then extend it by Z-bilinearity.

For φ ∈ R−n , we shall write φα = φ(x) for x ∈ Cα , α ∈ OPn ; hence φ(y) = −φα
+

for y ∈ Cα . Given two partitions α, β, we let α ∪ β denote the partition obtained
by collecting the parts of α and β together and rearranging them in a descending
order.
Lemma 3.32. Let φ ∈ R− −
m , ψ ∈ Rn , and γ ∈ OPm+n . Then

(φ · ψ)γ = ∑ z z
φα ψ β .
α,β∈OP,α∪β=γ α β
108 3. Schur duality

Proof. It can be checked directly that



C+ e
γ ∩ Bm,n = C+
α ×Cβ .
ˆ +
α,β∈OP,α∪β=γ

It follows from this, Lemma 3.30, and the standard induced character formula for
finite groups that
|Bem+n |
(φ · ψ)γ =
|Bem,n | · |C+
∑ (φ×ψ)(w)
ˆ
γ | w∈C+
γ

21+ℓ(γ) zγ
= ∑ φα ψβ |C+α | · |C+β |
2m+n+1 m!n! α,β∈OP,α∪β=γ

= ∑ 2m−ℓ(α) z−1
m!n!2m+n−ℓ(γ) α,β∈OP,α∪β=γ α m!2
n−ℓ(β) −1
zβ n!φα ψβ


= ∑ z z
φα ψβ ,
α,β∈OP,α∪β=γ α β

where we have used ℓ(γ) = ℓ(α) + ℓ(β). 

3.3.4. The characteristic map. Recall from (A.43) in Appendix A that the ring
ΓQ := Q ⊗Z Γ has a basis given by the power-sum symmetric functions pµ for
µ ∈ OP. Moreover, ΓQ is equipped with a bilinear form ⟨·, ·⟩ given in (A.52).
We define the characteristic map
ch : R−
Q −→ ΓQ
to be the linear map given by
ch(φ) = ∑ z−1
α φα p α , φ ∈ R−
n.
α∈OPn

Proposition 3.33. The characteristic map ch : R−


Q → ΓQ is an algebra homomor-
phism.

Proof. For φ ∈ R− −
m , ψ ∈ Rn , we have by Lemma 3.32 that

ch(φ · ψ) = ∑ z−1
γ (φ · ψ)γ pγ
γ∈OP

= ∑ ∑ z−1
γ
zα zβ
φα ψβ pγ = ch(φ)ch(ψ),
γ α,β∈OP,α∪β=γ

where we have used pγ = pα pβ . 

Denote by
⟨[M], [N]⟩ = dim HomBen (M, N)
for spin Ben -modules M, N. This defines a bilinear form ⟨·, ·⟩ on R−
n by Z-bilinearity.
The HomBen here can be either understood as in the category of Ben -modules (with
3.3. Representation theory of the algebra Hn 109

degree one morphism allowed) or in the category of (non-graded) Ben -modules, and
they give the same dimension. In light of super Schur’s Lemma 3.4, this is reduced
to a straightforward verification in the case when M ∼
= N is simple of type Q.
Then R− , and hence also R−
Q , carry a symmetric bilinear form, still denoted by
⟨·, ·⟩, which is induced from the ones on R− − −
n , for all n, such that Rn and Rm are
orthogonal whenever n ̸= m.
Lemma 3.34. For φ, ψ ∈ R−
n , we have

⟨φ, ψ⟩ = ∑ 2−ℓ(α) z−1


α φα ψα .
α∈OPn

Proof. Note that x ∈ C+ −1 ∈ C+ . By Lemma 3.30 and applying the


α implies that x α
standard formula for the bilinear form on characters of a finite group, we have
1
⟨φ, ψ⟩ =
e ∑
|Bn | x∈Be
φ(x−1 )ψ(x)
n

1
= ∑
2n+1 n! α∈OP α |φα ψα
2|C+
n

1
= ∑ 2n−ℓ(α) z−1
2n n! α∈OP α n!φα ψα
n

= ∑ 2−ℓ(α) z−1
α φα ψα .
α∈OPn


Proposition 3.35. The characteristic map ch : R− Q → ΓQ is an isometry, i.e., it

preserves the bilinear forms ⟨·, ·⟩ on RQ and ΓQ .

Proof. This follows from a direct computation using Lemma 3.34: for φ, ψ ∈ R−
n
we have
⟨ch(φ), ch(ψ)⟩ = ∑ z−1 −1
α zβ ⟨pα , pβ ⟩φα φβ
α,β∈OPn

= ∑ z−2
α 2
−ℓ(α)
zα φα ψα = ⟨φ, ψ⟩.
α∈OPn

3.3.5. The basic spin module. The algebra Hn acts on the Clifford algebra Cn by
the formula
ci .(ci1 ci2 . . .) = ci ci1 ci2 . . . , σ.(ci1 ci2 . . .) = cσ(i1 ) cσ(i2 ) . . . ,
for σ ∈ Sn . The Hn -module Cn is called the basic spin module. Let σ = σ1 . . . σℓ ∈
Sn be a product of disjoint cycles with cycle length of σi being µi , for i = 1, . . . , ℓ.
110 3. Schur duality

If I is a union of some of the supp(σi )’s, say I = supp(σi1 ) ∪ . . . ∪ supp(σis ), then


σ(cI ) = (−1)µi1 +...+µis −s cI . Otherwise, σ(cI ) is not a scalar multiple of cI .
Lemma 3.36. The value of the character ξn of the basic spin Ben -module at the
conjugacy class C+
α is given by

(3.22) ξnα = 2ℓ(α) , α ∈ OPn .

Proof. Let α = (α1 , α2 , . . .) ∈ OPn and ℓ(α) = ℓ. Let σ = σ1 . . . σℓ be an element in


Sn of cycle type α. The elements cI := ∏i∈I ci (which is defined up to a nonessen-
tial sign), for I ⊂ {1, . . . , n}, form a basis of the basic spin module Cn . Observe
that σcI = cI if I is a union of a subset of the supports supp(σ p ) for 1 ≤ p ≤ ℓ(α);
otherwise σcI is equal to ±cJ for some J ̸= I. Hence the character value ξnα , which
is the trace of σ on Cn , is equal to 2ℓ(α) . 

Below we will freely use the statements in Appendix A.3 on Schur Q-functions
Qλ . Recall the symmetric function qn defined by the generating function (A.40),
which is,
1 + xit
∑ qnt n = ∏ 1 − xit .
n≥0 i≥1

Lemma 3.37. Let n ≥ 1. We have


(1) ch(ξn ) = qn ;
(2) ⟨ξn , ξn ⟩ = 2;
(3) the basic spin Ben -module Cn is simple of type Q.

Proof. It follows by the definition of ch, Lemma 3.36, and (A.45) that
ch(ξn ) = ∑ 2ℓ(α) z−1
α pα = qn .
α∈OPn

Now pk (1, 0, 0, . . .) = 1 for each k ≥ 1, and hence, pα (1, 0, 0, . . .) = 1. Also, by


definition qn (1, 0, 0, . . .) = 2. Thus, specializing the identity ∑α∈OPn 2ℓ(α) z−1 α pα =
qn at (x1 , x2 , x3 , . . .) = (1, 0, 0, . . .) we obtain ∑α∈OPn 2ℓ(α) z−1
α = 2. We compute, by
Lemmas 3.34 and 3.36, that
⟨ξn , ξn ⟩ = ∑ 2−ℓ(α) z−1
α (2
ℓ(α) 2
) = ∑ 2ℓ(α) z−1
α = 2.
α∈OPn α∈OPn

As the Ben -module Cn is semisimple, to prove (3) it suffices to exhibit an odd


automorphism of the Ben -module Cn , by (2) and super Schur’s Lemma. Indeed, the
right multiplication with the element √1n (c1 + . . . + cn ) provides such an automor-
phism of order 2. 
Proposition 3.38. The characteristic map ch : R−
Q → ΓQ is an isomorphism of
graded vector spaces.
3.3. Representation theory of the algebra Hn 111

Proof. Since ΓQ is generated by qr for r ≥ 1 and ch is an algebra homomorphism


by Proposition 3.33, ch is surjective by Lemma 3.37. By Proposition 3.7 and Theo-
rem 3.29, the dimension of Q ⊗ R− n is |SPn |, which is the same as dim ΓQ , for each
n

n. Hence ch is an isomorphism. 

3.3.6. The irreducible characters. Using the algebra structure on R− , we define


the elements ξλ for a strict partition λ by the following recursive relations:
λ2
(3.23) ξ(λ1 ,λ2 ) = ξλ1 ξλ2 + 2 ∑ (−1)i ξλ1 +i ξλ2 −i ,
i=1
k
(3.24) ξλ = ∑ (−1) j ξ(λ ,λ ) ξ(λ ,...λ̂ ...,λ ) ,
1 j 2 j k
for k = ℓ(λ) even,
j=2
k
(3.25) ξλ = ∑ (−1) j−1 ξλ ξ(λ ,...λ̂ ...,λ ) ,
j 1 j k
for k = ℓ(λ) odd.
j=1

We emphasize that these are precisely the same recursive relations for the Schur
Q-functions Qλ ; see (A.48).

Lemma 3.39. We have ch(ξλ ) = Qλ and ⟨ξλ , ξµ ⟩ = 2ℓ(λ) δλµ , for λ, µ ∈ SP.

Proof. By Lemma 3.37, we have ch(ξn ) = qn = Q(n) . The general case of the first
identity follows since ch is a ring homomorphism, and in addition, ξλ and Qλ are
obtained from ξn ’s and qn ’s, respectively, by the same recursive relations.
The second identity follows from the fact that ch is an isometry and the formula
⟨Qλ , Qµ ⟩ = 2ℓ(λ) δλµ from (A.54). 

ℓ(λ)−δ(λ)
We define ζλ := 2− 2 ξλ , for λ ∈ SPn .

Lemma 3.40. The element ζλ lies in R−


n , for λ ∈ SPn .

Proof. We proceed by induction on ℓ(λ). For ℓ(λ) = 1, it is clear. Since Cn is a


simple spin Ben -module of type Q by Lemma 3.37, the induced module with char-
acter ξm ξn is a sum of two isomorphic copies of an honest module (the two odd
automorphisms of ξm and ξn give rise to an even automorphism of order 2); that is,
− lies in R− by (3.23). The general case
2 ξ ξ ∈ R . Hence, for ℓ(λ) = 2, 2 ξ
1 m n 1 (λ1 ,λ2 )

follows easily by induction using the recursive relations (3.24) and (3.25). 

Corollary 3.41. For strict partitions λ, µ, we have


{
λ λ 1 for ℓ(λ) even
⟨ζ , ζ ⟩ =
2 for ℓ(λ) odd,
⟨ζλ , ζµ ⟩ = 0, for λ ̸= µ.
112 3. Schur duality

Corollary 3.42. For each λ ∈ SPn , we have


ℓ(λ)−δ(λ)
Qλ = 2 2
∑ z−1 λ
α ζα p α .
α∈OPn

Also, for each α ∈ OPn , we have


ℓ(λ)+δ(λ)
pα = ∑ 2− 2 −ℓ(α)
ζλα Qλ .
λ∈SPn

Proof. The first identity follows from Lemma 3.39 and the definition of ζλ .
Write pα = ∑λ∈SPn aλα Qλ for some scalars aλα . Recall from (A.52) and (A.54)
that ⟨Qλ , Qµ ⟩ = 2ℓ(λ) δλµ and ⟨pα , pβ ⟩ = 2−ℓ(α) zα δαβ . Then,
ℓ(λ)−δ(λ) ℓ(λ)+δ(λ)
aλα = 2−ℓ(λ) ⟨pα , Qλ ⟩ = 2−ℓ(λ) 2 2 z−1 λ
α ζα ⟨pα , pα ⟩ = 2
− 2 −ℓ(α)
ζλα ,
where the middle equation uses the first identity of the corollary. 

Theorem 3.43. Let λ ∈ SPn and ℓ(λ) = ℓ. Then, ζλ is the character of a simple
spin Ben -module (which is to be denoted by Dλ ). Moreover, the degree of ζλ is equal
to
ℓ−δ(λ) n! λi − λ j
2n− 2 ∏
λ1 ! . . . λℓ ! i< j λi + λ j
.

Proof. Since ζλ ∈ R− by Lemma 3.40 and ⟨ζλ , ζλ ⟩ = 1 for ℓ(λ) even by Corol-
lary 3.41, ζλ or −ζλ , for ℓ(λ) even, is a simple character of type M. By Proposi-
tion 3.7 and Theorem 3.29, these are all simple characters of type M. Now since
⟨ζλ , ζλ ⟩ = 2 for ℓ(λ) odd by Corollary 3.41, we have two possibilities: (a) ζλ or
−ζλ is either simple of type Q, or (b) ζλ is of the form ±ζµ ± ζν with both ζµ and
ζν being simple of type M which means both ℓ(µ) and ℓ(ν) being even. Case (b)
cannot occur, otherwise it would contradict the linear independence of ζλ for all
λ ∈ SPn .
To show that ζλ instead of −ζλ for each λ ∈ SPn is a character of a simple
module, it suffices to know that ζλ(1n ) or ξλ(1n ) is positive. To that end, we claim that

n! λi − λ j
ξλ(1n ) = 2n ∏
λ1 ! . . . λℓ ! i< j λi + λ j
,

which is equivalent to the degree formula in the theorem.


The claim can be proved by induction on ℓ(λ) using the relations (3.23)–(3.25).
The initial case with ℓ(λ) = 1 is taken care by Lemma 3.36 and the case with
ℓ(λ) = 2 can be checked directly by using (3.23). As the induction is elementary
though lengthy (see [Jo2, proof of Proposition 4.13]), we will simply remark here
that the sought-for identity by using (3.24) and (3.25) precisely corresponds to the
3.4. Schur-Sergeev duality for q(n) 113

Laplacian type expansion of the classical Pfaffian identity (cf. Macdonald [Mac,
III.8, Ex. 5]) ( )
ti − t j ti − t j
Pf = ∏ .
ti + t j 1≤i< j≤2n ti + t j


3.4. Schur-Sergeev duality for q(n)


In this section, we formulate a double centralizer property for the actions of the
Lie superalgebra q(n) and of the algebra Hd on the tensor space (Cn|n )⊗d . We
obtain a multiplicity-free decomposition of (Cn|n )⊗d as a U(q(n)) ⊗ Hd -module.
The characters of the simple q(n)-modules arising this way are shown to be Schur
Q-functions (up to some 2-powers).

3.4.1. A double centralizer property. Recall from Section 3.2 (by setting m = n)
that we have a representation (V ⊗d , Φd ) of gl(n|n), hence of its subalgebra q(n),
and we also have a representation (V ⊗d , Ψd ) of the symmetric group Sd . More-
over, the actions of gl(n|n) and the symmetric group Sd on V ⊗d commute with
each other.
Note in addition that the Clifford algebra Cd acts on V ⊗d , also denoted by Ψd :
Ψd (ci ).(v1 ⊗ . . . ⊗ vd ) = (−1)(|v1 |+...+|vi−1 |) v1 ⊗ . . . ⊗ vi−1 ⊗ Pvi ⊗ . . . ⊗ vd ,
where P is given in (1.10), each vi ∈ V is assumed to be Z2 -homogeneous, and
1 ≤ i ≤ n.
Lemma 3.44. Let V = Cn|n . The actions of Sd and Cd above give rise to a repre-
sentation (V ⊗d , Ψd ) of Hd . Moreover, the actions of q(n) and Hd on V ⊗d commute
with each other.

Proof. To see the relation (3.20) holds, it suffices to check for σ = ( j, j + 1), 1 ≤
j ≤ d − 1. We compute that
Ψd ((i, i + 1))Ψd (ci ).(v1 ⊗ . . . ⊗ vd )
= Ψd ((i, i + 1))(−1)(|v1 |+...+|vi−1 |) v1 ⊗ . . . ⊗ Pvi ⊗ vi+1 ⊗ . . . ⊗ vd
= (−1)(|vi |+1)|vi+1 | (−1)(|v1 |+...+|vi−1 |) v1 ⊗ . . . ⊗ vi+1 ⊗ Pvi ⊗ . . . ⊗ vd ,
Ψd (ci+1 )Ψd ((i, i + 1)).(v1 ⊗ . . . ⊗ vd )
= Ψd (ci+1 )(−1)(|vi |·|vi+1 |) v1 ⊗ . . . ⊗ vi+1 ⊗ vi ⊗ . . . ⊗ vd
= (−1)(|vi |·|vi+1 |) (−1)(|v1 |+...+|vi−1 |+|vi+1 |) v1 ⊗ . . . ⊗ vi+1 ⊗ Pvi ⊗ . . . ⊗ vd .
Hence, we have Ψd ((i, i + 1))Ψd (ci ) = Ψd (ci+1 )Ψd ((i, i + 1)). This further implies
that Ψd ((i, i + 1))Ψd (ci+1 ) = Ψd (ci )Ψd ((i, i + 1)). A similar calculation shows that
Ψd (( j, j + 1))Ψd (ci ) = Ψd (ci )Ψd (( j, j + 1)) for j ̸= i, i − 1.
114 3. Schur duality

By the definitions of q(n) and of Ψd (ci ) via P, the action of q(n) commutes
with the action of ci for 1 ≤ i ≤ d. Since gl(n|n) commutes with Sd , so does the
subalgebra q(n) of gl(n|n). Hence, the action of q(n) commutes with the action of
Hd on V ⊗d . 
Theorem 3.45. The images Φd (U(q(n))) and Ψd (Hd ) satisfy the double central-
izer property, i.e.,
Φd (U(q(n))) =EndHd (V ⊗d ),
Endq(n) (V ⊗d ) = Ψd (Hd ).

Proof. Let g = q(n). We denote by Q(V ) the associative subalgebra of endomor-


phisms on V which (super)commute with the linear operator P. By Lemma 3.44,
we have
(3.26) Φd (U(g)) ⊆ EndHd (V ⊗d ).

We shall proceed to prove that Φd (U(g)) ⊇ EndHd (V ⊗d ).


By examining the action of Cd on V ⊗d , we see that the natural isomorphism
End(V )⊗d ∼= End(V ⊗d ) allows us to identify EndCd (V ⊗d ) ≡ Q(V )⊗d . As we recall
Hd = Cd o Sd , this further leads to the identification EndHd (V ⊗d ) ≡ Symd (Q(V )),
the space of Sd -invariants in Q(V )⊗d .
Denote by Yk , 1 ≤ k ≤ d, the C-span of the super-symmetrization
ϖ(x1 , . . . , xk ) := ∑ σ.(x1 ⊗ . . . ⊗ xk ⊗ 1d−k ),
σ∈Sd

for all xi ∈ Q(V ). Note that Yd = Symd (Q(V )) ≡ EndHd (V ⊗d ).


Let x̃ = Φ(x) = ∑di=1 1i−1 ⊗ x ⊗ 1d−i , for x ∈ g = Q(V ), and denote by Xk ,
1 ≤ k ≤ d, the C-span of x̃1 . . . x̃k for all xi ∈ q(n).
By the same argument as in the proof of Theorem 3.9 (Schur-Sergeev duality),
we have Yk ⊆ Xk for 1 ≤ k ≤ d. This implies that
EndHd (V ⊗d ) = Yd ⊆ Xd ⊆ Φd (U(g)).

Combining with (3.26), we have that Φd (U(g)) = EndHd (V ⊗d ) = EndB (V ⊗d ),


for B := Ψd (Hd ).
Note that the spin group algebra Hd , and hence also B, are semisimple super-
algebras, and so the assumption of Proposition 3.5 is satisfied. Therefore, we have
EndU(g) (V ⊗d ) = Ψd (Hd ). 

3.4.2. The Sergeev duality. Recall from (1.35) that, for a partition λ of d with
length ℓ(λ),
{
0, if ℓ(λ) is even,
δ(λ) =
1, if ℓ(λ) is odd.
3.4. Schur-Sergeev duality for q(n) 115

Recall further that Dλ stands for the simple Hd -module with character ζλ (see
Theorem 3.43), and L(λ) for the simple q(n)-module with highest weight λ (see
Section 2.1.6).
Theorem 3.46. Let V = Cn|n . As a U(q(n)) ⊗ Hd -module, we have

(3.27) ∼
V ⊗d = 2−δ(λ) L(λ) ⊗ Dλ .
λ∈SPd ,ℓ(λ)≤n

Here, 2−1 has the same meaning as in Proposition 3.5.

Proof. Let W = V ⊗d . It follows from the double centralizer property and the
semisimplicity of the superalgebra Hd that we have a multiplicity-free decomposi-
tion of the (q(n), Hd )-module W :

W∼ = 2−δ(λ) L[λ] ⊗ Dλ ,
λ∈Q(d,n)

where L[λ] is some simple q(n)-module associated to λ, whose highest weight (with
respect to the standard Borel) is to be determined. Also to be determined is the
index set Q(d, n) = {λ ∈ SPd | L[λ] ̸= 0}.
We shall identify a weight µ = ∑ni=1 µi εi occuring in W with a composition
µ = (µ1 , . . . , µn ) ∈ CP(d, n). We have the following weight space decomposition:

(3.28) W= Wµ ,
µ∈CP(d,n)

where Wµ has a linear basis ei1 ⊗ . . . ⊗ eid , with the indices satisfying the following
equality of sets:
{i1 , . . . , id } = {1, . . . , 1, , 1, . . . , 1, . . . , n, . . . , n, n, . . . , n}.
| {z } | {z }
µ1 µn

We have an Hd -module isomorphism:


(3.29) Wµ ∼ H
= IndCSd µ 1 = Hd ⊗CSµ 1.

bλµ for a composition µ and a strict partition λ defined via


Recall the integers K
the following symmetric function identity in (A.50) :
(3.30) qµ = ∑ K bλµ Qλ ,
λ∈SP,λ≥µ

bλλ = 1. To complete the proof of the theorem, we shall need the following.
where K
Lemma 3.47. Let µ = (µ1 , . . . , µn ) be a composition of d. We have the following
decomposition of Wµ as an Hd -module:

Wµ ∼ bλµ Dλ .
ℓ(λ)−δ(λ)
= 2 2 K
λ∈SP,λ≥µ

bλµ ∈ Z+ .
In particular, K
116 3. Schur duality

Proof. Decompose the Hd -module Wµ into irreducibles:



(3.31) Wµ ∼
= K̃λµ Dλ , for K̃λµ ∈ Z+ .
λ

Recall the characteristic map ch : R− → ΓQ from Section 3.3.4, where we have



denoted R− = n≥0 K(Hn -mod). By Lemmas 3.39 and 3.40, for λ ∈ SP,
l(λ)−δ(λ)
(3.32) ch([Dλ ]) = 2− 2 Qλ .
It follows by the ring structure on R− and Remark 3.31 that, for any composition
Hd
µ = (µ1 , µ2 , . . .) of d, IndCSµ 1 is equal to the product of the basic spin characters
ξµ1 · ξµ2 . . ., and hence by Lemma 3.37,
( )
(3.33) ch(Wµ ) = ch IndH d
CSµ 1 = qµ .

Applying the characteristic map to both sides of (3.31), and using (3.32), (3.33)
and (3.29), we obtain that
ℓ(λ)−δ(λ)
qµ = ∑ 2− 2 K̃λµ Qλ .
λ

bλµ .
ℓ(λ)−δ(λ)
It follows by a comparison of this identity with (3.30) that K̃λµ = 2 2 K 

We return to the proof of the theorem. Note that the simple q(n)-module L[λ]
⊕ [λ]
has a weight space decomposition L[λ] = µ∈CP(d,n),µ≤λ Lµ , and hence, λ ∈ Pd (n)
if L[λ] ̸= 0. Among all such µ, clearly λ corresponds to a highest weight. Hence,
we conclude that L[λ] = L(λ), the simple g-module of highest weight λ, and that
Q(d, n) = {λ ∈ SPd | ℓ(λ) ≤ n}. This completes the proof of Theorem 3.46. 

3.4.3. The irreducible character formula. A character of a q(n)-module with


weight space decomposition M = ⊕µ Mµ is
chM = tr |M x1H1 . . . xnHn = ∑ µ
dim Mµ · x11 . . . xnµn .
µ=(µ1 ,...,µn )

Theorem 3.48. Let λ be a strict partition of length ≤ n. The character of the


ℓ(λ)−δ(λ)
simple q(n)-module L(λ) is given by chL(λ) = 2− 2 Qλ (x1 , . . . , xn ).

Proof. By (3.28) and (3.29), we have V ⊗d ∼ H


= ⊕µ∈CPd (n) IndSµd 1. Using (3.33) and
applying the characteristic map ch and the trace operator tr x1H1 . . . xnHn simultane-
ously, which we will denote by ch2 , we obtain that
ch2 (V ⊗d ) = ∑ qµ (z)mµ (x),
µ∈Pd ,ℓ(µ)≤n

which can be written using (A.51) and (A.54) as


1 + xi z j
ch2 (V ⊗d ) = ∏ = ∑ 2−ℓ(λ) Qλ (x1 , . . . , xn )Qλ (z).
1≤i≤n,1≤ j 1 − xi z j λ∈SP
3.5. Notes 117

On the other hand, by applying ch2 to (3.27) and using (3.32), we obtain that
ℓ(λ)−δ(λ)
ch2 (V ⊗d ) = ∑ 2−δ(λ) chL(λ) · 2− 2 Qλ (z).
λ∈SPd ,ℓ(λ)≤n

Now the theorem follows by comparing the above two identities and noting the
linear independence of the Qλ (z)’s. 

3.5. Notes
Section 3.1. The classification of finite-dimensional simple associative superalge-
bras was due to Wall [Wal] over a general field, and it is somewhat simplified over
C in Józefiak [Jo1]. The basics on representation theory of superalgebras and finite
supergroups, including Wedderburn Theorem, Schur’s Lemma and the role of split
conjugacy classes, have been developed in [Jo1].
Section 3.2. The superalgebra generalization of Schur duality between general
linear Lie superalgebra gl(m|n) and the symmetric group Sd was due to Sergeev
[Sv1] and Berele-Regev [BeR], and the character of irreducible polynomial repre-
sentations of gl(m|n) was given in terms of super Schur polynomials. This general-
ization is intimately related to the combinatorics of supertableaux. Proposition 3.20
on the diagrammatic interpretation of the degree of atypicality of a polynomial
dominant weight of gl(m|n) was due to Cheng-Wang [CW5], and Corollary 3.21
appeared in Moens and van der Jeugt [MJ, Proposition 2.1]. The semisimplicity
of the category of polynomial modules of gl(m|n) was expected by experts. Our
proof here, which is based on a comparison of central characters, is borrowed from
Cheng-Kwon [CK].
Section 3.3. The algebra Hd is a twisted group algebra for a double cover Bed
of the hyperoctraheral group. The classification of the split conjugacy classes for
Bed (Theorem 3.29) was due to Read [Re]. We follow Józefiak [Jo3] to develop
systematically a superalgebra approach toward the characteristic map and spin rep-
resentation theory for Bed .
Section 3.4. The Sergeev duality is a version of Schur duality between the
queer Lie superalgebra q(n) and an algebra Hd , and it was outlined in [Sv1]. This
leads to a character formula in terms of Schur Q-functions for the irreducible poly-
nomial q(n)-modules [Sv1]. Here we present complete proofs which use exten-
sively the results from Sections 3.1 and 3.3.
Various results on symmetric functions including super Schur functions and
Schur Q-functions relevant to this Chapter are collected in Appendix A.
Chapter 4

Classical invariant theory

In this chapter, we give an introduction to the invariant theory for a group G,


which is one of the classical groups GL(V ), Sp(V ), or O(V ), where V is a finite-
dimensional vector space.
Traditionally, the G-invariants of a polynomial (or symmetric) algebra on U
are more widely studied, where U is a direct sum of copies of V or its dual. It has
gradually become clear that the algebra of G-invariants in an exterior algebra of U
can be developed in a parallel fashion to a large extent. In our presentation, we treat
the polynomial and exterior algebras in the united framework of supersymmetric
algebras. We formulate and establish the First Fundamental Theorem (FFT) in a
supersymmetric algebra setting which states that the subalgebra of G-invariants is
generated by a finite set of basic invariants of degree two. We also develop a tensor
version of FFT for each of the classical groups, which is used in the proof of the
FFT for supersymmetric algebras. In the case of type A, the tensor FFT is derived
from Schur duality, and it is indeed equivalent to Schur duality.
The FFT for supersymmetric algebras of classical groups will be used in an
essential way in the development of Howe duality in Chapter 5.

4.1. FFT for the general linear Lie group


In this section, we shall formulate and prove both the tensor and the polynomial
versions of the First Fundamental Theorem (FFT) of classical invariant theory for
the general linear Lie group. For a vector space U we shall denote the polynomial
algebra on U by P(U).

4.1.1. General invariant theory. Let U and V be finite-dimensional modules of


a classical (or more generally a reductive) Lie group G. Then as G-modules U ⊗d

119
120 4. Classical invariant theory

and P(U) with induced G-actions are completely reducible. The basic question of
classical invariant theory is to describe
(1) the space of tensor G-invariants (U ⊗d )G ;
(2) the algebra of polynomial G-invariants P(U)G ;
(3) the algebra of G-invariants in the tensor algebra P(U) ⊗ ∧(V ).
These different versions of G-invariants turn out to be closely related to each other.
Theorem 4.1. Let G be a classical (or reductive) group, and let U,V be finite-
dimensional rational G-modules. Then, the algebra of G-invariants in P(U) ⊗
∧(V ) is finitely generated as a C-algebra.

Proof. Set R := P(U) ⊗ ∧(V ). Under the assumptions of the theorem, the G-
module R is completely reducible, and hence the subalgebra J := (P(U) ⊗ ∧(V ))G
is a direct summand of the G-module R. Let us denote by ϖ : R −→ J the natural
projection which is G-equivariant.
The algebra R is naturally Z+ -graded by the total degree, denoted by deg, on
the polynomial subalgebra and the exterior subalgebra. Denote by J+ the subspace
of J which consists of G-invariant elements in R with zero constant terms, and
denote by ⟨J+ ⟩ the (two-sided) ideal of R generated by J+ . Then ⟨J+ ⟩, when
viewed as a submodule of the finitely generated P(U)-module R = P(U) ⊗ ∧(V ),
is finitely generated over P(U), by Hilbert basis theorem. In particular, ⟨J+ ⟩ as an
ideal of R is finitely generated. We can further take a set of generators φ1 , . . . , φn
of the ideal ⟨J+ ⟩ to be homogeneous elements, say deg φi = di ≥ 1 for each i.
To complete the proof of the theorem, we shall show that φ1 , . . . , φn generate J
as a C-algebra. Indeed, let φ ∈ J+ be a homogeneous element. Write φ = ∑i fi φi
for some fi ∈ R. Then,
φ = ϖ(φ) = ϖ(∑ fi φi ) = ∑ ϖ( fi )φi ,
i i

where deg ϖ( fi ) ≤ deg fi < deg φ. By induction on the degree, we can assume that
ϖ( fi ) lies in the C-subalgebra generated by φ1 , . . . , φn . 

For a general G-module U, it remains an open problem to describe a reasonable


set of generators for the algebra P(U)G . However, in the case when G is one of
the classical groups acting on the natural representation V , and U is a direct sum
of copies of V and copies of V ∗ , the problem turns out to have an elegant solution,
known as the First Fundamental Theorem (FFT) of classical invariant theory.

4.1.2. Tensor and multilinear FFT for GL(V ). Now let V be a finite-dimensional
vector space, and let G = GL(V ) be the general linear Lie group on V . Let U =
V ⊗d ⊗ (V ∗ )⊗k , the space of mixed tensors of type (d, k). Denote the representation
of G on such a U by ρd,k .
4.1. FFT for the general linear Lie group 121

First observe that λI ∈ G acts on V ⊗d ⊗ (V ∗ )⊗k by ρd,k (λI) = λd−k I. Hence,


(V ⊗d ⊗ (V ∗ )⊗k )G = 0 unless k = d.
Assume now that k = d, and we consider the G-representation ρd,d . For any
finite-dimensional G-module W , we have a canonical identification as G-modules:
W ⊗W ∗ ∼
= End(W ).
Set W = V ⊗d . We have (V ⊗d )∗ = (V ∗ )⊗d as G-modules. Then we have a canonical
identification as G-modules:
V ⊗d ⊗ (V ∗ )⊗d ∼
= End(V ⊗d ).
It follows that
( )G
(4.1) V ⊗d ⊗ (V ∗ )⊗d ∼ = EndG (V ⊗d ).

Recall the following commuting actions from Lemma 3.8 in Section 3.2 where
we have replaced the action of gl(V ) by G = GL(V ):
Φd Ψd
G y V ⊗d x Sd .
By Schur duality (Theorem 3.9), we have
(4.2) EndG (V ⊗d ) = Ψd (CSd ).

Via (4.1) and (4.2), Ψd (σ) for each permutation σ ∈ Sd transfers to a G-


( )G
invariant Θσ ∈ V ⊗d ⊗ (V ∗ )⊗d , which can be written down explicitly in terms
of dual bases. Let e1 , . . . , eN be a basis for V and e∗1 , . . . , e∗N be the dual basis of V ∗ .
We have
(4.3) Θσ = ∑ eiσ(1) ⊗ . . . ⊗ eiσ(d) ⊗ e∗i1 ⊗ . . . ⊗ e∗id .
1≤i1 ,...,id ≤N

Summarizing, we have proved the following.

Theorem 4.2 (Tensor FFT for GL(V )). Let G = GL(V ). There is no non-zero G-
invariants in the tensor space V ⊗d ⊗ (V ∗ )⊗k , for k ̸= d. The space of G-invariants
in V ⊗d ⊗ (V ∗ )⊗d is spanned by the Θσ for σ ∈ Sd .

For future applications, it will be convenient to formulate a multilinear version


of FFT, which is simply a dual version to the tensor FFT for GL(V ) above. We first
recall a standard fact from multilinear algebra, which follows from the universal
property of tensor product.

Lemma 4.3. Let V1 , . . . ,Vp be finite-dimensional vector spaces. Then, the dual
space (V1 ⊗ · · · ⊗ Vp )∗ can be naturally identified with the space of multilinear
functions on V1 ⊕ · · · ⊕Vp .
122 4. Classical invariant theory

( )∗
By Lemma 4.3, we shall regard the dual space V ⊗d ⊗ (V ∗ )⊗k as the space
of multilinear functions f : V d ⊕ V ∗k → C. There is no nonzero G-invariant mul-
tilinear functions, for k ̸= d. For, aI ∈ G sends (v, φ) = (v1 , . . . , vd , φ1 , . . . , φk ) to
(av1 , . . . , avd , a−1 φ1 , . . . , a−1 φk ); hence, if f is a G-invariant multilinear function,
then f (a(v, φ)) = ad−k f (v, φ).
Let k = d. Define the contraction ⟨ j|i∗ ⟩, for 1 ≤ i, j ≤ d, by
(4.4) ⟨ j|i∗ ⟩(v1 , . . . , vd , φ1 , . . . , φk ) := φi (v j ).
Given σ ∈ Sd , we let
fσ := ⟨1|σ(1)∗ ⟩⟨2|σ(2)∗ ⟩ · · · ⟨d|σ(d)∗ ⟩.
The tensor FFT for GL(V ) (Theorem 4.2) can now be reformulated as follows.
Theorem 4.4 (Multilinear FFT for GL(V )). Let G = GL(V ). Then there is no
nonzero G-invariant multilinear functions on V d ⊕ V ∗k , for k ̸= d. The space of
G-invariant multilinear functions on V d ⊕ V ∗d is spanned by the functions fσ for
σ ∈ Sd .
( ) ( ⊗d )∗
Proof. Thanks to V ∗∗ ∼
= V , we identify (V ∗ )⊗d ⊗V ⊗d ∼= V ⊗ (V ∗ )⊗d . Thus,
we have by (4.1) and (4.2) that
( )G ( )G
(V ∗ )⊗d ⊗V ⊗d ∼ = V ⊗d ⊗ (V ∗ )⊗d ∼ = Ψd (CSd ).
( )G
By Lemma 4.3, (V ∗ )⊗d ⊗V ⊗d can be identified with the space of G-invariant
multilinear functions on V ⊗d ⊗ (V ∗ )⊗d . Hence, the theorem follows as a dual ver-
sion of the tensor FFT for GL(V ) (see Theorem 4.2), where fσ is the counterpart
of Θσ in (4.3). 

4.1.3. Formulation of the polynomial FFT for GL(V ). Again let G = GL(V ).
Then G acts naturally on the dual space V ∗ and also on the direct sums V k and
V ∗m , for k, m ≥ 0. We have natural identifications of G-modules:
∼ ∼
V k −→ Hom(Ck ,V ), V ∗m −→ Hom(V, Cm ).
This leads to a G-equivariant isomorphism between polynomial algebras

P(V k ⊕V ∗m ) −→ P(Vm,k ),
where we have denoted
Vm,k := Hom(Ck ,V ) ⊕ Hom(V, Cm ).

We identify the Hom-space Hom(Ck , Cm ) with the space Mmk of m×k matrices
over C, and define the composition map
τ : Vm,k → Mmk , τ(x, y) = yx.
4.1. FFT for the general linear Lie group 123

Note that τ is G-equivariant, where we let G act on Mmk trivially. Hence, it induces
an algebra homomorphism
τ∗ : P(Mmk ) −→ P(Vm,k )G , τ∗ ( f ) = f ◦ τ.
The image of τ∗ consists of G-invariants, since for g ∈ G, x ∈ Vm,k , by the G-
equivariance of τ, we have τ∗ ( f )(g.x) = f (τ(g.x)) = f (τ(x)) = τ∗ ( f )(x).
Denote by xi j the (i, j)th matrix coefficient on Mmk , where 1 ≤ i ≤ m, 1 ≤ j ≤ k.
It follows by a direct computation that τ∗ (xi j ) coincides with the contraction ⟨ j|i∗ ⟩
(which is defined as in (4.4) with obvious modifications of indices), or equivalently,
τ∗ (xi j )(v1 , . . . , vk , v∗1 , . . . , v∗m ) = v∗i (v j ).
Theorem 4.5 (Polynomial FFT for GL(V )). The algebra of GL(V )-invariants in
P(V k ⊕V ∗m ) is generated by the contractions ⟨ j|i∗ ⟩, for 1 ≤ i ≤ m, 1 ≤ j ≤ k.

The proof of Theorem 4.5 will be given in Section 4.1.4 below.


Remark 4.6. The above theorem can be equivalently reformulated as the surjec-
tivity of the homomorphism τ∗ . One easily shows that the image of τ consists of all
the matrices Z ∈ Mmk such that rank(Z) ≤ min{m, k, dimV }. Under the assumption
that dimV ≥ min{m, k} the map τ is surjective. Therefore τ∗ is injective, and so τ∗
is actually an isomorphism. Equivalently, the algebra of polynomial G-invariants
in P(V ∗m ⊕V k ) is a polynomial algebra generated by the mk contractions ⟨ j|i∗ ⟩.

4.1.4. Polarization and restitution. Let P p (W ) be the set of degree p homoge-


neous polynomials on a finite-dimensional vector space W , and let f ∈ P p (W ). We
define a family of polynomials fr1 ...r p ∈ P(W p ), which are multi-homogeneous of
degree (r1 , . . . , r p ), by the following formula:

r
(4.5) f (t1 w1 + . . . + t p w p ) = fr1 ...r p (w1 , . . . , w p )t1r1 · · ·t pp ,
r1 ,...,r p

where ti ∈ C, wi ∈ W , and the sum is over ri ≥ 0 with r1 +. . .+r p = p. In particular,


f11...1 is multilinear.
Definition 4.7. The multilinear function f11...1 ∈ P(W p ) is called the polarization
of the polynomial function f ∈ P p (W ). We shall denote P f = f11...1 . On the
other hand, given a multilinear function F : W p → C, the polynomial function
RF ∈ P(W ), defined by RF(w) := F(w, . . . , w) for w ∈ W , is called the restitution
of F.

Let us denote by P(W p )(1,1,...,1) the space of multilinear functions on W p .


Proposition 4.8. The polarization map P : P p (W ) → P(W p )(1,1,...,1) , f 7→ P f , and
the restitution map R : P(W p )(1,1,...,1) → P p (W ), F 7→ RF, satisfy the following
properties:
(1) Both P and R are GL(W )-equivariant linear maps.
124 4. Classical invariant theory

(2) P f is S p -invariant, for f ∈ P p (W ).


(3) RP( f ) = p! f , for f ∈ P p (W ). In particular, R is surjective.

Proof. Parts (1) and (2) follow directly from the definitions.
Setting w1 = . . . = w p = w in (4.5), we obtain that


r
(t1 + . . . + t p ) p f (w) = fr1 ...r p (w, . . . , w)t1r1 · · ·t pp .
r1 ,...,r p

Comparing the coefficients of t1 · · ·t p , we conclude that


p! f (w) = f1,1,...,1 (w, . . . , w) = RP f (w),
whence (3). 
Corollary 4.9. Let G ⊆ GL(W ) be a classical group. Then, by restriction of the
restitution map to the subspace of G-invariants, we have a surjective linear map
R : P(W p )G
(1,1,...,1) −→ P (W ) .
p G

Proof. Since G is a classical group, all the representations of G involved are com-
pletely reducible. By Proposition 4.8, R : P(W p )(1,1,...,1) → P p (W ) is surjective
and G-equivariant, and so R sends a given G-isotypic component of P(W p )(1,1,...,1)
onto the corresponding isotypic component of P p (W ). The corollary follows by
considering the isotypic component of the trivial module. 

Now we are ready to derive the polynomial FFT for GL(V ) from the multilin-
ear FFT for GL(V ).

Proof of Theorem 4.5. Set G = GL(V ). It suffices to prove the following.


Claim. P p (V k ⊕V ∗m )G = 0, for p odd. The space P2d (V k ⊕V ∗m )G is spanned
by ⟨ j1 |i∗1 ⟩⟨ j2 |i∗2 ⟩ · · · ⟨ jd |i∗d ⟩, where 1 ≤ ia ≤ m, 1 ≤ ja ≤ k for each a = 1, . . . , d.
Let us set W = V k ⊕V ∗m , and let p be any positive integer to start with. Keeping
Lemma 4.3 in mind, we observe that P(W p )G (1,1,...,1) is a direct sum of spaces of
G-invariant multilinear functions on direct sums of p copies among V and V ∗ in
various order. By Theorem 4.4, we have P(W p )G (1,1,...,1) = 0, for p odd; moreover,
P(W )(1,1,...,1) is spanned by the functions of the form ⟨ j1 |i∗1 ⟩⟨ j2 |i∗2 ⟩ · · · ⟨ jd |i∗d ⟩.
2d G

Now the claim follows by applying Corollary 4.9. 

4.2. Polynomial FFT for classical groups


In this section, we formulate and establish the polynomial FFT for the orthogonal
and symplectic Lie groups in a uniform manner (see Theorem 4.13). A theorem of
Weyl allows us to reduce the polynomial FFT to a special basic case. This basic
case is then proved directly by induction.
4.2. Polynomial FFT for classical groups 125

4.2.1. A reduction theorem of Weyl. Given 0 ≤ s ≤ k, we have a natural inclu-


sion V s ⊂ V k by identifying (v1 , . . . , vs ) with (v1 , . . . , vs , 0, . . . , 0). We also have
the natural projection from V k to V s which sends (v1 , . . . , vs , . . . , vk ) to (v1 , . . . , vs ),
which allows us to view P(V s ) as a subalgebra of P(V k ). Note that V k is natu-
rally a GLk -module, with the action π given by the multiplication by g−1 on the
right: π(g)(v1 , . . . , vk ) = (v1 , . . . , vk )g−1 . The induced GLk -action, denoted by π′ ,
on P(V k ) is given by (π′ (g) f )(v1 , . . . , vk ) = f ((v1 , . . . , vk )g).
Though the goal of this subsection is Theorem 4.11 due to H. Weyl, it is natural
to work in a more general setting first. Given a subset S of a module M over a group
G, we denote by ⟨S⟩G the G-submodule of M generated by S.
Theorem 4.10. Let G be an arbitrary subgroup of GL(V ), and let N = dimV .
Assume that k ≥ N, and L is a GLk -submodule of P(V k ). Then, as a GLk -module,
L is generated by the intersection L ∩ P(V N ), i.e., L = ⟨L ∩ P(V N )⟩GLk .

Proof. We will assume k > N, since the case k = N is trivial.


Since P(V k ) is a complete reducible GLk -module, we may assume without
loss of generality that L is an irreducible GLk -module. Denote by Uk the subgroup
of GLk which consists of upper triangular k × k-matrices with all diagonal entries
being 1. Then P(V k )Uk is the space of highest weight vectors in P(V k ).
Claim. We have P(V k )Uk ⊆ P(V N ).
Note by the standard highest weight theory that LUk ̸= 0. Granting the claim,
we have LUk ⊆ P(V k )Uk ⊆ P(V N ), and hence, L ∩ P(V N ) ̸= 0. Then, it follows by
the irreducibility of L that L = ⟨L ∩ P(V N )⟩GLk .
So it remains to prove the claim. Recall dimV = N, and introduce the following
Zariski-open subset in V k :
Z = {(v = (v1 , . . . , vk ) ∈ V k | v1 , . . . , vN are linearly independent in V }.
Given v = (v1 , . . . , vk ) ∈ Z, we can find α1 , . . . , αk−1 ∈ C such that
α1 v1 + . . . + αk−1 vk−1 + vk = 0.
Hence, there exists u ∈ Uk such that π(u)(v1 , . . . , vk−1 , vk ) = (v1 , . . . , vk−1 , 0). For
example, the following element
 −1
1 0 ··· 0 α1
0 1 ··· 0 α2 
 
 .. .. .. .. .. 
u= . . . . . 

 .. .. .. .. 
. . . . αk−1 
0 ··· ··· 0 1
will do the job.
Let f ∈ P(V k )Uk . Then, f (v) = f (v1 , . . . , vk−1 , 0), for v ∈ Z. By induction on
k > N, we see that f (v) = f (v1 , . . . , vN , 0, . . . , 0), for v ∈ Z. Since the polynomial
126 4. Classical invariant theory

function f is determined by its restriction to the Zariski-open subset Z, f (v) =


f (v1 , . . . , vN , 0, . . . , 0), for all v ∈ V . This completes the proof of the claim, and
hence of the theorem. 
Theorem 4.11 (Weyl). Let G be an arbitrary subgroup of GL(V ), let N = dimV ,
and assume k ≥ N. Then, P(V k )G is generated by ⟨P(V N )G ⟩GLk . In particular,
if a subspace S generates the algebra P(V N )G , then ⟨S⟩GLk generates the algebra
P(V k )G .

Proof. Thanks to the commuting actions of G ⊆ GL(V ) and of GLk on P(V k ),


we see that P(V k )G is a GLk -module. The first part of the theorem follows by
specializing Theorem 4.10 to the case L = P(V k )G and noting that P(V N )G =
P(V k )G ∩ P(V N ).
Denote by A the subalgebra of P(V k ) generated by ⟨S⟩GLk . Then, A is con-
tained in the algebra P(V k )G , thanks to ⟨S⟩GLk ⊆ P(V k )G . On the other hand,
A ⊇ S, and so the algebra A contains the algebra P(V N )G which S generates. Since
it follows by definition that A is GLk -stable, A contains ⟨P(V N )G ⟩GLk , and so we
can apply the first part of the theorem to conclude that A contains P(V k )G . Hence,
A = P(V k )G . 

The following proposition will be useful later on.


Proposition 4.12. Let G be a subgroup of GL(V ), and let N = dimV . Assume
that S is a subspace of P(V N )G which generates the algebra P(V N )G . Then, for
k ≤ N, the algebra of invariants P(V k )G is generated by the restrictions Sk :=
{restrictions to V k of elements in S}. Moreover, Sk = S ∩ P(V k ) if S is GLN -stable.

Proof. The inclusion V k → V N and the projection V N → V k induce the restriction


map res : P(V N ) → P(V k ) and the inclusion P(V k ) ,→ P(V N ), respectively. Since
either of the compositions
res res
P(V k ) ,→ P(V N ) → P(V k ), P(V k )G → P(V N )G → P(V k )G
is the identity map, so res : P(V N )G → P(V k )G is surjective. Hence, P(V k )G is
generated by the restrictions Sk .
Now assume S is GLN -stable. In particular, TN acts on S semisimply, where Tk
denotes the diagonal subgroup of GLk , for k ≤ N. Note that TN = Tk × TN−k . Both
Sk and S ∩ P(V k ) can easily be seen to be equal to the subspace of S where TN−k
acts trivially, and hence they must be equal. 

4.2.2. The symplectic and orthogonal groups. Let V be a vector space equipped
with a non-degenerate symmetric or skew-symmetric bilinear form ω : V ×V → C.
We denote by G(V, ω) the subgroup of GL(V ) which preserves the form ω, that is,
G(V, ω) = {g ∈ GL(V ) | ω(gv1 , gv2 ) = ω(v1 , v2 ), ∀v1 , v2 ∈ V }.
4.2. Polynomial FFT for classical groups 127

The group G(V, ω) is called an orthogonal group and denoted by O(V ) if ω is


symmetric, and a symplectic group and denoted by Sp(V ) if ω is skew-symmetric.
The Lie algebra of the group G(V, ω), called the orthogonal and symplectic Lie
algebras respectively, is
g(V, ω) = {x ∈ gl(V ) | ω(xv1 , v2 ) = −ω(v1 , xv2 ), ∀v1 , v2 ∈ V }.
Let V = CN , which will mostly be viewed as the space of column vectors. A
non-degenerate symmetric (respectively, skew-symmetric) bilinear form relative to
the standard basis {e1 , . . . , eN } in CN is determined by its associated non-singular
symmetric (respectively, skew-symmetric) matrix J. That is, ω(v1 , v2 ) = vt1 Jv2 , for
v1 , v2 ∈ CN . Note that N is necessarily even in the skew-symmetric case. When a
form ω on V = CN has its associated matrix J, we have
G(V, ω) = {X ∈ GLN | X t JX = J}.
Different choices of non-singular symmetric (respectively, skew-symmetric)
N × N matrices J lead to isomorphic orthogonal (respectively, symplectic) groups
in different matrix forms. Some useful choices of the non-singular symmetric ma-
trices are the identity matrix IN , or the following anti-diagonal matrix
 
0 ··· 0 1
 0 · · · 1 0
 
(4.6) JN :=  . . . .  ,
 .. .. .. .. 
1 ··· 0 0
or  
( ) 0 Iℓ 0
0 Iℓ Iℓ 0 0 (for N = 2ℓ + 1).
(for N = 2ℓ),
Iℓ 0
0 0 1
When we choose ω to be the standard symmetric form on V = CN corresponding
to the identity matrix IN , O(V ) is simply the group ON of all N × N orthogonal
matrices X, i.e., X t X = IN .
On the other hand, some common choices for non-singular skew-symmetric
matrices with N = 2ℓ used to define the symplectic Lie group and algebra are
 
Ω 0 ··· 0
( ) 0 Ω ··· 0
0 Iℓ  
, or  . . . . ,
−Iℓ 0 . .
. . . . .
.
0 0 ··· Ω
( )
0 1
where Ω = . Another common choice of a non-singular skew-symmetric
−1 0
matrix is
( )
0 Jℓ
(4.7) .
−Jℓ 0
128 4. Classical invariant theory

The choices of (4.6) and (4.7) play prominent roles in the later Sections 5.3.2
and 5.3.1, respectively.

4.2.3. Formulation of the polynomial FFT. Let V = CN be equipped with a


symmetric (respectively, skew-symmetric) form ω+ (respectively, ω− ) whose as-
sociated matrix is J+ (respectively, J− ). It will be convenient to write G± (V ) for
G(V, ω± ) as this allows to treat the + and − cases uniformly.
Given k ∈ N, we denote by

k = {k × k complex symmetric matrices},


SM+
SM−
k = {k × k complex skew-symmetric matrices}.

We have a natural identification of G± (V )-modules between V k and the space


MNk of N × k complex matrices (on which G± (V ) acts by left multiplication). De-
fine maps
τ+ : V k ≡ MNk −→ SM+
k, τ+ (X) = X t J+ X,
τ− : V k ≡ MNk −→ SM−
k, τ− (X) = X t J− X.
It can be checked that τ± (X)t = ±τ± (X), so τ± are well-defined.
The map τ± is G± (V )-equivariant, where we let G± (V ) act on SM±
k trivially.
Indeed, for g ∈ G± (V ), X ∈ MNk , we have
τ± (gX) = (gX)t J± (gX) = X t gt J± gX = X t J± X = τ± (X).
Hence, the pullback via τ± gives rise to an algebra homomorphism
± (V )
τ∗± : P(SM±
k ) −→ P(V )
k G
, τ∗± ( f ) = f ◦ τ± ,
±
where P(V k )G (V ) denotes the algebra of G± (V )-invariant polynomials in V k . In-
deed, it follows from the G± (V )-equivariance of τ± that the image of τ∗± lies in the
G± (V )-invariant subalgebra of P(V k ).
Write X = (v1 , . . . , vk ) with vi ∈ V = CN . It follows by definition that
(X t J± X)i j = ω± (vi , v j ).
Denote by the same notation the restriction to SM± k the matrix coefficients xi j on
Mkk , for 1 ≤ i, j ≤ k. Introduce the G± (V )-invariant functions (i| j), for 1 ≤ i, j ≤ k,
by letting
(i| j)(v1 , . . . , vk ) := ω± (vi , v j ).
Note these functions are not independent in general, and
(i| j) = ±( j|i)
in the ± case, respectively. It follows by definitions that
τ∗± (xi j )(v1 , . . . , vk ) = (i| j).
4.2. Polynomial FFT for classical groups 129

Theorem 4.13 (Polynomial FFT for G± (V )). The homomorphism τ∗± : P(SM± k )→
± (V ) ± (V )
P(V ) k G is surjective. Equivalently, P(V )
k G is generated by the functions
(i| j), where 1 ≤ i ≤ j ≤ k in the + case and 1 ≤ i < j ≤ k in the − case.

The proof of Theorem 4.13 will be given in the subsequent subsections.


Remark 4.14. One can show that the image of τ± consists of all the matrices
Z ∈ SM± k such that rank(Z) ≤ min{k, dimV }. In particular, if dimV ≥ k, then τ± is
surjective. Under the assumption that dimV ≥ k, τ∗± is injective and so τ∗± is actu-
ally an isomorphism. Equivalently, the algebra of G± (V )-invariant polynomials in
V k is a free polynomial algebra generated by the k(k ± 1)/2 polynomials as listed
in Theorem 4.13.
Remark 4.15. The non-degenerate form ω± induces a canonical isomorphism of
G± (V )-modules: V ∗ ∼= V . Via this isomorphism, we can obtain from Theorem 4.13
a (seemingly more general) polynomial FFT for the algebra of G± (V )-invariant
polynomials in (V ∗ )m ⊕V k .

4.2.4. From basic to general polynomial FFT. Let us now refer to the statement
in the polynomial FFT for G± (V ) as formulated in Theorem 4.13 as FFT(k), for
k ≥ 1. Set N = dimV as before.
Assuming a basic case FFT(N) holds for now, let us complete the proof of
FFT(k) for all k. Denote by Sk the subspace spanned by the functions (i| j), where
1 ≤ i, j ≤ k.
FFT(N) =⇒ FFT(k) for k < N. (Indeed, the same argument below shows that
FFT(ℓ) =⇒ FFT(k) for all ℓ ≥ k ≥ 1.) The subspace SN is GLN -stable. Moreover,
SN ∩ V k is simply Sk . Now FFT(k) follows by applying FFT(N) and Proposi-
tion 4.12 for G = G± (V ) and S = SN .
FFT(N) =⇒ FFT(k) for k > N. According to FFT(N), SN generates the al-
gebra P(V N )G . Observe that ⟨SN ⟩GLk = Sk . Hence, by Weyl’s Theorem 4.11, the
algebra P(V k )G is generated by Sk , whence FFT(k).
This completes the proof of Theorem 4.13, modulo the proof of the basic case
FFT(N). We shall prove the basic case FFT(N) as Theorem 4.16 in the next sub-
section.

4.2.5. The basic case. In this subsection, we shall prove the following basic case
of the polynomial FFT for G± (V ) (that is, Theorem 4.13 for k = dimV ).
± (V )
Theorem 4.16. Let N = dimV . Then, the algebra P(V N )G is generated by the
functions (i| j), for 1 ≤ i, j ≤ N.

Proof. We proceed by induction on N. We treat the cases of orthogonal and sym-


plectic groups separately, though the overall strategy of the proof is the same.
130 4. Classical invariant theory

The orthogonal group case. Let V = CN , and take (x, y) = ∑Ni=1 xi yi , where
x = (x1 , . . . , xN ), y = (y1 , . . . , yN ) in CN , so O(V ) = ON . Denote
V1 := CN−1 × {0} ⊆ V
and denote by (·, ·)1 the bilinear form on V1 obtained by restriction from (·, ·) on V .
For N = 1, O1 = {±1}, Theorem 4.16 is clear.
Assume now that N > 1. We have an orthogonal decomposition V = V1 ⊕ CeN
with respect to (·, ·).
Let f ∈ P(V N )ON . Since −IN ∈ ON , f (X) = f (−IN · X) = f (−X), so f cannot
contain a nonzero odd degree term. Without loss of generality, let us assume f is
homogeneous of even degree 2d. Below, we will consider the restriction of f to
the Zariski-open subset of vectors v = (v1 , . . . , vN ) of V N given by the inequality
(vN , vN ) ̸= 0.
Let κ ∈ C such that κ2 = (vN , vN ). Since (vN , vN ) = (κeN , κeN ), there exists an
orthogonal matrix, g ∈ ON , such that gvN = κeN . Set
(4.8) v′i = κ−1 gvi , (i = 1, . . . , N − 1).
Then,
f (v) = f (gv) = κ2d f (v′1 , . . . , v′N−1 , eN ) = (vN , vN )d f (v′1 , . . . , v′N−1 , eN ).
Let
(4.9) v′i = v′′i + ti eN (i = 1, . . . , N − 1),
be the orthogonal decomposition in V = V1 ⊕ CeN for v′′i ∈ V1 . Using (4.8), we
compute that
(4.10) ti = (v′i , eN ) = (κ−1 gvi , κ−1 gvN ) = (vi , vN )/(vN , vN ).
i
Writing t I = t1i1 · · ·tN−1
N−1
for multi-indices I = (i1 , . . . , iN−1 ) and regarding ti as inde-
pendent variables, we have a decomposition of the following form:
(4.11) f (v′1 , . . . , v′N−1 , eN ) = ∑ fˆI (v′′1 , . . . , v′′N−1 )t I ,
I

where fˆI are polynomial functions uniquely determined by f . Noting from (4.10)
that ti is invariant under the transformation v′i 7→ gv′i for g ∈ ON−1 , we conclude
from (4.11) that fˆI is ON−1 -invariant for each I.
By induction hypothesis, there exist polynomials φI in the functions (i| j), 1 ≤
i ≤ j ≤ N − 1, such that fˆI = φI when evaluated on (v′′1 , . . . , v′′N−1 ). Using (4.8),
(4.9) and (4.10), we compute that
(vi , v j )(vN , vN ) − (vi , vN )(v j , vN )
(4.12) (v′′i , v′′j ) = (v′i , v′j ) − (ti eN ,t j eN ) = .
(vN , vN )2
Let us recall that we have been considering the restriction of f to the open subset
defined by (vN , vN ) ̸= 0. Plugging (4.12) into fˆI = φI , we have f (v) = (vN , vN )−p F1 ,
4.2. Polynomial FFT for classical groups 131

for some polynomial F1 in (vi , v j ), i ≤ j ≤ N, and some positive integer p, when-


ever (vN , vN ) ̸= 0.
We will be done, if we can show that the polynomial function (vN , vN ) p di-
vides F1 . Indeed, the same type of argument above allows us to conclude that
f (v) = (v1 , v1 )−q F2 , for some polynomial F2 in (vi , v j ), i ≤ j ≤ N, and some posi-
tive integer q, whenever (v1 , v1 ) ̸= 0. In this way, we obtain a polynomial equation
(v1 , v1 )q F1 = (vN , vN ) p F2 ,
which holds on a Zariski-open subset of V N given by the inequalities (v1 , v1 ) ̸= 0
and (vN , vN ) ̸= 0, and hence the polynomial equation must hold on V N . This im-
plies (vN , vN ) p divides F1 , since the polynomials (vN , vN ) and (v1 , v1 ) are relatively
prime.

The symplectic group case. Let V = CN be equipped with the standard basis
{e1 , . . . , eN }, where N = 2ℓ is even. For definiteness, let us take the symplectic
group SpN = Sp(V ) defined via the symplectic form ω on V :

ω(x, y) = ∑ (x2i−1 y2i − x2i y2i−1 ),
i=1

where x = (x1 , . . . , xN ), y = (y1 , . . . , yN ) in CN . We have an orthogonal decomposi-


tion V = V1 ⊕W with respect to the symplectic form ω, where
V1 := CN−2 × {0} ⊆ V, W := CeN−1 ⊕ CeN .
We shall denote by ω1 the bilinear form on V1 induced from ω on V , and regard
SpN−2 = Sp(V1 ) as a natural subgroup of SpN in this way.
Let f ∈ P(V N )SpN . Since −IN ∈ SpN , we may assume without loss of gen-
erality that f is homogeneous of even degree 2d, just as for the ON case above.
Below, we will consider the restriction of f to the Zariski-open subset of vectors
v = (v1 , . . . , vN ) of V N given by the inequality ω(vN−1 , vN ) ̸= 0.
Choose κ ∈ C such that κ2 = ω(vN−1 , vN ). Since ω(eN−1 , eN ) = 1 and so
ω(vN−1 , vN ) = ω(κeN−1 , κeN ), it is standard to show that there exists g ∈ SpN such
that
gvN−1 = κeN−1 , gvN = κeN .
Set
(4.13) v′i = κ−1 gvi , (i = 1, . . . , N − 2).
Then,
f (v) = f (gv) = κ2d f (v′1 , . . . , v′N−2 , eN−1 , eN )
= ω(vN−1 , vN )d f (v′1 , . . . , v′N−2 , eN−1 , eN ).
132 4. Classical invariant theory

When N = 2, we obtain a polynomial equation f (v1 , v2 ) = f (e1 , e2 )ω(v1 , v2 )d ,


which holds on a Zariski-open set ω(v1 , v2 ) ̸= 0 and hence holds everywhere. This
proves the theorem for N = 2.
Now assume N > 2. Let

(4.14) v′i = v′′i + si eN−1 + ti eN (i = 1, . . . , N − 2),

be the orthogonal decomposition in V = V1 ⊕ W for v′′i ∈ V1 . It follows by (4.13)


that
si = ω(v′i , eN ) = ω(κ−1 gvi , κ−1 gvN ) = ω(vi , vN )/ω(vN−1 , vN ),
(4.15)
ti = −ω(v′i , eN−1 ) = −ω(vi , vN−1 )/ω(vN−1 , vN ).
j j
N−2 i1 i
Write sJ t I = s11 · · · sN−2 t1 · · ·tN−2
N−2
associated to the multi-indices I = (i1 , . . . , iN−2 )
and J = ( j1 , . . . , jN−2 ). Regarding si and ti as independent variables, we have a
decomposition of the following form:

(4.16) f (v′1 , . . . , v′N−2 , eN−1 , eN ) = ∑ fˆIJ (v′′1 , . . . , v′′N−2 )sJ t I ,


I,J

where fˆIJ are polynomial functions uniquely determined by f . Noting from (4.15)
that si ,ti are invariant under the transformation v′i 7→ gv′i for g ∈ SpN−2 , we conclude
from (4.16) that fˆIJ is SpN−2 -invariant for every I, J.
By inductive assumption, there exist polynomials φIJ in the functions (i| j), 1 ≤
i ≤ j ≤ N − 2, such that fˆIJ = φIJ when evaluated at (v′′1 , . . . , v′′N−2 ). Using (4.13),
(4.14) and (4.15), we compute that
ω(v′′i , v′′j ) = ω(v′i , v′j ) + ti s j − sit j
(4.17) ω(vi , v j )ω(vN−1 , vN ) − ω(vi , vN−1 )ω(v j , vN ) + ω(vi , vN )ω(v j , vN−1 )
= .
ω(vN−1 , vN )2
Recall that we have been considering the restriction of f to the open subset defined
by ω(vN−1 , vN ) ̸= 0. Plugging (4.17) into fˆIJ = φIJ , we have f (v) = ω(vN−1 , vN )−p F,
for some polynomial F in ω(vi , v j )i≤ j≤N and some positive integer p, whenever
ω(vN−1 , vN ) ̸= 0. Repeating the same trick used in the orthogonal group case, we
see that the polynomial function ω(vN−1 , vN ) p divides F.
This completes the proof of the theorem. 

4.3. Tensor and supersymmetric FFT for classical groups


In this section, we will use the polynomial FFT for classical groups to derive the
corresponding tensor FFT. Finally, the supersymmetric algebra version of FFT is
derived from the tensor FFT.
4.3. Tensor and supersymmetric FFT for classical groups 133

4.3.1. Tensor FFT for classical groups. Let G = G(V, ω), where V is a vector
space equipped with a non-degenerate symmetric or skew-symmetric bilinear form
ω. Since V ∼
= V ∗ as a G-module, it suffices to consider the G-invariants on the ten-
sor space V instead of a mixed tensor space V ⊗m ⊗(V ∗ )⊗n , for m, n ≥ 0. Observe
⊗m

that −IV ∈ G acts by the scalar multiple by (−1)m on V ⊗m . Hence, (V ⊗m )G = 0


unless m = 2k is even.
Via the canonical identification of G-modules:
V ⊗V ∼
= V ∗ ⊗V ∼
= End(V ),
we obtain a natural G-module isomorphism:

T : V ⊗2k −→ End(V ⊗k ),
which sends u = v1 ⊗ v2 ⊗ . . . ⊗ v2k to T (u) satisfying
T (u)(x1 ⊗ . . . ⊗ xk ) = ω(x1 , v2 )ω(x2 , v4 ) . . . ω(xk , v2k )v1 ⊗ v3 ⊗ . . . ⊗ v2k−1 .
Clearly the identity IV ⊗k is G-invariant, and hence so is T −1 (IV ⊗k ). We describe
T −1 (IV ⊗k ) explicitly as follows. Take a basis { fi } for V and its dual basis { f i } for
V with respect to ω, i.e., ω( fi , f j ) = δi j . Define
N
θ = ∑ fi ⊗ f i , θk = θ
| ⊗ .{z
. . ⊗ θ} .
i=1
k

Then T −1 (IV ⊗k ) = θk .
Recall the following commuting actions:
Φ2k Ψ2k
GL(V ) y V ⊗2k x S2k .
Since G is a subgroup of GL(V ) which commutes with S2k , Ψ2k (σ)(θk ) is G-
invariant for any σ ∈ S2k .
Theorem 4.17 (Tensor FFT for G(V, ω)). Let G = G(V, ω), where ω is a non-
degenerate symmetric or skew-symmetric bilinear form on V . Then (V ⊗m )G = 0
for m odd. Moreover, (V ⊗2k )G has a spanning set {Ψ2k (σ)(θk ) | σ ∈ S2k }.

The tensor FFT for G(V, ω) affords an equivalent dual reformulation on the
G-invariants of (V ∗⊗m ) ∼
= (V ⊗m )∗ . We let (i| j) be the function such that
(i| j)(v1 , . . . , vm ) = ω(vi , v j ).
Denote by
θ∗k = (1|2)(3|4) . . . (2k − 1|2k)
the G-invariant function on (V ⊗m ) which corresponds to the identity under the

isomorphism T ∗ : V ∗⊗2k −→ End(V ⊗k ). Denote by Ψ∗2k the natural action of S2k
on (V ∗⊗2k ), which clearly commutes with the natural action of G.
134 4. Classical invariant theory

Theorem 4.18 (Tensor FFT for G(V, ω), a dual version). Let G = G(V, ω), where
ω is non-degenerate symmetric or skew-symmetric. Then (V ∗⊗m )G = 0 for m odd.
Moreover, (V ∗⊗2k )G has a spanning set {Ψ∗2k (σ)(θ∗k ) | σ ∈ S2k }.

Theorems 4.17 and 4.18 are equivalent, and we shall only present a detailed
proof of the tensor FFT as formulated in Theorem 4.18. The strategy of the proof
can be outlined as follows. First, we introduce an auxiliary polarization variable
X ∈ End V to reformulate this tensor FFT problem as one involving the basic case
of the polynomial FFT as in Theorem 4.16 (note End V ∼
= V N for N = dimV ). Then,
by applying Theorem 4.16, we transfer the problem at hand to an FFT problem for
GL(V ), for which Theorem 4.2 applies.

Proof of Theorem 4.18. Since −IV ∈ G acts on V ⊗m by a multiple of (−1)m , we


have (V ⊗m )G = 0, for m odd.
Now let m = 2k be even. We shall define an injective linear map
Φ : V ∗⊗m −→ Pm (End V ) ⊗V ∗⊗m , λ 7→ Φλ ,
where Φλ is defined as a function on End V × V ⊗m , which is of polynomial de-
gree m on End V and linear on V ⊗m , by letting Φλ (X, w) = ⟨λ, X ⊗m w⟩, for X ∈
End V, w ∈ V ⊗m . The injectivity follows since λ is recovered as Φλ (IV , −).
The space Pm (End V ) ⊗V ∗⊗m carries commuting actions of the groups G and
GL(V ) as follows: g. f (X, w) = f (g−1 X, w) for g ∈ G; h. f (X, w) = f (Xh, h−1 w)
for h ∈ GL(V ). Note that Φ is a G-equivariant map, i.e., g.Φλ = Φg.λ , for g ∈ G.
Also note that Φλ is GL(V )-invariant, i.e., h.Φλ = Φλ , for h ∈ GL(V ). Hence, we
obtain by restriction of Φ the following injective linear map (which can actually be
shown to be an isomorphism, though this stronger statement is not needed below)
( )GL(V )
Φ : (V ∗⊗m )G −→ (Pm (End V ))G ⊗V ∗⊗m .

For notational convenience below, let us fix V = CN (regarded as a space of


column vectors), and ω(x, y) = xt Jy, for x, y ∈ V . In this way, we identify End V as
the space MNN of N × N matrices.
The action of G on End V relevant to Φ above gives rise to a G-equivariant
isomorphism End V ∼= V N . When applying the basic case of the polynomial FFT
(Theorem 4.16), we obtain, for λ ∈ (V ∗⊗m )G , that
Φλ (X, w) = Fλ (X t JX, w),
where Fλ is a polynomial function on SMN± ×V ⊗2k , which is of polynomial degree
k on SMN± and linear on V ⊗2k (here we recall m = 2k).
We have a natural isomorphism SMN+ = ∼ S2 (V ∗ ) in the case of symmetric ω,
and SMN− ∼
= ∧2 (V ∗ ) in the case of skew-symmetric ω. Recall the action of GL(V )
on End V ∼ = MNN relevant to Φ above is given by left multiplication on MNN .
This action induces an action of GL(V ) on SMN± , which corresponds precisely
4.3. Tensor and supersymmetric FFT for classical groups 135

to the GL(V )-action on S2 (V ∗ ) (respectively, ∧2 (V ∗ )) coming from the GL(V )-


module structure on V ∗ . Hence, we have a GL(V )-invariant polynomial function
Fλ : S2 (V ∗ ) ×V ⊗2k → C in the case of symmetric ω (or Fλ : ∧2 (V ∗ ) ×V ⊗2k → C in
the case of skew-symmetric ω), which is linear on V ⊗2k and of polynomial degree
k on S2 (V ∗ ) (or ∧2 (V ∗ )). Thus, by the tensor FFT for GL(V ) (see Theorem 4.2),
Fλ can be written as a linear combination of the functions
( j) ( j)
(M, w) → ∑ ui1 Mi j v1 · · · ∑ uik Mi j vk
i, j i, j

where w is the tensor product of u1 , v1 , . . . , uk , vk ∈ CN in some order, and ui1 ’s


are the coordinates of u1 , and so on. Evaluating Fλ at M = J, we deduce that
λ = Fλ (J, −) is a linear combination of the (i| j)’s. 

4.3.2. From tensor FFT to supersymmetric FFT. Let W = W0̄ ⊕W1̄ be a finite-
dimensional vector superspace. We denote the supersymmetric algebra of W by

S(W ) = S(W0̄ ) ⊗ ∧(W1̄ ).

Let G = GL(V ). We set

(4.18) W0̄ = V m ⊕V ∗n , W1̄ = V k ⊕V ∗ℓ .

As a GL(V )-module, we have a decomposition S(W ) = ⊕r≥0 Sr (W ) by the total


degree, where Sr (W ) is a direct sum of the following tensor spaces

Sαβγδ :=Sα1 (V ) ⊗ . . . ⊗ Sαm (V ) ⊗ Sβ1 (V ∗ ) ⊗ . . . ⊗ Sβn (V ∗ )⊗


(4.19)
∧γ1 (V ) ⊗ . . . ⊗ ∧γk (V ) ⊗ ∧δ1 (V ∗ ) ⊗ . . . ⊗ ∧δℓ (V ∗ ),

where the summation is taken over all nonnegative integers αa , βb , γc , and δd such
that |α| + |β| + |γ| + |δ| = r. Here we have denoted α = (α1 , . . . , αm ), |α| = ∑a αa ,
and similarly for β, |β|, and so on. We regard the space Sαβγδ as a subspace of the
following mixed tensor space

T αβγδ :=V ⊗α1 ⊗ . . . ⊗V ⊗αm ⊗ (V ∗ )⊗β1 ⊗ . . . ⊗ (V ∗ )⊗βn ⊗


(4.20)
V ⊗γ1 ⊗ . . . ⊗V ⊗γk ⊗ (V ∗ )⊗δ1 ⊗ . . . ⊗ (V ∗ )⊗δℓ ,

(which is isomorphic to V ⊗(|α|+|γ|) ⊗ (V ∗ )⊗(|β|+|δ|) ). Hence, the space (Sαβγδ )GL(V )


can be viewed as a subspace of (T αβγδ )GL(V ) . By the tensor FFT for GL(V ) (The-
orem 4.2), we will assume below that |α| + |γ| = |β| + |δ| = N for some N, for
otherwise (T αβγδ )GL(V ) = 0.
Our goal is to understand the GL(V )-invariants in T αβγδ and then in Sαβγδ in
terms of the corresponding quadratic (i.e., degree two) GL(V )-invariants. Recall
from Theorem 4.2 the basic GL(V )-invariant Θ in V ⊗ V ∗ corresponding to the
136 4. Classical invariant theory

identity IV ∈ End(V ). We shall denote by ⟨a, i|b, j∗ ||⟩ the corresponding quadratic
GL(V )-invariant in the following space (which is ∼= V ⊗V ∗ ):

1α1 ⊗ . . . ⊗ 1αa−1 ⊗ (1i−1 ⊗V ⊗ 1αa −i ) ⊗ . . . ⊗ 1αm


⊗ 1β1 ⊗ . . . ⊗ 1βb−1 ⊗ (1 j−1 ⊗V ∗ ⊗ 1βb − j ) ⊗ . . . ⊗ 1βn ⊗ 1|γ| ⊗ 1|δ| ,

pairing the ith copy of V in V αa with the jth copy of V ∗ in (V ∗ )βb , where 1 ≤
a ≤ m, 1 ≤ i ≤ αa and 1 ≤ b ≤ n, 1 ≤ j ≤ βb . In similar notation, we have the
quadratic GL(V )-invariants ⟨a, i|||d, q∗ ⟩, ⟨|b, j∗ |c, p|⟩, and ⟨||c, p|d, q∗ ⟩, where 1 ≤
a ≤ m, 1 ≤ i ≤ αa , 1 ≤ b ≤ n, 1 ≤ j ≤ βb , 1 ≤ c ≤ k, 1 ≤ p ≤ γc , 1 ≤ d ≤ ℓ, 1 ≤ q ≤ δd .
By the tensor FFT for GL(V ) (Theorem 4.2), the GL(V )-invariants of T αβγδ are
spanned by Θσ for σ ∈ SN , and each Θσ is a suitable product of N invariants
among ⟨a, i|b, j∗ ||⟩, ⟨a, i|||d, q∗ ⟩, ⟨|b, j∗ |c, p|⟩, and ⟨||c, p|d, q∗ ⟩ such that each copy
of V and V ∗ in T αβγδ is used exactly in one pairing.
Note that the subspace (Sαβγδ )GL(V ) is obtained from the space (T αβγδ )GL(V )

by a (partial) supersymmetrization. Hence, applying the Sα × Sβ × S− γ × Sδ -
supersymmetrization operator ϖαβγδ to the Θσ gives us a spanning set {ϖαβγδ (Θσ )}
for (Sαβγδ )GL(V ) , where Sα = Sα1 × . . . × Sαm and by supersymmetrization we
mean the symmetrization over the subgroup Sα × Sβ and the anti-symmetrization
(denoted by the minus sign superscript above) over the subgroup Sγ × Sδ . Let us
denote by

(4.21) ⟨a|b∗ ||⟩, ⟨a|||d ∗ ⟩, ⟨|b∗ |c|⟩, ⟨||c|d ∗ ⟩

the quadratic GL(V )-invariants resulting from the supersymmetrizations of the


GL(V )-invariants ⟨a, i|b, j∗ ||⟩, ⟨a, i|||d, q∗ ⟩, ⟨|b, j∗ |c, p|⟩, and ⟨||c, p|d, q∗ ⟩, respec-
tively (note that these supersymmetrizations are independent of i, j, p, q, whence
the chosen notation). Then, from the product form of Θσ we derive that ϖαβγδ (Θσ )
is the corresponding product of N invariants among ⟨a|b∗ ||⟩, ⟨a|||d ∗ ⟩, ⟨|b∗ |c|⟩, and
⟨||c|d ∗ ⟩.

The cases for G = O(V ) or Sp(V ) can be pursued in an analogous manner. Set

(4.22) W0̄ = V m , W1̄ = V k .

We have a similar decomposition of the G-module S(W ) = ⊕α,γ Sαγ where Sαγ is
modified from (4.19) with β and δ being zero. The space Sαγ is a subspace of
the tensor space T αγ , where T αγ is modified from (4.20) with β and δ being zero.
Hence, the space (Sαγ )G can be viewed as a subspace of (T αγ )G . By the tensor FFT
(Theorem 4.17), we may assume below that |α| + |γ| is even, say, equal to 2N.
4.3. Tensor and supersymmetric FFT for classical groups 137

Our goal is to understand the G-invariants in T αγ and then in Sαγ in terms of


the corresponding quadratic (i.e., degree two) G-invariants. Recalling from Theo-
rem 4.17 the basic G-invariant θ in V ⊗V , we shall denote by (a, i|c, p) the corre-
sponding quadratic G-invariant in the following space (which is ∼
= V ⊗V ):
1α1 ⊗ . . . ⊗ 1αa−1 ⊗ (1i−1 ⊗V ⊗ 1αa −i ) ⊗ . . . ⊗ 1αm
⊗ 1γ1 ⊗ . . . ⊗ 1γc−1 ⊗ (1 p−1 ⊗V ⊗ 1γc −p ) ⊗ . . . ⊗ 1γk ,
pairing the ith copy of V in V αa with the pth copy of V in V γc , where 1 ≤ a ≤ m, 1 ≤
i ≤ αa and 1 ≤ c ≤ k, 1 ≤ p ≤ γc . In addition, we have G-invariants (a, i; a′ , i′ |),
which pairs the ith copy of V in V αa with the i′ th copy of V in V αa′ , where 1 ≤
a ≤ a′ ≤ m, 1 ≤ i ≤ αa , 1 ≤ i′ ≤ αa′ (and we impose i < i′ when a = a′ ); In similar
notation, we have G-invariants (|c, p; c′ , p′ ), which pairs the pth copy of V in V γc
with the p′ th copy of V in V γc′ , where 1 ≤ c ≤ c′ ≤ k, 1 ≤ p ≤ γc , 1 ≤ p′ ≤ γc′ (and
we impose p < p′ when c = c′ ).
Then by the tensor FFT (Theorem 4.17), the G-invariants of T αγ are spanned by
θσ for σ ∈ S2N , and each θσ is a suitable product of N invariants among (a, i|c, p),
(a, i; a′ , i′ ) and (c, p; c′ , p′ ) so that each V in T αγ is used exactly in one pairing.
The subspace (Sαγ )G is obtained from the space (T αγ )G by a (partial) super-
symmetrization. Hence, applying the Sα × S− γ -supersymmetrization operator ϖαγ
to the θσ gives us a spanning set {ϖαγ (θσ )} for (Sαγ )G . Let us denote by
(4.23) (a|c), (a; a′ |), (|c; c′ )
the quadratic G-invariants which resulted from the supersymmetrizations of the G-
invariants (a, i|c, p), (a, i; a′ , i′ |) and (|c, p; c′ , p′ ), respectively, where 1 ≤ a, a′ ≤ m,
a ≤ a′ , 1 ≤ c, c′ ≤ k, and c ≤ c′ . Note that (a; a|) = 0 for G = Sp(V ) and (|c; c) = 0
for G = O(V ). Hence, we derive from the product form of θσ that ϖαγ (θσ ) is a
product of N invariants among (a|c), (a; a′ |), (|c; c′ ).
Summarizing, we have established the following.
Theorem 4.19 (Supersymmetric FFT). Let G = GL(V ), O(V ), or Sp(V ). Let W
be as given in (4.18) for GL(V ) and in (4.22) for O(V ) or Sp(V ), respectively.
The algebra of G-invariant in the supersymmetric algebra S(W ) is generated by its
quadratic invariants, given in (4.21) for GL(V ) and in (4.23) for O(V ) or Sp(V ),
respectively.
Remark 4.20. Theorem 4.19 has two most important specializations:
(1) When W1̄ = 0, the supersymmetric algebra on W reduces to the symmetric
algebra S(W0̄ ). The corresponding FFT admits an equivalent reformulation as a
polynomial FFT which describes the G-invariants in a polynomial algebra on a
direct sum of copies of V (and V ∗ ).
(2) When W0̄ = 0, the supersymmetric algebra on W reduces to the exterior
algebra ∧(W1̄ ). The G-invariants in this case are known as skew invariants.
138 4. Classical invariant theory

4.4. Notes
Section 4.1. The materials here are mostly classical and standard. Weyl’s book
[We] on invariant theory was influential, where the terminology of First Funda-
mental Theorem (FFT) for invariants was introduced. See Howe [H1, H2]; also
see the notes of Kraft-Procesi [KP]. Theorem 4.1 on the finite-generation of an
invariant subalgebra of the supersymmetric algebra is somewhat novel, as it is usu-
ally formulated for a polynomial algebra only.
Section 4.2. The results here are classical. Our proof of Weyl’s theorem fol-
lows Kraft-Procesi [KP]. Weyl’s theorem allows us to reduce the proof of the
polynomial FFT for classical groups in general to a basic case. The basic case is
then proved by induction on dimensions, and our presentation here does not differ
much from Goodman-Wallach [GW]. In [GW], the general polynomial FFT was
derived using the tensor FFT instead of Weyl’s theorem, while in Weyl [We], the
Cappelli identity was used essentially in the proof of the polynomial FFT.
Section 4.3. The tensor FFT for classical groups can be found in Weyl [We].
The quick proof of the tensor FFT for the orthogonal groups here is due to Atiyah-
Bott-Patodi [ABP], and it works for the symplectic groups similarly (also see a
presentation in Goodman-Wallach [GW]). The supersymmetric FFT for classical
groups, which is a common generalization of the polynomial FFT and skew FFT,
is taken from Howe [H1]. The supersymmetric FFT will be used in next chapter
on Howe duality for Lie superalgebras.
Chapter 5

Howe duality

The goal of this chapter is to explain Howe duality for classical Lie groups, Lie
algebras and superalgebras, and applications to irreducible character formulas for
Lie superalgebras. Like Schur duality, Howe duality involves commuting actions
of a pair of Lie group G and superalgebra g′ on a supersymmetric algebra S(U), and
S(U) is naturally an irreducible module over a Weyl-Clifford algebra WC. In the
main examples, G is one of the three classical groups GL(V ), Sp(V ), or O(V ), and
g′ is a classical Lie algebra or superalgebra, and U is a direct sum of copies of the
natural G-module and its dual. According to the FFT for classical invariant theory
in Chapter 4, when applied to the G-action on the associated graded gr WC, the
basic invariants generating (gr WC)G turn out to form the associated graded space
for a Lie (super)algebra g′ . From this it follows that the algebra of G-invariants
WCG is generated by g′ .
The most important cases of S(U) are the two cases of a symmetric algebra
and of an exterior algebra, corresponding respectively to U1̄ = 0 and U0̄ = 0. In
these two cases, the Howe dualities only involve Lie groups and Lie algebras, and
they may be regarded as Lie theoretic reformulations of the classical polynomial
invariant and skew invariant theories, respectively, treated in Chapter 4.
Howe duality provides a realization of an important class of irreducible mod-
ules, called oscillator modules, of Lie (super)algebra g′ . The irreducible G-modules
and irreducible oscillator g′ -modules appearing in the decomposition of S(U) are
paired in “duality”, which is much stronger than a simple bijection. Character
formulas for the irreducible oscillator g′ -modules are then obtained, via a compar-
ison with Howe duality involving classical groups G and infinite-dimensional Lie
algebras, which we develop in detail.

139
140 5. Howe duality

5.1. Weyl-Clifford algebra and classical Lie superalgebras


In this section, we introduce the mixture of Weyl algebra and Clifford algebra uni-
formly in the framework of superalgebras. We formulate the basic properties of the
Weyl-Clifford algebra and its connections to classical Lie algebras and superalge-
bras. We then establish a general duality theorem, which works particularly well
in the setting of Weyl-Clifford algebra.

5.1.1. The Weyl-Clifford algebra. We shall denote by


S(U) = S(U0̄ ) ⊗ ∧(U1̄ )
the supersymmetric algebra on a vector superspace U = U0̄ ⊕ U1̄ . In a first round
of reading of the constructions below, the reader is advised to understand first the
two special cases with U1̄ = 0, and then U0̄ = 0, respectively.
The superspace S(U) is spanned by monomials of the form u1 . . . u p w1 . . . wq ,
for u1 , . . . , u p ∈ U0̄ and w1 , . . . , wq ∈ U1̄ ; Note that ui u j = u j ui , ui wk = wk ui , and
wk wℓ = −wℓ wk . The space S(U) is Z×Z2 -graded by letting the degree of a nonzero
element u1 . . . u p w1 . . . wq be (p + q, q̄), where q̄ denotes q modulo 2.
For x ∈ U0̄ (respectively, x ∈ U1̄ ), we define the left multiplication operator of
degree (1, 0̄) (respectively, (1, 1̄))
Mx : S(U) −→ S(U), y 7→ xy.

We regard U0̄∗ ⊆ U ∗ by extending u∗ ∈ U0̄∗ such that ⟨u∗ , w⟩ = 0 for all w ∈ U1̄ ,
and similarly we have U1̄∗ ⊆ U ∗ . In this way, we identify U ∗ ≡ U0̄∗ ⊕U1̄∗ as a vector
superspace. For u∗ ∈ U0̄∗ , we define the even derivation Du∗ of degree (−1, 0̄), and
for w∗ ∈ U1̄∗ , we define the odd derivation Dw∗ of degree (−1, 1̄) as follows:
Du∗ : S(U) −→S(U), Dw∗ : S(U) −→ S(U),
p

∑ ⟨u∗ , ui ⟩u1 . . . ûi . . . u p w1 . . . wq ,
D
u1 . . . u p w1 . . . wq −→
u

i=1
q

∑ (−1)k−1 ⟨w∗ , wk ⟩u1 . . . u p w1 . . . ŵk . . . wq ,
D
u1 . . . u p w1 . . . wq −→
w

k=1

where ui ∈ U0̄ and w j ∈ U1̄ . Here and below ûi means the omittance of ui , and so
on.
The linear operators Mu and Du∗ , for u ∈ U0̄ , u∗ ∈ U0̄∗ , in the superalgebra
End(S(U)) defined above are even, while Mw , Dw∗ ∈ End(S(U)), for w ∈ U1̄ , w∗ ∈
U1̄∗ , are odd. We shall denote by [·, ·] the supercommutator among linear operators.
Lemma 5.1. For x∗ , y∗ ∈ U ∗ and x, y ∈ U, we have
[Dx∗ , Dy∗ ] = [Mx , My ] = 0, [Dx∗ , My ] = ⟨x∗ , y⟩1.
5.1. Weyl-Clifford algebra and classical Lie superalgebras 141

Let us reformulate. Set


U = U ⊕U ∗ .
Then U is a superspace U = U 0̄ ⊕U
U 1̄ with
(5.1) U 0̄ = U0̄ ⊕U0̄∗ , U 1̄ = U1̄ ⊕U1̄∗ .
Let ⟨·, ·⟩′ be the even skew-supersymmetric bilinear form on the superspace U
defined as follows: for u1 , u2 ∈ U0̄ , u∗1 , u∗2 ∈ U0̄∗ , w1 , w2 ∈ U1̄ , and w∗1 , w∗2 ∈ U1̄∗ , we
set
⟨u1 + u∗1 , u2 + u∗2 ⟩′ = ⟨u∗1 , u2 ⟩ − ⟨u∗2 , u1 ⟩,
⟨w1 + w∗1 , w2 + w∗2 ⟩′ = ⟨w∗1 , w2 ⟩ + ⟨w∗2 , w1 ⟩,
⟨u1 + u∗1 , w1 + w∗1 ⟩′ = 0.
Define a linear map
ι : U −→ End(S(U))
by letting
ι(x) = Mx , ι(x∗ ) = Dx∗ , for x ∈ U, x∗ ∈ U ∗ .
It is straightforward to check that Lemma 5.1 affords the following equivalent re-
formulation.
Lemma 5.2. We have
(5.2) [ι(a), ι(b)] = ⟨a, b⟩′ 1, for a, b ∈ U .

The subalgebra in End(S(U)) generated by ι(U U ) will be denoted by WC(U U)


and will be referred to as the Weyl-Clifford algebra. When U1̄ = 0, WC(U U)
reduces to the usual Weyl algebra W(U U 0̄ ). Choose a basis x1 , . . . , xm for the m-
dimensional space U0̄ , with dual basis x1∗ , . . . , xm ∗ , so that S(U ) may be identified

with C[x1 , . . . , xm ]. The operators Mxi can then be identified with the multiplica-
tion operators xi , while the derivations Dxi∗ can be identified with the differential
operators ∂x∂ i ≡ ∂i . Hence, W(U U 0̄ ) can be identified with the algebra of polyno-
mial differential operators on C[x1 , . . . , xm ]. When U0̄ = 0, WC(U U ) is isomorphic
to the Clifford algebra Cl(U U 1̄ ). In the same spirit as above, let η1 , . . . , ηn be
a basis for the n-dimensional space U1̄ , and let η∗1 , . . . , η∗n be its dual basis, so
that ∧(U1̄ ) ≡ ∧(η1 , . . . , ηn ), the exterior superalgebra generated by η1 , . . . , ηn . The
multiplication operators Mηi correspond to left multiplications by ηi , and shall be
denoted accordingly by ηi . The derivations Dη∗i correspond to the odd differential
operators ∂η∂ i ≡ ði . Hence, Cl(U U 1̄ ) can be regarded as the superalgebra of differen-
tial operators on ∧(η1 , . . . , ηn ).
For general U, there is a algebra isomorphism WC(U ∼ W(U
U) = U )⊗Cl(U U ) and
0̄ 1̄
an identification S(U) ≡ C[x, η] := C[x1 , . . . , xm ]⊗∧(η1 , . . . , ηn ). The superalgebra
WC(U U ) can then be identified with the superalgebra of polynomial differential
operators on the superspace U, that is, the subalgebra of End(S(U)) generated by
142 5. Howe duality

the multiplication operators xi , η j , and the (super)derivativations ∂i , ð j , for 1 ≤ i ≤


m, 1 ≤ j ≤ n.

5.1.2. A filtration on Weyl-Clifford algebra. The linear map ι|U : U → WC(U U)


induces an injective algebra homomorphism κ : S(U) → WC(U U ), and similarly,
ι|U ∗ : U ∗ → WC(UU ) induces an injective algebra homomorphism κ∗ : S(U ∗ ) →
WC(U U ).

Proposition 5.3. The map m : S(U) ⊗ S(U ∗ ) → WC(U U ), x ⊗ y 7→ κ(x)κ∗ (y) is a


linear isomorphism. Moreover, the algebra WC(U U ) is isomorphic to the algebra
generated by ι(U
U ) subject to the relation (5.2).

Proof. The two statements in the proposition are equivalent. It follows by def-
inition and (5.2) that the map m is surjective. The injectivity of m is equivalent
to the claim that the “monomials” of an ordered basis for U and U ∗ form a basis
of WC(U U ), and the claim can then be established by a straightforward inductive
argument.
Alternatively, the proposition for WC(U U ) reduces to the well-known counter-
parts for W(U U 0̄ ) and Cl(U
U 1̄ ), thanks to WC(U U) ∼= W(U U 0̄ ) ⊗ Cl(U
U 1̄ ) (Note that
W(U U 0̄ ) and Cl(U
U 1̄ ) commute). 

We continue to identify WC(U U ) with the superalgebra of polynomial dif-


ferential operators generated by the multiplication operators xi , η j , and the (su-
per)derivativations ∂i , ð j , for 1 ≤ i ≤ m, 1 ≤ j ≤ n. By Proposition 5.3, WC(U U)
α β γ δ α α1 α
has a linear basis {x η ∂ ð | α, γ ∈ Z+ , β, δ ∈ Z2 }, where x = x1 · · · xn for α =
n n n

(α1 , . . . , αn ) and ηβ , ∂γ , ðδ are defined similarly. Define the degree deg xα ηβ ∂γ ðδ =


|α| + |β| + |γ| + |δ|, where |α| = α1 + . . . + αn and |β|, |γ|, |δ| are similarly defined.
The Weyl-Clifford algebra WC(U U ) carries a natural filtration {WC(U U )k }k≥0
by letting WC(U U )k be spanned by the elements xα ηβ ∂γ ðδ of degree no greater
than k. Clearly, WC(U U )k WC(U U )ℓ ⊆ WC(U U )k+ℓ , and so WC(U U ) is a filtered al-
gebra. Actually, W(U U 0̄ ) and Cl(U U 1̄ ) are filtered subalgebra of the filtered algebra
WC(U U ), and WC(U U ) = W(U U 0̄ ) ⊗ Cl(U U 1̄ ) as a tensor product of filtered algebras.
As usual, the associated graded algebra for WC(U U ) is defined to be


U) =
gr WC(U U )k /WC(U
WC(U U )k−1 .
k=0

For T = ∑α,β,γ,δ cαβγδ xα ηβ ∂γ ðδ , with cαβγδ ∈ C, of filtered degree j, i.e., T ∈


U ) j \WC(U
WC(U U ) j−1 , we define the symbol or the Weyl symbol of T to be

σ(T ) = ∑ cα,β,γ,δ xα ηβ yγ ξδ ∈ Sk (U
U ),
|α|+|β|+|γ|+|δ|=k
5.1. Weyl-Clifford algebra and classical Lie superalgebras 143

where we denote by yi and ξi the generators in S(U


U ) that correspond to ∂i and ði ,
respectively. We define a symbol map, for k ≥ 0,
σ : WC(U
U )k → ⊕kj=0 S j (U
U)
by sending T to σ(T ).
Proposition 5.4. Let G be a classical subgroup of GL(U0̄ ) × GL(U1̄ ). The symbol
map σ is a G-module isomorphism from WC(U U )k to ⊕kj=0 S j (U
U ) for each k. As a
G-module, the associated graded algebra gr WC(U U ) is isomorphic to the super-
symmetric algebra S(UU ).

Proof. The G-module WC(U U )k is semisimple and hence isomorphic to the associ-
ated graded ⊕ j=0 WC(U
k U ) j /WC(U U ) j−1 . For a fixed j, the linear span of {xα ηβ ∂γ ðδ |
|α| + |β| + |γ| + |δ| = j} as a G-module is isomorphic to WC(U U ) j /WC(UU ) j−1 ,
which in turn is isomorphic to S (Uj
U ). The proposition follows. 

5.1.3. Relation to classical Lie superalgebras. There is an intimate relation be-


tween the Weyl-Clifford algebra and classical Lie algebras/superalgebras, which
we will now formulate. Introduce G = G0̄ ⊕ G1̄ , where
G0̄ = S2 (ι(U
U 0̄ )) ⊕ ∧2 (ι(U
U 1̄ )), G1̄ = ι(U
U 0̄ ) ⊗ ι(U
U 1̄ ).
Proposition 5.5. (1) The subspace G of WC(U U ) is closed under the super-
commutator, and as Lie superalgebras G ∼ = spo(U U ). In particular, G0̄ is a
U 0̄ ) ⊕ so(U
Lie algebra isomorphic to sp(U U 1̄ ).
(2) Under the adjoint action, ι(U
U ) is isomorphic to the natural representation
of G ∼ U ).
= spo(U

Proof. We identify S2 (ι(U U 0̄ )) with the space of anti-commutators [ι(U U 0̄ ), ι(U


U 0̄ )]+ ,
that is, the linear span of [a, b]+ = ab + ba, for a, b ∈ ι(U U 0̄ ); note by (5.2) that
[a, b] = ab − ba ∈ C. Similarly we identify ∧2 (ι(U U 1̄ )) with [ι(UU 1̄ ), ι(U
U 1̄ )], the
linear span of [c, d] = cd − dc, for c, d ∈ ι(UU 1̄ ); note by (5.2) that [c, d]+ ∈ C.
We claim that S2 (ι(UU 0̄ )) is closed under the commutator, and moreover, we
2
have S (ι(U ∼
U 0̄ )) = sp(U
U 0̄ ) as Lie algebras. To see this, let x1 , . . . , xm be a basis for
U0̄ . Regard xi and ∂i as elements in WC(U U ) as before. We may identify S2 (ι(U U 0̄ ))
with the space spanned by elements of the form
(5.3) xi ∂ j + ∂ j xi , xi x j , ∂i ∂ j , 1 ≤ i, j ≤ m.
By a direct computation via this spanning set, one checks that S2 (ι(U U 0̄ )) is closed

under the commutator. Using that [b, a1 ] = ⟨b, a1 ⟩ by (5.2), one further computes
that, for a, b, a1 ∈ ι(UU 0̄ ),
[ ]
(5.4) [a, b]+ , a1 = a[b, a1 ] + [b, a1 ]a + [a, a1 ]b + b[a, a1 ]
= 2⟨b, a1 ⟩′ a + 2⟨a, a1 ⟩′ b.
144 5. Howe duality

Hence, ι(UU 0̄ ) is a representation of the Lie algebra S2 (ι(U U 0̄ )) under the adjoint
action. With respect to the ordered basis {x1 , . . . , xm , ∂1 , . . . , ∂m } for ι(U
U 0̄ ), we
can write down the matrices of (5.3), the span of which is precisely given by the
2m × 2m matrices of the m|m-block form
( )
A B
, B = Bt ,C = Ct .
C −At
U 0̄ )) ∼
This shows that S2 (ι(U U 0̄ ) and that ι(U
= sp(U U 0̄ ) is its natural representation.
Similarly, ∧ (ι(U
2 U )) is closed under the Lie bracket, ∧2 (ι(U
1̄ U )) ∼
1̄ = so(UU ) as

Lie algebras, and ι(U U 1̄ ) under the adjoint action of ∧2 (ι(U U 1̄ )) is its natural rep-
resentation. Since the case for ∧2 (ι(U U 1̄ )) is completely parallel to the case for
S2 (ι(U
U 0̄ )) above, we shall omit the details except recording the following identity
analogous to (5.4):
[ ]
(5.5) [c, d], c1 = 2⟨d, c1 ⟩′ c − 2⟨c, c1 ⟩′ d, ∀c, d, c1 ∈ ι(U U 1̄ ).

On the other hand, by a direct computation, we have that, for a, b ∈ U 0̄ and


c, d ∈ U 1̄ ,
1
[a ⊗ c, b ⊗ d]+ = ([a, b] ⊗ [c, d] + [a, b]+ ⊗ [c, d]+ ).
2
This implies that [G1̄ , G1̄ ]+ ⊆ G0̄ . Hence G is closed under the super-commutator
and so is a Lie subalgebra of WC(U U ).
We further observe that G preserves the bilinear form ⟨·, ·⟩′ on U . Indeed, it
follows by using (5.4) twice that, for a, b, a1 , b1 ∈ ι(U
U 0̄ ),
⟨[ ] ⟩′ ⟨ [ ]⟩′
[a, b]+ , a1 , b1 = − a1 , [a, b]+ , b1 .
Similarly, it follows by (5.5) that, for c, d, c1 , d1 ∈ ι(U
U 1̄ ),
⟨[ ] ⟩′ ⟨ [ ]⟩′
[c, d], c1 , d1 = − c1 , [c, d], d1 .
The remaining identities for the G-invariance of ⟨·, ·⟩′ can be verified similarly.
Since G0̄ = sp(UU 0̄ ) ⊕ so(UU 1̄ ) and G1̄ ∼ U 1̄ , it follows that G ∼
= U 0̄ ⊗U U ). Ad-
= spo(U
ditional direct computations on [[a, b]+ , c] and [[c, d]+ , a] together with (5.4) and
(5.5) show that ι(U U ) is indeed the natural representation of G under the adjoint
action. 
Proposition 5.6. The adjoint action of the Lie superalgebra G on WC(U U ) pre-
U ). The action of G0̄ can be lifted to an action by
serves the filtration on WC(U
automorphisms of the group Sp(UU 0̄ ) × O(U
U 1̄ ) on WC(U
U ).

Proof. The first statement follows from Proposition 5.5(2) and the definition of
the filtration on WC(U U ). Clearly the adjoint action of G0̄ on ι(U U ) lifts to an ac-
tion of the group Sp(U U 0̄ ) × O(U
U 1̄ ). As each finite-dimensional filtered subspace
WC(U U )k for k ≥ 1 is preserved by the adjoint action of G0̄ , the G0̄ -module WC(U U)
5.1. Weyl-Clifford algebra and classical Lie superalgebras 145

is semisimple and hence isomorphic to S(U


U ) by a standard induction argument.
The second statement follows. 

5.1.4. A general duality theorem. In this subsection, we establish a duality the-


orem in a general setting, which will be used subsequently.
Let G be one of the classical Lie groups GL(V ), Sp(V ), or O(V ). Assume
(ρ, L) is a rational representation of G of countable dimension, which means a
direct sum of finite-dimensional simple G-modules. Assume that a subalgebra
R ⊆ End(L) satisfies:
(i) R acts irreducibly on L.
(ii) For g ∈ G, T ∈ R, we have ρ(g)T ρ(g)−1 ∈ R. That is, R is a G-module
by conjugation.
(iii) As a G-module, R is a direct sum of finite-dimensional irreducible G-
modules.
As we shall see, in the main applications developed in this chapter, the algebra R is
taken to be the Weyl-Clifford algebra on a sum of copies of the natural G-module
V , and Conditions (i)-(iii) are easily verified.
Define the algebra of G-invariants in R to be
RG = {T ∈ R | ρ(g)T = T ρ(g), ∀g ∈ G}.
We write the G-module L as a direct sum of finite-dimensional simple G-modules

L(λ) of the isomorphism class λ as L ∼= λ∈G(ρ)
b
b
L(λ) ⊗ M λ , where G(ρ) denotes
the set of isomorphism classes of simple G-modules appearing in (ρ, L), and
M λ := HomG (L(λ), L)
denotes the multiplicity space of λ.
Then M λ is naturally a left RG -module by left multiplication. Indeed, for T ∈
RG , f ∈ M λ = HomG (L(λ), L), v ∈ L(λ), g ∈ G, we have
T f (g.v) = T ρ(g) f (v) = ρ(g)T f (v),
that is, T f ∈ Mλ.Hence, we have a decomposition of the CG ⊗ RG -module L
(which we shall simply refer to as a (G, RG )-module):

(5.6) L∼
= L(λ) ⊗ M λ .
b
λ∈G(ρ)

We shall need the following variant of the Jacobson density theorem, a proof
of which can be found in [GW, Corollary 4.1.6].
Proposition 5.7. Let L be a vector space of countable dimension. Assume that a
subalgebra R of End(L) acts on L irreducibly. Then, for any finite-dimensional
subspace X of L, we have R|X = Hom(X, L).
146 5. Howe duality

Theorem 5.8. Assume that (ρ, L) is a representation of a classical group G of


countable dimension, which is a direct sum of finite-dimensional simple G-modules.
Further assume that G, R, L satisfy Conditions (i)-(iii). Then the M λ ’s, for λ ∈
b
G(ρ), are pairwise non-isomorphic irreducible RG -modules.

Proof. We shall need the following G-invariant version of Proposition 5.7.


Claim. Let X ⊆ L be a finite-dimensional G-invariant subspace. Then RG |X =
HomG (X, L).
Indeed, by G-invariance of X, Hom(X, L) is a G-module, and the restriction
res
map R −→ Hom(X, L) given by r → r|X is a G-homomorphism. Since R and
Hom(X, L) are completely reducible G-modules, they contain RG and HomG (X, L)
as direct summands of trivial submodules; We shall denote by π the respective G-
equivariant projection maps. Hence, we have the following commutative diagram:
π / RG
(5.7) R
res res
 
π / HomG (X, L)
Hom(X, L)

Now, suppose that T ∈ HomG (X, L). Regarding T ∈ Hom(X, L), we may apply
Proposition 5.7 to find an element r ∈ R such that r|X = T . Then we have
T = π(T ) = π(r|X ) = π(r)|X ,
proving the claim.
Recall M λ = HomG (L(λ), L) by definition, for λ ∈ G(ρ).b Take two nonzero
−1
elements S, T ∈ HomG (L(λ), L). Then ψ = ST : T (L(λ)) → S(L(λ)) is an iso-
morphism of G-modules. It follows by the claim above that there exists u ∈ RG
which restricts to ψ : T (L(λ)) → L. Hence, uT : L(λ) → S(L(λ)) is an isomor-
phism of irreducible G-modules. By Schur’s Lemma, we must have S = cuT for
some scalar c, and this proves the irreducibility of the RG -module HomG (L(λ), L).
b
It remains to show that, for distinct λ, µ ∈ G(ρ), an RG -module homomorphism
φ : HomG (L(λ), L) → HomG (L(µ), L) must be zero. To that end, we take an arbi-
trary element T ∈ HomG (L(λ), L). Set S := φ(T ), U := T (L(λ)) ⊕ S(L(µ)), and
let p : U → S(L(µ)) ⊆ L be the natural projection, which is clearly G-equivariant.
By the above claim, there exists r ∈ RG such that p = r|U . Then, it follows from
pT = 0 that rT = 0 and then,
0 = φ(rT ) = rφ(T ) = rS = pS = S.
Hence we conclude that φ = 0 as expected. 

We will say the decomposition of the (G, RG )-module L as in (5.6) is strongly


multiplicity-free, in light of Theorem 5.8.
5.1. Weyl-Clifford algebra and classical Lie superalgebras 147

5.1.5. A duality for Weyl-Clifford algebras. Let U = U0̄ ⊕U1̄ be a vector super-
space, and let G be a classical Lie subgroup of GL(U0̄ ) × GL(U1̄ ). In all the exam-
ples which give rise to concrete Howe dualities in the next sections, U0̄ and U1̄ are
actually direct sums of the natural G-module. The induced G-module (ρ, S(U)) is
semisimple. Also, the G-module WC(U U ) ⊆ End(S(U)) by conjugation given by
−1
g.T = ρ(g)T ρ(g ), for g ∈ G and T ∈ WC(U U ), is semisimple.
U ), S(U) is irre-
Lemma 5.9. As a module over the Weyl-Clifford algebra WC(U
ducible.

Proof. Identify S(U) with C[x, η], and identify WC(U U ) as the superalgebra of
polynomial differential operators on U as in Section 5.1.2. Then, given any nonzero
element f in C[x, η], we can find a suitable constant-coefficient differential operator
D ∈ WC(U U ) such that D. f = 1. Now applying suitable multiplication operators in
WC(U U ) to 1, we obtain all the monomials xα ηβ which span C[x, η]. The lemma
follows. 

Then letting L = S(U) and R = WC(U U ), the triple (G, L, R) satisfies Condi-
tions (i)–(iii) in 5.1.4. Hence Theorem 5.8 is applicable in light of Lemma 5.9 and
gives us the following.
U )G )-module, S(U) is
Theorem 5.10. Retain the notations above. As a (G, WC(U
strongly multiplicity-free, i.e.,

S(U) ∼
= L(λ) ⊗ M λ .
b
λ∈G(ρ)

Here the M λ ’s are pairwise non-isomorphic irreducible WC(U U )G -modules, while


the L(λ)’s are pairwise non-isomorphic irreducible finite-dimensional G-modules.

Recall by Theorem 4.1 that the algebra S(U


U )G is finitely generated.
Proposition 5.11. Let {S1 , . . . , Sr } be a set of generators of the algebra S(U U )G ,
where G is a classical subgroup of GL(U0̄ ) × GL(U1̄ ). Suppose T j ∈ WC(U U )G is
chosen so that σ(T j ) = S j , for j = 1, . . . , r. Then {T1 , . . . , Tr } generate the algebra
WC(U U )G .

Proof. It follows from Proposition 5.4 that the symbol map σ restricts to an iso-
k to ⊕ j=0 S (U
U )G , for each k ≥ 0.
j
morphism from WC(U U )G k

Let a ∈ WC(U U )G
k . We shall show that a is generated by T1 , . . . , Tr by in-
duction on k. For k = 0 the claim is trivial. Suppose that k ≥ 1. By assump-
tion σ(a) = f (S1 , S2 , . . . , Sr ), for some polynomial f . Now consider the element
f (T1 , T2 , . . . , Tr ) ∈ WC(U k . We note that σ(a − f (T1 , . . . , Tr )) ∈ ⊕ j=0 S (U
U )G k−1 j
U ), and
hence a − f (T1 , . . . , Tr ) ∈ WC(U U )G
k−1 . By induction hyperthesis, a − f (T1 , . . . , Tr )
is generated by T1 , . . . , Tr , and hence so is a. 
148 5. Howe duality

U )G
Proposition 5.12. Suppose that the generators T1 , . . . , Tr for the algebra WC(U

can be chosen such that g := ⊕i=1 CTi is a Lie subalgebra of WC(U
r G
U ) under
the super commutator. Then the irreducible WC(U G λ
U ) -modules M ’s (see The-
orem 5.10) are pairwise non-isomorphic g′ -modules, and we have the following
(G, g′ )-module decomposition:

S(U) ∼
= L(λ) ⊗ M λ .
b
λ∈G(ρ)

Proof. By assumption, the action of g′ on S(U) extends to a representation of


the universal enveloping superalgebra U(g′ ), so we obtain an algebra homomor-
phism ρ′ : U(g′ ) → End(S(U)). The assumption that T1 , . . . , Tr generate the algebra
WC(U U )G implies that ρ′ (U(g′ )) = WC(U
U )G , and the proposition follows. 

We shall refer to (G, g′ ) in Proposition 5.12 as a Howe dual pair. In the


next sections, by specializing to the case when G is a classical group and U0̄ and
U1̄ are direct sums of the natural G-module, we shall be able to make an explicit
choice of the generators T1 , . . . , Tr so that Proposition 5.12 can be applied. Note
that S(U) = S(U0̄ ) when U1̄ = 0, and that S(U) = ∧(U1̄ ) when U0̄ = 0. In either
special case, the g′ appearing in Proposition 5.12 will be a Lie algebra.

5.2. Howe duality for type A and type Q


In this section, we first formulate Howe duality for the classical goup GL(V ) and
give a precise multiplicity-free decomposition as predicted by Proposition 5.12. We
find explicit formulas for the joint highest weight vectors in the above decompo-
sition, which will also be useful in Section 5.3. In addition, we formulate another
Howe duality between a pair of queer Lie superalgebras and provide an explicit
multiplicity-free decomposition.

5.2.1. Howe dual pair (GL(k), gl(m|n)). Take V = Ck , and hence identify GL(V )
with GL(k) and gl(V ) with gl(k). Given m, n ≥ 0, we consider the superspace
U = U0̄ ⊕ U1̄ with U0̄ = V m and U1̄ = V n , and identify naturally U ≡ V ⊗ Cm|n .
We let e1 , . . . , ek denote the standard basis for the natural gl(k)-module Ck , and let
ei , i ∈ I(m|n), denote the standard basis for the natural module Cm|n of the general
linear Lie superalgebra gl(m|n). We set, for 1 ≤ r ≤ k, 1 ≤ a ≤ m, 1 ≤ b ≤ n,
(5.8) xra := er ⊗ eā , ηrb := er ⊗ eb .
Then the set {xra , ηrb } is a basis for U. We will denote by C[x, η] the polynomial
superalgebra generated by the elements in (5.8), and from now on further identify

(5.9) S(U) ≡ S(V ⊗ Cm|n ) ≡ C[x, η].


5.2. Howe duality for type A and type Q 149

Introduce the following first order differential operators


m
∂ n

(5.10) E pq = ∑ xp j + ∑ η ps
∂xq j s=1 ∂ηqs
, 1 ≤ p, q ≤ k.
j=1
Also introduce the following first order differential operators
k
∂ k

Eī j¯ = ∑ x pi ∂x p j
, Eīs = ∑ x pi
∂η ps
,
p=1 p=1
(5.11)
k
∂ k

Est = ∑ η ps ∂η pt , Esī = ∑ η ps ∂x pi , 1 ≤ i, j ≤ m, 1 ≤ s,t ≤ n.
p=1 p=1

The natural commuting actions of gl(k) and gl(m|n) on V ⊗ Cm|n induce com-
muting actions on C[x, η]. The following lemma is standard and can be verified by
a direct computation.
Lemma 5.13. The commuting actions of gl(k) and gl(m|n) on S(Ck ⊗ Cm|n ) =
C[x, η] are realized by the formulas (5.10) and (5.11), respectively.

We shall denote by g′ the linear span of the elements in (5.11). Indeed, the
elements in (5.11) form a linear basis for g′ . Let U = U ⊕U ∗ as before.
Theorem 5.14. The g′ , which forms an isomorphic copy of gl(m|n), generates the
algebra of GL(V )-invariants in WC(U U ). Moreover, as a (GL(k), gl(m|n))-module,
S(Ck ⊗ Cm|n ) is strongly multiplicity-free.

Proof. Set G = GL(k). Recall from Proposition 5.4 that the associated graded of
the filtered algebra WC(UU ) is S(UU ), and that the action of gl(k) on S(U
U ) lifts to
an action of G. In light of (5.9), we identify WC(U U ) with the superalgebra of
polynomial differential operators on C[x, η], and so the elements in (5.11) may be
regarded as elements in WC(U U ).
Let {y pi } denote the basis in U0̄∗ dual to the basis {x pi } for U0̄ , and let {ξ pt }
denote the basis in U1̄∗ dual to the basis {η pt } for U1̄ . By the supersymmetric
First Fundamental Theorem for GL(V ) (see Theorem 4.19), the algebra S(U U )G =
( )G
S(U0̄ ⊕U0̄∗ ) ⊗ ∧(U1̄ ⊕U1̄∗ ) is generated by
k k
zī j¯ = ∑ x pi y p j , zts = ∑ η pt ξ ps ,
p=1 p=1
(5.12)
k k
zīt = ∑ x pi ξ pt , zt ī = ∑ η pt y pi , 1 ≤ i, j ≤ m, 1 ≤ s,t ≤ n.
p=1 p=1

Observe that the symbol map σ in 5.1.2 sends each Eab in (5.11) to zab for a, b ∈
I(m|n), and hence by Proposition 5.11, the Eab ’s form a set of generators for
WC(U U )G . Now the theorem follows from Proposition 5.12, since the Eab ’s gener-
ate the Lie superalgebra gl(m|n) by Lemma 5.13. 
150 5. Howe duality

Remark 5.15. When n = 0, S(U) = S(U0̄ ) becomes the symmetric algebra on


U0̄ = Ck ⊗Cm . When m = 0, S(U) = ∧(U1̄ ) is the exterior algebra on U1̄ = Ck ⊗Cn .
In these two important specializations of Theorem 5.14, g′ becomes a Lie algebra,
namely gl(m) and gl(n), respectively.

5.2.2. (GL(k), gl(m|n))-Howe duality. We will first find the multiplicity-free de-
composition of the (GL(V ), gl(m))-module S(Ck ⊗ Cm ) explicitly. To do that, we
will take advantage of the Schur duality which was established in Chapter 3. As
the formulation below will involve Lie algebra gl(k) for various k, we will add the
index k to denote by Lk (λ) the irreducible gl(k)-module of highest weight λ. For
dominant integral weight λ, the gl(k)-module Lk (λ) lifts to a GL(k)-module, which
will be denoted by the same notation.
Let us consider the natural action of gl(k) × gl(m) on Ck ⊗ Cm and its induced
action on the dth symmetric tensor Sd (Ck ⊗ Cm ), for d ≥ 0. As usual, we identify a
polynomial weight for gl(k) with a partition of length at most k. Recall Pd denotes
the set of partitions of d.
Theorem 5.16 ((GL(k), gl(m))-Howe duality). As a (GL(k), gl(m))-module, we
have ⊕
Sd (Ck ⊗ Cm ) ∼
= Lk (λ) ⊗ Lm (λ).
λ∈Pd
ℓ(λ)≤min{k,m}

Proof. We shall show that Schur duality implies Howe duality. By Schur duality
(see Theorem 3.10), we have
(5.13) ∼ ⊕ λ∈P Lk (λ) ⊗ Sλ ,
(Ck )⊗d = ∼ ⊕ µ∈P Lm (µ) ⊗ Sµ .
(Cm )⊗d =
d d
ℓ(λ)≤k ℓ(µ)≤m

Using (5.13) and denoting by ∆Sd the diagonal subgroup in Sd × Sd , we have


( )Sd
Sd (Ck ⊗ Cm ) ∼
= (Ck ⊗ Cm )⊗d
( )∆Sd

= (Ck )⊗d ⊗ (Cm )⊗d
⊕ ( )∆Sd

= Lk (λ) ⊗ Lm (µ) ⊗ Sλ ⊗ Sµ
λ,µ∈Pd
ℓ(λ)≤k,ℓ(µ)≤m


= Lk (λ) ⊗ Lm (λ).
λ∈Pd
ℓ(λ)≤min{k,m}

In the last step, we have used the well-known fact that Sλ ∼ = (Sλ )∗ , and hence by
Schur’s lemma
(Sλ ⊗ Sµ )∆Sd ∼
= HomSd (Sλ , Sµ ) ∼
= δλ,µ C.
This completes the proof of the theorem. 
5.2. Howe duality for type A and type Q 151

Remark 5.17. It is possible to derive Schur duality from (GL(k), gl(m))-Howe


duality, by making use of the well-known fact that the zero-weight subspace of
the GL(d)-module Ld (λ) for a partition λ of d is naturally an Sd -module, which
is isomorphic to the Specht module Sλ . Also, it is possible to derive the FFT for
GL(k) from Howe duality. Hence the FFT for the general linear group, Schur
duality, and Howe duality of type A are essentially all equivalent.

Now let us consider the induced action of gl(k) × gl(m) on the dth exterior
tensor space ∧d (Ck ⊗ Cm ), for 0 ≤ d ≤ km. Recall that λ′ denotes the conjugate
partition of a partition λ. We have the following skew version of Howe duality.

Theorem 5.18 (Skew (GL(k), gl(m))-Howe duality). As a (GL(k), gl(m))-module,


we have

∧d (Ck ⊗ Cm ) ∼
= Lk (λ) ⊗ Lm (λ′ ).
λ∈Pd
ℓ(λ)≤k,ℓ(λ′ )≤m

Proof. We shall use Schur duality to derive the skew Howe duality. For an Sd -
module M, let us denote by M Sd ,sgn the submodule of M that transforms by the
sign character. Using (5.13), we have
( )Sd ,sgn
∧d (Ck ⊗ Cm ) ∼
= (Ck ⊗ Cm )⊗d
( )∆Sd ,sgn

= (C k ⊗d
) ⊗ (C m ⊗d
)
⊕ ( )∆Sd ,sgn

= Lk (λ) ⊗ Lm (µ) ⊗ Sλ ⊗ Sµ
λ,µ∈Pd
ℓ(λ)≤k,ℓ(µ)≤m


= Lk (λ) ⊗ Lm (λ′ ).
λ∈Pd
ℓ(λ)≤k,ℓ(λ′ )≤m

In the last step, we have used the isomorphism Sµ ⊗ sgn ∼



= Sµ from (A.27), and
hence
( )∆Sd
(Sλ ⊗ Sµ )∆Sd ,sgn ∼
= Sλ ⊗ Sµ ⊗ sgn ∼
= HomSd (Sλ , Sµ ⊗ sgn) ∼
= δλ,µ′ C.

The natural action of gl(k) × gl(m|n) on Ck ⊗ Cm|n induces a natural action


on the supersymmetric algebra S(Ck ⊗ Cm|n ). Recall from (2.4) the weight λ♮ as-
sociated to a (m|n)-hook partition λ. The following theorem, which is a common
generalization of Theorems 5.16 and 5.18, provides an explicit form of the strongly
multiplicity-free decomposition in Theorem 5.14.
152 5. Howe duality

Theorem 5.19 ((GL(k), gl(m|n))-Howe duality). As a (GL(k), gl(m|n))-module,


we have ⊕
Sd (Ck ⊗ Cm|n ) ∼
= Lk (λ) ⊗ Lm|n (λ♮ ).
λ∈Pd
ℓ(λ)≤k,λm+1 ≤n

We shall omit the proof of Theorem 5.19, which is completely analogous to


the one for Theorem 5.16, now using the Schur-Sergeev duality (Theorem 3.10)
which we recall: as (gl(m|n), Sd )-module, (Cm|n )⊗d ∼
= ⊕λ∈Pd ,λm+1 ≤n L(λ♮ ) ⊗ Sλ .
Recall the character of the irreducible gl(k)-module Lk (λ) is given by the Schur
function sλ , while the character of the irreducible gl(m|n)-module Lm|n (λ♮ ) is given
the super Schur function hsλ (x; y); see Theorem 3.14. Now by taking the characters
of both sides of the isomorphism in Theorem 5.19 and summing over all d ≥ 0, we
immediately recover the super Cauchy identity (A.39) in Appendix A:
∏ j,k (1 + y j zk )
= ∑ hsλ (x; y)sλ (z).
∏i,k (1 − xi zk ) λ∈P
By setting the variables y j (respectively, xi ) to zero and noting that hsλ (x; 0) = sλ (x)
and hsλ (0; y) = sλ′ (y), we further recover the classical Cauchy identities as given in
(A.11), which correspond to the Howe duality in Theorem 5.16 and the skew-Howe
duality in Theorem 5.18, respectively.
Remark 5.20. When we derive the (GL(k), gl(m|n))-Howe duality from the Schur-
Sergeev duality, we could have replaced GL(k) by the Lie algebra gl(k) without any
change. Moreover, using Schur-Sergeev duality in the same fashion, one can derive
a more general (gl(k|ℓ), gl(m|n))-Howe duality on S(Ck|ℓ ⊗ Cm|n ).

5.2.3. Formulas for highest weight vectors. In this subsection, we shall find ex-
plicit formulas for the joint highest weight vectors in the (GL(k), gl(m|n))-Howe
duality decomposition (see Theorem 5.19). These formulas will play a key role in
the subsequent section on Howe duality for symplectic and orthogonal groups.
Recalling the notation from (5.10) and (5.11) the simple root vectors of gl(k)
and gl(m|n) are respectively (1 ≤ i ≤ k − 1, 1 ≤ s ≤ m − 1, 1 ≤ t ≤ n − 1):
m
∂ n

(5.14) Ei,i+1 = ∑ xi j ∂xi+1, j + ∑ ηis ∂ηi+1,s ,
j=1 s=1

(5.15)
k
∂ k
∂ k

Es̄,s+1 = ∑ x ps ∂x p,s+1 , Et,t+1 = ∑ η pt ∂η p,t+1 , Em,1 = ∑ x pm ∂η p1 .
p=1 p=1 p=1

Let λ be an (m|n)-hook partition of length at most k. We are looking for


the joint highest weight vector with respect to the standard Borel subalgebra of
gl(k) × gl(m|n), or equivalently the vector annihilated by (5.14) and (5.15), in
5.2. Howe duality for type A and type Q 153

the summand Lk (λ) ⊗ Lm|n (λ♮ ) in the decomposition of C[x, η]. Such a vector is
unique up to a scalar multiple, thanks to the multiplicity-free decomposition in
Theorem 5.19.
For 1 ≤ r ≤ m, define
 
x11 x12 · · · x1r
x21 x22 · · · x2r 
 
(5.16) ♢r := det  . .. .. ..  .
 .. . . . 
xr1 xr2 · · · xrr
The column determinant of an r × r matrix with possibly non-commuting
entries A = [ai j ] is defined to be
cdet A = ∑ (−1)ℓ(σ) aσ(1)1 aσ(2)2 · · · aσ(r)r .
σ∈Sr

In case m < k, we introduce the following determinant


 
x11 x12 · · · x1m η1t · · · η1t
x21 x22 · · · x2m η2t · · · η2t 
 
(5.17) ♢t,r := cdet  . .. . . .. .. .. ..  ,
 .. . . . . . . 
xr1 xr2 · · · xrm ηrt · · · ηrt
for m < r ≤ k and 1 ≤ t ≤ n, where the last (r − m) columns are filled with the same
vector. For notational convenience, we set ♢t,r := ♢r , for all t and 1 ≤ r < m.
Remark 5.21. The element ♢t,r is always nonzero. It reduces to (5.16) when
r ≤ m, and, up to a scalar multiple, reduces to η1t · · · ηrt when m = 0.

For a partition µ with µ1 ≤ k and 1 ≤ t1 ,t2 , . . . ,tµ1 ≤ n, we define


(5.18) ♢t1 ,t2 ,...,tµ1 ,µ′ := ♢t1 ,µ′1 ♢t2 ,µ′2 · · · ♢tµ1 ,µ′µ .
1

Now let r be fixed by the conditions > m and λ′r λ′r+1


≤ m for an (m|n)-hook
partition λ. Note that we must have 0 ≤ r ≤ n. Denote by λ≤r the subdiagram of
the Young diagram λ which consists of the first r columns of λ, i.e., the columns
of length greater than m.
Lemma 5.22. The vector ♢1,2,...,r,λ′≤r is a joint highest weight vector for the gl(k) ×
gl(m|n)-module Lk (λ≤r ) ⊗ Lm|n (λ♮≤r ) in the decomposition of C[x, η].

Proof. Set µ = λ≤r and ♢ = ♢1,2,...,r,µ′ in this proof. By Remark 5.21 and a direct
computation using (5.10), (5.11) and (5.14), we can verify the following.
(1) ♢ is non-zero.
(2) ♢ has weight (µ, µ♮ ) with respect to the action of gl(k) × gl(m|n).
(3) Each factor ♢i,λ′i in ♢ is annihilated by the operators in (5.14).
154 5. Howe duality

By the strongly multiplicity-free decomposition of Theorem 5.19 (recalling


S(Ck ⊗ Cm|n ) ≡ C[x, η]), we may identify Lm|n (µ♮ ) ≡ C[x, η]nµ1 , the space of highest
weight vectors in the gl(k)-module C[x, η] of highest weight µ (here n1 denotes the
nilradical of the standard Borel for gl(k)). It follows by (3) that ♢ is annihilated
by the operators in (5.14), and hence ♢ ∈ C[x, η]nµ1 by (1)-(3). Since ♢ has weight
µ♮ with respect to the action of gl(m|n) by (2) again, it must be the unique (up to
a scalar multiple) vector in Lm|n (µ♮ ) ≡ C[x, η]nµ1 of weight µ♮ . Hence it is a joint
highest weight vector in C[x, η] of weight (µ, µ♮ ). 

Theorem 5.23. Let λ be a partition such that ℓ(λ) ≤ k and λm+1 ≤ n. Then, a
highest weight vector of weight (λ, λ♮ ) in the gl(k) × gl(m|n)-module C[x, η] is
given by ♢1,2,...,λ1 ,λ′ = ♢1,λ′1 ♢2,λ′2 · · · ♢λ1 ,λ′λ .
1

Proof. The vector ♢1,2,...,λ1 ,λ′ is nonzero since each factor ♢i,λ′i is nonzero by Re-
mark 5.21 and distinct ηi j are involved in different ♢i,λ′i . Each factor ♢i,λ′i in
♢1,2,...,λ1 ,λ′ is a joint highest weight vector by a special case of Lemma 5.22, and
hence so is ♢1,2,...,λ1 ,λ′ . Also it is straightforward to verify by (5.10) and (5.11) that
♢1,2,...,λ1 ,λ′ has the correct weights (λ, λ♮ ). 

Remark 5.24. When n = 0, the highest weight vector in Theorem 5.23 does not in-
volve the odd generators ηi j . When m = 0, the vector becomes simply a monomial
in the anti-commuting variables ηi j .

5.2.4. (q(m), q(n))-Howe duality. In this subsection, using the (q(n), Hd )-Sergeev
duality in Theorem 3.46, we shall obtain a (q(m), q(n))-Howe duality in a similar
fashion as in 5.2.2.
Recall that by definition q(m) consists of X ∈ gl(m|m) super-commuting ( with )
A B
P given in (1.10), i.e., q(m) consists of the m|m-block matrices of the form .
B A
Recall from Remark 1.9 that we have another nonstandard realization of the Lie su-
peralgebra q(m) as e q(m) = {X ∈(gl(m|m))| XP − PX = 0}, which consists of the
A B
m|m-block matrices of the form . We let q(m) act on Cm|m as the non-
−B A
standard isomorphic copy e q(m), while we let q(n) act on Cn|n by linear maps in the
standard way, i.e. linear maps super-commuting with P. This induces an action of
q(m) × q(n) on Cm|m ⊗ Cn|n .
Define the linear map P∆ : Cm|m ⊗ Cn|n → Cm|m ⊗ Cn|n by

P∆ (v ⊗ w) := Pv ⊗ Pw, ∀v ∈ Cm|m , w ∈ Cn|n .

Lemma 5.25. The actions of P∆ and q(m) × q(n) on Cm|m ⊗ Cn|n commute with
each other.
5.2. Howe duality for type A and type Q 155

Proof. For Z2 -homogeneous X ∈ q(m), Y ∈ q(n), v ∈ Cm|m and w ∈ Cn|n we have


XP∆ (v ⊗ w) = X(Pv ⊗ Pw) = (XPv) ⊗ (Pw)
= PXv ⊗ Pw = (P ⊗ P)(Xv ⊗ w) = P∆ X(v ⊗ w),
Y P∆ (v ⊗ w) = Y (Pv ⊗ Pw) = (−1)|Y |+|Y ||v| Pv ⊗Y Pw
= (−1)|Y ||v| Pv ⊗ PY w = (−1)|Y ||v| (P ⊗ P)(v ⊗Y w)
= P∆Y (v ⊗ w).
The lemma follows. 

Recall that the finite group Πd is generated by a1 , . . . , ad with relations (3.17),


and this gives rise to the group Bed = Πd o Sd . The natural action of the group
Πd (respectively, Bed ) on (Cn|n )⊗d factors through an action of the Clifford algebra
Cd by Lemma 3.44 (respectively, the algebra Hd ), where ai acts as 1i−1 ⊗ P ⊗ 1d−i
with P given in (1.10). Hence the diagonal subgroups ∆Πd ⊂ Πd × Πd and ∆Bed ⊂
Bed × Bed act on the tensor product
(Cm|m )⊗d ⊗ (Cn|n )⊗d ∼
= (Cm|m ⊗ Cn|n )⊗d .
In this subsection, it is more convenient to talk about the groups Πd and Bed in place
of the algebras Cd and Hd .
Lemma 5.26. We have a natural identification:
( )∆(Bed )
(Cm|m ⊗ Cn|n )⊗d ∼
= Sd (Cmn|mn ),
e
where (·)∆(Bd ) denotes the space of ∆(Bed )-invariants.

Proof. Consider first the case when d = 1. We observe that (Cm|m ⊗ Cn|n )P con-
sists of elements of the form vm ⊗ vn + P(vm ) ⊗ P(vn ) and vm ⊗ P(vn ) + P(vm ) ⊗ vn ,

where vm ∈ Cm|0 and vn ∈ Cn|0 . Hence (Cm|m ⊗ Cn|n )P is isomorphic to Cmn|mn .
Since ∆Πd is an even subgroup of ∆(Bed ) generated by the d copies of P∆ ’s, we
have
( )∆(Bed ) (( )∆Πd )∆Sd
(Cm|m )⊗d ⊗ (Cn|n )⊗d ∼
= (Cm|m )⊗d ⊗ (Cn|n )⊗d
( )∆Sd
∼ ∆
= ((Cm|m ⊗ Cn|n )P )⊗d

= ((Cmn|mn )⊗d )Sd ∼
= Sd (Cmn|mn ).
This proves the lemma. 

By the Sergeev duality in Theorem 3.46, there is a commuting action of q(n)


and Bed on (Cn|n )⊗d . Hence there is a natural q(m) × q(n)-action on ((Cm|m ⊗
e
Cn|n )⊗d )∆(Bd ) , and by the identification in Lemma 5.26, in turn on Sd (Cmn|mn ).
156 5. Howe duality

We shall write down this action on Sd (Cmn|mn ) explicitly in terms of differential


operators.
Let x pi and ξ pi , for 1 ≤ p ≤ m and 1 ≤ i ≤ n, denote the standard coordinates
of Cmn|mn . We may then identify S(Cmn|mn ) with the polynomial superalgebra gen-
erated by the x pi and ξ pi , for 1 ≤ p ≤ m and 1 ≤ i ≤ n. Introduce the following first
order differential operators:

e pq =
n
∂ ∂
E ∑ (x p j ∂xq j + ξ p j ∂ξq j ),
j=1
(5.19) n
∂ ∂
∑ (x p j ∂ξq j − ξ p j ∂xq j ),
pq
E = 1 ≤ p, q ≤ m.
j=1

ei j =
m
∂ ∂
E ∑ (x pi ∂x p j + ξ pi ∂ξ p j ),
p=1
(5.20) m
∂ ∂
Ei j = ∑ (x pi ∂ξ p j + ξ pi ∂x p j ), 1 ≤ i, j ≤ n.
p=1

The following lemma is proved by a direct computation.

Lemma 5.27. The operators E e pq and E pq in (5.19), for 1 ≤ p, q ≤ m, form a copy of


q(m), while Eei j and Ei j in (5.20), for 1 ≤ i, j ≤ n, form a copy of q(n). Furthermore,
they define commuting actions of q(m) and q(n) on S(Cmn|mn ).

Recall from (1.35) that, for a partition λ of d with length ℓ(λ),


{
0, if ℓ(λ) is even,
δ(λ) =
1, if ℓ(λ) is odd.
Furthermore, recall that SPd denotes the set of strict partitions of d, and Ln (λ)
stands for the simple q(n)-module with highest weight λ, for λ ∈ SPd of length not
exceeding n (see Section 2.1.6).

Theorem 5.28 ((q(m), q(n))-Howe duality). As a (q(m), q(n))-module Sd (Cmn|mn )


admits the following strongly multiplicity-free decomposition:

Sd (Cmn|mn ) ∼
= 2−δ(λ) Lm (λ) ⊗ Ln (λ).
λ∈SPd
ℓ(λ)≤min{m,n}

Proof. By the Sergeev duality (Theorem 3.46), we have as a q(m) × Bed module

(Cm|m )⊗d ∼
= 2−δ(λ) Lm (λ) ⊗ Dλ .
λ∈SPd
ℓ(λ)≤m
5.3. Howe duality for symplectic and orthogonal groups 157

where we recall that Dλ denotes an irreducible Bed -module (or equivalently, an ir-
reducible Hd -module). Therefore, combined with Lemma 5.26, this gives us
e
(5.21) Sd (Cmn|mn ) ∼
= (((Cm|m )⊗d ) ⊗ ((Cn|n )⊗d ))∆(Bd )
⊕ e

= 2−δ(λ) 2−δ(µ) (Lm (λ) ⊗ Dλ ⊗ Ln (µ) ⊗ Dµ )∆(Bd )
λ,µ∈SPd
ℓ(λ)≤m,ℓ(µ)≤n
⊕ e

= 2−δ(λ) 2−δ(µ) (Lm (λ) ⊗ Ln (µ)) ⊗ (Dλ ⊗ Dµ )∆(Bd ) .
λ,µ∈SPd
ℓ(λ)≤m,ℓ(µ)≤n

The Bed -module Dλ is self-dual, i.e., (Dλ )∗ ∼


= Dλ , since by Theorem 3.43 and (3.23)-
(3.25) all the character values of D are real. Recalling Dλ is of type M if and only
λ

if δ(λ) = 0, we conclude by super Schur’s Lemma 3.4 that


e
dim(Dλ ⊗ Dµ )∆(Bd ) = dim HomBed (Dλ , Dµ ) = 2δ(λ) δλµ .
This together with (5.21) completes the proof of the theorem. 

Recall that the character of the irreducible q(n)-module Ln (λ) is given by the
Schur Q-function Qλ up to a 2-power; see Theorem 3.48. Now by taking the char-
acters of both sides of the isomorphism in Theorem 5.28 and summing over all
d ≥ 0, we recover the following Cauchy identity (A.54) in Appendix A:
1 + xi y j
∏ 1 − xi y j = ∑ 2−ℓ(λ) Qλ (x)Qλ (y).
i, j λ∈SP

5.3. Howe duality for symplectic and orthogonal groups


In this section, we use the First Fundamental Theorem (FFT) for classical Lie
groups in Chapter 4 to construct the Howe dual pairs (Sp(V ), osp(2m|2n)) and
(O(V ), spo(2m|2n)). We also obtain precise multiplicity-free decompositions of
these Howe dualities for symplectic and orthogonal groups. The case of orthogo-
nal groups, which are non-connected, requires some extra work.

5.3.1. Howe dual pair (Sp(V ), osp(2m|2n)). We take V = Ck , where k = 2ℓ is


assumed to be even in this subsection. We consider the superspace U = U0̄ ⊕ U1̄
with U0̄ = V m and U1̄ = V n , and identify naturally U ≡ V ⊗ Cm|n as before. We
continue to identify the supersymmetric algebra of U with C[x, η] as in (5.9).
Recall the skew-symmetric matrix (4.7). We let sp(k) = g(V, ω− ) be the sym-
plectic Lie subalgebra of gl(k), and let Sp(k) = Sp(V, ω− ) be the symplectic sub-
group of GL(k), that preserve the non-degenerate skew-symmetric bilinear form
ω− on V = Ck corresponding to the matrix (4.7). In order to write down explicitly
the Lie algebra sp(k), we introduce the following notation. For an ℓ × ℓ matrix
158 5. Howe duality

A = (ai j ), we denote by A♭ = Jℓ At Jℓ , the transpose of A with respect to the “oppo-


site” diagonal:

 
aℓ,ℓ aℓ−1,ℓ ··· a2,ℓ a1,ℓ
aℓ,ℓ−1 aℓ−1,ℓ−1 ··· a2,ℓ−1 a1,ℓ−1 
 
aℓ,ℓ−2 aℓ−1,ℓ−2 ··· a2,ℓ−2 a1,ℓ−2 
 
(5.22) A♭ =  . .. .. .. ..  .
 .. . . . . 
 
 aℓ,2 aℓ−1,2 ··· a2,2 a1,2 
aℓ,1 aℓ−1,1 ··· a2,1 a1,1

It is then straightforward to check that sp(k) consists precisely of the following


k × k matrices in ℓ|ℓ-block form:

{( ) }
A B
(5.23) sp(k) = | D = −A , B = B,C = C .
♭ ♭ ♭
C D

The standard Borel of sp(k) consists of the k × k upper triangular matrices of the
form (5.23) and so is compatible with the standard Borel of gl(k). Furthermore,
the Cartan subalgebra of sp(k) is spanned by the basis vectors Eii − Ek+1−i,k+1−i ,
for 1 ≤ i ≤ k/2.
The action of the Lie algebra sp(k) as the subalgebra of gl(k) on C[x, η] given
in (5.10) lifts to an action of the Lie group Sp(V ) = Sp(k). On the other hand,
the action ρ of gl(m|n) on C[x, η] with ρ(Eab ) = Eab , for a, b ∈ I(m|n), is modified
from (5.11) by a shift of scalars on the diagonal matrices:

k
∂ k k

Eī j¯ = ∑ x pi ∂x p j + 2 δi j , Eīs = ∑ x pi ∂η ps , 1 ≤ i, j ≤ m,
p=1 p=1
(5.24)
k
∂ k k

Est = ∑ η ps − δst ,
∂η pt 2
Esī = ∑ η ps ∂x pi , 1 ≤ s,t ≤ n.
p=1 p=1
5.3. Howe duality for symplectic and orthogonal groups 159

Introduce the following additional operators:


k
2 ( )
Iī j¯ = ∑ x pi xk+1−p, j − xk+1−p,i x p j ,
p=1
k
2 ( )
Iīs = ∑ x pi ηk+1−p,s − xk+1−p,i η ps ,
p=1

( k
2 )
Ist = ∑ η ps ηk+1−p,t − ηk+1−p,s η pt ,
p=1
(5.25)
( ∂
k
2
∂ ∂ ∂ )
∆ī j¯ = ∑ − ,
p=1 ∂x pi ∂xk+1−p, j ∂xk+1−p,i ∂x p j
k
2 ( ∂ ∂ ∂ ∂ )
∆īs = ∑ − , 1 ≤ i < j ≤ m,
p=1 ∂x pi ∂ηk+1−p,s ∂xk+1−p,i ∂η ps
k
2 ( ∂ ∂ ∂ ∂ )
∆st = ∑ − , 1 ≤ s ≤ t ≤ n.
p=1 ∂η ps ∂ηk+1−p,t ∂ηk+1−p,s ∂η pt
Let us denote by g′ the linear span of the elements in (5.24) and (5.25). Clearly,
the elements in (5.24) and (5.25) form a linear basis for g′ .
Lemma 5.29. (1) Under the super-commutator g′ forms a Lie superalgebra,
which is isomorphic to osp(2m|2n).
(2) The actions of Sp(k) and osp(2m|2n) on S(Ck ⊗ Cm|n ) commute.

Proof. (1) With U = Ck ⊗ Cm|n and U = U ⊕ U ∗ , we identify WC(U U ) with the


superalgebra of differential operators on S(U). The operators Eī j¯ and Est from
(5.24) can be written as
1 k ( ∂ ∂ ) 1 k ( ∂ ∂ )
Eī j¯ = ∑ x pi + x pi , Est = ∑ η ps − η ps ,
2 p=1 ∂x p j ∂x p j 2 p=1 ∂η pt ∂η pt
so that g′ can be regarded as a subspace inside G ∼
= spo(UU ) as in Proposition 5.5.
A direct computation using elements in (5.24) and (5.25) shows that g′ forms a Lie
superalgebra under the super-commutator. Moreover, a further direct computation
shows that g′ ∼
= osp(2m|2n), with identification of root vectors eα , for a root α, as
follows:
eδi −δ j = Eī j¯, eδi −εs = Eīs ,
eεs −εt = Est , eεs −δi = Esī , 1 ≤ i ̸= j ≤ m, 1 ≤ s ̸= t ≤ n.

eδi +δ j = Iī j¯, eδi +εs = Iīs ,


eεs +εt = Ist , e−δi −δ j = ∆ī j¯,
160 5. Howe duality

e−εs −εt = ∆st , e−δi −εs = ∆īs , 1 ≤ i < j ≤ m, 1 ≤ s ≤ t ≤ n.


We omit the details.
(2) Recall e1 , . . . , ek denote the standard basis for Ck . It follows by definition
of ω− that
ω− (e p , ek+1−p ) = −ω− (ek+1−p , e p ) = 1,
k/2
for 1 ≤ p ≤ k/2, and hence ∑ p=1 (e p ⊗ ek+1−p − ek+1−p ⊗ e p ) ∈ Ck ⊗ Ck is Sp(k)-
invariant. Thus, Iī j¯, Iīs , Ist defined above are Sp(k)-invariant in WC(U U ), where we
recall the definition of the basis (5.8) for U and we have regarded ∂x pi and ∂η∂ps as

elements in U ∗ . Similarly ∆ī j¯, ∆īs , ∆st defined above are Sp(k)-invariant in WC(UU ).
On the other hand, the elements in (5.24) are GL(k)-invariant (and hence Sp(k)-
invariant) in WC(U U ).
Hence, for g ∈ G, x ∈ g′ and z ∈ C[x, η], we have gx.z = xg.z, whence (2). 
Theorem 5.30 ((Sp(k), osp(2m|2n))-Howe duality, I). Sp(k) and osp(2m|2n) form
a Howe dual pair on S(Ck ⊗ Cm|n ), and the (Sp(k), osp(2m|2n))-module S(Ck ⊗
Cm|n ) is strongly multiplicity-free.

Proof. Set G = Sp(k). The action of sp(k) on S(Ck ⊗Cm|n ) clearly lifts to an action
of G. By Lemma 5.29, the actions of G and osp(2m|2n) on S(U) = S(Ck ⊗ Cm|n )
commute.
Let U = U ⊕ U ∗ . The associated graded for WC(U U ) is S(U
U ), and the action
of sp(k) on S(U U ) clearly lifts to an action of G. Let {y pi } denote the basis in U0̄∗
dual to the basis {x pi } for U0̄ , and let {ξ pt } denote the basis in U1̄∗ dual to the basis
{η pt } for U1̄ . By the supersymmetric First Fundamental Theorem ( (FFT) for G (see
Theorem 4.19), a generating set T for the algebra S(U U )G = S(U0̄ ⊕U0̄∗ ) ⊗ ∧(U1̄ ⊕
)G
U1̄∗ ) consists of zī j¯, zst , zīs , zsī in (5.12), Iī j¯, Iīs , Ist , and
(k
2 ) (
k
2 )
∑ pi k+1−p, j k+1−p,i p j ,
y y − y y ∑ pi k+1−p,s k+1−p,i ps ,
y ξ − y ξ
p=1 p=1
k
2 ( )
∑ ξ ps ξk+1−p,t − ξk+1−p,s ξ pt ,
p=1

for 1 ≤ i, j ≤ m, 1 ≤ s,t ≤ n. Observe that the symbol map σ sends the basis
elements (5.24) and (5.25) for g′ to the above elements in T for S(U
U )G . Now the
theorem follows from Proposition 5.12 and Lemma 5.29. 

5.3.2. (Sp(V ), osp(2m|2n))-Howe duality. Denote by u+ (respectively, u− ) the


subalgebra of osp(2m|2n) spanned by the ∆ (respectively, I) operators in (5.25).
The irreducible highest weight module L(osp(2m|2n), µ) below is understood to be
relative to the Borel subalgebra of osp(2m|2n) corresponding to the fundamental
system specified in the following Dynkin diagram:
5.3. Howe duality for symplectic and orthogonal groups 161

δ1 − δ2 ⃝
@ ⊗
@
⃝ ··· ⃝ ⃝ ··· ⃝
δ2 − δ3 δm − ε1 εn−1 − εn
−δ1 − δ2 ⃝
The root vectors associated to the simple roots −δ1 − δ2 , δi − δi+1 (1 ≤ i ≤
m − 1), δm − ε1 , εt − εt+1 (1 ≤ t ≤ n − 1), in the notations of (5.24) and (5.25), are
respectively
(5.26) ∆1̄2̄ , Eī,i+1 (1 ≤ i ≤ m − 1), Em1 , Et,t+1 (1 ≤ t ≤ n − 1).
Then gl(m|n) is a Levi subalgebra of osp(2m|2n) corresponding to the removal of
the simple root −δ1 − δ2 . We have a triangular decomposition
(5.27) osp(2m|2n) = u− ⊕ gl(m|n) ⊕ u+ .
In particular, the Lie superalgebras gl(m|n) and osp(2m|2n) share the same Cartan
subalgebra. An element f ∈ C[x, η] is called harmonic, if f is annihilated by the
+
subalgebra u+ . The space of harmonics C[x, η]u will be denoted by Sp H.
Recall that P(m|n) denotes the set of all (m|n)-hook partitions λ (i.e., λm+1 ≤
n), and also recall from (2.4) that, for an (m|n)-hook partition λ, λ♮ is a weight
for gl(m|n). Then λ♮ + ℓ1m|n can be regarded as a weight for the Lie superalgebras
gl(m|n) and osp(2m|2n), where we recall the definition of 1m|n from (1.29).
Theorem 5.31 ((Sp(k), osp(2m|2n))-Howe duality, II). Let k = 2ℓ. We have the
following decomposition of (Sp(k), osp(2m|2n))-modules:
⊕ ( )
S(Ck ⊗ Cm|n ) = L(Sp(k), λ) ⊗ L osp(2m|2n), λ♮ + ℓ1m|n .
λ∈P(m|n),ℓ(λ)≤ℓ

+
Proof. Observing that M u is a gl(m|n)-module for any osp(2m|2n)-module M
by the triangular decomposition (5.27), we may regard the space of harmonics
Sp H as an (Sp(k), gl(m|n))-module. Furthermore, since each graded subspace

of S(Ck ⊗ Cm|n ) as a gl(m|n)-module is a submodule of a tensor power of Cm|n


+
and hence is completely reducible, L(osp(2m|2n), µ)u is also a completely re-
ducible gl(m|n)-module, for any irreducible osp(2m|2n)-module L(osp(2m|2n), µ)
appearing in S(Ck ⊗ Cm|n ). It follows by the irreducibility of L(osp(2m|2n), µ) that
+
L(osp(2m|2n), µ)u is an irreducible gl(m|n)-module, and the highest weight vec-
tor of highest weight µ in L(osp(2m|2n), µ) remains to be a highest weight vector
+
in L(osp(2m|2n), µ)u . Therefore, we must have
L(osp(2m|2n), µ)u ∼
+
(5.28) = Lm|n (µ).

By Theorem 5.30, the (Sp(k), osp(2m|2n))-module S(Ck ⊗ Cm|n ) is strongly


multiplcity-free, and hence by (5.28), the (Sp(k), gl(m|n))-module Sp H is also
strongly multiplicity-free. To establish the explicit form of decomposition as stated
162 5. Howe duality

in the theorem, it suffices (and is indeed equivalent) to establish the following ex-
plicit decomposition of Sp H as an (Sp(k), gl(m|n))-module:

(5.29) Sp
H∼
= L(Sp(k), λ) ⊗ Lm|n (λ♮ + ℓ1m|n ).
λ∈P(m|n),ℓ(λ)≤ℓ

We first consider the limit case n = ∞ with the space of harmonics denoted by
Sp H∞ , where the condition λ ∈ P(m|n) reduces to λ ∈ P. Set ♢ = ♢1,2,...,λ1 ,λ′ . Ob-
serve by Theorem 5.23 that the vector ♢ associated to any partition λ with ℓ(λ) ≤ ℓ
is a joint sp(k) × gl(m|n)-highest weight vector and ♢ a polynomial independent of
the variables x pi and η pt for p ≥ ℓ + 1. Each ∆-operator in (5.25) is a second order
differential operator whose summands always involve differentiation with respect
to one of these variables for p ≥ ℓ + 1. Thus, ♢ is annihilated by u+ , and hence we
have ♢ ∈ Sp H∞ . One also easily checks that the vector ♢ has weight (λ, λ♮ + ℓ1m|n )
under the transformations of sp(k) × osp(2m|2n). We note that all the summands
on the right hand side of (5.29) indeed occur in Sp H∞ (with multiplicity one), and
all irreducible representations of Sp(k) occur. We conclude that (5.29) must hold
(when n = ∞).
Now consider the finite n case. We may regard S(Cd ⊗ Cm|n ) ⊆ S(Cd ⊗ Cm|∞ )
with compatible actions of osp(2m|2n) ⊆ osp(2m|2∞). We claim the following
equation on spaces of joint highest weight vectors holds:
(Sp H)n1 ×n(n) = (Sp H∞ )n1 ×n(∞) ∩ C[x, η],
where n1 (respectively, n(n)) denotes the nilradical of the standard Borel for sp(k)
(respectively, the nilradical generated by (5.26) for osp(2m|2n)). Note by definition
that (Sp H)n1 ×n(n) ⊇ (Sp H∞ )n1 ×n(∞) ∩ C[x, η]. Since each summand of the operators
in (5.24) and the ∆-operators in (5.25) that lie in osp(2m|2∞)\osp(2m|2n) must in-
volve ∂η∂ps for s > n, we conclude that (Sp H)n1 ×n(n) ⊆ (Sp H∞ )n1 ×n(∞) ∩ C[x, η], and
hence the claim is proved. Observe from the explicit formulas of the joint highest
vectors in Sp H∞ (see Theorem 5.23) that precisely those vectors corresponding to
(m|n)-hook partitions will not involve the variables η ps for s > n, and hence they
lie in (Sp H∞ )n1 ×n(∞) ∩ C[x, η]. This together with the claim above completes the
proof of the theorem. 

The irreducible osp(2m|2n)-modules appearing in the (Sp(k), osp(2m|2n))-


Howe duality decomposition above (for varied k) are called oscillator modules of
osp(2m|2n). The osp(2m|2n)-oscillator modules are infinite dimensional in gen-
eral, as the weights evaluated at the simple root −δ1 − δ2 is −λ1 − λ2 − k < 0.
Corollary 5.32 ((Sp(k), so(2m))-Howe duality). Let k = 2ℓ. We have the following
decomposition of (Sp(k), so(2m))-modules:
⊕ ( m )
S(Ck ⊗ Cm ) ∼
= L(Sp(k), λ) ⊗ L so(2m), ∑ i(λ + ℓ)δ i .
λ∈P,ℓ(λ)≤min{ℓ,m} i=1
5.3. Howe duality for symplectic and orthogonal groups 163

Corollary 5.33 ((Sp(k), sp(2n))-Howe duality). Let k = 2ℓ. We have the following
decomposition of (Sp(k), sp(2n))-modules:
⊕ ( n )
∧(Ck ⊗ Cn ) ∼
= L(Sp(k), λ) ⊗ L sp(2n), ∑ (λ′i − ℓ)εi .
λ∈P i=1
ℓ(λ)≤ℓ,ℓ(λ′ )≤n

5.3.3. Irreducible modules of O(V ). Take V = Ck . Let so(k) = g(V, ω+ ) be the


orthogonal Lie subalgebra of gl(k), and let O(k) = O(V, ω+ ) be the orthogonal
subgroup of GL(k), which preserve the non-degenerate symmetric bilinear form
ω+ on V = Ck associated to the k × k matrix Jk ; see (4.6). Let SO(V ) = O(V ) ∩
SL(V) denote the special orthogonal group of (V, ω+ ). The Lie algebras so(k), for
k = 2ℓ and k = 2ℓ + 1, consist precisely of the following k × k matrices in ℓ|ℓ- and
ℓ|1|ℓ-block forms, respectively (recall A♭ from (5.22)):
{( ) }
A B
(5.30) so(2ℓ) = | D = −A , B = −B,C = −C ,
♭ ♭ ♭
C D
  
 A x B 
(5.31) so(2ℓ + 1) =  y 0 z  | D = −A♭ , B♭ = −B,C♭ = −C ,
 
C w D
where in addition the ℓ × 1 matrices x = (xi1 ) and w = (wi1 ) and the 1 × ℓ matrices
y = (y1i ) and z = (z1i ) are related by z1i = −xℓ+1−i,1 and y1i = −wℓ+1−i,1 , for all
1 ≤ i ≤ ℓ. The standard Borel of so(k), for k even and odd, is the subalgebra
of upper triangular matrices in so(k), and hence is compatible with the standard
Borel of gl(k). The Cartan subalgebra of so(k) is spanned by the basis vectors
Eii − Ek+1−i,k+1−i , for 1 ≤ i ≤ ℓ, where k = 2ℓ or 2ℓ + 1 with ℓ ∈ Z+ . The following
fact is well-known (cf. [GW]).
Lemma 5.34. A finite-dimensional simple so(k)-module L(so(k), λ) of highest
weight λ = ∑ℓi=1 λi εi lifts to an SO(k)-module if and only if
(i) (λ1 , . . . , λℓ ) is a partition, for odd k = 2ℓ + 1;
(ii) (λ1 , . . . , λℓ−1 , |λℓ |) is a partition, for even k = 2ℓ.

Next we will describe the classification of simple O(k)-modules and give a


parametrization in terms of partitions. Let det denote the one-dimensional deter-
minant module of O(k).
First consider the case when k = 2ℓ + 1 is odd. In this case, −I ∈ O(k)\SO(k)
and O(k) is the direct product SO(k)×Z2 . Associated to a partition λ with ℓ(λ) ≤ ℓ,
we let L(O(k), λ) stand for the irreducible O(k)-module which restricts to the irre-
ducible SO(k)-module L(SO(k), λ) and on which the element −I acts trivially. We
shall denote the irreducible O(k)-module L(O(k), λ) ⊗ det by L(O(k), eλ), where the
Young diagram of the partition eλ is obtained from λ by replacing the first column
164 5. Howe duality

λ′1 by k − λ′1 . This way we have obtained a complete parametrization of the simple
O(k)-modules in terms of partitions µ with ℓ(µ) ≤ ℓ and µ′1 + µ′2 ≤ k = 2ℓ + 1.
Now consider the case when k = 2ℓ is even.

Proposition 5.35. Let k = 2ℓ be even. A simple O(k)-module L is exactly one of


the following (where all λi ∈ Z+ ):

(i) L is a direct sum of two irreducible SO(k)-modules of highest weights


(λ1 , · · · , λℓ−1 , λℓ ) and (λ1 , · · · , λℓ−1 , −λℓ ), where λℓ > 0. In this case,
L∼= L ⊗ det.
(ii) For λ = (λ1 , · · · , λℓ−1 , 0), there are exactly two non-isomorphic irreducible
O(k)-modules that restrict to an isomorphic irreducible SO(k)-module of
highest weight λ. If one of these modules is L, then the other one is
L ⊗ det.

We shall denote the simple O(k)-module in Proposition 5.35 (i) by L(O(k), λ).
Let τ ∈ O(k) \ SO(k) denote the element that switches the basis vector eℓ with
eℓ+1 and fixes all other standard basis vectors ei ’s of Ck . We declare L(O(k), λ) to
be the O(k)-module in Proposition 5.35(ii) on which τ acts on an SO(k)-highest
weight vector trivially. Note that τ transforms an SO(k)-highest weight vector in
the O(k)-module L(O(k), λ) ⊗ det by −1. We shall denote the simple O(k)-module
e where as before the Young diagram of the partition
L(O(k), λ) ⊗ det by L(O(k), λ),
eλ is obtained from λ by replacing the first column λ′ by k − λ′ . In this way, we
1 1
have obtained a complete parametrization of the simple O(k)-modules in terms of
partitions µ with ℓ(µ) ≤ ℓ such that µ′1 + µ′2 ≤ k = 2ℓ.
Summarizing the above discussion for k odd and even, we have obtained the
following uniform description.

Proposition 5.36. A complete set of pairwise non-isomorphic simple O(k)-modules


consists of L(O(k), µ), where µ runs over all partitions of length no greater than k
with µ′1 + µ′2 ≤ k.

5.3.4. Howe dual pair (O(k), spo(2m|2n)). We consider the superspace U = U0̄ ⊕
U1̄ with U0̄ = V m and U1̄ = V n , and identify naturally U ≡ V ⊗ Cm|n as before. We
continue the identification S(U) ≡ C[x, η] as in (5.9).
The action of the Lie algebra so(k) as the subalgebra of gl(k) on C[x, η] given
in (5.10) lifts to an action of Lie group O(V ) = O(k). On the other hand, recall
the action of gl(m|n) on C[x, η] given by (5.24). Introduce the following additional
5.3. Howe duality for symplectic and orthogonal groups 165

operators on C[x, η]:


k k
∂ ∂
Jī j¯ = ∑ x pi xk+1−p, j , ∇ī j¯ = ∑ ∂x pi ∂xk+1−p, j ,
p=1 p=1
k k
∂ ∂
(5.32) Jīs = ∑ x pi ηk+1−p,s , ∇īs = ∑ ∂x pi ∂ηk+1−p,s , 1 ≤ i ≤ j ≤ m,
p=1 p=1
k k
∂ ∂
Jst = ∑ η ps ηk+1−p,t , ∇st = ∑ ∂η ps ∂ηk+1−p,t , 1 ≤ s < t ≤ n.
p=1 p=1

Let us denote by g′ the linear span of the elements in (5.24) and (5.32). Note that
the elements in (5.24) and (5.32) form a linear basis for g′ .

Lemma 5.37. (1) Under the super-commutator g′ forms a Lie superalgebra,


which is isomorphic to spo(2m|2n).
(2) The actions of O(k) and spo(2m|2n) on S(Ck ⊗ Cm|n ) commute.

Proof. (1) Letting U = Ck ⊗ Cm|n and U = U ⊕ U ∗ we identify WC(U U ) with the



superalgebra of differential operators on S(U). We can then regard g as a subspace
inside G ∼= spo(U U ) as in Proposition 5.5. In a completely analogous fashion as in
the proof of Lemma 5.29, we can show that g′ ∼ = spo(2m|2n) with an explicit iden-
tification of root vectors with elements in (5.24) and (5.32). We omit the details.
(2) Recall the standard basis e1 , . . . , ek for Ck . By definition of ω+ , we have
ω+ (e p , ek+1−p ) = 1,
for 1 ≤ p ≤ k, and hence ∑kp=1 e p ⊗ ek+1−p ∈ Ck ⊗ Ck is O(k)-invariant. It follows
that Jī j¯, Jīs , Jst defined above are O(k)-invariant in WC(UU ). Dually, the operators
∇ī j¯, ∇īs , ∇st are also O(k)-invariant in WC(U U ). On the other hand, the elements in
(??) are clearly O(k)-invariant in WC(U U ). Hence (2) follows. 

Theorem 5.38 ((O(k), spo(2m|2n))-Howe duality, I). O(k) and spo(2m|2n) form a
Howe dual pair on S(Ck ⊗Cm|n ), and the (O(k), spo(2m|2n))-module S(Ck ⊗Cm|n )
is strongly multiplicity-free.

Proof. Set G = O(k). The action of so(k) on S(Ck ⊗ Cm|n ) lifts to an action of
G. By Lemma 5.37, the actions of G and spo(2m|2n) on S(U) = S(Ck ⊗ Cm|n )
commute.
Let U = U ⊕U ∗ so that the associated graded for WC(U U ) is S(UU ). Thus, the
action of so(k) on WC(U U ) lifts to an action of G. Let {y pi } denote the basis in U0̄∗
dual to the basis {x pi } for U0̄ , and let {ξ pt } denote the basis in U1̄∗ dual to the basis
{η pt } for U1̄ . By the supersymmetric First Fundamental Theorem ( (FFT) for G (see
Theorem 4.19), a generating set T for the algebra S(U U )G = S(U0̄ ⊕U0̄∗ ) ⊗ ∧(U1̄ ⊕
166 5. Howe duality

)G
U1̄∗ ) consists of zī j¯, zst , zīs , zsī in (5.12), Jī j¯, Jīs , Jst , and
k k k
∑ y pi yk+1−p, j , ∑ y pi ξk+1−p,s , ∑ ξ ps ξk+1−p,t ,
p=1 p=1 p=1
for 1 ≤ i, j ≤ m, 1 ≤ s,t ≤ n. The symbol map σ sends the basis elements (5.24)
and (5.32) for g′ to the above elements in T, and hence the theorem follows from
Proposition 5.12 and Lemma 5.37. 

5.3.5. (O(V ), spo(2m|2n))-Howe duality. Denote by u+ (respectively, u− ) the


subalgebra of spo(2m|2n) spanned by the ∇ (respectively, J) operators in (5.32).
The irreducible highest weight module L(spo(2m|2n), µ) below is understood to be
relative to the Borel subalgebra of spo(2m|2n) corresponding to the fundamental
system specified in the following Dynkin diagram:

⃝=⇒⃝ ⃝ ··· ⃝ ··· ⃝
−2δ1 δ1 − δ2 δ2 − δ3 δm − ε1 ε1 − ε2 εn−1 − εn

The root vectors associated to the simple roots −2δ1 , δi − δi+1 (1 ≤ i ≤ m −


1), δm − ε1 , εt − εt+1 (1 ≤ t ≤ n − 1), in the notations of (5.24) and (5.32), are
respectively
(5.33) ∇1̄1̄ , Eī,i+1 (1 ≤ i ≤ m − 1), Em̄1 , Et,t+1 (1 ≤ t ≤ n − 1).
Then gl(m|n) is a Levi subalgebra of spo(2m|2n) corresponding to the removal of
the simple root −2δ1 . We have a triangular decomposition
(5.34) spo(2m|2n) = u− ⊕ gl(m|n) ⊕ u+ .
In particular, the Lie superalgebras gl(m|n) and spo(2m|2n) share the same Cartan
subalgebra. For λ ∈ P(m|n), λ♮ + 2k 1m|n is a weight for gl(m|n), and hence can be
regarded as a weight for spo(2m|2n) as well.
Theorem 5.39 ((O(k), spo(2m|2n))-Howe duality, II). We have the following de-
composition of (O(k), spo(2m|2n))-modules:
⊕ ( k )
S(Ck ⊗ Cm|n ) = L(O(k), λ) ⊗ L spo(2m|2n), λ♮ + 1m|n .
λ∈P(m|n)
2
λ′1 +λ′2 ≤k

+ +
Proof. Denote the space of harmonics C[x, η]u by O H. Then M u is a gl(m|n)-
module for any spo(2m|2n)-module M by the triangular decomposition (5.34), so
that O H is an (O(k), gl(m|n))-module. Similar as in the proof of Theorem 5.31 it
is enough to establish the following decomposition of O H as an (O(k), gl(m|n))-
module:
⊕ ( k )
(5.35) O
H= ∼ L(O(k), λ) ⊗ Lm|n λ♮ + 1m|n .
λ∈P(m|n)
2
λ′1 +λ′2 ≤k
5.3. Howe duality for symplectic and orthogonal groups 167

The validity of (5.35) for finite n follows from the case when n = ∞ in a completely
parallel way as in the proof of Theorem 5.31. Hence it remains to establish (5.35)
in the limit case n = ∞, where the condition λ ∈ P(m|n) reduces to λ ∈ P.
Assume now n = ∞. Set ♢ = ♢1,2,...,λ1 ,λ′ , for λ ∈ P with λ′1 + λ′2 ≤ k. We
claim that ♢ ∈ O H, i.e, ♢ is annihilated by u+ . By Theorem 5.23 ♢ is a joint
so(k) × gl(m|n)-highest weight vector. To establish the claim, it suffices to show
that ♢ is annihilated by the simple root vector ∇1̄1̄ , since ♢ is clearly annihilated
by the remaining simple root vectors in (5.33) which lie in gl(m|n).
To show that ∇1̄1̄ (♢) = 0, we recall that ∇1̄1̄ = ∑kp=1 ∂x∂p1 ∂xk+1−p,1

and recall
from (5.18) that ♢ = ♢1,λ′1 · · · ♢λ1 ,λ′λ . We observe that each summand in the expan-
1
∂♢i,λ′ ∂♢ j,λ′
sion of the double differentiation ∇1̄1̄ (♢) is either of the form (i) i
∂x p1 ∂xk+1−p,1
j
···
∂2 ♢i,λ′
for i ̸= j, or of the form (ii) ∂x p1 ∂xk+1−p,1 · · · . Since λ′1 + λ′2 ≤ k by assumption,
i

we have λ′i + λ′j ≤ k for i ̸= j. Noting by definition (5.16) and (5.17) that the x pq
appearing in ♢i,λ′i satisfy the constraints p ≤ i, the derivative in (i) must be zero.
Expanding the determinant ♢i,λ′i along the first column, we obtain ♢i,λ′i = ∑ p x p1 A p ,
where A p is an expression not containing any xq1 . Therefore, the derivative in (ii)
must also be zero, and hence ∇1̄1̄ (♢) = 0.
When k = 2ℓ, recall the element τ ∈ O(k) \ SO(k) defined in Section 5.3.3.
When k = 2ℓ + 1, we let τ be the linear map on Ck that fixes e p , for all p ̸= ℓ + 1,
and that sends eℓ+1 to −eℓ+1 . By examining how τ transforms the so(k)-highest
weight vector ♢, for λ′1 + λ′2 ≤ k, one concludes that the O(k)-module generated
by ♢ is L(O(k), λ). Another direct computation implies that ♢ has spo(2m|2n)-
weight λ♮ + 2k 1m|n . Thus, all irreducible representations of O(k) occur, and all the
summands on the right hand side of (5.35) occur in the space of harmonics O H.
Therefore, (5.35) holds in the case when n = ∞. The theorem is proved. 

The simple spo(2m|2n)-modules appearing in the above (O(k), spo(2m|2n))-


Howe duality decomposition (for varied k) will be called oscillator modules of
spo(2m|2n). The spo(2m|2n)-oscillator modules are infinite dimensional in gen-
eral, as the weights evaluated at the simple root −2δ1 is −2λ1 − k < 0.
Corollary 5.40 ((O(k), sp(2m))-Howe duality). We have the following decompo-
sition of (O(k), sp(2m))-modules:
⊕ ( m
k )
S(Ck ⊗ Cm ) ∼
= L(O(k), λ) ⊗ L sp(2m), ∑ (λi + )δi .
′ ′ 2
ℓ(λ)≤m,λ1 +λ2 ≤k i=1

Corollary 5.41 ((O(k), so(2n))-Howe duality). We have the following decomposi-


tion of (O(k), so(2n))-modules:
⊕ ( n
k )
∧(Ck ⊗ Cn ) ∼
= L(O(k), λ) ⊗ L so(2n), ∑ (λ′i − )εi .
′ ′ ′ 2
ℓ(λ )≤n,λ1 +λ2 ≤k i=1
168 5. Howe duality

5.4. Howe duality for infinite-dimensional Lie algebras


In this section, we formulate Howe dualities between classical Lie groups and
infinite-dimensional Lie algebras in fermionic Fock spaces. The results of this
section will be applied in Section 5.5 to derived character formulas for the irre-
ducible oscillator modules of the Lie superalgebras arising in the Howe duality
decompositions in Section 5.3.

5.4.1. Lie algebras a∞ , c∞ and d∞ . We define the infinite-dimensional Lie alge-


b ∞ and its Lie subalgebras c∞ and d∞ of type C and D.
bras a∞ ≡ gl
The Lie algebra a∞ ≡ gl b ∞ . Denote by gl∞ the Lie algebra of all matrices
(ai j )i, j∈Z with ai j = 0 for all but finitely many i’s and j’s. Let the degree of the ele-

mentary matrix Ei j be j − i. This defines a Z-principal gradation gl∞ = j∈Z gl∞, j .
Denote by a∞ ≡ gl b ∞ = gl∞ ⊕ CK the central extension associated to the following
2-cocycle τ:
(5.36) τ(A, B) = Tr ([J, A]B) ,
where J = ∑ j≤0 Eii . Denoting by [·, ·]′ the Lie bracket on gl∞ , we introduce a
bracket [·, ·] on the central extension a∞ = gl∞ ⊕ CK by
[A + aK, B + bK] := [A, B]′ + τ(A, B)K, A, B ∈ gl∞ , a, b ∈ C.
By definition, a 2-cycle τ satisfies the following conditions: for A, B,C ∈ gl∞ ,
(i) τ(A, B) = −τ(B, A); (ii) τ(A, [B,C]) + τ(B, [C, A]) + τ(C, [A, B]) = 0.
The 2-cocyle condition on τ ensures that the bracket [·, ·] is skew-symmetric and
satisfies the Jacobi identity. More explicitly, we have the following commutation
relation on a∞ : for i, j, m, n ∈ Z,
( )
[Ei j , Emn ] = δ jm Ein − δin Em j + δ jm δin θ{i ≤ 0} − θ{ j ≤ 0} K,
where the Boolean characteristic function is defined to be
{
1, if the statement P is true,
(5.37) θ{P} =
0, if the statement P is false.
Remark 5.42. The Lie algebra a∞ can be “completed” in the following sense (see
[DJKM]). Denote by gl e ∞ the completed Lie algebra of matrices (ai j )i, j∈Z with
ai j = 0 for |i − j| ≫ 0. The formula for τ in (5.36) also makes sense for gl e ∞ and
thus we obtain a central extension of gl e ∞ . The τ is a 2-coboundary on gl∞ , leading
e ∞ is no
to a trivial central extension a∞ , as we have defined in this book. But τ on gl
e
longer a 2-coboundary, and it gives a non-trivial central extension of gl∞ . Similar
remarks apply to the classical subalgebras of gl∞ and gl e ∞ below in this section.
The completed Lie algebras and their central extensions are important because of
their relations to various subalgebras such as Heisenberg, Virasoro, and affine Lie
algebras, but they play no particular role in this book. Actually in Chapter 6, the
5.4. Howe duality for infinite-dimensional Lie algebras 169

uncompleted variants are preferred as we need to compare with finite rank Lie
algebras. The main results in this section make sense when formulated for these
completed Lie algebras.

We extend the Z-gradation of the Lie algebra gl∞ to a∞ by putting the degree
of K to be 0. In particular, we have a triangular decomposition
a∞ = a∞+ ⊕ a∞0 ⊕ a∞− ,
where ⊕
a∞± = gl∞,± j , a∞0 = gl∞,0 ⊕ CK.
j∈N

Denote by εi the element in (a∞0 )∗ determined by εi (E j j ) = δi j and εi (K) =


0, for i, j ∈ Z. Then the root system of a∞ is {εi − ε j | i, j ∈ Z, i ̸= j} with a
fundamental system {εi − εi+1 | i ∈ Z}. The corresponding set of simple coroots
is given by {Hia := Eii − Ei+1,i+1 + δi,0 K | i ∈ Z}. Let Λaj , for j ∈ Z, be the jth
fundamental weight for a∞ , i.e., Λaj (Hia ) = δi j , for all i ∈ Z, and Λaj (K) = 1. A
direct computation shows that
{ a
Λ0 − ∑0k= j+1 εk , for j < 0,
(5.38) Λj =
a
j
Λa0 + ∑k=1 εk , for j ≥ 1.
Let L(a∞ , Λ) denote the irreducible highest weight a∞ -module with highest weight
Λ ∈ (a∞0 )∗ with respect to the standard Borel subalgebra a∞+ ⊕ a∞0 . The level of
Λ is given by the scalar Λ(K) = ∑i∈Z Λ(Hia ).
The Lie algebra c∞ . Now consider the natural gl∞ -module V0 with basis {vi |
i ∈ Z} such that Ei j vk = δ jk vi . Let C be the following skew-symmetric bilinear
form on V0 :
C(vi , v j ) = (−1)i δi,1− j , ∀i, j ∈ Z.
Denote by c∞ the Lie subalgebra of gl∞ which preserves the bilinear form C:
c∞ = {a ∈ gl∞ | C(a(u), v) +C(u, a(v)) = 0, ∀u, v ∈ V0 }
= {(ai j )i, j∈Z ∈ gl∞ | ai j = −(−1)i+ j a1− j,1−i }.
Denote by c∞ = c∞ ⊕ CK the central extension of c∞ associated to the 2-cocycle
(5.36) restricted to c∞ . Then c∞ inherits from a∞ a natural triangular decomposition:
c∞ = c∞+ ⊕ c∞0 ⊕ c∞− ,
where c∞± = c∞ ∩ a∞± , c∞0 = c∞ ∩ a∞0 . Let {εi | i ∈ N} be the basis in (c∞0 )∗ dual
to the basis {Ei,i − E−i+1,−i+1 | i ∈ N} in c∞0 . The Dynkin diagram for c∞ with a
standard simple system is given as follows:

⃝=⇒⃝ ⃝ ··· ⃝ ⃝ ⃝ ···


β× := −2ε1 ε1 − ε2 ε2 − ε3 εi − εi+1
170 5. Howe duality

A set of simple coroots {Hic , i ≥ 0} for c∞ is


Hic = Eii + E−i,−i − Ei+1,i+1 − E1−i,1−i , i ∈ N,
H0c = E0,0 − E1,1 + K.
Then we denote by Λcj ∈ (c∞0 )∗ , for j ∈ Z+ , the jth fundamental weight of c∞ , i.e.,
Λcj (Hic ) = δi j for all i ∈ Z+ . Note that Λcj (K) = 1 for all j. A direct computation
shows that
j
(5.39) Λcj = ∑ εk + Λc0 , for j ≥ 1.
k=1

Denote by L(c∞ , Λ) the irreducible highest weight module of c∞ of highest weight


Λ ∈ (c∞0 )∗ , whose level is given by Λ(K) = ∑i≥0 Λ(Hic ).
The Lie algebra d∞ . We denote by d∞ the Lie subalgebra of gl∞ preserving
the following symmetric bilinear form D on V0 :
D(vi , v j ) = δi,1− j , ∀i, j ∈ Z.
Namely we have
d∞ = {a ∈ gl | D(a(u), v) + D(u, a(v)) = 0, ∀u, v ∈ V0 }
= {(ai j )i, j∈Z ∈ gl∞ | ai j = −a1− j,1−i }.

Denote by d∞ = d∞ ⊕ CK the central extension associated to the 2-cocycle (5.36)


b ∞:
restricted to d∞ . Then d∞ has a triangular decomposition induced from gl
d∞ = d∞+ ⊕ d∞0 ⊕ d∞− ,
where d∞± = d∞ ∩ a∞± and d∞0 = d∞ ∩ a∞0 .
The Cartan subalgebra d∞0 of d∞ is the same as c∞0 (both as subalgebras of
a∞ ), and we let ε j ∈ (d∞0 )∗ be defined in the same way as in the case of c∞ . The
Dynkin diagram with a standard fundamental system for d∞ is given as follows:
β× := −ε1 − ε2 ⃝
@
@
⃝ ⃝ ··· ⃝ ⃝ ⃝ ···
ε2 − ε3 ε3 − ε4 εi − εi+1
ε1 − ε2 ⃝
A set of simple coroots {Hid | i ≥ 0} for d∞ is given by
Hid = Eii + E−i,−i − Ei+1,i+1 − E1−i,1−i (i ∈ N),
H0d = E0,0 + E−1,−1 − E2,2 − E1,1 + 2K.
Denote by Λdj ∈ (d∞0 )∗ , for j ∈ Z+ , the jth fundamental weight of d∞ , i.e, Λdj (Hid ) =
δi j for all i ∈ Z+ . Denote by L(d∞ , Λ) the irreducible highest weight d∞ -module of
highest weight Λ, whose level is given by Λ(K) = 12 Λ(H0d )+ 12 Λ(H1d )+ ∑i≥2 Λ(Hid ).
5.4. Howe duality for infinite-dimensional Lie algebras 171

5.4.2. The fermionic Fock space. Recall from (A.62) that F denotes the fermionic
Fock space generated by the fermions ψ± (z), whose components ψ± n satisfy the
Clifford commutation relation (A.60).
We shall denote by Fℓ the fermionic Fock space of ℓ pairs of fermions

ψ±,p (z) = ∑ ψ±,p −r− 2 1


r z , 1 ≤ p ≤ ℓ.
r∈ 12 +Z

More precisely, we let Cbℓ be the Clifford algebra generated by ψ±,p


r , where 1 ≤
p ≤ ℓ and r ∈ 12 + Z, with anti-commutation relations given by:
−,q −,p −,q
r , ψs ]+ = δ p,q δr,−s I,
[ψ+,p r , ψs ]+ = [ψr , ψs ]+ = 0,
[ψ+,p +,q

for all r, s ∈ 12 + Z, 1 ≤ p, q ≤ ℓ. Then Fℓ is the simple Cbℓ -module generated by


the vacuum vector |0⟩, which satisfies the condition ψ±,p
r |0⟩ = 0 for all r > 0 and
1 ≤ p ≤ ℓ. Here I is identified with the identity operator on Fℓ .
Introduce a neutral fermionic field

∑ ϕr z−r− 2
1
ϕ(z) =
r∈ 21 +Z

whose components satisfy the following anti-commutation relation:


1
[ϕr , ϕs ]+ = δr,−s I, r, s ∈ + Z.
2
1
We denote by C bℓ+ 2 the Clifford algebra generated by ϕr and ψ±,p
r subject to the
additional anti-commutation relation [ϕr , ψ±,p
s ]+ = 0, where 1 ≤ p ≤ ℓ. By the Fock
1
1
space Fℓ+ 2 we mean the simple C bℓ+ 2 -module generated by the vacuum vector |0⟩,
which satisfies the condition ϕr |0⟩ = ψ±,p
r |0⟩ = 0 for all r > 0 and 1 ≤ p ≤ ℓ.
ℓ+ 1
The algebras Cb and C

b 2 are naturally filtered algebras by letting the degree
±,p
of each ψr and ϕr be 1 and the degree of I be 0. The associated graded algebras
are exterior algebras. We introduce natural 12 Z+ -gradations (called the principal
1
gradation) on Fℓ and Fℓ+ 2 by the eigenvalues of the degree operator d on Fℓ and
1
Fℓ+ 2 defined by
[ ]
d|0⟩ = 0, d, ψ±,p ±,p
−r = rψ−r , [d, ϕ−r ] = rϕ−r , ∀r, p.

1
Every graded subspace of Fℓ and Fℓ+ 2 with respect to the principal gradation is
finite dimensional.
Introduce the normal ordered product (denoted by : :)
172 5. Howe duality

{ −,q
−,q −ψs ψ+,p
r , if s = −r < 0,
r ψs :
:ψ+,p = +,p −,q
ψr ψs , otherwise,
(5.40) {
−ψs ψ−,p
+,q
r , if s = −r < 0,
:ψ−,p
r ψs :
+,q
= −,p +,q
ψr ψs , otherwise.
+,q −,p −,p
+,q ±,p −,q
±,p −,q
Also let :ψ+,p +,p q q
r ψs : = ψr ψs , :ψr ψs : = ψr ψs , :ψr ϕs : = ψr ϕs ,
for all p, q, r, s.

5.4.3. (GL(ℓ), a∞ )-Howe duality. Let



∑ Ei j zi−1 w− j = ∑ :ψ+,p (z)ψ−,p (w):.
i, j∈Z p=1

Equivalently, we have

(5.41) Ei j = ∑ :ψ+,p−i ψ−,p
1
2j−
:. 1
2
p=1

Consider the following generating functions

∑ e∗pq (n)z−n−1 = :ψ+,p (z)ψ−,q (z):, p, q = 1, . . . , ℓ.


n∈Z

Such generating functions are “vertex operators” in the theory of vertex algebras,
pq
but we will avoid such terminology in this book. We will only need e∗ (0) below,
pq pq
and hence introduce a short-hand notation E∗ ≡ e∗ (0). It follows that
(5.42) E∗pq = ∑ :ψ+,p −,q
−r ψr :, p, q = 1, . . . , ℓ.
r∈ 12 +Z

Lemma 5.43. (1) Sending Ei j to Ei j , for i, j ∈ Z, defines a representation of


the Lie algebra a∞ on Fℓ of level ℓ.
pq
(2) Sending E pq to E∗ in (5.42), for 1 ≤ p, q ≤ ℓ, defines an action of the Lie
algebra gl(ℓ), which lifts to an action of GL(ℓ), on Fℓ .
(3) GL(ℓ) and a∞ form a Howe dual pair on Fℓ .

Proof. (1) We need to check that


( )
[Ei j , Emn ] = δ jm Ein − δin Em j + δ jm δin θ{i ≤ 0} − θ{ j ≤ 0} ℓI.
Let us check in detail the case when m = j and n = i, as the remaining cases are
similar. Since clearly [Eii , Eii ] = 0, we can assume that i ̸= j. It follows from (5.41)
that
ℓ [ ]
−,p +,p −,p
[Ei j , E ji ] = ∑ ψ+,p
1 ψ
−i j− 1 ,ψ1 ψ
− j i− 1
2 2 2 2
p=1
5.4. Howe duality for infinite-dimensional Lie algebras 173

(ℓ )
+,p −,p +,p −,p +,p −,p +,p −,p
= ∑ −i j− − j i−
ψ 1 ψ 1 ψ 1 ψ
2
1 − ψ 1 ψ
− j i− 1 ψ
2
1 ψ
−i j−
2
1
2 2 2 2 2
p=1
(ℓ )
+,p −,p +,p −,p
= ∑ 1 −i i− 1 1 − j j− 1
ψ ψ − ψ
2
ψ
2 2 2
p=1
(ℓ ) ( )
+,p −,p +,p −,p
= ∑ 1 −i i− 1
:ψ ψ : − :ψ 1 ψ
− j j−
2
1 : + θ{i ≤
2
0} − θ{ j ≤ 0}
2
ℓI
2
p=1
( )
= Eii − E j j + θ{i ≤ 0} − θ{ j ≤ 0} ℓI.
pv pq uq
(2) We need to check that [E∗ , Euv
∗ ] = δqu E∗ − δ pv E∗ for 1 ≤ p, q, u, v ≤ ℓ.
Indeed, it follows by (5.42) that
[ +,p −,q ]
∗ ]= ∑
−,v
[E∗pq , Euv :ψ−r ψr :, :ψ+,u
−r ψr :
r∈ 12 +Z
( +,u −,q )
= ∑ δqu :ψ+,p −,v
−r ψr : − δvp :ψ−r ψr : = δqu E∗ − δ pv E∗ .
pv uq

r∈ 12 +Z
pq
As the operators E∗ have degree zero in the principal gradation, the action of gl(ℓ)
preserves each (principally) graded subspace of Fℓ , which is finite dimensional
and of integral weight. Hence the action of gl(ℓ) lifts to an action of the Lie group
GL(ℓ) on Fℓ .
(3) It follows by (5.41) and (5.42) that
[ ℓ ]
+,u −,u
[E∗pq , Ei j ] = ∑ −r ψr :, ∑ :ψ 1 −i ψ j− 1 :
:ψ+,p −,q
2 2
r∈ 12 +Z u=1
[ ] [ ]
−,q +,q −,q +,p −,q +,p −,p
= :ψ+,p
1 ψ
−i i− 1
:, :ψ 1 ψ
−i j− 1
: + :ψ 1 ψ
− j j− 1
:, :ψ 1 ψ
−i j− 1
:
2 2 2 2 2 2 2 2
−,q −,q
= :ψ+,p
1
−i
ψ j− 1 : − :ψ+,p
1
−i
ψ j− 1 : = 0.
2 2 2 2

Hence the actions of GL(ℓ) and a∞ on Fℓ commute.


The statement on Howe dual pair can be proved by a similar argument as for
Theorem 5.14, and we rephrase briefly the argument as follows. Set G = GL(ℓ),
U = Cℓ ⊗ C∞ ⊕ Cℓ∗ ⊗ C∞ , where C∞ is a vector space with a basis {wr | r ∈ − 21 −
Z+ }. The dual basis in C∞∗ is denoted by {w−r | r ∈ − 12 − Z+ }. We take a basis
{v+,p | 1 ≤ p ≤ ℓ} for Cℓ , and a dual basis {v−,p | 1 ≤ p ≤ ℓ} for Cℓ∗ . The Clifford
algebra Cbℓ can be naturally identified with WC(U U ) = C(UU ) for the purely odd
space U = U ⊕U ∗ by setting ψ±,p r ≡ v±,p ⊗ wr , so that its associated graded can be
identified with ∧(U U ). Observe that the images under the symbol map σ (defined in
5.1.2) of Ei j in (5.41) form a set of generators for ∧(UU )G , by the supersymmetric
First Fundamental Theorem of invariant theory for GL(V ) (see Theorem 4.19).
Now the claim on Howe dual pair follows from Proposition 5.12. 
174 5. Howe duality

Recall the well-known fact that a simple gl(ℓ)-module L(gl(ℓ), λ) of highest


weight λ = ∑ℓi=1 λi δi (with respect to the standard Borel subalgebra spanned by E pq
for p ≤ q) lifts to a simple GL(ℓ)-module L(GL(ℓ), λ) if and only if (λ1 , . . . , λℓ ) is
a generalized partition in the sense that all λi ∈ Z and λ1 ≥ . . . ≥ λℓ . By abuse of
notation, we shall identify a weight λ for gl(ℓ) with a generalized partition of the
form λ = (λ1 , . . . , λℓ ).
For a generalized partition λ = (λ1 , . . . , λℓ ) we define, for i ∈ Z,
{
|{ j | λ j ≥ i}|, for i ≥ 1,
λ′i :=
−|{ j | λ j < i}|, for i ≤ 0.
Theorem 5.44 ((GL(ℓ), a∞ )-Howe duality). As a (GL(ℓ), a∞ )-module, we have

(5.43) Fℓ ∼
= L(GL(ℓ), λ) ⊗ L(a∞ , Λa (λ)),
λ
where the summation is over all generalized partitions of the form λ = (λ1 , . . . , λℓ ),
and Λa (λ) := Λaλ1 + . . . + Λaλℓ = ℓΛa0 + ∑i∈Z λ′i εi .

Proof. By Lemma 5.43 and Proposition 5.12, we have a strongly multiplicity-free


decomposition of the (GL(ℓ), a∞ )-module Fℓ . To obtain the explicit decomposi-
tion as given in the theorem, we will exhibit an explicit formula for a joint highest
weight vector associated to each λ; see Proposition 5.45 below. The theorem fol-
lows as we observe by Proposition 5.45 that every simple GL(ℓ)-module appears
in the decomposition (5.43) of Fℓ . 

For 1 ≤ p ≤ ℓ and m ≥ 1, we denote


+,p +,p +,p
Ξ+,p
m := ψ−m+ 1 · · · ψ− 3 ψ− 1 ,
2 2 2
−,p −,p −,p
Ξ−,p
m := ψ−m+ 1 · · · ψ− 3 ψ− 1 ,
2 2 2

We make the convention that Ξ±,p


0 = 1.
Proposition 5.45. Given a generalized partition of the form λ = (λ1 , . . . , λℓ ), let
i, j be such that
λ1 ≥ · · · ≥ λi > λi+1 = · · · = λ j−1 = 0 > λ j ≥ · · · ≥ λℓ .
Then the joint highest weight vector in Fℓ associated to λ with respect to the stan-
dard Borel for gl(ℓ) × a∞ is
−, j −, j+1 −,ℓ
(5.44) vaλ = Ξ+,1 +,2 +,i
λ1 Ξλ2 · · · Ξλi · Ξ−λ j Ξ−λ j+1 · · · Ξ−λℓ |0⟩,

whose weights with respect to gl(ℓ) and a∞ are λ and Λa (λ), respectively.

Proof. The vector vaλ in (5.44) is indeed a highest weight vector for gl(ℓ) and a∞ ,
respectively, since applying any positive root vector in either gl(ℓ) or a∞ to vaλ is
either manifestly zero, or gives rise to two identical ψ∗∗
∗ in the resulting monomial,
5.4. Howe duality for infinite-dimensional Lie algebras 175

whence also zero. Another direct calculation shows that the highest weight of vaλ
for gl(ℓ) is (λ1 , . . . , λℓ ).
We recall that Emm = ∑ℓp=1 : ψ+,p ψ−,p1 :. When m ≥ 1 and n ≥ 1 one easily
2 −m m− 2
1

computes
+,q +,q −,q
[Emm , ψ−n+ 1 ] = δm,n ψ−n+ 1 , [Emm , ψ−n+ 1 ] = 0.
2 2 2

For m ≤ 0 and n ≥ 1 we have


−,q −,q +,q
[Emm , ψ−n+ 1 ] = −δ−m+1,n ψ−n+ 1 , [Emm , ψ−n+ 1 ] = 0.
2 2 2

This implies that the a∞ -weight of the vector vaλ equals ℓΛa0 + ∑i λ′i εi , which is also
equal to Λaλ1 + . . . + Λaλℓ by (5.38). 

5.4.4. (Sp(k), c∞ )-Howe duality. Let k = 2ℓ. It follows by (5.41) that


∑ (Ei, j − (−1)i+ j E1− j,1−i )zi−1 w− j
i, j∈Z
= ∑ℓp=1 (:ψ+,p (z)ψ−,p (w): + :ψ+,p (−w)ψ−,p (−z):) .
Equivalently, we have
ℓ( )
+,p −,p i+ j +,p −,p
(5.45) Ei, j − (−1)i+ j E1− j,1−i = ∑ :ψ 1 ψ
−i j− 1 :
2
− (−1) :ψ
2 j− 1 ψ 1
−i
: .
2 2
p=1

Consider the following generating functions


∑ ẽ pq (n)z−n−1 = :ψ−,p (z)ψ−,q (−z):,
n∈Z

∑ ẽ∗∗pq (n)z−n−1 = :ψ+,p (z)ψ+,q (−z):, p, q = 1, . . . , ℓ.


n∈Z
We introduce the following short-hand notations:
Ẽ pq ≡ ẽ pq (0), Ẽ∗∗
pq
≡ ẽ∗∗
pq
(0).
It follows that
∑ (−1)−r− 2 :ψ−,p −,q 1
Ẽ pq = −r ψr :,
r∈ 12 +Z
(5.46)
∑ (−1)−r− 2 :ψ+,p
1
Ẽ∗∗
pq
= −r ψr :.
+,q

r∈ 12 +Z
pq qp
Note that Ẽ pq = Ẽqp and Ẽ∗∗ = Ẽ∗∗ .
Lemma 5.46. Let k = 2ℓ.
(1) The formula (5.45) defines an action of Lie algebra c∞ on Fℓ of level ℓ.
pq pq
(2) The operators Ẽ pq and Ẽ∗∗ in (5.46) together with E∗ in (5.42), for 1 ≤
p, q ≤ ℓ, define an action of the Lie algebra sp(k) on Fℓ , which lifts to an
action of Sp(k).
176 5. Howe duality

(3) Sp(k) and c∞ form a Howe dual pair on Fℓ .

Proof. (1) This follows from Lemma 5.43(1), (5.45), and the definition of c∞ .
(2) Let us denote a standard basis for V = Cℓ ⊕ Cℓ∗ by v±,p , for p = 1, . . . , ℓ,
and denote a standard basis for C∞ by wr with r ∈ (− 21 − Z+ ). Letting U = V ⊗C∞ ,
we naturally identify Fℓ with ∧(U) by setting ψ±,p r ≡ v±,p ⊗ wr . Note that V is a
symplectic space with a natural pairing, and so is U. Hence, we have a natural
action of sp(V ) = sp(k) on Fℓ . The formula (5.46) is simply a precise way of
writing down this action in terms of coordinates. Indeed, comparing these formulas
with (5.23) we can see easily that the matrices there with B = C = 0 correspond
pq pq
to the E∗ ’s, while the ones with A = C = 0 correspond to the Ẽ∗∗ ’s, and the ones
with A = B = 0 correspond to the Ẽ ’s. Alternatively, it can be verified by a direct
pq

computation that the operators in (5.46) generate sp(k). Note that the action of
sp(k) preserves the principal gradation of Fℓ and every graded subspace of Fℓ is
finite dimensional. Thus, the action of sp(k) lifts to that of Sp(k).
pq
(3) As part of the proof of Lemma 5.43, we see that E∗ commutes with the
action of c∞ . We will now check that [Ẽuv , c∞ ] = 0, and leave the similar verification
that [Ẽuv
∗∗ , c∞ ] = 0 to the reader. Indeed, we have that
[ uv ]
Ẽ ,Ei, j − (−1)i+ j E1− j,1−i
[ ℓ ]
+,p −,p i+ j +,p −,p
= ∑ (−1)−r− 2 :ψ−u −v
∑ −i j−
1
ψ
−r r :, (:ψ 1 ψ 1 : − (−1) :ψ j− 1 ψ 1
−i
:)
2 2 2 2
r∈ 12 +Z p=1
[
= (−1)i ψ−v ψ−u1 + (−1)i ψ−u ψ−v1 + (−1) j ψ−u ψ−v + (−1) j ψ−v ψ−u ,
2 −i i− 2 2 −i i− 2 2−j 2−j
1 1 1
j− 12 1
j− 12
]
−,u +,v −,v i+ j +,u −,u i+ j +,v −,v
ψ+,u
1 ψ
−i j− 1 +ψ1 ψ
−i j− 1 − (−1) ψ j− 1 ψ1
−i
− (−1) ψ j− 1 ψ1
−i
2 2 2 2 2 2 2 2

=(−1)i ψ−v
1 ψ−u + (−1)i ψ−u
−i j− 1 1 ψ−v − (−1)i ψ−u
−i j− 1 1 ψ−v − (−1)i ψ−v
−i j− 1 1 ψ−u = 0.
−i j− 1
2 2 2 2 2 2 2 2

Hence the actions of Sp(k) and c∞ on Fℓ commute.


The Howe dual pair claim follows by the same type of argument as the one
given in the proof of Lemma 5.43(3) using now the First Fundamental Theorem of
invariant theory for Sp(k) (see Theorem 4.19). 

Theorem 5.47 ((Sp(k), c∞ )-Howe duality). As an (Sp(k), c∞ )-module, we have



Fℓ ∼
= L(Sp(k), λ) ⊗ L(c∞ , Λc (λ)),
λ∈P(ℓ)

where Λc (λ) := ℓΛc0 + ∑i≥1 λ′i εi = Λcλ1 + . . . + Λcλℓ . The joint highest weight vector
in Fℓ associated to λ with respect to the standard Borel for sp(k) × c∞ is vcλ :=
Ξ+,1 +,2 +,ℓ
λ1 Ξλ2 · · · Ξλℓ |0⟩.
5.4. Howe duality for infinite-dimensional Lie algebras 177

Proof. We observe that vcλ has sp(k)-weight λ, and c∞ -weight ℓΛc0 + ∑i≥1 λ′i εi , the
pq
latter of which by (5.39) equals Λcλ1 + . . . + Λcλℓ . Also Ẽ∗∗ annihilates vcλ , since in
both expressions only ψ+,p
pq
∗ ’s are involved and Ẽ∗∗ |0⟩ = 0. Since vλ is known to
c

be a joint (gl(ℓ), a∞ )-highest weight vector by Proposition 5.45, we see that vcλ is a
joint (sp(k), c∞ )-highest weight vector.
By Lemma 5.46 and Proposition 5.12, Fℓ is strongly multiplicity-free as an
(Sp(k), c∞ )-module. Since vcλ is a non-zero vector in Fℓ of sp(k)-weight λ, for every
λ ∈ P(ℓ), we also observe that all finite-dimensional Sp(k)-module appears in the
decomposition of Fℓ . So the multiplicity-free (Sp(k), c∞ )-module decomposition
of Fℓ follows. 
pp
The character of the Sp(k)-module Fℓ is the trace of the operator ∏ℓp=1 zEp∗ on
E −E
Fℓ , while that of the c∞ -module Fℓ is the trace of the operator ∏i∈N yi ii 1−i,1−i on
E −E pp
Fℓ . Computing the trace of ∏i∈N yi ii 1−i,1−i ∏ℓp=1 zEp∗ on both sides of the isomor-
phism in Theorem 5.47, we obtain the following character identity:
∞ ℓ
(5.47) p )= ∑
∏ ∏ (1 + yi z p )(1 + yi z−1 ch L(Sp(k), λ) ch L(c∞ , Λc (λ)).
i=1 p=1 λ∈P(ℓ)

5.4.5. (O(k), d∞ )-Howe duality. Let k = 2ℓ. It follows by (5.41) that


ℓ ( )
∑ (Ei, j − E1− j,1−i ) zi−1 w− j = ∑ :ψ+,p (z)ψ−,p (w): − :ψ+,p (w)ψ−,p (z): .
i, j∈Z p=1

Equivalently, we have the following operators on Fℓ :


ℓ ( )
−,p +,p −,p
(5.48) Ei, j − E1− j,1−i = ∑ :ψ+,p
1 ψ
−i j− 1 : − :ψ
j− 1 ψ1
−i
: .
2 2 2 2
p=1

Introduce the following generating functions


∑n∈Z e pq (n)z−n−1 = :ψ−,p (z)ψ−,q (z):,
∑n∈Z e∗∗ (n)z−n−1 = :ψ+,p (z)ψ+,q (z):.
pq

We will adopt the following short-hand notations:


Ě pq ≡ e pq (0), Ě∗∗
pq
≡ e∗∗
pq
(0), p, q = 1, . . . , ℓ.
It follows that
(5.49) Ě pq = ∑ :ψ−,p −,q
−r ψr :, Ě∗∗
pq
= ∑ :ψ+,p
−r ψr :.
+,q

r∈ 12 +Z r∈ 12 +Z
pq qp
Note that Ě pq = −Ěqp and Ě∗∗ = −Ě∗∗ .
Lemma 5.48. (1) The formula (5.48) defines an action of the Lie algebra d∞
on Fℓ of level ℓ.
178 5. Howe duality

pq pq
(2) The operators Ě pq , Ě∗∗ in (5.49) together with E∗ in (5.42), for 1 ≤ p, q ≤
ℓ, define an action of so(2ℓ) on Fℓ , which lifts to an action of O(2ℓ).
(3) O(2ℓ) and d∞ form a Howe dual pair on Fℓ .

Proof. The proof is completely analogous to that of Lemma 5.46 for type C. For
example, (2) can be easily obtained by comparing the formulas in (5.49) and (5.42)
with the one in (5.30), similar to the type C case. We leave the details to the
reader. 
1
Let k = 2ℓ + 1. We introduce the following operators on Fℓ+ 2 :
ℓ ( )
−,p +,p −,p
(5.50) Ei, j − E1− j,1−i = ∑ :ψ+,p
1 ψ
−i j− 1 : − :ψ
j− 1 ψ1
−i
: + :ϕ p1 −i ϕ pj− 1 :.
2 2 2 2 2 2
p=1

Equivalently, we have
∑ (Ei, j − E1− j,1−i ) zi−1 w− j
i, j∈Z
= ∑ℓp=1 (:ψ+,p (z)ψ−,p (w): − :ψ+,p (w)ψ−,p (z):) + :ϕ(z)ϕ(w):.

Introduce the following additional generating functions:


∑ e p (n)z−n−1 = :ψ−,p (z)ϕ(z):,
n∈Z

∑ e∗p (n)z−n−1 = :ψ+,p (z)ϕ(z):, p = 1, . . . , ℓ.


n∈Z

Introducing the short-hand notations Ě∗p = e∗p (0) and Ě p = e p (0), we have the fol-
lowing formulas:
(5.51) Ě p = ∑ :ψ−,p
−r ϕr :, Ě∗p = ∑ :ψ+,p
−r ϕr :.
r∈ 21 +Z r∈ 21 +Z

The following is a counterpart for k = 2ℓ + 1 of Lemma 5.48.


Lemma 5.49. (1) The formula (5.50) defines an action of the Lie algebra d∞
1
on F .ℓ+ 2

pq pq
(2) The operators Ě∗∗ , Ě pq in (5.49), Ě∗p , Ě p in (5.51), together with E∗ in
(5.42), for 1 ≤ p, q ≤ ℓ, define an action of the Lie algebra so(2ℓ + 1) on
1
Fℓ+ 2 , which lifts to an action of O(2ℓ + 1).
1
(3) O(2ℓ + 1) and d∞ form a Howe dual pair when acting on Fℓ+ 2 .

Proof. The proof is completely analogous to that of Lemma 5.46 in type C, and
will be omitted. 

Recall the partition parametrization of simple O(k)-modules from Proposi-


tion 5.36. The next theorem treats even and odd k uniformly.
5.5. Character formula for Lie superalgebras 179

Theorem 5.50 ((O(k), d∞ )-Howe duality). As an (O(k), d∞ )-module we have


⊕ ( )
F2 ∼
k
= L(O(k), λ) ⊗ L d∞ , Λd (λ) ,
λ∈P,λ′1 +λ′2 ≤k

where Λd (λ) := kΛd0 + ∑i≥1 λ′i εi .

Proof. By Lemmas 5.48 and 5.49, we have a strongly multiplicity-free decompo-


k
sition of the (O(k), d∞ )-module F 2 . The proof is completed by finding explicit
joint highest weight vectors with prescribed weights and noting that every simple
k
O(k)-module appears in the decomposition of F 2 , analogous to the proofs of The-
orems 5.44 and 5.47. We will refer the reader to [Wa, Theorem 3.2(2)] and [Wa,
Theorem 4.1(2)] for the precise form of these joint highest weight vectors for k
even and odd, respectively. 
E −E pp
Computing the trace of ∏i∈N yi ii 1−i,1−i ∏ℓp=1 zEp∗ on both sides of the identity
in Theorem 5.50, we obtain the following character identity (recall the Boolean
characteristic function (5.37)):
∞ ∞ ℓ
∏(1+yi )θ{k is odd} · ∏ ∏ (1 + yi z p )(1 + yi z−1
p )
i=1 i=1 p=1
(5.52)
= ∑ ch L(O(k), λ) ch L(d∞ , Λd (λ)).
λ∈P,λ′1 +λ′2 ≤k

5.5. Character formula for Lie superalgebras


In this section, we obtain character formulas for the irreducible oscillator mod-
ules of the Lie superalgebras of type osp constructed in Section 5.3, as a simple
application of the Howe dualities established in Section 5.4.

5.5.1. Characters for modules of Lie algebras c∞ and d∞ . For x ∈ {c, d}, let x∞0
be the Cartan subalgebra and W be the Weyl group of the Lie algebras x∞ defined
in Section 5.4. Let l be the Levi subalgebra of x∞ corresponding to the removal of
the simple root β× (as marked in the Dynkin diagrams in Section 5.4.1), and let
W0 denote the Weyl group of l. Let u+ and u− be respectively the nilradical and
opposite nilradical associated to l so that we have x∞ = u− ⊕ l ⊕ u+ . Let Wr0 be
the set of the minimal length representatives of the right cosets W0 \W of length r
for x∞ . It is well known that the Weyl group W can be written as W = W0W 0 with

W 0 = r≥0 Wr0 . For µ ∈ (x∞0 )∗ and w ∈ W we set
w ◦ µ := w(µ + ρx ) − ρx ,
where ρx ∈ (x∞0 )∗ is determined by ⟨ρx , Hix ⟩ = 1, for every simple coroot Hix .
Let λ be as in Theorem 5.47 in case of c∞ and as in Theorem 5.50 in case
of d∞ . Since ⟨Λx (λ), H xj ⟩ ∈ Z+ , for every j, it follows that ⟨w ◦ Λx (λ), Hix ⟩ ∈ Z+ ,
180 5. Howe duality

for w ∈ W 0 and Hix in the Levi subalgebra l. In our setting, noting that W0 is the
group of (finite) permutations of N, we may find explicitly a partition λw = λxw =
((λw )1 , (λw )2 , . . .) such that w ◦ Λx (λ) can be written as
{
kΛx + ∑ j>0 (λw ) j ε j , if x = d,
(5.53) w ◦ Λx (λ) = k 0x
2 Λ0 + ∑ j>0 (λw ) j ε j , if x = c.
Recall that sµ denotes the corresponding Schur function associated to a partition µ.
Proposition 5.51. Let λ be as in Theorem 5.47 in case of c∞ and as in Theorem 5.50
in case of d∞ . We have the following character formula:
1 ∞
ch L(x∞ , Λx (λ)) = ∑ (−1)r ∑ 0 sλw (y1 , y2 , . . .),
Dx r=0
for x = c, d,
w∈W r

where we have set yi = e−εi


for i ≥ 1, and
{
∏1≤i≤ j (1 − yi y j ), for x = c,
Dx :=
∏1≤i< j (1 − yi y j ), for x = d.

Proof. For any such λ, set Λ = Λx (λ). The module L(x∞ , Λ) is integrable and hence
affords the Weyl-Kac character formula
eσ(Λ+ρx )−ρx
(5.54) ch L(x∞ , Λ) = ∑ (−1)ℓ(σ) ∏α∈Φ+ (1 − e−α )
,
σ∈W

where Φ+ denotes the set of positive roots. Let ρl ∈ (x∞0 )∗ be the element deter-
mined by ρl (H j ) = 1, for all j ∈ N, and ρl (H0x ) = 0. Set ρu := ρx − ρl . Then for
τ ∈ W0 we have τ(ρu ) = ρu . Let Φ+ = Φ+ l ⊔ Φu , where Φl and Φu are the subsets
+ + +

of positive roots lying in l and u, respectively. We now rewrite (5.54) as follows:


eτw(Λ+ρx )−ρx
ch L(x∞ , Λ) = ∑ ∑ (−1)ℓ(τ) (−1)ℓ(w) ∏α∈Φ + (1 − e
−α )
w∈W 0 τ∈W0
(−1)ℓ(w) τw(Λ+ρx )−τ(ρx )+τ(ρx )−ρx
ℓ(τ) e
= ∑ ∑ Dx
(−1)
∏α∈Φ+l (1 − e−α )
w∈W 0 τ∈W0
( )
ℓ(w) τ w(Λ+ρx )−ρx +τ(ρl )−ρl
(−1) e
= ∑ ∑ (−1)ℓ(τ)
w∈W 0 τ∈W0
D x ∏α∈Φ+l (1 − e−α )
(−1)ℓ(w) eτ(w◦Λ+ρl )−ρl
= ∑ D x ∑ (−1)ℓ(τ)
∏α∈Φ+ (1 − e−α )
w∈W 0 τ∈W0 l

1 ∞
= x ∑ (−1)r ∑ sλw (y1 , y2 , . . .).
D r=0 w∈W 0 r

In the last identity we have used the fact that l is of type A+∞ , and so the irreducible
l-character of integrable highest weight w ◦ Λ is simply the Schur function sλw . 
5.5. Character formula for Lie superalgebras 181

5.5.2. Characters for oscillator osp(2m|2n)-modules. Recall hsλ denotes the su-
per Schur function defined in Appendix (A.35). Below, by the character of a mod-
ule of Sp(k) or of osp(2m|2n), we mean the trace of the operator ∏1≤p≤ℓ zEp −E
pp k+1−p.k+1−p

E E
or ∏1≤i≤m ∏1≤ j≤n xi īī y j j j , respectively. We have the following character formula
for the irreducible oscillator osp(2m|2n)-modules appearing in the Howe duality
decomposition in Theorem 5.31.

Theorem 5.52. Let x = {x1 , . . . , xm } and y = {y1 , . . . , yn }. For λ ∈ P(ℓ) with


λm+1 ≤ n, we have the following character formula:
ch L(osp(2m|2n), λ♮ + ℓ1m|n )
( ) ∏1≤i≤m (1 + xi ys ) · ∑∞
r=0 (−1) ∑w∈Wr0 hsλw (y; x)
r
x1 · · · xm ℓ
= · 1≤s≤n .
y1 · · · yn ∏1≤i< j≤m ∏1≤s≤t≤n (1 − xi x j )(1 − ys yt )

E E
Proof. Computing the trace of the operator ∏ℓp=1 zEp −E
pp k+1−p.k+1−p
∏m n
i=1 ∏ j=1 xi y j
īī jj

on both sides of the isomorphism in Theorem 5.31, we obtain that


( )
x1 · · · xm ℓ ℓ m n (1 + y j z−1 p )(1 + y j z p )
∏ ∏ ∏ −1
y1 · · · yn p=1 i=1 j=1 (1 − xi z p )(1 − xi z p )
=
(5.55)
∑ ch L(Sp(2ℓ), λ) ch L(osp(2m|2n), λ♮ + ℓ1m|n ).
λ∈P(ℓ)
λm+1 ≤n

Next, by formal algebraic manipulations starting from (5.47), we shall obtain


a new identity with the same left-hand side as (5.55). Replacing ch L(c∞ , Λc (λ)) in
(5.47) by the expression in Proposition 5.51, we obtain an identity of symmetric
functions in variables y1 , y2 , . . .. We replace yn+i by xi , for all i ≥ 1, and then apply
to both sides of the new identity the standard involution ωx (see (A.7)) on the ring of
symmetric functions in the variables x1 , x2 , . . .. In the process we use the identities
(A.22) and (A.7):
( )
ωx ∏ (1 − xi x j )−1 = ∏ (1 − xi x j )−1 ,
1≤i≤ j 1≤i< j
( )
ωx ∏(1 − ys xi )−1 = ∏(1 + ys xi ).
i≥1 i≥1

Finally, on the resulting identity we set x j = 0 for j ≥ m + 1 and multiply both sides
( )ℓ
by xy11···x
···yn
m
. In this way, we have obtained the following identity which shares the
same left-hand side as (5.55):
( )
x1 · · · xm ℓ ℓ m n (1 + y j z−1 p )(1 + y j z p )
∏ ∏ ∏ = ∑ ch L(Sp(2ℓ), λ)×
y1 · · · yn p=1 i=1 j=1 (1 − xi z−1
p )(1 − xi z p ) λ∈P(ℓ)
λm+1 ≤n
182 5. Howe duality

( )ℓ
x1 ···xm
y1 ···yn ∏1≤i≤m (1 + xi ys ) ∑∞
r=0 (−1) ∑w∈Wr0 hsλw (y; x)
r
1≤s≤n
.
∏1≤i< j≤m ∏1≤s≤t≤n (1 − xi x j )(1 − ys yt )
The theorem follows by comparing this identity with (5.55) and noting the linear
independence of the characters ch L(Sp(2ℓ), λ). 

5.5.3. Characters for oscillator spo(2m|2n)-modules. The goal of this subsec-


tion is to find a character formula for the irreducible oscillator spo(2m|2n)-modules,
which appear in the (O(k), spo(2m|2n))-Howe duality decomposition (Theorem 5.39).
First, let k = 2ℓ + 1 be odd and let λ be a partition with λ′1 + λ′2 ≤ k. Following
the notations of Section 5.3.3 we denote by eλ the partition obtained from λ by
replacing the first column by k − λ′1 . Since the restrictions to SO(k) of the modules
L(O(k), λ) and L(O(k), eλ) are isomorphic, the usual character will not distinguish
them. Recall −I ∈ O(k)\SO(k). So we define the (enhanced) character ch M of an
O(k)-module M to be the trace of the operator ε−I ∏ℓp=1 zEp −E
pp k+1−p,k+1−p
, where ε
is an additional formal variable such that ε = 1. The character of an spo(2m|2n)-
2
Eīī E j j
module is defined as usual to be the trace of the operator ∏m n
i=1 ∏ j=1 xi y j .
Theorem 5.53. Let k = 2ℓ + 1. Let x = {x1 , . . . , xm } and y = {y1 , . . . , yn }. For
λ ∈ P(m|n) with λ′1 + λ′2 ≤ k, we have the following character formula:
( k )
ch L spo(2m|2n), λ♮ + 1m|n =
2
( ) 2k ∏1≤i≤m (1 + xi ys ) · ∑∞
r=0 (−1) ∑w∈Wr0 hsλw (y; x)
r
x1 · · · xm
· 1≤s≤n .
y1 · · · yn ∏1≤i≤ j≤m ∏1≤s<t≤n (1 − xi x j )(1 − ys yt )
Proof. From the formula of the highest weight vector of L(O(k), λ) in the proof of
Theorem 5.39, we observe that
{
ε|λ| ch L(so(k), λ), if ℓ(λ) ≤ ℓ,
ch L(O(k), λ) = |λ|
ε ch L(so(k), eλ), if ℓ(λ) > ℓ.
e
Since ε|λ| = ε|λ|+1 , we conclude that the (enhanced) characters ch L(O(k), λ), where
λ′1 + λ′2 ≤ k, are linear independent.
E E
Computing the trace of the operator ε−I ∏ℓp=1 zEp −E
pp k+1−p,k+1−p
∏i, j xi īī y j on
jj

both sides of the isomorphism in Theorem 5.39, we obtain the following character
identity:
( )k
x1 · · · xm 2 ℓ m n (1 + εy j z−1 p )(1 + εy j z p )(1 + εy j )
y1 · · · yn ∏ ∏ ∏ −1
p=1 i=1 j=1 (1 − εxi z p )(1 − εxi z p )(1 − εxi )
(5.56) ( k )
= ∑ ch L(O(k), λ) ch L spo(2m|2n), λ♮ + 1m|n .
λ′ +λ′ ≤k
2
1 2
λm+1 ≤n
5.5. Character formula for Lie superalgebras 183

Analogous to the proof of Theorem 5.52, by means of algebraic manipulations


starting from (5.52), we shall obtain a new identity with the same left-hand side as
(5.56). Replacing ch L(d∞ , Λd (λ)) in (5.52) by the expression in Proposition 5.51
and taking into account the eigenvalue of the element −I, we obtain an identity
of symmetric functions in the variables y1 , y2 , . . .. We replace yn+i by xi , for all
i ≥ 1, and then apply the involution ωx on the ring of symmetric functions in the
variables x1 , x2 , . . .. Finally, we set x j = 0 for j ≥ m + 1 and multiply both sides of
( ) 2k
the resulting identity by xy11···x m
···yn . In this way we obtain

( ) 2k
x1 · · · xm ℓ (1 + εy j z−1
p )(1 + εy j z p )(1 + εy j )
y1 · · · yn ∏ ∏ (1 − εx z−1
p )(1 − εx z
p=1 i, j i i p )(1 − εxi )

= ∑ ch L(O(k), λ)×
λ′1 +λ′2 ≤k
λm+1 ≤n
( ) 2k ∏1≤i≤m (1 + xi ys ) ∑∞
r=0 (−1) ∑w∈Wr0 hsλw (y; x)
r
x1 · · · xm 1≤s≤n
.
y1 · · · yn ∏1≤i≤ j≤m ∏1≤s<t≤n (1 − xi x j )(1 − ys yt )

The theorem follows by comparing this identity with (5.56) and noting that the
(enhanced) characters ch L(O(k), λ) are linearly independent. 

We now turn to the case when k = 2ℓ is even. In this case, the trick of intro-
ducing the extra variable ε does not work, and we obtain the following.

Theorem 5.54. Let k = 2ℓ. Let x = {x1 , . . . , xm } and y = {y1 , . . . , yn }. For λ ∈


P(m|n) with λ′1 + λ′2 ≤ k, we have the following character identity:

ch L(spo(2m|2n), λ♮ + ℓ1m|n ) + ch L(spo(2m|2n), eλ♮ + ℓ1m|n ) =


( )
( ) ∏1≤i≤m (1 + xi ys ) · ∑∞
r=0 (−1) ∑w∈Wr0 hsλw (y; x) + hseλw (y; x)
r
x1 · · · xm ℓ
· 1≤s≤n .
y1 · · · yn ∏1≤i≤ j≤m ∏1≤s<t≤n (1 − xi x j )(1 − ys yt )

Under the additional assumption that λℓ = 0, eλ = λ, and the theorem becomes

ch L(spo(2m|2n), λ♮ + ℓ1m|n ) =
( ) ∏1≤i≤m (1 + xi ys ) · ∑∞
r=0 (−1) ∑w∈Wr0 hsλw (y; x)
r
x1 · · · xm ℓ 1≤s≤n
· .
y1 · · · yn ∏1≤i≤ j≤m ∏1≤s<t≤n (1 − xi x j )(1 − ys yt )
184 5. Howe duality

E E
Proof. Computing the trace of the operator zEp −E
pp k+1−p,k+1−p
∏i, j xi īī y j on both
jj

sides of the isomorphism in Theorem 5.39, we obtain


( )
x1 · · · xm ℓ ℓ m n (1 + y j z−1 p )(1 + y j z p )
∏ ∏ ∏ −1
y1 · · · yn p=1 i=1 j=1 (1 − xi z p )(1 − xi z p )
(5.57)
= ∑ ch L(O(k), λ) ch L(spo(2m|2n), λ♮ + ℓ1m|n ).
λ′1 +λ′2 ≤k
λm+1 ≤n

In the same way as in the proof of Theorem 5.53 by a formal algebraic ma-
nipulation based on (5.52), we obtain the following identity that shares the same
left-hand side as (5.57):
( )
x1 · · · xm ℓ ℓ m n (1 + y j z−1 p )(1 + y j z p )
∏ ∏ ∏ −1
= ∑ ch L(O(k), λ)×
y1 · · · yn p=1 i=1 j=1 (1 − xi z p )(1 − xi z p ) λ′ +λ ′ ≤k
1 2
λm+1 ≤n
( ∞
) ∏1≤i≤m (1 + xi ys ) · ∑r=0 (−1)r ∑w∈W 0 hsλw (y; x)
x1 · · · xm ℓ 1≤s≤n
r
· .
y1 · · · yn ∏1≤i≤ j≤m ∏1≤s<t≤n (1 − xi x j )(1 − ys yt )
The theorem follows by comparing this identity with (5.57) and noting that the
characters of L(O(k), λ) and L(O(k), eλ) coincide. 

5.6. Notes
Section 5.1. The materials on Weyl-Clifford algebra and connections to classi-
cal Lie algebras/superalgebras are standard. The general duality theorem, The-
orem 5.8, is taken from Goodman-Wallach [GW, 4.2.1], and it can be regarded
as an abstract generalization of the original formulation (see Theorem 5.10 and
Proposition 5.12) of Howe duality [H1] (Howe’s paper as a preprint was dated in
1976).
Section 5.2. The (GL(k), gl(m))-Howe duality (respectively, its skew version)
in Theorems 5.16 and 5.18 (due to Howe [H1, H2]) offers a representation theoretic
interpretation of the classical Cauchy identity (respectively, its dual version). It is
equivalent to Schur duality as well as the First Fundamental Theorem of invariant
theory for general linear groups.
The Howe dual pair (GL(k), gl(m|n)) (Theorem 5.14) already appeared in
Howe’s classical paper [H1], and the precise multiplicity-free (GL(k), gl(m|n))-
Howe duality decomposition (Theorem 5.19) was obtained later independently by
Sergeev [Sv2] and Cheng-Wang [CW1]. The formula for the joint highest weight
vectors in the (GL(k), gl(m|n))-Howe duality (Theorem 5.23) was due to Cheng-
Wang [CW1], and it unifies the formulas of Howe [H2] in two (non-super) special
cases when m or n is zero; a formula of highest weight vectors for the more general
5.6. Notes 185

(gl(k|ℓ), gl(m|n))-Howe duality was also obtained in loc. cit.. The (q(m), q(n))-
Howe duality was due to Sergeev [Sv2] and Cheng-Wang [CW2] independently,
and it gives a representation theoretic interpretation of the Cauchy identity for
Schur Q-functions.
Section 5.3. The Howe dual pairs (Sp(k), osp(2m|2n)) and (O(k), spo(2m|2n))
on S(Ck ⊗ Cm|n ) (Theorems 5.30 and 5.38) were formulated in Howe [H1]; they
were based on, and in turn can be regarded as a Lie theoretic reformulation of, the
First Fundamental Theorem of invariant theory for classical groups. The orthogo-
nal group O(k) is disconnected. We follow Howe [H2] to present a parametrization
of the simple O(k)-modules in terms of partitions. The strongly multiplicity-free
decompositions in (Sp(k), osp(2m|2n))- and (O(k), spo(2m|2n))-Howe dualities
(Theorems 5.31 and 5.39) were obtained in Cheng-Zhang [CZ2], where the high-
est weight vector formula in Section 5.2 plays a key role in the proofs. The short
proofs presented in this book follow Cheng-Kwon-Wang [CKW]. We easily re-
cover the Howe duality decompositions in the non-super setting, which were due
to Howe [H2] with somewhat different arguments, by setting m or n to zero.
There are additional Howe dualities involving the spin groups (see Howe [H2])
as well as spo(2m|2n + 1) (see Cheng-Kwon-Wang [CKW, Appendix A]). There
is also a type A Howe duality involving infinite-dimensional irreducible modules of
gl(n) or more generally gl(m|n) (see Kashiwara-Vergne [KaV] and Cheng-Lam-
Zhang [CLZ]). We do not treat these cases in the book and refer the reader to the
original papers for details.
Section 5.4. Howe duality between classical groups and infinite-dimensional
Lie algebras was systematically developed by Wang [Wa], and it has been used
in Kac-Wang-Yan [KWY] for the study of representation theory of classical Lie
subalgebras of W1+∞ . The (GL(ℓ), a∞ )-Howe duality (Theorem 5.44) can also be
recovered as a limit case of a duality for affine Lie algebras of type gl given earlier
by I. Frenkel [Fr], and the proof here follows [Wa]. The (Sp(k), c∞ )-Howe duality
and (O(k), d∞ )-Howe duality (Theorems 5.47 and 5.50) are due to Wang [Wa],
where one can find several more Howe dualities in bosonic and fermionic Fock
spaces not covered in the book.
Section 5.5. Irreducible characters for the oscillator modules of Lie superalge-
bras were computed in Cheng-Zhang [CZ2] following the approach of Cheng-Lam
[CL1]. A character formula of Enright [En] played an essential role there. The
simpler and more elementary approach presented in the book bypasses Enright’s
formula, via a comparison with Howe duality involving infinite-dimensional Lie
algebras in Section 5.4, and follows Cheng-Kwon-Wang [CKW].
The reader is referred to Cheng-Kwon-Wang [CKW], where the calculation of
the characters via a comparison of two Howe dualities has been refined to compute
the corresponding u− -homology groups with coefficients in the oscillator modules.
The approach there follows the strategy of Aribaud [Ar] and Cheng-Kwon [CK].
Chapter 6

Super duality

In this chapter, we develop a super duality approach to obtain a complete and con-
ceptual solution of the irreducible character problem in certain parabolic Bernstein-
Gelfand-Gelfand categories for general linear and ortho-symplectic Lie superal-
gebras. These parabolic categories contain all the finite-dimensional irreducible
modules of these Lie superalgebras.
Super duality is an equivalence of categories between parabolic categories for
Lie superalgebras and their Lie algebra counterparts at an infinite-rank limit. A
weak version of such a category equivalence on the Grothendieck group level,
which is established first in an elementary way, already implies a solution of the
irreducible character problem of these Lie superalgebras in terms of Kazhdan-
Lusztig polynomials of classical types. It is further shown that the corresponding
Kostant u-homology groups, or equivalently the Kazhdan-Lusztig-Vogan polyno-
mials, are matched perfectly under super duality. Via the truncation functors which
we introduce, the Kazhdan-Lusztig solution for the irreducible character problem
at infinite-rank limit provides us a complete solution to the irreducible character
problem in the corresponding parabolic BGG categories for finite-dimensional ba-
sic Lie superalgebras.

6.1. Lie superalgebras of classical types


This section starts with a discussion of the infinite-rank Lie superalgebras g, g and
e
g, whose representation theories we will investigate in later sections. Here, g is a
classical Lie algebra, g is a basic Lie superalgebra, and e
g is an auxiliary Lie super-
algebra which contains both g and g as subalgebras and which plays the role of an

187
188 6. Super duality

intermediary between them. We present explicit matrix realizations of these Lie su-
peralgebras together with their finite-dimensional counterparts, e
gxn , gxn , and gxn , for
n ∈ N. All these superalgebras implicitly depend on a fixed type x = a, b, b• , c, d.

6.1.1. Head, tail, and master diagrams. For m ∈ Z+ , consider a vector space
with basis {ε−m , . . . , ε−1 } ∪ {εr |r ∈ 12 N}, and a symmetric bilinear form (·|·) given
by
1
(εr |εs ) = (−1)2r δrs ,r, s ∈ {−m, . . . , −1} ∪ N.
2
We introduce the following notation for roots which we shall need shortly:
α× := ε−1 − ε1/2 , α j := ε j − ε j+1 , −m ≤ j ≤ −2,
(6.1) 1
β× := ε−1 − ε1 , αr := εr − εr+1/2 , βr := εr − εr+1 , r ∈ N.
2

We
 define
the Tn
tail diagrams to be the following three Dynkin diagrams
,
Tn
, and
e with prescribed fundamental systems denoted by Π(Tn ), Π(Tn ),
Tn

e n ), for n ∈ N:
and Π(T

Tn ⃝ ⃝ ⃝ ··· ⃝ ⃝

β× β1 β2 βn−2 βn−1
 ⊗
Tn ⃝ ⃝ ··· ⃝ ⃝

α× β1/2 β3/2 βn−5/2 βn−3/2
 ⊗ ⊗ ⊗ ⊗ ⊗
en
T ···

α× α1/2 α1 αn−1 αn−1/2

Corresponding to these three diagrams we have the Lie superalgebras gl(1 + n),
gl(1|n), and gl(1 + n|n). For n = ∞, we have analogous diagrams that are the
Dynkin diagrams of the corresponding infinite-rank Lie superalgebras.
 
We shall choose another Dynkin diagram  k  , called a head diagram, to
connect with one of the three tail diagrams to produce the following three new
Dynkin diagrams, which will be called the master diagrams (n ∈ N ∪ {∞}):
     
Tn en
(6.2)
k

k
Tn
k
T
  
The three master diagrams will be denoted by
gn
,
gn
, and
e
gn
respectively,
and the associated Lie superalgebras gn , gn , and e
gn will be introduced subsequently.
We denote the fundamental systems corresponding to the three master diagrams by
(6.3) Πn := Π(k) ⊔ Π(Tn ), Πn := Π(k) ⊔ Π(Tn ), Π e n ).
e n := Π(k) ⊔ Π(T

When n = ∞, we shall make it a convention to drop the subscript n, and


denote
  these
master diagrams, fundamental systems, and Lie superalgebras by g , g , and



e

e
g , Π(T), Π(T), and Π(T),g, g, and e
g, respectively.
6.1. Lie superalgebras of classical types 189


The head diagram

k used in this book is always chosen to be one of the
Dynkin diagrams
kx x  we will add the superscript x to
defined below. Accordingly,
x
the general notations in 6.1.1 to write Πn ,
gn , gxn , and so on, when it is needed to

b,x b, c, d and m ≥ 1, introduce the Lie (super)algebras
specify the type x. For x = a,
x
k with Dynkin diagrams
k and prescribed fundamental systems denoted by
Π(k ) as follows (see (6.1) for notation of α j ):
x


ka ⃝ ⃝ ⃝ ··· ⃝ ⃝ ⃝

α−m α−m+1 α−3 α−2

kb ⃝⇐=⃝ ⃝ ··· ⃝ ⃝ ⃝

−ε−m α−m α−3 α−2
•  y⇐=⃝
kb ⃝ ··· ⃝ ⃝ ⃝

−ε−m α−m α−3 α−2

kc ⃝=⇒⃝ ⃝ ··· ⃝ ⃝ ⃝

−2ε−m α−m α−3 α−2

α−m ⃝
 @
@
kd ⃝ ⃝ ··· ⃝ ⃝ ⃝

α−m+1 α−3 α−2

−ε−m −ε−m+1

We have the following identifications of Lie algebras: ka = gl(m), kb = so(2m + 1),



kb = osp(1|2m), kc = sp(2m), and kd = so(2m). The Lie superalgebra osp(1|2m)
behaves like the finite-dimensional semisimple Lie algebras, as every finite-dimensional
osp(1|2m)-module is completely reducible by Corollary 2.30.

6.1.2. The index sets. Let us introduce some notations and conventions for in-
dex sets, which will be needed for defining the Lie superalgebras gx , gx , and e
gx
associated to the master diagrams (6.2).
For m ∈ Z+ , we introduce the following totally ordered set eIm :
3 1 1 3
··· < < 1 < < −1
| < ·{z
· · < −m} < 0 < −m
| · · · < −1} < < 1 < < · · ·
< {z
2 2 2 2
m m

Further introduce the following subsets of eIm :


{ }
Im := −1, . . . , −m, 0, −m, . . . , −1 ∪ {1, 2, 3, . . .} ∪ {1, 2, 3, . . .},
| {z } | {z }
m m
{ } {1 3 5 } {1 3 5 }
Im := −1, . . . , −m, 0, −m, . . . , −1 ∪ , , ,... ∪ , , ,... ,
| {z } | {z } 2 2 2 2 2 2
m m
{ 1 3 }
eI+ := − m, . . . , −1, , 1, , 2, . . . .
m
2 2
190 6. Super duality

For X = Im , Im , or eI+
m , define

X× := X \ {0}, X+ := X ∩ eI+
m.

6.1.3. Infinite-rank Lie superalgebras. We shall provide explicit matrix realiza-


gx , for x = a, b, b• , c, d. Actually
tions of infinite-rank Lie superalgebras gx , gx , and e
gx for x = a, b, c, d are Lie algebras.
Lie superalgebras of type a. For m ∈ Z+ , let Vem be the infinite-dimensional
superspace over C with ordered basis {vi |i ∈ eIm }, whose Z2 -grading is specified as
follows:
1
|vr | = |vr | = 0̄ (r ∈ Z \ {0}), |vs | = |vs | = 1̄ (s ∈ + Z+ ).
2
The parity of the vector v0 is to be specified. With respect to this basis, a linear
map on Vem may be identified with a complex matrix (ars )r,s∈eIm . Let gl(Vem ) denote
the Lie superalgebra consisting of (ars )r,s∈eIm with ars = 0 for all but finitely many
ars ’s. Denote as usual by Ers ∈ gl(Vem ) the elementary matrix with 1 at the rth row
and sth column and zero elsewhere.
The vector spaces Vm and V m are defined to be the subspaces of Vem with ordered
basis {vi } indexed by Im and Im , respectively. The subspaces of Vm , V m , and Vem
× e× ×
with basis vectors vi , with i indexed by I× ×
m , Im , and Im , are denoted by Vm , V m and
Vem× , respectively. Similarly, the subspaces with basis vectors vi , for i indexed by
I+
+ e+ + e+
m , Im and Im , are denoted by Vm , V m and Vm , respectively. We summarize these
+

vector superspaces together with the index sets for their bases in Table 1 below:

Table 1

×
Superspaces Vm V m Vem Vm× V m Vem× Vm+ V m Vem+
+
×
Index sets Im Im eIm Im eI× Im eI+
+

m m I+
m m

×
Let W be one of the superspaces Vem , Vem× , Vem+ ,Vm ,Vm× ,Vm+ ,V m ,V m or V m , regarded
+

as subspaces of Vem . They give rise to Lie superalgebras gl(W ) as subalgebras of


gl(Vem ). The standard Cartan subalgebra of gl(W ) is spanned by the basis {Err },
with corresponding dual basis {εr }, where r runs over the index set corresponding
to W . The standard Borel subalgebra of gl(W ) is spanned by Ers , with r ≤ s.
The fundamental systems corresponding to the standard Borel subalgebras of
the Lie superalgebras gl(Vem+ ), gl(Vm+ ), and gl(V m ) are precisely Π
+ e a , Πa , and Πa , re-

spectively,
 and 
the corresponding Dynkin diagrams are the master diagrams
e
ga ,
g a
g
and
a e + + +
, respectively. Therefore, gl(Vm ), gl(Vm ), and gl(V m ) are matrix


realizations of the Lie superalgebras e ga , ga , and ga , respectively. We summarize
this in the Table 2 below.
6.1. Lie superalgebras of classical types 191

Table 2. Matrix forms for Lie superalgebras of type a

Lie superalgebras ega ga ga


gl(Vem+ ) gl(Vm+ ) gl(V m )
+
Matrix forms

a e a we have the following pos-


Associated to the fundamental systems Πa , Π and Π
itive systems:
a e+
Φ = {εr − εs |r < s (r, s ∈ eI+
m )},
a
Φ+ = {εi − ε j |i < j (i, j ∈ I+
m )},
+ +
a
Φ = {εr − εs |r < s (r, s ∈ Im )}.

Lie superalgebras of types b• , c. We set |v0 | = 1̄. For m ∈ Z+ , define a non-


degenerate skew-supersymmetric bilinear form (·|·) on the superspace Vem by
(vr |vs ) = (vr |vs ) = 0, (vr |vs ) = δrs = −(−1)|vr |·|vs | (vs |vr ), r, s ∈ eI+
m,
(6.4)
(v0 |v0 ) = 1, (v0 |vr ) = (v0 |vr ) = 0, r ∈ eI+
m.

We obtain non-degenerate skew-supersymmetric bilinear forms by restriction, which


×
are again denoted by (·|·), on the subspaces Vem× , Vm , Vm× , V m and V m .
×
Let W be one of the superspaces Vem , Vem× , Vm , Vm× , V m and V m , respectively.
Recall from Chapter 1 that the Lie superalgebra spo(W ) is the subalgebra of gl(W )
preserving the form defined in (6.4). That is, for ε ∈ Z2 ,
spo(W )ε = {T ∈ gl(W )ε | (T v|w) = −(−1)ε|v| (v|Tw), ∀v, w ∈ W }.
The standard Cartan subalgebra of spo(W ) has a basis {Er := Err − Er,r }, with
corresponding dual basis {εr }, where r runs over the index sets eI+ e+ + +
m , Im , Im , Im ,
+ +
Im , Im , respectively. The standard Borel subalgebra of spo(W ) is obtained by
taking the intersection of spo(W ) with the standard Borel subalgebra of gl(W ).
One checks in a straightforward fashion that the fundamental systems associated

e b• , Π
to these standard Borel subalgebras are precisely Π e c , Πb• , Πc , Πb , Πc , re-
spectively. Thus, we have obtained matrix realizations of the Lie superalgebras,
for m ≥ 1, in Table 3 below.

Table 3. Matrix forms for Lie superalgebras of types b• , c

• • •
Superalgebras e
gb gb gb e
gc gc gc
×
Matrix forms spo(Vem ) spo(Vm ) spo(V m ) spo(Vem× ) spo(Vm× ) spo(V m )
192 6. Super duality

x e x , for x = b• , c, we have the


Associated to the fundamental systems Πx , Π and Π
following positive systems:
Φ = {±εr − εs |r < s (r, s ∈ eI+
b• e + e+
m )} ∪ {−2εi (i ∈ Im )} ∪ {−εr (r ∈ Im )},
+

c e+
Φ = {±εr − εs |r < s (r, s ∈ eI+
m )} ∪ {−2εi (i ∈ Im )},
+

b•
Φ+ = {±εi − ε j |i < j (i, j ∈ I+
m )} ∪ {−εi , −2εi (i ∈ Im )},
+

c
Φ+ = {±εi − ε j |i < j (i, j ∈ I+
m )} ∪ {−2εi (i ∈ Im )},
+

b• + + +
Φ = {±εr − εs |r < s (r, s ∈ Im )} ∪ {−2εi (−m ≤ i ≤ −1)} ∪ {−εr (r ∈ Im )},
+ +
c
Φ = {±εr − εs |r < s (r, s ∈ Im )} ∪ {−2εi (−m ≤ i ≤ −1)}.

Lie superalgebras of types b, d. Now we set |v0 | = 0̄. Define a supersymmet-


ric bilinear form (·|·) on the superspace Vem by
(vr |vs ) = (vr |vs ) = 0, (vr |vs ) = δrs = (−1)|vr |·|vs | (vs |vr ), r, s ∈ eI+
m,
(6.5)
(v0 |v0 ) = 1, (v0 |vr ) = (v0 |vr ) = 0, r ∈ eI+
m.
By restriction, we obtain non-degenerate supersymmetric bilinear forms, which
×
will also be denoted by (·|·) on the subspaces Vem× , Vm , Vm× , V m and V m .
×
Let W be one of the spaces Vem , Vem× , Vm , Vm× , V m and V m , respectively. Re-
call from Chapter 1 that the Lie superalgebra osp(W ) is the subalgebra of gl(W )
preserving the form given by (6.5). Introduce a short-hand notation:
(6.6) Er := Err − Er̄r̄ .
The standard Cartan subalgebra of osp(W ) has the basis {Er }, with correspond-
ing dual basis {εr }, where r runs over the index sets eI+ e+ + + + +
m , Im , Im , Im , Im , Im , re-
spectively. As before, the standard Borel subalgebra of osp(W ) is obtained by
intersecting osp(W ) with the standard Borel subalgebra of gl(W ). One computes
that the associated fundamental systems are precisely Π e b, Π
e d , Πb , Πd , Πb , Πd ,
respectively. Thus, we have obtained the following matrix realizations of Lie su-
peralgebras, for m ≥ 1, in Table 4.

Table 4. Matrix forms for Lie superalgebras of types b, d

Superalgebras egb gb gb e
gd gd gd
×
Matrix forms osp(Vem ) osp(Vm ) osp(V m ) osp(Vem× ) osp(Vm× ) osp(V m )

e x , Πx , Πx , for x = b, d, we have the fol-


Associated to the fundamental systems Π
lowing positive systems:
Φ = {±εr − εs |r < s (r, s ∈ eI+
b e+ e+
+
m )} ∪ {−2εs (s ∈ I0 )} ∪ {−εr (r ∈ Im )},
6.1. Lie superalgebras of classical types 193

d e+
Φ = {±εr − εs |r < s (r, s ∈ eI+
+
m )} ∪ {−2εs (s ∈ I0 )},
b
Φ+ = {±εi − ε j |i < j (i, j ∈ I+
m )} ∪ {−εi (i ∈ Im )},
+

d
Φ+ = {±εi − ε j |i < j (i, j ∈ I+
m )},
+ + + +
b
Φ = {±εr − εs |r < s (r, s ∈ Im )} ∪ {−2εs (s ∈ I0 )} ∪ {−εr (r ∈ Im )},
+ + +
d
Φ = {±εr − εs |r < s (r, s ∈ Im )} ∪ {−2εs (s ∈ I0 )}.

e + , x Φ+ , and x Φ+ for
6.1.4. The case of m = 0. Let x = a, b, b• , c, d. The sets x Φ
m = 0 still make sense, and they are the positive systems for the Lie superalge-
bras e
gx , gx and gx whose Dynkin diagrams and fundamental systems are as follows.
Some of the Dynkin diagrams in the case of m = 0 differ somewhat from the coun-
terparts in the case when m ≥ 1, and this is why we treat the two cases separately.
For m = 0, the corresponding Lie superalgebras are realized similarly as in Sec-
tion 6.1.3, and we shall keep the same notations as before.
 ⊗ ⊗ ⊗ ⊗ ⊗ ⊗
e
ga ··· ···

α1/2 α1 α3/2 αr αr+1/2 αr+1

 ⊗ ⊗ ⊗ ⊗ ⊗
y
⇐= ··· ···
e
gb

−ε1/2 α1/2 α1 αr−1/2 αr αr+1/2

•  ⊗ ⊗ ⊗ ⊗ ⊗
e
gb ⃝⇐= ··· ···

−ε1/2 α1/2 α1 αr−1/2 αr αr+1/2


α1/2

 @ ⊗ ⊗ ⊗ ⊗ ⊗
@
e
gc ··· ···

⊗ α1 α3/2 αr−1/2 αr αr+1/2


−ε1/2 − ε1

 ⊗ ⊗ ⊗ ⊗ ⊗
e
gd ⃝=⇒ ··· ···

−2ε1/2 α1/2 α1 αr−1/2 αr αr+1/2


ga ⃝ ⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

β1 β2 β3 βn βn+1 βn+2


gb ⃝⇐=⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

−ε1 β1 β2 βn−1 βn βn+1

•  y
gb ⇐=⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

−ε1 β1 β2 βn−1 βn βn+1
194 6. Super duality


gc ⃝=⇒⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

−2ε1 β1 β2 βn−1 βn βn+1

β1 ⃝
 @
@
gd ⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

β2 β3 βn−1 βn βn+1
−ε1 − ε2 ⃝

ga ⃝ ⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

β1/2 β3/2 β5/2 βr βr+1 βr+2

 y
gb ⇐=⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

−ε1/2 β1/2 β3/2 βr−1 βr βr+1

• 
gb ⃝⇐=⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

−ε1/2 β1/2 β3/2 βr−1 βr βr+1

β1/2 ⃝
 @
@
gc ⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

β3/2 β5/2 βr−1 βr βr+1
−ε1/2 − ε3/2 ⃝

gd ⃝=⇒⃝ ⃝ ··· ⃝ ⃝ ⃝ ···

−2ε1/2 β1/2 β3/2 βr−1 βr βr+1

We have the identifications of Dynkin diagrams when m = 0:


  •     •    

=
ga
ga ,
gb ,
gc
=
gb =
gd
,
gb ,
gc
=
gb =
gd
.

6.1.5. Finite-dimensional Lie superalgebras. For m ∈ Z+ , let W be one of the


×
superspaces Vem , V m , Vm , Vem+ , V m , Vm+ , Vem× , V m , and Vm× . For n ∈ N, let Wn stand for
+

the subspace of W spanned the vectors vr ∈ W with n ≤ r ≤ n. We consider the Lie


superalgebras gl(Wn ), for W = Vem+ ,V m ,Vm+ , and the Lie superalgebras spo(Wn ) and
+
×
osp(Wn ), for W = Vem , Vem× ,Vm ,Vm× ,V m ,V m . As in the case of n = ∞, these provide
matrix realizations
x  xof Lie superalgebras
x  gxn , gxn , and e
gxn , whose associated Dynkin
diagrams
gn ,
gn , and
e
gn are given in (6.2); see Table 5 below. Note that gxn
are actually Lie algebras, for x = a, b, c, d.

6.1.6. Central extensions. We assume that x = b, b• , c, d in this subsection. We


shall replace the Lie superalgebras gx , gx and egx and their finite-dimensional ana-
logues by their central extensions, and study the representations of their central ex-
tensions instead. As these central extensions are trivial, we can easily recover the
representations of gx , gx and e
gx from those of their central extensions. The central
6.1. Lie superalgebras of classical types 195

Table 5. Identifications for finite-dimensional Lie (super)algebras

Type x gxn gxn e


gxn
a gl(m + n) gl(m|n) gl(m + n|n)
b so(2(m + n) + 1) osp(2m + 1|2n) osp(2(m + n) + 1|2n)
b• osp(1|2(m + n)) spo(2m|2n + 1) spo(2(m + n)|2n + 1)
c sp(2(m + n)) spo(2m|2n) spo(2(m + n)|2n)
d so(2(m + n)) osp(2m|2n) osp(2(m + n)|2n)

extensions will be more conceptual and convenient for formulations of truncation


functors and super duality in later sections. For x = a, the Lie superalgebras ga , ga ,
and e
ga are already suitable for super duality later on, and so their central extensions
will not be needed.
Let m ∈ Z+ . The supertrace Str defined in Section 1.1 makes sense for gl(Vem ).
For X,Y ∈ gl(Vem ), we have
(6.7) Str(XY ) = (−1)|X||Y | Str(Y X).
Let I := E00 + ∑r≤ 1 Err . This allows us to define τ : gl(Vem ) × gl(Vem ) → C by
2

τ(A, B) := Str([I, A]B) = Str(IAB − AIB), A, B ∈ gl(Vem ).


It is easy to show that τ is a (trivial) 2-cocycle, and hence defines a central extension
b Vem ) of gl(Vem ) by the one-dimensional center CK. That is, we have gl(
gl( b Vem ) =
gl(Vem ) ⊕ CK as a vector superspace, with Lie bracket [X, [] + τ(X,Y )K,
b Yb ] = [X,Y
b
where we have used Xb to denote the element in gl(Vem ) that corresponds to the
element X ∈ gl(Vem ). This central extension is trivial, since it is straightforward to
check that an isomorphism φ from Lie superalgebra gl( b Vem ) to the direct sum of Lie
e
superalgebras gl(Vm ) ⊕ CK is given by
(6.8) b = X − Str(IX)K,
φ(X) φ(K) = K.
×
Now for W being one of Vem× ,Vm ,Vm× ,V m ,V m , the restrictions of τ to the subal-
gebras gl(W ) give rise to respective central extensions, which in turn induce central
extensions on the ortho-symplectic subalgebras. The central extension of gx , gx and
e
gx arising this way will also be denoted by gx , gx and e gx , respectively, by abuse of
notation. We hope this will not cause confusion, as only these central extensions
will be used in the remainder of this chapter.
We note that gx and gx are naturally subalgebras of e gx . The standard Cartan
g will be denoted by h , h and e
x x
subalgebras of g , g and e
x x x x hx , respectively. hx , h
or e
hx has a basis {K, Ebr } with dual basis {Λ0 , εr }, where r runs over the index sets
Im , Im , or eI+
+
m , respectively. Here Λ0 is determined by
+

Λ0 (K) = 1, Λ0 (Ebr ) = 0,
196 6. Super duality

for all admissible r in each case. The C-span of Λ0 and the εr ’s in the dual of hx
(respectively, h and e hx ) will be denoted by hx∗ (respectively, h and e
x x∗
hx∗ ), where r
lies in I+
+ e+
m (respectively, Im and Im ).
In case x = a it will also be advantageous to set Ebr ≡ Err , for r > 0. For nota-
tional convenience later on, we shall declare Λa0 to be 0.
From now on, we shall adopt a convention of dropping the superscript x.
We
write g, g,
 and e
g for gx , gx , and e
gx , with associated Dynkin diagrams


g ,
g

e
g
, and

, respectively, where x denotes a fixed type among a, b, b , c, d.

6.2. The module categories


In this section, we define the parabolic Bernstein-Gelfand-Gelfand (BGG) cate-
gories of modules for the Lie superalgebras e g, g, and g, and also their finite-rank
counterparts. We introduce the truncation functors to relate the module categories
of the infinite-rank Lie superalgebras with their finite-rank counterparts. It is shown
that the truncation functors send irreducible and parabolic Verma modules to irre-
ducible and parabolic Verma modules accordingly, or to zero.

6.2.1. Category of polynomial modules revisited. Recall from 6.1.5 that Ve0,k +
is
the (k|k)-dimensional superspace spanned by vr , for r = 1/2, 1, 3/2, . . . , k − 1/2, k,
for k ∈ N, and Ve0+ = Ve0,∞
+
. We introduce the short-hand notations
el+ = gl(Ve + ), el+ = gl(Ve + ), k ∈ N.
0 k 0,k

In this subsection, which is an infinite-rank counterpart of 3.2.6 in Chapter 3, we


shall show that the category of polynomial el+ -modules is semisimple. This will be
needed in 6.2.2, where el+ appears as a direct summand of the Levi subalgebra for
e
g.
Let λ = (λ1 , λ2 , . . .) be a partition. For j ∈ N, we denote
(6.9) θ(λ) j = max{λ j − j, 0}, θ(λ) j−1/2 = max{λ′j − j + 1, 0}.
One recognizes (θ(λ)1/2 , θ(λ)3/2 , . . . | θ(λ)1 , θ(λ)2 , . . .) as the modified Frobenius
coordinates of λ′ of Example 2.53(3). Define
λθ := ∑ θ(λ)r εr .
r∈ 12 N

(For a definition of λθ in a general setting see (6.15) below.)


The standard Borel subalgebra of el+ k corresponds to the fundamental system
consisting of (all odd) simple roots εr − εr+1/2 , for r = 1/2, 1, 3/2, . . . , k − 1/2. By
Theorem 2.52, the irreducible polynomial representations of el+ ∼
k = gl(k|k) are pa-
rameterized by (k|k)-hook partitions, and that the set of highest weights of these
6.2. The module categories 197

modules is precisely {λθ | λ ∈ P(k|k)}. We have natural inclusions of Lie superal-


gebras:
el+ ⊂ el+ ⊂ · · · ⊂ el+ ⊂ el+ ⊂ · · · ,
1 2 k−1 k
compatible with their respective standard Borel subalgebras.

Let k ∈ N ∪ {∞}. Let V = µ∈h∗ Vµ be an h-semisimple el+ k -module, where
Vµ = 0 unless µ satisfies (−1) (µ, εr ) ∈ Z+ , for r ∈ 2 N, and (µ, εr ) = 0, for r ≫ 0.
2r 1

Let n ∈ N with n < k. We form a subspace of V ,



trknV := Vν ,
ν

where ν satisfies (ν, εr ) = 0, for r > n. (trkn as a functor will be defined and studied
systematically in a general setting in 6.2.5).
Let λ be a (k|k)-hook partition. The character of the irreducible el+ k -module
e + θ
L(lk , λ ) is given by hsλ (x1 , . . . , xk ; y1/2 , y3/2 , . . . , yk−1/2 ) according to Theorem 3.14,
where x j = eε j and y j−1/2 = eε j−1/2 for j ∈ N. Hence the character of the el+ n -module
( + θ )
e
trn L(lk , λ ) is equal to hsλ (x1 , . . . , xn ; y1/2 , y3/2 , . . . , yn−1/2 ) or zero, depending
k

on whether λ is an (n|n)-hook partition or not. Hence, for n < k, we have


{
( + θ )
e ∼ L(el+ θ
n , λ ), for a (n|n)-hook partition λ,
(6.10) trn L(lk , λ ) =
k
0, otherwise.

From now on, the natural inclusions L(el+ θ e+ θ


n , λ ) ⊆ L(lk , λ ) for any λ ∈ P(n|n) will
be understood in the sense of the above isomorphism, for n < k. One checks that
∪ e+ θ e+ θ
k≥n L(lk , λ ) is an irreducible highest weight l -module of highest weight λ .
Hence,

L(el+ , λθ ) = L(el+ θ
k , λ ).
k≥n

The L(el+ , λθ ), for λ ∈ P, has character given by the super Schur function hsλ .
( )
Lemma 6.1. Let λ, µ ∈ P. Then Exte1l+ L(el+ , λθ ), L(el+ , µθ ) = 0.

Proof. Consider a short exact sequence of el+ -modules


(6.11) 0 −→ L(el+ , λθ ) −→ E −→ L(el+ , µθ ) −→ 0.
First suppose that λ = µ. Then by weight consideration, the two-dimensional µθ -
weight subspace of E is a highest weight space, and thus E contains two proper
submodules and the short exact sequence must split.
Now suppose that λ ̸= µ. Choose n > max{|λ|, |µ|}. Applying tr∞ n to every term
in (6.11), we obtain by (6.10) the following short exact sequence of polynomial el+
n-
modules
(6.12) 0 −→ L(el+ θ ∞ e+ θ
n , λ ) −→ trn E −→ L(ln , µ ) −→ 0.
198 6. Super duality

By Theorem 3.25, (6.12) splits. Hence, we can find a singular vector w in the el+ n-
∞ e+ e + θ
module trn E so that U(ln )w = L(ln , µ ). Since there are no weight vectors in E
of weight µθ + εi − εi+1/2 , for i ≥ n, w is a singular vector in E with respect to the
standard Borel subalgebra of el+ too.
Now we consider the el+ -submodule L generated by w in E. We claim that
L∼ = L(el+ , µθ ), and so (6.11) splits. For if L is not isomorphic to L(el+ , µθ ), then L
contains L(el+ , λθ ), and in particular contains a highest weight vector v of weight
λθ . This implies that µθ − λθ is a positive integral combination of positive roots of
el+ . By the choice of n, µθ − λθ has to be a positive integral combination of posi-
tive roots of el+ e+ e+ e+ θ
n , and hence v ∈ U(ln )w and U(ln )w ⊇ L(ln , λ ). This contradicts
U(el+ e+ θ
n )w = L(ln , µ ). So (6.11) splits, and the lemma is proved. 

Lemma 6.2. As an el+ -module, we have



(Ve0+ )⊗d ∼
= L(el+ , λθ )⊕dλ ,
λ∈Pd

where dλ is the dimension of the Specht module of Sd corresponding to λ.

Proof. Set W = (Ve0+ )⊗d . Fix an integer n ≥ d. By Theorem 3.10, we have an


isomorphism of el+
n -modules

tr∞ e + ⊗d ∼
n W = (V0,n ) = L(el+ θ ⊕dλ
n ,λ ) .
λ∈Pd

Now let w be a singular vector in the el+ ∞ θ


n -module trn W of weight λ , for a fixed
λ ∈ Pd . Observe, as in the proof of Lemma 6.1, that w is also a singular vector
with respect to the standard Borel subalgebra of el+ in W . This implies that for each
λ ∈ Pd , the module L(el+ , λθ ) appears in W as a composition factor with multiplicity
at least dλ . A comparison of el+ -characters shows that each L(el+ , λθ ) appears in W
with multiplicity exactly dλ , and they are all the composition factors of W . Now
the lemma follows from Lemma 6.1. 

The notions of polynomial weights and polynomial modules for el+ can be de-
fined just as in Definition 3.23. As in Proposition 3.24, the irreducible polynomial
el+ -modules are precisely L(el+ , λθ ), for λ ∈ P. The following theorem is an infinite-
rank analogue of Theorem 3.25, and it follows easily from Lemmas 6.1 and 6.2.

Theorem 6.3. The category of polynomial modules of el+ = gl(Ve0+ ) is a semisimple


tensor category.

In the extreme case when k = 0, we have e g = gl(Ve0+ ), and the category of


e defined below in 6.2.3.
polynomial modules of gl(Ve0+ ) is simply the category O
6.2. The module categories 199

6.2.2. Parabolic subalgebras and dominant weights. Recall the Lie superalge-
bras g, g, and e
g from Section 6.1.6, which implicitly depends on a fixed m ∈ Z+ .
We shall in addition fix an arbitrary subset Y0 of Π(k), with the convention that
Y0 = 0/ when m = 0. Let Y , Y , and Ye be the following subsets of Π, Π, and Π,e
respectively:
Y = Y0 ⊔ Π(T) \ {β× }, Y = Y0 ⊔ Π(T) \ {α× }, e \ {α× }.
Ye = Y0 ⊔ Π(T)

As fixing Y0 also fixes the sets Y , Y , and Ye , we will make a convention of suppress-
ing them from notations below.
Let l, l, and el be the standard Levi subalgebras of g, g, and e g corresponding
to the subsets Y , Y , and Ye , respectively. The standard Borel subalgebras of g, g, and
e
g, spanned by the central element K and upper triangular matrices, are denoted by
b, b, and eb, respectively. Let p = l + b, p = l + b, and ep = el + e
b be the corresponding
parabolic subalgebras with nilradicals u, u, and e u and opposite nilradicals u− ,
ū− , and e
u− , respectively.
Let − λ = (λ−m , . . . , λ−1 ) ∈ Cm and let + λ be a partition. Recall also that hα
denotes the coroot for a root α. Associated to a given d ∈ C and a tuple (− λ; + λ)
such that ⟨∑−1
i=−m λi εi , hα ⟩ ∈ Z+ for all α ∈ Y0 ⊆ Π(k), we define the weights
−1
(6.13) λ := dΛ0 + ∑ λi εi + ∑ + λ j ε j ∈ h∗ ,
i=−m j∈N
−1
+ ′ ∗
(6.14) λ♮ := dΛ0 + ∑ λi εi + ∑ λs+ 1 εs
2
∈h ,
i=−m s∈ 21 +Z+
−1
(6.15) λθ := dΛ0 + ∑ λi εi + ∑ θ(+ λ)r εr ∈ e
h∗ .
i=−m r∈ 12 N

These weights λ, λ♮ , λθ will be referred to as dominant weights. We denote by



P+ ⊂ h∗ , P ⊂ h and Pe+ ⊂ e
+
h∗ the sets of dominant weights of the forms (6.13),
(6.14) and (6.15), for all d ∈ C, respectively. We will also identify an element
λ ∈ P+ of the form (6.13) with the tuple (dΛ0 , − λ, + λ). By definition we have
bijective maps
+
♮ : P+ −→ P , λ 7→ λ♮ ,
θ : P+ −→ Pe+ , λ 7→ λθ .

For λ ∈ P+ , let L(l, λ) denote the highest weight irreducible l-module of highest
weight λ. We extend L(l, λ) to a p-module by letting u act trivially. Define the
parabolic Verma g-module ∆(λ) as
∆(λ) := Indgp L(l, λ) = U(g) ⊗U(l) L(l, λ),
and denote its unique irreducible quotient g-module by L(λ).
200 6. Super duality

Similarly, for λ ∈ P+ , we define the irreducible l-module L(l, λ♮ ), the parabolic


Verma g-module ∆(λ♮ ) and its unique irreducible quotient g-module L(λ♮ ); we also
similarly define the irreducible el-module L(el, λθ ), the parabolic Verma e g-module
e θ ) and its unique irreducible quotient e
∆(λ g-module L(λe θ ).

e We now introduce a version of parabolic BGG


6.2.3. The categories O, O and O.
category O of g-modules.
Definition 6.4. Let O be the category of g-modules M such that M is a semisimple
h-module with finite-dimensional weight subspaces Mγ , γ ∈ h∗ , satisfying
(i) M decomposes over l as a direct sum of L(l, µ) for µ ∈ P+ ;
(ii) there exist finitely many weights λ1 , λ2 , . . . , λk ∈ P+ (depending on M)
such that if γ is a weight in M, then γ ∈ λi − ∑α∈Π Z+ α, for some i. (Recall
that Π is the fundamental system for g.)
∗ +
Analogously we define the category O of g-modules using h , P , l, Π, and
e of e
define the category O g-modules using e h∗ , Pe+ ,el, Π.
e The morphisms in O, O and
e
O are all (not necessarily even) g-, g- and e
g-homomorphisms, respectively.
Proposition 6.5. Let µ ∈ P+ . The following statements hold.
(1) The restrictions to l of the g-modules ∆(µ) and L(µ) decompose as direct
sums of L(l, ν), for ν ∈ P+ .
(2) The restrictions to l of the g-modules ∆(µ♮ ) and L(µ♮ ) decompose as direct
sums of L(l, ν♮ ), for ν ∈ P+ .
(3) The restrictions to el of the e e θ ) and L(µ
g-modules ∆(µ e θ ) decompose as di-
rect sums of L(el, νθ ), for ν ∈ P+ .

Proof. (1) As a Lie (super)algebra we have l ∼ = k0 ⊕ gl(V0+ ), where k0 is a Levi


subalgebra of k corresponding to Y0 (our convention is that K ∈ k0 ), and V0+ was
defined in 6.1.3. Hence an irreducible l-module of highest weight ν ∈ P+ is iso-
morphic to a tensor product of a finite-dimensional irreducible k0 -module and a
polynomial gl(V0+ )-module. Note that the opposite nilradical u− as an l-module is
a direct sum of irreducible modules with highest weight ν ∈ P+ . More explicitly,
we have the following isomorphisms of l-modules:
 m∗

 C ⊗V0+ ⊕ u− 0, for x = a

 −
 C ⊗V0 ⊕ ∧ (V0 ) ⊕ u0 , for x = b
2m+1 + 2 +

u− ∼= C2m|1 ⊗V0+ ⊕ S2 (V0+ ) ⊕ u−


0, for x = b•

 −

 C ⊗V0 ⊕ S (V0 ) ⊕ u0 ,
2m + 2 +
for x = c
 2m −
C ⊗V0 ⊕ ∧ (V0 ) ⊕ u0 ,
+ 2 +
for x = d.
+∼ ∞
Here V0 = C is a purely even space and it is isomorphic to the natural module of
gl(V + ) ∼
0 = gl(∞), u− is the opposite nilradical of k corresponding to the Levi sub-
0
algebra k0 , and furthermore Cm , C2m , C2m+1 , and C2m|1 are the natural k-modules,
6.2. The module categories 201

on which k0 acts semisimply by restriction. It is evident that the tensor product


of two irreducible l-modules with highest weights in P+ decomposes into a di-
rect sum of irreducibles with highest weights in P+ . Recall S(U) stands for the
supersymmetric algebra of a superspace U. Since, as l-modules, we have
∆(µ) ∼
= S(u− ) ⊗ L(l, µ),
it follows that ∆(µ) is a direct sum of irreducible l-modules with highest weights in
P+ . Since L(µ) is a quotient of ∆(µ), the same holds for L(µ). This proves (1).
+
(2) The Levi subalgebra of g is l = k0 ⊕ gl(V 0 ), which is isomorphic to l. The
opposite nilradical ū− is a direct sum of irreducible l-modules with highest weights
of the form ν♮ , where ν ∈ P+ . Explicitly, we have

Cm∗ ⊗V 0 ⊕ u−
+

 0, for x = a

 −

 C
2m+1 + +
⊗V 0 ⊕ S (V 0 ) ⊕ u0 , for x = b
2
−∼
⊗V 0 ⊕ ∧2 (V 0 ) ⊕ u−
+ +
u = C 2m|1
0, for x = b•




+ +
C ⊗V 0 ⊕ ∧ (V 0 ) ⊕ u0 ,
2m 2 −
for x = c

 2m + + −
C ⊗V 0 ⊕ S (V 0 ) ⊕ u0 ,
2 for x = d.

A main difference from (1) is that V 0 ∼


+
= C∞ here is a purely odd space, yet the
+
notations for the exterior and the symmetric squares of V 0 here are understood in
the non-super sense. Now we have ∆(λ♮ ) = S(ū− ) ⊗ L(l, λ♮ ), and so a verbatim
argument as in (1) establishes (2).
(3) The Levi subalgebra el of eg is isomorphic to k0 ⊕ gl(Ve0+ ). As an el-module,
u− is isomorphic to a direct sum of irreducible modules, each of which is a tensor
e
product of a finite-dimensional irreducible k0 -module and an irreducible polyno-
mial module of gl(Ve0+ ). Recall the polynomial modules of gl(Ve0+ ) were shown to
form a semisimple tensor category in Theorem 6.3. We have


 Cm∗ ⊗ Ve0+ ⊕ u−0, for x = a



 C 2m+1 ⊗ Ve +
⊕ ∧ e
2 (V +
) ⊕ u −
, for x = b
0 0 0
−∼ e 2 e+ −
u = C 2m|1 ⊗ V0 ⊕ S (V0 ) ⊕ u0 ,
+
for x = b•



 C2m ⊗ Ve0 ⊕ S (Ve0 ) ⊕ u0 ,
+ 2 + −
for x = c

 ∼ 2m e +
= C ⊗ V0 ⊕ ∧2 (Ve0+ ) ⊕ u− 0 , for x = d.
Here the exterior squares are understood in the super sense for the superspace Ve0+ .
So each irreducible el-submodule of e u− has highest weight lying in Pe+ . Thus, by
e θ ), for λ ∈ P+ , is also a direct sum of irreducible el-modules with
Theorem 6.3, ∆(λ
highest weights in Pe+ , and so is its quotient L(λ
e θ ). This proves (3). 

As an immediate consequence of Proposition 6.5 we have the following.


Corollary 6.6. Let λ ∈ P+ .
(1) The modules ∆(λ) and L(λ) lie in O.
202 6. Super duality

(2) The modules ∆(λ♮ ) and L(λ♮ ) lie in O.


e θ ) and L(λ
(3) The modules ∆(λ e
e θ ) lie in O.

6.2.4. The categories On , On and O e n . SD:subsec:catOn


For n ∈ N, recall the sets Πn , Πn , Πe n of simple roots for the Dynkin diagrams
(6.2) and the associated finite-dimensional Lie superalgebras gn , gn and e gn from
Section 6.1.5. These Lie superalgebras gn , gn and e gn can be identified naturally
with the subalgebras of g, g and eg generated by the central element K and the root
vectors associated to the fundamental system in (6.2). Moreover, we have natural
inclusions gn ⊂ gn+1 , gn ⊂ gn+1 , and e gn ⊆ e gn+1 , with g = ∪n gn , g = ∪n gn , and
e
g = ∪negn . The standard Cartan and Borel subalgebras of gn are
hn = h ∩ gn , b n = b ∩ gn ,
respectively. Similarly, we write hn and e hn ( bn and e
bn ) for the standard Cartan
(Borel) subalgebras of gn and e
gn , respectively.
Recall the notations λ ∈ P+ , λ♮ , and λθ from (6.13), (6.14) and (6.15). Given
λ ∈ P+ with + λ j = 0 for j > n, we may regard λ as a weight in h∗n in a natural

way. Similarly, for λ ∈ P+ with + λ j = 0 for j > n, we regard λ♮ as a weight in

hn . Finally, for λ ∈ P+ with θ(+ λ) j = 0 for j ∈ 21 N with j > n, we regard λθ as a

weight in e h∗n . These weights λ, λ♮ , λθ in h∗n , hn , and e
h∗n respectively with the above
constraints will be called the dominant weights. The subsets of dominant weights

in h∗n , hn , and e h∗n will be denoted by Pn+ , Pn , and Pen+ , respectively.
+

Corresponding to a fixed Y0 ⊆ Π(k), the Levi and parabolic subalgebras of the


finite-rank Lie superalgebra gn are
ln = l ∩ gn , pn = p ∩ gn ,
respectively. This allows us to define the corresponding parabolic Verma and irre-
ducible gn -modules ∆n (µ) and Ln (µ) with highest weight µ ∈ Pn+ . The correspond-
ing category of gn -modules is denoted by On , which is defined as in Definition 6.4,
now with h, l, P+ and Π therein replaced by hn , ln , Pn+ and Πn , respectively.
The statements in the previous paragraph admit obvious counterparts for the
Lie superalgebras gn and e gn as well. We introduce the self-explanatory notations
∆n (υ), Ln (υ), On , l, p for gn , and ∆ e n el, e
en (ν), Ln (ν), O p for e
+
gn , where υ ∈ Pn and
ν ∈ Pen+ .

6.2.5. Truncation functors. Let n < k ≤ ∞. For M ∈ Ok , we write M = γ Mγ ,
where γ ∈ ∑−1
i=−m Cεi + ∑0< j≤k Z+ ε j + CΛ0 , according to its weight space decom-
position. The truncation functor trkn : Ok → On is defined by

trkn (M) = Mν ,
ν
6.3. The irreducible character formulas 203

where the summation is over ν ∈ ∑−1 i=−m Cεi + ∑0< j≤n Z+ ε j +CΛ0 . When it is clear
from the context we shall also write trn instead of trkn . Analogously, truncation
ek → O
functors trkn : Ok → On and trkn : O e n are defined. These functors are obviously
exact. (The notation trkn used earlier in 6.2.1 corresponds to an extreme case here
when k = 0.)
Recall from (6.6) and Section 6.1.6 the notation E j and Ebj for basis elements
of the Cartan subalgebras.
Proposition 6.7. Let n < k ≤ ∞ and X = L, ∆.
{
( ) Xn (µ), if ⟨µ, Ebj ⟩ = 0, ∀ j > n,
(i) For µ ∈ Pk we have trn Xk (µ) =
+
0, otherwise.
{
+ ( ) X n (µ), if ⟨µ, Ebj ⟩ = 0, ∀ j > n,
(ii) For µ ∈ Pk we have trn X k (µ) =
0, otherwise.
{
( ) Xen (µ), if ⟨µ, Ebj ⟩ = 0, ∀ j > n,
(iii) For µ ∈ Pek+ we have trn Xek (µ) =
0, otherwise.

Proof. We shall only show (i), as similar arguments prove (ii) and (iii). Since
trkn ◦ trlk = trln , it suffices to show (i) for k = ∞.
First suppose that ⟨µ, Ebn+1 ⟩ > 0. Then every weight ν of L(l, µ) also satisfies
⟨ν, Ebn+1 ⟩ > 0. Recall from the proof of Proposition 6.5 that as an l-module u−
decomposes into a direct sum of irreducibles with highest weights in P+ . Thus,
every weight υ in ∆(µ) = S(u− )⊗L(l, µ) must also satisfy ⟨υ, Ebn+1 ⟩ > 0. Therefore,
trn (∆(µ)) = 0, and hence trn (L(µ)) = 0.
Now suppose that ⟨µ, Ebn+1 ⟩ = 0. Since µ ∈ Pk+ , this implies that ⟨µ, Ebj ⟩ = 0
for all j > n. Let l′ and p′ denote the standard Levi and parabolic subalgebras
of g corresponding to the removal of the vertex βn of the Dynkin diagram of g.
Then l′ ∼ = gn ⊕ gl(W ), where W is the subspace of V0+ spanned by the vectors vi ,
i > n. Consider the parabolic Verma module Indgp′ Ln (µ), where Ln (µ) extends to an
l′ -module by a trivial action of gl(W )-module and is then extended to a p′ -module
in a trivial way. Clearly, L(µ) is the unique irreducible quotient of Indgp′ Ln (µ).
Since trn (Indgp′ Ln (µ)) = Ln (µ) and trn (L(µ)) is a nonzero gn -module (as it contains
a non-zero vector of weight µ), we conclude that trn (L(µ)) = Ln (µ).
To complete the proof, we observe that ∆(µ) ∼ = S(u− ) ⊗ L(l, µ), and that the
gn -module trn (∆(µ)) has highest weight µ with character equals ch ∆n (µ). From
these we conclude that trn (∆(µ)) = ∆n (µ). 

6.3. The irreducible character formulas


e → O and T : O
In this section, functors T : O e → O are introduced, and they are
shown to send irreducible modules and parabolic Verma modules to irreducible and
204 6. Super duality

parabolic Verma modules, respectively. The main ingredients for studying these
functors are two sequences of odd reflections on the standard Borel subalgebra of
e
g, leading to new Borel subalgebras which are “approximately compatible” with
the standard Borel subalgebras of g and g. As a consequence, the solution of the
irreducible character problem in O via the classical Kazhdan-Lusztig theory also
e
provides a complete solution to the irreducible character problem for O and O.

6.3.1. Two sequences of Borel subalgebras of e g. Let us recall briefly the basics
on odd reflections from Chapter 1, Sections 1.3 and 1.4, which are applied now to
the Lie superalgebra e g. Under the odd reflection with respect to an isotropic odd
simple root α in a fundamental system Π e ′ , the resulting new fundamental system
e α is given by (1.30) as follows:
Π
e α = {β ∈ Π
Π e ′ | (β, α) = 0, β ̸= α} ∪ {β + α | β ∈ Π e ′ , (β, α) ̸= 0} ∪ {−α}.

Let us denote the Borel subalgebra corresponding to Π e ′ by b′ and the resulting


new Borel subalgebra by bα . According to Lemma 1.36, an irreducible e g-module
of b′ -highest weight λ is also a bα -highest weight module with bα -highest weight
equal to either λ or λ − α, depending on whether or not (λ, α) = 0. It is instructive
to review Examples 1.31 and 1.37 to see how the highest weights change under a
sequence of odd reflections, as they offer in a simpler setting a similar pattern to
what we shall see below.
Recall from (6.1) the odd roots αr and the even roots βr , for r ∈ 12 N. Recall
from Section 6.1 that the standard Dynkin diagram associated to eg is given by
 ⊗ ⊗ ⊗ ⊗ ⊗ ⊗
(6.16)
k ··· ···
α× α1/2 α1 αn−1/2 αn αn+1/2

Fix n ∈ N. We shall specify a sequence of n(n+1)


2 odd reflections, which will trans-
form the standard diagram (6.16) into a new diagram with a new fundamental sys-
tem of the form (6.18) below. The ordered sequence of n(n+1) 2 odd roots in use
are:
ε1/2 − ε1 ,
ε3/2 − ε2 , ε1/2 − ε2 ,
(6.17) ε5/2 − ε3 , ε3/2 − ε3 , ε1/2 − ε3 ,
··· ,··· ,··· ,··· ,
εn−1/2 − εn , εn−3/2 − εn , ..., ε3/2 − εn , ε1/2 − εn .
Let us go into the details. First, applying the odd reflection corresponding to α1/2 =
ε1/2 − ε1 , we obtain the following diagram:
 ⊗ ⊗ ⊗ ⊗ ⊗

k ⃝ ⃝
α3/2 α5/2
···
β× ε1 − ε1/2 β1/2 α2 α3
6.3. The irreducible character formulas 205

Next, applying to it consecutively the two odd reflections corresponding to α3/2 =


ε3/2 − ε2 and α1/2 + α1 + α3/2 = ε1/2 − ε2 , we obtain the following two diagrams:

 ⊗ ⊗ ⊗ ⊗ ⊗

k ⃝ ⃝
α5/2
···
β× ε1 − ε1/2 ε1/2 − ε2 −α3/2 ε3/2 − ε5/2 α3

 ⊗ ⊗ ⊗

k ⃝ ⃝ ⃝ ⃝
α5/2
···
β× β1 ε2 − ε1/2 β1/2 β3/2 α3

Then, to the last diagram we apply consecutively the three odd reflections corre-
sponding to α5/2 = ε5/2 − ε3 , α3/2 + α2 + α5/2 = ε3/2 − ε3 , and α1/2 + α1 + α3/2 +
α2 + α5/2 = ε1/2 − ε3 and obtain the following three diagrams:

 ⊗ ⊗ ⊗ ⊗

k ⃝ ⃝ ⃝ ⃝
α7/2
···
β× β1 ε2 − ε1/2 β1/2 ε3/2 − ε3 −α5/2 β5/2

 ⊗ ⊗ ⊗ ⊗

k ⃝ ⃝
ε2 − ε1/2 ε1/2 − ε3 ε3 − ε3/2
⃝ ⃝
α7/2
···
β× β1 β3/2 β5/2

 ⊗ ⊗

k ⃝ ⃝ ⃝
ε3 − ε1/2
⃝ ⃝ ⃝
α7/2
···
β× β1 β2 β1/2 β3/2 β5/2

Continue this way until we finally apply the n odd reflections corresponding to
αn−1/2 , αn−3/2 + αn−1 + αn−1/2 , . . ., ∑2n−1 i=1 αi/2 , which equal εn−1/2 − εn , εn−3/2 −
εn , . . . , ε1/2 − εn , respectively. The resulting new Borel subalgebra for e g in the end
will be denoted by e c
b (n), and the corresponding positive system will be denoted
e
by Φ (n). The corresponding fundamental system, denoted by Π
c e c (n), is listed in
the following Dynkin diagram:
  ⊗ ⊗ ⊗
(6.18)
k
Tn ⃝ ··· ⃝ ···
εn − ε1/2 β1/2 βn−1/2 αn+1/2 αn+1


The crucial point here is that the subdiagram to the left of the first in (6.18) is
the Dynkin diagram of gn and the precise detail of the remaining part of (6.18) is
not needed.
On the other hand, starting with the standard Dynkin diagram (6.16) of e g , we
may apply a different sequence of n(n+1)
2 odd reflections associated to the following
206 6. Super duality

n(n+1)
ordered sequence of 2 odd roots:
ε1 − ε3/2 ,
ε2 − ε5/2 , ε1 − ε5/2 ,
(6.19) ε3 − ε7/2 , ε2 − ε7/2 , ε1 − ε7/2 ,
··· ,··· ,··· ,··· ,
εn − εn+1/2 , εn−1 − εn+1/2 , ..., ε2 − εn+1/2 , ε1 − εn+1/2 .
Since the procedure here is completely parallel to the previous one for the sequence
(6.19), we skip the details. The resulting new Borel subalgebra for e g in the end
es
will be denoted by b (n), and the corresponding positive system will be denoted by
e s (n). The corresponding fundamental system, denoted by Π
Φ e s (n), is listed in the
following Dynkin diagram:

  ⊗ ⊗ ⊗
(6.20) Tn+1

k
⃝ ··· ⃝ ···
εn+1/2 − ε1 β1 βn αn+1 αn+3/2

Again the crucial fact here is that the subdiagram to the left of the odd simple root
εn+1/2 − ε1 in (6.20) is the Dynkin diagram of gn+1 .
+
Recall that Φ+ and Φ denote the standard positive systems of g and g, re-
spectively. Also, b and b denote the corresponding standard Borel subalgebras of
g and g, respectively. By definition, g and g are naturally subalgebras of e
g.
e c (n), and b = e
Proposition 6.8. We have Φ+ ⊂ Φ
+ e s (n), and
bc (n) ∩ g. Also, Φ ⊂ Φ
b=ebs (n) ∩ g.
e c (n) and Π ⊂ Φ
Proof. It suffices to check that Π ⊂ Φ e s (n), where Π and Π are the
standard fundamental systems of g and g, respectively. These inclusions can then
be observed directly from (6.18) and (6.20). 

6.3.2. Odd reflections and highest weight modules. Recall the standard Levi
subalgebra el of e
g with nilradical e u− from Section 6.2.2.
u and opposite nilradical e
Lemma 6.9. The sequences of odd reflections (6.17) and (6.19) leave the set of
roots of e u− invariant.
u and the set of roots of e

Proof. The isotropic odd roots used in the sequences of odd reflections (6.17) and
(6.19) are of the form αi + αi+1/2 + . . . + α j for 0 < i < j, hence they are all roots
of el. Since e u− are both invariant under the action of el, the lemma follows by
u and e
applying Lemma 1.26. 

We denote by e bec (n) and e


bes (n) the Borel subalgebras of the Levi subalgebra el
l l
corresponding to the fundamental systems Π e c (n) ∩ ∑ e Zα and Πe s (n) ∩ ∑ e Zα,
α∈Y α∈Y
6.3. The irreducible character formulas 207

respectively. By Lemma 6.9 we have, for all n ∈ N,


(6.21) e
bc (n) = e
becl (n) + e
u, e
bs (n) = e
besl (n) + e
u.

Below we will regard P+ ⊂ Pe+ and P ⊂ Pe+ , just as we regard h∗ ⊂ e


+
h∗ and

h ⊂e
h∗ .
Proposition 6.10. Let λ = (dΛ0 , − λ, + λ) ∈ P+ , where + λ = (+ λ1 , + λ2 , . . .) is a
partition. Let n ∈ N.
(1) Suppose that ℓ(+ λ) ≤ n. Then the highest weight of L(el, λθ ) with respect
to the Borel subalgebra e
bec (n) is λ.
l
(2) Suppose that 1 ≤ n. Then the highest weight of L(el, λθ ) with respect to

the Borel subalgebra e


bes (n) is λ♮ .
l

Proof. The proof of (2) using the sequence (6.19) is completely parallel to the
proof of (1) which uses (6.17). We shall only prove (1).

We observe that the odd  e
reflections (6.17) only affect the tail diagram
T and
leave the head diagram
k unchanged, and the summand
−1
(6.22) λ|k := dΛ0 + ∑ λi εi
i=−m

of λθ in (6.15) is unchanged by these odd reflections.


We will show more generally by induction on k that, after applying the first
k(k + 1)/2 odd reflections of the sequence (6.17), the highest weight of L(l, λθ )
with respect to the Borel subalgebra e
bec (k) becomes
l
k k

(6.23) λ[k] =λ|k + ∑ + λi εi + ∑ ⟨+ λi − k⟩εi−1/2
i=1 i=1

+ ∑ ⟨+ λ j − j + 1⟩ε j−1/2 + ∑ ⟨+ λ j − j⟩ε j ,
j≥k+1 j≥k+1

where ⟨q⟩ := max{q, 0} for q ∈ Z. Part (1) follows as a special case of (6.23) for
k = n. Note that λ[0] = λθ = λ|k + ∑ j≥1 ⟨λ′j − j + 1⟩ε j−1/2 + ∑ j≥1 ⟨λ j − j⟩ε j .
Suppose that k = 1. If ℓ(+ λ) < 1, then + λ = 0 and λθ = λ|k . So in particular
⟨λθ , Eb1/2 + Eb1 ⟩ = 0, and thus by Lemma 1.36, the e bec (1)-highest weight is λ[1] = λθ .
l
If ℓ(+ λ) ≥ 1, then ⟨λθ , Eb1/2 + Eb1 ⟩ > 0, as we recall that

λθ = λ|k + + λ1 ε1/2 + (+ λ1 − 1)ε1 + · · · .
Hence, by Lemma 1.36, the e
bec (1)-highest weight after the odd reflection with re-
l
spect to ε1/2 − ε1 is

λ[1] = λ|k + + λ1 ε1 + (+ λ1 − 1)ε1/2 + · · · ,
208 6. Super duality

proving (6.23) in the case k = 1.


Now by the induction hypothesis, suppose that λ[k] is the new highest weight
after applying the first k(k + 1)/2 odd reflections of the sequence (6.17).
If ℓ(+ λ) ≤ k, then λ[k] = λ|k + ∑ki=1 + λi εi . Therefore, for 1 ≤ i ≤ k + 1, we have
⟨λ[k] , Ebi−1/2 + Ebk+1 ⟩ = 0, and by Lemma 1.36, the odd reflections with respect to
εi−1/2 − εk+1 leave λ[k] unchanged. Thus we have λ[k+1] = λ[k] , and in this case we
are done.
Now assume that ℓ(+ λ) ≥ k + 1. Let s = + λk+1 . We further separate this into
two cases (i) and (ii) below.

(i) Suppose that + λk+1 ≥ k + 1. Then + λk+1 ≥ k + 1, and hence we can rewrite
k k
′ ′
λ[k] = λ|k + ∑ + λi εi + ∑ (+ λi − k)εi−1/2 + (+ λk+1 − k)εk+1/2
i=1 i=1

+ ( λk+1 − k − 1)εk+1 +
+
∑ ⟨+ λ j − j + 1⟩ε j−1/2 + ∑ ⟨+ λ j − j⟩ε j .
j≥k+2 j≥k+2

Now we apply the next (k + 1) odd reflection in the sequence (6.17) consecutively.
As ⟨λ[k] , Ebk+1/2 + Ebk+1 ⟩ > 0, by Lemma 1.36 we calculate the new weight after the
odd reflection with respect to εk+1/2 − εk+1 to be
k k
′ ′
λ[k,1] = λ|k + ∑ + λi εi + ∑ (+ λi − k)εi−1/2 + (+ λk+1 − k − 1)εk+1/2
i=1 i=1

+ ( λk+1 − k)εk+1 +
+
∑ ⟨+ λ j − j + 1⟩ε j−1/2 + ∑ ⟨+ λ j − j⟩ε j .
j≥k+2 j≥k+2

Now ⟨λ[k,1] , Ebk−1/2 + Ebk+1 ⟩ > 0, so by Lemma 1.36 we calculate the new weight
after the odd reflection with respect to εk−1/2 − εk+1 to be
k k−1
′ ′
λ[k,2] =λ|k + ∑ + λi εi + ∑ (+ λi − k)εi−1/2 + (+ λk − k − 1)εk−1/2
i=1 i=1
+ ′
+ ( λk+1 − k − 1)εk+1/2 + (+ λk+1 − k + 1)εk+1

+ ∑ ⟨+ λ j − j + 1⟩ε j−1/2 + ⟨+ λ j − j⟩ε j . ∑
j≥k+2 j≥k+2

Continuing this way, after a total of (k + 1) odd reflections we end up with the
weight
k+1 k+1

λ[k,k+1] =λ|k + ∑ + λi εi + ∑ (+ λi − k − 1)εi−1/2
i=1 i=1

+ ∑ ⟨+ λ j − j + 1⟩ε j−1/2 + ∑ ⟨+ λ j − j⟩ε j ,
j≥k+2 j≥k+2

which is exactly λ[k+1] . The induction step is completed in this case.


6.3. The irreducible character formulas 209


(ii) Now suppose that + λk+1 = s < k + 1. Since + λ is a partition, the weight
(6.23) becomes
k s

λ[k] = λ|k + ∑ + λi εi + ∑ (+ λi − k)εi−1/2 ,
i=1 i=1

where + λ′ − k
> 0, for i ≤ s. It follows by Lemma 1.36 that odd reflections with
i
respect to εk+1/2 − εk+1 , · · · , εs+1/2 − εk+1 leave λ[k] unchanged, while odd refec-
tions with respect to εs−1/2 − εk+1 , · · · , ε1/2 − εk+1 do affect λ[k] . In a similar way,
the new weight after these (k + 1) odd reflections is calculated to be
k+1 s

λ[k,k+1] = λ|k + ∑ + λi εi + ∑ (+ λi − k − 1)εi−1/2 ,
i=1 i=1

which is again equal to λ[k+1] . The induction step is completed. 

Proposition 6.11. Let n ∈ N and λ = (dΛ0 , − λ, + λ) ∈ P+ . Let Ve (λθ ) be a highest


g-module of highest weight λθ with respect to the standard Borel subalgebra
weight e
e
b. The following statements hold.
(1) Suppose that ℓ(+ λ) ≤ n. Then Ve (λθ ) is a highest weight module of highest
weight λ with respect to the Borel subalgebra e bc (n).

(2) Suppose that + λ1 ≤ n. Then Ve (λθ ) is a highest weight module of highest
weight λ♮ with respect to the Borel subalgebra e bs (n).

Proof. By definition, ∆(λe θ ) = U(e g) ⊗U(ep) L(el, λθ ), and hence its e


g-quotient Ve (λθ )
contains a unique copy of L(el, λθ ) that is annihilated by e u. By Proposition 6.10(1),
L(el, λθ ) has e
bc (n)-highest weight λ, and let us denote by v a corresponding e
el bc (n)-
el
highest weight vector. Thus, by (6.21), v is also a e bc (n)-singular vector in Ve (λθ )
of weight λ. The vector v clearly generates the l-module L(el, λθ ) and hence the
e
g-module Ve (λθ ), proving the first statement of (1).
e
The proof of (2) based on Proposition 6.10(2) is similar and hence omitted. 

6.3.3. The functors T and T . By definition, g and g are naturally subalgebras of


g, l and l are subalgebras of el, while h and h are subalgebras of e
e h. Also, we have

inclusions of the restricted duals h∗ ⊆ eh∗ and h ⊆ eh∗ .
Given a semisimple e h-module M e = ⊕ e∗ M eγ , we form the following sub-
γ∈h
e
spaces of M:
⊕ ⊕
(6.24) e :=
T (M) eγ ,
M and e :=
T (M) eγ .
M
γ∈h∗ ∗
γ∈h

e is an h-submodule of M,
Note that T (M) e and T (M)
e is an h-submodule of M.
e One
⊕ e
e e e e
checks that if M = γ∈eh∗ Mγ is an l-module, then T (M) is an l-submodule of M
210 6. Super duality

e is an l-submodule of M.
and T (M) e Furthermore, if M e is a e e is
g-module, then T (M)
a g-submodule of Me and T (M)
e is a g-submodule of M. e
The direct sum decomposition in Me gives rise to the natural projections
e −−−−→ T (M)
TMe : M e and e −−−−→ T (M)
T Me : M e

that are h- and h-module homomorphisms, respectively. If fe : Me→Ne is an e


h-
e
homomorphism, then the following maps induced by restriction of f
T [ fe] : T (M)
e −−−−→ T (N)
e and T [ fe] : T (M)
e −−−−→ T (N)
e

are also h- and h-module homomorphisms, respectively. Also if fe : M e→N e is a


e e e
g-homomorphism, then TMe and T [ f ] (respectively, T Me and T [ f ]) are g-module (re-
spectively, g-module) homomorphisms. It follows that T and T define exact func-
tors from the category of e
h-semisimple eg-modules to the category of h-semisimple
g-modules and the category of h-semisimple g-modules, respectively. Also, we
have the following commutative diagrams:
e e
e −−−f−→
M e
N e −−−f−→
M e
N
   
T T T T
(6.25) y Me y Ne y Me y Ne
e e
e −−T−[−
T (M)
f]
e
→ T (N) e −−T−[−
T (M)
f]
e
→ T (N)

Lemma 6.12. For λ ∈ P+ , we have


( ) ( )
T L(el, λθ ) = L(l, λ), T L(el, λθ ) = L(l, λ♮ ).

Proof. We prove the first formula using a comparison of the characters. The proof
of the second formula is similar and will be omitted.
Let λ = (dΛ0 , − λ, + λ) ∈ P+ . To be consistent with the notation of λ|k in (6.22),
it is convenient to make a convention that k contains the central element K, and let
L(el ∩ k, λ|k ) denote the irreducible el ∩ k-module of highest weight λ|k . It follows by
the discussion in Section 6.2.1 that
(6.26) ch L(el, λθ ) = ch L(el ∩ k, λ|k ) hs+ λ′ (x1/2 , x3/2 , . . . ; x1 , x2 , . . .),
where xr := eεr for r ∈ 12 N.
Note that el ∩ k = l ∩ k. Now L(el, λθ ) is completely reducible as an l-module,
since a polynomial module of gl(Ve0+ ) = gl(∞|∞) is completely reducible when
restricted to gl(∞). On the character level, applying T to L(el, λθ ) corresponds to
setting x1/2 , x3/2 , x5/2 , . . . in the character formula (6.26) to zero. Thus, by (A.34),
( )
T L(el, λθ ) is an l-module with character being ch L(el∩k, λ|k ) s+ λ (x1 , x2 , . . .), which
is precisely the character of L(l, λ). This proves the first formula. 
6.3. The irreducible character formulas 211

e to O, and T is an exact functor


Proposition 6.13. T is an exact functor from O
e to O.
from O

Proof. In light of Proposition 6.12, it suffices to show that if M e then T (M)


e ∈ O, e ∈O
e e
and T (M) ∈ O. We shall only show that T (M) ∈ O, as the argument for T (M) ∈ Oe
is analogous.
First consider the case when head diagram is non-degenerate, i.e., m > 0 (see
Section 6.1.1). Let M e Then there exists ζθ , ζθ , . . . , ζθ , with ζ1 , . . . , ζr ∈ P+ ,
e ∈ O.
1 2 r
such that any weight of M e is bounded by some ζi . Ignoring dΛ0 recall that we have
ζ j = (− ζ j , + ζ j ), where + ζ j is a partition with |+ ζ j | = k j . For each ζ j , let Pj be the
following finite subset of h∗ :
Pj := {(− ζ j , µ) | µ ∈ P with |µ| = k j }.
e and P(M) := ∪rj=1 Pj .
Set M := T (M)
Claim. Given any weight ν of M, there exists γ ∈ P(M) such that γ − ν ∈ Z+ Π.
e we
It suffices to prove the claim for ν ∈ P+ . Since ν is also a weight of M,
θ e
have ζi − µ ∈ Z+ Π, for some i. Thus
ζθi − ν = pα× + − κ + + κ,
where p ∈ Z+ , − κ ∈ ∑α∈Π(k) Z+ α, and + κ ∈ ∑β∈Π(T)\α e ×
Z+ β. This implies that + ν
is a partition of size ki + p, and hence there exists γ ∈ P(M) such that + ν is obtained
from the partition + γ by adding p boxes to it. For every such a box, we record the
row number in which it was added in the multiset J with |J| = p. Then we have
ν = γ − − κ − ∑ (ε−1 − ε j ),
j∈J

and hence ν < γ. Thus, we conclude that M ∈ O.


The limit case m = 0 now can be proved case by case by slight modification of
the argument above. 
e is a highest weight e
Proposition 6.14. Let λ ∈ P+ . If Ve (λθ ) ∈ O g-module of highest
θ e θ e θ
weight λ , then T (V (λ )) and T (V (λ )) are highest weight g- and g-modules of
highest weights λ and λ♮ , respectively.

Proof. We will only prove the statement for T (Ve (λθ )), as the case of T (Ve (λθ )) is
analogous.
By Proposition 6.11, for n > ℓ(+ λ), Ve (λθ ) is a e
bc (n)-highest weight module of
highest weight λ, and thus a nonzero vector v in Ve (λθ ) of weight λ (unique up to
a scalar multiple and independent of n) is a ebc (n)-highest weight vector of Ve (λθ ).
Evidently we have v ∈ T (Ve (λθ )), and by Proposition 6.8, v is a b-singular vector
in T (Ve (λθ )).
212 6. Super duality

The g-module T (Ve (λθ )), regarded as an l-module, is completely reducible by


Proposition 6.5 and Lemma 6.12. To complete the proof, it suffices to show that
every vector w ∈ T (Ve (λθ )) of weight µ = (− µ, + µ) ∈ P+ lies in U(n− )v, where n−
denotes the opposite nilradical of b.
To that end, we choose n such that n > max{ℓ(+ λ), ℓ(+ µ)}. Then we have
w ∈ U(e nc (n)− denotes the opposite nilradical of e
nc (n)− )v, where e bc (n). Now the
condition n > max{ℓ( λ), ℓ( µ)} imply that
+ +

−1 n−1
(6.27) λ−µ = ∑ ai εi + ∑ b j ε j , ai , b j ∈ Z.
i=−m j=1

(We emphasize that only ε j with positive integer indices j appear in (6.27).) But
e c (n), i.e.,
λ − µ is also a Z+ -linear combination of simple roots from Π
λ−µ = ∑ aα α,
e c (n)
α∈Π

with all aα ∈ Z+ and finitely many aα > 0. We have the following


Claim. λ − µ is a Z+ -linear combination of Πn = Π(k) ⊔ {β−1 , β1 , . . . , βn−1 }.
The claim then implies that w ∈ U(n− )v, and the proposition follows.
It remains to prove the claim. Assume on the contrast that there were some
α∈Π e c (n) \ Πn = {εn − ε1/2 , β1/2 , . . . , βn−1/2 , αn+1/2 , αn+1 , . . .} with aα ̸= 0. If
aεn −ε 1 ̸= 0, then we must also have aα ̸= 0, for α = β1/2 , . . . , βn−1/2 , αn+1/2 by the
2
constraint (6.27). Similarly, if aβi−1/2 ̸= 0 for some 1 ≤ i ≤ n, then we must also
have aα ̸= 0, for α = βi+1/2 , . . . , βn−1/2 , αn+1/2 as well. In all cases, ⟨λ − µ, Ebr ⟩ ̸= 0,
for some r ≥ n, contradicting (6.27). 

6.3.4. Character formulas. The following theorem is the first main result of this
chapter.
Theorem 6.15. Let λ ∈ P+ . We have
( ) ( θ )
e θ ) = ∆(λ), T L(λ
T ∆(λ e ) = L(λ);
( ) ( θ )
e θ ) = ∆(λ♮ ), T L(λ
T ∆(λ e ) = L(λ♮ ).

Proof. We will prove only the statements involving T , and the statements involving
T can be proved in the same way.
Let us write ∆(λ) = U(u− ) ⊗C L(l, λ) and ∆(λ e θ ) = U(e u− ) ⊗C L(el, λθ ). We

observe that all the weights in U(u ), L(l, λ), U(e u ), and L(el, λθ ) are of the form

∑ j<0 a j ε j + ∑r>0 br εr with br ∈ Z+ (possibly modulo dΛ0 ). Since also T (U(e u− )) =


( ) ( )
U(u− ), it follows by Lemma 6.12 that ch T ∆(λ e θ ) = ch ∆(λ). Since T ∆(λ e θ)
is(a highest) weight module of highest weight λ by Proposition 6.14, we have
e θ ) = ∆(λ).
T ∆(λ
6.4. Kostant homology and KLV polynomials 213

e = L(λ
Set M e θ ). Suppose that M := T (M) e is not irreducible. Since by Propo-
sition 6.14 the M is a highest weight g-module, it must have a b-singular vector,
say w of weight µ ∈ P+ , inside M that is not a highest weight vector. We choose
n > max{ℓ(+ λ), ℓ(+ µ)}. By Proposition 6.11, the e g-module M e is a e
bc (n)-highest
weight module of highest weight λ, say with a highest weight vector vλ . By Propo-
sition 6.14, vλ is a b-highest weight vector of M, and hence w ∈ U(a)vλ , where a
is the subalgebra of n− generated by the root vectors corresponding to the roots in
−Π = −Π(k) ⊔ {−β−1 , −β1 , . . . , −βk }, for some k.
Now choose q > max{n, k + 1}. Note that vλ is also a e bc (q)-highest weight
vector of the eg-module M e of weight λ. Since w is b-singular, it is annihilated by
the root vectors corresponding to the roots in Π(k) ⊔ {β−1 , β1 , β2 , . . .}. But w is
also annihilated by the root vectors corresponding to the roots in Π e c (q) comple-
mentary to the subset Π = Π(k) ⊔ {β−1 , β1 , β2 , . . . , βq−1 }, since these root vectors
commute with a and w ∈ U(a)vλ . It follows that w is a e bc (q)-singular vector in M,e
e
contradicting the irreducibility of M. 

It is standard to write
(6.28) ch L(λ) = ∑ aµλ ch ∆(µ),
µ∈P+

for λ ∈ P+ and aµλ ∈ Z. The triangular transition matrix (aµλ ) in Theorem 6.16
or its inverse is known in the Kazhdan-Lusztig theory, since the Kazhdan-Lusztig
polynomials determine the composition multiplicities of parabolic Verma modules
in the parabolic BGG category O of g-modules; see, e.g., Soergel [So, p. 455]. The
infinite-rank of g here does not cause any difficulty in light of Proposition 6.7. We
have the following character formulas for g and e g.
Theorem 6.16. Let λ ∈ P+ . Then
(1) ch L(λ♮ ) = ∑µ∈P+ aµλ ch ∆(µ♮ ),
e θ ).
e θ ) = ∑µ∈P+ aµλ ch ∆(µ
(2) ch L(λ

Proof. Follows immediately from Theorem 6.15 and (6.28). 


Remark 6.17. Theorem 6.16 together with Proposition 6.7 provide a complete
solution à la Kazhdan-Lusztig to the irreducible character problem in the category
On for the finite-rank basic Lie superalgebras of types gl and osp.

6.4. Kostant homology and KLV polynomials


In this section, we formulate and study the Kostant homology groups of the Lie su-
peralgebras eu− , u− and ū− , with coefficients in modules belonging to the respective
e We show that the functors T : O
categories O, O, and O. e → O and T : O e → O match
214 6. Super duality

perfectly the corresponding Kostant homology groups (and also Kostant cohomol-
ogy groups). Such matchings are then interpreted as equalities of the Kazhdan-
e
Lusztig-Vogan polynomials for the categories O, O, and O.
Some basic facts on Lie algebra homology and cohomology needed in this
section can be found in the book of Kumar [Ku].

6.4.1. Homology and cohomology of Lie superalgebras. Let L = L0̄ ⊕ L1̄ be a


vector superspace, and let T(L) be the tensor superalgebra of L. Then T(L) =
⊕∞
n=0 T (L) is a Z-graded associative superalgebra. Recall that the exterior alge-
n

bra of L is the quotient superalgebra Λ(L) := T(L)/J, where J is the homogeneous


two-sided ideal of T(L) generated by elements of the form x ⊗ y + (−1)|x||y| y ⊗ x,

where x and y are Z2 -homogeneous elements of L. Then Λ(L) = ∞ n=0 Λ L is also
n

a Z-graded associative superalgebra. For elements x1 , x2 , . . . , xk in L, the image of


the element x1 ⊗x2 ⊗· · ·⊗xk under the canonical quotient map from T k (L) to Λk (L)
will be denoted by x1 x2 · · · xk . We emphasize that the exterior algebra defined and
used in this subsection is always understood in the super sense.
Now let L = L0̄ ⊕ L1̄ be a Lie superalgebra of countable dimension. For an
L-module V we define

Cn (L,V ) = ∧n (L) ⊗V, C• (L,V ) = Cn (L,V ).
n≥0

Define the boundary operator d := ⊕n dn : C• (L,V ) → C• (L,V ) by


(6.29) dn (x1 x2 · · · xn ⊗ v) :=

∑ (−1)s+t+|xs | ∑i=1 |xi |+|xt | ∑ j=1 |x j |+|xs ||xt | [xs , xt ]x1 · · · xbs · · · xbt · · · xn ⊗ v
s−1 t−1

1≤s<t≤n
n
+ ∑ (−1)s+|xs | ∑i=s+1 |xi | x1 · · · xbs · · · xn ⊗ xs v,
n

s=1

for xi ∈ L homogeneous and v ∈ V . Here [xs , xt ] ∈ L denotes the supercommutator


and yb indicates that the term y is omitted as usual. It is standard to check that
d 2 = 0 (for Lie algebras, see Kumar [Ku]; for Lie superalgebras see e.g. [KK] for
a proof).
The kth Lie superalgebra homology group of L with coefficient in V , de-
noted by Hk (L,V ), is defined to be the kth homology group of the following chain
complex:
d d d d
· · · −→ Cn (L,V ) −→
n
Cn−1 (L,V ) −→ · · · −→
2
L ⊗V −→
1
V −→
0
0,
that is, Hk (L,V ) = Ker dk /Im dk+1 , for k ≥ 0.
Let us be more precise about the dual of a possibly infinite-dimensional space,
as it is relevant to the Lie (super)algebra cohomology to be defined below.
6.4. Kostant homology and KLV polynomials 215

Now suppose that L is a Lie superalgebra (or simply a vector superspace) which
contains a sequence of finite-dimensional subalgebras (or subspaces) Lk , for k ∈ N,
such that Lk ⊆ Lk+1 for each k and that L = ∪k Lk . Further suppose that L contains
a natural basis B such that B ∩ Lk is a basis for Lk , for all k. Then with respect to
the basis B, we may extend by zero to regard Lk∗ ⊆ Lk+1 ∗ for each k. In this way, the
restricted dual of L, denoted by L , is given by L := ∪k Lk∗ . All the examples of
∗ ∗

Lie superalgebras in this book satisfy these assumptions, and there is such a basis
for L consisting of root or weight vectors.
Fix n ∈ Z+ , and let V be a vector superspace of countable dimension. Then,
∧n Lk ⊆ ∧n Lk+1 , and the basis B of L induces a natural basis, denoted formally by
∧n B, of ∧n L which is compatible with ∧n Lk for all k. Regard Hom(∧n Lk ,V ) ⊆
Hom(∧n Lk+1 ,V ) by an extension by zero with respect to the basis ∧n B, for each k.
In this way, we define the (restricted) Hom-space

(6.30) Hom(∧n L,V ) := Hom(∧n Lk ,V ).
k

In the same sense, a (restricted) C-multilinear map f : L × · · · × L → V means the


extension by zero of a C-multilinear map f : Lk × · · · × Lk → V for some k.
The Lie superalgebra cohomology is defined as follows. Let L be a Lie su-
peralgebra satisfying the assumptions above. Take an L-module V . For n ∈ Z+ ,
let

(6.31) Cn (L,V ) = Hom(∧n L,V ), C• (L,V ) = Cn (L,V ),
n≥0

(see (6.30)). An element f ∈ Cn (L,V ) is identified with a restricted C-multilinear


n
z }| {
map f : L × · · · × L → V , which is skewsupersymmetric in the following sense: for
1 ≤ i ≤ n − 1,

f (x1 , . . . , xi , xi+1 , . . . , xn ) = −(−1)|xi |·|xi+1 | f (x1 , . . . , xi+1 , xi , . . . , xn ).

Define the coboundary operator ∂ = ⊕n≥0 ∂n : C• (L,V ) → C• (L,V ), where


(6.32)

(∂n f )(x1 , . . . , xn+1 ) :=


n+1
∑ (−1)s+1+|x |(| f |+∑ i=1 |xi |)
s−1
s
xs f (x1 , . . . , xs−1 , xbs , xs+1 , . . . .xn+1 )
s=1

+ ∑(−1)s+t+|xs | ∑i=1 |xi |+|xt | ∑ j=1 |x j |+|xs ||xt | f ([xs , xt ], x1 , . . . , xbs , . . . , xbt , . . . , xn+1 ).
s−1 t−1

s<t

It is straightforward to verify that ∂2 = 0. The kth (restricted) Lie superalgebra


cohomology group of L with coefficient in the L-module V , denoted by H k (L,V ),
216 6. Super duality

is by definition the kth cohomology group of the complex


∂−1 ∂ ∂ ∂n−1 ∂
(6.33) 0 −→ V −→
0
C1 (L,V ) −→
1
· · · −→ Cn (L,V ) −→
n
Cn+1 (L,V )−→ · · · ,
that is, H k (L,V ) := Ker ∂k /Im ∂k−1 .
Remark 6.18. If we use the cochain complex (6.31) above by interpreting the
Hom-space in the non-restricted sense to be the space of all linear maps from ∧n L
to V , the resulting nth Lie superalgebra cohomology group will be denoted by
Hn (L,V ). This version of cohomology will only appear once in this book, as
it is used mildly in the proof of Theorem 6.32. Of course, the two versions of
cohomology coincide when L is finite dimensional.

6.4.2. Kostant u− -homology and u-cohomology. Recall the Lie algebra g with
parabolic subalgebra p, Levi subalgebra l, nilradical u, and opposite radical u−
from Section 6.2.2. Let M ∈ O. Then we have M = ⊕µ∈h∗ Mµ , with dim Mµ < ∞. In
this case, the restricted dual M ∗ of M is given by M ∗ = ⊕µ∈h∗ (Mµ )∗ . Let τ : g → g
be the Chevalley automorphism of g from Section 1.1.3 in Chapter 1. We denote
M ∗ with a τ-twisted g-action by M ∨ , i.e.,
M ∨ := ⊕µ∈h∗ (Mµ )∗ ,
where the action of g is now given by (x · f )(v) = (−1)|x|·| f |+1 f (τ(x) · v). Then
M ∨ lies in O, for M ∈ O. We shall refer M ∨ the dual of M in O, and sending
each M ∈ O to M ∨ defines an exact functor, called the duality functor, on O. The
following is a summary of some standard properties of the duality functor on O, cf.
Humphreys [Hum, Theorem 3.2].
Proposition 6.19. The duality functor is a an exact and contravariant functor,
inducing a self-equivalence on the category O. For each M ∈ O, M and M ∨ have
the same character, and hence they have the same composition factor multiplicities.
In particular, we have L(λ)∨ ∼
= L(λ), for λ ∈ P+ .

Let M ∈ O. We note that u and u− are l-modules so that the chain and cochain
complexes of the Lie superalgebra u or u− with coefficients in M are semisimple
l-modules. It is a standard fact that the boundary and coboundary operators are
l-invariant, so the homology and cohomology groups of u and u− are naturally
semisimple l-modules.
Lemma 6.20. Let M ∈ O and n ∈ Z+ . As semisimple l-modules, we have
∼ Hn (u, M ∗ )∗ .
H n (u, M) =

Proof. We have the following isomorphism of l-modules for each n ≥ 0:


Hom(∧n u, M) ∼
= Hom(∧n u ⊗ M ∗ , C).
The Hom’s here are understood as usual in the restricted sense; see (6.30). Fur-
thermore, these isomorphisms are compatible with the coboundary operator d of
6.4. Kostant homology and KLV polynomials 217

the complex C• (u, M) and the coboundary operator Hom(d, C) of the complex
Hom (C• (u, M ∗ ) , C). Since the operators are also l-module homomorphisms, we
obtain the corresponding isomorphisms of the cohomology groups as l-modules.


Lemma 6.21. Let M ∈ O and n ∈ Z+ . As semisimple l-modules, we have

Hn (u, M ∗ )∗ ∼
= Hn (u− , M ∨ ).

In particular, we have Hn (u, L(λ)∗ )∗ ∼


= Hn (u− , L(λ)).

Proof. We use a superscript τ to indicate the l-action obtained by twisting the usual
l-action by the Chevalley automorphism τ. Since uτ ∼ τ
= u− and (M ∗ ) = M ∨ , we see
that
( i )τ
∧ u ⊗ M∗ ∼ = ∧i u− ⊗ M ∨ .

Furthermore, these isomorphisms of chain complexes commute with the respective


boundary operators as well. Thus, we obtain an l-module isomorphism:

Hn (u, M ∗ )τ ∼
= Hn (u− , M ∨ ).

Now since the l-modules Hn (u, M ∗ )τ and Hn (u, M ∗ )∗ are direct sums of simple l-
modules with highest weights in P+ and they have the same character, we have an
l-module isomorphism Hn (u, M ∗ )τ ∼
= Hn (u, M ∗ )∗ . The lemma follows. 

e of Proposition 6.19, Lem-


Clearly the counterparts for categories O and O
mas 6.20 and 6.21 hold with verbatim arguments.

Theorem 6.22. Let M ∈ O, M e M ∈ O, and n ∈ Z+ . We have


e ∈ O,
( )
Hn u− , M ∨ ∼ = H n (u, M) , as l-modules;
( ) ( )

Hn ū− , M ∼ = H n u, M , as l-modules;
( ) ( )
Hn e e∨ ∼
u− , M = Hn e e , as l-modules.
u, M
( ) ( )
− ∼
= H n (u, L(λ)), Hn ū− , L(λ♮ ) ∼= H n u, L(λ♮ )
In particular,
( ) Hn (u( , L(λ)) )
we have
and Hn e e θ) =
u− , L(λ ∼ Hn e e θ ) , for λ ∈ P+ .
u, L(λ

e are analogous.
Proof. We only prove the statements for O, while those for O and O
Combining Lemma 6.20 and Lemma 6.21, we obtain an l-module isomorphism
∼ H n (u, M) . This implies that Hn (u− , L(λ)) ∼
Hn (u− , M ∨ ) = = H n (u, L(λ)), thanks to
L(λ)∨ ∼
= L(λ) by Proposition 6.19. 
218 6. Super duality

6.4.3. Comparison of Kostant homology groups. We start with setting up nota-


e → O and T : O → O from (6.24). For M
tions. Recall the functors T : O e set
e ∈ O,
e e
M = T (M) ∈ O and M = T (M) ∈ O. Denote by
de: ∧(e e → ∧(e
u− ) ⊗ M e
u− ) ⊗ M,
d : ∧(u− ) ⊗ M → ∧(u− ) ⊗ M, d : ∧(ū− ) ⊗ M → ∧(ū− ) ⊗ M
e C• (u− , M), and
u− , M),
the boundary operators of the three chain complexes C• (e
− e d, and d are homomorphisms of el-, l-, and
C• (ū , M), respectively. Note that d,
l-modules, respectively.
The l-module ∧(u− ) is a direct sum of L(l, µ), µ ∈ P+ , each appearing with
finite multiplicity. As an l-module, ∧(ū− ) is a direct sum of L(l, µ♮ ), µ ∈ P+ , each
appearing with finite multiplicity. By Theorem 6.3, the el-module ∧(e u− ) is also a
direct sum of L(el, µ ), µ ∈ P , each appearing with finite multiplicity. The l-module
θ +

∧(u− ) ⊗ M , the l-module ∧(ū− ) ⊗ M, and the el-module ∧(e u− ) ⊗ M e are completely
reducible by Theorem 6.3.

Lemma 6.23. For M e∈O e and λ ∈ P+ , set M = T (M)


e ∈ O and M = T (M) e ∈ O.
Then,
( ) ( )
u− ) ⊗ M
(1) T ∧ (e e = ∧(u− ) ⊗ M, and T ∧ (e e θ ) = ∧(u− ) ⊗ L(λ).
u− ) ⊗ L(λ
[ ]
Moreover, T de = d.
( ) ( )
u− ) ⊗ M
(2) T ∧ (e e = ∧(ū− ) ⊗ M, and T ∧ (e e θ ) = ∧(ū− ) ⊗ L(λ♮ ).
u− ) ⊗ L(λ
[ ]
Moreover, T de = d.

Proof. We will only prove (1), as the argument for (2) is parallel.
( )
It follows from eu− ∩ g = u− and the definition of T that T ∧ (e u− ) = ∧(u− ).
Now, since all modules involved have weights of the ( form ∑i<0) ai εi + ∑r>0 br εr
(possibly modulo dΛ0 ) with br ∈ Z+ , we see that T ∧ (e u− ) ⊗ Me and ∧(u− ) ⊗ M
( )
have the same character. Hence, it follows from T ∧ (e u− ) ⊗ Me ⊇ ∧(u− ) ⊗ M that
( ) ( )
T ∧ (e e = ∧(u− ) ⊗ M. By Theorem 6.15, T L(λ
u− ) ⊗ M e θ ) = L(λ), and so we
( )
have T ∧ (e e θ ) = ∧(u− ) ⊗ L(λ).
u− ) ⊗ L(λ
¿From the definitions of d,e d, and d in (6.29), we have d(v) e = d(v) for all
[ ]
v ∈ ∧(u− ) ⊗ M and d(w)e = d(w) for all v ∈ ∧(ū− ) ⊗ M. Thus, T de = d and
[ ]
T de = d. This completes the proof of (1). 

Proposition 6.24. For M e set M = T (M)


e ∈ O, e ∈ O and M = T (M)
e ∈ O. Suppose
∼ ⊕ e e
that ∧(e−
u )⊗M e = µ∈P+ L(l, µ )
θ m(µ) , as l-modules. Then

(1) ∧(u− ) ⊗ M ∼= µ∈P+ L(l, µ)m(µ) , as l-modules.
∼ ⊕ + L(l, µ♮ )m(µ) , as l-modules.
(2) ∧(ū− ) ⊗ M = µ∈P
6.4. Kostant homology and KLV polynomials 219

Proof. We only prove (1) as (2) is similar. By Lemma 6.23, we have a surjec-
tive l-module homomorphism T : ∧(e e → ∧(u− ) ⊗ M. By Lemma 6.12,
u− ) ⊗ M
( )
T L(el, µθ ) = L(l, µ), for all µ ∈ P+ . Now (1) follows from the exactness of T . 

By Lemma 6.23 and (6.25), we have the following commutative diagram:


de
e −−→ ∧n (e de
e −−→ ∧n−1 (e e −−→ · · · de
u− ) ⊗ M
· · · −−→ ∧n+1 (e u− ) ⊗ M u− ) ⊗ M
  
(6.34) T n+1 − e T n − e T n−1 − e
y ∧ (eu )⊗M y ∧ (eu )⊗M y ∧ (eu )⊗M
d d d
· · · −−→ ∧n+1 (u− ) ⊗ M −−→ ∧n (u− ) ⊗ M −−→ ∧n−1 (u− ) ⊗ M −−→ · · ·
Thus T induces an l-homomorphism from Hn (e e to Hn (u− ; M). Similarly, T
u− ; M)
e to Hn (ū− ; M). We have the following
u− ; M)
induces an l-homomorphism from Hn (e
more precise result.
Theorem 6.25. For M e set M = T (M)
e ∈ O, e ∈ O and M = T (M)
e ∈ O. Let n ∈ Z+ .
Then,
(1) T (Hn (e e ∼
u− ; M)) = Hn (u− ; M), as l-modules.
(2) T (Hn (e e ∼
u− ; M)) = Hn (ū− ; M), as l-modules.

Proof. We shall only prove (1), as the argument for (2) is similar.
We regard ∧(u− ) ⊗ M ⊆ ∧(e e By Lemma 6.23 and (6.34), we have
u− ) ⊗ M.
( )
e = Ker de∩ ∧(u− ) ⊗ M = Ker d,
T (Ker d)
and ( )
e = Im de∩ ∧(u− ) ⊗ M = Im d.
T (Im d)
Since T is an exact functor, we have
( )
⊕ ⊕
T u− , M)
Hn (e e (Im d)
e = T (Ker d)/T e = Ker d/Im d = Hn (u− , M).
n≥0 n≥0
This completes the proof. 
Corollary 6.26. Let M e λ ∈ P+ , and n ∈ Z+ . Furthermore, set T (M)
e ∈ O, e = M and
e
T (M) = M. We have
(1) T (H n (e e ∼
u, M)) = H n (u, M), and thus T (H n (e e θ ))) ∼
u, L(λ = H n (u, L(λ)), as
l-modules;
(2) T (H n (e e ∼
u, M)) = H n (u, M), and thus T (H n (e e θ ))) ∼
u, L(λ = H n (u, L(λ♮ )), as
l-modules.
(3) Moreover, we have
( ( ))
Homl (L(l, µ), H n (u, M)) ∼
= Homel L(el, µθ ), H n e e
u, M
( ( ))

= Homl L(l, µ♮ ), H n ū, M ,
220 6. Super duality

and so in particular
( ( ))
Homl (L(l, µ), H n (u, L(λ))) ∼
= Homel L(el, µθ ), H n e e θ)
u, L(λ
( ( ))

= Homl L(l, µ♮ ), H n ū, L(λ♮ ) .

Proof. We shall prove (1). We have the following isomorphisms:


( ) ( ) ( )
T H (en e ∼
u, M) = T Hn (e e ∨ ∼ e
u− , M ) = Hn u− , T (M )∨

( ) ( )
∼ e ∨ ∼ n
= Hn u− , T (M) = H u, T (M) . e

The first and the last isomorphisms follow from Theorem 6.22. The second iso-
morphism is due to Theorem 6.25(1), while the third isomorphism is due the com-
patibility of the functors T and ·∨ . This proves the first statement of (1). Setting
Me = L(λ
e θ ) and using Theorem 6.15 we get the second statement of (1). The first
( )
“∼
=” in (3) follows from (1) and that T L(el, µθ ) = L(l, µ) (see Lemma 6.12).
The argument for (2) is similar to (1), and the second “∼ =” in (3) follows from
(2) and Lemma 6.12. 

6.4.4. Kazhdan-Lusztig-Vogan (KLV) polynomials. For a basic Lie superalge-


bra, the Weyl group of its even subalgebra does not completely control the linkage,
as we have seen in Section 2.2, Chapter 2. So we cannot expect its associated
Hecke algebra and corresponding Kazhdan-Lusztig polynomials to play a funda-
mental role for the category O. We instead adopt an approach via homological
algebra.
Definition 6.27. The (parabolic) Kazhdan-Lusztig-Vogan polynomials (KLV
e are defined as follows: for
polynomials, for short) in the categories O, O and O
µ, λ ∈ P ,
+

∞ ( )
ℓµλ (q) := ∑ (−q)−n dim Homl L(l, µ), Hn (u− , L(λ)) ,
n=0
∞ ( )
ℓµ♮ λ♮ (q) := ∑ (−q)−n dim Homl L(l, µ♮ ), Hn (ū− , L(λ♮ )) ,
n=0
∞ ( )
ℓeµθ λθ (q) := ∑ (−q)−n dim Homel L(el, µθ ), Hn (e
u− , L(λθ )) .
n=0

Recall from (6.28) that ch L(λ) = ∑µ∈P+ aµλ ch ∆(µ), for λ ∈ P+ . By definition,
ℓµλ (q), ℓµ♮ λ♮ (q), ℓeµθ λθ (q) are power series in q−1 .

Proposition 6.28. The ℓµλ (q), ℓµ♮ λ♮ (q), ℓeµθ λθ (q) are polynomials in q−1 . Moreover,
we have ℓµλ (1) = aµλ .
6.4. Kostant homology and KLV polynomials 221

Proof. We will only show that ℓµλ (q) is a polynomial in q−1 , and similar arguments
can be applied to ℓµ♮ λ♮ (q), ℓeµθ λθ (q).
Take λ ∈ P+ , and recall that Cn (u− , L(λ)) = ∧(ū− ) ⊗ L(λ). Since Cn (u− , L(λ))
has finite-dimensional weight spaces, the module L(l, µ) appears with finite multi-
plicity in Cn (u− , L(λ)), for each n. Also observe that, for a given weight µ ∈ P+ ,
µ cannot be a weight of Cn (u− , L(λ)), for n ≫ 0. This proves that ℓµλ (q) is a
polynomial in q−1 .
We can now evaluate the polynomial ℓµλ (q) at q = 1. Applying the Euler-
Poincaré principle to the chain complex C• (u− , L(λ)) of h-modules, we have
ch L(λ) ∑ (−1)n ch ∧n (u− )
n≥0

= ∑ (−1)n ch Hn (u− , L(λ))


n≥0
( )
= ∑ (−1)n ∑ dim Homl L(l, µ), Hn (u− , L(λ)) ch L(l, µ)
n≥0 µ∈P+

= ∑ ℓµλ (1)ch L(l, µ).


µ∈P+

The above identity can be rewritten as


(6.35) ch L(λ) = ∑ ℓµλ (1)ch ∆(µ),
µ

using the following character formula of ∆(µ):


ch L(l, µ)
ch∆(µ) = n ch ∧n (u− )
,
∑n≥0 (−1)
The equality ℓµλ (1) = aµλ now follows by a comparison of (6.35) with (6.28) and
noting the linear independence of ch ∆(µ). 

The following reformulation of Theorem 6.25 is a generalization of the char-


acter formulas in Theorem 6.16.

Theorem 6.29. For µ, λ ∈ P+ we have ℓµλ (q) = ℓeµθ λθ (q) = ℓµ♮ λ♮ (q).

6.4.5. Stability of KLV polynomials. For n ∈ N recall the category On of finite-


dimensional gn -modules from Section 6.2.3. We have denoted the corresponding
finite-dimensional Levi subalgebra by ln , and we shall denote the corresponding
finite-dimensional nilradical and opposite radical by un and u− n , respectively. Re-
call also that the set {λ ∈ P+ |ℓ(+ λ) ≤ n} is denoted by Pn+ , while Ln (λ) ∈ On
denotes the irreducible gn -module of highest weight λ ∈ Pn+ . We shall also write
L∞ (λ) ≡ L(λ), P∞+ ≡ P+ et cetera. For k ∈ N ∪ ∞ with k > n, recall from Sec-
tion 6.2.5 that the truncation functor trkn : Ok → On is a gn -module homomorphism.
222 6. Super duality

Comparing the characters we obtain that, for λ ∈ Pk+ ,


{
Ln (ln , λ), if λ ∈ Pn+ ,
(6.36) trkn (Lk (lk , λ)) =
0, otherwise.
The following lemma is easy.
Lemma 6.30. For n, k ∈ N ∪ ∞ with k > n, we have the following ln -module iso-
morphisms.
( )
(1) trkn ∧(u− −
k ) = ∧(un ).
( )
(2) trkn ∧(u− −
k ) ⊗ Lk (λ) = ∧(un ) ⊗ Ln (λ).

We now employ the same method as in Section 6.4.3 to study the effect of the
truncation functors trkn on the corresponding u− k -homology groups with coefficients
in the module Lk (λ), for λ ∈ Pk . Let d : ∧(u−
+ (n) −
n ) ⊗ Ln (λ) → ∧(un ) ⊗ Ln (λ) denote
the corresponding boundary operator of the chains of the finite-dimensional Lie
algebra u−n with coefficients in Ln (λ). It is evident that d |∧(u−
(k)
n )⊗Ln (λ)
= d (n) , and
furthermore we have d (n) trkn = trkn d (k) . The proof of Theorem 6.25 can be adapted
to prove the following.
Theorem 6.31. Let λ ∈ Pk+ and i ∈ Z+ . We have
{
k
( −
) Hi (u−
n , Ln (λ)) , if λ ∈ Pn+ ,
trn Hi (uk , Lk (λ)) =
0, otherwise.

For λ, µ ∈ Pn+ , the expression


∞ ( )
ℓµλ (q) := ∑ (−q)−i dim Homln L(ln , µ), Hi (un − , Ln (λ)) ,
(n)

i=0
equals the parabolic Kazhdan-Lusztig polynomial of the Lie algebra gn via Vogan’s
homological interpretation. Theorem 6.31 and (6.36) imply that the Kazhdan-
Lusztig polynomials for finite-rank semisimple or reductive Lie algebras admit a
remarkable stability, i.e., we have
(n)
ℓµλ (q) = ℓµλ (q), λ, µ ∈ Pn+ .

Similar argument leads to analogous stability of the KLV polynomials in cate-


gories On and Oe n , which coincide for n ≫ 0 with ℓe θ θ (q) and ℓ ♮ ♮ (q), respectively.
µ λ µλ
Theorem 6.29 then further allows us to identify suitably these KLV polynomials for
finite-rank Lie (super)algebras gn , gn , and e
gn .

6.5. Super duality as an equivalence of categories


In this section, we establish super duality, which is formulated as an equivalence of
the two module categories O and O. More precisely, we shall show that the functors
6.5. Super duality as an equivalence of categories 223

T :Oe → O and T : O e → O are equivalences of categories. The category O, as a


variant of BGG category for a classical Lie algebra, has been well understood, and
super duality leads to new insight on the module category Oe for Lie superalgebras.
Some basic facts on the category theory and homological algebra needed in
this section can be found in the book of Mitchell [Mi].

6.5.1. Extensions à la Baer-Yoneda. In this subsection, we shall review the Baer-


Yoneda extension Ext1 in a general abelian category C. This is needed later on for
e since there may not be enough projective objects. A reader
the categories O, O, O,
can also accept the exact sequence (6.40) and then safely skip this subsection. We
shall omit most of the proofs, and refer to Mitchell [Mi, Chapter VII] for details.
For two objects A,C ∈ C let E, E ′ be short exact sequences of the form
i j
(6.37) E: 0 −→ A −→ B −→ C −→ 0
i′ j′
E′ : 0 −→ A −→ B′ −→ C −→ 0.
We say that E ∼ E ′ if there exists f : B → B′ such that i′ = f ◦i and j′ ◦ f = j. By the
5 lemma f is an isomorphism, and so ∼ is indeed an equivalence relation. The set
of extensions of degree 1 of C by A is by definition the set of equivalence classes
of exact sequences of the form (6.37) with respect to the relation ∼, and will be
denoted by Ext1C (C, A). We shall write E ∈ Ext1C (C, A) for the class represented by
the exact sequence E by abuse of notation.
Given an object C′ ∈ C, a morphism γ : C′ → C, and an element E ∈ Ext1C (C, A),
we have the following diagram:
C′

γ
y
E: 0 −−−−→ A −−−−→ B −−−−→ C −−−−→ 0
Then, there exists a unique element E ′ ∈ Ext1C (C′ , A) such that the following dia-
gram is commutative:
E′ : 0 −−−−→ A −−−−→ B′ −−−−→ C′ −−−−→ 0
 
β γ
y y
E : 0 −−−−→ A −−−−→ B −−−−→ C −−−−→ 0
Indeed, B′ is simply the pullback of γ and the morphism B → C. The morphism
A → B′ is then uniquely determined by the commutativity of the left square and
the exactness of the first row of the diagram. We denote E ′ ∈ Ext1 (C′ , A) by Eγ.

Letting γ vary, we obtain a map E· : HomC (C′ ,C) → Ext1C (C′ , A) defined by γ → Eγ.
On the other hand, letting E vary, one shows that a morphism γ : C′ → C induces
a well-defined map ·γA : Ext1C (C, A) → Ext1C (C′ , A). That is, for a fixed object A in
224 6. Super duality

C Ext1C (·, A) is a contravariant functor. With this notation ·γA = Ext1C (γ, A) we shall
write ·γA = ·γ, when A is clear from the context.
Dually, suppose we are given an object A′ ∈ C, a morphism α : A → A′ , and an
element E ∈ Ext1C (C, A), so that we have the following diagram:
E: 0 −−−−→ A −−−−→ B −−−−→ C −−−−→ 0

α
y
A′
We obtain uniquely an element E ′ ∈ Ext1C (C, A) such that the following diagram is
commutative:
E : 0 −−−−→ A −−−−→ B −−−−→ C −−−−→ 0
 
α β
y y
E ′ : 0 −−−−→ A′ −−−−→ B′ −−−−→ C −−−−→ 0
The resulting element E ′ is denoted by αE. Fixing E we thus obtain a map
·E
·E : HomC (A, A′ ) → Ext1C (C, A′ ) given by α → αE. On the other hand, letting
E vary, we can show that a fixed morphism α : A → A′ induces a well-defined map

C α· : ExtC (C, A) → ExtC (C, A ) so that ExtC (C, ·) is a covariant functor. With this
1 1 1

notation we have C α· = ExtC (C, α). Thus, Ext1C (·, ·) gives a bi-functor on C. We
1

shall write C α· = α·, when C is clear from the context. For an extension E, with
morphisms α and γ as above, we have (αE)γ = α(Eγ).
We define an abelian group structure on Ext1C (C, A). For A ∈ C, we define a
morphism ∆ = ∆A : A → A ⊕ A, ∆(a) := (a, a), and a morphism ∇ = ∇A : A ⊕ A → A
by ∇(a, a′ ) := a + a′ . Given two elements E, E ′ ∈ Ext1C (C, A) we let E ⊕ E ′ be the
obvious element in Ext1C (C ⊕C, A ⊕ A). To define the group operation we consider
(6.38) ∇A (E ⊕ E ′ )∆C ,
which is an exact sequence of the form (6.37). Define E + E ′ to be the ele-
ment in Ext1C (C, A) corresponding to (6.38). It can be shown that the operation
+ : Ext1C (C, A) × Ext1C (C, A) → Ext1C (C, A) is well-defined, abelian, and associa-
tive. Furthermore, it can be shown that the equivalence class of the split exact
sequence E0 is an additive identity and that (−1A )E is the additive inverse of E.
For A,C ∈ C, it is convenient to set Ext0C (C, A) := HomC (C, A).
Let X be an object in C. Applying the functor HomC (X, ·) to the exact sequence
(6.37) gives us an exact sequence
(6.39) 0 → HomC (X, A) → HomC (X, B) → HomC (X,C).
Now for each γ ∈ HomC (X,C), we obtain Eγ ∈ Ext1C (X, A). It can be shown that
the map E· extends (6.39) to the following exact sequence:
(6.40) 0 −→HomC (X, A) −→ HomC (X, B) −→ HomC (X,C)
6.5. Super duality as an equivalence of categories 225

E· i· j·
−→Ext1C (X, A) −→ Ext1C (X, B) −→ Ext1C (X,C).

6.5.2. Relating extensions in O, O, and O.e Now we return to the categories O, O,


e
and O, of g-, g-, and e
g-modules, respectively. Let A, e and let Ee ∈ Ext1 (C,
e Ce ∈ O e A).
e
e
O
Then we have by Section 6.5.1 an exact sequence in O e of the form
Ee : e −→ Be −→ Ce −→ 0.
0 −→ A
Applying the exact functors T and T , we obtain exact sequences in O and O, re-
spectively. It can be easily seen that this way the functors T and
( T define )natu-
1 e e 1 e T (A)
e and
ral homomorphisms of abelian groups from Ext e (C, A) to ExtO T (C),
( ) O
ExtO1 e T (A)
T (C), e , respectively. The respective elements are denoted by T [E]
e and
e
T [E].
Recall the Lie algebra g with parabolic subalgebra p, Levi subalgebra l and
nilradical u from Section 6.2.2. We have the following relative Koszul resolution
for the trivial p-module, which is a straightforward superalgebra generalization of
the standard one (see Knapp-Vogan [KV, II.7]):
∂ ∂ ε
(6.41) · · · −→Ck −→
k
Ck−1 −→ · · · −→
1
C0 −→ C −→ 0.
Here Ck := U(p) ⊗U(l) ∧k (p/l) is a p-module with p acting on the left of the first
factor, and ε is the augmentation map from U(p) to C. The p-homomorphism ∂k is
given by
(6.42)

∂k (a ⊗ x1 x2 · · · xk )

∑ (−1)s+t+|xs | ∑i=1 |xi |+|xt | ∑ j=1 |x j |+|xs ||xt | a ⊗ [xs , xt ]x1 · · · b


xs · · · b
s−1 t−1
:= xt · · · xk
1≤s<t≤k
k
+ ∑ (−1)s+1+|xs | ∑i=1 |xi | axs ⊗ x1 · · · b
s−1
xs · · · xk .
s=1
Here a ∈ U(p) and xi are homogeneous elements in p and xi denotes xi + l in p/l.
For λ ∈ P+ , we extend the irreducible l-module L(l, λ) trivially to a p-module.
For k ≥ 0, Dk := Ck ⊗L(l, λ) is a p-module by a diagonal p-action. Tensoring (6.41)
with L(l, λ) gives us the following exact sequence of p-modules, where ∂k ⊗ 1 is
simply denoted by ∂k again for k > 0, and ∂0 := ε ⊗ 1:
∂ ∂ ∂
(6.43) · · · −→ Dk −→
k
Dk−1 −→ · · · −→
1
D0 −→
0
L(l, λ) −→ 0.
Theorem 6.32. Let λ ∈ P+ , V ∈ O, Ve ∈ O, e and V ∈ O. We have, for i = 0, 1,
( )
Homl L(l, λ), H i (u,V ) ∼ = ExtiO (∆(λ),V ),
( )
u, Ve ) ∼
Homel L(el, λθ ), H i (e e θ ), Ve ),
= ExtiOe (∆(λ
226 6. Super duality

( )
Homl L(l, λ♮ ), H i (u,V ) ∼ i
= ExtO (∆(λ♮ ),V ).

Proof. We shall only prove the first isomorphism for l for i = 0, 1, and the remain-
ing two isomorphisms can be proved in an analogous fashion.
Let V ∈ O. Recall the (unrestricted) cohomology group Hi (u,V ) from Re-
mark 6.18 in contrast to the (restricted) cohomology group H i (u,V ) used mostly in
this chapter. Let Hi (u,V )ss be the maximal h-semisimple submodule of H i (u,V ).
We have the following
∼ H i (u,V ).
Claim. Hi (u,V )ss =
To see this, observe by definition that we have a natural injective map from
ι
H i (u,V ) → Hi (u,V )ss . Since the coboundary operator ∂i is an h-homomorphism
and Hom(∧ (u),V ) is a direct product of finite-dimensional h-weight spaces, Ker ∂i
i

and Im ∂i are also direct products of finite-dimensional h-weight spaces. Now given
x ∈ Ker ∂i representing a cohomology class in Hi (u,V )ss , there exists an element
y ∈ Im ∂i−1 such that x − y ∈ Ci (u,V ) ∩ Ker ∂i . We have then ι(x − y) ∈ x + Im ∂i−1 ,
and hence ι is also a surjection. This proves the claim.
Now since L(l, λ) is h-semisimple and Hi (u,V ) is a direct product of finite-
dimensional h-weight spaces, it follows by the claim that
( ) ( )
(6.44) Homl L(l, λ), Hi (u,V ) ∼ = Homl L(l, λ), H i (u,V ) .

which allows us to freely switch between H i and Hi below.


Unraveling the definition, we have H 0 (u,V ) = V u , the u-invariants of V . Thus,

Homl (L(l, λ), H 0 (u,V )) ∼


= Homl (L(l, λ),V u )
∼ Homp (L(l, λ),V ) ∼
= = HomO (∆(λ),V ).

where the last “∼ =” is due to Frobenius reciprocity. This proves the isomorphism
for l with i = 0 in the theorem.
( )
It remains to show that Homl L(l, λ), H 1 (u,V ) ∼ = Ext1O (∆(λ),V ). We shall
construct the isomorphism explicitly.
( )
First, take f¯ ∈ Homl L(l, λ), H 1 (u,V ) . Recall H 1 (u,V ) = Ker d1 /Im d0 , with
Ker d1 ⊆ Hom(u,V ). Let f ∈ Homl (L(l, λ), Hom(u,V )) ∼ = Homl (u ⊗ L(l, λ),V )
be a representative,
( which, by
) Frobenius reciprocity, is regarded as an element in
Homp Indpl (u ⊗ L(l, λ)) ,V .
Applying the exact functor U(g) ⊗U(p) - to (6.43), we obtain an exact sequence

( ) ∂′2 ∂′1
U(g) ⊗U(l) ∧2 (p/l) ⊗ L(l, λ) −→U(g) ⊗U(l) (p/l ⊗ L(l, λ)) −→
∂′
U(g) ⊗U(l) L(l, λ) −→
0
∆(λ) −→ 0,
6.5. Super duality as an equivalence of categories 227

where ∂′i denotes the differentials induced from ∂i . This gives rise to the following
diagram
∂′
U(g) ⊗U(l) (p/l ⊗ L(l, λ)) −−−1−→ U(g) ⊗U(l) L(l, λ) −−−−→ ∆(λ) −−−−→ 0

f
y
V
Taking the pushout E f of f and ∂′1 , we obtain the following commutative diagram
U(g) U(g)
IndU(l) (p/l ⊗ L(l, λ)) −−−−→ IndU(l) L(l, λ) −−−−→ ∆(λ) −−−−→ 0
 
f 
y y
α
V −−−−→ Ef −−−−→ ∆(λ) −−−−→ 0
Since f represents a cocycle, we have f ◦∂′2 = 0. This implies that α is an injection.
Now since V and ∆(λ) lie in O, so does E f , and hence we obtain an element in
Ext1O (∆(λ),V ). One verifies in a straightforward manner that if f is a coboundary,
then E f ∼
= V ⊕ ∆(λ), and hence the map f¯ → E f¯ is well-defined.
Conversely, suppose that we have an extension of the form
0 −→ V −→ E −→ ∆(λ) −→ 0.
Note that E ∈ O, and hence there exists an l-submodule of E that is mapped iso-
morphically to L(l, λ) ⊆ ∆(λ). We denote this l-module also by L(l, λ) and note
that uL(l, λ) ⊆ V . This allows us to define an element fE ∈ Homl (u ⊗ L(l, λ),V ) ∼
=
Homl (L(l, λ), Hom(u,V )) by fE (u ⊗ w) = uw. Now fE corresponds to a cocycle in
Homl (L(l, λ), Hom(u,V )), because V is a submodule of the u-module E. Also, one
sees that if E is the trivial extension, then fE must be a coboundary. Thus, sending
E to f¯E gives us a well-defined map Ext1O (∆(λ),V ) → Homl (L(l, λ), H1 (u,V )) ∼ =
Homl (L(l, λ), H 1 (u,V )); see (6.44).
It can be
( checked1 that these
) two1 maps are inverses to each other. Hence, we
have Homl L(l, λ), H (u,V ) ∼ = ExtO (∆(λ),V ). 

e
It makes sense to apply T and T to the short exact sequences (i.e., Ext1e ) in O.
O

Corollary 6.33. Let λ, µ ∈ P+ .


We have, for i = 0, 1,
( ( )) ( )
e θ ), L(µ
T ExtiOe ∆(λ e θ ) = ExtiO ∆(λ), L(µ) ,
( ( )) ( )
e θ ), L(µ
T ExtiOe ∆(λ e θ ) = Exti ∆(λ♮ ), L(µ♮ ) .
O

Proof. We shall only prove the identity for T , as the proof for T is analogous.
By Theorem 6.32, for i = 0, 1, we have
( ) ( )
ExtiO ∆(λ), L(µ) ∼ = Homl L(l, λ), H i (u, L(µ)) ,
228 6. Super duality

( ) ( )
(6.45) ExtiOe ∆(λ e θ) ∼
e θ ), L(µ = Homel L(el, λθ ), H i (e e θ )) .
u− , L(µ
( )
Recall that T L(el, λθ ) = L(l, λ) by Lemma 6.12, and that T (H n (e e θ ))) ∼
u, L(λ =
n
H (u, L(λ)) by Corollary 6.26. Now the identity for T in the corollary follows
by applying the functor T to (6.45) and the naturality of T . 
f e f . The following proposition is standard and well-
6.5.3. Categories O f , O and O
known. For a proof the reader may consult Kumar [Ku, Lemma 2.1.10]. Recall Π
denotes the standard fundamental system of g.
Proposition 6.34. Let M ∈ O. Then there exists a (possibly infinite) increasing
filtration of g-modules
(6.46) 0 = M0 ⊆ M1 ⊆ M2 ⊆ · · · ⊆ Mi−1 ⊆ Mi ⊆ · · ·
such that

(1) M = i≥0 Mi ;
(2) Mi /Mi−1 is a highest weight module of highest weight νi with νi ∈ P+ ,
for i ≥ 1;
(3) the condition νi − ν j ∈ ∑α∈Π Z+ α implies that i < j;
(4) for any weight µ of M, there exists an r ∈ N such that (M/Mr )µ = 0.
e
e ∈ O.
Similar statements hold for M ∈ O and M

Let O f denote the full subcategory of O consisting of finitely generated U(g)-


f
modules. The categories O and O e f are defined in a similar fashion.

Corollary 6.35. Let M ∈ O. Then M ∈ O f if and only if there exists a finite increas-
ing filtration 0 = M0 ⊆ M1 ⊆ M2 ⊆ · · · ⊆ Mk = M of g-modules such that Mi /Mi−1
is a highest weight module of highest weight νi with νi ∈ P+ , for 1 ≤ i ≤ k. Similar
statements hold for M ∈ O and M e
e ∈ O.

Proof. If M is a g-module with a finite filtration of the form (6.46), then clearly
M is finitely generated. Conversely, suppose that u1 , u2 , . . . , un are vectors in M of
weights ν1 , ν2 , . . . , νn generating M over U(g). By Proposition 6.34(4), M has a
(possibly infinite) filtration of the form (6.46) such that there exists r ∈ N with Mr
containing all vectors in M of weights ν1 , ν2 , . . . , νn . Thus ui ∈ Mr , for 1 ≤ i ≤ n,
and hence Mr = M. The proofs for e g and g are analogous and omitted. 

6.5.4. Lifting highest weight modules. The following proposition is the converse
of Proposition 6.14.
Proposition 6.36. (1) Suppose that V (λ) is a highest weight g-module of
g-module Ve (λθ ) of
highest weight λ ∈ P+ . Then there is a highest weight e
highest weight λ such that T (Ve (λ )) = V (λ).
θ θ
6.5. Super duality as an equivalence of categories 229

(2) Suppose that U(λ♮ ) is a highest weight g-module of highest weight λ♮


with λ ∈ P+ . Then there is a highest weight e e θ ) of highest
g-module U(λ
e θ )) = U(λ♮ ).
weight λθ such that T (U(λ

Proof. We shall only prove (1), as (2) is similar.


Let W be the kernel of the surjective g-homomorphism from ∆(λ) to V (λ).
Now Theorem 6.15 says that T (∆(λ e θ )) = ∆(λ). By the exactness of the functor T ,
it suffices to prove that W lifts to a submodule W e θ ) such that T (W
e of ∆(λ e ) = W.
We recall the standard Borel subalgebra b and standard fundamental system
Π for g. Recall the sequence of Borel subalgebras e bc (n) and corresponding fun-
damental systems Π e (n) from Section 6.3.1, and that Πn = Π(k) ⊔ Π(Tn ) from
c

Section 6.1.1. We have the following.


Claim. Let vµ be a b-singular vector of weight µ ∈ P+ in a e g-subquotient of
e ). Then vµ is a e
∆(λ θ bc (n)-singular vector, for n ≫ 0.
This claim can be seen as follows. There exists n ≫ 0 such that λ − µ is a
non-negative sum of even simple roots lying in Πn , and such that every weight in
e θ ) is of the form λ − ν, where ν is a non-negative sum of simple roots in Π
∆(λ e c (n).
e θ ).
e c (n) \ Πn , then µ + α cannot a weight in ∆(λ
Therefore, if α is a simple root in Π
This proves the claim.
There is an increasing filtration of g-modules for W , 0 = W0 ⊆ W1 ⊆ W2 ⊆ · · · ,
satisfying the properties of Proposition 6.34. For each i > 0, let wi be a weight
vector of weight νi in Wi such that wi +Wi−1 is a non-zero b-highest weight vector
of Wi /Wi−1 . Now by Theorem 6.15, ∆(λ e θ ) = ⊕µ∈P+ L(el, µθ )m(µ) if and only if
⊕ e θ ), there is a highest weight
∆(λ) = µ∈P+ L(l, µ)m(µ) . Then, regarding ∆(λ) ⊆ ∆(λ
vector w ei of the el-module U(el)wi ∼= L(el, νθi ) with respect to the Borel subalgebra
e
b ∩el, for each i. Let Wei be the eg-submodule of ∆(λ e θ ) generated by w e1 , w
e2 , . . . , w
ei
e
for i ≥ 1, and set W0 = 0.
We will prove by induction on i that T (W ei ) = Wi , for all i. The case i = 0 is
trivial. Assume that i ≥ 1 and T (W ei−1 ) = Wi−1 . By construction, wi + W ei−1 is a b-
singular vector in Wi /Wi−1 . It follows by the claim above that wi + Wi−1 is a e
e e e bc (n)-
singular vector of weight νi for n ≫ 0. By (6.21), we have e b=e b ∩el + e
u⊆e b ∩el +
e c
b (n). Thus, by the construction of w ei above, w ei + W e
ei−1 is a b-highest weight vector
e e −
of Wi /Wi−1 . Now U(u )L(l, νi ) = Wi /Wi−1 and U(e u− )L(el, νθi ) = Wei /W
ei−1 , and thus
T (Wei /W
ei−1 ) = Wi /Wi−1 . This together with the inductive assumption implies that
e
T (Wi ) = Wi .

e=
Finally, setting W e we obtain that T (W
i≥1 Wi ,
e ) = W. 

6.5.5. Super duality and strategy of proof. Recall the functors T and T from
Section 6.3.3. The following is a main result of this chapter.
230 6. Super duality

Theorem 6.37. e → O is an equivalence of categories.


(1) T : O
e → O is an equivalence of categories.
(2) T : O
(3) The categories O and O are equivalent.

The equivalence of categories in Theorem 6.37(3) is called super duality.


Define an equivalence relation ∼ on e h∗ by letting µ ∼ ν if and only if µ − ν lies
e of e
in the root lattice ZΦ g. For each such equivalence class [µ], fix a representative
[µ] ∈ e
o ∗
h and declare [µ]o to have Z2 -grading 0̄. For ε = 0̄, 1̄, set
{ ⟨ ⟩ }
e
h∗ε = µ ∈ e h∗ | ∑ µ − [µ]o , Ebr ≡ ε (mod 2) , for x = a, b, c, d,
r∈1/2+Z+
{ m ⟨ ⟩ ⟨ ⟩ }
e
h∗ε = µ ∈ e
h∗ | ∑ µ − [µ]o , Eb−i + ∑ µ − [µ]o , Ebr ≡ ε (mod 2) , for x = b• .
i=1 r∈N

e is a semisimple e ⊕
Recall that Ve ∈ O h-module with Ve = γ∈eh∗ Veγ . Then Ve gives rise
g-module Ve ′ :
to a Z2 -graded e
⊕ ⊕
(6.47) Ve ′ := Ve0̄ Ve1̄ , Veε := Veµ (ε = 0̄, 1̄),
µ∈e
h∗ε

which is compatible with the Z2 -grading on e


g.
e e e and O
We define O and O to be the full subcategories of O
0̄ f ,0̄ e f , respectively,
consisting of objects equipped with Z2 -gradation given by (6.47). Note that the
morphisms in O e 0̄ and O
e f ,0̄ are of degree 0̄. It is clear that for a given Ve ∈ O, e Ve ′
defined in (6.47) is isomorphic to Ve in O. e It then follows by definition that the
e e
two categories O and O are equivalent. Similarly, O
0̄ e f and Oe f ,0̄ are equivalent
categories.
0̄ f ,0̄
We can now analogously define O0̄ , O f ,0̄ , O and O to be the respective full
f
subcategories of O, O f , O and O consisting of objects equipped with Z2 -gradation
∼ O, O0̄ =
(6.47). Similarly, O0̄ = ∼ O, and also O f ,0̄ =
∼ O f , O f ,0̄ =
∼ Of .

Since O0̄ =∼ O, O0̄ ∼ = O and O e 0̄ ∼ e it suffices to prove Theorem 6.37(1) for


= O,
T :Oe → O and T : O
0̄ 0̄ e → O . In order to keep notations simple we will from
0̄ 0̄

now on drop the superscript 0̄ and use O, O, e O f and O e f to denote the respective
0̄ e 0̄ e f ,0̄ . Henceforth, when we write ∆(λ ef,
categories O , O
f ,0̄
, O and O e θ ), L(λ
e θ) ∈ O
λ ∈ P+ , we will mean the corresponding modules equipped with the Z2 -gradation
(6.47). Similar convention applies to ∆(λ♮ ) and L(λ♮ ).
In the remainder of this subsection, we outline the strategy of the proof of
Theorem 6.37. The detailed proof will be carried out in the next subsection. We
shall restrict our proof for T entirely, since the arguments for the functors T and
T are completely parallel. First, we study the relation between the morphisms
6.5. Super duality as an equivalence of categories 231

e and O under T . In particular, we show in Proposition 6.42


in the categories O
that T induces an isomorphism between morphisms of the two categories. In con-
junction with Theorem 6.32 that relates extensions with homology groups, this
isomorphism is then used to show that T induces an isomorphism between the first
extension groups of the two categories. These, together with well-known facts
from homological algebra, imply that T is an equivalence of categories.

6.5.6. The proof of super duality. We shall freely use the following short-hand
notations without further notices. Set M = T (M) e and M = T (M), e for M
e ∈ O.e Sim-
e ′ e ′′ e e
ilar convention applies to the images of T and T of M , M , N ∈ O. In addition, set
V (λ) = T (Ve (λθ )) and V (λ♮ ) = T (Ve (λθ )), for a highest weight eg-module Ve (λθ ) of
highest weight λθ with λ ∈ P+ ; see Proposition 6.14 for a justification of notations.
e and let
e ∈ O,
Lemma 6.38. Let N
e
Ee : e ′ −→ e e ′′ −→ 0
i
(6.48) 0 −→ M M−→M

be an exact sequence of e e
g-modules in O.
(1) The exact sequence (6.48) induces the following commutative diagram
with exact rows. (We will use subscripts to distinguish various maps in-
duced by T .)
( ′′ ) ( ) ( ′ )
0 −−→ HomOe M e ,N e −−→ Hom e M, e Ne −−→ Hom e M e ,N e
O O
  
T e ′′ e T e e T e ′ e
y M ,N y M,N y M ,N
( ′′ ) ( ) ( )
0 −−→ HomO M , N −−→ HomO M, N −−→ HomO M ′ , N

·Ee ( ′′ ) ( ) ( ′ )
−−→ Ext1Oe Me ,N e −−→ Ext1 M,e N e −−→ Ext1 M e ,N e
e
O e
O 

T 1 T 1 T 1
y Me ′′ ,Ne y M,e Ne y Me ′ ,Ne
e
·T (E) ( ) ( ) ( )
−−−→ Ext1O M ′′ , N −−→ Ext1O M, N −−→ Ext1O M ′ , N .

(2) The analogous statement holds replacing T by T in (1), M by M, et cetera.

Proof. By (6.40), the rows are exact, and it remains to show that the following
diagram is commutative:
( ′ ) ·Ee ( ′′ )
HomOe Me ,N e −−− −→ Ext1Oe Me ,N e
 
T e ′ e T 1
(6.49) y M ,N y Me ′′ ,Ne
( ) ·T (E)
e ( )
HomO M ′ , N −−−−→ Ext1O M ′′ , N
232 6. Super duality

( ′ ) ( ′′ )
Let fe ∈ HomOe M e . Then feEe ∈ Ext1 M
e ,N
e
O
e ,N
e is the bottom exact row of
the following commutative diagram:

f′ −−−ei−→
0 −−−−→ M e −−−−→
M e ′′ −−−−→ 0
M
 
(6.50) e 
yf y
e −−−−→ Fe −−−−→ M
0 −−−−→ N e ′′ −−−−→ 0

Here we identify feEe with the module F,


e which is the pushout of fe and ei. Applying
the functor T to (6.50) gives us a commutative diagram with exact rows:
TMe ′ ,Me [ei]
0 −−−−→ M ′ −−−−→ M −−−−→ M ′′ −−−−→ 0
 
T [ fe] 
y Me ′ ,Ne y
e −−−−→ M ′′ −−−−→ 0
0 −−−−→ N −−−−→ T (F)
e ≡ T 1 ′′ [ feE],
We conclude that T (F) e is the pushout of T e ′ e [ fe] and T e ′ e [ei]. There-
Me ,Ne M ,N M ,M
fore, we obtain that

TMe1 ′′ ,N [ feE] e ≡ T e ′ e [ fe]T [E],


e ≡ T (F)
M ,N
e

i.e., the diagram (6.49) is commutative. 


e N
Lemma 6.39. Let M, e We have
e ∈ O.
e N)
(1) T : HomOe (M, e → HomO (M, N) is an injection,
e N)
(2) T : HomOe (M, e → Hom (M, N) is an injection.
O

Proof. By Lemma 6.12, any el-isomorphism φ e : L(el, µθ ) → L(el, µθ ), with µ ∈ P+ ,


e] : L(l, µ) → L(l, µ) and T [φ
induces isomorphisms T [φ e] : L(l, µ♮ ) → L(l, µ♮ ). The
lemma follows. 

Lemma 6.40. Let Ve (λθ ) be a highest weight e


g-module of highest weight λθ with
e Then we have
e ∈ O.
λ ∈ P+ , and let N
(1) T : HomOe (Ve (λθ ), N)
e −→ HomO (V (λ), N) is an isomorphism.
(2) T : HomOe (Ve (λθ ), N)
e −→ Hom (V (λ♮ ), N) is an isomorphism.
O

Proof. Consider the commutative diagram with exact rows


e θ ) −−−−→ Ve (λθ ) −−−−→ 0
e −−−−→ ∆(λ
0 −−−−→ M
  
T Te θ Te θ
yM e y ∆(λ ) y V (λ )
0 −−−−→ M −−−−→ ∆(λ) −−−−→ V (λ) −−−−→ 0.
6.5. Super duality as an equivalence of categories 233

This gives rise to the following commutative diagram with exact rows:
( ) ( ) ( )
0 −−−→ HomOe Ve (λθ ), N e −−−→ Hom e ∆(λe θ ), Ne −−−→ Hom e M, e Ne
O O
  
Te θ e Te θ e T e e
y V (λ ),N y ∆(λ ),N y M,N
( ) ( ) ( )
0 −−−→ HomO V (λ), N −−−→ HomO ∆(λ), N −−−→ HomO M, N .
Corollary 6.26(3) and Theorem 6.32 imply that T∆(λ e is an isomorphism. By
e θ ),N
Lemma 6.39, TM, e is an injection. Now a standard diagram chasing implies that
eN
TVe (λθ ),Ne is an isomorphism. 

Lemma 6.41. Let Ve (λθ ) be a highest weight e


g-module of highest weight λθ with
e Then we have
e ∈ O.
λ ∈ P+ , and let N
(1) T : Ext1Oe (Ve (λθ ), N)
e → Ext1 (V (λ), N) is an injection.
O

(2) T : Ext1Oe (Ve (λθ ), N)


e → Ext1 (V (λ♮ ), N) is an injection.
O

Proof. Let
e
(6.51) e −→ Ee −→
0 −→ N
f
Ve (λθ ) −→ 0
be an exact sequence of e
g-modules. Suppose that (6.51) gives rise to a split ex-
T [ fe]
act sequence of g-modules 0 → N → E → V (λ) → 0. Thus there exists ψ ∈
HomO (V (λ), E) such that T [ fe] ◦ ψ = 1V (λ) . By Lemma 6.40, there exists ψ e∈
e θ e e
e ] = ψ. Thus T [ f ◦ ψ
HomOe (V (λ ), E) such that T [ψ e
e ] = T [ f ] ◦ T [ψ
e ] = 1V (λ) . By
e e
Lemma 6.40 again, we have f ◦ ψ = 1Ve (λθ ) , and hence (6.51) splits. 

e N
Proposition 6.42. Let M, e We have
e ∈ O.
e N)
(1) T : HomOe (M, e −→ HomO (M, N) is an isomorphism.
e N)
(2) T : Hom e (M, e −→ Hom (M, N) is an isomorphism.
O O

Proof. First we assume that M e∈O e f . We proceed by induction on the length of


e If M
filtration of M. e is a highest weight module, then it is true by Lemma 6.40.
e e
Let 0 = M0 ⊂ M1 ⊂ M e2 ⊂ · · · ⊂ M
ek = M e be an increasing filtration of e
g-modules
such that M ei /M
ei−1 is a highest weight module of highest weight νθ with νi ∈ P+ ,
i
for 1 ≤ i ≤ k. Let Zei := Me i /M
ei−1 and Zi := T (Zei ).
Consider the following commutative diagram with exact top row of e g-modules
and exact bottom row of g-modules, for i ≥ 1.
eιi
ei−1 −−−
0 −−−−→ M −→ ei −−−−→
M Zei −−−−→ 0
  
(6.52) T e T e Te
y Mi−1 y Mi y Zi
ι
0 −−−−→ Mi−1 −−−i−→ Mi −−−−→ Zi −−−−→ 0
234 6. Super duality

The sequence (6.52) induces the following commutative diagram with exact rows.
( ) ( )
0 −−−−→ HomOe Zei , N e −−−−→ Hom e M
O
ei , N
e
 
Te e T e e
y Zi ,N y M,N
( ) ( )
0 −−−−→ HomO Zi , N −−−−→ HomO Mi , N

·eιi ( ) ( )
−−−− → HomOe Mei−1 , N
e −−−−→ Ext1 Zei , Ne
e
O
 
T e e T 1
y Mi−1 ,N y Zei ,Ne
·ιi ( ) ( )
−−−− → HomO Mi−1 , N −−−−→ Ext1O Zi , N .
The map TMei−1 ,Ne is an isomorphism by induction. The map TZe1 ,Ne is an injection
i
by Lemma 6.41. Also TZei ,Ne is an isomorphism by Lemma 6.40. Now a standard
diagram chasing implies that TMei ,Ne is an isomorphism.
Now we consider the general case (i.e. M e By Proposition 6.34, we may
e ∈ O).
choose an increasing filtration of e g-modules 0 = M e0 ⊂ M e1 ⊂ M
e2 ⊂ · · · such that

e e e e
i≥0 Mi = M and Mi /Mi−1 is a highest weight module of highest weight νi with
θ

νi ∈ P+ , for i ≥ 1. Then the direct limit of {M ei } is lim M e ∼ e


−→ i = M and the inverse
e e ∼ e e ∼
limit is ← − HomOe (Mi , N) = HomOe (M, N). Similarly we have lim
lim −→ Mi = M and

←− HomO (Mi , N) = HomO (M, N). Furthermore, we have the following commu-
lim
tative diagram (where φ = lim
←−TMei ,Ne ):
( ) ∼ ( )
HomOe M, e N e −−− =
−→ lim Hom e Mei , Ne
 ←− O

T e e φ
y M,N y
( ) ∼
= ( )
HomO M, N −−−−→ lim ←− Hom O M i , N
Since φ is an isomorphism, so is TM, eN e . 

e N
Lemma 6.43. Let M, e We have
e ∈ O.
e N)
(1) T : Ext1Oe (M, e → Ext1 (M, N) is an injection,
O
e N)
(2) T : Ext1Oe (M, e → Ext1 (M, N) is an injection.
O

Proof. The proof is virtually identical to the proof of Lemma 6.41, where we use
Proposition 6.42 in place of Lemma 6.40. 
Proposition 6.44. Let Ve (λθ ) be a highest weight e
g-module of highest weight λθ
with λ ∈ P , and let N
+ e
e ∈ O. Then we have
(1) T : Ext1Oe (Ve (λθ ), N)
e → Ext1 (V (λ), N) is an isomorphism.
O

(2) T : Ext1Oe (Ve (λθ ), N)


e → Ext1 (V (λ♮ ), N) is an isomorphism.
O
6.5. Super duality as an equivalence of categories 235

Proof. Consider the following commutative diagram with exact rows:


e θ ) −−−eπ−→ Ve (λθ ) −−−−→ 0
e −−−−→ ∆(λ
0 −−−−→ M
  
(6.53) T Te θ Te θ
y Me y ∆(λ ) y V (λ )
π
0 −−−−→ M −−−−→ ∆(λ) −−−−→ V (λ) −−−−→ 0.
The sequence (6.53) induces the following commutative diagram with exact rows:
( ) ( ) ( )
e θ ), N
HomOe ∆(λ e −−−−→ Hom e M, e Ne −−−−→ Ext1 Ve (λθ ), N e
O e
O 
 
Te θ e T e e T 1
y ∆(λ ),N y M,N y Ve (λθ ),Ne
( ) ( ) ( )
HomO ∆(λ), N −−−−→ HomO M, N −−−−→ Ext1O V (λ), N

·e
π ( ) ( )
e θ ), N
−−−−→ Ext1Oe ∆(λ e −−−−→ Ext1 M,e N e
e
O
 
T 1 T 1
y ∆(λ
e θ e
),N y M,e Ne
·π ( ) ( )
−−−−→ Ext1O ∆(λ), N −−−−→ Ext1O M, N .
The map TM,1 is an injection by Lemma 6.43. The map T 1
eNe e is an isomorphism
e θ ),N
∆(λ
by Lemma 6.33. Also T∆(λ e and TM,
e θ ),N eN e are isomorphisms by Lemma 6.42. Now a
standard diagram chasing implies that TVe1(λθ ),Ne is an isomorphism. 

Proof of Theorem 6.37. Let C, C′ be abelian categories. Recall from Mitchell


[Mi, II.4] that by definition a full and faithful functor F : C 7→ C′ is an equiva-
lence of categories, if it satisfies the representative property that for every M ′ ∈ C′
there exists M ∈ C with F(M) ∼ = M′.
Proposition 6.42 implies that the functor T is full and faithful. Now for every
M ∈ O, there is a filtration of g-modules for M, 0 = M0 ⊂ M1 ⊂ M2 ⊂ · · · , with
Mi ∈ O satisfying the properties of Proposition 6.34. The filtration {Mi } of M
ei } with M
lifts to a filtration {M e f such that T (M
ei ∈ O ei ) ∼
= Mi by induction using
Propositions 6.36 and 6.44. Set M e := ∪i≥0 M e ∼
ei . It follows that T (M) = M.
We need to prove that M e ∈ O.e To do that we proceed as in the proof of the
claim in Proposition 6.13. Suppose that ζ1 , . . . , ζr ∈ P+ such that every weight
of M is bounded by ζi , for some i. For j = 1, . . . , r, define Pj and P(M) as in
Proposition 6.13. Now since T (M) e = M, it suffices to prove that if ν ∈ P+ and
ν ≤ ζi for some i, then there exists γ ∈ P(M) such that νθ < γθ . Let |+ ζi | = ki . We
write
ζi − ν = − κ + p(ε−1 − ε1 ) + + κ,
where p ∈ Z+ , − κ ∈ ∑α∈Π(k) Z+ α, and + κ ∈ ∑β∈Π(T)\β× Z+ β. Then + ν is a par-
tition of size ki + p. Hence, there exists γ ∈ P(M) such that the partition + ν is
obtained from + γ by adding p boxes. We record the sub-indices of the boxes in
236 6. Super duality

θ(+ γ) \ θ(+ ν) in the multiset J θ . Then we have γθ − νθ = − κ + ∑s∈J θ (ε−1 − εs ),


and hence γθ > νθ . This implies that all the weights of M e are bounded above by
θ e
the finite set {γ |γ ∈ P(M)}, and hence M ∈ O.e
We conclude that the functor T satisfies the representative property. Hence,
e
T : O → O is an equivalence of categories. This proves (1).
The proof of (2) is entirely parallel to (1), while (3) follows from (1) and (2)
(see e.g. [Mi, II.10]). 
Remark 6.45. Super duality helps to provide a category theoretical explanation
and new proofs for results obtained in earlier chapters by other means: the irre-
ducible polynomial character formula for gl(m|n) in terms of super Schur functions
via Schur-Sergeev duality in Chapter 3, and also the irreducible character formulas
for the oscillator modules of Lie superalgebras via Howe duality in Chapter 5.

6.6. Notes
This chapter is an exposition on the formulation and proof of super duality, which
is an equivalence of categories O and O, in Cheng-Lam-Wang [CLW], which in
turn was built on Cheng-Lam [CL2] for type A. The super duality conjecture for
type A was first formulated in Cheng-Wang-Zhang [CWZ] and more generally in
Cheng-Wang [CW4], motivated and based in part by Brundan [Br1]. There is a
different approach of Brundan-Stroppel [BrS] to the special case of super duality
conjecture in [CWZ].
Section 6.1. This preliminary section follows closely Cheng-Lam-Wang [CLW].
Section 6.2. The truncation functors and their basic properties regarding Verma
and irreducible modules (Proposition 6.7) have counterparts in the algebraic group
setting, which was developed by Donkin [Don]. In the super duality context, the
truncation functors first appeared in Cheng-Wang-Zhang [CWZ], and then gener-
alized in [CW4, CL2, CLW].
Section 6.3. The odd reflection approach leading to Theorem 6.15 and the
character formulas in Theorem 6.16 was developed in [CL2, CLW]. It is remark-
able that, when combined with Proposition 6.7, this solves completely the irre-
ducible character problem for modules in the category On , which include all finite-
dimensional simple modules of types gl and osp. This approach can be generalized
in a straightforward manner to provide irreducible characters of Kac-Moody super-
algebras.
Totally different approaches and solutions for the finite-dimensional irreducible
character problem have been developed in Serganova [Sva], Brundan [Br1], and
Brundan-Stroppel [BrS] for gl(m|n), and in Gruson-Serganova [GS] for osp.
Section 6.4. We follow Tanaka [Ta] for a formula of the (co)boundary operator
for Lie superalgebra (see also Iohara-Koga [IK]). The results on the identification
6.6. Notes 237

of Kazhdan-Lusztig-Vogan polynomials in the categories O, O and O e are due to


[CL2, CLW]. The study of u-cohomology groups were initiated in a fundamental
paper of Kostant [Kos]. We have followed Vogan’s homological interpretation of
the Kazhdan-Lusztig polynomials [Vo, Conjecture 3.4] (also see Serganova [Sva]
in the setting of finite-dimensional module category for gl(m|n)), as the usual ap-
proach via Hecke algebras is not applicable for Lie superalgebras. The Kazhdan-
Lusztig conjecture [KL] for the BGG category of a semisimple Lie algebra was
proved in Beilinson-Bernstein [BB] and Brylinski-Kashiwara [BK]. The polyno-
mials ℓµλ (q) for the category O coincide with the usual parabolic Kazhdan-Lusztig
polynomials. Theorem 6.22, which describes an explicit connection between u− -
homology and u-cohomology groups, is based on [Liu, Section 4].
Section 6.5. The formulation of super duality (Theorem 6.37) follows Cheng-
Lam-Wang [CLW] (also see [CWZ, CW4, CL2] for earlier formulation). The
proof here using Ext1 only is somewhat different from the original one, and it is
based on Theorem 6.32 which is adapted from Rocha-Caridi and Wallach [RW,
§7, Theorem 2]. The definition of the module subcategory O e 0̄ and its variants for
type A was first given in [CL2, Section 2.5], and in general in [CLW, Section 5.2].
The tilting modules, formulated earlier by Ringel and Donkin in different con-
texts, can be shown to exist in the categories On , On , On (cf. Brundan [Br1], Cheng-
Wang-Zhang [CWZ] and Cheng-Wang [CW2] for type A). The functors T : O e →O
and T : Oe → O can be shown to respect the tilting modules (see Cheng-Lam-Wang
[CLW2]). We do not treat the tilting modules in the book, and refer the reader to
the original papers for details.
Appendix A

Symmetric functions

A.1. The ring Λ and Schur functions


A.1.1. The ring Λ. A partition λ = (λ1 , λ2 , . . . , λℓ ) is a sequence of decreasing
nonnegative integers λ1 ≥ λ2 ≥ . . . ≥ λℓ ≥ 0. Each λi is called a part of λ, and the
number of nonzero parts of λ is called the length of λ, which is denoted by ℓ(λ).
The sum of all its nonzero parts is called the size of λ, and is denoted by |λ|. We
sometimes use the notation λ = (1m1 2m2 . . .) to denote a partition λ, where mi is the
number of parts of λ that are equal to i for i ≥ 1. We often identify a partition λ
with its associated Young diagram, and denote by λ′ the conjugate partition whose
Young diagram is transposed to the Young diagram of λ.
Let P denote the set of all partitions. For λ, µ ∈ P, we write λ ⊇ µ if λi ≥ µi for
all i ≥ 1. The dominance order on P, denoted by ≥, is the partial order defined
by

λ≥µ ⇔ |λ| = |µ| and λ1 + . . . + λi ≥ µ1 + . . . + µi , for all i ≥ 1.

Consider the Z-module of formal power series Z[[x]] in the set of infinite inde-
terminates x := {x1 , x2 , x3 , . . .}. For α = (α1 , α2 , . . . , αk ) ∈ Zk+ we write

xα := x1α1 x2α2 · · · xkαk ∈ Z[[x]].

The symmetric group Sn acts on Z[[x]] by permuting the indeterminates, i.e.,


if σ ∈ Sn and f ∈ C[[x]], then

σ f (x1 , x2 , . . .) := f (xσ(1) , xσ(2) , . . .),

where by definition σ(i) = i, for all i > n. Since the action of Sn+1 is compatible
with that of Sn , we obtain an action of the direct limit, denoted by S∞ , on Z[[x]].

239
240 A. Symmetric functions

The monomial symmetric function associated to a partition λ is


mλ (x) = mλ := ∑ σxλ ∈ Z[[x]],
σ∈S∞ /Sλ

where Sλ is the stabilizer subgroup of the monomial xλ in S∞ .


Definition A.1. The ring of symmetric functions Λ is the Z-span of the elements
{mλ | λ ∈ P} in Z[[x]].

We remark that Λ is closed under multiplication and hence is indeed a ring.


Also note that Λ is naturally graded by degree. We denote the Z-submodule of

degree n by Λn so that Λ = n≥0 Λn . Note that Λn is a free Z-module of rank
equal to the number of partitions of n. For a field F (which is often taken to be Q
or C), we denote the base change by ΛF = F ⊗Z Λ and ΛnF = F ⊗Z Λn .
There are several other bases for Λ that we use freely in this book.
Definition A.2. Let n ∈ N. We define
pn := m(n) = ∑ xin ,
i≥1

en := m(1n ) = ∑ xi1 xi2 · · · xin ,


i1 <i2 <···<in

hn := ∑ mλ = ∑ xi1 xi2 · · · xin ,


λ⊢n i1 ≤i2 ≤···≤in

that are called the nth power, elementary, and complete symmetric function,
respectively. Furthermore, we set e0 = h0 = 1, and er = hr = 0, for r < 0.

Let t be a formal indeterminate. The generating series of the power, elemen-


tary, and complete symmetric functions are as follows:
d 1
P(t) := ∑ pnt n−1 = dt ln ∏ 1 − xit ,
n≥1 i≥1

E(t) := ∑ ent n
= ∏(1 + xit),
n≥0 i≥1
1
H(t) := ∑ hnt n = ∏ 1 − xit .
n≥0 i≥1

It follows that
E(−t)H(t) = 1,
( )
pr t r
E(t) = exp ∑ (−1)r−1 ,
(A.1) r≥1 r
( )
pr t r
H(t) = exp ∑ .
r≥1 r
A.1. The ring Λ and Schur functions 241

Definition A.3. For a partition λ = (λ1 , λ2 , . . . , λℓ ) we define


ℓ ℓ ℓ
pλ = ∏ pλi , eλ := ∏ eλi , hλ := ∏ hλi .
i=1 i=1 i=1

(A.2) The set {pλ | λ ⊢ n} forms a linear basis for ΛnQ , for n ∈ N. Moreover,
ΛQ = Q[p1 , p2 , . . .].
The sets {eλ | λ ⊢ n} and {hλ | λ ⊢ n} are Z-bases for Λn , for n ∈ N. Moreover,
Λ = Z[e1 , e2 , . . .] = Z[h1 , h2 , . . .].

Proof. For a partition λ, a monomial appearing in pλ is of the form


(A.3) xiλ11 xiλ22 · · · xiλℓℓ ,
for some i1 , i2 , . . . , iℓ ∈ N. Write xα = xαj11 · · · xαjkk for the expression (A.3) such that
α1 ≥ α2 ≥ · · · ≥ αk and all the ji ’s are distinct. Then clearly α ≥ λ. Hence, we
have
pλ = ∑ cλµ mµ ,
µ≥λ

with cλµ ∈ Z+ and cλλ > 0. This implies that the matrix (cλµ ), with sub-indices or-
dered compatibly with the dominance order, is an invertible upper triangular matrix
with coefficients in Z. Since {mλ | λ ⊢ n} is a basis for ΛnQ , so is {pλ | λ ⊢ n}.
For a partition λ, every summand in eλ′ is of the form
(A.4) xi1 · · · xiλ′ x j1 · · · x jλ′ · · · xkℓ · · · xkλ′ ,
1 2 ℓ

where i1 < i2 < · · · < iλ′1 , j1 < j2 < · · · < jλ′2 , et cetera. Let us write the expression
in (A.4) as xα = xαj11 xαj22 · · · xαjss , such that α1 ≥ α2 ≥ · · · ≥ αs and all the ji ’s are
distinct. Note that λ ≥ α. Hence we conclude that
eλ′ = ∑ aλµ mµ ,
µ≤λ

with aλµ ∈ Z+ and aλλ = 1. This implies that the set {eλ | λ ⊢ n} is a Z-basis for
Λn .
It follows by (A.1) that, for all n ∈ N,
n
(A.5) ∑ (−1)r er hn−r = 0.
r=0
Thus we can express en as a polynomial of hr ’s and er−1 ’s, for r ≤ n. By induction
on n it follows that every en is a polynomial of {h1 , . . . , hn } with integer coefficients.
Therefore, the Z-module Z[h1 , . . . , hn ]∩Λn = ∑λ⊢n Zhλ contains the Z-span of {eλ |
λ ⊢ n}. Hence ∑λ⊢n Zhλ = Λn , and so {hλ | λ ⊢ n} is a Z-basis for Λn . 

We note the following immediate consequence of (A.2).


242 A. Symmetric functions

(A.6) The sets {p1 , p2 , . . .}, {e1 , e2 , . . .}, and {h1 , h2 , . . .} are algebraically inde-
pendent in Λ.
By (A.6), we can define a ring homomorphism ω : Λ → Λ by letting
ω(en ) = hn , ∀n ∈ N.
(A.7) We have ω(hn ) = en and ω(pn ) = (−1)n−1 pn . Thus, ω is an involution of the
ring Λ.

Proof. We prove the first statement by induction on n. From (A.5) we obtain


n n−1
hn = − ∑ (−1)r er hn−r , en = − ∑ (−1)n−r er hn−r .
r=1 r=0

This implies by induction hypothesis that


n
ω(hn ) = − ∑ (−1)r hr en−r = en .
r=1

Also, it follows by (A.1) that


d d
ω (P(t)) = ω (ln H(t)) = ln E(t) = −P(−t),
dt dt
from which we conclude that ω(pn ) = (−1)n−1 pn . 

For a partition λ = (1m1 2m2 . . .) of n, we let


zλ = ∏ imi mi !,
i≥1

which is the order of the centralizer of a permutation in Sn of cycle type λ. We


compute that
( r)
pr t
(A.8) H(t) = ∏ exp = ∑ z−1 |λ|
λ pλt .
r≥1 r λ∈P

Equivalently, for each n ≥ 1, we have


hn = ∑ z−1
λ pλ .
λ⊢n

Applying ω to (A.8), we obtain that


(A.9) E(t) = ∑ (−1)|λ|−ℓ(λ) z−1 |λ|
λ pλt .
λ∈P

Equivalently, for each n ≥ 1, we have


en = ∑ (−1)|λ|−ℓ(λ) z−1
λ pλ .
λ⊢n
A.1. The ring Λ and Schur functions 243

A.1.2. Schur functions. We first suppose that x = {x1 , x2 , . . . , xn }, and denote the
ring of symmetric polynomials in n variables by Λn . There is a natural ring homo-
morphism πm,n : Λm → Λn , for m > n, given by sending each x j for j > n to 0, and
Λ is the inverse limit of Λn .
For a partition λ of length no greater than n, we let
λn
aλ := ∑ λ1
sgn(σ)xσ(1) λ2
xσ(2) · · · xσ(n) ,
σ∈Sn

where sgn(σ) denotes the sign of the permutation σ. Then aλ is skew-symmetric


in the sense that
τaλ = sgn(τ)aλ , ∀τ ∈ Sn .

Let ρ = (n − 1, n − 2, . . . , 1, 0), so that we have


( )
λn λ +n− j
aλ+ρ = ∑ λ1 +n−1 λ2 +n−2
sgn(σ)xσ(1) xσ(2) · · · xσ(n) = det xi j
1≤,i, j≤n
.
σ∈Sn

In particular, aρ is the Vandermode determinant. Since aλ+ρ = 0, when xi = x j


for i ̸= j, it follows that aλ+ρ is divisible by xi − x j , for all i ̸= j. Therefore the
expression
aλ+ρ
sλ :=

is a symmetric polynomial in x1 , x2 , . . . , xn , called the Schur polynomial associated
to λ. Noting that sλ (x1 , . . . , xn , 0) = sλ (x1 , . . . , xn ), the inverse limit, denoted by
sλ ∈ Λ, exists and it is called the Schur function associated to λ.
Let An denote the Z-module of skew-symmetric polynomials in x1 , x2 , . . . , xn .
Then aλ+ρ , as λ runs over partitions with ℓ(λ) ≤ n, forms a Z-basis An . Since
every element in An is divisible by aρ , the map Λn → An given by multiplication
by aρ is a Z-module isomorphism. Thus, the sλ ’s, as λ runs over all partitions
with ℓ(λ) ≤ n, form a Z-basis for Λn . Therefore, we conclude that the set of Schur
functions associated with partitions of size k is a basis for Λk , for each k ≥ 0.
(A.10) For a partition λ we have
( )
sλ = det hλi −i+ j 1≤i, j≤n , n ≥ ℓ(λ),
( )
sλ = det eλ′i −i+ j , n ≥ ℓ(λ′ ).
1≤i, j≤n

In particular, we have ω(sλ ) = sλ′ .


(k)
Proof. Let er denote the rth elementary symmetric polynomial in the variables
x1 , x2 , . . . , xbk , . . . , xn , i.e., with xk deleted. Set
( )
(k)
M := (−1)n−i en−i .
1≤i,k≤n
244 A. Symmetric functions

Let α = (α1 , . . . , αn ) ∈ Zn+ . Let


( )
Aα = xαj i , Hα := (hαi −n+ j )1≤i, j≤n ,
1≤i, j≤n

and let
n−1
∑ er t = ∏(1 + xit).
(k) r
E (k) (t) =
r=0 i̸=k

Then
1
H(t)E (k) (−t) = .
1 − xk t
This implies that
n
∑ hα −n+ j (−1)n− j en− j = xkα .
(k) i
i
j=1

This equation can be recast in a matrix form as Aα = Hα M. Taking the determinant


of both sides with α = λ + ρ, we obtain that
aλ+ρ = det Hλ det M.
/ we obtain that aρ = det M, and so aλ+ρ = det Hλ+ρ aρ , which is an
Setting λ = 0,
equivalent form of the first identity in (A.10) we wish to prove.
( )
Now the matrices H = (hi− j )0≤i, j≤N and E = (−1)i− j ei− j 0≤i, j≤N are inverses
of each other by (A.5) with determinant equal to 1. This gives a relationship be-
tween the minors of H and the cofactors of E t . Exploiting this relationship one can
prove the following identity:
( ) ( )
det hλi −i+ j 1≤i, j≤n = det eλ′ i −i+ j 1≤i, j≤n ,
for n ≥ max{ℓ(λ), ℓ(λ′ )}. The second identity in (A.10) now follows by the first
identity in (A.10) and the identity above.
The formula ω(sλ ) = sλ′ follows by applying ω to the first identity and using
the second identity with λ and λ′ switched. 

Take two sets of independent variables x = {x1 , x2 , . . .} and y = {y1 , y2 , . . .}.


Applying (A.8) to the set of variables {xi y j } and setting t = 1, we obtain that
1
∏ (1 − xi y j ) = ∑ z−1
λ pλ (x)pλ (y).
i, j λ
(A.11) The following identities hold:
1
∏ 1 − xi y j = ∑ hλ (x)mλ (y),
i, j λ∈P

∏(1 + xi y j ) = ∑ eλ (x)mλ (y),


i, j λ∈P
A.1. The ring Λ and Schur functions 245

1
∏ 1 − xi y j = ∑ sλ (x)sλ (y),
i, j λ∈P

∏(1 + xi y j ) = ∑ sλ (x)sλ (y). ′


i, j λ∈P
(The last two are known as Cauchy identities.)

Proof. We compute
( )
1
∏ 1 − xi y j = ∏ H(y j ) = ∏ ∑ hr (x)yrj = ∑ hα (x)yα1 1 yα2 2 · · · ,
i, j j j r≥0 α

where the last sum is over all compositions α = (α1 , α2 , . . .). Hence it equals
∑λ hλ (x)mλ (y), proving the first identity in (A.11). The second identity follows
by applying ω in x variables to the first one and using (A.10).
Assume that the number of variables of x and y both equal to n. The gen-
eral
( case is−1proved
) by letting n go to infinity, as usual. Consider the n × n matrix
(1 − xi y j ) . If we multiply the ith row of this matrix by ∏nj=1 (1 − xi y j ), then
we obtain a matrix whose (i, k)th entry is equal to
n
∏(1 − xi yr ) = ∑ (−1)n− j en− j (y)xi
(k) n− j
.
r̸=k j=1

That is, it is the (i, k)th entry of the matrix Aρ (x)M(y), where Aρ and M are as in
the proof of (A.10). This implies that
( ) 1
(A.12) det (1 − xi y j )−1 = aρ (x)aρ (y) ∏ .
i, j 1 − xi y j

Next, we compute that


( ) ∞ n
det (1 − xi y j )−1 = det( ∑ (xi y j )k ) = ∑ ∑ sgn(σ) ∏ xiαi yασ(i)
i
,
k=0 α σ∈Sn i=1

where the first summation is over compositions α. Thus we can write


( )
det (1 − xi y j )−1 = ∑ xα aα (y) = ∑ ∑ (−1)ℓ(τ) xτ(λ+ρ) aλ+ρ (y)
α λ τ∈Sn
(A.13) = ∑ aλ+ρ (x)aλ+ρ (y)
λ
where the summations over λ are taken over partitions of length no greater than n.
The first Cauchy identity follows now by combining (A.12) and (A.13). The
second Cauchy identity follows by applying ω in y variables to the first Cauchy
identity and using (A.10). 

We define a Z-valued bilinear form (·, ·) on Λ by declaring


(A.14) (hλ , mµ ) = δλµ .
246 A. Symmetric functions

This extends to a Q-valued bilinear form on ΛQ .


(A.15) Let {uλ | λ ∈ P} and {vλ | λ ∈ P} be two bases for ΛQ . Then the following
conditions are equivalent.
(1) (uλ , vµ ) = δλµ .
(2) ∑λ uλ (x)vλ (y) = ∏i, j (1−x1 i y j ) .

Proof. We write
uλ = ∑ aλν hν , vµ = ∑ bµη mη , aλν , bµ,η ∈ Q.
ν η

Then (1) is equivalent to ∑ν aλν bµν = δλµ , which is equivalent to ∑λ aλν bλη = δνη .
Now (2) is equivalent to ∑λ ∑ν,η aλν bλ,η hν (x)mη (y) = ∑λ hλ (x)mλ (y) by using
(A.11). This is equivalent to ∑λ aλν bλη = δνη . 

It follows from (A.15) and (A.11) that


(A.16) (sλ , sµ ) = δλµ .

A.1.3. Skew Schur functions. For partitions µ and ν, we write the product sµ sν
as
sµ sν = ∑ cλµν sλ , cλµν ∈ Z.
λ

The cλµν associated with a triple of partitions are the Littlewood-Richardson coeffi-
cients, and they allow us to define the skew Schur function associated to partitions
λ, µ by the formula
sλ/µ := ∑ cλµν sν .
ν
This definition is equivalent to the following identity:
(A.17) (sµ f , sλ ) = ( f , sλ/µ ), ∀ f ∈ Λ.
This can be seen by checking for f = sν . It follows from the definitions that sλ/0/ =
sλ , and by (A.10) that
ω(sλ ) = sλ′ .
Since ω is a ring isomorphism, we obtain by the definition of the Littlewood-

Richardson coefficients above that cλµ′ ν′ = cλµν . It now follows from the definition
of the skew Schur functions that
ω(sλ/µ ) = sλ′ /µ′ .
Let µ ⊆ λ be partitions which we may regard as Young diagrams. The skew
diagram λ/µ is obtained from λ by removing the sub-diagram µ. A tableau of
shape λ/µ is a filling of the boxes of λ/µ with elements from the set N. A tableau
T of λ/µ is called semistandard, if the entries in T are strictly increasing from
A.1. The ring Λ and Schur functions 247

top to bottom along each column and (weakly) increasing from left to right along
each row. Given such a semistandard tableau T with entries t ∈ T , we define xT :=
∏t∈T xt .
The skew Schur function sλ/µ admits the following combinatorial description,
a proof of which can be found in [Mac, (5.12)].
(A.18) For partitions λ ⊇ µ, we have
sλ/µ = ∑ xT ,
T
where the summation is over all semistandard tableaux of shape λ/µ.
The following is an easy consequence of (A.18).
(A.19) Let x = {x1 , x2 , . . .} and y = {y1 , y2 , . . .} be two independent sets of vari-
ables. We have
sλ (x, y) = ∑ sµ (x)sλ/µ (y).
µ⊆λ

Let µ ⊆ λ be partitions. Recall the weight of a semistandard tableau T of shape


λ/µ is (1m1 2m2 · · · ), where mi is the number of times i appears in T . Denote by
Kλ−µ,ν the number of semistandard tableaux of shape λ/µ and weight ν. Then
(A.18) implies that
(A.20) sλ/µ = ∑ Kλ−µ,ν mν .
ν∈P
It follows from (A.17) and (A.20) that
sµ hν = ∑ Kλ−µ,ν sλ .
λ

Taking ν = (r) to be the one-part partition we obtain the Pieri’s formula:


(A.21) sµ hr = ∑ sλ ,
λ
where the summation is over partitions λ such that λ/µ is a skew diagram of size r
whose columns all have length at most one.
Recall that a partition is called even, if all its parts are even. We have the
following two symmetric function identities ([Mac, I.5, Example 5]):
∑ sµ = ∏ (1 − xi x j )−1 ,
µ even 1≤i≤ j

∑ sν′ = ∏ (1 − xi x j )−1 .
ν even 1≤i< j

Since the involution ω interchanges sλ with sλ′ we conclude from these identities
that
( )
(A.22) ω ∏ (1 − xi x j )−1 = ∏ (1 − xi x j )−1 .
1≤i≤ j 1≤i< j
248 A. Symmetric functions

A.1.4. The Frobenius characteristic map. Let Rn denote the Grothendieck group
of the category of Sn -modules, which has [Sλ ] for λ ∈ Pn as a Z-basis. Let


R= Rn .
n=0

Then R is a graded algebra with multiplication given by


S
×Sn ( f ⊗ g),
f g = IndSmm+n for f ∈ Rm , g ∈ Rn .

The Frobenius characteristic map chF : R → Λ is the Z-linear map such that, for
n ≥ 0,
(A.23) chF (χ) = ∑ z−1
µ χµ pµ , χ ∈ Rn ,
µ∈Pn

where χµ denotes the character value of χ at a permutation of cycle type µ. Re-


call 1n and sgnn denote the trivial and sign representations or characters of Sn ,
respectively. The following basic properties of chF are well known.
(A.24) The characteristic map chF is an isomorphism of graded rings. Moreover,
for n ≥ 0 and λ ∈ Pn , we have that
( )
chF IndCS n
CSλ 1n = hλ ,
( CSn )
chF IndCS λ
sgnn = eλ ,
chF ([Sλ ]) = sλ .

Proof. One can use the induced character formula to prove that chF is a ring homo-
morphism. By (A.8) and (A.9) we have respectively chF (1n ) = hn and chF (sgnn ) =
en . This immediately implies the first two identities since chF is a homomorphism.
It is known (e.g., as a special case of Lemma 3.11) that

(A.25) CSn
IndCSµ
1n ∼
= Kλµ Sλ ,
λ≥µ

where the Kostka number Kλµ is equal to the number of semistandard λ-tableaux
of content µ. Note that Kµµ = 1.
Also from (A.18) we conclude that
sλ = ∑ Kλµ mµ .
µ≤λ

Using the relations of dual bases (A.14) and (A.16), this identity admits the fol-
lowing dual version:
(A.26) hµ = ∑ Kλµ sλ .
λ≥µ
A.2. Supersymmetric Polynomials 249

It follows by comparing (A.25) and (A.26) and using the first identity in (A.24) that
chF ([Sλ ]) = sλ . Since [Sλ ] and sλ for λ ∈ P form Z-bases for R and Λ respectively,
we conclude that chF is an isomorphism. 

(A.27) For µ ∈ Pn , we have an Sn -module isomorphism: Sµ ⊗ sgnn ∼



= Sµ .

Proof. By (A.23) and (A.10), we have that

chF ([Sµ ⊗ sgnn ]) = ∑ (−1)n−ℓ(µ) z−1


µ χµ pµ
µ∈Pn
( )
=ω ∑ z−1
µ χµ p µ
µ∈Pn

= ω(sµ ) = sµ′ = chF ([Sµ ]).

Now the isomorphism follows from (A.24) that chF is an isomorphism. 

A.2. Supersymmetric Polynomials


A.2.1. The ring of supersymmetric polynomials. Let x = {x1 , x2 , . . . , xm } and
y = {y1 , y2 , , . . . , yn } be independent indeterminates. A polynomial f in Z[x, y] is
called supersymmetric, if the following conditions are satisfied:
(1) f is symmetric in x1 , . . . , xm .
(2) f is symmetric in y1 , . . . , yn .
(3) The polynomial obtained from f by setting xm = yn = t is independent of
t.
Denote by Λ(m|n) the ring of supersymmetric polynomials in Z[x, y]. For a field F,
we denote the base change by Λ(m|n),F = F⊗Z Λ(m|n) . We call an element f ∈ Z[x, y]
satisfying (1) and (2) above doubly symmetric.
For k ≥ 1 set
m n
σrm,n := ∑ xir − ∑ yrj .
i=1 j=1

Then σrm,n is supersymmetric. Let


m n
m,n := ∏ ∏ (xi − y j ).
i=1 j=1

Note that m,n is supersymmetric. Also, if g is doubly symmetric, then gm,n is


supersymmetric.
(A.28) Let q be doubly symmetric. Then qm,n is a polynomial in {σrm,n | r ≥ 1}.
250 A. Symmetric functions

Proof. We proceed by induction on n (for arbitrary m), with the case n = 0 being
clear.
Let n ≥ 1. We may assume, without loss of generality, that q is homogeneous
of degree k ≥ 0, and we proceed by induction on k. Consider
qm,n |yn =0 = q|yn =0 x1 · · · xm m,n−1 .

By induction hypothesis on n, there exists a polynomial g in {σ1m,n−1 , . . . , σrm,n−1 }


such that
qm,n |yn =0 = g(σ1m,n−1 , . . . , σrm,n−1 ).
Note that
g(σ1m,n , . . . , σrm,n )|xm =yn =t = qm,n |xm =yn =0 = 0.

Thus xm − yn divides g(σ1m,n , . . . , σrm,n ). Since g(σ1m,n , . . . , σrm,n ) is doubly symmet-


ric, it follows that
(A.29) g(σ1m,n , . . . , σrm,n ) = q∗ m,n ,
for some doubly symmetric polynomial q∗ . Now
q∗ m,n |yn =0 = g(σ1m,n−1 , . . . , σrm,n−1 ) = qm,n |yn =0 .
Hence yn divides q∗ − q. Since q∗ − q is symmetric in y1 , . . . , yn , we conclude that
there exists a doubly symmetric polynomial f such that
(A.30) qm,n = q∗ m,n + y1 · · · yn f m,n ,
Note that f has degree k − n if k ≥ n and f = 0 if k < n.
In case when k < n we have f = 0. Hence we are done by (A.29).
Suppose that k ≥ n. Write f as a polynomial in terms of σ1m,0 , · · · , σm m,0 and
σ0,n , · · · , σ0,n . Replace these σ’s in the same polynomial by σm+1,0 , · · · , σm
1 n 1
m+1,0
and σ10,n , · · · , σn0,n , and denote by f ∗ the resulting homogeneous doubly symmetric
polynomial in Z[x1 , . . . , xm+1 , y1 , . . . , yn ]. Since f ∗ has degree k − n < k, we have
by induction hypothesis on k that there exists some polynomial h such that
f ∗ m+1,n = h(σ1m+1,n , . . . , σsm+1,n ).
This implies that
y1 · · · yn f m,n = (−1)n ( f ∗ m+1,n ) |xm+1 =0 = (−1)n h(σ1m,n , . . . , σsm,n ),
and hence the (A.28) follows from (A.30). 

(A.31) The algebra Λ(m|n),Q of supersymmetric polynomials is generated by σrm,n


for all r ≥ 1.
A.2. Supersymmetric Polynomials 251

Proof. We proceed by induction on n to show that every supersymmetric polyno-


mial p ∈ Q[x, y] is a polynomial in {σrm,n | r ≥ 1}. The case when n = 0 is clear.
Let n ≥ 1 and p ∈ Q[x, y] be a supersymmetric polynomial. Then, the poly-
nomial p|xm =yn =0 in x1 , . . . , xm−1 and y1 , . . . , yn−1 is supersymmetric. By induction
hypothesis on n, there exists a polynomial f such that
p|xm =yn =0 = f (σ1m−1,n−1 , . . . , σrm−1,n−1 ).
Set f ∗ := f (σ1m,n , . . . , σrm,n ). Then (p − f ∗ )|xm =yn =t = 0, and hence xm − yn divides
p − f ∗ . By double symmetry m,n divides p − f ∗ , and hence there exists a doubly
symmetric polynomial q such that
p = f ∗ + qm,n .
By (A.28), qm,n is generated by {σrm,n | r ≥ 1}, and hence so is p. 

We can define the ring of supersymmetric functions Λm|∞ and prove the fol-
lowing n = ∞ version of (A.31).
(A.32) The algebra Λ(m|∞),Q of supersymmetric functions is generated by σrm,∞ for
all r ≥ 1.

Proof. Let f ∈ Λ(m|∞),Q . Assume that f is homogeneous of degree n. Then f(n) :=


f |xn+1 =xn+2 =···=0 is supersymmetric and hence by (A.31)
f(n) = g(σ1m,n , . . . , σrm,n ).
Consider the element
f − g(σ1m,∞ , . . . , σrm,∞ ) ∈ Λ(m|∞),Q .
The expression f − g(σ1m,∞ , . . . , σrm,∞ ) is homogeneous of degree n and contains no
monomials made up of x1 , . . . , xm , y1 , . . . , yn only. It follows by the symmetry of the
y variables that f − g(σ1m,∞ , . . . , σrm,∞ ) = 0. 

A.2.2. Super Schur functions. Define Λ e (m|n) to be the subring of the ring of dou-
bly symmetric polynomials over Z in the two sets of variables {x1 , . . . , xm } and
{y1 , . . . , yn } consisting of polynomials f such that f |xm =t=−yn is independent of t.
Evidently f (x, y) ∈ Λ e m|n if and only if f (x, −y) is supersymmetric, hence we have
an isomorphism of rings

= e
(A.33) ρy : Λ(m|n) −→ Λ (m|n) ,

e (m|∞) accordingly, which is


given by ρy ( f (x, y)) = f (x, −y). We define the ring Λ
isomorphic to Λ(m|∞) .
Regard Λ as the ring of symmetric functions in x1 , x2 , . . ., and identify the vari-
ables xm+i = yi , for i ≥ 1. By applying the involution, denoted by ωy , to the set
252 A. Symmetric functions

e m|∞ .
of infinite variables {y1 , y2 , . . .} we obtain a ring homomorphism ωy : Λ → Λ
Indeed, by (A.7), we have
(A.34) ωy (pr ) = x1r + . . . + xm
r e m|∞ .
+ (−1)r−1 (yr1 + yr2 + . . .) ∈ Λ
Since the pr ’s, for r ≥ 1, are algebraically independent generators of Λ by (A.2) and
(A.6), we conclude that ωy (ΛQ ) ⊆ Λ e m|∞,Q , and hence ωy (Λ) ⊆ Λe m|∞,Q ∩ Z[x, y] =
e m|∞ . Now (A.32) and the fact that ρ−1
Λ −1
y (ωy (pr )) = σm,∞ , r ≥ 1 , imply that ρy ◦
r

ωy : ΛQ → Λm|∞,Q is a ring isomorphism. This implies that ωy : ΛQ → Λ e m|∞,Q is an


isomorphism. The inverse isomorphism ω−1 e
y : Λm|∞,Q → ΛQ , which is again given
by the involution ω on the y variables, clearly satisfies that ω−1 e
y (Λm|∞ ) ⊆ Λ. Hence,
ωy : Λ → Λe m|∞ is a ring isomorphism.
Define the super Schur function hsλ associated to a partition λ by
(A.35) hsλ (x; y) = ∑ sµ (x)sλ /µ (y).
′ ′

µ⊆λ

The definition of hsλ (x; y) makes sense for x and y being finite and infinite. By
(A.19), we have for y infinite
( )
(A.36) ωy sλ (x, y) = hsλ (x; y).
(A.37) The map ωy : Λ → Λ e m|∞ is an isomorphism of rings. Also, the set {hsλ | λ ∈
e m|∞ .
P} is a Z-basis for Λ
The map Λe (m|∞) → Λ
e (m|n) obtained by setting the variables yi = 0, for i > n, is
a ring epimorphism. Furthermore, it sends the super Schur function corresponding
to an (m|n)-hook partition to the respective super Schur polynomial corresponding
to the same (m|n)-hook partition. The other super Schur functions are sent to zero.
Evidently, the super Schur polynomials corresponding to (m|n)-hook partitions are
linearly independent. Thus we have the following.
(A.38) The set {hsλ }, as λ runs over all (m|n)-hook partitions, forms a Z-basis for
e m|n .
Λ
The the Cauchy identities in (A.11) admit the following super generalization.
(A.39) Let x = {x1 , x2 , . . .}, y = {y1 , y2 , . . .}, z = {z1 , z2 , . . .} be three sets of (possi-
bly infinite) indeterminates. The following identities hold:
∏ j,k (1 + y j zk )
= ∑ hsλ (x; y)sλ (z).
∏i,k (1 − xi zk ) λ∈P

Proof. If suffices to prove the claim in the case when all three sets of indetermi-
nates are infinite. A variant of the Cauchy identity in (A.11) can be written as
1 1
∏ (1 − xi zk ) ∏ (1 − y j zk ) = ∑ sλ (x, y)sλ (z).
i,k j,k λ∈P
A.3. The ring Γ and Schur Q-functions 253

Now the claim follows by applying the involution ωy on the y variables to both
sides of the above equation and using (A.36). 

A.3. The ring Γ and Schur Q-functions


A.3.1. The ring Γ. Let x = {x1 , x2 , . . .}. Define a family of symmetric functions
qr = qr (x), r ≥ 0, via a generating function
1 + txi
(A.40) Q(t) := ∑ qr (x)t r = ∏ 1 − txi .
r≥0 i

Note that q0 (x) = 1, and that Q(t) satisfies the relation


(A.41) Q(t)Q(−t) = 1,
which is equivalent to the identities:

∑ (−1)r qr qs = 0, n ≥ 1.
r+s=n

These identities are vacuous for n odd. When n = 2m is even, we have


m−1
1
(A.42) q2m = ∑ (−1)r−1 qr q2m−r − 2 (−1)m q2m .
r=1
Let Γ be the Z-subring of Λ generated by the qr ’s:
Γ = Z[q1 , q2 , q3 , . . .].

The ring Γ is graded by the degree of functions: Γ = n≥0 Γn , where Γn = Γ ∩ Λn .
We set
ΓQ = Q ⊗Z Γ, ΓC = C ⊗Z Γ.
For any partition µ = (µ1 , µ2 , . . .), we denote
qµ = qµ1 qµ2 . . . .

We shall denote by OP and SP the sets of odd and strict partitions, respectively.
(A.43) The ring Γ and ΓQ enjoy the following remarkable properties:
(1) ΓQ is a polynomial algebra with polynomial generators p2r−1 for r ≥ 1.
(2) ΓQ is a polynomial algebra with polynomial generators q2r−1 for r ≥ 1.
(1′ ) {pµ | µ ∈ OP} forms a linear basis for ΓQ .
(2′ ) {qµ | µ ∈ OP} forms a linear basis for ΓQ .

Proof. Recall the generating function for the power-sums: P(t) = ∑r≥1 pr t r−1 . We
have
Q′ (t)
= (ln Q(t))′ = P(t) + P(−t) = 2 ∑ p2r+1t 2r .
Q(t) r≥0
254 A. Symmetric functions

It follows that Q′ (t) = 2Q(t) ∑r≥0 p2r+1t 2r , from which we deduce


rqr = 2(p1 qr−1 + p3 qr−3 + . . .),
with the last term on the right-hand side being pr−1 q1 for r even and pr for r odd.
By using induction on r, we conclude that (i) each qr is expressible as a poly-
nomial in terms of ps ’s with odd s; (ii) each pr with odd r is expressible as a
polynomial in terms of qs ’s, which can be further restricted to the odd s (note that
each qr can be written as a polynomial of qs with odd s, by applying (A.42) and an
induction on r).
So,
ΓQ = Q[p1 , p3 , . . .] = Q[q1 , q3 , . . .].
Since the pr ’s (for r odd) are algebraically independent by (A.6), we have proved
(1). (2) follows from (1) by the above equation and a dimension counting.
Clearly, (1′ ) is equivalent to (1), while (2′ ) is equivalent to (2). 
(A.44) {qµ | µ ∈ SP} forms a Z-basis for Γ. Moreover, for any partition λ, we have
qλ = ∑ aµλ qµ ,
µ∈SP,µ≥λ

for some aµν ∈ Z.

Proof. We claim that qλ lies in the Z-span of {qµ | µ ∈ SPn , µ ≥ λ}, for each par-
tition λ of a given n. This can be seen by induction downward on the dominance
order on λ as follows. The initial step for λ = (n) is clear, as (n) is strict. For non-
strict λ, we have λi = λ j = m for some i < j and some m > 0. Applying (A.42) to
rewrite qλ easily provides the inductive step.
According to a formula of Euler, |SPn | = |OPn |. Now (A.44) follows from
Part (2) of (A.43) and the above claim. We record here the short proof of Euler’s
formula:
∏ (1 + qr )(1 − qr )
∑ |SPn |qn = ∏(1 + qr ) = r≥1∏r≥1 (1 − qr )
n≥0 r≥1

∏r≥1 (1 − q2r ) 1
= = = ∑ |OPn |qn .
∏r≥1 (1 − qr ) ∏r≥1,odd (1 − qr ) n≥0

(A.45) We have qn = ∑α∈OPn 2ℓ(α) z−1
α pα .

Proof. As seen in the proof of (A.43) above, we have


( )
1
Q(t) = exp 2 ∑ p2r−1t 2r−1 .
r≥1 2r − 1
Now (A.45) follows by comparing the coefficients of t n on both sides. 
A.3. The ring Γ and Schur Q-functions 255

A.3.2. Schur Q-functions. We shall define the Schur Q-functions Qλ , for λ ∈ SP.
Let
Q(n) = qn , n ≥ 1.
Consider the generating function
t1 − t2
Q(t1 ,t2 ) := (Q(t1 )Q(t2 ) − 1) .
t1 + t2
By (A.41), Q(t1 ,t2 ) is a power series in t1 and t2 , and we write
Q(t1 ,t2 ) = ∑ Q(r,s)t1rt2s .
r,s≥0

The following can be checked easily from the definition by noting Q(t1 ,t2 ) =
−Q(t2 ,t1 ).
(A.46) We have Q(r,s) = −Q(s,r) , Q(r,0) = qr . In addition,
s
Q(r,s) = qr qs + 2 ∑ (−1)i qr+i qs−i , r > s.
i=1

We recall the following classical facts (see Wikipedia). Any 2n × 2n skew-


symmetric matrix A = (ai j ) satisfies det(A) = Pf(A)2 , where Pf(A) is the Pfaffian
of A given by
Pf(A) = ∑ sgn(σ)aσ(1)σ(2) . . . aσ(2n−1)σ(2n) ,
σ
summed over σ ∈ S2n such that σ(2i − 1) < σ(2i) and σ(2i − 1) < σ(2i + 1) for all
admissible i. Equivalently, P f (A) can be defined as the above sum over the whole
group S2n divided by 2n n!.
Definition A.4. The Schur Q-function Qλ , for λ ∈ SP with ℓ(λ) ≤ 2n, is defined
to be the Pfaffian of the 2n × 2n skew-symmetric matrix (Q(λi ,λ j ) ).

Example A.5. For indeterminates t1 , . . . ,t2n , let A = (ai j ) be the 2n × 2n skew-


t −t
symmetric matrix with ai j = tii +t jj . Then,
ti − t j
Pf(A) = ∏ .
1≤i< j≤2n ti + t j

λ2n
Remark A.6. Equivalently, Qλ is the coefficient of t1λ1 . . .t2n in
Q(t1 , . . . ,t2n ) := Pf (Q(ti ,t j )).
It can be shown that Q(t1 , . . . ,t2n ) is alternating in the sense that
Q(tσ(1) , . . . ,tσ(2n) ) = sgn(σ)Q(t1 , . . . ,t2n ), ∀σ ∈ S2n .
It follows that Qλ = 0 unless all parts of λ are distinct.
256 A. Symmetric functions

(A.47) For λ ∈ SP with ℓ(λ) ≤ n, Qλ is equal to the coefficient of t1λ1 t2λ2 . . . in


ti − t j
Q(t1 ) . . . Q(tn ) ∏ ,
1≤i≤ j≤n ti + t j
ti −t j
where it is understood that ti +t j = 1 + 2 ∑r≥1 (−1)r (ti−1t j )r .
The statement of (A.47) can be found in [Mac, (8.8), pp.253]. The following
recursive relations for Qλ are obtained directly by the Pfaffian definition of Qλ and
the Laplacian expansion of Pfaffians (see Wikipedia).
(A.48) For a strict partition λ = (λ1 , . . . , λm ),
m
Qλ = ∑ (−1) j Q(λ ,λ ) Q(λ ,...,λ̂ ,...,λ
1 j 2 j m)
, for m even,
j=2
m
Qλ = ∑ (−1) j−1 Qλ Q(λ ,...,λ̂ ,...,λ
j 1 j m)
, for m odd.
j=1

(A.49) For λ ∈ SPn , we have


Qλ = q λ + ∑ aλµ qµ .
µ∈SPn ,µ>λ

It follows from the recursive relation (A.48) that Qλ is equal to qλ plus a linear
combination of qν for ν > λ (not necessarily strict). Now (A.49) follows since each
qν is expressible as a linear combination of qµ with strict partitions µ ≥ ν by (A.42)
(this fact has been used in the induction step in the proof of (A.44)).
As an immediately corollary of (A.44) and (A.49), we have
(A.50) The Qλ for all strict partitions λ form a Z-basis for Γ. Moreover, for any
partition µ, we have
qµ = ∑ Kbλµ Qλ ,
λ∈SPn ,λ≥µ

where K bλµ ∈ Z and Kbλλ = 1. (Clearly, one can further assume that µ is a composi-
tion instead of a partition in the statement).
bλµ ≥ 0
Indeed from representation-theoretic consideration it can be seen that K
(cf. Chapter 3, Lemma 3.47) .

A.3.3. The inner product on Γ. Let x = {x1 , x2 , . . .} and y = {y1 , y2 , . . .} be two


independent sets of variables.
(A.51) We have
1 + xi y j
∏ 1 − xi y j = ∑ 2ℓ(α) z−1
α pα (x)pα (y)
i, j α∈OP

= ∑ mµ (x)qµ (y).
µ∈P
A.3. The ring Γ and Schur Q-functions 257

Proof. Let xy = {xi y j | i, j = 1, 2, . . .}. It follows by (A.40) that


1 + xi y j
∏ 1 − xi y j = ∑ qn (xy) = ∑ 2ℓ(α) z−1
α pα (xy).
i, j n≥0 α∈OP

Now the first identity follows since pα (xy) = pα (x)pα (y). We further compute that

1 + xi y j
∏ 1 − xi y j = ∏ ∑ qr (y)xir i
i

i, j i ri =0

= ∑ mµ (x)qµ (y).
µ∈P

The second identity is proved. 

We define an inner product ⟨·, ·⟩ on ΓQ by letting


(A.52) ⟨pα , pβ ⟩ = 2−ℓ(α) zα δαβ .
(A.53) Let {uλ }, {vλ } be dual bases for ΓQ with respect to ⟨·, ·⟩. The following are
equivalent:
(i) ⟨uλ , vµ ⟩ = δλµ∀λ, µ;
1 + xi y j
(ii) ∑ uλ (x)vλ (y) = ∏ 1 − xi y j .
λ i, j

Such an equivalence can be established by the same standard argument as


(A.15), now based on (A.51).
(A.54) We have
⟨Qλ , Qµ ⟩ = 2ℓ(λ) δλµ , λ, µ ∈ SP.
Equivalently, we have
1 + xi y j
∏ 1 − xi y j = ∑ 2−ℓ(λ) Qλ (x)Qλ (y).
i, j λ∈SP

A rather nontrivial direct proof of (A.54) can be found in Jozefiak [Jo2]. Alter-
natively, one first works on the generality of Hall-Littlewood symmetric functions
Qλ (x;t) for all partitions λ, and then (A.54) can be obtained as a specialization at
t = −1. For details we refer to [Mac].
¿From the Hall-Littlewood approach, the Schur Q-function, for a strict parti-
tion λ with ℓ(λ) = ℓ and m ≥ ℓ, is equal to
( zi + z j )
λℓ
(A.55) Qλ (z1 , . . . , zm ) = 2ℓ ∑ w z λ1
· · · zℓ ∏ ∏ ,
1≤i≤ℓ i< j≤m zi − z j
1
w∈S /S m m−ℓ

where the symmetric group Sm acts by permuting the variables z1 , . . . , zm and Sm−ℓ
is the subgroup acting on zℓ+1 , . . . , zm .
258 A. Symmetric functions

A.3.4. A characterization of Γ. Denote by Γm the counterpart of Γ in m variables,


which consists of symmetric polynomials obtained by setting all but the first m
variables to be zero in the functions in Γ. We define
 := ∏ (zi + z j ) ∈ Γm .
1≤i< j≤m

(A.56) Let f be any symmetric polynomial in m variables. Then f  lies in Γm .

Proof. Let λ be a partition with ℓ(λ) ≤ m and let ρ = (m − 1, m − 2, . . . , 1, 0) so


that λ + ρ is a strict partition. Then Qλ+ρ (z1 , . . . , zm ) ∈ Γ. Now by (A.55) we see
that the Schur Q-polynomial Qλ+ρ (z1 , . . . , zm ), up to a 2-power, is equal to
( zi + z j )
∑ w zλ11 +m−1 · · · zλmm ∏ zi − z j
w∈Sm i< j≤m
(zi + z j )
= ∏ (z i − z j ) ∑ sgn(w)w(zλ11 +m−1 · · · zλmm )
i< j≤m w∈Sm
aλ+ρ
aρ i<∏
= (zi + z j ) = sλ (z1 , . . . , zm ).
j≤m

Since the Schur polynomials sλ (z1 , . . . , zm ) form a basis for the space Λm , this
proves the lemma. 

We have the following characterization of Γ.


(A.57) Let g ∈ Λ. Then g ∈ Γ if and only if it satisfies the cancelation property that
g|zi =−z j =t is independent of t for some i ̸= j.

Proof. Clearly, the power sums pr for r odd satisfy the above cancelation property,
and so does any element in Γ.
To prove the converse, it suffices to do so in the setting Γm of m variables. We
shall show that if g ∈ Λm is a symmetric polynomial in z1 , . . . , zm such that g|zi =−z j =t
is independent of t, then g can be written as a polynomial in the odd power sums
{p2k+1 | k ∈ Z+ }. Observe that g|zi =−z j =t is independent of t for some particular
choice of i, j with i ̸= j if and only it is independent of t for all (nonidentical) pairs
i, j.
We first note by (A.56) that q ∈ Γm for any symmetric polynomial q. We
proceed by induction on m. If m = 0, 1, then the cancelation property is vacuous.
Assume that m ≥ 2. Let p′2k+1 denote the (2k + 1)st power sum in the variables
z1 , . . . , zm−2 . Then the polynomial g|zm−1 =−zm =0 is symmetric in z1 , . . . , zm−2 , and
it satisfies the cancelation property. By induction hypothesis on m, there exists a
polynomial f such that
g|zm−1 =−zm =0 = f (p′1 , . . . , p′2r+1 ).
A.4. Boson-Fermion Correspondence 259

Set f ∗ := f (p1 , . . . , p2r+1 ). Then (g − f ∗ )|zm−1 =−zm =t = 0, and hence zm−1 + zm


divides g − f ∗ . Since g − f ∗ is symmetric in z1 , . . . , zm ,  divides g − f ∗ , and thus
there exists a symmetric polynomial q such that
g = f ∗ + q.
By (A.56), q is generated by the odd power sums in the variables z1 , . . . , zm , and
hence so is g. 

A.4. Boson-Fermion Correspondence


A.4.1. The Maya diagrams. The Maya diagrams are by definition the functions
f : 21 + Z → {±} such that f (n) = + and f (−n) = −, for n ≫ 0. In other words,
a Maya diagram is an assignment of the signs ± to the unit intervals on the real
line associated to the lattice Z with fixed asymptotic at infinity. These interval are
naturally labeled by 1/2 + Z, and we shall identify them.
A Maya diagram f can be reconstructed by knowing the subset m of 12 + Z
which is the preimage of + for f , which is simply an increasing sequence m =
{m j } j≥1 in 21 + Z, such that m1 < m2 < m3 < . . . and m j = m j−1 + 1 for j ≫ 0. In
this appendix, we shall identify Maya diagrams with such sequences m.
Example A.7. Consider the following Maya diagrams:
− − − − + + + + + +
ma : ··· p p p p p p p p p p ···
−3 −2 −1 0 1 2 3 4

− − + + + + + + + +
mb : ··· p p p p p p p p p p ···
−3 −2 −1 0 1 2 3 4

− + − + + – + – + +
mc : ··· p p p p p p p p p p ···
−3 −2 −1 0 1 2 3 4

The Maya diagram mb is obtained from ma by changing signs from − to + at inter-


vals − 32 and − 12 . The diagram mc is obtained from ma by changing signs from −
to + at intervals − 52 , − 12 and from + to − at 32 , 72 . In terms of infinite sequences, we
have ma = 21 + Z+ , mb = (− 32 , − 12 , 12 , 23 , 52 , . . .), and mc = (− 52 , − 12 , 12 , 25 , 92 , 11
2 , . . .).

A.4.2. Partitions. To each Maya diagram m = {m j } j≥1 , we assign an integer,


called charge,
( 1)
ℓm = lim m j − j +
j→∞ 2
and a partition
(A.58) λm = (1/2 − m1 + ℓm , 3/2 − m2 + ℓm , 5/2 − m3 + ℓm , . . .).
260 A. Symmetric functions

Recall that P denotes the set of all partitions. The following can be viewed as the
Boson-Fermion correspondence at combinatorial level.
(A.59) The map m 7→ (ℓm , λm ) is a bijection between the set of Maya diagrams and
the set Z × P of “charged partitions”.

Proof. We will use primarily the example of mc in Example A.7 to illustrate how
to visualize the Young diagram associated to λmc . Let us draw a new diagram
following the signs on a Maya diagram from far left to far right as follows. We
move to the southeast direction by one unit for each unit interval with a minus
sign on the Maya diagram of mc , and move to northeast by one unit for each unit
interval with a plus sign on. As a result, we obtain the zigzag path in the first
diagram below. The path is labeled by the unit intervals on the Maya diagram.
Extending the two line segments coming from minus infinity and going to plus
infinity, respectively, a finite bounded region emerges. This region can be further
partitioned by solid lines as in the second diagram, and from which we read off the
partition λmc = (3, 2, 2, 1) starting from the southwest row. It is easy to identify the
partition obtained this way for a Maya diagram m with the partition λm defined in
(A.58). The vertical dashed line on the first diagram indicates that the intersection
of the two infinite line segments corresponds to the coordinate 0 separating the two
unit intervals − 21 and 12 . From this we read off the charge ℓmc = 0.

@
@ ·
@ 1 3 5 7 ··
2 @
2 2 @2
..
@ @ @
9
. − 25 − 32 − 12 2

@
−2
7 @
@ @
@
@
@
@
@
0

@
@ ·
@ ··
@
@ @
..
@ @
9
. 2

@ @ @ @ @
−2
7
@ @ @ 2
7

@ @
− 25@ @ @ 2
5

@ @
− 32 @ @ 2
3

@
− 12 @ 1
2

Observe that uniformly shifting every sign on a Maya diagram m one unit in-
terval to the left gives rise to a Maya diagram whose charge is ℓm − 1. For example,
the mb can be obtained from ma by such a uniform shifting to the left twice in a
A.4. Boson-Fermion Correspondence 261

row, and so ℓmb = ℓma − 2. Such a shifting clearly corresponds to a uniform shift-
ing of the labels on the corresponding zigzag path in the first diagram. Hence, it
suffices to show that the set of Maya diagrams with charge 0 is in bijection with the
set of partitions via m 7→ λm . This bijection follows since we can easily reconstruct
the labeled zigzag path (and hence the Maya diagram) from a given partition λ by
first using a dashed diagonal line on the upside-down Young diagram for λ (whose
first southwest row has λ1 boxes) to locate the coordinate 0 point. 

A.4.3. Femions and fermionic Fock space. Let F be the vector space with a
linear basis given by the symbols |m⟩ parameterized by the Maya diagrams m.
Note that different Maya diagrams are related by changing signs at finitely
many intervals, and let us consider changing signs to a Maya diagram at one in-
terval at a time. Accordingly, we define the linear operators ψ+ −
n , ψn on F, for
n ∈ 2 + Z, as follows:
1

{
b i , mi+1 , . . .⟩, if mi = −n for some i,
(−1)i−1 | . . . , mi−1 , m
ψn |m⟩ =
+
0, otherwise,
{
(−1) | . . . , mi , n, mi+1 , . . .⟩, if mi < n < mi+1 for some i,
i
ψ−n |m⟩ =
0, otherwise,
where the hat m b i denotes the removal of the term. That is, ψ+
n corresponds to the

sign change at −n if the initial sign at −n is +, and ψn corresponds to the sign
change at n if the initial sign at n is −.

Remark A.8. Let V be a vector space with a standard basis vi , i ∈ 21 + Z, and let
{v∗i } denote its dual basis. It is instructive to regard each vector |m⟩ as a semi-
infinite wedge vm := vm1 ∧ vm2 ∧ vm3 ∧ · · · . The signs in the above definition of
ψ±n can be explained naturally by identifying ψn as the contraction operator v−n
+ ∗

and ψ− n as the exterior product with vn . In this way, the Maya diagram model is
equivalent to a semi-infinite wedge model of the fermionic Fock space.

We shall refer to | 12 , 23 , 25 , . . .⟩ as the vacuum vector of F and denote it by


|0⟩. Let us quote Dirac from his book “The Theory of Quantum Mechanics”: The
perfect vacuum is a region where all the states of positive energy are unoccupied
and all those of negative energy are occupied. We took the liberty in switching
“positive” with “negative” in this exposition.
Our convention is to consider the operators ψ±
n as odd operators and the com-
mutators among them to be anti-commutators: [A, B]+ = AB + BA.
(A.60) The following anti-commutation relations hold for all m, n ∈ 12 + Z:

n , ψm ]+ = δn,−m I,
[ψ+ n , ψm ]+ = 0,
[ψ+ +
[ψ− −
n , ψm ]+ = 0.

In particular, (ψ+ 2 − 2
n ) = (ψn ) = 0.
262 A. Symmetric functions

Proof. Follows by a direct verification using the definitions of ψ±


n . It corresponds
to the fact that the process of changing signs at two (possibly identical) intervals
are interchangeable. 

One also verifies directly the following.


(A.61) We have ψ+ −
n |0⟩ = ψn |0⟩ = 0, for n > 0.

Denote by C the Clifford (super)algebra generated by ψ±


n , for n ∈ 2 +Z, subject
1

to the relations in (A.60).


(A.62) A linear basis for F can be given by
(A.63) ψ− − −
−p1 ψ−p2 . . . ψ−pr ψ−q1 ψ−q2 . . . ψ−qs |0⟩,
+ + +

for p1 > p2 > . . . > pr > 0 and q1 > q2 > . . . > qs > 0 with r, s ≥ 0. The basis
element (A.63) is equal to m (up to a sign), where m is obtained from 12 + Z+ by
deleting q1 , . . . , qs and adding −p1 , . . . , −pr .
The Fock space F is an irreducible module of the Clifford algebra C generated
by the vacuum vector |0⟩.

Proof. The identification of the vector (A.63) with m up to a sign follows from the
definitions of ψ±
n , which correspond to the additions and removals of the + signs.
The irreducibility follows from the fact that a Maya diagram can be transformed
into another by changing the signs at one interval at a time. 
Example A.9. We have |mc ⟩ = −ψ− −
−5/2 ψ−1/2 ψ−7/2 ψ−3/2 |0⟩, for m from Exam-
+ + c

ple A.7.

We define the half-integral Frobenius coordinates for a partition λ as fol-


lows. Draw the Young diagram for λ in the English convention, and a diagonal
line from the vertex (0, 0) which divides the Young diagram into two halves. We
then record the lengths of rows of the upper half and the columns of the lower half
as (p1 , p2 , . . . , pr |q1 , q2 , . . . , qr ), where each pi , qi lie in 12 + Z+ , p1 > p2 > . . . >
pr > 0 and q1 > q2 > . . . > qr > 0. The sequence (p1 − 12 , p2 − 12 , . . . , pr − 12 |q1 −
2 , . . . , qr − 2 ) is Frobenius’ original notation for the partition λ (see Macdonald
1 1

[Mac, p.3]). For example the modified Frobenius coordinates for the partition
(3, 2, 2, 1) is ( 52 , 12 | 72 , 23 ).
It turns out that the indices pi , q j in (A.63) (at least when r = s) have natural
interpretations in terms of partitions.
(A.64) Let m be a Maya diagram of charge 0. Write (p1 , p2 , . . . , pr |q1 , q2 , . . . , qr )
for the modified Frobenius coordinates for the partition λm . Then, up to a sign,
|m⟩ is equal to ψ− − −
−p1 ψ−p2 . . . ψ−pr ψ−q1 ψ−q2 . . . ψ−qr |0⟩.
+ + +

r2
The sign in (A.64) can be computed to be (−1)q1 +...+qr − 2 .
A.4. Boson-Fermion Correspondence 263

Proof. The proof follows from two observations which work for any general Maya
diagram of charge 0, and we will illustrate below by the running example mc from
Example A.7: (i) In the diagrams appearing in the proof of (A.59), the 2 labels
corresponding to the + sign to the left of the dashed vertical line are − 25 , − 21 , and
the opposites of the 2 labels corresponding to the − sign to the right of the dashed
vertical line are − 32 , − 72 ; they are exactly the indices of the ψ± s in the formula
|mc ⟩ = −ψ− −
−5/2 ψ−1/2 ψ−7/2 ψ−3/2 |0⟩ by (A.62). (ii) The modified Frobenius coor-
+ +

dinates for the partition (3, 2, 2, 1) are ( 52 , 21 | 72 , 32 ), and they precisely correspond to
the 2 + 2 labels specified in (i). 

We form generating functions (called fermionic vertex operators) ψ± (z) in a


formal variable z by letting

∑ −n− 2
ψ− (z) = ∑ ψ− −n− 2
1 1
(A.65) ψ+ (z) = ψ+
nz , nz .
n∈ 21 +Z n∈ 12 +Z

A.4.4. Charge and energy. We define the energy operator L0 as a linear opera-
tor on F which diagonalizes the basis elements (A.63) with eigenvalue p1 + . . . +
pr + q1 + . . . + qs . In addition, we define a charge operator α0 as a linear operator
on F which diagonalizes the basis elements (A.63) with eigenvalue s − r. The ele-
ment (A.63) will be said of having energy p1 + . . . + pr + q1 + . . . + qs and charge
s − r. We have the following charge decomposition:

F= F(ℓ) ,
ℓ∈Z

where F(ℓ)
is spanned by the basis vectors (A.63) of charge ℓ. We shall see that the
notion of charge here matches with the notion of charge ℓm for a Maya diagram m.
(A.66) For each ℓ ∈ Z, F(ℓ) has a basis |m⟩ parameterized by the Maya diagrams
m = {m j } j≥1 that satisfy ℓm = ℓ.

Proof. By definition, ψ+n of charge 1 corresponds to the removal of a term from


a sequence m = (m1 , m2 , . . .), which increases the integer ℓm by 1 . On the other
hand, the operator ψ−n of charge −1 corresponds to the creation of a new sequence
(i.e. Maya diagram) by the insertion of a term into a sequence m, which decreases
the integer ℓm by 1. This means that the two notions of charge are identical up to
constant. This constant is zero once we observe that |0⟩ = |ma ⟩ is clearly in F(0)
and also ℓma = 0 by definition. 

For ℓ ∈ Z, we let
 +
 ψ 12 −ℓ . . . ψ− 32 ψ− 12 |0⟩,
+ +
 if ℓ > 0
(A.67) |ℓ⟩ = |0⟩, if ℓ = 0

 ψ−1 . . . ψ− 3 ψ− 1 |0⟩,
+ℓ − −
if ℓ < 0.
2 2 2
264 A. Symmetric functions

In particular, |1⟩ = ψ+
−1
|0⟩ and |-1⟩ = ψ−
−1
|0⟩. The following is easily verified.
2 2

(A.68) For each ℓ ∈ Z, the element |ℓ⟩ ∈ F(ℓ) has the minimal energy, which is
ℓ2 /2, among all vectors (A.63) of charge ℓ.

A.4.5. From Boson to Fermion. The Heisenberg Lie algebra Heis is the Lie alge-
bra generated by hk (k ∈ Z) and a cental element c with the commutation relation:
(A.69) [hm , hn ] = mδm,−n c, m, n ∈ Z.
In particular, h0 is central. The Heisenberg algebra Heis acts irreducibly on a
polynomial algebra B(ℓ) := C[p1 , p2 , . . .] in infinitely many variables p1 , p2 , . . ., by
letting h−k (k > 0) act as the multiplication operator by kpk , hk (k > 0) act as the
differentiation operator ddpk , c acts as the identity I, and h0 act as ℓI, for ℓ ∈ C. The
Heis-module B(ℓ) is irreducible, since any nonzero polynomial in B(ℓ) can be trans-
formed to a nonzero constant polynomial by a suitable sequence of differentiation
operators and then produce arbitrary monomials by the multiplication operators.
The Heis-module B(ℓ) is a highest weight module in the sense that hn .1 = 0 for
n > 0, and any highest weight Heis-module generated by a highest weight vector v
such that c.v = v and h0 .v = ℓv is isomorphic to B(ℓ) .
Similarly as in (5.40) we introduce the normal ordered product
{ −
− −ψ+ n ψm , if n = −m < 0
:ψ+ ψ : = −
m n
ψm ψn , otherwise.
+

Define the bosonic vertex operator α(z) = ∑k∈Z αk z−k−1 by letting


(A.70) α(z) = :ψ+ (z)ψ− (z):,
which is equivalent to defining componentwise
(A.71) αk = ∑ −
n ψk−n :,
:ψ+ k ∈ Z.
n∈ 12 +Z

Note that α0 = ∑n>0 ψ+ − − +


−n ψn − ∑n<0 ψn ψ−n is a well-defined linear operator on F.
The following is (one half of) the Boson-Fermion correspondence.
(A.72) Let ℓ ∈ Z. Heisenberg algebra Heis acts on F(ℓ) by letting c 7→ I, α0 7→ ℓ · I,
and hk 7→ αk , for k ∈ Z. That is, the following commutation relation holds:
(A.73) [αm , αn ] = mδm,−n I, m, n ∈ Z.
Moreover, the Heis-module F(ℓ) is irreducible and it is isomorphic to B(ℓ) .

Proof. We shall be free to use the following (anti-)commutator identities:


[AB,C] = A[B,C]+ − [A,C]+ B = A[B,C] + [A,C]B.
A.4. Boson-Fermion Correspondence 265

By a direct computation using the anti-commutator identity, we have

1
[αk , ψ± ±
n ] = ±ψk+n , k ∈ Z, n ∈ + Z.
2
The identity (A.73) follows by another computation using the commutator identity.
By definition the operators αk have charge 0. Hence, αk (F(ℓ) ) ⊆ F(ℓ) , that is,
F(ℓ) is a Heis-module.
For p, q > 0 and m of charge ℓ, ψ− ′
−p ψ−q |m⟩ = |m ⟩ if it is nonzero, where m
+ ′

is obtained from m by replacing a term q appearing in m by a new term −p, and


the charge of m′ remains to be ℓ. Recalling the definition (A.58) of the partition
λm , we see that |λm′ | = |λm | + p + q. This implies that the energy of m is equal to
|λm | up to a universal constant. The constant is determined to be ℓ2 /2 by recalling
(A.68) and noting λ|ℓ⟩ = 0. / Hence, the energy of the element |m⟩ is |λm | + ℓ2m /2.
From this it follows that the q-dimension (or graded dimension) of F(ℓ) is given by
ℓ2
L0 q2
tr|F(ℓ) q = ∞ .
∏k=1 (1 − qk )

On the other hand, note that the energy operator L0 satisfies (and is indeed
characterized by) the following properties: (i) L0 |0⟩ = 0; (ii) [L0 , ψ± ±
−n ] = nψ−n , for
all n ∈ 2 + Z. It follows from (A.71) that
1

(A.74) [L0 , α−k ] = nα−k , k ∈ Z.

Note that αk |ℓ⟩ = 0, for k > 0, and α0 |ℓ⟩ = ℓ|ℓ⟩. Hence, the irreducible Heis-
submodule of F(ℓ) generated by |ℓ⟩ is isomorphic to B(ℓ) and has q-dimension
ℓ2
q2
equal to
∏∞ k , the same as the graded dimension of F(ℓ) . It follows that the
k=1 (1−q )
Heis-module F(ℓ) is irreducible. 

Remark A.10. From the proof, we have [α0 , ψ± ±


n ] = ±ψn for n ∈ 2 + Z. This
1

together with α0 |0⟩ = 0 implies that α0 here can be identified with the charge
operator defined earlier.

The other half of the Boson-Fermion correspondence allows one to reconstruct


the fermions ψ± (z) in terms of the boson α(z). We will formulate the statement
but skip its proof. Denote by S : F(ℓ) → F(ℓ+1) the shift operator which sends |ℓ⟩ to
|ℓ + 1⟩ and commutes with the action of αk for all k ̸= 0. Then

ψ± (z) = S±1 : exp(± α(z)dz):
z− j z− j
= S±1 z±α0 e∓ ∑ j<0 j α j ∓ ∑ j>0
e j αj
.
266 A. Symmetric functions

A.4.6. Jacobi triple product identity.


(A.75) Let q, y be formal variables. The Jacobi triple product identity holds:
∞ ℓ2
∏(1 − qk )(1 + qk− y)(1 + qk− 2 y−1 ) = ∑q
1 1
2 2 yℓ .
k=1 ℓ∈Z

Proof. We compute the trace of the operator qL0 yα0 on F in two ways.
First, by the linear basis (A.62) of F and knowing each fermionic operator ψ±
−ℓ
contributes ℓ to the energy and ±1 to the charge, we have
tr|F qL0 yα0 = ∏ (1 + qn y)(1 + qn y−1 ).
n∈ 12 +Z+

On the other hand, F(ℓ) for each ℓ ∈ Z is identified with the bosonic Fock space
B(ℓ) by (A.72). Using (A.68), we compute that
ℓ2
q 2 yℓ
L0 α0
tr|F q y = ∑ tr|B (ℓ)
L0 α0
q y =∑ ∞ .
ℓ∈Z ∏k=1 (1 − q )
k
ℓ∈Z
The Jacobi triple product identity follows now by equating the two formulas for
the trace and clearing the denominator. 

A.5. Notes
Section A.1. The materials on symmetric functions and the Frobenius characteristic
map are fairly standard, and they can be found in Macdonald [Mac] in possibly
different order.
Section A.2. The characterization of supersymmetric polynomials in A.2.1 ap-
peared in Stembridge [St]. The results on super Schur functions and super Cauchy
identity in A.2.2 can be found in [BeR, Sv1].
Section A.3. The materials can be found in Macdonald [Mac] and in Józefiak
[Jo2]. The characterization of the ring Γ in A.3.4 appeared in Pragacz [Pr, Theo-
rem 2.11], and is used in Section 2.3.2.
Section A.4. The materials on Boson-Fermion correspondence are standard.
The use of Maya diagrams follows Miwa-Jimbo-Date [MJD]. Additional applica-
tions of Boson-Fermion correspondence, most notably to soliton equations, can be
found in Kac-Raina [KR] and Miwa-Jimbo-Date [MJD].
Bibliography

[Ar] F. Aribaud, Une nouvelle démonstration d’un théorème de R. Bott et B. Kostant,


Bull. Soc. Math. France 95 (1967), 205–242.
[ABP] M. Atiyah, R. Bott, and V.K. Patodi, On the heat equation and the index theorem, In-
vent. Math. 19 (1973), 279V330.
[AYY] S. Azam, H. Yamane, and M. Yousofzadeh, Classification of Finite Dimensional Irre-
ducible Representations of Generalized Quantum Groups via Weyl Groupoids, preprint,
arXiv:1105.0160.
[BB] A. Beilinson and J. Bernstein, Localisation de g-modules, C.R. Acad. Sci. Paris Ser. I Math.
292 (1981), 15–18.
[BeR] A. Berele and A. Regev, Hook Young Diagrams with Applications to Combinatorics and to
Representations of Lie Superalgebras, Adv. Math. 64 (1987), 118–175.
[BL] I.N. Bernstein and D.A. Leites, A formula for the characters of the irreducible finite-
dimensional representations of Lie superalgebras of series gl and sl, (Russian) C. R. Acad.
Bulgare Sci. 33 (1980), 1049–1051.
[BKN] B. Boe, J. Kujawa, and D. Nakano, Cohomology and support varieties for Lie superalge-
bras, Trans. Amer. Math. Soc. 362 (2010), 6551–6590.
[Br1] J. Brundan, Kazhdan-Lusztig polynomials and character formulae for the Lie superalgebra
gl(m|n), J. Amer. Math. Soc. 16 (2003), 185–231.
[Br2] J. Brundan, Kazhdan-Lusztig polynomials and character formulae for the Lie superalgebra
q(n), Adv. Math. 182 (2004), 28–77.
[BK] J. Brundan and A. Kleshchev, Projective representations of symmetric groups via Sergeev
duality, Math. Z. 239 (2002), 27–68.
[BrK] J. Brundan and J. Kujawa, A new proof of the Mullineux conjecture, J. Algebraic Combin. 18
(2003), 13–39.
[BrS] J. Brundan and C. Stroppel, Highest Weight Categories Arising from Khovanov’s Diagram
Algebras IV: the general linear supergroup, preprint, arXiv:0907.2543.
[BK] J.L. Brylinski and M. Kashiwara, Kazhdan-Lusztig conjecture and holonomic systerms,
Invent. Math. 64 (1981), 387–410.
[CaL] B. Cao and L. Luo, Generalized Verma Modules and Character Formulae for osp(3|2m),
preprint, arXiv:1001.3986.

267
268 Bibliography

[Car] R. Carter, Lie Algebras of Finite and Affine Type. Cambridge Studies in Advanced Mathe-
matics 96. Cambridge University Press, Cambridge, 2005. xviii+632 pp.
[CK] S.-J. Cheng and J.-H. Kwon, Howe duality and Kostant homology formula for infinite-
dimensional Lie superalgebras, Int. Math. Res. Not. 2008, Art. ID rnn 085, 52 pp.
[CKL] S.-J. Cheng, J.-H. Kwon, and N. Lam, A BGG-type resolution for tensor modules over
general linear superalgebra, Lett. Math. Phys. 84 (2008), 75–87.
[CKW] S.-J. Cheng, J.-H. Kwon, and W. Wang, Kostant homology formulas for oscillator modules
of Lie superalgebras, Adv. Math. 224 (2010), 1548–1588.
[CL1] S.-J. Cheng and N. Lam, Infinite-dimensional Lie superalgebras and hook Schur functions,
Commun. Math. Phys. 238 (2003), 95–118.
[CL2] S.-J. Cheng and N. Lam, Irreducible characters of general linear superalgebra and super
duality, Commun. Math. Phys. 280 (2010), 645–672.
[CLW] S.-J. Cheng, N. Lam and W. Wang, Super duality and irreducible characters of ortho-
symplectic Lie superalgebras, Invent. Math. 183 (2011), 189–224.
[CLW2] S.-J. Cheng, N. Lam and W. Wang, Super duality for general linear Lie superalgebras and
applications, preprint (2011).
[CLZ] S.-J. Cheng, N. Lam, and R.B. Zhang, Character formula for infinite dimensional unitariz-
able modules of the general linear superalgebra, J. Algebra 273 (2004), 780–805.
[CW2] S.-J. Cheng and W. Wang, Remarks on Schur-Howe-Sergeev duality, Lett. Math. Phys. 52
(2000), 143–153.
[CW1] S.-J. Cheng and W. Wang, Howe duality for Lie superalgebras, Compositio Math. 128
(2001), 55–94.
[CW3] S.-J. Cheng and W. Wang, Lie subalgebras of differential operators on the super circle,
Publ. Res. Inst. Math. Sci. 39 (2003), 545–600.
[CW4] S.-J. Cheng and W. Wang, Brundan-Kazhdan-Lusztig and Super Duality Conjectures, Publ.
Res. Inst. Math. Sci. 44 (2008), 1219–1272.
[CW5] S.-J. Cheng and W. Wang, Dualities for Lie superalgebras, Lecture notes for Shanghai
summer school 2009, Surveys of Modern Mathematics 2, International Press and Higher
Education Press, pp. 1–46, arXiv:1001.0074.
[CWZ2] S.-J. Cheng, W. Wang and R.B. Zhang, A Fock space approach to representation theory of
osp(2|2n), Transform. Groups 12 (2007), 209–225.
[CWZ] S.-J. Cheng, W. Wang, and R.B. Zhang, Super duality and Kazhdan-Lusztig polynomials,
Trans. Amer. Math. Soc. 360 (2008) 5883–5924.
[CZ1] S.-J. Cheng and R.B. Zhang, Analogue of Kostant’s u-homology formula for general linear
Superalgebras, Int. Math. Res. Not. 2004, 31–53.
[CZ2] S.-J. Cheng and R.B. Zhang, Howe duality and combinatorial character formula for or-
thosymplectic Lie superalgebras, Adv. Math. 182 (2004), 124–172.
[DJKM] E. Date, M. Jimbo, M. Kashiwara and T. Miwa: Transformation groups for soliton equa-
tions. III. Operator approach to the Kadomtsev-Petviashvili equation, J. Phys. Soc. Japan
50 (1981), 3806–3812. Transformation groups for soliton equations. IV. A new hierarchy of
soliton equations of KP-type, Phys. D 4 (1981/82), 343–365.
[DES] M. Davidson, E. Enright, R. Stanke, Differential Operators and Highest Weight Represen-
tations, Mem. Amer. Math. Soc. 94 (1991) no. 455.
[Don] S. Donkin, On tilting modules for algebraic groups, Math. Z. 212 (1993), 39–60.
[En] T. Enright, Analogues of Kostant’s u-cohomology formulas for unitary highest weight mod-
ules, J. Reine Angew. Math. 392 (1988), 27–36.
Bibliography 269

[FSS] L. Frappat, A. Sciarrino and P. Sorba, Dictionary on Lie algebras and superalgebras. Aca-
demic Press, Inc., San Diego, CA, 2000.
[Fr] I. Frenkel, Representations of affine Lie algebras, Hecke modular forms and Kortweg-de
Vries type equations, Lect. Notes Math. 933 (1982), 71–110.
[Ger] J. Germoni, Indecomposable representations of osp(3, 2), D(2, 1; α) and G(3). Colloquium
on Homology and Representation Theory (Spanish) (Vaquerias, 1998). Bol. Acad. Nac.
Cienc. (Cordoba) 65 (2000), 147–163.
[GW] R. Goodman and N. Wallach, Representations and invariants of the classical groups. Ency-
clopedia of Mathematics and its Applications, 68. Cambridge University Press, Cambridge,
1998.
[Go] M. Gorelik, The Kac construction of the centre of U(g) for Lie superalgebras, J. Nonlinear
Math. Phys. 11 (2004), 325–349.
[GS] C. Gruson and V. Serganova, Cohomology of generalized supergrassmannians and char-
acter formulae for basic classical Lie superalgebras, Proc. Lond. Math. Soc. 101 (2010),
852–892.
[H1] R. Howe, Remarks on classical invariant theory, Trans. Amer. Math. Soc. 313 (1989), 539–
570.
[H2] R. Howe, Perspectives on invariant theory: Schur duality, multiplicity-free actions and
beyond, The Schur Lectures, Israel Math. Conf. Proc. 8, Tel Aviv (1992), 1–182.
[HLT] P.-Y. Huang, N. Lam, and T.-M. Tu, Super duality and homology of unitarizable modules
of Lie algebras, Pub. RIMS (to appear), arXiv:1012.1087.
[Hum] J. E. Humphreys, Representations of Semisimple Lie Algebras in the BGG Category O,
Graduate Studies in Mathematics, 94. American Mathematical Society, Providence, RI,
2008.
[IK] K. Iohara and Y. Koga, Second homology of Lie superalgebras, Math. Nachr. 278 (2005),
1041–1053.
[Jo1] T. Józefiak, Semisimple superalgebras, In: Algebra–Some Current Trends (Varna, 1986),
pp. 96–113, Lect. Notes in Math. 1352, Springer-Berlag, Berlin-New York, 1988.
[Jo2] T. Józefiak, Characters of projective representations of symmetric groups, Expo. Math. 7
(1989), 193–247.
[Jo3] T. Józefiak, A class of projective representations of hyperoctahedral groups and Schur Q-
functions, Topics in Algebra, Banach Center Publ., 26, Part 2, PWN-Polish Scientific Pub-
lishers, Warsaw (1990), 317–326.
[K4] V. Kac, Classification of simple Lie superalgebras (in Russian), Funkcional. Anal. i Priložen
9 (1975), 91–92.
[K1] V. Kac, Lie superalgebras, Adv. Math. 26 (1977), 8–96.
[K2] V. Kac, Representations of classical Lie superalgebras. Differential geometrical methods
in mathematical physics, II (Proc. Conf., Univ. Bonn, Bonn, 1977), pp. 597–626, Lecture
Notes in Math. 676, Springer, Berlin, 1978.
[K5] V. Kac, Laplace operators of infinite-dimensional Lie algebras and theta functions, Proc.
Nat. Acad. Sci. USA 81 (1984), 645–647.
[K3] V. Kac, Infinite dimensional Lie algebras. Third edition. Cambridge University Press, Cam-
bridge, 1990.
[KR] V. Kac and A. Raina, Bombay Lectures on Highest Weight Representations of Infinite Di-
mensional Lie Algebras, Advanced Series in Mathematical Physics 2. World Scientific Pub-
lishing Co., Inc., Teaneck, NJ, 1987.
270 Bibliography

[KW] V. Kac and M. Wakimoto, Integrable highest weight modules over affine superalgebras and
Appell’s function, Commun. Math. Phys. 215 (2001), 631–682.
[KWY] V. Kac, W. Wang, and C. Yan, Quasifinite representations of classical Lie subalgebras of
W1+∞ , Adv. Math. 139 (1998) 56–140.
[KK] S.-J. Kang and J.-H. Kwon, Graded Lie superalgebras, supertrace formula, and orbit Lie
superalgebras, Proc. London Math. Soc. 81 (2000), 675–724.
[KaV] M. Kashiwara and M. Vergne, On the Segal-Shale-Weil representations and harmonic poly-
nomials, Invent. Math. 44 (1978), 1–47.
[KL] D. Kazhdan and G. Lusztig, Representations of Coxeter groups and Hecke algebras, Invent.
Math. 53 (1979), 165–184.
[KV] A. Knapp and D. Vogan, Cohomological Induction and Unitary Representations, Princeton
Mathematical Series, 45. Princeton University Press, Princeton, NJ, 1995.
[Kos] B. Kostant, Lie Algebra Cohomology and the generalized Borel-Weil Theorem,
Ann. Math. 74 (1961), 329–387.
[KP] H. Kraft and C. Procesi, Classical Invariant Theory, A Primer. 1996, 128 pp. Available at
http://www.math.unibas.ch/ kraft/Papers/KP-Primer.pdf.
[Ku] S. Kumar, Kac-Moody groups, their flag varieties and representation theory. Progress in
Mathematics, 204. Birkhauser Boston, Inc., Boston, MA, 2002.
[LZ] N. Lam and R.B. Zhang, Quasi-finite modules for Lie superalgebras of infinite rank,
Trans. Amer. Math. Soc. 358 (2006), 403–439.
[LSS] D. Leites, M. Saveliev, and V. Serganova, Embedding of osp(N/2) and the associated non-
linear supersymmetric equations. Group theoretical methods in physics, Vol. I (Yurmala,
1985), 255–297, VNU Sci. Press, Utrecht, 1986.
[Liu] L. Liu, Kostant’s Formula for Kac-Moody Lie Algebras, J. Algebra 149 (1992), 155–178.
[Mac] I. G. Macdonald, Symmetric functions and Hall polynomials, Second Edition, Oxford Math-
ematical Monographs. Oxford Science Publications. The Clarendon Press, Oxford Univer-
sity Press, New York, 1995.
[Ma] Yu. Manin, Gauge field theory and complex geometry, Grundlehren der mathematischen
Wissenschaften 289, Second Edition, Springer-Verlag, Berlin, 1997.
[MM] J. Milnor and J. Moore, On the Structure of Hopf Algebras, Ann. Math. 81 (1965), 211–264.
[Mi] B. Mitchell, Theory of categories, Pure and Applied Mathematics XVII, Academic Press,
New York-London 1965.
[MJD] T. Miwa, M. Jimbo, and E. Date, Solitons. Differential equations, symmetries and infinite-
dimensional algebras, Cambridge Tracts in Mathematics 135. Cambridge University Press,
Cambridge, 2000.
[MJ] E. Moens and J. van der Jeugt, A determinantal formula for supersymmetric Schur polyno-
mials, J. Algebraic Combin. 17 (2003), 283–307.
[Mu] I. Musson, Lie superalgebras, Clifford algebras, induced modules and nilpotent orbits,
Adv. Math. 207 (2006), 39–72.
[Naz] M. Nazarov, Capelli identities for Lie superalgebras, Ann. Sci. École Norm. Sup. 30 (1997),
847–872.
[Pe] I. Penkov, Characters of typical irreducible finite-dimensional q(n)-modules, Funct. Anal.
App. 20 (1986), 30–37.
[PS] I. Penkov and V. Serganova, Cohomology of G/P for classical complex Lie supergroups G
and characters of some atypical G-modules, Ann. Inst. Fourier 39 (1989), 845–873.
[PS2] I. Penkov and V. Serganova, Characters of finite-dimensional irreducible q(n)-modules,
Lett. Math. Phys. 40 (1997), 147–158.
Bibliography 271

[Po] N. Popescu, Abelian categories with applications to rings and modules. London Mathemat-
ical Society Monographs 3, Academic Press, London-New York, 1973.
[Pr] P. Pragacz, Algebro-geometric applications of Schur S- and Q-polynomials. Topics in in-
variant theory (Paris, 1989/1990), 130–191, Lecture Notes in Math. 1478, Springer, Berlin,
1991.
[Re] E. W. Read, The α-regular classes of the generalized symmetric groups, Glasgow Math. J.
17 (1976), 144–150.
[RW] A. Rocha-Caridi and N. Wallach: Projective modules over graded Lie algebras I, Math.
Z. 180 (1982), 151–177.
[Ro] L. Ross, Representations of graded Lie algebras, Trans. Amer. Math. Soc. 120 (1965) 17–
23.
[Sa1] B. Sagan, Shifted tableaux, Schur Q-functions, and a conjecture of R. Stanley, J. Com-
bin. Theory Ser. A 45 (1981), 62–103.
[San] J. Santos, Foncteurs de Zuckermann pour les superalgébres de Lie, J. Lie Theory 9 (1999),
69–112.
[Sch] M. Scheunert, The Theory of Lie superalgebras, Lect. Notes in Math. 716. Springer, Berlin,
1979.
[SNR] M. Scheunert, W. Nahm, and V. Rittenberg, Classification of all simple graded Lie algebras
whose Lie algebra is reductive I, II. Construction of the exceptional algebras, J. Math. Phys.
17 (1976), 1626–1639, 1640–1644.
[Sva] V. Serganova, Kazhdan-Lusztig polynomials and character formula for the Lie superalge-
bra gl(m|n), Selecta Math. (N.S.) 2 (1996), 607–651.
[Sva2] V. Serganova, Kac-Moody superalgebras and integrability, In: Developments and trends in
infinite-dimensional Lie theory, 169–218, Progr. Math. 288, Birkhäuser, 2011.
[Sv4] A. Sergeev, The centre of enveloping algebra for Lie superalgebra Q(n, C).
Lett. Math. Phys. 7 (1983), no. 3, 177–179.
[Sv1] A. Sergeev, The tensor algebra of the identity representation as a module over the Lie
superalgebras gl(n, m) and Q(n), Math. USSR Sbornik 51 (1985), 419–427.
[Sv3] A. Sergeev, The invariant Polynomials of simple Lie superalgebras, Represent. Theory 3
(1999), 250–280 (electronic).
[Sv2] A. Sergeev, An Analog of the Classical Invariant Theory, I, II, Michigan J. Math. 49 (2001),
113–146, 147–168.
[SVe] A. Sergeev and A. Veselov, Grothendieck rings of basic classical Lie superalgebras, Annals
of Math. 173 (2011), 663-703.
[SW] B. Shu and W. Wang, Modular representations of the ortho-symplectic supergroups, Proc.
London Math. Soc. 96 (2008), 251–271.
[So] W. Soergel, Character formulas for tilting modules over Kac-Moody algebras, Represent.
Theory (electronic) 2 (1998), 432–448.
[St] J. Stembridge, A characterization of supersymmetric polynomials, J. Algebra 95 (1985),
439–444.
[Su] Y. Su, Composition factors of Kac modules for the general linear Lie superalgebras, Math.
Z. 252 (2006), 731–754.
[SZ] Y. Su and R.B. Zhang, Character and dimension formulae for general linear superalgebra,
Adv. Math. 211 (2007), 1–33.
[Ta] J. Tanaka, On homology and cohomology of Lie superalgebras with coefficients in their
finite-dimensional representations, Proc. Japan Acad. Ser. A Math. Sci. 71 (1995), 51–53.
272 Bibliography

[J] J. Van der Jeugt, Character formulae for Lie superalgebra C(n), Comm. Algebra 19,
(1991), 199–222.
[JHKT] J. Van der Jeugt, J.W.B. Hughes, R. C. King and J. Thierry-Mieg, Character formulas for
irreducible modules of the Lie superalgebras sl(m/n), J. Math. Phys. 31 (1990), 2278–
2304.
[Vo] D. Vogan, Irreducible characters of semisimple Lie Groups II: The Kazhdan-Lusztig Con-
jectures, Duke Math. J. 46 (1979), 805–859.
[Wal] C.T.C. Wall, Graded Brauer groups, J. Reine Angew. Math. 213 (1964), 187–199.
[Wa] W. Wang, Duality in infinite dimensional Fock representations, Commun. Contem. Math. 1
(1999), 155–199.
[WZ] W. Wang and L. Zhao, Representations of Lie superalgebras in prime characteristic I, Proc.
London Math. Soc. 99 (2009), 145–167.
[We] H. Weyl, The classical groups. Their invariants and representations. Fifteenth printing.
Princeton Landmarks in Mathematics. Princeton University Press, Princeton, NJ, 1997.
[Zou] Y. Zou, Categories of finite-dimensional weight modules over type I classical Lie superal-
gebras, J. Algebra 180 (1996), 459–482.
Index

εδ-sequence, 22 degree
atypicality, 62, 70
even, 2
adjoint action, 3
odd, 2
adjoint map, 3
derivation, 3
algebra
diagram
Clifford, 141
head, 188
Hecke-Clifford, 106
master, 188
univeral enveloping, 30
Maya, 259
Weyl, 141
charge of, 259
Weyl-Clifford, 141
skew, 246
tail, 188
bilinear form dimension, 2
even, 6 duality
invariant, 3 super, 230
odd, 6
skew-supersymmetric, 6 energy, 263
supersymmetric, 6 extension, 223
boundary, 214 exterior algebra, 10

central character, 52 FFT, 119


character formula, 213 multilinear
characteristic map, 108 GL(V ), 122
Frobenius, 248 polynomial
charge, 263 GL(V ), 123
Chevalley automorphism, 5 O(V ), Sp(V ), 129
coboundary, 215 supersymetric
cohomology, 215 O(V ), Sp(V ), 137
Kostant, 216 tensor
restricted, 215 GL(V ), 121
unrestricted, 226 O(V ), Sp(V ), 133
column determinant, 153 First Fundamental Theorem, 119
conjugacy class Frobenius coordinates, 72
split, 89 half-integral, 262
contraction, 122 modified, 74
coroot, 14 functor

273
274 Index

T , 209 module
T , 209 Kac, 40
duality, 216 oscillator
truncation, 202 osp(2m|2n), 162
fundamental system, 19 spo(2m|2n), 167
parabolic Verma, 199
group polynomial, 100, 198
orthogonal, 127 spin, 89
symplectic, 127 basic, 109
type M, 86
harmonic, 161 type Q, 86
highest weight, 32 multiplicity-free
highest weight vector, 33 strongly, 146
homology, 214
Kostant, 216 nilradical, 199
homomorphism opposite, 199
Harish-Chandra, 52 normal ordered product, 171
Lie superalgebra, 3
module, 2 odd reflection, 26
Howe dual pair, 148 odd trace, 18
operator
ideal, 2 boundary, 214
charge, 263
Jacobi triple product identity, 266 coboundary, 215
energy, 263
Kac module, 40
Kazhdan-Lusztig-Vogan polynomial, 220 parity, 2
partition, 239
level, 169 size of, 239
Lie superalgebra, 3 dominance order, 239
e
g, g, g, 196 even, 247
basic, 13 generalized, 174
Cartan type hook, 45
H(n), 12 length of, 239
S(n), 11 odd, 103
W (n), 10 part of, 239
e
S(n), 11 strict, 103
classical, 13 Pfaffian, 255
exceptional Pieri’s formula, 247
D(2, 1, α), 9 Poincaré-Birkhoff-Witt Theorem, 30
F(4), 10 polarization, 123
G(3), 10 polynomial
general linear, 4 doubly symmetric, 249
ortho-symplectic, 6 Kazhdan-Lusztig-Vogan, 220
periplectic, 9 Schur, 243
queer, 8 supersymmetric, 249
solvable, 31 positive system, 19
special linear, 5
type a, 190 reflection
type b• , 191 odd, 26
type b, 192 restitution, 123
type c, 191 restricted dual, 215
type d, 192 ring of symmetric functions, 240
linkage principle root
gl,osp, 62 even, 14
q, 70 isotropic, 15
Littlewood-Richardson coefficient, 246 odd, 14
Index 275

simple, 19 half-integer, 44
root system, 13 highest, 32
integer, 44
Schur function, 243 linked
skew, 246 gl,osp, 60
super, 252 q, 69
Schur polynomial, 243 polynomial, 100
strongly multiplicity-free, 146 Weyl group, 14
subalgebra, 2 Weyl symbol, 142
Borel, 20 Weyl vector, 20
Cartan, 13
Levi, 199
parabolic, 199
subspace, 2
super duality, 230
superalgebra, 2
semisimple, 86
simple, 2
supercharacter, 51
superdimension, 2
supergroup, 88
superspace, 2
supertableau, 94
content of, 94
supertrace, 5
supertranspose, 5
symbol map, 143
symmetric function
complete, 240
elementary, 240
involution ω, 242
monomial, 240
power, 240
ring of, 240
system
fundamental, 19
positive, 19
root, 13

tableau, 246
semistandard, 246
trace
odd, 18
typical, 62, 70

univeral enveloping algebra, 30

vector
highest weight, 33
singular, 65
vacuum, 261
virtual character, 42

weight
dominant, 199, 202
dominant integral, 41
extremal, 71

Das könnte Ihnen auch gefallen