Groups Quantum
Groups Quantum
Editorial Board
I.H. Ewing F.W. Gehring P.R. Halmos
Graduate Texts in Mathematics
Quantum Groups
With 88 Illustrations
Editorial Board
J.H. Ewing F. W. Gehring P.R. Halmos
Department of Department of Department of
Mathematics Mathematics Mathematics
Indiana University University of Michigan Santa Clara University
Bloomington, IN 47405 Ann Arbor, MI 48109 Santa Clara, CA 95053
USA USA USA
987654321
ISBN 978-1-4612-6900-7
Preface
Christian Kassel
March 1994, Strasbourg
Preface v
II
Tensor Products 23
1 Tensor Products of Vector Spaces 23
2 Tensor Products of Linear Maps 26
3 Duality and Traces . . . . . . . . 29
4 Tensor Products of Algebras .. 32
5 Tensor and Symmetric Algebras 34
6 Exercises 36
7 Notes ............... 38
viii Contents
III
The Language of Hopf Algebras 39
1 Coalgebras.. 39
2 Bialgebras . . . . . . . . . . . . 45
3 Hopf Algebras . . . . . . . . . . 49
4 Relationship with Chapter I. The Hopf Algebras GL(2)
and SL(2). . . . . . . . . . . 57
5 Modules over a Hopf Algebra. . . . . . . . . . . . . . . 57
6 Comodules......................... 61
7 Comodule-Algebras. Coaction of SL(2) on the Affine Plane 64
8 Exercises 66
9 Notes.............................. 70
IV
The Quantum Plane and Its Symmetries 72
1 The Quantum Plane . . . . . . . . . . . . . 72
2 Gauss Polynomials and the q- Binomial Formula 74
3 The Algebra Mq(2) . . . . . . . . . . 77
4 Ring-Theoretical Properties of Mq(2) . 81
5 Bialgebra Structure on Mq(2) . . . . . . 82
6 The Hopf Algebras GLq(2) and SLq(2) 83
7 Coaction on the Quantum Plane 85
8 Hopf *-Algebras 86
9 Exercises 88
10 Notes . . . . . . 90
V
The Lie Algebra of SL(2) 93
1 Lie Algebras . . . . . 93
2 Enveloping Algebras . . 94
3 The Lie Algebra .5[(2) . 99
4 Representations of .5[(2) 101
5 The Clebsch-Gordan Formula. 105
6 Module-Algebra over a Bialgebra. Action of .5[(2) on the
Affine Plane . . . . . . . . . . . . . . . . . . . . . . . 107
7 Duality between the Hopf Algebras U(.5[(2)) and SL(2) 109
8 Exercises 11 7
9 Notes............................ 119
VI
The Quantum Enveloping Algebra of .5[(2) 121
1 The Algebra Uq(.5[(2)) . . . . . . . . . . . . . 121
2 Relationship with the Enveloping Algebra of .5[(2) 125
3 Representations of Uq . . . . . . . . . . . . . . . . 127
4 The Harish-Chandra Homomorphism and the Centre of Uq 130
Contents ix
VII
A Hopf Algebra Structure on Uis[(2)) 140
1 Comultiplication............. 140
2 Semi simplicity . . . . . . . . . . . . . . 143
3 Action of Uq(.s[(2)) on the Quantum Plane 146
4 Duality between the Hopf Algebras Uq(.s[(2)) and SLq(2) 150
5 Duality between Uq (.s[(2))-Modules and SL q(2)-Comodules 154
6 Scalar Products on Uq (.s[(2))-Modules 155
7 Quantum Clebsch-Gordan . 157
8 Exercises 162
9 Notes............ 163
IX
Drinfeld's Quantum Double 199
1 Bicrossed Products of Groups . . . . . . 199
2 Bicrossed Products of Bialgebras . . . . . 202
3 Variations on the Adjoint Representation 207
4 Drinfeld's Quantum Double . . . . . . . . 213
5 Representation-Theoretic Interpretation of the
Quantum Double . . . . 220
6 Application to Uq(.s[(2)) . 223
7 R- Matrices for U q 230
8 Exercises 236
9 Notes . . . . . . . 238
x Contents
XI
Tensor Categories 275
1 The Language of Categories and Functors . 275
2 Tensor Categories . . . . . . . 281
3 Examples of Tensor Categories . . . . . . . 284
4 Tensor Functors . . . . . . . . . . . . . . . 287
5 Turning Tensor Categories into Strict Ones 288
6 Exercises 291
7 Notes . . . . . . . . . . . . . . . . . . . . . 293
XII
The Tangle Category 294
1 Presentation of a Strict Tensor Category 294
2 The Category of Tangles . . . . . . . . . 299
3 The Category of Tangle Diagrams . . . . 302
4 Representations of the Category of Tangles 305
5 Existence Proof for Jones-Conway Polynomial 311
6 Exercises 313
7 Notes . . . . . . . . . . . . . . . . . . . . . . . 313
XIII
Braidings 314
1 Braided Tensor Categories . . . . . 314
2 The Braid Category . . . . . . . . . 321
3 Universality of the Braid Category. 322
4 The Centre Construction . . . . . . 330
5 A Categorical Interpretation of the Quantum Double 333
6 Exercises 337
7 Notes........................... 338
Contents xi
XIV
Duality in Tensor Categories 339
1 Representing Morphisms in a Tensor Category 339
2 Duality . . . . . . . . . . . . . . 342
3 Ribbon Categories . . . . . . . . 348
4 Quantum Trace and Dimension . 354
5 Examples of Ribbon Categories. 358
6 Ribbon Algebras 361
7 Exercises 365
8 Notes . . . . . . 366
XV
Quasi-Bialgebras 368
1 Quasi-Bialgebras . . . . . 368
2 Braided Quasi-Bialgebras 371
3 Gauge Transformations . 372
4 Braid Group Representations . 377
5 Quasi-Hopf Algebras. 379
6 Exercises 381
7 Notes . . . . . . . . . 381
XVII
Drinfeld and Jimbo's Quantum Enveloping Algebras 403
1 Semisimple Lie Algebras . . . . . . 403
2 Drinfeld-Jimbo Algebras . . . . . . 406
3 Quantum Group Invariants of Links 410
4 The Case of s[(2) 412
5 Exercises 418
6 Notes....... 418
xii Contents
XVIII
Cohomology and Rigidity Theorems 420
1 Cohomology of Lie Algebras . . . . . . . . . . 420
2 Rigidity for Lie Algebras . . . . . . . . . . . . 424
3 Vanishing Results for Semisimple Lie Algebras 427
4 Application to Drinfeld-Jimbo Quantum Enveloping Algebras 430
5 Cohomology of Coalgebras . . . . . . . . . . . . . . . . .. 431
6 Action of a Semisimple Lie Algebra on the Cobar Complex 434
"1 Computations for Symmetric Coalgebras . . . . . . . . 435
8 Uniqueness Theorem for Quantum Enveloping Algebras 442
9 Exercises................. 446
10 N o t e s . . . . . . . . . . . . . . . . . . . 446
11 Appendix. Complexes and Resolutions. 447
XIX
Monodromy of the Knizhnik-Zamolodchikov Equations 449
1 Connections . . . . . . . . . . . . . . . . . . . . 449
2 Braid Group Representations from Monodromy . 451
3 The Knizhnik-Zamolodchikov Equations. 455
4 The Drinfeld-Kohno Theorem 458
5 Equivalence of Uh(g) and Ag,t . . . . . . 461
6 Drinfeld's Associator . . . . . . . . . . . 463
7 Construction of the Topological Braided Quasi-Bialgebra Ag,t 468
8 Verification of the Axioms 471
9 Exercises . . . . . . . . . . . 479
10 N o t e s . . . . . . . . . . . . . 479
11 Appendix. Iterated Integrals 480
XX
Postlude. A Universal Knot Invariant 484
1 Knot Invariants of Finite Type . . . . . 484
2 Chord Diagrams and Kontsevich's Theorem. 486
3 Algebra Structures on Chord Diagrams . . . 491
4 Infinitesimal Symmetric Categories. . . . . . 494
5 A Universal Category for Infinitesimal Braidings 496
6 Formal Integration of Infinitesimal Symmetric Categories 498
7 Construction of Kontsevich's Universal Invariant 499
8 Recovering Quantum Group Invariants 502
9 Exercises 505
10 Notes . . . . 505
References 506
Index 523
Part One
Quantum 8L(2)
Chapter I
Preliminaries
(l.3)
(1.4)
then there exists a unique algebra structure on the quotient vector space
AI I such that the canonical projection from A onto AI I is a morphism of
algebras.
4. We endow the product set A = DiEI Ai of a family (Ai)iEI of algebras
with the unique algebra structure such that the canonical projection from
A to Ai is an algebra morphism for all i E I. The algebra A is called the
product algebra of the family (Ai)iEI.
5. Given an algebra A we can form the algebra A[x] of all polynomials
~~=o aix i where n is any non-negative integer and the algebra A[x, X-I]
n .
of all Laurent polynomials ~i=m aix~ where m, n E Z.
6. For any positive integer n we denote by Mn(A) the algebra of all
n x n-matrices with entries in A.
7. The space End(V) of linear endomorphisms of a vector space V is an
algebra with product given by the composition and unit by the identity
map id v of V.
Given an algebra A, a left A-module or, simply, an A-module is a vector
space V together with a bilinear map (a, v) f---> av from A x V to V such
that
a(a'v) = (aa')v and Iv = v (l.5)
1.1 Algebras and Modules 5
for all a, a' E A and v E V. One similarly defines a right A-module using
a bilinear map from V x A to V. A right A-module is nothing else than a
left module over the opposite algebra AOP. Therefore we need only consider
left modules which shall for simplicity be called modules in the sequel.
If V and V' are A-modules, a linear map f : V --+ V'is said to be
A-linear or a morphism of A-modules if
(1.8)
PROOF. Clearly, (i) =} (ii) and (iii) =} (iv). We also have (i) =} (iii): it
suffices to define p as the canonical projection from Vi EB V" onto V'.
Similarly, (ii) =} (iv).
Assertion (iii) =} Assertion (i). Let V" = Ker (p); it is a submodule of V.
The relations v = p( v) + (v - p( v)) and p2 = P prove that V is the direct
sum Vi and V". Similarly, (iv) =} (ii).
Assertion (ii) =} Assertion (v). Assuming (ii), we have to prove that
any finite-dimensional A-module V is semisimple. We may also assume
that dim(V) > O. Consider a non-zero submodule VI of V of minimal
dimension; it has to be simple. By (ii) there exists a submodule VI such
that V ~ VI EB VI and dim(VI) < dim(V). Iterating this procedure, we
build a sequence (Vn)n>O of simple submodules and a sequence (vn )n>O of
submodules such that
V n ~ Vn+ I EB V n+ I and dim(Vn+1) < dim(Vn).
Since the dimension of vn is strictly decreasing, there exists an integer p
such that VP = {O}. The module V is a direct sum of simple modules:
V~VIEB"'EBVp'
It remains to be shown that Assertion (v) implies Assertion (i). Let
Vi C V be a pair of finite-dimensional A-modules. By (v)
V=EB1I;
iEI
is a direct sum over a finite index set I of simple submodules 11;. Let J be
a maximal subset of I such that
Vi n (EB Yj) = {o}. (1.9)
jEJ
If i tf- J, then
Vi n (11; EB EB Yj) =I {O},
jEJ
hence
11; n (V' + EB Yj) =I {O}.
jEJ
Since 11; is simple, this implies that
11; c Vi + EB Yj
jEJ
for all i tf- J. This holds also for all i E J. Consequently, for the sum V of
all 11; we must have
V = Vi + EB
Yj. (1.10)
jEJ
As a consequence of (1.9-1.10), we get V = Vi EB V" where V" is the
submodule EB jEJ Yj. 0
I.2 Free Algebras 7
Formula (2.1) equips k{X} with an algebra structure, called the free algebra
on the set X. The unit is the empty word: 1 = 0. In the sequel we shall
mainly consider free algebras on finite sets. If X = {Xl' ... ' Xn} we also
denote k{X} by k{xl, ... ,xn }.
Free algebras have the following universal property.
PROOF. It is enough to define Ion any word of X. For the empty word we
set 1(0) = 1. Otherwise, if x il ' ... , xi p are elements of X, we define
(2.3)
algebra k[XI' ... ,xn ] in n variables with coefficients in the ground field k.
As a corollary of (2.4) we have
HOmAlg(k[xl> ... ,xn],A) ~ {(al' ,an) E An I aiaj = aja i for all (i,j)}
(2.5)
for any algebra A.
In the next sections we shall see more examples where families of elements
subject to "universal" algebraic relations are represented by quotients of
free algebras.
(3.1)
HOmAlg(k[x], A) ~ A. (3.2)
The algebra k[x] is called the affine line and the set HOmAlg(k[x], A) is
called the set of A-points of the affine line. Now A has an abelian group
structure. We wish to express it in a universal way using the affine line
k[x]. The abelian group structure of A consists of three maps, namely the
addition + : A2 -+ A, the unit 0 : {O} -+ A, and the inverse - : A -+ A,
satisfying the well-known axioms which express the fact that the addition
is associative and commutative, that it has 0 as a left and right unit and
that
(-a) + a = a + (-a) = 0
for all a E A. These laws do not depend on the particular commutative
algebra A. It will therefore be possible to express them universally.
To this end, let us introduce the affine plane k[x', x"] with the bijection
obtained from (3.1) for n = 2. An element of Hom Alg (k [x' , x"], A) is called
an A-point of the affine plane. The set HOmAlg(k, A), reduced to the single
point rJ A' will be denoted by {O}.
Proposition 1.3.2. Let ~ : k[x] ---> k[x', x"], c : k[x] ---> k, S: k[x] ---> k[x]
be the algebra morphisms defined by
Under the identifications (3.2-3.3), the morphisms
to the maps +, and - respectively.
~, c and S correspond
k[x', x", x'-I, X"-I] = k[x', y', x", Y"l/ (x' y' - 1, x" y" - 1)
A k[x x-I]
u ., ---> k[x' " x" X"-I]
X,-I" co k[x X-I] ---> k
<C., ,
by
~(x) = x'x", c(x) = 1, S(x) = X-I. (3.6)
Then the morphisms ~, c and S correspond respectively to the multipli-
cation in A x, to the unit 1 and to the inverse under the identifications
(3.4-3.5). Here again, the morphisms ~,c, S equip k[x, X-I] with a co com-
mutative Hopf algebra structure.
10 Chapter 1. Preliminaries
(4.1)
Proposition 1.4.1. Let .6. : M(2) --+ M(2)@2 be the algebra morphism
defined by
Then for any commutative algebra A, the morphism .6. corresponds to the
matrix multiplication in M 2 (A) under the identifications (4.1-4.2).
The proof is easy and left to the reader. It is convenient to rewrite the
formulas for .6. in Proposition 4.1 in the compact matrix form
( f(a) f(b))
f(c) f(d) .
PROOF. We give it only for GL(2). Similar arguments work for SL(2). Let
(~ ~) be a matrix in GL 2 (A). Since A is commutative, there exists a
unique algebra morphism f : M(2)[tJ -+ A such that
f(a) = a, f(b) = (3, f(e) = "1, f(d) = 8 and f(t) = (a8 - (3"1)-1.
Now,
and
SL(2)2 = GL(2)2 /(t' -1, t" -1) = M(2)2/(a'd' -b' e' -1, a" d" -b" e" -1).
corresponding to the units of the groups GL 2(A) and SL 2(A) and the
algebra morphisms
c ( ac db) = (1 0)
0 1 and S (ac db) = (ad - bc) -1 ( d
-c -b)
a .
(5.2)
for all i,j EN. The elements of Ai are said to be homogeneous of degree i.
I.6 Graded and Filtered Algebras 13
I = EB InA;
i;:>O
and the quotient algebra AI I is graded with (AI I)i = Ad (I n Ai) for all i.
PROOF. It suffices to show that I = E9i>O I n Ai' First observe that the
sum has to be direct since the subspaces Ai form a direct sum. Therefore,
it remains to be checked that I = 2:i>O I n Ai' The ideal I is generated by
homogeneous elements xi of degree d~. Consequently, if x E I then
for some ai' b;EA. Now, a; = 2: j ai and bi = 2: j bi, where ai and bi are
homogeneous elements of degree j. It follows that
x = L 7
ai x i b
i,j,k
Example 2. The polynomial algebra k[XI' ... ,xnl is graded as the quotient
of the free algebra A = k{ Xl' ... ,xn } (graded as in Example 1) by the ideal
I generated by the degree-2 homogeneous elements xix j -XjXi where i and
j run over all integers between 1 and n. The generators Xl' ... 'X n are of
degree one.
The algebras M(2) and M(2)2 of Section 4 are graded as polynomial
algebras. On the contrary, the ideals defining the algebras GL(2) and 5L(2)
are not generated by homogeneous elements. Though not graded, GL(2)
and 5L(2) are filtered algebras in the sense of the following definition.
For any filtered algebra A there exists a graded algebra S = gr(A) defined
by
Fi(A) = EB Aj
0:Scj:Sci
for all i E N. We have gr(A) = A.
Example 5. Let A:J ... :J Fl(A):J Fo(A) be a filtered algebra and I be a
two-sided ideal of A. The quotient-algebra A/lis filtered with
1. 7 Ore Extensions
Let R be an algebra and R[t] be the free (left) R-module consisting of all
polynomials of the form
Theorem 1.7.1. (a) Assume that R[t] has an algebra structure such that
the natural inclusion of R into R[t] is a morphism of algebras, and we have
deg(PQ) = deg(P) + deg(Q) for any pair (P, Q) of elements of R[t]. Then
R has no zero-divisors and there exist a unique injective algebra endomor-
phism a of R and a unique a-derivation 8 of R such that
for all a E R.
(b) Conversely, let R be an algebra without zero-divisors. Given an injec-
tive algebra endomorphism a of R and an a-derivation 8 of R, there exists
a unique algebra structure on R[t] such that the inclusion of R into R[t] is
an algebra morphism and Relation (7.2) holds for all a in R.
The algebra defined by Theorem 7.1 (b), denoted R [t, a, 8], is called the
Ore extension attached to the data (R, a, 8).
PROOF. (a) Let a, b be non-zero elements of R, hence of degree 0 in R[t].
We have deg(ab) = deg(a) + deg(b) = 0, which implies that ab =f. O. Conse-
quently, R has no zero-divisors.
Let us now prove the existence and the uniqueness of the endomorphisms
a and 8. Take any non-zero element a of R and consider the product tao
We have deg(ta) = deg(t) + deg(a) = 1. By definition of R[t] there exist
uniquely determined elements a(a) =f. 0 and 8(a) of R such that
8 0 0 0
a 8 0 0
0 a 8 0
T= 0 0 a 8
0 0 0 a
PROOF. For any integer i ~ 1, let e i be the infinite column vector whose
entries are all zero, except for the i-th one which is equal to the unit 1 of
R. We may apply the matrix T of endomorphisms to ei . Since 8(1) = 0 and
a(l) = 1 we get
(7.7)
for all i ~ 1. Now, let P L:~=o ait i be an element of R[t] such that
<p(P) = O. We wish to show that all elements ao,"" an are zero. Apply
<p(P) to the vector column e1 . By (7.7) we get
n n
0= <p(P)(e 1 ) = L (iiiI)T i (e 1 ) = L iii ei+1'
i=O i=O
The set {eJ i>l being free, we have iii = 0 for all i. Since R has a unit, we
get a i = 0 for-all i. Hence, P = O. D
allows one to lift the algebra structure of S to R[t]. Relation (7.2) holds in
R[t] in view of Lemma 7.3. 0
Corollary 1.7.4. Under the hypotheses of Theorem 7.1 (b), the following
holds.
(a) For all i with 0:::; i :::; m + n we have
i P
with an i= O. Using (7.10) once again, we get another relation of the form
18 Chapter 1. Preliminaries
which, by Part (b), implies that an(a n ) = O. The map a being an isomor-
phism, we get an = 0, hence a contradiction. D
1. 8 Noetherian Rings
Proposition 1.8.1. Let A be a ring. The following two statements are
equivalent.
(i) Any left ideal I of A is finitely generated, i. e., there exist aI' ... , an
in I such that I = Aa 1 + ... + Aan .
(ii) Any ascending sequence II C 12 C 13 C ... c A of left ideals of A is
finite, i. e., there exists an integer r such that Ir+i = IT for all i ~ O.
PROOF. Let us first show that (i) implies (ii). Consider an ascending se-
quence II C 12 C 13 C ... of left ideals of A. The union of these ideals is
a left ideal I which, by (i), is generated by a finite number a 1 , . .. ,an of
elements of A. By definition of the union there exists an integer r such that
aI' ... ,an all belong to the ideal IT' It follows that I C IT C Ir+i C I for
all i ~ O.
We now establish the converse. Let I be a left ideal that is not finitely
generated and a 1 be an element of I. The left ideal II = Aa 1 is contained
in I and II =J I. Therefore, we can find an element a 2 E 1\ Aa 1 . We have
II C 12 = Aa 1 + Aa 2 C I and II =J 12 =J I. Proceeding inductively, we find
an infinite strictly ascending sequence II C ... In C 171 +1 C ... I of left
ideals. D
Proof of Theorem 8.3. Let I be a left ideal of the Ore extension R[t, a, 15].
We have to prove that I is finitely generated. Given an integer d ;::: 0, define
Id as the union of {O} and of all elements of R which appear as leading
coefficients of degree d elements of I. One checks easily that Id is a left
ideal of R.
On the other hand, if a is the leading coefficient of some polynomial P,
then a(a) is the leading coefficient of tF. Consequently, a(Id) is included
in I MI . We therefore have the ascending sequence
The coefficient of t d in Q is
a- L riad-n(ad,.J = 0,
O::;i::;p
which shows that the degree of Q is < d. We can therefore apply the
induction hypothesis and conclude as above. 0
I. 9 Exercises
1. (Schur's lemma) Prove that any A-linear map between simple A-
modules is either zero or an isomorphism. Deduce that the A-linear
endomorphisms of a simple A-module form a skew-field.
Prove that
1 1
P(k{x1,,x n }) =-- and P(k[x1,, Xn]) = (1 _ t)n
I - nt
(b) Show that any trace on R[t,o], i.e., any linear map T on R[t, oj
such that T(XY) = T(YX) for any pair (x, y) of elements of R[t, 0], is
zero.
where
1.10 Notes
Ore extensions were introduced by Ore in [Ore33]. They are also called
"skew polynomial rings" in [Coh71] [MR87] (see also [Cur52]). One of Ore's
motivations was to find a large class of non-commutative algebras that
are embeddable into a skew-field. As is well-known, this is possible for
any commutative integral domain, but not for a general non-commutative
algebra. Ore proved that any algebra obtained from a skew-field by iterated
Ore extensions can itself be embedded into some skew-field (see Proposition
0.8.4 in [Coh71]). For more details on Noetherian rings, we refer the reader
to [Lan65] and [MR87]. The examples given in [MR87], 2.11 show that the
non-commutative version of Hilbert's basis theorem is no longer true if the
endomorphism a is not assumed to be bijective.
Chapter II
Tensor Products
(1.4)
kV~V~Vk
The tensor product also commutes with the direct sum of spaces. Let
(Ui)iEI be a family of vector spaces indexed by a set I. Recall that there
exists a vector space EBiEI Ui , called the direct sum of the family (Ui ), and
linear maps qi : Ui ---> EBiEI Ui such that for any vector space V, the linear
map
(1.5)
iEI iEI
given by f ---> (f 0 qi)i is an isomorphism.
Proposition 11.1.4. We have
(1.6)
iEI iEI
PROOF. By Corollary 1.2 and (1.5) we have the chain of isomorphisms
Hom((EB Ui ) Q9 V, W) ~
Hom( EB Ui , Hom(V, W))
iEI iEI
~
IIHom(Ui,Hom(V, W))
iEI
~
II Hom(Ui Q9 V, W)
iEI
~
Hom(EB (Ui Q9 V), W).
iEI
These hold for any vector space W. A classical argument given in full detail
in the second proof of Proposition 5.1 (c) allows one to conclude. 0
Recall also the notion of a direct product of vector spaces. Let (V,)iEI
be a family of vector spaces indexed by a set I. There exists a vector space
[LEI v" called the direct product of the family (V,)iEI' and linear maps
Pi : DiEI V, ---> V, such that for all vector spaces U, the map
Corollary 11.1.5. Let {Ui}iEI be a basis of the vector space U and {Vj}jEJ
be a basis of V. Then the set {u i Q9 Vj}(i,j)EIXJ is a basis of the tensor
product U Q9 V. Consequently, we have dim(U Q9 V) = dim(U) dim(V).
26 Chapter II. Tensor Products
U 0 V S;! EB k (u i 0 V j)'
(i,j)ElxJ
o
Let us define the notion of a free module over an algebra A using the
tensor product. It is a module of the form A 0 V where V is a vector space
and A acts on A 0 V by
a(a' 0 v) = aa' 0 v
>. : Hom(U, U') 0 Hom(V, V') ........ Hom(V 0 U, U' 0 V') (2.2)
defined by
(>'(f09))(v0u) = f(u) 0g(v). (2.3)
The reasons for the switch of U and V in (2.2) will become apparent in
III.5.2 and in Chapter XIV. The main result of this section is the following.
11.2 Tensor Products of Linear Maps 27
A : Hom( ku i , U' ) (X) Hom(V, V') ---+ Hom(V (X) ku i , U' (X) V')
A' : U' (X) Hom(V, V') ---+ Hom(V, U' (X) V') (2.4)
defined by
A'(U' (X) f)(v) = u' (X) f(v)
is an isomorphism. By assumption, we also have U' = EBiEI' ku~ for some
finite basis {uiLEI'. We again use (1.6-1.7) and the fact that the direct
product over the finite set I' is the same as the direct sum. We get
and
Hom(V, U' (X) V') ~ II Hom(V, ku~ (X) V').
iEI'
This allows us to break A' into the direct product of the maps
A' : ku~ (X) Hom(V, V') ---+ Hom(V, ku~ (X) V').
In this special case, A' is given by A' (u~ (X) f) (v) = u~ (X) f (v), which is clearly
an isomorphism. Hence, so is the map A' of (2.4), which concludes the proof.
There are similar arguments in the remaining two cases. D
U I Vi U* 16! V*
lid@id@A
AV,W@AU,V
o
1 AU,W
(3.1)
(3.4)
(3.5)
PROOF. Immediate. o
30 Chapter II. Tensor Products
(3.8)
f+ : X* W -7 V* Y and fX: V Y* -7 X W*
by
f+(x i w j ) =L fk; v k Ye (3.9)
k,
and
r(v i y j ) = L fi~j xk w e (3.10)
k,e
if
f(v i Wj) = L fi~e xk Ye (3.11)
k,e
Lemma 11.3.3. The definitions of f+ and fX are independent of the choice
of bases. We also have
tr(f) =L ff (3.14)
(3.17)
( 4.2)
for all a E A and b E B. The tensor product of algebras enjoys the following
universal property.
This proves the uniqueness assertion. As for the existence of the map f @ g,
one checks that the previous formula defines an algebra morphism. This
1I.4 Tensor Products of Algebras 33
o
We apply Proposition 4.1 to a situation encountered in Chapter I.
where X' U X" denotes the disjoint union of the two copies and where
X' X" - X" X' is the two-sided ideal generated by all elements of the form
x' x" - x" x' with x' E X' and x" E X".
~(a)=aa+bc, ~(b) = a b + b d,
~(c)=ca+dc, ~(d) = c b + d d.
34 Chapter II. Tensor Products
(5.2)
(5.3)
in view of (5.3). This proves the uniqueness of f. As for its existence, one
checks immediately that the previous formula defines an algebra morphism
from T(V) into A.
(c) By Corollary 1.5, if {eiLEI is a basis of V, then {ei1 ein }i 1 ,,,. ,inEI is
a basis of the vector space Tn (V). When n runs over the set of non-negative
integers we get a basis of T(V) which is clearly in bijection with Ii basis of
k{I}. This bijection induces an isomorphism between both vector spaces.
The product on T(V) corresponds to the concatenation in k{ I} under this
isomorphism.
Let us give another, less pedestrian, proof of Part (c). By (1.5), (1.8),
(5.4) and (1.2.2) we have the following chain of natural bijections:
HOmAlg(T(V) , A) ~
Hom(V,A)
~
Hom(EB ke i , A)
iEI
~
II Hom(kei,A)
iEI
~
Homset(I, A)
~
HOmAlg(k{I}, A).
Let a be the composition of these bijections. First, take A = T(V) and
define <p = a(idT(v)); this is an algebra morphism from k{I} to T(V). Now
take A = k{I} and define 'ljJ = a-l(idk{I}); this is an algebra morphism
from T(V) to k{I}. We claim that <p and 'ljJ are isomorphisms between T(V)
and k{I}. First, observe that the bijection a is natural, which means that
for any algebra morphism f : A --- A' we have
foa(w)=a(fow)
for any w E HOmAlg(T(V) , A). Let us now compose <p and 'ljJ. On the one
hand, we get
f(x)f(y) = f(y)f(x)
for any pair (x, y) of elements of V, there exists a unique algebra morphism
f: S(V) ---> A such that f 0 iv = f
(c) If I is an indexing set for a basis of V, then the symmetric algebra
S(V) is isomorphic to the polynomial algebra k[I] on the set I.
(d) If V' is another vector space, we have an algebra isomorphism
HOmAlg(S(VEElV'),A) ~ Hom(VEElV',A)
~ Hom(V,A) x Hom (V' , A)
~ HOmAlg(S(V) , A) x HOmAlg(S(V'), A)
~ HOmAlg(S(V) S(V'), A).
We then successively take A to be S(V EEl V') and S(V) S(V'), which
produces isomorphisms between these algebras, as in the second proof of
Part (c) of Proposition 5.1. 0
II.6 Exercises
1. If f and l' [resp. g and g'] are composable linear maps, show that
(A0A')n= E9 A i 0Aj.
i+j=n
6. (Exterior algebra) For any vector space V we define the exterior alge-
bra (or Grassmann algebra) A(V) as the quotient A(V) = T(V) / I' (V)
of T(V) by the two-sided ideal I' (V) generated by the elements x 0 x
where x runs over V. If Xl"" ,X n are elements of V, denote by
Xl 1\ .. . 1\ xn the class of Xl 0 ... 0 xn in A(V). The subspace of A(V)
generated by the elements Xl 1\ ... 1\ xn is denoted N'(V). Let iv
be the canonical map from V = TI (V) to A(V). Prove the following
statements.
(a) The algebra A(V) is graded such that An(v) is the subspace of
degree n homogeneous elements.
(b) For any algebra A and any linear map f : V -+ A satisfying
f(X)2 = 0 for all X E V, there exists a unique algebra morphism
! : A(V) -+ A such t.hat ! 0 iv = f.
(c) Let I be an ordered set indexing a basis {eJiEI of V. Then the
set {e i1 1\ ... 1\ einL,<...<inEl is a basis of N'(V).
(d) Assume V of finite dimension d. Prove that
II.7 Notes
For more details on the tensor, symmetric and exterior algebras as well as
on the subspaces S~(V) and A~(V) of Exercise 7, see [Bou70], Chap. 3.
Chapter III
The Language of Hopf Algebras
(1.3)
A
commutes, where T A A is the flip switching the factors: T A A (a0a') = a' 0a.
A morphism of algebras f : (A, /1, 1]) ---> (A', /1', 1]') is a linear map f from
A to A' such that
/1' 0 (J 0 J) = f 0 /1 and f 0 1] = 1]'. (1.4)
We now get the definition of a coalgebra by systematically reversing all
arrows in the previous diagrams.
Definition 111.1.1. (a) A coalgebra is a triple (C, 6., E) where C is a vector
space and 6. : C ---> C 0 C and E :. C ---> k are linear maps satisfying the
following axioms (Coass) and (Coun).
(Coass): The square
(1.5)
commutes.
(Coun): The diagram
k0C ~ C0C ~ C0k
""~ r~ /'~ (1.6)
C
commutes. The map 6. is called the coproduct or the comultiplication while
E is called the counit of the coalgebra. The squares (1.5-1.6) express that
the coproduct 6. is coassociative and counital.
If, furthermore, the triangle (Cocomm)
C
(1.7)
Te,c
~
commutes, where Te,e is the flip, we say that the coalgebra C is cocommu-
tative.
(b) Consider two coalgebras (C, 6., E) and (C', 6.', E'). A linear map f
from C to C' is a morphism of coalgebras or a coalgebra morphism if
(J 0 J) 0 6. = 6.' 0 f and E = E' 0 f. (1.8)
IILl Coalgebras 41
Then (C,,0. op, c) is a coalgebra which we call the opposite coalgebra and
denote by ccoP.
The next result relates algebras and coalgebras.
PROOF. Let (C,,0., c) be a coalgebra. Recall the map A: C*C* --+ (CC)*
of Corollary II.2.2. Set); = A 0 T c* C*' Define A = C*, JL = ,0.* 0 ); and
7] = 10* where the superscript * on 'a linear map indicates its transpose.
Then (A, JL, 7]) is an algebra (use the commutative diagrams (1.1-1.2) and
(1.5-1.6)). D
PROOF. Let (A, JL, 7]) be a finite-dimensional algebra. Then the map); from
A * A * to (A A) * is an isomorphism, which allows us to define ,0. by
42 Chapter III. The Language of Hopf Algebras
~= >:-1 op,*. We also set c = T]*. Using the commutative diagrams (1.1-
1.2) and (1.5-1.6), one checks that (A*,~,c) is a coalgebra. 0
Example 4. (The matrix coalgebra) Let A = Mn(k) be the algebra of n x n-
matrices with entries in k. Denote by Eij the matrix with all entries equal
to 0, except for the (i, j)-entry which is equal to 1. The set of matrices
Eij (1 S; i,j S; n) is a basis of Mn(k). Let {Xij} be the dual basis. Then
A * is the co algebra defined by
n
~(Xij) = :L>ik Xkj and C(Xij) = 8ij (1.11 )
k=l
Indeed, we have
and
xij(p,(Ekt Emn))
DtmXij(Ekn)
8tmDikDjn
L DikDtp8pmDjn
p
It is clear that
000/(I00+00I) =0/100/1.
Similarly, the counit factors through a map ~ : 0/1 - t k. Then clearly, the
triple (0/1, A,~) is a coalgebra. It is called the quotient-coalgebra. We shall
give examples later.
Notation 1.6. We now present Sweedler's sigma notation which we shall
use continually in the sequel. If x is an element of a coalgebra (0, A, e), the
element A (x) of 0 0 0 is of the form
(1.13)
In order to get rid of the subscripts, we henceforth agree to write the sum
(1.13) in the form
A(x) = LX'
0x". (1.14)
(x)
or L:(x) x(1) X(2) x(3) x(4). More generally, let D. (71) : C ----) c(n+1)
"(71)
u = ("d
u 1 e 0 (n-l) ) 0 ,,(71-1)
u = (d
1 e;s?(n-l) U") 0 U,,(71-1) . (1.18)
Using the conventions (1.14), the condition (1.6) for counitality may be
reformulated for any x E C as
L x(l) E(X(2) x(3) x(4) x(5) = L x(1) X(2) x(3) X(4). (1.22)
(x) (x)
for all x E c.
111.2 Bialgebras 45
III. 2 Bialgebras
Let H be a vector space equipped simultaneously with an algebra structure
(H,fL,ry) and a coalgebra structure (H,~,c). Let us discuss two compati-
bility conditions between these two structures. We give H H the induced
structures of a tensor product of algebras (see II.4) and of a tensor product
of co algebras (see Section 1, Example 5).
Theorem 111.2.1. The following two statements are equivalent.
(i) The maps fL and ry are morphisms of coalgebras.
(ii) The maps ~ and care morphisms of algebras.
PROOF. It consists essentially in writing down the commutative diagrams
expressing both statements. The fact that fL is a morphism of coalgebras is
equivalent to the commutativity of the two squares
HH ....!!:... H HH
1(id<SlT<SIid) (ll.<SIll.) lll. l~
(HH)(HH) ~ HH H
whereas the fact that ry is a morphism of coalgebras is expressed by the
commutativity of the two diagrams
k ...!!..... H
~ id ,/ c o
k
Observe that these four commutative diagrams are exactly the same as the
following four diagrams whose commutativity express the fact that ~ and
care morphisms of algebras:
(H H) (H H)
1(~<SI~)(id<SlT<SIid)
HH
46 Chapter III. The Language of Hopf Algebras
and e<8le 7)
H0H ------+ k0k k --t H
l~ lid \, id ./e
e
H --t k k
Definition III.2.2. A bialgebra is a quintuple (H, {t, 'fl, A, c) where (H, {t, 'fl)
is an algebra and (H, A, c) is a coalgebra verifying the equivalent conditions
of Theorem 2.1. A morphism of bialgebras is a morphism for the underlying
algebra and coalgebra structures.
In the sequel, we shall mainly use Condition (ii) of Theorem 2.1 to define
a bialgebra structure. Using the conventions of 1.6, we see that the condition
A(xy) = A(x)A(y) is expressed for any pair (x, y) of elements in a bialgebra
by
L(xy)' 0 (xy)" = x'y' 0 x"y".L (2.1)
(xy) (x)(y)
We also have
HCOP = (H , f"'"I/. 71
'n
AOP
"
c)
and c(xy) = 1 = c(x)c(y), which implies that the maps A and care
morphisms of algebras. Thus k[X] becomes a bialgebra.
If, in addition, X is a finite set, then the dual of k[X] also is a bialgebra.
We have already observed that the algebra structure of the dual is the
usual algebra structure of the space of k-valued functions on X. An easy
III.2 Bialgebras 47
computation shows that the comultiplication and the co unit on the algebra
of functions are given by
Example 3. (The bialgebra M(n)) Let M(n) = k[xll, ... ,xnnl be the
polynomial algebra in n 2 variables {xijh:::;i,j:::;n' For all i,j, set
n
(2.5)
n-I
= 1 VI ... Vn + LL V(]"(l) ... V(]"(p) V(]"(p+l) ... V(]"(n) +V I ... Vn 1 (2.6)
p=1 (]"
where a runs over all permutations oj the symmetric gmup Sn such that
a(l) < a(2) ... < a(p) and a(p + 1) < a(p + 2) ... < a(n).
Such a permutation a is called a (p, n - p)-shuffie.
PROOF. By universality of the tensor algebra, there exist unique algebra
morphisms 6. : T(V) -+ T(V) T(V) and c : T(V) -+ k such that their
restrictions to V are given by the formulas of the theorem. Now consider
several elements vI' ... , vn in V. Formula (2.5) is a trivial consequence of
the multiplicativity of c.
Let us now compute 6.( vI ... v n ). We shall do this by induction on n.
Formula (2.6) holds for n = 1 by definition. Suppose it holds up to n-1 ~ 1.
Then we have the series of equalities
6.( VI ... vn )
= 6.(vI v n _ I )6.(v n )
48 Chapter III. The Language of Hopf Algebras
n-2
(10 VI' .. Vn - I +L L Vo-(I) ... v(J(p) 0 V(J(p+l) ... v(J(n-l)
p=1 (J
where (J runs over all (p, n -1- p)-shuffies of Sn-l' Let us rewrite the last
sum in the form
n-2
10 VI' .. Vn +L L Vp(l) ... Vp(p) 0 Vp(p+l) ... Vp(n_l)Vn
p=1 p
+ VI' .. Vn - 1 0 Vn + Vn 0 VI' .. Vn - 1
n-l
+ LL VT(l) ... VT(p_I)V n 0 VT(p) ... VT(n-l) + VI'" Vn 0 1
p=2 T
where p runs over all (p, n - 1 - p)-shuffies of Sn-I and T runs over all
(p - 1, n - p )-shuffies permuting the set {I, ... , n} \ {p}. Now observe that
if (J E Sn is a (p, n - p )-shuffie, then either d n) = n, hence the restriction
p of (J to Sn-l is a (p, n -1- p)-shuffie, or (J(p) = n, hence T = (J acting on
{l, ... , n} \ {p} is a (p -1, n - p)-shuffie. This completes the proof of (2.6).
It remains to prove the coassociativity, the counitality and the cocom-
mutativity of ~. The counitality results from an easy computation using
(2.5) and (2.6). The co commutativity is a consequence of the fact that the
permutation
(
p+1
1
p+2
2 ...
...
p
n
p+1
1
p+2
2
...
...
n)
P
6.(xy) = (1 x + x 1)(1 y + y 1) = 1 xy + x y + y x + xy l.
We deduce
6.([x, y]) = 1 [x, y] + [x, y] 1,
which implies that [x, y] is primitive. D
As a consequence of Proposition 2.7, we see that for any set {Xl' ... ,x n }
of primitive elements in a bialgebra, 6.(XI' ... ,x n ) is given by Formula
(2.6) of Theorem 2.4 after replacing Vi by Xi.
This proves that the convolution is associative. The map 7] 0 e: is a left unit
for the convolution in view of
which results from (1.21). One proves similarly that 7] 0 e: is a right unit.
(b) Let a, bE A and a, (3 E C*. Then for x E C we have
This proves that AC,A preserves the product. As for the unit, we have
S * id H = id H *S = 7] 0 e:.
111.3 Hopf Algebras 51
A Hopf algebra with an antipode 8 will be denoted by (H, j.l, ry,~, c:, 8).
Using Sweedler's convention 1.6, we see that an antipode satisfies the
relations
L:: x'8(x") = c:(x)l = L:: 8(x')x" (3.3)
(x) (x)
for all x E H. In any Hopf algebra we have relations such as
L:: X(l) 0 X(2) 0 8(x(3) 0 x(4) 0 X(5) L:: X(l) 0 c:(x(2) 0 x(3) 0 X(4)
(x) (x)
L:: x(1) 0 X(2) 0 x(3) .
(x)
The first equality follows from (3.3), i.e., by definition of the antipode
while the second one follows from (1.21), i.e., from the Axiom (Coun).
Such computations will be performed later without further explanations.
We state the counterpart of Example 1 of Section 2.
L:: 0'(x')0"(8x")
(a) (x)
0(L:: x'(8x"))
(x)
o(ryc:(x))
c:*ry*(o)(x).
(3.5)
PROOF. (a) Let us start with (3.4). Define maps v, pin Hom(H H, H) by
(p*p,)(xy) = L p((xy)')p,((xy)//)
(xy)
L S(x'y')x//y//
(x)(y)
L S((xy)')(xy)//
(xy)
7]E(XY)
III.3 Hopf Algebras 53
2:= x'y'S(yl)S(X")
(x)(y)
2:= x'c(y)S(x")
(x)
ryc(x)ryc(y)
ryc(xy) ,
which is the same.
Applying (id * S)(x) = ryc(x) to x = 1, one gets S(l) = 1. This proves
(3.4).
Let us deal with (3.5). It is equivalent to prove t:J..oS = (SS)ot:J.. P. We
set p = t:J..oS and v = (SS)ot:J.. P. These are linear maps from H to HH.
We wish to show that p = v. This will follow from p*t:J.. = t:J..*v = (ryry)c,
which we prove now. On the one hand, by (1.21)
The fourth and seventh equalities follow from (3.3), the sixth one from
(1.21).
We also derive
S(L: S(X')S2(x ll ))
(x)
S(L: S(x')x ll )
(x)
S(E(x)l)
E(X) 1.
One proves that (i) is equivalent to (iii) in a similar fashion.
(c) Recall Relations (3.3): we have
L: S(x")x' = 1JE(x),
(x)
1JE(x) = L: S(x")x'
(x)
Corollary 111.3.5. Let H = (H, jL, 'fI,~, c:, 5) be a Hopf algebra. Then
HOPCOP = (H , ,...., ,." ~op " c: 5)
/lOP 'Yl
PROOF. It is enough to check that if (3.3) holds for x and y, then it holds
for the product xy. Now, by (3.3-3.4)
L x'(Ly'5(yll))5(x")
(x) (y)
(L x'5(x"))c:(y)
(x)
c:(x)c:(y)
c:(xy).
Use the previous lemma to show that the following provide examples of
Hopf algebras.
56 Chapter III. The Language of Bopf Algebras
6.(x[V, w]y) = L (x'[v, w]y' 121 xl/yl/ + x'y' 121 xl/[v, w]yl/)
(x)(y)
Similarly, the counit axiom follows from c:(t) = 1 and from the matrix
equalities
(~ !) (~ ~) = (~ !) = (~ ~) (~ !). (4.2)
( ae db)( S(a)
S(e)
S(b)
S(d)
)=( S(a)
S(e)
S(b))( a b
S(d) c d
)=( c:(a)
c:(e)
C:(b))
c:(d) .
(4.3)
As for t, we have tS(t) = S(t)t = c:(t) = 1 since S(t) = C 1 = ad - be.
The antipode is an involution due to the fact that GL(2) and SL(2) are
both commutative. This can also be checked directly on Formula (1.5.2)
defining S.
Proposition 111.5.2. Let (A, /-L, T/, A, c, S) be a Hopf algebra and U, U', V
and V' be A-modules such that, either U or U', and, either V or V', are
finite-dimensional vector spaces. Then the linear map
of (II.2.2) is A-linear if, in addition, the flip TU* v' : U* V' ---. V' U*
is A-linear. In particular, the maps '
are A-linear.
Zl (A(a(fg)))(vu)
L A(a' f a" g) (v u)
(a)
L (a' f) (u) (a" 9) (v )
(a)
= L (a')' f(S((a')")u) (a")'g(S((a")")v)
(a)
Z2 (aA(fg))(vu)
We used (3.6) for the fourth equality. Observe that Zl -I=- Z2 in general.
(b) Let V' = k be given the trivial action. Replacing a'" in Zl [resp. a"
in Z2] by c:(a"') [resp. by c:(a")] and using (1.21), we get
= O:(c(a)v)
= c(a)o:(v)
by the rightmost relation (3.3) and by (5.6). This implies that the evalua-
tion map is A-linear.
(b) The coevaluation map 0v is A-linear as the composition of the unit
TJ : k -; End(V) and of Av\;. The latter is A-linear by Proposition 5.2. So
is the map TJ : k -; End(V) following
(aTJ(l))(v) (aidv)(v)
2: a'idv(S(a")v)
(a)
2: a'S (a")v
(a)
c(a)v
(TJ(a1))(v)
for all v E V and a E A. Here we used the leftmost relation (3.3).
(c) For the composition map, one uses Lemma 11.2.5. 0
III. 6 Comodules
Algebras act on modules, coalgebras coact on comodules. This section is
devoted to the definition of the latter concept. Let A be an algebra. Recall
that an A-module is a pair (M, JLM) where M is a vector space and JLM :
A0M -; M is a linear map such that the following axioms (Ass) and (Un)
hold.
(Ass): The square
(6.1)
commutes.
(Un): The diagram
'T/0id
--+
(6.2)
commutes.
A morphism of A-modules f : (M, JLM) -; (M', JLM') is a linear map f
from M to M' such that
(6.3)
62 Chapter III. The Language of Hopf Algebras
(6.4)
L>.0id
~
commutes.
(Co un): The diagram
c0id
kN f---
(6.5)
N
commutes.
(b) Let (N, Ll N) and (N', Ll NI) be C -comodules. A linear map f from N
to N' is a morphism of C-comodules if
Actually, the comodules we have just defined are left comodules. One
similarly defines a right C-comodule N, using a map N C ---7 N subject
to relations parallel to (6.4-6.5). A right C-comodule is the same as a (left)
comodule over the opposite coalgebra ccoP.
The composition of two morphisms of comodules is another morphism
of comodules. Similarly, the inclusion of a subcomodule into a co module is
a morphism of comodules. Let us give a few examples of comodules.
Example 1. Let C be a coalgebra. Then (C, Ll) is a C-comodule.
Example 2. Let C be a coalgebra and C* the dual vector space equipped
with the dual algebra structure of Proposition 1.2. If (N, LlN) is a C-
comodule, then the dual vector space N* has the structure of a right C*-
module given by the composition of the maps
(M Q9 N) Q9 P ~ M Q9 (N Q9 P) and k Q9 M ~ M ~ M Q9 k
Notation 6.3. It is often convenient to use for comodules the same kind
of notation as was introduced for coalgebras in Section 1. Let (C, fl., s) be
a co algebra and (N, fl. N ) be a C-comodule. By convention we shall write
fl.N(X) =L Xc Q9 xN (6.11)
(xl
64 Chapter III. The Language of Hopf Algebras
Definition III. 7.1. Let (H, J-LH' TiH' 6. H , EH) be a bialgebra and (A, J-L A, TiA)
be an algebra. We say A is an H -comodule-algebra if
(a) the vector space A has an H-comodule structure given by a map
6. A : A --t H Q9 A, and
(b) the structure maps J-L A : A Q9 A --t A and TiA : k --t A are morphisms
of H -comodules, the tensor product AQ9A and the ground field k being given
the H -comodule structures described in Section 6.
PROOF. It is similar to the proof of Theorem 2.1. We first express the fact
that J-L A is a morphism of H-comodules with the commutative square
(7.1)
idMA
)
III.7 Comoduie-Algebras. Coaction of SL(2) on the Affine Plane 65
where u = (/-LH @ id@ id) 0 (id @ TA.H @ id) 0 (D. A @ D. A ). The fact that 7)A
is a morphism of H-comodules is equivalent to the commutativity of the
square
k
I~ (7.2)
k@k
Now, Diagrams (7.1-7.2) are exactly the same as Diagrams (7.3) below
which express the fact that D.A is a morphism of algebras:
A@A (H @ A) @ (H @ A) k@k
I~A Iv IrlHT)A (7.3)
A H@A H@A
for all a, b E A.
We now show that the affine plane k[x, y] defined in 1.3 possesses an
comodule-algebra structure over the bialgebras M(2) and 8L(2).
D.A ( ~ ) = (~ ~) @ ( ~ ) .
(~ !)Q9(~ !)Q9(~)
( (~ Q9 id) ~ A) ( ~ )
0
III.8 Exercises
1. (Tensor product of coalgebras) Let (0, ~, c) and (0', ~', c:') be coalge-
bras. Show that the linear maps 7T : 0 Q9 0' ---> 0 and 7T' : 0 Q9 0' ---> 0'
defined by 7T(CQ9C') = c:'(c')c and 7T'(CQ9C') = c(c)c' are morphisms
III.8 Exercises 67
Find an antipode.
3. (Tensor coalgebra) Let V be a vector space.
(a) Show that the canonical isomorphisms v(n+m) ~ Vn Vm
endow T' (V) = EBn>O vn with a coalgebra structure, called
the tensor coalgebra of V.
(b) Let Pv be the canonical projection of T' (V) onto V. Prove that
for any co algebra 0 and any linear map f : 0 --+ V, there exists
a unique morphism of coalgebras f : 0 --+ T' (V) such that
f=Pvof.
(c) Using the notation of Chapter II, Exercise 7, define the subspace
8'(V) = EBn>O 8~(V) [resp. A'(V) = EBn>O A~(V)] of T'(V)
generated by all symmetric [resp. antisymffietric] tensors. Show
that 8' (V) and A' (V) are subcoalgebras of T' (V).
(d) Let 0 be a cocommutative coalgebra and f be a linear map from
o to V. Prove the existence and the uniqueness of a coalgebra
morphism f : 0 --+ 8' (V) such that f = Pv 0 f.
4. (Graded dual) The graded dual vector space of a graded vector space
V = EBn>O Vn is the graded vector space Vg*r = EBn>O V;. Let
W = EBn;'O Wn be another graded vector space. Show -that there
is a grading on the tensor product V W such that
(V W)n = ED Vi Wj .
i+j=n
(ff')(x) = L f(y)f'(y-Ix)
yEG
where x E G and f, f' E O(G). Show that O(G) has a Hopf algebra
structure such that the linear map f 1-+ 2:xEG f(x)x is a Hopf algebra
isomorphism from O(G) to the group Hopf algebra C[G]. Determine
the unit, the comultiplication, the counit and the antipode of O(G).
~(t) = t t, ~(x) = 1 x + x t,
c(t) = 1, c:(x) = 0, S(t) = t, S(x) = tx
endow H with a Hopf algebra structure whose antipode is of order 4.
III.8 Exercises 69
(a) Set 1jJ0 = TiE, and 1jJrL = id;;' (convolution of n morphisms all
equal to the identity of H) if n > 0 and 1jJrL = 3*rL if n < O.
Prove that each map 1jJ" is an endomorphism of algebras [resp.
of coalgebras] when H is commutative [resp. co commutative]
and that, in both cases, we have 1jJrL o1jJm = 1jJrL+m for any pair
(n, m) of integers.
(b) Let H = k [G] be a group. Show that 1jJrL is the coalgebra endo-
morphism given by 1jJrL(g) = grL (g E G).
(c) Let H = 3(V) be a symmetric algebra. Then 1jJrL(x) = ndx for
any x E 3 d (V).
(d) Show that if H = 3L(2), then the algebra endomorphism 1jJrL is
determined by the matrix identity
if n > 0
and by
1jJrL(a) 1jJrL(b) ) _ ( d
(
1jJrL(c) - -c if n < O.
1jJrL(c)
III. 9 Notes
The concept of a Hopf algebra was developed by algebraic topologists ab-
stracting the work of Hopf [Hop41] on manifolds admitting a product (such
as Lie groups). A basic reference is the famous article [MM65] by Milnor
and Moore. Hopf algebras also came up in the representation theory of Lie
groups and algebraic groups (see [Abe80] [DG70] [Hoc81] [Ser93]). For ab-
stract Hopf algebras, we refer to Abe's and Sweedler's monographs [Abe80]
[Swe69].
All examples of bialgebras given in this chapter turn out to be either
commutative or cocommutative, except for the Hopf algebra of Exercise 7
which is due to Sweedler. Not many examples of non-commutative, non-
cocommutative bialgebras were known before the "quantum group" era
(nevertheless, see [Par81]' [Rad76], [Swe69], pages 89-90, [Taf71]' [TW80J).
This has dramatically changed in the 1980's with the appearance of quan-
tum groups. For details on the order of the square of the antipode of a Hopf
algebra, see [Rad76][Taf71][TW80].
III.9 Notes 71
AO = {a E A* IJ-l*(a) E A* 0A*}.
N = {m E MIfl.M(m) = 10m}.
It turns out that N is a subcomodule, but not a submodule of M. Put the
induced comodule structure on the free H-module H 0 M. Then H 0 M
becomes a bimodule. The structure theorem for bimodules can be stated
as follows: if H is a Hopf algebra, then the map x 0 m 1---+ xm from H 0 N
to M is an isomorphism of H-bimodules. For details, see [Abe80] [Swe69].
Chapter IV
The Quantum Plane and Its
Symmetries
This new relation defines the quantum plane. In Section 2 we derive a few
identities well-known to combinatorialists and to the experts in the theory
of linear q-difference equations. Next, investigating the self-transformations
of the quantum plane, we build a bialgebra Mq(2) and Hopf algebras GLq(2)
and SLq(2), which are one-parameter deformations of the bialgebras 1v1(2) ,
G L(2), and S L(2) defined in Chapter 1. The bialgebras obtained in this
way are our first examples of quantum groups. They have the peculiarity
of being neither commutative nor cocommutative.
(1.1 )
IV.1 The Quantum Plane 73
(1.2)
n-l qn - 1
(
n ) =l+q+"'+q =--. (2.1)
q q- 1
(2.3)
(k)!q (n - k)!q'
( ~ )q = ( n ~ k )q' (2.4)
PROOF. Relations (2.4-2.5) follow from easy computations. For Part (a),
one proceeds by induction on n using (2.5). D
(x +y)n = L
O:Sk:Sn
PROOF. Because of the universal property of the quantum plane, it suffices
to prove the statement in kq[x, y]. Expanding (x+y)n and using (1.2), we see
that the monomials in the expansion are all scalar multiples of monomials
of the form xkyn-k. We therefore have
(x +y)n = L
O:Sk:Sn
(2.6)
Relation (2.6) clearly holds for n = 1. It thus suffices to check that the
L (~)' xkyn-k
O:Sk:Sn
L ( n-l)'
k x
k+l n-l-k
y
O:Sk:Sn-l
=L (( n-l)'k-l +q
k (n-l )')
k
k n-k
xy.
(2.7)
D
The q-exponential is an invertible formal series, but, in contrast to the
case q = 1, we have eq(z)-l i= eq ( -z). In order to compute the inverse
of eq(z), we consider the algebra of formal series k[[zll and the algebra
End(k[[z]]) of linear endomorphisms of k[[zll. Define two elements Z and
Tq of End(k[[z]]) by (ZJ)(z) = zJ(z) and (TqJ)(Z) = J(qz). An easy com-
putation shows that (Z, Tq) is an End(k[[z]])-point of the quantum plane,
which is to say we have the following lemma.
PROOF. Lemma 2.5 implies (-ZTq)Z = qZ( -ZTq). Using Proposition 2.4,
we get the following identity in End(k[[z]]):
eq(Z(l- Tq)) = eq(Z) 0 eq(-ZTq). (2.10)
Let us apply both sides of (2.10) to the constant formal series 1. On the
one hand, we have eq(Z(l - Tq))(l) = 1 because (1 - Tq)(l) = O. On the
other hand, by (2.9) we get
and
PROOF. One proceeds as in the proof of Proposition 2.6, but now with the
operator identity (aTq)( -ZTq) = q (-ZTq)(aTq). By Propositions 2.2 and
2.4, we get
and
eq((a - Z)Tq) = eq( -ZTq) 0 eq(a)
in End(k[[z]]). Applying again these identities to the constant formal series
1, we get the desired relations in view of eq( -ZTq)(l) = eq(z)-l, which was
~~d~~. 0
Using x" and y" in a similar fashion leads to three more relations obtained
from (3.5-3.6) by exchanging band c, namely
Definition IV.3.2. The algebra Mq(2) is the quotient of the free algebra
k{ a, b, c, d} by the two-sided ideal Jq generated by the six relations (3.2-3.4)
of Theorem 3.l.
(3.11)
and
A C
B D )( ~ )
and where R' is the tensor product algebra
A'
m=(~ ~) and m' = ( C' B' )
D'
be two R-points of Mq(2) such that the elements A, B, C, D commute with
the elements A', B', C', D'.
80 Chapter IV. The Quantum Plane and Its Symmetries
,
m m =
(All
Gil
BII)
D" =
(A'
G'
B') ( A
D' G ~)
is an R-point of Mq (2).
(b) We have Detq(m'm) = Detq(m') Detq(m) in R.
(c) The quadruple
(_q~lG -!B)
PROOF. (a) We use the reformulation of Theorem 3.1 stated a few lines
ahead of Proposition 3.3. Let R' be the tensor product algebra
and ( XII)
y" -
(A'B' D'G') ( Y
X )
By definition, the elements X, Y of R' commute with the other variables A,
A', etc. It results from Theorem 3.1 that the pairs (X', Y') and (X", ylI)
are R'-points of the quantum plane. Now, by hypothesis, the elements A',
B', G', D' of R' commute with X' and y' and the elements A, B, G, D
commute with X" and yll. By a second application of Theorem 3.1,
A'
( G' B') ( X' ) = (All BII) ( X )
D' y' Gil D" Y
and
A" Gil ) ( X )
B" D" Y
are R'-points of the quantum plane. It follows that m'm is an R-point of
Mq(2).
(b) This follows from computations we leave to the reader. A more con-
ceptual method is suggested as an exercise at the end of this chapter.
(c) Define A' = D, B' = -qB, G' = _q-1G, and D' = A. Then Relations
(3.8-3.10) imply
Theorem IV.4.1. The algebra Mq(2) is Noetherian and has no zero divi-
sors. A basis for the underlying vector space is given by the set of monomials
i . k C
{a l)1 cd} i,j,k,C?O .
We shall prove this theorem by building a tower
The last step consists in building A4 out of A 3 . This is the only step
involving a non-zero derivation. First, one checks that
(4.1)
uniquely determined by
and
SLq(2) = Mq(2)/(det q - 1) = GLq(2)/(t - 1).
Given an algebra R, we define an R-point of GLq(2) [resp. of SL q(2)J as
an R-point m = (A, B, C, D) of Mq(2) whose quantum determinant
PROOF. (a) We first have to show that ~ and e are well-defined on GLq(2)
and on SLq(2). For SLq(2) this results from the following computations:
by (5.4)
~(detq - 1) = (detq - 1) detq + 1 (detq - 1)
and e(detq -1) = o. A similar argument works for GL q(2) provided we set
The coassociativity and counit axioms hold for GLq(2) and for SLq(2) since
they already hold for Mq(2).
(b) It remains to check that GLq(2) and SLq(2) have antipodes. Set
By Proposition 3.4 (c), the quadruple (S' (a), S' (b), s' (c), S' (d)) is a Mq(2)OP-
point of Mq(2). Consequently, S' defines a morphism of algebras from
Mq(2) to Mq(2)OP. Next, we extend S' to GLq(2) and to SLq(2) by setting
S'(t) = t. This is a well-defined algebra morphism because
S' (t)S' (detq) = (S' (d)S' (a) - q-1 S' (c)S' (b)) S' (t) = (ad - q- 1bc) t = 1.
~)
o
In contrast to the inversion in a group and to the antipode of GL(2)
and of SL(2), the antipode S of GLq(2) and of SLq(2) is in general not
involutive. Indeed, from (6.1) we derive
s2n(a)
(
s2n(c)
for any positive integer n. Fix such an n and let q be a root of unity of
order exactly n. Then we obtain two examples of Hopf algebras for which
the square of the antipode has order n. For results on the order of S2
previous to the quantum group era, see [Rad76] [Taf71] [TW80].
(7.1)
PROOF. We use Proposition III. 7.2. We first check that (7.1) defines an
algebra morphism 6 A from A to Mq(2) 0 A. It is enough to verify that
86 Chapter IV. The Quantum Plane and Its Symmetries
(e x + d y)(a x + b y)
qae x 2 + (be + qda) xy + qbd y2
q (ae x 2 + (q-1be + ad) xy + bd y2)
q (a x + b y)(e x + d y)
q~A(X)~A(Y)'
Definition IV.8.l. Let (H, IL, 71, il, c, S) be a complex Hopf algebra. We
say that H is a Hopf *-algebra if there exists an antilinear involution * on
H sat'isfying the two conditions
(i) the map * is an antimorphism of real algebras, i.e., an algebra mor-
phism from H into HOP, as well as a morphism of real coalgebras, and
(ii) we have S(S(x)*)* = x for all x E H.
Two Hopf *-algebra structures *1 and *2 on H are equivalent if there
exists a Hopf algebra automorphism r.p of H such that r.p(X*l ) = r.p(x )*2 for
all x in H.
We wish to show that the Hopf algebras GLq(2) and SLq(2) have natural
Hopf *-algebra structures given by matrix transposition. We shall need the
following equivalent formulation.
The second equality follows from * being an involution while the third
one follows from Definition 8.1 (ii). Conversely, define * = SI from an
automorphism I as in Lemma 8.2. It is an involution by Lemma 8.2 (ii).
Let us check Condition (ii) of Definition 8.1. We have
We now present the main result of this section. We freely use the notation
of the previous sections. Recall the inverse t of the element detq = ad-q- 1 bc
of GL q(2). In SLq(2) we have t = 1.
PROOF. By Theorem 3.1, the transpose ( ab dc) of the matnx. (ac db)
is an Mq(2)-point of Mq(2). Consequently, there exists a unique antilinear
algebra endomorphism, of Mq (2) defined by the matrix identity
t (~ ~) C 1
(~ ~).
We conclude the proof by recalling that * = 5f. o
IV.9 Exercises
1. (Gauss) Show that
if n is odd
"~ (_l)k ( nk ) = { (1- q)(l- q3)0 ... (1- qn-l)
if n is even.
O:Sk:Sn q
m+k
( n+m+1 )
m+1 q
m
IV.9 Exercises 89
Oq(4) = zn-l,
(n).q (n - l).q
for all n ?: 1. Deduce that the q-exponential eq(z) is, up to a mul-
tiplicative constant, the only formal series solution of the equation
Oq(f) = J.
5. Let Aq[~, 17] be the algebra k{~, 17} /(e, 172, ~17 + q 17~). Set
(b) Check that (a~ + b17)(C~ + d17) = detq ~17. Deduce Part (b) of
Proposition 3.4.
(c) Find a M q(2)-comodule-algebra structure on Aq[~,17]'
90 Chapter IV. The Quantum Plane and Its Symmetries
IV.I0 Notes
The content of Section 2 on q-identities, as well as Exercises 1-3, is classical.
We borrowed it from [And76]' Chap. 3 and from [Cig79].
The q-exponential is an example of a q-hypergeometric series or basic
hypergeometric series, i.e., of a formal series L:n>O anz n such that each
quotient an+Iia n is a rational function of qn (where q is a complex param-
eter different from 0 and from 1). Basic hypergeometric series first appeared
in a note published by Heine [Hei46] in 1846. Since, q-analogues of most
classical functions and identities have been found. F.H. Jackson [Jac10]
introduced the q-differentiation operator 8q and its inverse which is the q-
integration. Nowadays, q-series appear in combinatorics, in number theory,
in statistical mechanics, and in the theory of Lie algebras. There are many
monographs on this vast subject, e.g., [GR90] [Sla66].
The operator Tq introduced in Section 2 is fundamental in the theory of
linear q-difference equations with polynomial coefficients. Such an equation
is a functional equation of the form
n
L Pi(Z)J(qi z ) = Q(z)
i=O
where Po(z), ... , Pn(z), Q(z) are polynomials and J(z) is a function. Using
the operator Tq , one can rewrite the equation above as (L:~=o PiT~)(J) = Q.
The articles by Adams [Ada29] and by Trjitzinsky [Trj33] are two classical
references on the formalism of the q-difference equations.
Sections 3, 5 and 6 are taken from Manin's book [Man88]. With Section
3 we entered the heart of the subject of Part I of this book. The bialge-
bras Mq(2), GLq(2), and SLq(2) of Sections 5-6 depend on one parameter.
There also exist two-parameter versions such as the algebra M p ,q(2) gen-
erated by four generators a, b, c, d and the six relations
IV.lO Notes 91
ba = pab, db = qbd,
ea = qae, de = ped,
be = pq-1eb, ad - da = (q-l - p) eb.
It has the same bialgebra structure as M(2). With the additional relation
ad - p-1be = 1, one gets the Hopf algebra SL p,q(2) of [AST91].
In higher dimension n > 2, Faddeev, Reshetikhin, Takhtadjian [RTF89]
defined the bialgebra Mq(n) generated by the generators (T/)l~i,j~n and
the relations
T:nT k
~J
= TkTm
J~'
rkTm _ TmT k
~J J~
= (q-l _ q) TmT k ~J
for i < j and k < m. The comultiplication and the counit are given by
n
b.(T/) =L Tik (8) T1 and (T/) = bij
k=l
The algebra Mq(n) is an iterated Ore extension and, like Mq(2), it possesses
a remarkable grouplike central element that is
for x, y ELand x', y' E L'. The canonical injections of Land L' into L ffi L'
and the canonical projections of L ffi L' onto Land L' are morphisms of
Lie algebras.
2. Given a Lie algebra L, we define the opposite Lie algebra LOP as the
vector space L with Lie bracket [-, - p given by r
[x, yr p = [y, x] = -[x, y].
The linear map op(x) = -x is a Lie algebra isomorphism from L to LOP.
3. Let I be an ideal of a Lie algebra L. There exists a unique Lie al-
gebra structure on the quotient vector space L/ I such that the canonical
projection from L onto L/ I is a morphism of Lie algebras.
4. Let f : L ---7 L' be a morphism of Lie algebras. Its kernel Ker (1) is
an ideal of L, the image f(L) is a sub algebra of L', and the induced map
L /Ker (1) ---7 f (L) is an isomorphism of Lie algebras.
5. Let A be an (associative) algebra. Set [a, b] = ab - ba for a, bE A. It
is easy to show that this bilinear map is antisymmetric and satisfies the
Jacobi identity. We also have [a, bc] = [a, b]c+ bra, c] for all a, b, c E A. This
Lie algebra will be denoted by L(A).
For any vector space V, we denote the Lie algebra L(End(V)) of all
endomorphisms of V by g[(V). When V is of finite dimension n, then g[(V)
is isomorphic to the Lie algebra g[(n) = L(Mn(k)) of n x n-matrices with
entries in the field k. It is clear that the commutator of two matrices with
zero trace is of trace zero. Consequently, the vector space $[( n) of traceless
n by n matrices is a Lie subalgebra of g[(n).
ideal of the tensor algebra T(L) generated by all elements of the form
xy - yx - [x, y] where x, yare elements of L. We define
U(L) = T(L)/I(L).
The above generators of I(L) are not homogeneous for the grading of T(L)
defined in II.5. Therefore there is no grading on the enveloping algebra
compatible with the grading of the tensor algebra. Nevertheless, U(L) is
filtered as a quotient algebra of T(L).
We define a map i L as the composition of the canonical injection of L
into T(L) and of the canonical surjection of the tensor algebra onto the
enveloping algebra. By definition of i L, we have i L([x, y]) = xy - yx, which
shows that i L is a morphism of Lie algebras.
Example 1. If L is an abelian Lie algebra, then U (L) coincides with the
symmetric algebra S(L). In particular, if L is the zero Lie algebra {O}, then
U({O}) = k. We also have U(LOP) = U(L)OP.
We now state the universal property of U (L).
Corollary V.2.2. (a) For any morphism of Lie algebras f : L ---> L',
there exists a unique morphism of algebras U(f) : U(L) ---> U(L') such that
U(f) 0 i L = i L' 0 f. We also have U(id L ) = idu(L)'
(b) If l' : L ---> L" is another morphism of Lie algebras, then
PROOF. (a) Apply Theorem 2.1 to A = U(L') and to the morphism of Lie
algebras iu 0 f.
(b) We have
Corollary V.2.3. Let Land L' be Lie algebras and L EB L' their direct
sum. Then
U(L EB L') ~ U(L) U(L').
PROOF. We first construct an algebra morphism cp from U(L EB L') to the
algebra U(L) U(L'). For any x ELand x' E L', set
This formula defines a linear map f from L EB L' into U(L) U(L'). Let us
show that f is a morphi3m of Lie algebras. For x, y ELand x', y' E L' we
have
'lj;( tp(x, x')) = 'lj;(x 1) + 'lj;(1 x') = iL!J)L' ((x, 0) + (0, x')) = i L!J)L' (x, x').
Consequently, 'lj; 0 tp = id. A similar argument shows that tp 0 'lj; = id. D
Corollaries 2.2 and 2.3 allow us to put a Hopf algebra structure on the
enveloping algebra U(L). Indeed, a comultiplication ~ on U(L) is defined
by ~ = tp 0 U (8), where 8 is the diagonal map x 1--+ (x, x) from L into L ffi L
and tp is the isomorphism U(L ffi L) ~ U(L) U(L) that was built in the
proof of Corollary 2.3. The counit is given by c = U(O) where 0 is the zero
morphism from L into the zero Lie algebra {O}. Finally, the antipode is
defined by S = U(op) where op is the isomorphism from L onto LOP of
Example 1. 2.
+X I Xn 1
where (J runs over all (p, q)-shujjles of the symmetric group Sn' and
S(X I X 2 X n ) = (_1)nXnX2XI
LffiL
lid!J)8
8!J)id
-----> LffiLffiL
The formula for ~ results from Theorem III.2.4. The definition of Sand
Lemma III.3.6 imply that S is an antipode for U(L). 0
For the sake of completeness, we give two additional important properties
of enveloping algebras.
Theorem V.2.5. Let L be a Lie algebra.
(a) The algebra U(L) is filtered as a quotient of the tensor algebra T(L)
(graded as in II.5) and the corresponding graded algebra is isomorphic to
the symmetric algebra on L:
gr U(L) ~ S(L).
(2.1)
for x E L, v E V, and v' E V'. According to nI.5, the Lie algebra acts on
Hom(V, V') by
(xf)(v) = xf(v) - f(xv), (2.4)
which can also be expressed as p(x)(f) = [p(x), f] for f E Hom(V, V'). In
particular, if V'is the trivial module k, then L acts on the dual vector
space V* = Hom(V, k) by
H = (~ ~1)' 1= (~ ~)
form a basis of 9((2). Their commutators are easily computed. We get
and
[I, X] = [I, Y] = [I, H] = O. (3.1)
The matrices of trace zero in 9[(2) form the subspace s[(2) spanned by
the basis {X, Y, H}. Relations (3.1) show that s[(2) is an ideal of 9[(2) and
that there is an isomorphism of Lie algebras
which reduces the investigation of the Lie algebra 9[(2) to that of s((2).
The enveloping algebra U = U(s[(2)) of s[(2) is isomorphic to the algebra
generated by the three elements X, Y, H with the three relations
XP Hq = (H - 2p)qXp, yP Hq = (H + 2p)qyP,
o
Proposition V.3.2. The set {Xiyj HkL,j,kEN is a basis of U(s[(2)).
We also have
1 1
[X,C] X[X, Y] + [X, Y]X + "2 [X, H]H + "2 H[X, H]
XH+HX-XH-HX=O.
PROOF. The third relation is trivial; the first two result from Lemma 3.1.
o
We now state the theorem describing simple finite-dimensional U-modules.
102 Chapter V. The Lie Algebra of 8(2)
0 n 0 0
0 0 n-l 0
p(n)(X) =
0 0 1
0 0 0 0
0 0 0 0
1 0 0 0
p(n)(Y) = 0 2 0 0
0 0 n 0
and
n 0 0 0
0 n-2 0 0
p(n)(H) =
0 0 -n+2 0
0 0 0 -n
Let us determine the action of the Casimir element on the simple module
V(n).
Lemma V.4.5. Any central element of U acts by a scalar on the sim-
ple module V(n). In particular, the Casimir element C acts on V(n) by
multiplication by the scalar n(n2+2) , which is non-zero when n > O.
PROOF. Let Z be a central element in U. It commutes with H which decom-
poses V (n) into a direct sum of one-dimensional eigenspaces. Consequently,
the operator Z is diagonal with the same eigenvectors {v = v o , ... , v n } as
H. In particular, there exist scalars aD, ... , an such that ZVp = apvp for
all p. Now
o
We finally show that any finite-dimensional U-module is a direct sum of
simple U-modules.
l.b. If the submodule V' is simple of dimension> 1, then Lemma 4.5 im-
plies that the Casimir element C acts on V' as a scalar a i- O. Consequently,
the operator Cia is the identity on V'. Now V IV' is one-dimensional, hence
a trivial module. Therefore C sends V into the submodule V', which means
that the map Cia is a projector of V onto V'. As Cia commutes with any
element of U, the map Cia is a morphism of U-modules. By Proposition
1.1.3, the submodule V" = Ker (Cia) is a supplementary submodule to V'.
2. General case. We are now given two finite-dimensional modules V' c V
without any restriction on the codimension. We shall reduce the situation
to the codimension-one case by considering vector spaces W' C W defined
as follows: W [resp. W'] is the subspace of all linear maps from V to V'
whose restriction to V' is a homothety [resp. is zero]. It is clear that W' is
of co dimension one in W. In order to reduce to Part 1, we have to equip
Wand W' with U-module structures. We give Hom(V, V') the U-module
structure defined by Relation (2.4). Let us check that Wand W' are U-
submodules. For fEW, let a be the scalar such that f(v) = av for all
v E V'; then for any x E L, we have
PROOF. It is enough to prove that, for all p with 0 :S: p :S: m, the module
V (n) 0 V (m) contains a highest weight vector of weight n+m - 2p. In effect,
if so, there exists a non-zero morphism of modules from V (n + m - 2p) into
106 Chapter V. The Lie Algebra of SL(2)
V(n) V(m). The module V(n+m- 2p) being simple, the kernel of such a
morphism has to be zero, which means that the morphism is an embedding
ofV(n+m-2p) into V(n)V(m). Thesubmodules V(n+m-2p) being
simple and of distinct highest weights, their sum in V(n) V(m) is direct.
Thus, the right-hand side of the Clebsch-Gordan formula embeds into the
left-hand side. To conclude, it suffices to check that both sides have the same
dimension. Now the dimension of V(n+m) EEl V(n+m - 2) EEl EEl V(n - m)
equals
m
L (n +m - 2p + 1) (n+ l)(m+ 1)
p=o
dim(V(n)) dim(V(m))
dim(V(n) V(m)).
PROOF. Set
P
a
t
= (_l)i (m-p+i)!(n-i)!
(m - p)!n! and w = L aivi V~~i'
i=O
Now,
(ki(n - i + 1) + (ki_l (m - p + i)
(_l)i (m - p + i)!(n - i)! (n _ i + 1)
(m - p)!n!
i-l (m - p + i - l)!(n - i + I)! ( .)
+ (- l) (m _ p )'.n., m - p +z
o.
D
Remark 5.3. (a) One deduces from Proposition 5.1 that the adjoint repre-
sentation V(2) is related to V(O) and V(l) by
(b) The dual module V(n)* is isomorphic to the simple module V(n)
(prove it). Consequently, we have the U-linear isomorphisms
and
xl = c(x)l (6.2)
108 Chapter V. The Lie Algebra of 8L(2)
(xy)(ab) x(y(ab))
x(L (y1a)(y"b))
(y)
L (x'(YI(a))) (x"(Y"(b)))
(x)(y)
L ((x'YI)a) ((x"y")b)
(x)(y)
L ((xy)la) ((xy)"b).
(xy)
o
The following examples show that module-algebra structures appear in
a number of situations.
Example 1. Let 'P be an automorphism of an algebra A. Consider the
algebra k[Z] of the group of integers with the bialgebra structure described
in IIL2, Example 2. If k[Z] acts on A by sending a generator of Z on 'P,
then A becomes a module-algebra over k[Z].
Let us describe module-algebras over enveloping algebras.
We now return to the Lie algebra .5[(2) and show how the affine plane
becomes a module-algebra over the enveloping algebra U (.5[(2)).
Theorem V.6.4. Define an action of the Lie algebra .5[(2) on the polyno-
mial algebra k[x, yl by
oP oP oP oP
X P = x oy , Y P = y ox ' H P = x ox - y oy
[X,YlP = x~
oy
(y OP)
ox
_y~
ox
(x OP)
oy
oP 02p oP 02p
x- + xy-- - y- - yx--
ox oyox oy oxoy
HP.
One similarly shows that [H,XlP = 2XP and [H, YlP = -2YP.
In order to conclude that we have a module-algebra structure, it is enough
in view of Lemma 6.3 to check that the generators X, Y, H act on k[x, yl
as derivations, which is clearly the case.
(b) Fix a non-negative integer n and set v = xn E k[x, Yln' Clearly, v is
a highest weight vector of weight n. For all p :::: 0 we have
v = ~ yPv = ( n ) xn-pyp
p p! P
if p ::; nand vp = 0 if p > n. Since the monomials {v p }p generate k[x, yln'
the latter is a .5[(2)-module generated by a highest weight vector of weight
n. Hence, by Theorem 4.4, it is isomorphic to the simple module V(n). D
Let us motivate this definition. Let 'P be the linear map from U to the
dual vector space H* defined by
Similarly, 'IjJ( x) (u) = < u, x > defines a linear map from H to U*. From
Proposition 111.1.2 we know that the dual spaces U* and H* carry natural
algebra structures. If, in addition, the vector space H is finite-dimensional,
then the dual space H* has a natural bialgebra structure induced by the one
on H (see 111.2, Example 1). We are now ready to state a characterization
for duality between bialgebras.
Proposition V.7.2. Given bialgebras U and H and a bilinear form < , >
on U x H, the bilinear form realizes a duality between U and H if and only
if the linear maps 'P and 'IjJ are morphisms of algebras.
If, moreover, H is finite-dimensional, then the bilinear form realizes a
duality if and only if'P is a morphism of bialyebras.
We shall say that the duality between U and H is perfect when both
maps 'P and 'IjJ are injective. In case U and H are finite-dimensional, a
perfect duality between them induces isomorphisms of bialgebras between
U and H* and between Hand U*.
PROOF. Let us express that 'P is a morphism of algebras. Recall that the
unit of H* is equal to the counit of H and that the product of two linear
V.7 Duality between the Hopf Algebras U(s[(2)) and 5L(2) 111
(af3)(x) = L a(x')f3(x")
(xl
for all x E H. Then the relations 'P(1) = 1 and 'P(uv) = 'P(u)'P(v) imply
< 1, x> = 'P(l)(x) = E(X) and
< UV,X > 'P(UV)(x) = ('P(u)'P(v))(x)
L 'P(u)(x')'P(v)(x") = L < u, x' >< v, x" > .
(x) (x)
It results that Relations (7.1) and (7.3) of Definition 7.1 are equivalent
to the fact that 'P is a morphism of algebras. By symmetry, we see that
Relations (7.2) and (7.4) are equivalent to the fact that 7jJ is a morphism
of algebras.
Now assume that H is finite-dimensional. Then the dual space H* is
a bialgebra. We have already expressed the fact that 'P is a morphism of
algebras. Let us express that it is a morphism of coalgebras. On one hand,
the relation E'P = E expressing that 'P preserves the co unit reads
Thus, the map 'P is a morphism of coalgebras if and only if Relations (7.2)
and (7.4) are satisfied. D
A(u) B(U))
p(u) = ( C(u) D(u)
112 Chapter V. The Lie Algebra of 8L(2)
p(u) =
( A(u)
C(u)
B(u) ) = ( < u, a>
D(u) <u,c>
< u,b >
< u,d > ).
Let us express that p(uv) = p(u)p(v). We have
Expanding this matrix product, we get exactly the four desired conditions
since, as we know from Chapter I, the coproduct on M(2) is defined by the
matrix relation
Lemma V.7.4. If Conditions C(x) and C(y) hold, then so does C(xy).
PROOF. Relation (7.2), and Conditions C(x) and C(y) imply that
L < u', x' > < v', x" > < u", y' > < v", y" > .
(n)(v)(x)(y)
L < u', x' > < u", y' > < v', x" > < v", y" >
(n)(v)(x)(y)
< uV,xy >.
D
The duality between M(2) and U is not perfect: the morphism 7j; is not
injective as the following lemma shows.
for any pair (u, v) of elements of U. On the other hand, by (7.3) we have
< 1, ad - be > = c (ad - be) = 1.
This implies that the linear map u f--+ < u, ad - be > is a morphism of
algebras from U to k. To show that this morphism coincides with the
co unit c, it suffices to check that both maps have the same values on the
generators X, Y and H. Now we have
< X,ad - be >
= c(a) < X, d > + < X, a> c(d) - c(b) < X, e > - < X, b > c(e)
= O=c(X).
Similarly, we get < Y, ad - be >= 0 = c(Y). Finally,
< H,ad - be >
c(a) < H,d > + < H,a > c(d) - c(b) < H,e > - < H,b > c(e)
= -1 + 1 = 0 = c(H).
o
As a consequence of the previous lemma, the morphism of algebras 1/J :
M(2) ---> U* factors through SL(2) = M(2)/(ad - be - 1). We still denote
by 1/J the induced morphism of algebras from SL(2) to U* and by < , >
the corresponding bilinear form.
Theorem V.7.6. The bilinear jorm < u,x >= 1/J(x)(u) realizes a duality
between the Hopj algebras U and SL(2).
PROOF. We already know that 1/J is a morphism of algebras. By Proposition
7.2 we are left with showing that cp : U ---> SL(2)* is a morphism of algebras
too. Now, the projection from M(2) onto SL(2) dualizes to an injective
morphism from SL(2)* into M(2)*. It is clear that, when composing the
latter with cp, we get the morphism of algebras cp : U ---> M(2)* investigated
earlier. Consequently, cp : U ---> SL(2)* is a morphism of algebras. This
shows that we have a duality between bialgebras.
It remains to examine the antipodes and to check Relation (7.5). Let us
start with the generators. In the abridged matrix form we have
(~ ~1)
d
<X, ( -e ~b) >
S(a) S(b) ) >.
<X, ( S(e) S(d)
One proceeds similarly with Y, H, and 1.
V.7 Duality between the Hopf Algebras U(.s[(2)) and SL(2) 115
< S(u),x > = < u,S(x) > and < S(v),x > = < v,S(x) >
for all x E SL(2), then < S(uv), x> = < uv, S(x) >. Similarly, let x, y be
elements of SL(2). If
< S(u),x > = < u,S(x) > and < S(u),y > = < u,S(y) >
for all u E U, then < S(u), xy > = < u, S(xy) >.
PROOF. Theorem III.3.4 (a) and Definition 7.1 imply that
f(xiyn-i) = Oni
is a highest weight vector with weight n ofthe U-module k[x, Yl~, which will
imply that k[x, Yl~ contains a submodule isomorphic to the simple module
V(n). Since
dim(V(n)) = n + 1 = dim(k[x, Yl~),
we get k[x, Yl~ ~ V(n).
116 Chapter V. The Lie Algebra of SL(2)
(7.9)
for all u E U and for all i such that 0 :s: i :s: n. Indeed, by definition of f,
by III.6, Example 2 and by Lemma III.7.4 we have
(uf)(Xiyn-i)
(u f)(t:.A(xiyn-i))
ta~G)(n~i)
t~
r=Os=O
G) (n~i)
ta~ G) (n~i)
Let us apply Relation (7.9) to H. A straightforward computation using
(7.2-7.3) and the definition of the bilinear form yields
< x,d >= c(c) < x,d- 1 > + < X,C > c(d- 1 ) = o.
Consequently,
< X,aid > = c(a)i < x,d> + < X,a i > c(d) = o.
o
V.8 Exercises 11 7
V.8 Exercises
1. Let L be a Lie algebra. Show that [L, L] is an ideal of L and that the
quotient Lie algebra Lab = L/[L, L] is abelian. Prove that if f is a
morphism of Lie algebras from L to any abelian Lie algebra V, then
r
there exists a unique linear map b from Lab into V such that f is
r
the composition of b and of the canonical projection from L onto
Lab.
Opf(x) = x( L xd(Xi)).
l::;i::;d
Deduce that, when p(Op) is invertible, there exists a vector v in
V such that f(x) = xv for all X in L.
118 Chapter V. The Lie Algebra of 8L(2)
6. Find all invariant symmetric bilinear forms of 5[(2) (as defined in the
previous exercise; assume that the field k is of characteristic zero).
12. (Bialgebra structure on the quantum plane) (a) Show that the formu-
las
equip the free algebra k{x,y} and the quantum plane kq[x,y] with a
bialgebra structure.
(b) Prove that an algebra R is a module-algebra over the bialgebra
k{x,y} [resp. over kq[x,y]] if and only if R possesses an algebra
endomorphism 7 and a 7-derivation 0 [resp. 7 and 0 such that the
relation 07 = q70 holds].
(c) Find all kq[x, y]-algebra structures on the polynomial algebra k[z]
(consider only the ones for which 7 is an automorphism). In par-
ticular, show that, when 7 is the algebra automorphism 7 q of k[z]
considered in IV.2, then 0 is necessarily a scalar multiple of Oq (see
Exercise 4 in Chapter IV).
V.9 Notes
There exist numerous textbooks on the theory of Lie algebras. See, for
instance, [Bou60][Dix74][Hum72][Jac79][Ser65][Var74]. The content of this
chapter is essentially taken from these sources. We found the proof of The-
orem 4.6 in Serre's book [Ser65]. As for Definition 7.1, we took it from
[Tak81]. Let us supplement the content of this chapter with the following
remarks.
(Free Lie algebras) Let X be a set. Consider the smallest Lie sub algebra
(X) of the free algebra k{X} containing X. Denote by ix the injection
of X into (X). The free Lie algebra (X) enjoys the following universal
property: For any set-theoretic map f from X into a Lie algebra L, there
exists a unique morphism of Lie algebras J : (X) ----+ L such that f = JOi x'
It follows from this universal property, from Proposition 1.2.1, and from
Theorem 2.1 that there is an isomorphism of algebras
U((X)) ~ k{X}.
A description of bases for (X) may be found in [Bou60], Chap. 2. See also
[Reu93].
(Primitive elements of the enveloping bialgebra) Any Lie algebra L is
contained in the Lie algebra of primitive elements of its enveloping algebra.
In characteristic zero, this embedding is an equality:
L = Prim(U(L)).
When applied to free algebras, one gets (X) ~ Prim(k{ X}) (see [Bou60],
Chap. 2).
(Real forms) A real form of a complex Lie algebra L is a real Lie subal-
gebra LR of L such that the embedding of the complexification LR EB iLR
into L is an isomorphism of complex Lie algebras. Here i denotes a square
root of -1. To any real form of L, one associates its conjugation, which is
the antilinear involutive endomorphism of Lie algebras a given by
a(x + iy) = x - iy
for all x, y E L R . Conversely, given any such involution of L, we obtain a
real form by
LR = {x ELI a (x) = x}.
For any real form of L with conjugation a, we define a Hopf *-algebra
structure on the enveloping algebra U (L) by * = SoU (a). In other words,
we have 1* = 1 and
for all Xl' ... 'X n E L. Conversely, suppose we have a Hopf *-algebra struc-
ture on the enveloping algebra U (L). Since * is a coalgebra morphism, it
120 Chapter V. The Lie Algebra of 5L(2)
preserves the Lie sub algebra of primitive elements, which is L (we are in
characteristic zero). It is easy to check that the subspace of all elements
x of L such that x = -x* is a real form of L. We thus see that the real
forms on a complex Lie algebra L are in one-to-one correspondence with
the Hopf *-algebra structures on U(L).
For instance, the real Lie sub algebra su(2) of 2 x 2-matrices M in s((2)
such that M = _t M is a real form of s((2). The vectors A = ~(X - Y),
B = ~(X + Y), iH form a real basis of su(2) such that
This proves that su(2) is isomorphic to the Lie algebra so(3) of real anti-
symmetric 3 x 3-matrices.
(Duality) Theorem 7.6 asserts the existence of a Hopf algebra morphism
from SL(2) to U(s((2))*. This morphism is actually an isomorphism from
SL(2) to the restricted dual U(s((2))o. This holds, more generally, for
any simply-connected algebraic group in characteristic zero (see [Abe80]
[Hoc8l] PS9lb] [Swe69]).
Chapter VI
The Quantum Enveloping Algebra
of 1[(2)
These q-analogues are more symmetric than the ones defined in IV.2, as
shown by the relations
[-n] = -[n] and [m + n] = qn[m] + q-m[n]. (1.2)
122 Chapter VI. The Quantum Enveloping Algebra of .5[(2)
Observe that, if q is not a root of unity, then [n] =1= 0 for any non-zero
integer. This is not so when q is a root of unity. In that case, denote by d
its order, i.e., the smallest integer> 1 such that qd = 1. Since we assume
q2 =1= 1, we must have d > 2. Define also
d if d is odd
{ (1.3)
e = d/2 when d is even.
Let us agree that d =e= 00 when q is not a root of unity. Now it is easy
to check that
[n] = 0~n == 0 modulo e. (1.4)
We also have the following versions of factorials and binomial coefficients.
For integers 0 :::; k :::; n, set [OJ! = 1,
if k > 0, and
[ n ] [n]! (1.6)
k - [k]![n - k]!
These q-analogues are related to those of IV.2 by
and
[ ~ ] = q-k(n-k) ( ~ ) q2 (1.8)
(x + y)n = t
k=O
l(n-k) [ ~ ] xkyn-k. (1.9)
-(m-1) K _ m-1 K- 1
[m] p m - 1 q ~
q_ q 1
m-1 K _ -(m-1) K- 1
[m] q q -1 p m - 1,
q-q
-(m-1) K _ m-1 K- 1
[m] q q Em - 1
q _ q-1
m-1 K _ -(m-1) K- 1
[m] E m - 1 q q _
q_ q 1
PROOF. The first two relations result trivially from Relations (1.11). The
third one is proved by induction on musing
(1.13)
Let us take as given for a moment that there exists an aI-derivation 8 such
that
8(F) = K - ~-1 and 8(K) = O.
q_ q 1
j-l
8(Fj KR) = L Fj- 1 8(F)(q-2iK)KR (1.14)
i=O
PROOF. We must check that, for all j, mEN and all , nEZ, we have
(1.15)
Let us compute the right-hand side of (1.15) using (1.11), (1.13), and (1.14).
We have
m-1
i=O
j-1
+ L pi-18(F)(q-2iK)KRFmKn
i=O
VI.2 Relationship with the Enveloping Algebra of 5[(2) 125
L
m-l
q-2-2(m-l) FJ+m- 18(F)(q-2iK)KHn
i=O
L
j-l
+ q-2Rm F m+j- 18(F)(q-2i-2m K)KHn
i=O
L
m-l
q-2Rm F m+j- 18(F)(q- 2i K)KHn
i=O
L
j+m-l
+ q-2Rm Fm+j-18(F)(q-2iK)KHn
i=m
j+m-l
q-2Rm ( L FJ+m- 18(F) (q- 2i K)K Hn )
i=O
q-2m8(FJ+m KHn)
8(Fj K . F m Kn).
KK- 1 = K- 1K = 1, (2.1)
and
?/J(E) = E, ?/J(F) = F, ?/J(K) = K, ?/J(L) = [E, F].
It is clear that r.p gives rise to a well-defined morphism of algebras from Uq
to U~. Let us show that ?/J : U~ ----+ Uq is well-defined too. It suffices to check
that the images under ?/J of the defining Relations (2.1) hold in the algebra
Uq . This is clearly true for Relations (2.1 ~ 2.2) and for [E, F] = L. For the
remaining relation in (2.3) we have
(q - q~l)?/J(L) = (q - q~l)[E, F] = K - K~l.
VI. 3 Representations of Uq
We assume in this section that the complex parameter q is not a root of
unity. Our aim is to determine all finite-dimensional simple Uq-modules
under this assumption by closely following the methods of Section V.4.
For any Uq-module V and any scalar A =I=- 0, we denote by VA the subspace
of all vectors v in V such that K v = Av. The scalar A is called a weight of
V if VA =I=- {O}.
o
Definition VI.3.2. Let V be a Uq-module and A be a scalar. An element
v =I=- 0 of V is a highest weight vector of weight A if Ev = 0 and if K v = Av.
A Uq-module is a highest weight module of highest weight A if it is generated
by a highest weight vector of weight A.
Proposition VI.3.3. Any non-zero finite-dimensional Uq-module V con-
tains a highest weight vector. Moreover, the endomorphisms induced by E
and F on V are nilpotent.
PROOF. Since k = C is algebraically closed and V is finite-dimensional,
there exists a non-zero vector wand a scalar a such that K w = aw. If
Ew = 0, the vector w is a highest weight vector and we are done. If not,
let us consider the sequence of vectors Enw where n runs over the non-
negative integers. According to Lemma 3.1, it is a sequence of eigenvectors
with distinct eigenvalues; consequently, there exists an integer n such that
Enw =I=- 0 and En+1w = O. The vector Enw is a highest weight vector.
In order to show that the action of E on V is nilpotent, it suffices to check
that 0 is the only possible eigenvalue of E. Now, if v is a non-zero eigen-
vector for E with eigenvalue A =I=- 0, then so is Knv with eigenvalue q-2n A.
The endomorphism E would then have infinitely many distinct eigenvalues,
which is impossible. The same argument works for F. 0
Kv = cqn-2Pv (3.1)
P P'
(3.2)
and
(3.3)
VI.3 Representations of Uq 129
We shall see in VII.I that Pc,a may be identified with the counit of a Hopf
algebra structure on Uq . It will imply that the module VI,a is trivial and
that any trivial Uq-module is isomorphic to a direct sum of copies of VI,a'
On the other hand, the module V_I a is not trivial.
On the (n + 1)-dimensional module Vc n' the generators E, F and K act
by operators that can be represented o~ the basis {va, vI' ... , v n } by the
matrices
0 [n] 0 0
0 0 [n-I] 0
Pc,n(E) = c
0 0 I
0 0 0 0
0 0 0 0
I 0 0 0
Pc,n(F) = 0 [2] 0 0
0 0 [n] 0
and
qn 0 0 0
0 qn-2 0 0
Pc,n(K) = c
0 0 q-n+2 0
0 0 0 q-n
(3.4)
and EVa = O.
Lemma VI.3.6. Relations (3.4-3.5) define a Uq-module structure on V(>.).
The element va generates V(>.) as a Uq-module and is a highest weight vec-
tor of weight >..
130 Chapter VI. The Quantum Enveloping Algebra of 5[(2)
KK-1v =
P
KEK-1v p =
We also have
[E,F]vp =
By analogy with the classical case, the highest weight Uq-module V('\)
is called the Verma module of highest weight .\. It enjoys the following
universal property.
It results from I = FUq n U;: that I is a two-sided ideal and that the
projection 'P from U;: onto k[K, K- I ] is a morphism of algebras. The map
'P is called the Harish-Chandra homomorphism. It permits one to express
the action of the centre Zq on a highest weight module.
zv = 'P(z)(A)V.
132 Chapter VI. The Quantum Enveloping Algebra of .5[(2)
z = cp(z) + I: FiPiE i .
i>O
PROOF. For any integer n > 0, consider the Verma module V(qn-1). By
(3.5) we have
EV n =
q-(n-1)qn-1 - qn-1 q-(n-1)
-1
_
Vn -
.
q-q
Thus, Vn is a highest weight vector of weight qn-1-2n = q-n-1. By Propo-
sition 4.4, a central element z acts on the module generated by vn as the
mUltiplication by the scalar <p(z)(q-n-1); but, since vn is in V(qn-1), the
element z also acts as the scalar <p(z)(qn-1). In other words, we have
Now,
K n + K- n = (K + K- 1)n + (terms of degree < n).
One concludes by applying the induction hypothesis. o
We are ready to state the main theorem.
Theorem VI.4.8. When q is not a root of unity, the centre Zq of Uq is
a polynomial algebra generated by the element C q . The restriction of the
Harish-Chandra homomorphism to Zq is an isomorphism onto the subalge-
bra of k[K, K- 1] generated by qK + q-1 K- 1.
PROOF. We already know that the restriction of <p to the centre is injective.
We are left with determining its image. By Lemmas 4.6 and 4.7, the latter
is contained in the subalgebra of k[K,K- 1] generated by qK + q- 1K- 1.
Consider the central element C q defined above. By (4.2) we know that
1
<p(Cq ) = (q-q _1)2(qK +q- 1K- 1),
134 Chapter VI. The Quantum Enveloping Algebra of ,5[(2)
which proves that the image of Zq is the whole subalgebra and that Cq
generates the centre. The latter is a polynomial algebra because the powers
of qK + q-l K- 1 are linearly independent for obvious reasons of degree.
o
The first big difference with the generic case appears in the following
statement.
Before we prove this proposition, we state two lemmas. The first one
implies that the centre of Uq is much bigger when q is a root of unity than
when it is not. The second one is a special case of a general statement on
finite-dimensional modules.
Proof of Proposition 5.2. Let us assume that there exists a simple finite-
dimensional module V of dimension > e. We shall prove that V has a
non-zero submodule of dimension::; e. Hence, a contradiction.
(a) Suppose there exists a non-zero eigenvector v E V for the action
of K such that Fv = O. We claim that the subspace V' generated by
v, Ev, ... ,Ee- 1v is a submodule of dimension::; e. It is enough to check
that V'is stable under the action of the generators E, F, K. This is clear
for K. Let us check that V'is stable under E. The vector E(EPv) = EP+1 V
belongs to V' if p < e - 1. If p = e - 1, we have
where c 1 is a scalar in view of Lemmas 5.3 and 5.4. Finally, V'is stable
under F thanks to Fv = 0 and Lemma 1.3.
(b) Now, suppose there is no non-zero eigenvector v E V for the action of
K such that Fv = O. Let v be a non-zero eigenvector for the action of K. We
have Fv -=I- O. We claim that the subspace V" generated by v, Fv, ... , F e- 1v
is also a submodule of dimension::; e. Again, V" is clearly stable under K.
It is also stable under F since the vector F(FPv) = FP+1 v belongs to V"
if P < e - 1. If p = e - I, we have
where c2 is another scalar, again in view of Lemmas 5.3 and 5.4. The scalar
c2 is not zero; otherwise, there would exist an integer p < e such that FP v
would be an eigenvector for K killed by F, which would contradict our
assumption.
In order to check that V" is stable under E, we use the central element
Cq defined in Section 4. By Lemma 5.4, it acts on V by multiplication by
a scalar c3 . By definition of Cq we get for p > 0
E(FPv) EF(FP-1 V )
(c q _ q-1(q_q-1)2
K + qK- 1 ) (FP-1 V )
which show;:; that E(FPv) sits in V". When p = 0, we use the same argu-
ment after observing that v = C;-l Fev. 0
(5.1)
(5.2)
(5.4)
(5.5)
(5.6)
if 0 :S p < e - 1 and by Fv o = 0, EV e_ 1 = ev o , and KV e_ 1 = p,q-2 ve _ 1
otherwise. These formulas determine another Uq-module, denoted V(p" e).
The following theorem which we admit without proof closes the list of
all simple finite-dimensional Uq-modules when q is a root of unity.
Theorem VI.5.5. Any simple Uq-module of dimension e is isomorphic to
a module of the following list:
(i) V(A, a, b) with b -I 0,
(ii) V(A, a, 0) where A is not of the form qj-I for any 1 :S j :S e - 1,
(iii) V(ql-j, e) with e -I 0 and 1 :S j :S e - l.
It should be added that all modules V (A, a, b) and V(p" e), including the
ones that are not in the list of Theorem 5.5, are indecomposable.
In the situation under investigation, the algebra Uq possesses an inter-
esting finite-dimensional quotient-algebra.
Definition VI.5.6. The algebra U q is the quotient of the algebra Uq by
the two-sided ideal generated by the central elements E e , Fe, and K e - 1.
We let it act on the vectors vp of the canonical basis of the module V(A, 0, 0)
(check that this module is killed by E e and Fe, but in general not by K e -1).
We assume that A is neither zero, nor a root of unity. Since Evo = 0, we
have
(5.8)
Since v o, ... , v e - I are linearly independent, Relation (5.8) implies that
s-r
2: aO,j,Hr AR = 0 (5.9)
=0
Assume Z -=I- 0, hence Y -=I- O. Denote by d(Z) [resp. by 8(Z)] the degree in
K [resp. the degree in K- 1 ] of the non-zero element Z of Uq written in the
above-mentioned basis. Relation (5.11) implies that
d(Z) = d(Y) +e and 8(Z) = 8(Y). (5.12)
Now, by definition of Z, we have
o :::; 8(Z) :::; d(Z) < e. (5.13)
Combining (5.12-5.13), we get d(Y) < 0 :::; 8(Z) = 8(Y). This is impossible;
hence, Z = O. 0
VI.6 Exercises
1. Compute [Ei, Fj] in Uq .
VI. 7 Notes
The algebra Uq = Uq (g[(2)) is due to Kulish and Reshetikhin [KR81]. Drin-
feld [Dri85][Dri87] and Jimbo [Jim85] independently generalized this con-
struction by defining an algebra Uq(g) for any complex semisimple Lie
algebra (more generally, for any symmetrizable Kac-Moody Lie algebra) g.
VI.7 Notes 139
K i Ej K~l
i = q aE
'J .i' K i Fj Ki~l = q
~aF
'J j'
K-K-:-l
[Ei' Fj ] = bij q-q
' ~'l'
We assume in this chapter that the field k is the field of complex numbers
and that q is not a root of unity. We now equip the algebra Uq = Uis[(2))
defined in Chapter VI with a Hopf algebra structure. Then we prove that
any finite-dimensional Uq-module is a direct sum of the simple modules de-
scribed in VI.3. We show later that Uq acts naturally on the quantum plane
of IV.1 and that it is in duality with the Hopf algebra SLq(2) of Chapter
IV. We shall also build scalar products on the simple finite-dimensional
Uq-modules. We describe the quantum Clebsch-Gordan formula and give
the main properties of the quantum Clebsch-Gordan coefficients.
VII. 1 Comultiplication
We resume the notation of the previous chapter. Set
[6.(E),6.(F)]
(1 I8l E +E I8l K)(K- 1 I8l F +F I8l 1)
- (K- 1 I8l F +F I8l 1)(1 I8l E +E I8l K)
K- 1 I8l EF + F I8l E + EK- 1 I8l K F + EF I8l K
- K- 1 I8l FE - K- 1 E I8l F K - F I8l E - FE I8l K
K- 1 I8l [E, F] + [E, F] I8l K
K- 1 I8l (K - K- 1 ) + (K - K- 1 ) I8l K
q _ q-l
6.(K) - 6.(K-l)
q _ q-l
and
and S2(K) = K. 0
tt
r=Os=O
qr(i-r)+s(j-s)-2(i-r)(j-s) [ ~ ] [ ~ ]
6.(~)i6.(F)j6.(K)R
(1 @ ~ + ~ @K)i(K-l @F+F@ 1)1(KR @K R).
Now,
(~@ K)(1 @~) = q2 (1 @ ~)(~ @ K)
and
(K- 1 @ F)(F @ 1) = q2 (F @ 1)(K- 1 @ F).
Applying Relation (VI.1.g), we get
6.(~)i = t
r=O
qr(i-r) [ ; ] ~i-r @ ~r K i - r
and
6.(F)j = ts=o
qs(j-s) [ ~ ] F S K-(j-s) @ Fj-s.
VII.2 Semisimplicity
In this section we shall prove that any finite-dimensional Uq-module is the
direct sum of simple Uq-modules when q is not a root of unity, which we
assume in this chapter. Let us start with a technical lemma on the simple
modules Vc:.n of VI.3.
144 Chapter VII. A Hopf Algebra Structure on Uq (s[(2))
PROOF. Define
q + q-l
C = Cq - c (q-q -1 )2
and on ~I,n by
or, equivalently,
(qn+2 _ cc/)(qn - cc/) = 0,
which would be contrary to the assumptions. D
eigenvalue cq2 -I=- c, hence it is zero. Let us prove that EV2 is zero too.
Indeed, writing EV2 = >'v 1 + /-w 2, we have
which implies /-lc(q2 - 1) = 0 and >.c(q2 - 1) = /-la. Thus, >. = /-l = O. One
proves in a similar way that F acts as 0 on V. Since [E, F] acts as 0, we
have K = K- 1 on V. In particular, since K- 1 V2 = cV 2 - av 1 , we have
a = -a, hence a = O. In this situation K is also diagonalizable and we
reach the same conclusion as before.
We now assume that dim (V') = p > 1 and that the assertion to be
proved holds in dimension < p. There is the following alternative: either
V'is simple, or it is not.
La. If V'is not simple, one uses the same argument as in Part La of the
proof of Theorem V.4.6.
Lb. Suppose now that the submodule V'is simple of dimension> 1. The
one-dimensional quotient module V/V' has weight c = 1. Let us consider
the operator C of Lemma 2.1; it acts by 0 on V/V'. Consequently, we have
CV c V'. On the other hand, C acts on V' as multiplication by a scalar
a -I=- O. It follows that Cia is the identity on V'. Therefore the map Cia is
a projector of V onto V'. This projector is Uq-linear since C is central. By
Proposition 1.1.3, the submodule V" = Ker (C / a) meets the requirements.
2. General case. We are now given finite-dimensional modules V' c V
without any restriction on the codimension. We shall reduce to the codimen-
sion-one case by considering vector spaces W' C W defined as follows: W
[resp. W'] is the subspace of all linear maps from V to V' whose restriction
to V'is a homothety [resp. is zero]. It is clear that W' is of co dimension one
in W. In order to reduce to Part 1, we have to equip Wand W' with Uq -
module structures. We give Hom(V, V') the Uq-module structure defined
in II1.5. Let us check that Wand W' are submodules of Hom(V, V'). For
fEW, let a be the scalar such that f (v) = av for all v E V'; then for all
x E Uq and v E V', we have
which implies f(K v) = O. This proves that KV" c V". Similarly, V" is
stable under K- 1 . On the other hand, we have for v, hence for K v in V",
o c(E)f(Kv) = (Ef)(Kv
f(S(E)Kv) + Ef(K- 1 Kv) = - f(Ev) + Ef(v).
Consequently, f(Ev) = 0, which implies that V" is stable under the action
of E. A similar computation shows that FV" c V". The subspace V" is
therefore a submodule. D
(3.1)
(3.3)
or to
(3.4)
It is well-known that, if 8 is a derivation of a commutative algebra, then
aR8 is a derivation too. In a non-commutative situation, this is no longer
the case. Nevertheless, the following assertion holds.
Lemma VII.3.1. Let 8 be a (cr, T)-derivation of A and a be an element of
A. If there exist algebra automorphisms cr' and T' of A such that
then the linear endomorphism aR8 is a (cr', T)-derivation and aT8 is a (cr, T')-
derivation.
PROOF. This follows from straightforward computations. D
aq (xmyn)
= [m 1 x m-l y n and (3.6)
ax
for all m, n 2: O. Let us describe all commutation relations between the
endomorphisms xc' Xr , YRl Yr' ax' ay' Oq/ox, Oq/oy. We say that a commu-
tation relation between two endomorphisms u and v is trivial if uv = vu.
ayYe,r = qYe,ray'
Oq Oq
oY a y = qa y oY ,
Oq Oq
oY xr = qXr oY ,
We also have
PROOF. (a) This part results from easy, but fastidious computations.
(b) First observe that, if Relation (3.3) holds for two elements a, a' of A,
then it holds for their product aa'. Indeed, we have
8aa~
a(a)8a~ + 8(a)Ta~
a(a)a(a')8 + a(a)8(a')T + 8(a)T(a')T
a(aa')8 + 8(aa')T.
148 Chapter VII. A Hopf Algebra Structure on Uq (s[(2))
We are reduced to checking Relation (3.3) for Oq/ox and Oq/oy in the case
when a = x and a = y. For Oq/ox we have
-1 Oq (OqX) -1 Oq Oq
(0' x 0' y)( X) C ox + ox / x = q Xc ox + 0' x = ox Xf
-1 Oq (Oqy) Oq Oq
(O'x O'y)(Y)c ox + ox fO'x = qypox = ox Yp .
(3.7)
(a) Formulas (3.7) define the structure of a Uq-module-algebra on kq[x, y].
(b) The subspace kq[x,Yln of homogeneous elements of degree n is a Uq -
submodule of the quantum plane. It is generated by the highest weight vector
xn and is isomorphic to the simple module VLn .
Theorem 3.3 is the quantum version of Theorem V.6.4. It shows that the
quantum plane contains all finite-dimensional simple Uq-modules.
PROOF. (a) We first show that the formulas (3.7) equip kq[x, yl with a
Uq-module structure. In other words, we have to check Relations (VI.l.lO-
l.12). We use Proposition 3.2.
Relation (l.10) is trivially verified. For Relation (l.11) we have
Oq Oq Oq Oq
[E,F] XfoyYrOx -YrOxXeoy
-1 Oq Oq Oq -1 Oq Oq Oq
q XeYr oyox +XeUyOX -q YrXCoxoy -YrUx oy
Oq Oq
XCUy Ox - YrUx OY
Uy(U x - u;l) - Ux(U y - u;l)
q _ q-1
UxU y-1 - UyU x-1
q _ q-1
K-K-1
q _ q-1 .
ul = c(u)l, (3.8)
and
K(PQ) = K(P)K(Q), (3.9)
E(PQ) = + E(P)K(Q),
PE(Q) (3.10)
F(PQ) = K- 1(P)F(Q) + F(P)Q, (3.11)
for any pair (P, Q) of elements of the quantum plane. Relation (3.8) fol-
lows easily from (3.5-3.7) and Relation (3.9) from the fact that K acts as
an algebra automorphism. By Lemma 3.1 and by Proposition 3.2(b), the
endomorphism Xc ~~ is a (id, u x u;l )-derivation and Yr ~~ is a (u;lu y , id)-
derivation, which implies Relations (3.10-3.11).
(b) We have Exn = 0, Kxn = qnxn, and
1 P( n) _ -P [n]! n-p p
lP]! F x - q [p]![n _ p]! x y.
A(u) B(u) )
p(u) = ( C(u) D(u) . (4.2)
(iii) If u = F2 K R, we have
(4.8)
(vii) If u = E2 K C, we have
(ix) If u = E2 F2 K R, we have
DA - AD = (q - q-l) BG.
From the above observations, we see that it is enough to perform the check-
ing for u = Kf, which is trivial, and for u = EF KR. In the latter case, (4.7)
implies
(DA - AD)(u) = q - q-l = (q _ q-l) (BG)(u).
D
As a consequence of Lemma 4.1 and of IV.3, there exists a unique mor-
phism of algebras 'Ij; from Mq(2) into U; such that
The duality between Mq(2) and Uq is not perfect, just as in the classical
case.
Theorem VII.4.4. The bilinear form < u, x> = 't/J(x)(u) realizes a dual-
ity between the Hopf algebras Uq and SLq(2).
PROOF. We use the same argument as in the proof of Theorem V.7.6. The
only difference lies with the antipodes. We first check Relation (V.7.5) for
the generators. Using the condensed matrix form, we have
For F we have
1, n-i
r=Os=O
i n-i
r=Os=O
i n-i
LL
r=Os=O
Cr,s
VII.6 Scalar Products on Uq (s[(2))-Modules 155
< E,a i > c(a)< E,a i - 1 > + < E,a >< K,a i - 1 >
< E, a i- 1 > = ... = < E, a > = O.
Similarly, if j > 0 we get
< E,d >= c(c) < E,d- 1 > + < E,c >< K,d- 1 >= O.
Consequently,
< E,aid > = c(a)i < E,d > + < E,a i >< K,d > = O.
o
(Vi,v i ) -_ q-(n-i-l)i [ ni ] .
156 Chapter VII. A Hopf Algebra Structure on Uq (.sl(2))
PROOF. Let us first assume that there exists a scalar product on VE n such
that (v, v) = 1. Let us show that (vi' Vj ) is necessarily of the pre~cribed
form. By definition and by (6.1) we have
1 1 . 1 1
(Vi' Vj ) = [ill (Pv, vj ) = [ill (v, T(F)'vj ) = [ill (v, (EK- )'vj ).
~qi(H1) (v K- i Eiv.)
[i]! ' ,
e i qi(H1) [n]! (v K-iv)
[i].
This proves the uniqueness of the scalar product. Let us now prove its
existence.
Clearly, there exists a non-degenerate symmetric bilinear form such that
.
( v" v).) =
q-(n-i-1)i [ ni ] J: ..
U,). (6.2)
(7.1)
we need give this formula only for the modules V1,n, henceforth denoted
for simplicity by Vn .
One proves Theorem 7.1 in the same way as Proposition V.5.l. It suffices
to check that the module Vn 129 V m contains a highest weight vector of weight
qn+m-2 p for any integer p such that 0 :::; p :::; m.
v(n+m-2p) = '""'
p
(_l)i [-
m p + ~']'[. n -
~.
']'
q-i(m-2 p+i+1) v(n) 129 v(m)
~ [m - p]![n]! 'p-'
,=0
PROOF. It is clear that v~n) 129 v~~~ has weight qn-2i+m-2(p-i) = qn+m-2P.
Let us prove that Ev(n+m-2p) = O. Recall that .6.(E) = 1129 E + E 129 K. It
follows that
Ev(n+m-2p)
xv(n) Q9v(m)
,-1 p-'
{ (n) (m)} .
Vi Q9 Vj O~i~n, O~j~m ,
(n+m-2p) _ ~ Fk (n+m-2p)
Vk - [k]! v
n m n+m- 2p ]
[
i j k
[7 ; n +~- 2p ] =0 (7.3)
[
n m n + m - 2p ] = [j + I]q-(n-2i) + [i] [n m n+m- 2P ]
i j +1 k+1 [k + 1] i j k .
(7.4)
VII.7 Quantum Clebsch-Gordan 159
we get
[k + 1]Vk~~m~2p) = FVkn+m~2p)
= ""'
~
(Y.
, ([p - i + k + 1]q~(n~2i)v(n)
, v(m) p~,+k+1
i
+ [~. + 1] Vi+1
(n) (m))
Vp~i+k
= ""'
~ ,
(Y. ([p - i + k + 1]q~(n~2i) + [i])v(n)
, v(m) p~,+k+I'
i
PROOF. The first two assertions are clear. Let us prove the last one. If
WI = VI v~ and W 2 = V 2 v~, we have
(~(X)(VI v~), V2 v~)
(w l ,T(x)w2),
160 Chapter VII. A Hopf Algebra Structure on Uq (.s[(2))
o L q-i(n-i-l)-j(m-j-l) [ 7] [7]
',J
X [7 ; n +~- 2p ] [7 7 n +~- 2q ]
when p i= q or k i= R, and
L q-i(n-i-l)-j(m-j-l) m
j
n +m-
k
2p ]2
i,j
(n)
vi
(m) _
Q9 Vj - q
-i(n-i-l)-j(m-j-l) [ ~." ] [mJ. ]
m n+m-2p [n n: n+ ~ - 2p ]
x L L qk(n+m-2 p -k-l) Z J vin + m - 2P ).
PROOF. (a) Arguing as in the proof of Theorem 6.2, one shows that
( V (n+m-2p) ,V(n+m-2 P)) -- 0
k
for some scalars, and ,'. Now, if k 2: f, the vector Ekv~n+rn-2q) is zero or
is a scalar multiple of the highest weight vector v(n+m-2 q ), which brings us
back to a previous case.
(b) Let us compute (vkn+rn-2p),v~n+rn-2q)). It is equal to
m
J
n+m - 2p
k ][~ m
s
n+m - 2q
f
L n ; n + ~ - 2p ] [ 7 mj n+m - 2q
f
H.i=p+k
q-2(n-1-1)-J(rn-J -l) [ n ] [ ; ] [ n m n +m - 2p ]
j k
Hj=p+k
n m n +m - 2q ]
x [ i J f .
On the other hand, we have
( (n+rn-2 p )
Vk
(n+m-2 q )) _ " s:
,vR - upqukR q
-k(n+m-2p-k-l) [ n + mk
- 2p ]
Applying (6.2), one gets the desired explicit expression for 'pk. o
For more details on the quantum Clebsch-Gordan coefficients, see [KR89]
[KK89] [Vak89] where they are expressed in terms of q-Hahn polynomi-
als, i.e., of certain orthogonal q-hypergeometric series (see also [GR90],
Chap, 7). Koelink-Koornwinder and Vaksman showed that the orthogonal-
ity relations of the q- Hahn polynomials were equivalent to the orthogonality
relations of the quantum Clebsch-Gordan coefficients. The corresponding
property for the classical Clebsch-Gordan coefficients was known already
(see [Koo90]).
162 Chapter VII. A Hopf Algebra Structure on Uq (s((2))
VII.8 Exercises
1. Compute S(EiFjKR) in Uq .
VII.9 Notes
The Hopf algebra structure of Uq (s((2)) is due to Sklyanin [Sk185]. The
Drinfeld-Jimbo algebras Uq(g) also have a non-commutative, non-cocom-
mutative Hopf algebra structure. In the cases A, D, E considered in VI. 7,
it is given on the generators (Ei' Fi , Kih~i~R by
Universal R-Matrices
Chapter VIII
The Yang-Baxter Equation and
(Co )Braided Bialgebras
Part II is centered around the now famous Yang-Baxter equation whose so-
lutions are the so-called R-matrices. We introduce the concept of braided
bialgebras due to Drinfeld. These are bialgebras with a universal R-matrix
inducing a solution of the Yang-Baxter equation on any of their mod-
ules. This provides a systematic method to produce solutions of the Yang-
Baxter equation. There is a dual notion of cobraided bialgebras. We show
how to construct a cobraided bialgebra out of any solution of the Yang-
Baxter equation by a method due to Faddeev, Reshetikhin and Takhtadjian
[RTF89]. We conclude this chapter by proving that the quantum groups
GLq(2) and SLq(2) of Chapter IV can be obtained by this method and
that they are cobraided.
which is equivalent to
(1.1)
for any VI' V 2 E V. The flip satisfies the Yang-Baxter equation because of
the Coxeter relation (12)(23)(12) = (23)(12)(23) in the symmetry group
53'
Here is a way to generate new R-matrices from old ones.
(AC 16> id v ) = A(C 16> id v ), (id v 16> AC) = A(id v 16> c),
Wo = Vo Q9 Vo and t = Vo Q9 VI - q-1V 1 Q9 Vo
1 2
and w2 = [2] F Wo = VI Q9 VI
for i = 1,2. This completes the proof of the first assertion in Proposition
1.3.
The second assertion results from tedious computation. Let us give some
details. We first observe that the matrix <I> of cp with respect to the basis
{vo Q9 vo, Vo Q9 VI' VI Q9 Vo, VI Q9 vd is given by
~~ (
n
A 0 0
0 a 'Y
0 'Y (3
0 0 0
where
q-l A + qp qA+q-lp A-P
a= (3= , 'Y = [2]'
[2] [2]
The automorphisms cp Q9 id and id Q9 cp can be expressed, respectively, by
the 8 x 8-matrices <1>12 and <1>23 in the basis consisting of the elements
VoQ9VoQ9Vo, VO Q9VOQ9Vl' VOQ9Vl Q9Vo, VOQ9Vl Q9V 1, VI Q9VoQ9Vo, VI Q9VOQ9Vl'
VI Q9 VI Q9 vo, and VI Q9 VI Q9 VI of V Q9 V Q9 V where
170 Chapter VIII. The Yang-Baxter Equation and (Co)Braided Bialgebras
A 0 0 0 0 0 0 0
0 A 0 0 0 0 0 0
0 0 a 0 'Y 0 0 0
0 0 0 a 0 'Y 0 0
<1>12 = 0 0 0 0 0 0
'Y /3
0 0 0 'Y 0 /3 0 0
0 0 0 0 0 0 A 0
0 0 0 0 0 0 0 A
and
A 0 0 0 0 0 0 0
0 a 'Y 0 0 0 0 0
0 'Y /3 0 0 0 0 0
0 0 0 A 0 0 0 0
<1>23 = 0 0 0 0 A 0 0 0
0 0 0 0 0 a 'Y 0
0 0 0 0 0 'Y /3 0
0 0 0 0 0 0 0 A
Now, <1>12<1>23<1>12 - <1>23<1>12<1>23
0 0 0 0 0 0 0 0
0 K -a/3'Y 0 0 0 0 0
0 -a/3'Y L 0 a/3'Y 0 0 0
0 0 0 -K 0 a/3'Y 0 0
0 0 a/3'Y 0 M 0 0 0
0 0 0 a/3'Y 0 -L a/3'Y 0
0 0 0 0 0 a/3'Y -M 0
0 0 0 0 0 0 0 0
where K = a((A - a)A - 'Y 2 ), L = a/3(a - (3) and M = /3b 2 + A(/3 - A)).
Suppose that we have proved that K, Land M are multiples of a/3'Y. Then
2. If qA + q-l/1 = 0, then
~~qA (y 0
q-I _ q
1
0
0
1
0
0
0
0
0
q-I
)
3. If q-I A + q/1 = 0,
U
then
~ ~q-'A
n
0 0
0 1
1 q _ q-I
0 0
It is clear that Cases 2 and 3 are equivalent within a change of basis after
exchanging q and q-I. As we shall see in the next example, the minimal
polynomial of <I> is of degree:::; 2.
Example 3. We now give an important class of R-matrices with quadratic
minimal polynomial. Such R-matrices will be used in Chapter XII to con-
struct isotopy invariants of links in R 3 .
Let V be a finite-dimensional vector space with a basis {e l , ... , eN}' For
two invertible scalars p,q and for any family {rij}l:S;i,j:S;N of scalars in k
such that r ii = q and rijr ji = P when i =I- j, we define an automorphism c
ofV@V by
c(ei @ ei )
if i < j
c(e i @ ej )
if i > j.
Proposition VIII.1.4. The automorphism c is a solution of the Yang-
Baxter equation. Moreover, we have
Definition VIII.2.1. Let (H, /-L, 'TI, 6., E) be a bialgebra. We call it quasi-
cocommutative if there exists an invertible element R of the algebra H H
such that for all x E H we have
(2.1)
where yjkj) = x~j) for any j ::::; p and y;k) = 1 otherwise. For instance, if
R = 2:i Si t i , then R31 will be the element of HQ93 given by
R31 = L ti 1 Si'
i
(2.3)
and
(2.4)
174 Chapter VIII. The Yang-Baxter Equation and (Co)Braided Bialgebras
and
L si rgy (td rgy (t i )" = L SiSj rgy tj rgy t i (2.6)
i,(t;) i,j
Example 1. Cocommutative bialgebras are braided with universal R-matrix
R=lrgyl.
Here is a non-trivial example.
Example 2. (Sweedler's four-dimensional Hopf algebra) Let H be the al-
gebra generated by two elements x, y and relations
x 2 = 1, y2 = 0, yx + xy = 0.
The set {I, x, y, xy} forms a basis of the under lying vector space. There is
a unique Hopf algebra structure on H such that
Observe that the antipode 5 is of order 4 and that, for any a E H, we have
S2(a) = xax- 1 . Set
where), is any scalar. It is easy to show that R,\ satisfies the conditions
of Definition 2.2, thus endowing H with the structure of a braided Hopf
algebra for any scalar ),. Observe that R-;.l = TH,H(R,\).
We now investigate a few properties of universal R-matrices. The follow-
ing lemma will be useful later. It shows how to form new quasi-cocommut-
ative Hopf algebras from a given one.
(H, fLOP, 7),.6., c, 5- 1 ,5, R- 1 ), (H, fL, 7), .6. oP, c, S-l, 5, R- 1 )
(b) If, furthermore, (H, /-L, 1), 11, c:, S, s-1, R) is braided, then so is
(H, /-L, 1), 11oP, C:, s-1, s, TH,H(R)).
PROOF. (a) As a result of Corollary III.3.5, we see that (H, /-Lop, 1), 11, C:, S-l)
and (H,/-L,1},ll oP ,c,S-l) are Hopfalgebras. In (H,/-LP,1), 11,10, S-l), Rela-
tion (2.1) reads ll oP (x) = R- 1 1l(x)R, whereas it becomes
Similarly, one shows that Relation (2.4) for R implies Relation (2.3) for
TH,H(R). 0
Theorem VIII.2.4. Let (H, /-L, 1), 11, c:, R) be a braided bialgebra.
(a) Then the universal R-matrix R satisfies the equation
(2.7)
and we have
(10 lSi idH )(R) = 1 = (idH lSi c)(R). (2.8)
(b) If, moreover, H has an invertible antipode, then
i,j,k i,j,k
(2.12)
176 Chapter VIII. The Yang-Baxter Equation and (Co)Braided Bialgebras
and
R12(b. id)(R)
(b. 0P id)(R)R12
(TH ,H id)(b. id)(R)R12
(TH ,H id)(R13R23)R12
o
VIII.2 Braided Bialgebras 177
(2.15)
l: (a{3)(si) t; = l: (a (3)(~(Si)) ti
Q9
Now, we have
so using (2.5) we get ~(AR(a)) = Li,j a(tjtj) si Q9 Sj' On the other hand,
178 Chapter VIII. The Yang-Baxter Equation and (Co)Braided Bialgebras
L AR(a') AR(a//)
(a)
L Sia'(t i ) sja//(tj)
i,j,(a)
L a(titj) Si Sj
i,j
(3.2)
The latter two equalities hold only when H has an invertible antipode.
Proposition VIII.3.1. Under the previous ."t,ypotheses,
(a) the map c~w is an isomorphism of H-modules, and
(b) for any triple (U, V, W) of H -modules, we have
c~v,w = (c~,w id v )(iduc~,w), c~,vw = (idvc~,w )(c~,v id w )
and
VIll.4 The Square of the Antipode in a Braided Hopf Algebra 179
PROOF. (a) We have to prove that c~w is H-linear. Now, by (2.1), for any
x E H we have '
c~w(x(v
, Q9 w)) TV,W (R~(x)(v Q9 w))
TV,W (~OP(x)R(v Q9 w))
~(X)TV,W (R(V Q9 w))
x(c~,w(v Q9 w)).
(b) We prove the second and the last relations, leaving the first one to
the reader. For U E U, v E V and W E W we get using (2.6)
i,j
c~,VW(U Q9 v Q9 w).
As for the last relation in Part (b) of Proposition 3.1, we have
(c~,wQ9idu)(idvQ9c~,w)(c~,vQ9idw)(uQ9vQ9w) = L tktjWQ9SktiVQ9SjSiU
ifj,k
and
(idwQ9c~,v )(c~,wQ9idv )(iduQ9c~,w )(UQ9VQ9w) = L tjtiWQ9tksiVQ9SkSju,
i,j,k
Both right-hand sides are equal in view of (2.11). An alternative proof will
be given in XIII.l. 0
(4.1)
(4.2)
(4.3)
PROOF. Let us first show that S2 (x)u = ux for all x. If y belongs to H 0 H,
Relation (2.1) implies the equality
(t. 0P 0 id)(y)(R (1) = (R (1)(t. 0 id)(y)
in H 0 H 0 H. When y = t.(x) for some x E H, we get
L S(ti)SiX' @ S2(XI')S(X") = UX @ 1,
i,(x)
(4.7)
Define the biduals V** and **V by V** = (V*)* and **V = *(*V). The
reader is invited to prove the following proposition.
Proposition VIII.4.4. Under the hypotheses of Proposition 4.3, the map
v r--; < -, uv > [resp. the map v r--; < -, u-1v >] from V to V** [resp. to
** V] is an H -linear injective map.
(4.8)
(4.10)
for all a E H. In view of (4.10) it is enough to show that ~(U)R21R = uu.
By (4.10) again and by Theorem III.3.4 we have
L ~(S(ti))~(si)R21R
L (S S)(~OP(ti))~(Si)R21R
L (S S)(~OP(ti))R21R~(si)
We now let the algebra H4 act on H H on the right by
Z,]
(S id)(R;-/ R 21 )
(S id)(l 1)
1 1.
184 Chapter VIII. The Yang-Baxter Equation and (Co)Braided Bialgebras
Hence,
Next,
and
(u 1)(id 5)(R- 1 R)
(u 1)(id 5)(1 1)
(u 1).
Finally, we have
r * r = r * r = E, (5.1)
(ii) we have
fLOP = r * fL * r, (5.2)
(iii) and
VIII.5 A Dual Concept: Cobraided Bialgebras 185
where * is the convolution operation on linear forms, and the linear forms
r 12 , r 23 and r 13 are defined by
r 12 =r c, r 23 = c r, r 13 = (c r)(TH,H id H ).
The linear form r is called the universal R-form of H. A Hopf algebra is
cobraided if the underlying bialgebra is.
This definition is dual to Definition 2.2. More precisely, Relation (5.2)
is dual to Relation (2.1), whereas Relations (5.3) correspond to Relations
(2.3-2.4). Conditions (5.1-5.3) can be reexpressed in the following way. For
any triple (x, y, z) of elements of H we have
(i)
l..:: r(x' y')r(x" y") = l..:: r(x' y')r(x" y") = c(x)c(y), (5.4)
(x)(y) (x)(y)
(ii)
yx = l..:: r(x' y')x"y"r(x lll y"') , (5.5)
(x)(y)
(iii)
where /-Lv and /-Lw are the actions of H on V and W respectively and where
we have identified R with the linear map from k to H H, sending 1 to R.
Let H be a cobraided bialgebra with universal R-form r. Given the H-
comodules V and W with respective coact ions Av : V ~ H V and
Aw : W ~ H W, we define the linear map
cv,w: VW ~ WV
186 Chapter VIII. The Yang-Baxter Equation and (Co)Braided Bialgebras
(this was obtained by reversing the arrows and interchanging V and W).
Using the conventions of nI.6 we can rewrite this definition for any v E V
and W E W as
Proposition VIII.5.2. (a) Under the previous hypotheses, the map cvw
is an isomorphism of H -comodules. '
(b) If U is a third H -comodule, we have
and
CU,v0W = (id v Q9 Cu,w)(cu,v Q9id w )
Moreover, we have
v, cv,w )(v Q9 w)
(c w 0
L (WH)(VH)VV Q9W W
(v)(w)
v Q9w.
VIII.5 A Dual Concept: Cobraided Bialgebras 187
The second equality follows from the coassociativity of the coactions while
the third one is a consequence of Relation (5.1) and the last one follows from
v
the counitarity of the coactions. One proves that C W is a right inverse to
Cv W in a similar way. '
We now prove that Relation (5.2) implies that cv,w is a map of comod-
ules, namely we have
This is equivalent to
L r(wH0uHVH)Ww0uu0Vv
(u)(v)(w)
cuv,w(u 0 v 0 w).
The second equality follows from the co associativity, and the third one from
Relation (5.7). One proves that cu,vw = (id v 0 cu,w)(cu,v 0 id w ) in a
similar way.
The last relation of Proposition 5.2 is a consequence of the previous
relations and of the naturality of the maps cr. We leave the proof to the
reader. A proof in a more general context will be given in XIII.l. 0
188 Chapter VIII. The Yang-Baxter Equation and (Co)Braided Bialgebras
Pick a family of indeterminates T!, where i and j both run over the set
{I, ... ,N}.
Definition VIII.6.2. The algebra A(c)is the quotient of the free algebra
F generated by the family (T!hsi,jSN by the two-sided ideal I(c) generated
by all elements err where
e:nJ. n = ~
~
ckTmT n _
ij k
~
~
TkTcmn
i j k (6.1)
lSk,SN lSk,SN
and i, j, m and n run over the indexing set.
PROOF. It is clear that the above formulas define unique algebra maps
A :F -+ F &; F and c: F -+ k.
k,,p,q k,,p,q
p,q k,,p,q
p,q k,,p,q
p,q p,q
and
c(c::;n)
k, k,
L C~fOkmOn - L 0ikOjcrt
k, k,
crr - cfr = o.
o
3. We now let A(c) coact on V. Define a linear map Av from V to
A(c) &; Von the basis {vih:5i:5N by
m,n
which is zero by definition of A(c).
5. We now establish the universality of A(c). Let (A', b..~) be a pair satis-
fying the conditions of Theorem 6.1. Then there exists a family (U{)l$;i,j$;N
of elements of A' uniquely determined by
b..~(Vi) = L u{ Vj.
l$;j$;N
'~
" ck1umu n _ ' " uku1cmn
ij k l ~ i j kl
l$;k,l$;N l$;k,l$;N
for all i, j, m and n. From this it is clear that the map! from F to A'
defined by !(T!) = u{ for all i and j extends to a bialgebra map factoring
through A(c). Let us check the relation b..~ = (f idv)b.. v . For any i we
have
l$;j$;N l$;j$;N
(6.4)
for all i,j,m and n.
rn,n
"6 ckf
'J
r(Tq
P
0 TmT
k f
n) - "
6 r(Tqp 0 rkTf)
'J
cmn
k
k,f k.f
m,n
and
r(x (9 yz) =L r(x' (9 y) r(x" (9 z). (6.6)
(x)
PROOF. The proof is similar to the proof of Lemma 6.5. Use the fact that
c- 1 is also a solution of the Yang-Baxter equation. 0
p,q p,q
The second equality results from the fact that c- 1 is the inverse of c. The
general case follows from the next lemma.
PROOF. We give the proof for the couple (x, yz). The proof for (xy, z) is
similar. In view of Relation (5.7) and Relations (6.6-6.7) we have
c(y)c(x)c(z)
c(x)c(yz).
o
The relation L(x),(y) r(x' @y')r(x" @y") = c(x)c(y) is proved similarly.
3. Condition (ii): We have to check that for any x and y in A(c) we have
p,q p,q
p,q
p,q
L r(x' @ y' z')x"y" z" L r( x' @ z')r( x" @ y')x'" y" z"
(x)(y)(z) (x)(y)(z)
194 Chapter VIII. The Yang-Baxter Equation and (Co)Braided Bialgebras
q 0 0 0 )
-1/2
( 0 q 0 0 (7.1)
q 000 1
o 0 1 q _ q-l
where ql/2 is an invertible scalar. This matrix has been displayed in Section
1 where we proved it was an R-matrix. The FRT construction associates
to e a cobraided bialgebra A(e) which we now describe.
Proposition VIII.7.1. The bialgebra A(e) associated to the R-matrix (7.1)
is isomorphic to the bialgebra Mq(2) of Definition IV.3.2.
)
o 0 o
( ~~!~~!
ae
ea
bd
db
ad
eb
~~)(60
be
da 0
q
o
o
0
0
1
o
1
q _ q-l
VIII.7 Application to GLq(2) and SLq(2) 195
An easy computation shows that these relations are equivalent to the six
relations
ba qab, db qbd,
ca qac, dc qcd,
cb bc, da-ad (q - q-l)bc,
defining the algebra Mq(2). This identifies A(c) and Mq(2) as algebras.
The corresponding comultiplications are clearly the same (compare (6.2)
and Theorem IV.5.1). 0
From this and from Theorem 6.4, we deduce the following important
result on Mq(2).
Corollary VIII.7.2. The bialgebra Mq(2) has a unique structure as a co-
braided bialgebra with universal R-form r determined by
where oX = q-l/2.
It is easy to check that the coaction of A(c) on the two-dimensional vector
space V coincides with the coaction of Mq(2) on the elements of degree I
of the quantum plane kq[x,y] (see IV.7).
We now show that GLq(2) and SLq(2) are cobraided with the same
universal R-form. Since SLq(2) is a quotient of GLq(2), it is enough to
prove this for SLq(2). We start with the following lemma.
Lemma VIII.7.3. For all x E Mq(2) we have
PROOF. Recall that SLq(2) is the quotient of Mq(2) by the ideal I generated
by the element detq -1. Now, Lemma 7.3 is equivalent to the statement that
r((detq-1)x) =r(x(detq-1)) =0
VIII. 8 Exercises
1. Consider a matrix of the form
p 2 a = pa 2 + abc, q2 a = qa 2 + abc,
p2 d = pd 2 + dbc, q2 d = qd 2 + dbc.
2. Consider the Hopf algebra H of Section 2, Example 2. Show that
there exists an automorphism lP of the Hopf algebra H such that
(lP lP) (R)J = R)..' if and only if there exists a non-zero scalar fL such
that )..' = fL2)...
6. Let (H, fL, ry,~, c, 8, 8-1, r) be a cobraided Hopf algebra with invert-
ible antipode 8.
8. Let A be the algebra k{s, t, e 1}/(S2, st+ts). Show that the following
formulas define a unique cobraided Hopf algebra structure on A:
~(t) = t t, ~(s) = s 1 + e 1 s,
10. Let c be the R-matrix of Proposition 1.4. Prove that A(c) is the
algebra generated by (T!)l$.i,j$.N and the relations
VIII. 9 Notes
The Yang-Baxter equation first came up in a paper by Yang [Yan67] as a
factorization condition of the scattering S-matrix in the many-body prob-
lem in one dimension and in work of Baxter [Bax72] [BaxS2] on exactly solv-
able models in statistical mechanics. It also played an important role in the
quantum inverse scattering method created around 1975-79 by Faddeev,
Sklyanin, Takhtadjian [FadS4] for the construction of quantum integrable
systems. Attempts to find solutions of the Yang-Baxter equation in a sys-
tematic way have led to the theory of quantum groups (see [DriS7]). Many
papers in the literature are devoted to the construction of R-matrices, e.g.,
[DriS5] [DriS7] [JimS6a] [JimS6b] [KSSO], to quote but a few.
The concept of a quasi-co commutative and of a braided (or quasi-tri-
angular) Hopf algebra is due to Drinfeld [DriS7] [DriS9a]. For a review, see
[Maj90b]. The four-dimensional Hopf algebra of Example 2 of Section 2
is due to Sweedler. The universal R-matrices R). were found by Radford
[Rad93a].
The dual concept of cobraided bialgebras appears in [Hay92] [LT91]
[Maj91b] [Sch92]. Cobraided bialgebras have properties dual to braided
bialgebras. We gave some of them in Exercises 5-7.
The FRT construction is due to Faddeev, Reshetikhin and Takhtadjian
[RTFS9]. The bialgebras M p,q(2) and Mq(n) of IV.I0 can be obtained by
this method (see Exercise 10). In Sections 5-6 we followed the treatment
proposed by [LT91].
Exercise 1 is taken from [Kau91] and Exercise 2 from [Rad93a]. The
cobraided Hopf algebra of Exercise S was found by Pareigis [ParSl] before
the advent of quantum groups.
Chapter IX
Drinfeld's Quantum Double
x =yz. (1.1)
zy = (z . y) zY. (1.2)
(3(z, y) = zY
for all y, y' E Hand z, z' E K. This group structure is wlled the bicrossed
product of Hand K. Furthermore, the groups Hand K can be identified
respectively with the subgroups H x {I} and {I} x K of H I><l K, and every
element (y, z) in H I><l K can be written uniquely as the product of an
element of H x {I} and an element of {I} x K:
(y,z) = (y,l)(l,z)
where y E Hand z E K.
(b) Conversely, let G be a group and H, K be subgroups of G such that
the multiplication on G induces a set-theoretic bijection from H x K onto
G. Then the pair (H, K) is necessarily matched and the previous bijection
induces a group isomorphism from the bicrossed product H I><l K onto G.
Y , =Z -1 . ( zy ') =z -1 .y -1 ,
z' = (ZZ-1.y-')-1.
Set (y', z')(y, z) = (Y, Z) where y' and z' are given the above values. We
have to show that (Y, Z) = (1,1). Multiplying the last identity by (y, z) on
the left, we get
( y,z ) -1 -_ ( -1.
z -1 ( z-1.y-l)_1)
y , z .
(y, l)(y', 1) = (yy', 1), (1, z)(l, z') = (1, zz') and (y, 1)(1, z) = (y, z),
which proves the remaining assertions of Part (a).
For the proof of Part (b), it suffices to review the arguments that led us
to Definition 1.1. D
z . y = y and zY = Z.
for all y, y' E Hand z E K. Then (H, K) is a matched pair and the
bicrossed product H [XI K is isomorphic to the semidirect product of K
by H. In this case, the identity (1, z)(y, 1)(1., z)-1 = ((z y), 1) proves that
H x {I} is a normal subgroup of H [XI K and that the action of K on H
corresponds to the conjugation in the bicrossed product.
a : G x X ---> X.
a . (xy) = ~
~ (a I . x I ) (a , ,x" . y), (2.1)
(a)(x)
a1=c(a)1, (2.2)
(ab)X = L ab'.x' b" X" , (2.3)
(b)(x)
F = c(x)l, (2.4)
L a'x' a" . x" = L a" x" a' . x' (2.5)
(a) (x) (a) (x)
uA( ax ) = "'"
L-t a I X I r59a " X " and c(a x) = c(a)c(x)1 (2.6)
(a)(x)
in X, and
its coproduct by
foraEAandxEX.
If the bialgebras X and A have antipodes, respectively denoted S x and
SA' then the bicrossed product is a Hopf algebra with antipode S given by
PROOF. The above formulas show that we equipped the bicrossed product
with the coalgebra structure of the tensor product of coalgebras X and A.
It is then clear that i x and i A are co algebra morphisms. It remains to be
proved that X ~ A has an algebra structure and that the coproduct and
the counit, as well as the embeddings ix and i A , are algebra morphisms.
Let us start with the associativity of the product. An easy but tedious
computation using Relations (2.1) and (2.3) and the fact that both a and
IX.2 Bicrossed Products of Bialgebras 205
in view of (2.6) and (2.7). To conclude, we show that Relation (2.5) implies
that the coproduct is a morphism of algebras. We have
.6. ( (x a) (y b)) = L x' (a' . y') a"'Y'" b' x" (a" . y") a""Y"" b".
(x)(a)(y)(b)
defines an antipode on the bialgebra X [Xl A. Using the fact that SA and
Sx are antipodes, we get
c(a)(L X'SX(X") 1)
(x)
c(a)c(x)l 1
c(xa)ll.
Similarly, we have
Proposition IX.3.1. The map (a, x) f---+ a x endows H with the structure
of a left module-algebra on the bialgebra H. We denote by adH the thus
208 Chapter IX. Drinfeld's Quantum Double
defined H -module, and we call this action the left adjoint representation
of H. Similarly, the map (x, a) f-+ x a endows H with the structure of a
right module-algebra on the bialgebra H. We denote by Had the H-module
defined this way, and we call this action the right adjoint representation of
H.
PROOF. We give the proof for the left adjoint representation. We first check
that (a,x) f-+ a x puts an H-module structure on H. Indeed, we have
1 x = x and
a 1 = L a'S(a") = c(a)1
(a)
and
o
Example 1. (Conjugacy in a group) Let G be a group and k[G] be the
corresponding Hopf algebra. The left adjoint representation of k[G] is given
by the formula
a x = axa- 1
for a,x E G.
Example 2. (Adjoint representation of a Lie algebra) Let L be a Lie alge-
bra and U(L) be its enveloping algebra equipped with its canonical Hopf
algebra structure (see V.2). The left adjoint representation of U(L) is given
by the formula
a x = ax - xa
for a, x E L. The corresponding representation of L is called the adjoint
representation of the Lie algebra L.
We now wish to deduce the so-called coadjoint representations of H on
the dual vector space H* from the above-defined adjoint representations.
We use the following lemma.
IX.3 Variations on the Adjoint Representation 209
< ab, f > = L < b, f' > < a, f" > (3.2)
(f)
These actions will be called the left and right coadjoint representations
of H. Applying Corollary 3.3 to the Hopf algebra
By Corollary III.3.5, the Hopf algebra (HCOP)* is isomorphic via the map
S* to the Hopf algebra (HOP)* = (H*, A *, c*, (J.L 0P ) * ,TJ*, (S-l)*, S*). This
isomorphism induces a right action of (HOP)* on H. We summarize this
with the following statement.
Proposition IX.3.4. Under the hypotheses of Corollary 3.3, there exists
a unique right (HOP)* -module-coalgebra structure on H given for a E H
and f E H* by
af = L
f(S-l(a"')a')a".
(a)
PROOF. Let f, g E H* and a E H. By Corollary 3.3, the action of (HCOP)*
on H is given by
<af,g> = L <alll,j"><a",g><a',S*(j'
(f)(a)
af = L f(S(a')a"')a".
(a)
o
We now state the main result of this section. It will allow us to construct
Drinfeld's quantum double in the next section.
a(a @ f) = a f = L f(S-l(a")?a')
(a)
and
(3(a@ f) = af = L f(S-l(a"')a')a"
(a)
where a E Hand f EX. Then the pair (H, X) of Hopf algebras is matched
in the sense of Definition 2.2.
L f'(S-1(a(2))x'a(1))!"(S-1(a(5))a(3))(a(4). g)(X")
(a)(f)(x)
L f' (S-l (a (2) )x' a(1))!" (S-l (a(6))a (3) )g(S-l (a(5) )x" a(4))
(a)(f)(x)
L 1(S-1 (a(6) )a(3) S-l (a(2) )x' a(1) )g(S-l (a(5) )x" a(4))
(a)(x)
( ab ) f = '""""
~ ab't' bIIf" .
(b)(f)
Now,
L a b'f b" f" L f' (S-l (b")S-l (a'")a' b')!" (S-l (b" lll )bill) a"b""
(b)(f) (a)(b)(f)
= L c(bl)J(S-l(bll)S-l(a'l)alb')alb"1
(a) (b)
= L J(S-l((ab)II/)(ab)') (ab)"
(ab)
= (ab)f.
IX.4 Drinfeld's Quantum Double 213
L a' I' 161 a" . f" L 1'(S-l(a ll )a' )a" 161 fl(S-l(alll)?a"l)
(a)(f) (a)(f)
We first give a more explicit description of D(H), then in the next sub-
section we prove that the quantum double is a braided Hopf algebra in the
sense of VIII.2.
As a vector space, we have D(H) = X 0 H. The unit of D(H) is 10 l.
Its counit and its comultiplication are given by
and
6.(f 0 a) = L (f' 0 a') 0 (f" 0 a") (4.2)
(a)(J)
Computing the right-hand side using the formulas of Theorem 3.5, we get
L 1 g'(S-l(a")?a') 0 g"(S-l(a"lII)a"')a""b
(a) (g)
We shall use Relation (4.4) in order to simplify our notations and write
f a instead of f Q9 a = i x (f)i H (a) whenever a is an element of Hand f is
a linear form on H. Under this convention, the multiplication in D(H) is
determined by the straightening formula
af = L f(S-l(a"')?a')a" (4.5)
(a)
where f E X and a E H.
When H is cocommutative, the bicrossed product construction of the
double of H can be reduced to the crossed product construction of Section
2, Example 3, as shown in the following statement.
af L f(S-l(a"')a')a"
(a)
L f(S-l(a"')a")a'
(a)
L f(l)s(a")a'
(a)
s(f)a,
af = L f(S-l(al')?a')a"
(a)
L f(S-l(a'I)?a")a'
(a)
L (a" f)a'
Ca)
L (a' . f)a".
Ca)
iEI iEI
Theorem IX.4.4. Under the previous hypotheses, the Hopf algebra D(H)
equipped with the element R = ~iEI (1 e i ) (e i 1) E D(H) D(H) is
braided.
R= L(10ei)0((eioS)01).
iEI
Consider an element ~ = b 0 U 0 C 0 v in H 0 X 0 H 0 X. Let us pair it
with RR using the duality between H and X. We get
i,jEI
c(b)v(l) L u( (L ei(c')) (L e (S(c j ll
))) eiej)
(c) iEI jEI
c(b)V(l)U(L c'S(c"))
(c)
c(b)v(l)c(c)u(l)
< 1 0 1 0 1 0 1, ~ > .
Consequently, RR = 1 0 1 0 1 0 1. One proves that R is a left inverse of R
in a similar way.
(2) We now check that (Do 0 id)(R) = R13R23 or, equivalently, that
Therefore,
i,jEI
c(a)c(b)v(l) L t(e')u(e")
(c)
L 1" (b )u( a"" ei )l' (e')e i (S-l (a"') e" a')v( a")
(a)(c)(f), iEI
L c(a"')f(be')u(e"a')v(a")
(a)(c)
L f(be')u(e"a')v(a").
(a) (c)
On the other hand, we have
< R~(f a),~ >
L < (1 ei)(f' a') (e i 1)(f" a"), ~ >
(a)(f),iEI
= L f(be')u(e"a')v(a")
(a)(c)
(4.11)
220 Chapter IX. Drinfeld's Quantum Double
which proves again that D(k[G]) is the crossed product of k[G] by itself,
where the algebra acts on itself by conjugation. Its universal R-matrix is
R= L g0 e g . (4.12)
gEG
Despite the fact that the quantum double is not cocommutative when G is
not abelian, its antipode is involutive, which implies that the element
u = L eg_,g (4.13)
gEG
H0H0V0H H0V
1 idH0TH,v0idH lTH,v
H0V0H0H V0H
1!lV0!l 1fl.vidH
V0H V0H0H
commutes.
IX.5 Representation-Theoretic Interpretation of the Quantum Double 221
'\'""
~ a I Vv Q9 a /I VH = '\'""
~
(
a 1/)
v v Q9 (1/)
a v H aI (5.2)
(a) (v) (a)(v)
(5.3)
2: a(aiv)j(ai )
< a, (2: j(ai)ai)v >
< a, fv >
< aj,v >
since j = I:i j(ai)a i . Incidentally, this observation implies that ~v is
independent of the choice of the basis.
In order to complete the proof that V is a crossed H-bimodule, we have
to check Relation (5.2) using (5.1). If a E H, v E V and j E H*, then
222 Chapter IX. Drinfeld's Quantum Double
L a'(aiv)f(a"a i )
(a),i
L 1"(a")a'((L f'(ai)ai)v)
(a) (f) .
L f(a""S~l(a"')?a')(a"v)
(a)
L c(a"')f(?a') (a"v)
(a)
L f(?a') (a"v)
(a)
L 1'(a') f"a"v
(a) (f)
L ai(a"v)f(aia')
(a),i
This, being true for any linear form f, implies (5.2). In the previous series
of equalities, we used the comultiplication on H*, Relation (5.1), the fact
that S~l is a skew-antipode, that c is a counit, and that f = Li f( ai)a i.
(b) Conversely, let V be a crossed H-bimodule. We show that V can be
given a D(H)-module structure. Observe that if (V, ~v : V ---> V (9 H) is
a right H-comodule, then V becomes a left module over the dual algebra
X = H* by
IX.6 Application to Uq (.s1(2)) 223
aU v}
(5.5)
We assume until the end of this chapter that q is a root of unity of order
d in the field k where d is an odd integer > 1. Let us resume the notation
of VL1. Recall
qn _ q-n
[n] = q _ q-l '
which is defined for any integer n, and the corresponding q-factorials [n]!.
We have [n] 1= 0 if 0 < n < d and [d] = O.
In VL5 we defined the algebra U q as the quotient of Uq by the two-
sided ideal generated by the three elements Ed, F d, Kd - 1. We proved in
Proposition VL5.7 that the finite set {EiFjK}o::;i,j,e::;d_l was a basis of
the underlying vector space of U q . We endow the algebra U q with a Hopf
algebra structure.
[ d ] _ [d]! _ 0
r - [r]![d - r]! -
(6.1)
We now apply the quantum double construction of Section 4 to the
Hopf algebra H = B q . We first determine X = (B~P)* as a Hopf al-
gebra. Consider the linear forms a and ry on Bq defined on the basis
{EmKnh:S:m,n:S:d_I by
< a, Em K n > = DmO q2n and < ry, Em K n > = DmI . (6.2)
Proposition IX.6.2. The following relations hold in the Hopf algebra X:
ad = 1, ryd = 0, arya- I = q-2ry,
~(a) =a a, ~(71) = 1 71 + 71 a,
c:(a) = 1, c:(ry) = 0,
8(a) = ad-I, 8(71) = _rya d- I .
Moreover, the set {ryiaj}O:S:i,j:S:d-I forms a basis of X.
PROOF. We start with the following lemma.
Lemma IX.6.3. For all integers i, j, m, n, we have
< ryia j , Em K n > = 8mi (i)!q2 q2 j (i+n).
PROOF. By Proposition VII.1.3, if a and j3 are linear forms on H, then the
product a(3 in X is given by
r~ r; ) ( q2
= 8mz. (,;)' n
'q2 q2 (i+ )
j o
226 Chapter IX. Drinfeld's Quantum Double
(4) Let us prove the last assertion of Proposition 6.2. As the dimension
of X is d 2 , it is enough to show that the set {1]i o:1}O::;i,j::;d_l is linearly
independent. Suppose there exists a relation of the form
L Aij1]i a j = O.
O::;i,j::;d-l
xa =L a(S-l(x"')?x') x".
(x)
Hence,
E(3 = -(3(K- 1E?) + (3(K- 1?) E + (3(K- 1?E) K. (6.5)
Proposition 6.4 is then a consequence of (6.4-6.5) and of the following
lemma. 0
PROOF. The surjectivity of X follows from the fact that the image of the
basis {T)ia J Ek Ki} generates U q'
In order to show that X is a map of algebras, it is enough to check that
the images under X of the generators E, K, a, T) satisfy the relations of
Proposition 6.4. Observe that (6.6) implies
X(E) = E, X(K) = K,
q _ q-1
x(a) = K, X(T)) = 2 FK.
q
Now, by definition of U q we have
x(K)x(a) x(a)x(K),
q _ q-l
X(K)x(T)) ~C::-2-KFK=q-2 X(T))X(K),
q
x(E)x(a) q-2 x(a)x(E).
IX.6 Application to Uq (s (( 2) ) 229
Finally, we get
-1
X(E)x(TJ) q-q EFK
q2
q - q-1 FEK + ~ (K _ K-1)K
q2 q2
'1
-1
_q-2(1_ q - q FKE-K 2)
-1
q- ; (1 Q9 F K + FK Q9 K)
q
+ X(TJ) Q9 x(a)
X(l) Q9 X(TJ)
(XQ9X)(~(TJ)).
Similarly,
X(S(TJ)) -X(TJa- 1 )
q _ q-1
q2 F
-1
q- ; S(K)S(F)
q
-1
q-q S(FK)
q2
S(X(TJ))
D
(q -l)k
- q k(k-l)/2+2k(i-j)-2ij Ek Ki Fk Kj.
[k]! q
OSi,j.kSd-l
(7.2)
Apply Relation (7.2) to the vector Em K n: using Lemma 6.3, we obtain the
linear system of equations
lI ij
"'kR Ukm
I: (m)''q2 q2C(k+n)
OSk,CSd-l
An argument similar to the one that proved the linear independence of the
family {r]i o:1}OSi,jSd-l in Proposition 6.2 shows that f.L~R = 0 for m i' i.
Computing the coefficients f.L~~ that are solutions of the linear system
lI ij q2R(Hn) = __1_ /5
"'~R (m)!q2 In
O<:;i,j,k<:;d-1
We now determine the coefficients Ci,j,k' Theorem 7.1 will follow from
Lemma 7.2. 0
PROOF. (a) We first express Ci,j,k in terms of C = co,o,o using the relations
~OP(x)R = R~(x)
0<:;i,j,k<:;d-1
+
O<:;i,j,k<:;d-1
and
R~(E)
0<:;i,j,k<:;d-1
+
O<:;i,j,k<:;d-1
"L-t q2j C.
t,],k
Ek Ki Epk Kj
O<:;i,j,k<:;d-1
2j-(k-1)
"[k]q
o -1
c. EkKipk-1Kj+1
',],k
+
k<d-1
_ ,], -
O<i . q - q
232 Chapter IX. Drinfeld's Quantum Double
2j+(k-1)
+ '"
~
[k] q -1
c
,,),k
Ek Ki Q9 F k - 1 Kj-l
OSi,j,kSd-1 q - q
+
OSi.,j,kSd-1
OSi,j,kSd-1
+
OSi,j,kSd-1
OSi,j,kSd-1
2k+(k-1)
'~
" [k] q -1
c
',],k
E k - 1 K i + 1 Q9 Fk Kj-1
OSi,j,ksd-1 q - q
2k-(k-1)
+ '"
~
[k] q -1
C
',],k
E k - 1 K i - 1 Q9 Fk Kj-1
OSi,j,kSd-1 q - q
+ L
OSi,j,ksd-1
R6.(F)
OSi,j,kSd-1
+ q -2i C. Ek F Ki rV. Fk Ki
,t,],k '6l
OSi,j,kSd-1
C.
',],k
= q-2(k+i) C .
I,)-l,k' (7.4)
q2(k+1)+k
-2j
q CH1 ,j,k Ci,j,k - [k + 1J q _ q-1 Ci - 1,j+1,k+1
q2(k+1)-k
+[k + 1J q _ q-1 Ci +l,j+1,k+1' (7.5)
qk+2
+[k + 1J q _ q-1 CH1 ,j+1,k+l'
hence
-1
q - q k-4j-2
CH1 ,j+1,k+1 = [k + 1J q Ci,j,k'
or, equivalently,
-1
q - q k-4j+l
Ci,j,k = [kJ q Ci - 1,j-1,k-1'
Therefore, we get
( -1)k
q - q qk(k+ 1)/2-4kj+2k(k-1)+k C
Ci,j,k = [kJ! i-k,j-k,O
( -1)k
q - q k(k-1)/2+2k(k-2j) -2(i-k)(j-k)
~! q q C
where + ... stands for a sum of monomials containing only positive pow-
ers of E or of F. We now use (.6. id)(R) = R 13 R 23 , which is Relation
(VIII.2.3). We have
whereas
O~i,e,m,j<d
O~i,e,m,j<d
Now, 2:0d
_J
qNj vanishes except when N is a multiple of d, in which case
the sum equals d. Therefore,
Kv(n)
p = qn-2 p v(n)
p' Ev(n)
p = [n - p + 1] v(n)
p-l' Fv(n)
p = [p + 1] v(n)
p+ 1 .
cR (v(n) 0 v(m))
Vn,VTn P r
eR
Vn,Vm
(v(n)
p
v(m))
r
1 (q_q-I)k [n-p+k]![r+k]! k(k-I)/2
q
d [k]! [n - p]![r]!
O"Si,j,k"Sd-1
x Qnm(k) v(m) v(n)
pr r+k p-k
where
Q;;:'(k) = L q2(i-j)k-2ij+i(n-2p)+j(m-2r) ,
a"Si,j<d
which we can rewrite as
where i runs over the set of all integers in [0, d - 1] such that
2i = m - 2r - 2k + ad.
As 2 is invertible modulo d, there exists only one integer i satisfying these
conditions. Therefore,
Q;;:'(k) = d q 2ik+i(n-2 p) = d q k(m-n)-pm-rn-2(k- p)(k+r)+(m+ad)n/2. 0
IX.8 Exercises
1. Let H be a bialgebra and C a co algebra. Prove that C is a module-
coalgebra over H if and only if there exists an H-module structure
on C such that the comultiplication ,0. : C ~ C @ C and the counit
c: C ~ k of Care H-module maps for the tensor product H-module
structure on C @ C and for the trivial H-module structure on k.
7. Let G be a finite group. (a) Show that a left module over the quantum
double D(k[G]) is a left G-module V with a decomposition of the form
V = EB9EG Vg such that hVg C Vhgh-l for all g, hE G.
IX.8 Exercises 237
(a) Prove that there exists a Hopf algebra structure on E for which
the product is the convolution of I1L3, the unit is 'f} 0 c, the
coproduct ~', the counit c', and the antipode S' are given by
EE E PH@PH* IH H*
. \-1
IS /lH,H.
IX.9 Notes
The quantum double construction is due to Drinfeld [Dri87]. Our pre-
sentation is inspired from [Maj90a] [Tak81] (see also [RSTS88]). Radford
[Rad93a] proved that the quantum double is a minimal braided Hopf al-
gebra, i.e., it has no proper braided Hopf subalgebras. Conversely, any
minimal braided Hopf algebra is finite-dimensional and is a quotient of the
quantum double of some Hopf algebra. More generally, if H is braided with
universal R-matrix R, consider the subspace A of H generated by all ele-
ments of the form (idH 0 o:)(R) where 0: is any linear form on H. Radford
showed that the subspace A can be given the structure of a Hopf subalge-
bra, and that there exists a map of braided Hopf algebras from D(A) to H
whose image is a minimal braided Hopf subalgebra of H.
In Section 4 we proved that the quantum double of H was isomorphic to
a crossed product when H is cocommutative. This is true more generally
when H is braided. For more details, see [Maj91a].
Exercise 11 presents a construction dual to Drinfeld's quantum double,
yielding cobraided Hopf algebras. We took it from Takeuchi [Tak92a] where
a dual version of Theorem 5.2 is also given (see also [PW90] [RSTS88]).
The term "crossed bimodule" is due to [Yet90]. It was called a "quantum
Yang-Baxter module" in [Rad93b].
The Hopf algebra Uq has been considered in [Lus90a][Lus90b]. A compu-
tation of its universal R-matrix was performed in [KM91] using a different
method. It also appears in work by Reshetikhin-Turaev [RT91] constructing
quantum invariants for 3-dimensional manifolds.
Part Three
Low-Dimensional
Topology and
Tensor Categories
Chapter X
Knots, Links, Tangles, and Braids
We now embark into a topological digression which will lead us into the
world of knots. The reason for the presence of this chapter in a book de-
voted to quantum groups is the close relationship between the newly dis-
covered invariants of links (such as the celebrated Jones polynomial) and
R-matrices. This relationship will become more precise in Chapter XII. In
this one we proceed to describe several classes of one-dimensional subman-
ifolds of the three-dimensional space R 3 , such as knots, links, tangles, and
braids. Since there are excellent textbooks on knot theory, we shall not
prove all assertions that can be found elsewhere. Nevertheless, all results
pertaining to the matter of this book, namely those connecting topological
problems with the algebra of quantum groups, will be proved in detail.
After defining knots and links in R 3 , we recall the classical problem of
their classification up to isotopy. Traditionally, one approaches this problem
by constructing algebraic isotopy invariants. One major step in this direc-
tion was undertaken in the 1920's by Alexander, who associated a polyno-
mial to each isotopy class of oriented links. The Alexander polynomial was
used to distinguish many links and has been a powerful tool in knot theory
since.
In the summer of 1984 Vaughan Jones found a different one-variable
polynomial which distinguished knots that the Alexander polynomial could
not distinguish [Jon85]. Shortly after, a new invariant appeared, the so-
called Jones-Conway polynomial, which is a two-variable generalization of
both the Alexander and the Jones polynomials. Another aim of this chapter
is to establish the existence and the main properties of the Jones-Conway
polynomial.
242 Chapter X. Knots, Links, Tangles, and Braids
"(
h t,-
)
=
{h(2t,-)
h'(2t-1,-)oh(1,-)
if
if
~ t ~ 1/2
1/2 ~ t ~ 1
defines an isotopy between Land L". In other words, the relation '" i is
transitive. 0
The reader will find a proof of this result in [BZ85], Prop. 1.10. As a
consequence, we shall suppress the subscripts i and c from the symbols '""i
and,"" c and henceforth speak of isotopic or equivalent links.
We end this section with a definition of a trivial link.
om = 00 ... 0 (m times).
Trivial links of the same order, but with different orientations, are always
isotopic. Therefore we need not specify the orientation of a trivial link.
(b) (The linking number) This is a more refined invariant which dates
back to Gauss. Let us consider two connected components L1 and L2 of
a link L. Consider a diagram of L (to be defined in Section 3). It shows
crossings of L1 and of L 2 . We associate to each crossing P an integer
E(P) = 1 defined as in Figure 2.2.
x x
E(P)=+l E(P) =-1
where P runs over all crossings of L1 and L 2 . This number does not de-
pend on the projection and is an isotopy invariant for links of order 2. For
instance, we have lk( 00) = 0 for the trivial link with two components,
and lk(H) = 1 for the Hopf link H drawn in Figure 2.3. It follows that the
Hopf link is not trivial.
(c) (The fundamental group of a link) Define 7r( L) = 7r1 (R3 \ L) as the
fundamental group of the complement of the link in R3 (the definition
of the fundamental group is given in the Appendix to this chapter). For
the trivial knot, the group 7r( 0) is isomorphic to Z. More generally, the
group of the trivial link of order m is isomorphic to the free group F m on
m generators. By the very definition of isotopy, the fundamental group of
a link is an isotopy invariant. It is a very powerful invariant as one can
see from a theorem of Dehn's which asserts that a link L of order m is
trivial if and only if 7r( L) = Fm' In general, the group of a link is non-
abelian. Though it is possible to give a presentation of 7r(L) by generators
and relations from a plane projection of L, it is very difficult to use this
246 Chapter X. Knots, Links, Tangles, and Braids
Lemma X.3.3. Any link diagram may be turned after appropriate changes
of crossings into a link diagram representing a trivial link in R3.
PROOF. Consider a link diagram. Pick a vertex and start moving along the
link, leaving a trail of red paint on the edges. At each crossing point, make
the red edge into an overcrossing edge unless the other edge is already red,
in which case the first edge is made into an undercrossing one. Apply this
procedure to each connected component. The resulting link diagram (ob-
tained from the original ones by a series of changes of crossings) represents
a trivial link. 0
PROOF. We sketch the proof. For details, see [BZ85]. Let L be a link in
R3. Consider the set S of all possible linear projections of R3 onto a fixed
plane. Given a projection 7r of S, there exists a homeomorphism h of R3
such that 7ro(h(L)) = 7r(L). It is therefore enough to show that the subset
Sreg of those projections 7r of S such that 7r(L) is a regular link projection
is not empty. Now S is in bijection with R2. Therefore we can transport
the topology of R2 onto S. What we shall actually prove, is that Sreg is
dense in S for this topology.
Let 7r be an element of S \ Sreg. Then in the projection 7r(L) we may
have the following singularities: some crossing point may be of order 2:: 3
or some vertex may sit in the interior of some edge. This happens when the
direction of the projection 7r passes through three edges or when it passes
through a vertex and an edge. In the first case, the direction sweeps over a
portion of a quadric; this projects to a part of a conic. In the second case,
it determines by projection a segment of the plane. Identifying S with R 2 ,
we see that S \ Sreg is contained in a finite number of straight lines and
conics of the plane. Therefore Sreg is dense in S. 0
248 Chapter X. Knots, Links, Tangles, and Braids
Lemma X.3.5. Two generic diagrams are isotopic if and only if they are
obtained from one another by a finite number of operations belonging to the
following set:
(A) a generic isotopy,
(B) an isotopy interchanging the order of the vertices with respect to the
height,
(C) a Reidemeister Transformation (0), and
(D) an isotopy in the neighbourhood of a local maximum as depicted in
Figure 3.9 and its images under reflection in the plane of the page, in a
horizontal line and in a vertical line.
Lemma X.3.6. Two generic diagrams are isotopic if and only if they are
obtained from one another by a finite number of the following operations:
(A) a generic isotopy,
(B) an isotopy interchanging the order of the vertices with respect to the
height,
(C) a Reidemeister Transformation (0), and
(E) an isotopy in the neighbourhood of a crossing point as shown in
Figure 3.10 and its images under reflection in the plane of the page.
X.3 Link Diagrams 251
XL\l
Figure 3.10. An isotopy in the neighbourhood of a crossing point
Figure 3.11.
In Figure 3.11 the first operation is of type (0), the second one of type (A)
and (B) while the third one is of type (E).
Figure 3.12.
In Figure 3.12 the first and fourth operations are of type (E), the second
one of type (0), and the third one of type (A) and (B). Reflecting the
previous transformations with respect to the plane of the paper or with
respect to a vertical line takes care of their images under these reflections.
As for the reflections with respect to a horizontal line (they involve local
252 Chapter X. Knots, Links, Tangles, and Braids
minima), our assertion follows from the set of operations of Figure 3.13
where the first and last ones are Transformations (C), the middle one is
authorized by Figure 3.11, and the remaining ones are of type (A) and (B).
o
Figure 3.13.
X X
X+
Figure 4.1.
X
Theorem X.4.2. There exists a unique map L 1-+ P L from the set of all
oriented links in R3 to the ring Z[x, x-I, y, y-l] of two-variable Laurent
polynomials such that
(i) if L '" L', then P L = PL"
(ii) the value of P on the trivial knot is 1, and
(iii) whenever (L+, L_, Lo) is a Conway triple, we have
(4.1)
(4.2)
(4.3)
PROOF. The mirror image of the Conway triple (L+,L_,Lo) is the triple
(L_, L+, Lo). Consequently, we have
(4.4)
where (L+,L_,Lo) runs over all Conway triples of /C. The A-module Y is
called the skein module of R 3 .
PROOF. Figure 4.2 implies that (Qi8l n , O'81 n , QiSIn+l) is a Conway triple for
all n 2: 1. By definition of Y we get
Assuming Proposition 4.7 and using the ring map (q,m : A ----> C defined
by (q,m(x) = qm and (q,m(y) = q-q-l, which furnishes C with a A-module
structure, we see that IJ>m,q extends by linearity to a unique A-linear map
IJ>~,q from A[KJ to C. Now for any Conway triple (L+,L_,L o), we have
for any integer m > 1 and any complex number q that is not a root of
unity. Since IJ>m,q(O) =f:. 0, we have f(qm,q_q-I) = O. Since this is true for
an infinite number of distinct powers of q, the polynomial f is divisible by
the polynomial y - (q - q-I). The latter assertion holds for infinitely many
complex numbers q, which is possible only if the polynomial f is zero. This
proves the injectivity of Q. D
Application 4.8. We end this section with the computation of the Jones-
Conway polynomial of the right-handed trefoil knot K and of the Hopf
link H. Figures 4.3-4.4 show that (H, 00, 0) and (K, 0, H) are Conway
triples.
Figure 4.3.
X.5 Tangles 257
Figure 4.4.
By (4.1) we have
-1
xPH = X
-1
Poo + yPo = x
-1 X - X
y
+ y.
Hence,
PH = (x- 1 - X- 3 )y-1 + x- 1 y.
A similar computation yields
for the right-handed trefoil knot. By Corollary 4.3, we see that the Jones-
Conway polynomial of the mirror image K is given by
This proves that the trefoil knot is not isotopic to its mirror image, a fact
already observed by Dehn [Deh14] in 1914.
X.5 Tangles
This section is devoted to the concept of tangles which generalizes the no-
tion of links. Tangles will be used extensively in Chapter XII, in particular
for the proof of Proposition 4.7.
For any integer n > 0, we set [n] = {1, 2, ... ,n}. When n = 0, we agree
that [0] is the empty set. We denote by I the closed interval [0, 1] and by
R 2 the real plane.
The boundary condition in Definition 5.1 means that the tangle intersects
the two boundary planes of R2 x I transversally. Observe that a link in
R2 x I is a tangle of type (0,0). Figure 5.1 shows an example of a tangle
that is not a link.
-~--y-~---
~
___Y' ___
Figure 5.1. A tangle
Given a tangle L of type (k, f), we define two finite sequences s(L) and
b(L) consisting of + and - signs. If k = 0, then s(L) = 0 is the empty set
by convention. Similarly if = 0, we set b(L) = 0. In the general case, we
define
S(L)=(Cl"",Ck) and b(L) = ('rfl""''rfc)
where ci = + [resp. 'rfi = + ] if the point (i, 0, 0) [resp. the point (i, 0,1)]
is an endpoint [resp. an origin] of L. We have ci = - and 'rfi = - in the
remaining cases.
Let us give a few examples of tangles that shall be used in the sequel.
1. We denote the polygonal arcs [(1,0,1), (1,0,0)] and [(1,0,0), (1,0,1)] by
1 and i respectively. We have s(1) = (+), b(1) = (+), s(i) = (-), and
b(i) = (-).
2. The tangles X+ and X_ of Figure 4.1 can be defined by
and
n= [(2,0,0), (3/2,0,1/2)] U [(3/2,0,1/2), (1,0,0)]. (5.3)
X.5 Tangles 259
We have s(n) = (-, +), b(n) = 0, s(n) = (+, -), and b(n) = 0 (see
Figure 5.2).
v
Figure 5.3).
Clearly, if Land L' are combinatorially equivalent, they have the same
boundaries and are of the same type. Isotopies are defined as follows.
Again if Land L' are isotopic, they have the same boundaries and are
of the same type. The isotopy is shown to be an equivalence relation for
tangles in the same way as it was for links (see Lemma 1.5). We have the
following counterpart of Proposition 1.6.
As for links, we shall suppress the subscripts i and c from the symbols
and rv c and we shall henceforth speak of isotopic or equivalent tangles.
rv i
A crossing point ofrr is a point of the tangle projection sitting in the inter'ior
of at least two edges. The order of a crossing point P is the number of
distinct edges in the interior of which P sits.
(b) A tangle projection is regular if each crossing point is of order exactly
2.
Lemma X.5.7. Two generic tangle diagrams are isotopic if and only if
they are obtained from one another by a finite number among the following
operations:
(A) a generic isotopy,
(B) an isotopy interchanging the order of the vertices with respect to the
height,
(C) a Reidemeister Transformation (0), and
(E) an isotopy in the neighbourhood of a crossing point as shown in
Figure 3.10 and their images under reflection in the plane of the page.
idb(L) 0 L rv L rv L 0 ids(L)'
X.6 Braids
We now consider a special class of tangles, called braids. Fix an integer
n'?1.
1 i i +1 n
X\
Figure 6.2. The braid O"i
1 i+I n
><
Figure 6.3. The braid 0";1
X.6 Braids 265
C 0 id v @(n-2) if i = 1,
idV@(i-l) 0 C 0 idv@(n-i-l) if 1 < i < n -1, (6.3)
id v @(n-2) 0 C if i=n-1.
PROOF. First it is clear that equivalent braids have the same permutation.
Thus the map factors through Bn' It is a morphism of groups because we
have CJ(L' 0 L) = CJ(L') 0 CJ(L), CJ(1n) = id, and CJ(L-l) = CJ(L)-l. The
permutation of the braid CJi is the transposition (i, i + 1). The surjectivity
of the map follows from the fact that such transpositions generate the
symmetric group Sn' 0
This lemma is not surprising in view of Moore's theorem which gives the
following presentation in terms of generators and relations for the symmet-
ric group Sn: it is generated by the n - 1 transpositions Si = (i, i + 1) and
by Relations (6.1-6.2) where CJ i has been replaced by si' as well as by the
additional relations s; = 1 for i = 1, ... ,n - l.
One big difference between symmetric groups and braid groups is that
the former are finite groups while the latter are infinite groups when n > l.
Moreover, the group Bn has no torsion, that is to say, all elements i- In
have infinite order.
endowed with the subset topology of en. The symmetric group Sn acts on
Yn by permutation of the coordinates. Let Xn = Yn/ Sn be the quotient
space with the quotient topology. The space Xn is the configuration space
of n distinct points in e. Consider the following set p = {I, 2, ... ,n} of n
distinct points in X n .
such that for all tEl we have fi(t) "I fj(t) whenever i "I j and
1(0) = (1,2, ... ,n) and {ll(I),12(1), ... ,ln(1)} = {1,2, ... ,n}.
Proposition X.6.12. Two braids with n strands are equivalent if and only
il the corresponding loops are isotopic in X n .
We can transpose the composition of tangles on the level of loops. Let I
[resp. 1'] be the loop in Xn corresponding to a braid L [resp. L']. It is easy
to see that the loop I l' = ((f I'h, ... , (f I')n) corresponding to the braid
L' 0 Lx, composed in the sense of tangles, (see Section 5) is given for all i
by
if 0:::; t :::; 1/2,
( ') () {Ii (2t)
II i t = 1~(i)(2t -1) if 1/2 :::; t :::; 1
where u = u(f) is the permutation of I. The loop I l' is called the product
of the loops I, and 1'. We have u(f1') = u(f') 0 u(f).
Given a loop I = (fl' ... ,in) corresponding to a braid L, the loop I-I
defined by
li-l(Z) = la--1(i)(I- z),
where again u = u(f), corresponds to the inverse braid L -1. The loop
corresponding to the braid In is the constant map z f--+ (1,2, ... , n).
X.7 Exercises 269
We have the following lemma which is the counterpart for loops of Lem-
mas 5.10-5.11.
X.7 Exercises
1. (Centre of the braid group) Let n be an integer> 2. Show that the cen-
tre of the braid group Bn is generated by the element (0'1'" O'n_l)n.
if j = i,
if j = i + 1,
if j ~ i, i + 1.
Prove that there exists a morphism X of the braid group Bn into the
group of automorphisms of Fn such that X(O'i) = Xi for all i.
X.8 Notes
Classical references on knot theory are [Bir74][BZ85][Kau87a][Rei32][RoI76].
The Jones polynomial VL was defined in [Jon85] [Jon87]. Its two-variable
extension PL appeared in a number of papers written almost simultane-
ously [FYH+85] [Hos86] [LM87] [PT87] (see also [HKW86] [Kau91]). For
Theorem 4.2 we followed the proof given by Thraev in [Tur89].
(Smooth tangles) There is a version of tangles and isotopies where piece-
wise-linear maps are replaced by C= maps and the boundary condition
of Definition 5.1 is replaced by a transversality condition. Such smooth
tangles project to smooth tangle diagrams. It may be shown that smooth
isotopy classes of smooth tangles are in bijection with isotopy classes of
tangles as defined in Section 5 (see [BZ85]).
(Framed tangles) Let us define a normal vector field on a smooth tangle L
as a C= vector field on L that is nowhere tangent to L and that is given by
the vector (0, -1,0) at all points of the boundary 8L. One may suggestively
think of a tangle with a normal vector field as a tangled ribbon defined as
follows: one edge of the ribbon is the tangle itself whereas the other one
is obtained from the first one by a small translation along the vector field.
A framing of the tangle L is a homotopy class of normal vector fields on
L where two normal vector fields are said to be homotopic if they can be
deformed into one another within the class of normal vector fields. One can
extend the concept of isotopy from tangles to tangles with framings. Isotopy
classes of tangles with framings are called framed tangles or ribbons. We
x.s Notes 271
shall see in Chapter XIV that ribbons give rise to an interesting categorical
structure, the so-called ribbon categories.
Framed tangles can be represented by tangle diagrams in the sense of
Section 5 just as ordinary tangles are. Take a tangle diagram. By defini-
tion, it represents the following framed tangle: the underlying tangle is the
tangle represented by the diagram and the framing is determined by the
constant normal vector field (0, -1, 0) that is perpendicular to the plane of
the diagram and points to the reader. Any framed tangled may be repre-
sented by a planar diagram in such a way. We already know this for the
underlying unframed tangle. To represent a general framed tangle with a
vector field whirling around it, it is enough to know how to represent a ver-
tical tangle around which the vector field turns by an angle of 27T or of - 27T.
The corresponding ribbons appear in Figure 8.1 and may be represented
by the diagrams of Figure 8.2.
Figure 8.1.
Figure 8.2.
The link a, called the closure of (J, is isotopic to the one represented in
Figure 8.4. Alexander [Ale23] showed that any link in R3 was equivalent
to the closure of some braid. Non-equivalent braids may have equivalent
closures. Define an equivalence relation ~ on the set of all braids by: (J ~ (J'
if (J and (J' have equivalent closures. Then Markov's theorem ([Mar36]; for
a proof, see [Bir74]) states that ~ is the equivalence relation generated by
conjugation in the braid groups and by relations of the form (J ~ in ((J )(J;1
where (J E Bn and in is the morphism of Bn to Bn+l defined by in ((Ji) = (Ji
for i = 1, ... , n -1. As a consequence, any family (fn : Bn ----t C)n>O of set-
theoretic maps with values in a set C such that for all n and all (J, T E Bn
is a homotopy from fl to f2
We define 7f 1 (X,*) as the set of homotopy classes in L*X. We have the
following lemma.
PROOF. (a) If h [resp. h'] is a homotopy from f to 9 [resp. from l' to g']'
then (s, t) f---t (h(s, - )h'(s, - ))(t) is a homotopy from ff' to gg'.
(b) A homotopy from (f 1') f" to f (f' f") is given by
f( S~I) if 0< -
t < 8+1
- 4 '
1'(4t - s - 1) if s+1 < < t s+2
4 - - 4 '
f" (4t2~;2) if 05+2
4 -
< < t -
1.
if O<t<s+1
- - 2 '
if s+1
2 -
<t <
-
1
XL1.1 Categories
Definition XL1.1. A category C consists
(1) of a class Ob(C) whose elements are called the objects of the category,
(2) of a class Hom(C) whose elements are called the morphisms of the
category, and
(3) of maps
such that
276 Chapter XI. Tensor Categories
s(id v ) = b(id v ) = V,
(b) Jor any morphism J E Hom( C), we have
idb(f) 0 J = J 0 ids(f) = J,
(c) Jor any morphisms J, g, h satisJying b(J) = s(g) and b(g) = s(h), we
have
(h 0 g) 0 J= h 0 (g 0 J).
Here Hom( C) x Ob(C) Hom( C) denotes the class of couples (J, g) of com-
posable morphisms in the category, i.e., such that b(J) = 8(g). The conven-
tional notation for the composition of J and 9 is go J or gJ. The object
s(J) is called the source of the morphism J and b(J) is its target. For the
identity morphism of an object V we write id v . We denote by Homc(V, W)
the class of morphisms of the category C whose source is the object V and
whose target is the object W. If J E Homc(V, W), we write
J: V ---+ W or v-.L..w.
A morphism from an object V to itself is called an endomorphism of V.
The class of all endomorphisms of V is denoted End(V). A morphism J
from V to W in the category is an isomorphism if there exists a morphism
g: W ---+ V such that go J = id v and Jog = id w .
Everybody knows (or at least uses) the category Set of sets and the
category Gr of groups. We have already made use of the category Vect(k)
[resp. of Vectf(k)] consisting of vector spaces [resp. of finite-dimensional
vector spaces] and of linear maps over a field k. In Chapter I we used the
category Alg of algebras and the category A-Mod of left A-modules where
A is an algebra. We have also considered the category Cog of coalgebras.
We define the product oj two categories C and V as the category C x V
whose objects are pairs of objects (V, W) E C x V and whose morphisms
are given by
It is clear that if F : C --> C' and G : C' --> C" are functors, then the
composition G F is a functor from C to C". For any category C, there exists
a functor ide, called the identity functor of C, which is the identity on the
objects and on the morphisms in C. The inclusion of a subcategory in a
category is a functor.
F(V) ~~ G(V)
IF(f) 1G(f)
F(W) 2~ G(W)
commutes.
If, furthermore, Tf(V) is an isomorphism of c' for any object V in C, we
say that Tf : F --> G is a natural isomorphism.
278 Chapter XI. Tensor Categories
rt : id v -+ FG and B: GF -+ idc
GF(V) ~ V
1GF(f) 11
GF(V/) ~ v/
commutes. It results that if F(f) = F(f'), hence GF(f) = GF(f'), then
we have f = 1'. Therefore, the functor F is faithful. Using the natural
isomorphism rt in a similar way, we prove that G is faithful too. Now con-
sider a morphism 9 : F(V) -+ F(V/). Let us show that 9 = F(f) where
f = B(V') 0 G(g) 0 B(V)-l. Indeed,
e(V/) 0 GF(f) 0 B(V)-l = f = B(V/) 0 G(g) 0 B(V)-l.
Since F is fully faithful, there exists a unique morphism G(g) from G(W)
to G(W') such that
One checks easily that this defines a functor G from V into C and that
1] : id v ---> FG is a natural isomorphism. In order to show that F and G
are equivalences of categories, we have only to find a natural isomorphism
() : GF ---> idc. We define ()(V) : GF(V) ---> V for any object V E Ob(C) as
the unique morphism such that F(()(V)) = 1](F(V))-l. It is easily checked
that this formula defines a natural isomorphism. D
and
G(W) G(7)(W))I(GFG)(W) (}(G(W))IG(W)
s(f 129 g) = s(f) 129 s(g) and b(f 129 g) = b(f) 129 b(g),
(c) if l' and g' are morphisms such that s(f') = b(f) and s(g') = b(g),
then
(f' 129 g') 0 (f 129 g) = (f' 0 1) 129 (g' 0 g), (2.1)
(d) and
idv@w = id v 129 id w . (2.2)
Relation (2.1) implies that
j 129 9 = (f 129 idb(g)) 0 (ids (f) 129 g) = (idb(f) 129 g) 0 (f 129 ids(g))' (2.3)
This means that, for any triple (U, V, W) of objects of C, there exists an
isomorphism
auvw .
, : (U 129 V) 129 W ---+ U 129 (V 129 W) (2.4)
such that the square
aU,V,lV
(U 129 V) 129 W U 129 (V 129 W)
1
----t
luvwx (U 0 V) 0 (W 0 X) (2,6)
1 aU,V,W0X
U 0 ((V 0 W) 0 X) idu<8lav,w,x
U 0 (V 0 (W 0 X) )
such that
Iv TV
I0V ---+ V V0I ---+ V
1 idI <8I1 11 [resp . 1 / <81 id I 11 ] (2.8)
I0V' ~ V' V'0I ~ V'
and
(V Q9 W) Q9 I ~ V Q9 (W Q9 1)
'\., rVQ1IW ,/' idv<8Irw
VQ9W
commute for any pair (V, W) of objects of C.
PROOF. Consider the diagram
( U Q9 (I Q9 V)) Q9 W
/ ' a<8lidw
( (U Q9 1) Q9 V) Q9 W
'\., (ru<8lid v) (id U <8lIVV
<8Iidw <8Iidw
I
(U Q9 V) Q9 W
a la a
U Q9 (V Q9 W)
/ ' idvQ1IW
ru<81 id u <81 "
(lv<8Iidw)"
(U Q9 1) Q9 (V Q9 W) id u Q9 lV<8IW U Q9 ((I Q9 V) Q9 W)
'\., a ,/' idu <8Ia
UQ9 (I Q9 (V Q9 W) )
Here we dropped the subscripts of the associativity constraint a for sim-
plicity. The outside hexagon commutes by the Pentagon Axiom (2.6). The
naturality (2.5) of a implies the commutativity of the two middle squares
whereas (2.9) implies the commutativity of the top square and of the lower
left triangle. Consequently, the lower right triangle commutes as well. Set-
ting U = I, we get
id J Q9 (lV<8IW 0 a) = id J Q9 (lv Q9 id w )
This relation, together with the naturality of the left unit constraint (2.8)
and the fact that l is an isomorphism, implies lV<8IWoa = lvQ9id w , which ex-
presses the commutativity of the upper triangle in the statement of Lemma
XI.2.2. A similar proof works for the other triangle. 0
284 Chapter XI. Tensor Categories
Lemma XI.2.3. Let I be the unit of a tensor category. For any object V
we have
of Proposition 11.1.3. The pentagon and triangle axioms are clearly satisfied.
There are some important examples of subcategories of Vect(k) preserv-
ing the tensor structure. For instance, if G is a group, then the category
k[G]-Mod of representations of Gover k, or, equivalently, of k[G]-modules,
is a subtensor category of Vect(k) where the tensor product U (9 V of two
G-modules and the field k are given the following G-structures:
for 9 E G, u E U, v E V and)", E k.
We know from Chapter III that the group algebra k[G] is an associative
algebra over k with a comultiplication and a counit. We now investigate
such types of algebras. Let A be an associative unital k-algebra with a
morphism of algebras ~ : A ----+ A (9 A, called the comultiplication, and a
morphism of algebras E : A ----+ k, called the counit. Let us denote by A-
Mod the category of left A-modules (alias, representations of A). If U, V
are left A-modules, the tensor product U (9 V is a left A (9 A-module. The
comultiplication allows to pull back this A (9 A-module structure into an
A-module structure. It is given by
PROOF. Let (A, 'P, r], L.\, c) be a bialgebra. It follows from Proposition III. 5. 1
that (A-Mod, 0,1 = k, a, l, r) is a tensor category.
Conversely, let (A, 'P, r], L.\, c) be an algebra with comultiplication and
counit. Suppose that (A-Mod, 0, 1= k, a, l, r) is a tensor category. Let us
prove that L.\ is coassociative and that 10 is a counit in the sense of Definition
III. 1. 1.
Let us start with the coassociativity of L.\. Consider the associativity
constraint aA A A- By hypothesis, it is A-linear, which means that for
a,u,v,w E A ~~ have
aA,A,A (a (( u 0 v) 0 W) ) = a aA,A,A ( (u 0 v) 0 w) .
By definition of the associativity constraint, this can be reexpressed as
Setting u = v = w = 1 E A, we get
90 h = Pn,m(g, h) E G n+m
Relations (3.4) are satisfied. The category GL(k) becomes a tensor category
in this way.
(b) A family of subgroups {Gn}n of GLn(k) preserved by the maps Pn m
gives also rise to a strict tensor category. For instance, take the family 'of
symmetric groups Sn' the latter being realized as a subgroup of GLn(k) via
the permutation matrices. The resulting category S is a tensor category.
F(au,v,w)
F((U0V)0W) I F(U0 (V 0 W))
IF(u)
(4.1)
10 F(U) ----> F(U)
and rF(U)
F(U) 0 I ~ F(U)
commute for all objects (U, V, W) in C. The tensor functor (F, 'Po, '(2) is
said to be strict if the isomorphisms 'Po and 'P2 are identities of V.
(b) A natural tensor transformation 'rJ : (F, 'Po, '(2) -7 (F', 'P~, 'P~) be-
tween tensor functors from C to V is a natural transformation 'rJ : F -7 F'
288 Chapter XI. Tensor Categories
such that the following diagrams commute for each couple (U, V) of objects
in C:
1
/ 'Po
convention. If S = (VI'.' Vk ) and S' = (Vk+I' ... ' Vk+n) are nonempty
sequences of S, we denote by S * S' the sequence
(5.1)
obtained by placing S' after S. We also agree that S * 0 = S = 0 * S. To
any sequence S of S, we assign an object F(V) of C defined inductively by
where all opening parentheses are placed on the left-hand side of VI.
c
We are now ready to define the category str : its objects are the elements
of S, i.e., the finite sequences of objects of C, and its morphisms are given
by
Homes,,(S, S') = Homc(F(S) , F(S')).
This defines a category whose identities and composition are taken from C.
c
The rest of the section is devoted to the proof that str is a strict tensor
category equivalent to C.
Proposition XI.5.1. The categories cstr and C are equivalent.
PROOF. The map F defined above on the objects of str extends to a c
c
functor F : str -+ C which is the identity on morphisms, hence fully
faithful. As any object in C is clearly isomorphic to the image under F of a
sequence of length one, we see that F is essentially surjective. This proves
the proposition in view of Proposition 1.5. Observe that G(V) = (V) defines
a functor G : C -+ cstr which is the inverse equivalence to F. Indeed, we
have FG = ide and 0 : GF -+ ides" via the natural isomorphism
and
rp(5,5' * (V)) = (rp(5,5')@id v )oa-;;}S),F(S'),v' (5.3)
The following lemma will be used in the proof of Theorem 5.3.
The first and last equalities follow from (5.3), the second and fifth ones
from the naturality of the associativity constraint, i.e., from Relation (2.5),
the third from the Pentagon Axiom (2.6), and the fourth one from the
induction hypothesis. 0
XI.6 Exercises 291
We can now define the tensor product f * f' of two morphisms f : S --; T
c
and f' : S' --; T' of str . By definition, f is a morphism from F(S) to F(T)
and f' is another one from F(S') to F(T') in C. We define the tensor
product f * g in cstr by the commutative square
<pcs,S')
F(S) F(S') )
F(S * S')
1 f @f'
<pCT,T')
1 f *f' (5.4)
F(T) F(T') )
F(T * T') .
c
Theorem XI.5.3. Equipped with this tensor product str is a strict tensor
c
category. The categories C and str are tensor equivalent.
XI. 6 Exercises
1. Let f be a pre-ordered set, i.e., a set with a binary relation s:; such
that x s:; x, and (x s:; y and y s:; z) =? x s:; z. Set Ob(..1) = f,
Hom(..1) = {(x, y) E f x fix s:; y}, s(x, y) = x, b(x, y) = y, and
(y, z) 0 (x, y) = (x, z). Show that these data define a category ..1.
2. Prove that the class of all categories form a category Cat whose ob-
jects are the categories and whose morphisms are the functors.
3. Prove that the class of functors form a category Funet whose objects
are the functors and whose morphisms are the natural transforma-
tions between functors.
(a)
HOmAlg(k[G], A) ~ HomGl(G, A X)
5. Let I be a set and Vect I the category whose objects are families
(Ui)iEI of vector spaces indexed by I. The set of morphisms in Vece
from (UJi to ell;)i is the product set TIiEI Hom(Ui , Vi). For any vec-
tor space U, we set 6.(U) = (Ui)i where Ui = U for all i E I. Show
that 6. defines a functor from Vect to Vect I and that the direct
sum EB and the direct product II of vector spaces define functors
EB, II : Vect I -+ Vect. Prove that the diagonal functor 6. is right
adjoint to the functor EB and left adjoint to the functor II.
that these constraints satisfy the Pentagon and the Triangle Axioms
if and only if a, A, p satisfy the functional equations
XI. 7 Notes
Tensor categories were introduced in 1963 by Benabou [Ben63]. See also
[Mac63] where constraints as well as the Pentagon and the Triangle Axioms
were defined. Tensor categories are also called monoidal categories in the
literature. Our terminology is taken from Joyal and Street [JS91a] [JS93].
Lemma 2.2 is due to Kelly [Ke164]. For a proof of Mac Lane's coherence
theorem, see [Mac63] [Mac71]. Exercise 8 was taken from [Ke164].
Chapter XII
The Tangle Category
if and only if one may pass from a to b by operations replacing any subword
of the form c by a subword of the form d where (c, d) belongs to R.
Example 1. The abelian group Z2 has a presentation (F, R) where
(1.1)
where the symbols 0 and Q9 in the right-hand sides denote the composition
and the tensor product in the tensor category C respectively. The subwords
of a 0 b and those of a Q9 b consist of the word itself, the subwords of a and
those of b.
The class of words in F is the union of all words of positive rank. We
introduce an equivalence relation on words.
Definition XII.1.1. Two words a and a' in F are equivalent if there exist
words a o = a, a 1, ... ,ak = a' such that for all i, the word ai+l is obtained
from a i by replacing a subword x of one of them by a subword y of the other
where x and yare the two sides of any of the following relations:
([idv1 Q9 [fl] Q9 [idwD 0 ([idvl Q9 [J21 Q9 [idwD 0'" 0 ([idvl @ [fd @ [idwD
~ ([idv1 Q9 ([f l ] 0 [f2 10 0 [Jk-ID Q9 [id w ]) 0 ([id v ] @ [fk1 @ [idwD
~ (([id v ] Q9 ([Jl] 0 [J2] 0 0 [Jk-lD)@ [idw1)0 (([id v ] @ [h]) @ [id w ])
where the words bl ... bk , cl , . . . ,ce are of the form [id v ] 0 [f] 0 [id w ] for
some f E :F. Set S = s(b k ) and T = b(c l ). Then by (1.3) and Part(a) we
get
for some f, l' E:F. The last two equivalences follow from (1.7). D
Composing and tensoring words are operations that are compatible with
the above-defined equivalence relation. Denote by M(F) the class of equiv-
alence classes of words in F. We define a strict tensor category C(F) as
follows. The objects of C(F) are the objects of C whereas M(F) is the
class of morphisms in C(F). The identity, source, and target maps for C(F)
are given by
The composition and the tensor product of words have already been de-
fined.
The map sending a word a to the morphism 0: of C is a strict tensor
functor from C(F) to C. When this functor is an equivalence of categories,
we say that the strict tensor category C is free on the class F. In view of
Proposition XL1.5, this is equivalent to
Observe that the endomorphisms of the unit object 0 are exactly the
tangles without boundaries, i.e., the links in the space R2xlO, 1[. This ob-
servation will be crucial in Sections 4-5.
We now state the main theorem of this section. It involves the "ele-
mentary" tangles defined by (X.5.1-5.5). We shall also use the following
conventions: 1= id(+) and 1= id(_), and XY is short for X Y when X
and Yare elements of the generating set below.
(n n) 0 (I n 1n) 0 (n X n) 0 (n 1 U l) 0 (n u)
= (n n) 0 (n 1 n l) 0 (I U 1n) 0 (U n), (2.3)
(n X n) 0
X+ 0 X_ = X_ 0 X+ =11, (2.4)
(X+ 1) 0 (1 X+) 0 (X+ 1) = (1 X+) 0 (X+ 1) 0 (1 X+), (2.5)
(1 n) 0 (X l) 0 (1 u) =1, (2.6)
The proof of Theorem 2.2 will be given at the end of Section 3. Figures
2.2-2.9 illustrate Relations (2.1-2.8).
J '"
(
)
Figure 2.8. Relation (2.7)
}
! n
Figure 2.9. Relation (2.8)
302 Chapter XII. The Tangle Category
xx xx xx
They differ from the tangles X only by their orientations.
Geometrically, this means that Ii = II and a' = II' are generic tangle dia-
grams and that a = an and a' = an' where we use the notation introduced
in Part (a) of this proof. By assumption, II and II' are isotopic diagrams.
Thus, they can be obtained from each other by a finite sequence of oper-
ations taken from the Transformations (A), (B), (C), and (E) of Lemma
X.5.7. In order to show that the words an and an' are congruent modulo
R, it is therefore enough to check that the above-mentioned transforma-
tions do not change the congruence class of words. Let us verify this case
by case.
(A) If II and II' are generically isotopic, then an = all"
(B) If II and II' differ by a Transformation (B), then an "" an' in view
of Relation (1.8).
(C) If II differs from II' by a Reidemeister Transformation (0), then
an ""n an' thanks to (2.1-2.2).
(E) This case is taken care of by Relations (3.1-3.4). 0
~I
<\
Figure 3.2. Congruence when L = Z
We now deal with Transformation (III): When all strands are oriented
downwards, it follows from Relation (2.5) and its inverse. In the remaining
cases, one proceeds by reducing to the previous case as for Transformation
(II) above. For details, see [Tur94]' 1.4.5.
XII.4 Representations of the Category of Tangles 305
Here we made use of the partial transpose and of the partial trace defined
in 11.3. We shall also use the evaluation maps ev v , ev v * and the coevalu-
ation maps bv , bv * of 11.3 where we identify any finite-dimensional vector
space with its bidual. We are now ready to state the main theorem of this
section.
(4.2.a)
(4.2.b)
c+ c - = c - c+ = 1dVV' (4.3.d)
(c Q9 idv)(id v Q9 c)(c Q9 id v ) = (id v Q9 c)(c Q9 idv)(id v Q9 c), (4.3.e)
(id v Q9 d')(c Q9 idw)(id v Q9 b) = id v , (4.3.f)
gh = idvw, and hg = idwv (4.3.g)
where the linear maps 9 : W Q9 V ---+ V Q9 Wand h : V Q9 W ---+ W Q9 V are
defined by
(4.3.h)
and
(4.3.i)
The data (V, W, b, b', d, d', c, c-) where V, Ware finite-dimensional vector
spaces and b, b', d, d', c, c- are linear maps satisfying Relations (4.3.a-i) will
be called a representation data for the tangle category T.
Conversely, by Proposition 1.4 and Theorem 2.2, any representation data
(V, W, b, b', d, d', c, c-) for T gives rise to a unique tensor functor F : T ---+ V
such that F((+)) = V, F((-)) = W, and such that Relations (4.2.c-e)
hold.
These considerations imply that Theorem 4.2 is a consequence of the
following proposition. 0
and c- = c- 1 . Then (V, V*, b, b', d, d', c, c-) is a representation data for T.
There is a converse statement whose formulation and proof are left to
the reader. Before we prove Proposition 4.3, we give a corollary to Theorem
4.2, and we state two lemmas which will be used in the proof of Proposition
4.3 (they may also be used to establish the converse statement).
Let (c, JL) be an enhanced R-matrix on a finite-dimensional vector space
V and F be the unique strict tensor functor from T to V satisfying Relations
(4.2.a-b). Let u be a braid with n strands. Since u is a tangle, we may
evaluate F On u. We get an automorphism F(u) of Vn. Similarly, F can
308 Chapter XII. The Tangle Category
be evaluated On the link a- which is the closure of the braid (J (see X.8).
We get an endomorphism F(a-) of the ground field k, i.e., a scalar. In the
following corollary, we express the automorphism F((J) and the scalar F(a-)
in terms of the representation p~ of the braid group Bn associated to the
R-matrix c in Corollary X.6.9.
Corollary XII.4.4. With the previous notation, we have
F((J) = p~((J) and F(a-) = tr(fLem 0 p~((J)) (4.4)
for any braid (J of Bn'
PROOF. (a) It suffices to prove the first statement for the generators (J l' ... ,
(In-l of Bn" Using the notation of X.6, we have
(Ji = 1i - 1 X+ 1n - i- 1
in the tangle category. Applying the tensor functor F, we get by X.6.2 and
by (4.2.a)
F((Ji) = id V0 (i-l) c id v0 (n-i-l) = p~((J).
(b) We first express the closure a- in the tangle category. We have
and
Un = (l ... 1 U r ... no ... 0 ... (l uno u.
Therefore
F(a-) = F(n n) 0 (F((J) idV*0n) 0 F(UrJ.
Now, it is easy to check that (4.2.a-b) imply that
Lemma XII.4.5. Under the previous hypotheses and with the previous
notation, we have
(didw)(id w b) = id w ~ DB = 1, (4.5.b)
(d' idv)(id v b') = id v ~ D' B' = 1, (4.5.c)
(id w d')(b' id w ) = id w ~ B'D' =1 (4.5.d)
where 1 represents various matrix units.
Proof of Proposition 4.3. Let (c, p,) be an enhanced R-matrix on the finite-
dimensional vector space V. In order to apply Lemmas 4.5-4.6, we set
W = V*, a = id v , and (3 = (p,*)-1. Pick a basis {v 1, ... , v m } of V along
with its dual basis {v 1 , ... ,vm}. Define a matrix B' by
Since p, and (3 are isomorphisms, the matrix B' is invertible. Let D' be its
inverse. By definition of b, b', d, d' we have
which is equivalent to
T j*T = T(p,* C!9 p,*)-1 j* (p,* C!9 p,*)T.
1
[((3C!9id v )Tv,v* (CTv,v) x (id v C!9(3-1 ) [(idC!9a*) (Tv,vc'f) x Tv*,v (a -1 C!9id V* ) * 1
is equal to idv*v. Replacing a and (3 by their values, we are reduced to
proving
((p,*)-1 C!9 id v )TV,v * (CTV,v) X (id v C!9 p,*)(TV,vC'f)XTv*,v = idv*v. (4.7)
Relation (4.7) is equivalent to
(id v C!9 (p,*)-1) (CTV,v )X (id v C!9 p,*)(TV,vC'f) X = idvv*. (4.8)
Taking transposes and using Lemma II.3.3, we see that (4.8) is equivalent
to (4.1.c). D
XII.5 Existence Proof for Jones-Conway Polynomial 311
A;;,2 tr2 (crrJid (>9 fLm)) - A;;,I(q - q-l) tr 2 (id (>9 fLm)
(5.4)
PROOF. We apply Theorem 4.2 to the pair (cm,/-lm) of Lemma 5.1. Rela-
tions (4.2.a) imply the desired forms for Fm,q(U) and for Fm,/U). Relation
(5.1) translates immediately to
For the trivial knot, we observe that it is the closure of the trivial braid in
B 1 . We may then appeal to Corollary 4.4, which yields Fm,q(O) = tr(/-lm)'
We conclude with (5.2). 0
XII.6 Exercises
1. Consider the strict tensor category whose objects are the nonnegative
integers and whose morphisms are the isotopy classes of all braid
diagrams in R x [0, 1]. Show that it is generated by the morphisms
X+, X_ and the relations X+ 0 X_ = X_ 0 X+ = id.
2. Let c E Aut(Vl VI) be an R-matrix as in VIII.2, Example 2. Find
all automorphisms J.L of VI such that (c, J.L) is an enhanced R- matrix.
3. Compute the trace ofthe automorphism (J.Lm J.Lm)cm where (c m , J.L m )
is the enhanced R-matrix of Lemma 5.1. Deduce the value of the
functor F m,q of Proposition 5.2 on the trefoil knot and on the Hopf
link (Hint: use Corollary 4.4 and (5.1)).
XII.7 Notes
The results of this chapter are essentially due to Turaev [Tur89] whose
exposition we followed closely, and to Yetter [Yet88]. Enhanced R-matrices
already appear in [Tur88], though in a slightly different form.
In XIV.5.1 we shall build a strict tensor category Rout offramed tangles
or ribbons (defined in X.8). A presentation ofR by generators and relations
is given in [FY89] [Tur89].
Chapter XIII
Braidings
c: Q9 -> Q9T.
This means that, for any couple (V, W) of objects of the category, we have
an isomorphism
cv,w : V Q9 W -> W Q9 V (1.1)
XIII.l Braided Tensor Categories 315
(1.2)
/ au,v,w ~ av,w,u
(1.5)
and
(1.6)
Let us investigate the relationship of the braiding with the unit con-
straints.
316 Chapter XIII. Braidings
and
(1. 7)
(U 129 V) 129 W
,,/ cu,vidw "" au, v, w
(V 129 U) 129 W UQ9 (V 129 W)
1 av,u,w liducv,w
But this is a special case of the commutative square (1.2) (expressing the
functoriality of the braiding) where f is replaced by Cuy and g by id w
o
We give a few examples of braidings.
318 Chapter XIII. Braidings
Proposition XIII.1.4. Let (H, IL, "1, Ll, c) be a bialgebra. The tensor cat-
egory H -Mod is braided if and only if the bialgebra H is braided.
PROOF. Let (H, IL, "1, Ll, c, R) be a braided bialgebra with universal R-
matrix R. In VIlL3 we defined isomorphisms c{J w from V W to W V
by ,
c~,w(v w) = TV,W (R(v w))
where v E V and w E W. Proposition VIlL3.1 implies that the family c is
a braiding.
Conversely, let (H, IL, "1, Ll, c) be a bialgebra. Suppose that there exists a
braiding c in the tensor category H-Mod. Define an invertible element R
in HH by
(1.9)
Let us show that R is a universal R-matrix for H.
If v, ware elements of H-modules V, W, the naturality of the braiding
implies the commutativity of the square
HH ~ HH
lV0
w 1 w0v
VW ~ WV
where ii : H ----; V and w : H ----; Ware the H-linear maps defined by
ii(1) = v and w(1) = w. This implies that
LlOP(a)R = RLl(a)
for all a E H.
XIII.1 Braided Tensor Categories 319
R12R13R23 = R23 R 13 R 12
of Theorem VIII. 2.4 (a) .
(1.13)
for all objects V, W in the category. If (1.13) holds, we call the braiding c a
symmetry for the category. Notice that the commutativity of the hexagon
(HI) and the commutativity of the hexagon (H2) are equivalent in a sym-
metric tensor category.
We give two examples of symmetric tensor categories.
where 0'1' ... ,0'm+n-l are the generators of Bm+n defined in X.6. The braid
cn m is represented in Figure 2.1. Observe that the permutation of the braid
cn :m is the permutation sn,m of Proposition 1.6.
n m
PROOF. We have to prove that the family (c n m)n m>O is functorial with
respect to all morphisms in B and satisfies Rel~tio~s [1.5) and (1.6).
Let us start with the functoriality. Since any morphism in B is an element
of a braid group, it is enough to check the functoriality with respect to the
generators O'i' More precisely, we must prove that for all i, j such that
1 :::; i :::; n - 1 and 1 :::; j :::; m - 1 we have
Both sides of this relation are represented by the braid diagrams of Figure
2.2. It is clear that one can pass from one braid diagram to the other
by repeated applications of the Reidemeister Transformation (III), which
322 Chapter XIII. Braidings
proves the equality. The reader may replace this topological proof by an
algebraic one using the braid relations of Lemma X.6.4.
Figure 2.2.
v>
0
,~
n m p n m p
n m p n m p
(V0 V) 0 V
/0"0 id v "" av,v,v
(V 0 V) 0 V V0 (V0 V)
1 av,v,v l idv0 0"
V0(V0V) V 0 (V0V)
l idv0 0" 1 av,'v,v
V0 (V0 V) (V0V)0V
1 av,'v,v 10"0 id v
(V 0 V) 0 V (V0V) 0V
10"0 idv 1 av,v,v
(V0V) 0 V V0 (V0 V)
~ av,V,V / idv00"
V0(V0V)
commutes.
The commutativity of this dodecagon is equivalent to
and
v = a(c/ 0 id)a- 1(id 0 (J1)a((J1 0 id),
we have to prove that u = v. Now, by Definition X1.4.1 we have
(3.3)
where IP2 = IP2(V, V). Now IP2 is a natural isomorphism. Therefore the
squares
'P2(VV,v)
F(V 0 V) 0 F(V) F((V0V)0V)
1
)
F(u)idp(v) IF(Uid v )
'P2(VV.V)
F(V 0 V) 0 F(V) ) F((V0V)0V)
and
'P2(V,vV)
F(V) 0 F(V 0 V) F(V0(V0V))
1 1
)
U = (id0IP21)IP21F(id0(J)F(a)F((J0id)
F(a- 1)F(id 0 (J)F(a)IP2(IP2 0 id)
(id 0 IP21 )IP2 1F ((id 0 (J )a( (J 0 id)a- 1(id 0 (J)a) IP2( IP2 0 id).
Similarly, we have
The equality u = v results from the fact that (J satisfies Relation (3.1). 0
VV
1
1@1
V'V'
(3.5)
(3.6)
(3.7)
(ry(l) ry(1))'P21F(c1,l)'P2
'P; -lry(2)F(C 1,1)'P2
'P; -1 F'(C 1,1)ry(2)'P2
'P; -IF'(C 1,1)'P;(ry(1) ry(l))
a'(ry(l) ry(l)).
The first and last equalities follow by definition of a and a', the second
and fourth ones by definition of a natural tensor transformation (Definition
XI.4.1), and the third one by Definition XI.1.3.
We can state the first universality property of B.
Theorem XIII.3.3. For any tensor category C the functor, defined above,
8: Tens(B, C) ----> YB(C) is an equivalence of categories.
C. We wish to construct a natural tensor transformation 7]j from (F, i.po, i.p2)
to (F', i.p~, i.p~) such that 7]j(1) = f. We proceed as follows. If n = 0,1, we
set 7]j(O) = i.p~i.po1 and 7]j(l) = f. If n > 1 we define 7]j(n) inductively by
(3.8)
(3.9)
for any integer n ~ 0 and any element g of the braid group En' that
7] j (0 )i.po = i.p~ (this holds by definition of 7] j (0)), and that for all n, m ~ 0
The full faithfulness of 8 follows from 7]j(1) = f and from 7]T)(1) = 7].
The first relation holds by definition. Let us check the second one. We shall
prove
(3.11)
by induction on n. This is clear for n = 0,1. If n > 1, we use the fact that
7]and 7]T)(1) are natural tensor transformations to write
7]T)(l)((n - 1) 1)
i.p~(7]T)(1)(n-1) 7]1)(1)(1))i.p2
i.p; (7](n - 1) 7](1)) i.p2
7]((n - 1) 1)
7]( n).
PROOF. If such a functor F exists, then (2.2) implies F(n) = V<8I n and
F( ai ) -- F('d<8l(i-l)
1 I
,0,
'6' CI,1
,0,
'6'
'd<8l(n-i-I)) _ 'd<8l(i-I),o,
1 I-I V '6' a
,0,
'6'
'd<8l(n-i-l)
1 V
for 1 :::; i :::; n - 1. This proves the uniqueness of F in view of the fact that
aI' a z ,"" a n - l generate Bn as a group.
Let us prove the existence of F. Set F(n) = V<8I n . Define automorphisms
cl ,, cn - 1 of F(n) by
ci -_ 1'd<8l(i-l),o,
'6' a
,0,
'6'
'd<8l(n-i-l)
1
when 1 :::; i :::; n-1. The automorphism a being a Yang-Baxter operator, the
automorphisms ci satisfy the braid group relations (X.6.1-6.2). It follows
from Theorem X.6.5 that there exists a unique morphism of groups F from
the braid group Bn to Aut(F(n)) such that F(a i ) = ci for all i. The functor
F is a strict tensor functor from l3 into C and we have F(CI,I) = c i = a.
o
Definition XIII.3.6. A tensor functor (F, 'Po, 'P2) from a braided tensor
category C to a braided tensor category V is braided if, for any pair (V, V')
of objects of C, the square
We denote by Br(C, V) the category whose objects are the braided ten-
sor functors from C to V and whose morphisms are the natural tensor
transformations.
Theorem XIII.3.7. For any braided tensor category C the functor 8' :
Br(l3, C) --t C defined by 8' (F) = F(l) is an equivalence of categories.
It is clear that the left-hand side sits inside the right-hand side. We have
to prove the opposite inclusion. Let f : F(l) = V --> F'(l) = V' be an
element of Homd8'(F), 8'(F')). We wish to prove that f is a morphism
in the category YB(C), which means that the square (3.5) has to commute
with (f = 'P;-1 F(Cl,l)'P2 and (f' = 'P; -1 F'(C 1,1)'P;. We have
The second and fourth equalities follow from (3.12) whereas the third one
follows from the naturality of the braiding c in C.
Now 8' is fully faithful in view of the isomorphisms
where the first one follows by definition, the second one from the full faith-
fulness of 8 (Theorem 3.3), and the last one has just been proved.
Essential surjectivity of 8'. Let V be an object of C. Since C is braided, the
automorphism Cv v is a Yang-Baxter operator by Theorem 1.3. According
to Theorem 3.3, the functor 8 is essentially surjective, which means that
there exists a tensor functor (F, 'Po, 'P2) : B --> C along with an isomorphism
ex: V --> F(l) such that
(3.13)
The commutativity of (Co,o), (C1,0) and (CO,I) is left to the reader. Let
us first check that (C 1,1) commutes. We have
'P2 (id 12> F( Cn,I)) 'P2 1'P2 (F( cn,m) 12> id) 'P2 1'P2
'P2(id 12> 'P2)(id 12> c P(n),P(I)(id 12> 'P2 1)'P2 1
'P2 ('P2 12> id) (cP(n),P(m) 12> id)( 'P2 1 12> id)'P2 1 'P2
'P2(id 12> 'P2)(id 12> c P(n),P(I) a (cP(n),P(m) 12> id)
('P2 1 12> id)'P2 1 'P2
'P2(id 12> 'P2)CP(n),P(m)P(I) a ('P2 1 12> id)'P2 1 'P2
'P2 CP(n),P(m+l)(id 12> 'P2) a ('P2 1 12> id)'P2 1 'P2
'P2 cP(n),P(m+l) .
cx,v : X V ---- V X
defined for all objects X in C such that for all objects X, Y in C we have
(4.1)
(4.2)
XV ~
lfl8i i d v (4.3)
YV ~
(4.4)
XIII.4 The Centre Construction 331
Cv,W : (V, C,v) 129 (W, c_,W) ----7 (W, C,W) 129 (V, C_,v).
The first and fourth equalities follow from (4.4), the second one from (4.1),
and the third one by (XI.2.3), i.e., by the naturality of the tensor product.
2. Let f: (V,c_,v) ----7 (W,c_,w) and l' : (V',C_,vI) ----7 (W',c,w') be
morphisms in Z(C). We claim that so is f 129 1'. Let us check Relation (4.2)
for f 129 1'. We have
The first and fifth equalities follow from (4.4) and from (XI.2.3), the second
and fourth ones from (4.2), and the third one from (XI.2.1).
Now it is clear that the tensor product is well-defined on the objects and
on the morphisms of Z(C). It is functorial and satisfies all the required
axioms because it already does so in the original category C. Thus, the
category Z(C) is a strict tensor category. We next show that it is braided.
3. Let us start by proving that cv,w is a morphism in Z(C). We have to
check Relation (4.2) for cv,w' namely
332 Chapter XIII. Braidings
PROOF. Let us first prove the existence of Z(F). For any object V of C we
set
Z(F)(V) = (F(V), C-,F(V))
where C,F(V) is defined for all objects X in C' by cX,F(V) = F(C F -l(X),v)'
Here C v is the braiding in C. Relation (4.1) is satisfied because F is a
tensor f~nctor. Therefore Z(F)(V) is an object in Z(C').
If f : V --+ V' is a morphism in C, set Z(F)(f) = F(f). Relation (4.2)
is satisfied because of the naturality of the braiding in C. This proves that
Z(F) is a functor. Clearly, IT 0 Z(F) = F. Let us now check that Z(F) is
a braided tensor functor. It preserves tensor products because of (1.5) and
(4.4). It also respects braidings. Indeed, we have
Corollary XIII.4.4. For any strict braided tensor category C there exists
a unique braided tensor functor Z from C to Z (C) such that II 0 Z = ide.
~v (v) = L Vv vA E V A (5.1)
(v)
AV ~ VA
l~~v l~v~
XV ~ VX.
334 Chapter XIII. Braidings
L Vv Q9 (vA)'x Q9 (vA)"y
(v)
L (vv ) v Q9 (vv ) A X Q9 v A y.
(v)
Cx,v(X Q9 v) = L Vv Q9 v AX
(v)
L ai . v Q9 aix
(v),i
Tx,v(R(XQ9v)).
o
We prove Theorem 5.1 in five steps.
1. We first define a functor F from Z(A-Mod) to D(A)-Mod. Let (V, C_ v)
be an object of Z(A-Mod). By Lemma 5.2 and Theorem IX.5.2, the vector
space F(V, c_ v) = V is a left D(A)-module. Recall from IX.5 that the
action of D(A) on V is determined by
CX0)y,v(x c>9 y c>9 v) = (cx,v c>9 id y ) ((id x c>9 cy,v )(x c>9 y c>9 v)).
The left-hand side is equal to
XIII. 6 Exercises
1. Let H be a braided bialgebra with universal R-matrix R. Show that
the category H-Mod is symmetric if and only if TH,H(R) = R- 1 .
2. Let C be a strict tensor category. Show that one gets a definition for
a braiding equivalent to Definition 1.1 if one replaces the hexagons
(HI) and H(2) by the square
cU,vid x0y
VUXY
1 idv0UCX, Y
VUYX
for all objects U, V, X, Y.
3. Resume the notation of Exercise XI.S. Define a commutativity con-
straint c by c(vw) = "((n,p)(wv) where v and w are homogeneous
vectors of respective degrees nand p, and where "( is a function with
values in k \ {o}. Show that c is a braiding if and only if the functional
equations
"((m, n + p) = a(n,p, m)-l"((m,p)a(n, m,p),,((m, n)a(m, n,p)-l
and
"((m + n,p) = a(p, m, n)"((m,p)a(m,p, n)-l"((n,p)a(m, n,p)
are satisfied for all integers m, n, p.
c
4. Given a tensor category C, define the reverse category rev as the
category C with tensor product given by V rev W = W V. Prove
that, if C is braided with braiding c, then (id, 'Po = id, 'P2 = c) is a
tensor functor between C and crev .
5. Let C be a braided tensor category. Show that one can equip the
c
strict tensor category str of XI. 5 with a braiding such that the tensor
equivalences constructed in XI.5 between both categories are braided
functors.
338 Chapter XIII. Braidings
6. (Presentation of the braid category) Show that the strict tensor cat-
egory B is generated by the morphisms 0"1' 0"1 1 of B2 and by the
relations 0"10"1 1 = 0"1 1 0"1 = id 2 and
XIII. 7 Notes
Braided tensor categories were introduced by Joyal and Street [JS91a]
[JS93]. They generalize the concept of a symmetric tensor category which
appeared in the 1960's in the work of Benabou [Ben64] and Mac Lane
[Mac63], and was extensively studied in relation to algebraic geometry and
algebraic topology (see, e.g., [DeI90] [DMS2] [KLSO] [Mac63] [SR72]).
The content of Section 3 is taken from [JS93]. Lemma 3.5 is the analogue
of Theorem XII.4.2 for braids. We found the example of crossed G-sets in
[FYS9]. Exercise 4 is from [JS93].
The centre construction of Section 4 is due to Drinfeld (unpublished), to
Joyal and Street [JS91c], and to Majid [Maj91b].
Chapter XIV
Duality in Tensor Categories
~
~
Figure 1.4. A morphism f : Ul ... Urn --+ Vl ... Vn
~
' 10gV' ~
U V
cf?'~'
1
U
9
V
Figure 1.5.
is in Figure 1.6.
XIV.1 Representing Morphisms in a Tensor Category 341
~ 1 tb
CPl
Figure 1.6. The identity (XI.2.3)
This leads to the following "partial isotopy principle": for any figure pre-
senting a morphism of C, the part of the figure lying to the left (or to the
right) of a vertical line may be pushed up or down without changing the
corresponding morphism in C. We shall use this principle frequently and
without any further explanation in the sequel.
Assume now that the tensor category is braided with a braiding c. For
any pair (V, W) of objects in C we represent cv,w and its inverse cv,\.v
respectively by the pictures in Figure 1.7.
V
x x W
cv,W
W
-1
cv,W
V
V W V W W V W V
Figure 1.8. Invertibility of cv, w
~ VU VU
X
>(:~
~~~ ~I
t v
l~
y~
~X
U V W UV W UV W U V W
Figure 1.11. Proof of Theorem XIII. 1.3
XIV.2 Duali ty
We now abstract the notion of duality introduced in I1.3.
f'\v
dv
Figure 2.2. The morphisms bv and d v
The above data are enough to extend duality to a functor and to derive
adjunction formulas of the type proved in Chapter II.
Let us first define the transpose 1* : V* -+ U* of a morphism f : U -+ V
in C by
1* = (d v Q9 idu*)(id v * Q9 f Q9 idu*)(id v * Q9 bu ) (2.2)
With our graphical conventions we can represent the transpose 1* of a
morphism f : U -+ V as in Figure 2.4.
Relations (2.1) imply that (JU)D = f and (l)U = g. A similar proof works
for the other adjunction formula. We invite the reader to give a graphical
proof.
(c) We define a morphism Av,w : W* Q9 V* ---+ (V Q9 W)* by
VQ9W WV
A
" "
WV VQ9W
Av,w A-I
v,w
Figure 2.5. The morphisms AV,w and AV,\v
wv
" "
wv
wv wv
" " " "
id",-~
lvtZlw
I id v0;l
wv
W v W V
.. ProojojA-V,W
Figure 26 1
0 AV,W = id W*0V*
346 Chapter XIV. Duality in Tensor Categories
VW
"
VW VW
" "
'T~ v v
! idvw!
VW VW
W9W
"
Note that Figures 1.6 and 2.3 are used in these graphical proofs. 0
b~ :I -+ * V V and d~: V *V -+ I
XIV.2 Duality 347
Right duality has properties analogous to the ones stated for left duality
in Proposition 2.2. We leave their formulation to the reader. In particular,
right duality implies that the functor V 0 - [resp. the functor - 0 *V] is
left adjoint to the functor *V 0 - [resp. to the functor - 0 V].
In general, right duality is different from left duality unless we add extra
hypotheses on C. Nevertheless, it may happen that C is autonomous, i.e.,
it has left and right duality. In this case, there are isomorphisms
*(V*) ~ V ~ (*V)*
for any object V. We refer to [JS93] for a proof. Hint: the first isomorphism
is a consequence of the following natural isomorphisms
the first one being implied by the right duality and the second one by the
left duality.
Example 1. Let A be a Hopf algebra with antipode S. The category A-
Mod f of left A-modules that are finite-dimensional over the ground field k
is a tensor subcategory of A-Mod. For any left A-module, endow the dual
vector space V* = Hom(V, k) with the A-action given by
bv (1) = 2:= Vi 0vi and dV(v i 0Vj) =< vi,v j > (2.8)
where {vJ i is any basis of V and {Vi L is the dual basis in V*. The map bv
is the coevaluation map and the map d v is the evaluation map of II.3. It was
proved in Proposition III.5.3 that bv and d v are A-linear. By Proposition
II.3.1 they satisfy Relations (2.1), endowing A-Mod f with the structure of
a tensor category with left duality.
Suppose, furthermore, that the antipode S is invertible. For any left A-
module V, denote by *V the same dual vector space now equipped with
the left A-action given for all a E A, v E V and all linear forms j on V by
(2.10)
using the same conventions as above. One checks that b~ and d~ are A-
linear and that they satisfy Relations (2.5), endowing A-Mod j also with
the structure of a tensor category with right duality. In other words, the
category A-Mod j is autonomous when A is a Hopf algebra with invertible
antipode.
CV*,w
V \ w W. (;'J" w
Figure 3.3. Proof of Proposition 3.1
~
w -
V
The first equality follows from (2.1), the second one from Figure 1.8, and
the last one from the naturality of the braiding c_,w.
We go one step further by introducing the concept of a ribbon category.
Definition XIV.3.2. Let (C, , I) be a strict braided tensor category with
left duality.
(a) A twist is a family ()v : V -t V of natural isomorphisms indexed by
the objects V of C such that
()V0W = (()v ()w)cw,vcv,w (3.1)
and
(3.2)
for all objects V, W in C.
(b) A ribbon category is a strict braided tensor category with left duality
and with a twist.
The naturality of the twist means that for any morphism f : V - t W
we have ()wf = j()v. Using the graphical conventions of Sections 1-2,
Relations (3.1-3.2) may be represented as in Figures 3.4 and 3.5.
VW V W
V*
Figure 3.5. Relation (3.2)
350 Chapter XIV. Duality in Tensor Categories
PROOF. (a) See Figure 3.6. All equalities follow by naturality of the braiding
and of the twist.
~~
1
(5
1
IB~
l~ )
x
~~
X '~ X
...:..
V W V W V W
><
~J
~\
...:.. ...:..
~ ...
V W
Figure 3.6. Proof of Lemma 3.3 (a)
(3.4)
and
d~ = dvcv,v* (e v id v .). (3.5)
We shall agree to represent b;" and d;" as in Figure 3.7.
V~U
b'v d'v
Figure 3.7. The morphisms b~ and d~
Let us prove that the morphisms b~ and d;" equip C with the structure
of a category with right duality, where *V = V*. Before we give a precise
statement, we shall prove the following technical lemma.
Lemma XIV.3.4. For any object V of a ribbon category, we have
e~? v,\;
(d v idv)(id v * c )(cv,v.b v id v )
(dvcv,v* idv)(id v cv,vb v )
(id v dv cv,v * ) (c v,\; id v * )(id v bv )'
V V V
Figure 3.8. The equalities of Lemma 3.4
It is clear from the pictures that the naturality of the braiding implies the
last two equalities. So it is enough to prove the first one.
By naturality of the twist and by Lemma 3.3 (b), we get
(3.6)
Let us denote by f the second term of the equalities in Lemma 3.4. Figure
3.9 shows that the right-hand side of (3.6) is equal to (e~ f id v * )b v
352 Chapter XIV. Duality in Tensor Categories
...:...
c
X C!!
~
...:...
evv* bv
I~Srl
rn
v
Therefore, we have f = e 2 , as desired.
As for Figure 3.9, the first equality follows from (3.1), the second one
from Proposition 3.1 and from (3.2), the third one from the naturality of
the tensor product, the fourth one from (2.1), and the last one from the
naturality of the braiding. 0
and
(3.8)
XIV.3 Ribbon Categories 353
PROOF. (a) By (3.4-3.5) and the naturality of the tensor product, we have
The first equality is by definition, the second one by (2.1), the third one
by naturality of the braiding, the fourth one by Lemma 3.4, the fifth one
follows from Figure 1.8, and the last one from (2.1). 0
V*
L~
Figure 3.11. Isomorphisms between V and V* *
Definition XIV.4.1. Let C be a ribbon category with unit I. For any ob-
ject V of C and any endomorphism f of V, we define the quantum trace
trq(f) of f as the element
I~V@V* evfid~V@V*~V*@V~I.
This notion coincides with the usual trace when C is the category Vect f (k)
(see Proposition II.3.5). Graphical representations of trq(f) are given in
Figure 4.1.
v
Figure 4.2. The proof oftrq(fg) = trq(gf)
The first equality is by definition, the second one by (2.1), the third
one by naturality of the braiding, the fourth one by (2.1), the fifth one by
naturality of the braiding, the sixth one by naturality of the twist, and the
last one by definition.
(b) We know from Proposition XI.2.4 that the composition in End(I)
coincides with the tensor product. Therefore, it is equivalent to prove that
trq(f 0 g) = trq(f) 0 trq(g). The proof of the latter is in Figure 4.3. The
first and last equalities in that diagram on the next page follow from the
definition, the second one from (3.1), the third one from (2.1), the fourth,
sixth, and seventh ones by naturality of the braiding.
(c) The proof of tr q(f) = tr q(f*) is in Figure 4.4 two pages on.
356 Chapter XIV. Duality in Tensor Categories
~~~~
SVwV
~ c@
V
c@
W
Figure 4.3. The proof of trq(f 0 g) = trq(f) 0 trq(g)
XIVA Quantum Trace and Dimension 357
I@
V*
The first equality in the diagram on the previous page follows by defini-
tion, the second one by (3.2) and by Proposition 3.1, the third one by (3.8),
the fourth one by naturality of the braiding, the fifth and the sixth ones
by (2.1), the seventh one by definition of d'v, the eighth one by naturality
of the twist, the ninth one by Lemma 3.4. 0
Definition XIV.4.3. Let C be a ribbon category with unit I. For any ob-
ject V of C we define the quantum dimension dimq(V) as the element
~W7 A
V /A~
Figure 5.1. The framed tangles be and de
It is easy to check that the maps be and de satisfy Relation (2.1), thus
equipping R with the structure of a strict braided tensor category with left
duality. Observe that the transpose L * of a ribbon L is isotopic to the ribbon
obtained by rotating L through an angle 7r around an axis perpendicular
to the plane of projection.
We define a twist on R as follows: 8(+) is the left ribbon of Figure X.8.1
oriented downwards (also represented by the left tangle diagram of Figure
X.8.2). The right ribbon of Figure X.8.1 oriented downwards defines the
inverse of 8(+) (it is represented by the right tangle diagram of Figure
X.8.2). To define the twist for an arbitrary object, we use Relations (3.1-
3.2). Check that, if c is of length n, then 8e is obtained by twisting by an
angle of 27r the plane containing n vertical fiat ribbons.
Quantum trace and quantum dimension are defined in the ribbon cat-
egory R by the formulas of Section 4. One can check using Reidemeister
Transformations (1') and (II) that if L is a ribbon with s(L) = b(L), then
its quantum trace trq(L) is the closure of L drawn in Figure 5.2. Quantum
dimensions are trivial links with the framing pointing to the reader.
The category R has two universal properties similar to the ones given for
the category B in XIII.3. For the first one which corresponds to Theorem
XII.4.2, we refer to PS91c] [Tur89]. We state the second one paralleling
Corollary XIII.3.8.
360 Chapter XIV. Duality in Tensor Categories
and
du2Jn (gl"'" gn' h 1,, h n ) = I5 g, . h, .. . l5 gn ,h n
Relations (2.1) are satisfied. A twist Bc((n is defined inductively on n by
(3.1) and its initial value Bc = id z1cJ ' One checks that the quantum di-
mension of an object is its cardinality:
Apply Theorem 5.1 to the ribbon category Z[Xa(G)] and to the object
G. We get an endomorphism Fa(L) of the unit object {I}, i.e., an integer
for any (framed) link L. We invite the reader to use the algorithm described
above to compute this isotopy invariant for a few simple links. For instance,
Fa (L) is equal to the number of couples (g1' g2) E G x G with g1g2 = g2g1 if
L is the Hopflink, and with glg2g)"1 = g:;l glg2 if L is the trefoil knot. Freyd
and Yetter [FY89] proved for a general link L that Fa(L) is the number of
group homomorphisms from the fundamental group of L to G.
(6.1)
(6.2)
(6.4)
PROOF. (a) Let D be a ribbon algebra with the distinguished central ele-
ment O. Braiding and duality in D-Mod f are given as in XIlL1.3 and XlV.2,
362 Chapter XIV. Duality in Tensor Categories
(0- 1 0-1)(R21R)(vw)
~(O-l)(V w)
0vw(v w).
L S(Sj)S(tk)Sktj V
j,k
L S(Sj)utj v
j
L S(Sj)S2(tj)UV
j
J
S(U)UV = uS(U) V.
D
d'v(vQ9Oo) = L < tiOo, Sie-Iv > = L < a, S(ti)Sie-1v > = < a, ue-1v > .
Therefore,
shows that the element u).. corresponding to the universal R-matrix R).. is
independent of the parameter A and is given by u).. = x = S(u)..). Therefore
u)..S(u)..) = x 2 = l. One checks that H is a ribbon algebra with 8 = l.
Example 2. This is due to Reshetikhin and Turaev [RT91]. It deals with the
Hopf algebra U q' i.e., the finite-dimensional quotient of U/5 [( 2)) considered
in VI. 5 for q a root of unity. In IX.6-7 we proved that U q was a braided Hopf
algebra and we computed its universal R-matrix. We resume the notation
and conventions of Chapter IX. In particular, we assume that q is a root
of unity of odd order d > l.
PROOF. The centrality of 8 follows from (6.4) and from the fact that we
also have
S2(X) = KxK- 1
for all x E U q. It is immediate to check that s( 8) = l. As for ~ (8), we have
qn+l _ q-n-l
dim (V)=qn+ qn-2+ ... + q-n+2+ q -n= =[n+1].
q n q _ q-l
XIV.7 Exercises 365
XIV.7 Exercises
l. For any braided Hopf algebra D, define an algebra D(8) as the quo-
tient of the polynomial algebra D[8] by the two-sided ideal generated
by 8 2 - uS(u). Show that D(8) has a unique Hopf algebra structure
such that the natural inclusion of D in D(8) is a Hopf algebra map
and that ~(8) = (R21R)-1(8 8), c(8) = 1, and S(8) = 8. Prove
that D(8) is a ribbon algebra.
2. Under the hypotheses of the previous exercise, show that the category
of left D(8)-modules is equivalent to the category whose objects are
pairs (V, 8v ) where V is a left D-module and 8v is a D-linear auto-
morphism of V such that for all v in V we have 8V2 (v) = uS(u) v,
and whose morphisms (V, 8v ) ---t (W, 8w ) are the D-linear maps f
from V to W such that f8 v = 8w f.
3. Using the definitions and the notation of 5.2, compute 8c @n for n > l.
4. Given a finite abelian group A and a commutative ring K, let K(A) be
the commutative K-algebra of K-valued functions on A. It has a basis
{eo} aEA over K such that the multiplication is given by eoe b = 8a,b
for all a, b E A.
(a) Show that there is a unique Hopf algebra structure on K(A) such
that for all a E A we have
XIV.8 Notes
The graphical calculus described in Section 1 was advocated in many pa-
pers, e.g., [FY89] [FY92] [JS91a] [Kau91] [RT90] [RT91].
The concept of duality in a tensor category appeared in the classical
references quoted in Chapter XIII. The examples presented in this book
require distinguishing carefully between left and right duality. In Section 2
we followed Joyal and Street's treatment of duality as proposed in [JS93]
(see also [FY89]). There Joyal and Street also introduced the concept of
a twist in a strict braided tensor category and the concept of a ribbon
category. Actually, they called the latter tortile tensor categories. The name
used here was coined by Turaev [Tur92].
Definition 4.1 is due to Turaev [Tur92] generalizing previous definitions
of [KL80] and [FY89]. We devised a proof of Theorem 4.2 highlighting the
power of the graphical calculus of Section 1 (a different proof can be found
in [Tur94]).
Ribbon algebras were invented by Reshetikhin and Turaev [RT90] who
also showed that the quantum groups of Drinfeld and Jimbo gave birth to
ribbon algebras.
The construction of the ribbon algebra D(B) of Exercise 1 is taken from
[RT90]. Exercise 4 is due to Turaev: this example does not produce any
interesting isotopy invariant. Exercise 5 is from [JS91b]. Exercise 6 is due
to the author.
There exists an elaboration of the centre construction of XIII.4, to be
found in [KT92], which assigns to any strict tensor category C with left
duality a ribbon category D(C). It is related to the quantum double of
a finite-dimensional Hopf algebra A with in-.'ertible antipode and to the
construction of Exercise 1 by the equivalence of ribbon categories
category C', there is a natural bijection between the set of strict braided
tensor functors preserving duality and twist from R( C) to C' and the set of
strict tensor functors from C to C'. In particular, if C is a ribbon category,
then the identity functor of C corresponds to a functor Fe : R( C) -> C
preserving tensor products, braidings, duality and twists. For more details
on the category R(C), see [FY89] [JS93] [RT90] [RT91] [Tur92] [Tur94].
The existence of the functor Fe allows one to find isotopy invariants for
framed links with values in the endomorphism monoid of the unit object
of the ribbon category C. Proceed as at the end of 5.1. The main difference
is that we are now permitted to colour the connected components of a link
with different objects of the category rather than with one single object.
Chapter xv
Quasi-Bialgebras
XV.l Quasi-Bialgebras
In XI.3.1 we introduced the notion of an algebra (A, 6., c) with comultipli-
cation and counit: it is an associative unital k-algebra A with a morphism
of algebras 6. : A --> A A (the comultiplication) and a morphism of alge-
bras c : A --> k (the counit). We observed that the classical tensor product
on Vect( k) restricted to a tensor product on the category A-Mod of left
A-modules, for which I = k is a unit.
(1.8)
of A. Let us check that <1>, I and r satisfy the conditions of the proposi-
tion. First, these elements are invertible because the constraints are iso-
morphisms.
We next prove Relation (1.2). This is done as in the proof of Proposition
XIII.1.4: by functoriality of the associativity constraint, for all u E U, v E V
and w E W, we have the commutative square
(AQ9A)Q9A ~ A Q9 (A Q9 A)
1
(u<Slv)<SI w 1 u<Sl(v<Slw)
(UQ9V)Q9W ~ U Q9 (V Q9 W)
a( <1>(uQ9 (v Q9 w)))
( (id Q9 ~) (~( a) ) <I> ) ( U Q9 (v Q9 w) ) .
(2.4)
(2.6)
The naturality of the braiding implies that for any pair V, W of A-modules,
the braiding Cv w is of the form (2.5). As a consequence of the A-linearity
of cA A' we get 6. 0P (a)R = R6.(a) for all a E A, which is equivalent to
Relation (2.1). The commutativity of the hexagons (HI) and (H2) in XIII. 1
implies Relations (2.2) and (2.3), as follows from an easy computation
using (2.5).
By (XIIL1.13) the category A-Mod is symmetric if cw,vcv,w = idvw
for all V and W. Now
PROOF. This counterpart of Theorem VIIL2.4 (a) follows from (1.9), (2.5),
and from Theorem XIIL1.3. D
(3.2)
The first and last equalities follow by definition, the third one from Relation
(1.1).
Relation (1.2): For all a E A we have
Now,
<l>123(~ Q9 id Q9 id)(F121)F121
F 34 (id Q9 id Q9 ~)(F23)<1>234(id Q9 ~ Q9 id)(F231)F23 1
F 23 (id Q9 ~ Q9 id) (F23 (id Q9 ~)(F)<I>(~ Q9 id)(F- 1)F121) F23 1
F 23 (id Q9 ~ Q9 id)(Fd<l>123(~ Q9 id Q9 id)(F121 )F121
(<I>Fh34 (id Q9 ~F Q9 id)(<I>F) (<I>F)123'
which proves (3.4). The first and last equalities follow from (3.2-3.3), the
second and sixth ones from the fact that ~ is an algebra morphism, the
third one holds because F12 and F34 commute, the fourth one follows by
applying (1.1) to a = F and F-l, and the fifth one from (1.3).
Relation (1.4): Using the definition of <I> F and Relations (1.4) and (3.1), we
immediately get (id Q9 c Q9 id)(<I>F) = FF- 1 = 1 Q91. 0
When F is an gauge transformation on A, then so is F- 1 and we have
(3.5)
If F' is another gauge transformation, then so is the product F F' and we
XV.3 Gauge Transformations 375
have
(3.6)
Definition XV.3.3. Two quasi-bialgebras (A,.6, 10, <1 and (A', .6 / , 10 ' , <1>/)
are equivalent if there exist a gauge transformation F on A' and an iso-
morphism a : A ---) A~ of quasi-bialgebras.
Lemma XV.3.4. Under the previous hypothesis, the triple (id, id, iff) is
a tensor functor from the tensor category A-Mod to the tensor category
ArMod.
PROOF. Recall Definition XI.4.1. We have to check Relations (XI.4.1-4.3),
namely if2(k, V) = if2(V, k) = id v and
if2(U, VI29W) (id u l29if2(V, W)) a{;,v,w = au,v,w if2(UI29V, W)(if2(U, V)l29id w )
(3.8)
where a F is the associativity constraint induced by <l>F. The first set of
equalities follows from (3.1) and (3.7). Let us prove (3.8). For all u E U,
v E V and w E W we have
The first and last equalities follow from (1.5) and (3.7), and the second one
from the definition of <I> F. 0
We state the first main result of this section. Let A and A' be equivalent
quasi-bialgebras with a gauge transformation F on A' and an isomorphism
a : A ---) A~ of quasi-bialgebras. The map a induces a strict tensor functor
(a*,id,id) from A~-Mod to A-Mod as explained in Example 2 of XI.4.
Since a is an isomorphism, a* is a tensor equival~nce.
Theorem XV.3.5. The tensor functor (a* , id, iff) is a tensor equivalence
between A'-Mod and A-Mod.
376 Chapter XV. Quasi-Bialgebras
Lemma XV .3.7. Under this hypothesis, the tensor functor (id, id, cpf) is
a braided tensor equivalence from A-Mod to ArMod.
PROOF. In view of Theorem 3.5, it is enough to show that (id, id, cpf) is
braided in the sense of Definition XIII.3.6. We must check that we have
cpf oc{;,v = cu,v ocpf. The latter is equivalent to F-1(RFb = (RF-1b,
which follows from (3.9). D
PROOF. This follows from Relation (XI.5.4) which expresses the tensor
c
product of morphisms in the strict category str in terms of the tensor
product of morphisms and of the associativity constraint in C. We also use
Relation (XI.5.3) in the following special cases:
When C = A-Mod is the braided category of left modules over the braided
quasi-bialgebra A, the braiding C is given by
CVV(vl
,
V2) = (R(VI v2)) 21 .
and if i > 1
Ci( VI ... Vn ) = <Pi 1((Ri,Hl <Pi)(V 1 ... Vn )) Hl,i (4.2)
378 Chapter XV. Quasi-Bialgebras
(4.3)
(R~2(VI Q9 ... Q9 vn )) 21
((F21 1(a Q9 a)(R)Fd(Vl Q9 ... Q9 Vn )) 21
It should be clear that the statement of Theorem 4.2 depends on the way
we put parentheses on V0 n . Other systems of parenthesizing give rise to
different, but equivalent braid group representations.
XV.5 Quasi-Hopf Algebras 379
and
We shall write (A,~, c, <1>, S, a, (3) to express the complete data of a quasi-
Hopf algebra. As in XIV.2 consider the category A-Modf of left A-modules
that are finite-dimensional vector spaces over the ground field k. Equip it
with the tensor category structure induced by ~ and <1>. For any object V of
A-Mod f consider the objects V* and *V as defined in Example 1 ofXIV.2.
We define maps bv : k ----t V Q9 V*, d v : V* Q9 V ----t k, b;,r : k ----t *V Q9 V, and
d;,r : V Q9 *V ----t k by
Proposition XV.5.2. The maps bv , d v , b;,r and d;,r are A-linear and the
composite maps
PROOF. The first statement follows from Relations (5.1), and the second
one from Relations (5.2). 0
(5.4)
X,Y,zEG gEG h
Then (DW( G), bo, c, CP, S, (X, (J) is a braided quasi-Hopf algebra with univer-
sal R-matrix R in the sense of Definitions 1.1, 2.1 and 5.1.
XV.7 Notes 381
XV.6 Exercises
1. Let (A,~, c, S) be a Hopf algebra and F = 2:i fi gi E A A be a
gauge transformation such that
Set
u= L S(i{3S(Zi))S(t j )as j Xi
i,j
XV.7 Notes
Quasi-bialgebras, quasi-Hopf algebras, and gauge transformations were in-
vented by Drinfeld [Dri89b][Dri90][Dri89c] in relation with his treatment
382 Chapter XV. Quasi-Bialgebras
Quantum Groups
and Monodromy
Chapter XVI
Generalities on Quantum
Enveloping Algebras
In order to state the main results of Part IV, we need the concept of a
quantum enveloping algebra. This requires the use of formal series and of
h-adic topology. The chapter is completed by an appendix on inverse limits.
(1.1)
PROOF. The formal series f is invertible if and only if there exists another
series g = Ln>o bnhn such that fg = 1. From (1.1) we see that this
is equivalent to -the existence of an infinite family (b o, bi , ... ) of complex
numbers such that aobo = 1 and
(1.3)
for all n > O. The relation aobo = 1 shows that the invertibility of ao is a
necessary condition for f to be invertible. This condition is also sufficient
since the family (bo,b i , ... ) can be determined inductively from bo = a 1 o
and Relations (1.3). 0
Lemma 1.1 may be interpreted as saying that the ring K is a local ring
whose maximal ideal is the ideal (h) generated by h.
For any integer n > 0 consider the algebra Kn = C[h]/(hn) of truncated
polynomials obtained as the quotient of the algebra of complex polynomi-
als in one variable by the ideal generated by h n. There is a morphism of
algebras 7rn from K to Kn sending a formal series f = Ln>o anh n to the
class of L~:~ akhk modulo (hn). This map is surjective a~d its kernel is
the ideal h n K generated by h n in the ring of formal series. Consequently,
7rn induces an isomorphism of algebras
(1.4)
~ k '
k=O
XVI.1 The Ring of Formal Series and h-Adic Topology 387
and we have Pn(Jn) = fn-l for all n > O. Hence a~n) = a~n-l) for k running
from 0 to n - 2. We can therefore define a formal series f = Ln>o anhn
by an = a~n+2) = a~n+3) = .... We have rr(J) = (In)n' - 0
(1.5)
As a consequence, we get
(1.6)
a trivial fact already used in the proof of Proposition 1.2. We also have
They form an inverse system of K-modules, and we may consider the in-
verse limit
(2.1)
which has a natural structure as a K-module. The inverse limit 1\/[ has a
natural topology, the inverse limit topology, for which it is easy to see as
in Section 1 that the family of submodules (h n l'vf)n is a family of open
neighbourhoods. The module M is called the h-adic completion of .M.
The projections in : IV! --+ Mn induce a unique K-linear map i : AI --+ /1.1
such that TIn 0 i = in for all n. The kernel of i is given by
Ker (i) = n h M.
n>O
n
For any ~dule M the module M/(nn>O hnM) is separated and the
completion M is complete. Indeed, consider the projection TIn : IV[ --+ IV!n'
Its kernel is hnAI, which implies the isomorphism of modules
(2.2)
Taking inverse limits, we get IV! = /1.1, which proves that IV! is complete.
Any separated, complete K-module will be equipped with the topologL
called the h-adic topology,--soming from the inverse limit topology on IV!
via the isomorphism M ~ M.
We now describe an important class of separated, complete K-modules.
It includes K itself, viewed as a K-module by left multiplication. Take any
complex vector space V. Define V[[h]] as the set of all formal series
(2.3)
Taking inverse limits yields a K-linear map goo between the corresponding
inverse limits. Since V[[h] and N are separated and complete, we get a
map, still denoted goo' from V[[h]] to N. This map restricts to g on V. 0
Topologically free modules can be characterized in a simple way. Recall
that a K-module M is torsion free if hm i- 0 when m is any non-zero
element of M.
Taking inverse limits and using the fact that M and V[[h]] are separated
and complete, we get
n n
o
We end with a caveat. We have V[[h]] ~ V C[[h]] only if V is a finite-
dimensional vector space. There is no such isomorphism when V is infinite-
dimensional, in which case V[[h]] is strictly bigger than V C[[h]]. Indeed,
take an infinite family (en)nEN of linearly independent vectors; then the
element Ln::::o enh n of V[[h]] does not belong to V C[[h]].
MN~NM. (3.2)
We also have
KM~M~MK, (3.3)
which means that K serves as a unit for completions.
The topological tensor product is functorial as can be seen from the
definition: if f : M ----> M' and 9 : N ----> N' are K-linear maps, then there
exists a K-linear map
fg : MN ----> M'N'
V[[h]]W[[h]] = (V Q9 W)[[h]].
are isomorphisms, where the first one is induced by (1.4) and the second
one is given by mQ9f f--* fm (the inverse map being induced by m f--* m(91).
Applying this to M Q9 K N where M = V[[h]] and N = W[[h]] we get
(4.5)
where TA,A' : A'0A' --; A''0A is the flip. In other words, the product in the
tensor product algebra A'0A' is given by
(4.9)
for all a E A,
(c'0idA)A = idA = (idA'0c)A, (4.11)
(idA'0idA'0A) (<I (A'0idA'0idA)(<I = <I>234 (idA'0A'0id A) (<I <I> 123 ,
(4.12)
and
(4.13)
When <I> = 1'01'01, we call A a topological bialgebra.
A morphism f : (A, p" 7], A, c, <I --; (A', p,', 7]', A', c', <I>') of topological
quasi-bialgebras is a morphism f between the underlying topological alge-
bras such that
on the space of formal series with coefficients in Ao where /-L, 7), 6. and c
are the unique K-linear maps such that 7)(J) = f7)o(l) = f1 for all f E K,
/-L(a0a') = ILo(a (>9 a'), 6.(a) = 6. o(a), c(a) = co(a)
for all a, a' E Ao. We call Ao[[h]] the trivial topological braided quasi-
bialgebra associated to Ao.
Example 4. Let A = (A, IL, 7),6., c, <P, R) be a topological braided quasi-
bialgebra. Since (A0A)/h(A0A) ~ A/hA (>9 A/hA, the K-linear maps
IL, 7), 6., c induce C-linear maps
fl : A/ hA (>9 A/ hA ---+ A/ hA, fl: C ---+ A/ hA,
LS. : A/hA ---+ A/hA (>9 A/hA, E: A/hA ---+ C.
Define <I? as the class of <P modulo (A/hA)3 and R as the class of R modulo
(A/hA)2. Then A = (A/hA, fl, fl, LS., E, <I?, R) is a braided quasi-bialgebra.
Another concept we need to adapt is the concept of a topological A-
module M over a topological algebra A = (A, /-L, 7)). It is a left K-module
with a K-linear map ILM : A0M ---+ M such that
ILM 0 (IL0id M ) = ILM 0 (id A 0ILM) and ILM 0 (7)0id M ) = id M . (4.19)
We shall write ILM(a0m) = am for a E A and m E M. The definition of a
morphism of topological A-modules is left to the reader.
Let M and N be topological A-modules. Then their topological tensor
product M0N is a topological A0A-module. If A has a comultiplication
6. : A ---+ A0A, we can pull back the A0A-module structure on M0N to
a topological A-module structure given by
a(m0n) = 6.(a)(m0n) (4.20)
for all a E A, m in M, and n in N.
XVI.5 Quantum Enveloping Algebras 395
(4.22)
and
Definition XVI.5.1. A quantum enveloping algebra (QUE) for the Lie al-
gebra g is a topological braided quasi-bialgebra A = (A, p" 'f/,~, c, cI>, R) such
that A is a topologically free module, the induced braided quasi-bialgebra
A = (A/hA, jl, i), L5., if, <1>, R) as in Example 4 of Section 4 coincides with
the trivial braided quasi-bialgebra structure of U(g) and the map 'f/ is triv-
ially extended from i).
Therefore
A = U(g)[[h]] (5.1)
as a K-module. From Proposition 3.2 we derive
for all n > O. By Proposition 2.3 (b) we know that the maps p, 'T/, ~ and
C are determined by their restrictions to U(g) U(g), C, U(g) and U(g)
respectively. For elements a, a' E U (g), we have
'T/(f) = 11 (5.6)
for all 1 E C[[h]]. Finally, by Proposition 3.2 again, the elements <P and R
can be written
<P = L<Pn hn (5.7)
n2:0
and
R= L Rn hn (5.8)
n2:0
where (<pn)n>O and (Rn)n>o are families of elements of U(g)03 and U(g)0 2
respectively such that -
in view of the fact (stated in V.g) that the subspace of primitive elements
in U (g) is 9 provided that the ground field is of characteristic zero.
We now associate another invariant to a QUE A. If R is its universal
R-matrix and R21 is the image of R under the flip, the formula
(5.11)
t21 = t. (5.12)
(5.13)
(5.14)
which implies that .6. 0(x i ) = Xi 01 + 1 0 xi for all i. Since the element xi is
primitive for the comultiplication .6. 0 of U(g), it belongs to g. Consequently,
t belongs to 9 0 Ug. Relation (5.12) implies actually that t is in 9 0 g.
Let us check the invariance of t. By (4.15) applied twice, we get
(5.15)
for all a E A. Identifying the coefficients of h, we obtain .6. o(x)t = t.6. o(x)
for all x E g.
398 Chapter XVI. Generalities on Quantum Enveloping Algebras
At this point, the only explicit quantum enveloping algebras for a given
Lie algebra g we know are the trivial QUE U(g)[[h]] constructed from U(g)
as explained in Example 3 of Section 4 and their gauge-transforms. Since
the universal R-matrix of such a QUE is 1 1, the corresponding canonical
2-tensor vanishes: t = O.
We now present an example of a QUE with a non-zero canonical2-tensor.
We shall see more non-trivial examples in Chapters XVII and XIX.
Example 1. (A quantum enveloping algebra associated to the Heisenberg Lie
algebra) We consider the 3-dimensional Lie algebra g with the set {x, y, z}
as a basis and with Lie bracket determined by
[x,y] = z and [x,z] = [y,z] = O.
The symmetric 2-tensor t = z (51 z is invariant because z is central in the Lie
algebra. We claim that there exists a QUE whose classical limit is (g, t).
Indeed, take the trivial bialgebra A = U(g) [[hll as in Example 3 of Section
4, except that we set R = e ht / 2 and <f> = 1~1~1. In order to make sure that
A is a topological braided bialgebra, we have to check Relations (4.15-4.17).
The first one follows from the fact that t is invariant. Relations (4.16-4.17)
with <f> = 1~1~1 are equivalent to
eh (t 13+ t 12)/2 = ehh3/2eht12/2 and e h(h3+t23)/2 = ehh3/2eht23/2. (5.17)
Relations (5.17) hold because the elements t 12 , t 13 and t 23 commute with
one another, due to the centrality of z. Now,
R21 R = e ht == 1 (51 1 + ht mod h 2
shows that t = z z is the canonical 2-tensor of A.
(1+
n>O
L Cn an h n f= 1 + ah. (6.1)
XVI.6 Symmetrizing the Universal R-Matrix 399
PROOF. Any formal series 1 + I:n>O cnanh n of the above form defines an
element of the inverse limit A = lim A/hn A, hence of A since A ~ A by
n
hypothesis. Equation (6.1) is equivalent to the system of equations
n-1
2c 1 = 1 and 2c n +L cpcn _ p = 0 (6.2)
p=l
if n > l. This system has a unique solution as can be seen by an easy
induction. 0
The unique element 1+ I:n>O cnanh n satisfying (6.1) is called the square
root of the element 1 + ah and is denoted by (1 + ah)1/2. Its inverse will be
denoted by (1 + ah)-1/2.
(6.3)
which can also be written in the form
(6.4)
We claim that
1/2
F = ( R(RR)-1/2 ) (6.5)
is a solution of (6.4) where we use the notation defined after Lemma 6.l.
The element F is invertible and congruent to 1 Q9 1 modulo h since R is.
In order to prove the claim, we observe that ReRR) = (RR)R implies
R(RR)n = (RR)n R for all n 2 1, hence
XVI. 7 Exercises
1. Show that
1+ !!. + '"' (_1)n-1 (2n - 3)!! hn
2 ~ 2nn!
n2:2
is a square root of 1 + h in the algebra C[[hll of formal series where
(2n - 3)!! = TI~:i (2k - 1).
2. Let M = V[[hll and N = W[[h]] be topologically free modules.
Show that HomK(M, N) is a topologically free module isomorphic to
Hom(V, W)[[h]]. Deduce that if P is a third topologically free module,
then
Hom K (M0N,P) ~ HomK(M,HomK(N,P)).
3. Let g be a Lie algebra and t E g0g such that [t12' td = [t 13 , t 23 ] = 0
in U(g)3. Consider the gauge transformation F = e ht . Show that
(U g[[hllh is a topological bialgebra.
4. Show that the inverse systems of abelian groups and the maps of
inverse systems form a category Inv such that li-Ill is a functor from
n
Inv to the category Ab of abelian groups. Prove that Ii-Ill is left adjoint
n
to the functor assigning to each abelian group A the constant inverse
system (An,prJ where An = A and Pn = idA for all n.
5. Let (Cn)n>O be a denumerable family of abelian groups. Consider
the inverse system (An,p n ) where An = Co X ... X C n and Pn is
the natural projection. Prove that the inverse limit of this system is
isomorphic to the direct product of all groups Cn.
6. Let (An,p n ) be an inverse system of abelian groups. Use the fact that
its inverse limit can be expressed as the kernel of an endomorphism
of TIn An to prove that for any abelian group C there is a natural
isomorphism
Hom(C,li-Ill An) ~ li-Ill Hom(C, An)
n n
XVI.9 Appendix. Inverse Limits 401
7. (The ring of p-adic integers) Given a prime p consider the inverse sys-
tem of rings (Zjpnz) equipped with the natural projections induced
by the inclusions of ideals (pn) C (pn-l). Show that the inverse limit
Zp is a ring with a unique maximal ideal. Prove that the inverse limit
topology on Zp can be defined by an ultrametric distance and that
the ring of natural integers Z forms a dense subring of Zp in which
all integers prime to p are invertible.
XVI. 8 Notes
The material of Sections 1-4 is standard. For details on h-adic topology
and completions, read [Bou61], III and [Mat70], Chap. 9. The concept of
a quantum enveloping algebra and the content of Sections 5-6 are due to
Drinfeld (see [Dri87] and [Dri89b], Section 3). Exercise 3 is taken from
[Enr92].
(9.2)
for all n > O. The inverse limit has the following universal property.
Proposition XVI.9.1. For any abelian group C and any given family
Un : C ~ An)n>O of morphims of groups such that Pn 0 fn = fn-l for all
n > 0, there exists a unique morphism of groups
The inverse limit is functorial. Define a map from the inverse system
(An'Pn) to the inverse system (A~,p~) as a family Un : An ----t A~)n>O of
morphisms of groups such that p~ 0 in = in-l 0 Pn for all n > O. -
Proposition XVI.9.2. Under the previous hypothesis, there exists a
unique morphism of groups
n n n
The inverse limit of any inverse system (An' Pn) possesses a natural topol-
ogy called the inverse limit topology. It is obtained as follows. Put the dis-
crete topology on each An' i.e., the topology for which each subset is an
open set. The inverse limit topology on 1,Lrp. An is the restriction of the di-
n
rect product topology on ITn>O An' In other words, a basis of open sets of
the inverse limit is given by the family of all subsets 7r;;l(Un ) where n runs
over the non-negative integers and Un is any subset of An' By definition of
this topology, the structural maps 7rn from 1,Lrp. An to An are continuous.
n
Moreover, a map i from a topological set to 1,Lrp. An is continuous with
n
respect to the inverse limit topology if and only if the map 7rn 0 i into An
is continuous for all n ~ O.
One may replace the word "abelian group" by "ring", "module" ... in the
above definition. The statements ofthe Appendix remain true in this case, a
fact we have consistently used in this chapter without further explanation.
Chapter XVII
Drinfeld and Jimbo's Quantum
Enveloping Algebras
< y,x >p=< X,y >p and < [x,y],z >p=< x, [y,z] >p (1.2)
404 Chapter XVII. Drinfeld-Jimbo Quantum Enveloping Algebras
j j
(1.4)
Expanding the left-hand side of (1.4) gives < [Xi' X], x j >ad = aij(x) whereas
we have < Xi' [X, xj] >ad = -(3ji(X) for the right-hand side. D
yi = 2.::: Bjixj
j
[C,x] L[xixi,X]
L xi[Xi,X] +L [Xi,X]xi
o
by Lemma 1.1. D
of g @ g. This element will playa central r61e in Chapter XIX. It enjoys the
following properties.
(1.8)
406 Chapter XVII. Drinfeld-Jimbo Quantum Enveloping Algebras
(1.11)
and
L
l-aij
k=O
(_l)k C -kaij ) Yik1jYiI-aij-k = O. (1.12)
The Cartan matrix for .5[(2) is the 1 x I-matrix A = (2) with D = (1).
In this case, the presentation above reduces to the formulas (V.3.2).
We end this summary by a few words on the representation theory of a
semisimple Lie algebra g. Any finite-dimensional g-module is semisimple,
i.e., is the direct sum of simple modules. The finite-dimensional simple g-
modules are classified by the set of dominant weights: a dominant weight is a
linear form A on the subspace f) of g spanned by HI"'" Hn such that A(Hi)
is a non-negative integer for all i = 1, ... ,n. For every dominant weight A,
there exists a unique finite-dimensional simple g-module VA generated by
an element VA' called a highest weight vector, such that
(1.13)
for all i = 1, ... ,n. All finite-dimensional simple g-modules are of this form.
The Casimir element C acts by a positive scalar on every simple g-module
VA of dimension> 1, i.e., with A =I- O. We have proved these facts for 5[(2)
in Chapter V. In the case of .5[(2) the set of dominant weights is in bijection
with N, the dominant weight A corresponding to the integer n being defined
by A(H) = n.
We then take the inverse limit of the maps f n . The uniqueness of 1 results
from the fact that the K-subalgebra generated by X is dense in K(X). D
(2.2)
and if i -=I j
and
(2.4)
sinh(hdi H;/2) _
sm 2 = Hi
. h(JLd i j) mod h.
We refer to [Dri87] for a proof. Let us make a few remarks. First, if we
set h = in Relations (2.1-2.4) and (2.6-2.9), we recover the enveloping
algebra of 9 in Serre's presentation. In other words, we have an isomorphism
of algebras
(2.lO)
The fact that Uh(g) is a topologically free K-module is not straightforward.
It can be proved by constructing a Poincare-Birkhoff-Witt-type basis. One
has also to check that (2.6-2.9) define morphisms of algebras D..h and ch'
For D..h this follows from (2.5) and the q-binomial formula of Proposition
IV.2.2.
The topological bialgebra Uh (g) has an antipode Sh determined by
Note that the comultiplication of Uh(g) is not cocommutative and that the
antipode is not involutive. Nevertheless, for all a E Uh(g) we have
(2.12)
where p = 2:~=1 J-liHi' the scalars J-li being determined from the inverse
A-I of the Cartan matrix by J-li = 2:7=1 (A-l)jidj.
More importantly, Theorem 2.4 implicitly states that Uh (g) has a uni-
versal R-matrix, which we denote by R h . Drinfeld proved that Rh is of the
form
(2.13)
(2.16)
410 Chapter XVII. Drinfeld-Jimbo Quantum Enveloping Algebras
and generated by an element VA' called a highest weight vector, such that
(2.17)
for all i = 1, ... ,n, as in the classical case. Rosso [Ros88] proved that any
topologically free Uh(g)-module W with dim(W/hW) < 00 was a direct
sum of modules of the form VA' We shall give an explanation of this fact
in XVIII.4.
(3.1)
Then Proposition 3.1 follows from Proposition VIII.4.5 and from the fact
that uS (u) has a unique square root whose constant term is 1.
Relation (3.2) is reduced in [Dri89a], Proposition 5.1 to showing that
both terms have the same action on all modules of the form VA' It is
enough to evaluate the ~ntral elements S(u)u and 02 = e- 2hP u 2 on a
highest weight vector of VA' Since u can be expressed in such a way that
the generators Xi killing the highest weight vector appear to the right of
Yi, we see that the actions of S(u)u and of 02 are the same as the actions
of the elements obtained from the part of Rh corresponding to I! = 0 in
Formula (2.13). A simple computation shows then that S(u)u and of 0 2 act
by the same scalar on VA' For more details, see [Dri89a], Section 5. 0
(3.3)
(3.4)
(3.6)
of Uh(g) satisfying
Ch==C modh. (3.7)
Proposition 5.1 of [Dri89a] asserts that
(3.8)
and
sinh(hH/2) e hH / 2 _ e- hH / 2
[X, Y] = sinh(h/2) = eh/2 _ e- h/ 2 (4.2)
KE=lEK (4.6)
X(E) = E, X(H) = H,
(q - q-l )[E, F]
K-K-l
K _ e- hH/ 2
X(K - e- 2 .6. 1 ).
o
The final step of the proof of Theorem 4.2 goes as follows: By Propo-
sitions 4.3-4.4 we know that any topological Uh-module becomes a topo-
logical crossed Bh-bimodule via x. In view of Relation (IX.5.5) a universal
R-matrix for Uh is given by
qn(n-l)/2 _
Rh = 2:
m,n?O
I[ ]
m. n q.
I m
X(H En)@X(b.r;' b.~). (4.10)
By definition of X we get
"
(q - q -l)n qn(n-l)/2 _hm H m En @ H m F n
~ m![n]! 4m .
m,n2:0 q
for all p > o. In the inverse limit the family (b.~,n,p)p assembles to form a
K-linear endomorphism b.~,n of M. Now the sum in (4.11) is finite, which
implies that b.~n(x) vanishes modulo h for m and n large enough.
The counitarity of b. M yields b.~o = id M whereas the coassociativity
gives
b.i,j b. m,n
M M
= -2nj (j + n)
q n
,,(jh)t
~ 2tt!
(i + t) m -
i
b.i+m-t,j+n
M
(4.12)
q2 t2:0
for all i, j, m, n after using the classical binomial formula as well as the
q-binomial formula of Proposition IV.2.2. Here we agree that b.~n = 0
416 Chapter XVII. Drinfeld-Jimbo Quantum Enveloping Algebras
when m or n < O. Set .6. 1 = .6.~0 and .6. 2 = .6. f;.:/. By (4.12) and (VI.1.7)
we get
.6. m ,O = ~(.6.1,0)m = ~.6.m, (4.13)
M m! M m! 1
n(n-l) n(n-1)/2
.6.O,n = q (.6. O,l)n = q .6. n (4.14)
M (n)q2! M [n]q! 2'
n(n-l)/2
.6. m,n - .6. m,O .6. O,n - q .6. m .6. n (4.15)
M - M M - m![n]q! 1 2'
and
.6. m,n E
M
+ .6. m,n-l
M
K= "'" ~
~2tt!
E.6. m-t,n + "'" (_2)r (M + r).6. m+r,n-1
M ~ r M
t20 r20
(4.17)
for all m, n ~ O. Specializing the exponents m and n to 0 and 1 in (4.16)
gives Relations (4.8) whereas setting m = 1 and n = 0 in (4.17) gives
[E,.6. 1] = -~E. When we set m = 0 and n = 1 in (4.17), then necessarily
t = 0 and we get
0 [n]q 0 0
0 0 [n -1]q 0
Pn(X) =
0 0 1
0 0 0 0
0 0 0 0
1 0 0 0
Pn(Y) = 0 [2]q 0 0
0 0 [n]q 0
and
n 0 0 0
0 n-2 0 0
Pn(H) =
0 0 -n+2 0
0 0 0 -n
Therefore,
+ e(n-2)h/2 + ... + e-(n-2)h/2 + e- nh / 2
e nh / 2
q 0 0 0 )
-1/2
( 0 q 0 0
(4.20)
q 0 0 0 1 '
o 0 1 q _ q-1
XVII. 5 Exercises
1. Compute Pc in Formula (2.13) for Rh when g = (0, ... ,1, .. , ,0) where
1 occurs exactly once.
XVII. 6 Notes
A full account of the theory of semisimple Lie algebras can be found, for
instance, in [Bou60][Dix74][Hum72][Jac79][Ser65][Var74]. See [Bou60] for
the complete list of Cartan matrices.
XVII.6 Notes 419
In the special case 9 = 13[(2), the algebra U h(13[(2)) had previously been
constructed by Kulish and Reshetikhin [KR81] with the Hopf algebra struc-
ture found by Sklyanin [Sk185].
Drinfeld devised the quantum double construction precisely in order to
find a universal R-matrix for Uh(g). This method was applied by Drinfeld
[Dri87] himself to give an explicit form of Rh in the case 13[(2) and by Rosso
[Ros89] in the case 13[(n). Expressions of the universal R-matrix in the
general case are due to Kirillov-Reshetikhin [KR90] and to Levendorsky-
Soibelman [LS90].
The representation theory of Uhg was elucidated by Lusztig [Lus88] and
Rosso [Ros88].
Chapter XVIII
Cohomology and Rigidity
Theorems
for all x, y E g and m E M. It was shown in V.2 that a left g-module is the
same as a left module over the enveloping algebra U(g) of g.
For n > 0, let Cn(g, M) = Hom(A ng, M) be the space of all antisym-
metric n-linear maps from g to M. An n-linear map f is antisymmetric
if f(Xu(l)' ,xcr(n)) = c(O")f(Xl' ... ,x n ) for all Xl' ... 'X n E g and all
permutations 0" of the set {I, ... , n}. If n = 0, we set CO(g, M) = M.
XVIII.1 Cohomology of Lie Algebras 421
for all Xl" .. 'Xn+l E g. The hat ~ on a letter means that it has been
omitted. If f belongs to eO(g, M) = M, we set (of)(x) = xf. A classical
computation using the Jacobi identity and the definition of a g-module
gives the following.
(1.3)
which is called the n-th cohomology group of the Lie algebra 9 with coeffi-
cients in the g-module M.
Let us describe Hn(g, M) in degree n = 0,1,2. In degree 0 we have
xf(y, z)+yf(z, x)+zf(x, y)- f([x, V], z)- f([y, z], x)- f([z, x], y) =0 (1.5)
We shall see in the next section that 2-cocycles appear when we "deform"
Lie algebras and their enveloping algebras.
The second cohomology group H2(g, M) has also an interpretation in
terms of extensions of g. These are defined as follows. Let 9 be a Lie algebra
and M be a left g-module. An extension of the Lie algebra 9 with kernel
M is a Lie algebra ~ together with a surjective morphism p : ~ --+ 9 of Lie
algebras such that
(i) the kernel of p (which is a Lie ideal in ~) is M, and
(ii) for any x E ~ and m E M, we have
[x,m] = -[m,x] =p(x)m. (1. 7)
Theorem XVIII.2.1. Let fI and fI' be Lie algebras. Suppose given two
morphisms a and a' of topological algebras from U(fI) [[hll to U(fI')[[h]] such
that a == a' modulo h. If H1(g, U(g')) = 0, there exists an invertible element
F E U(fI')[[hll with F == 1 modulo h such that a'(x) = Fa(x)F-1 for all
x E U(fI)[[hll.
The class modulo h of a (and of a') is an algebra morphism a o from U(g)
to U(g'). We give U(fI') a left I-module structure by setting xu = [ao(x),u]
where x E I and U E U(g'). The cohomological condition in Theorem 2.1
refers precisely to this module structure.
PROOF. Since a is C[[hll-linear, it is determined by its restriction on U(g).
Write the latter in the form
where (an)n is a family of linear maps from U(g) to U(g'). The map a
preserves the unit, which implies that a o(1) = 1 and an (1) = 0 if n > O. It
also preserves the product, which is equivalent to the relations
(2.2)
and
an(xy) = L ap(x)aq(y) (2.3)
p+q=n
if n > O. In particular, we have
(2.4)
Suppose now that x and yare elements of fl. Then Relation (2.4) implies
that
(2.5)
In view of our definition of the fI-action on U (fI') and of (1.4), we see that
a 1 is a 1-cocycle of I with values in U(g'). Since H1(g, U(g')) = 0, the map
a is a 1-coboundary, which means that there exists an element u 1 E U(g')
such that
(2.6)
for all x E g. Set
(2.7)
XVIII.2 Rigidity for Lie Algebras 425
tL = L tLn hn (2.13)
n;::O
(2.14)
for all x E U(g) and all n > O. The associativity of the product tL is
expressed by
(2.15)
for all x, y, z E U(g). Expanding tL with (2.13), we obtain the equivalent
system of equations
for all x, y, z E U(g) and all n ~ O. Let N be the smallest integer n > 0
(if it exists) such that tL n i- O. If no such integer exists, we have tL = tLo,
which means that A coincides with U(g)[[h]] as a topological algebra and
the theorem is proved. If N exists, let us rewrite (2.16) for n = N. Using
the customary notation for the product in U(g), we get
for all x, y, z E U(g). In other words, tLN satisfies Condition (1.9) of Corol-
lary 1.3 with M = U(g). Since H2(g, U(g)) vanishes, we may apply Corol-
lary 1.3, which yields a linear endomorphism aN of U(g) with aN(l) = 0
and
(2.18)
for all x, y E U(g). Define a C[[hll-linear automorphism a of U(g)[[h]] by
XVII1.3 Vanishing Results for Semisimple Lie Algebras 427
(2.19)
Since 0: == id mod h N , we have 11' == 11 mod hN. Let us compute 11' modulo
h N + 1 . Relation (2.18) implies that
for all Y1"'" Yn -1 E g. If f E CO(g, M), set hf = O. Using (3.2) and (l.2),
we get
where
Zi = L ([XklYi]J(X k 'Y1, ... ,Yi, ... ,Yn)+Xkf([X k'Yi],Yl, ... ,Yi, ... ,Yn))'
k
Relation (3.1) will be proved if we show that all Zi vanish. Using the linear
forms ak and (3kf of XVII.l, we get
Let us equip U(g) with the adjoint representation of g for which the Lie
algebra acts on U(g) on the left by Xl1 = Xl1-l1X = [x,u] where x E g and
11 E U(g). If u = Xl ... xn with xl' . .. ,x n belonging to g, an easy induction
shows that
n
X.U = L Xl .. Xi - l [x, xi]xi+l ... Xn (3.4)
'i=l
430 Chapter XVIII. Cohomology and Rigidity Theorems
U(g) ~ EB sn(g)
n::o-O
U(g). By Theorem 2.1 and by the vanishing of H1(g, U(g (see Corollary
3.3), there exists an element F == 1 mod h in U(g)[[h]] such that we have
(a' 0 a- 1)(u) = FuF- 1 for all u E U(g) [[h]]. Replacing u by a(a) yields
the conclusion. 0
6~+16~ (Xl Q9 ... Q9 Xn) = Xl Q9 ... Q9 Xi - l Q9 (~Q9 ide )(~(Xi)) Q9 x'i +1 Q9' .. Q9 x n
~(l)Q9XIQ9"'@Xn
1 Q9 1 @ Xl Q9 .. Q9 Xn
6~+16~ (Xl Q9 ... @ XrJ
606
j=O i=O
i<j j<O,i
,",(-1)i+j(6
L.." n+1 6n _6 n+l 6n-
j i i j l)
i<j
o
by Lemma 5.1. o
The natural isomorphisms Tn(c) Q9 TTn(C) ~ Tn+Tn(C) induce an asso-
ciative graded product on T(C) = EBn>O Tn(c). This product is compati-
ble with the differential 6 in the follOWIng sense.
Lemma XVIII.5.3. If wE Tn(c) and w' E TTn(C), then
We shall use the term differential graded algebra for a graded algebra
with a differential satisfying Relation (5.1). It follows from Lemma 5.3
that the product on T- (C) induces an associative graded product on the
cohomology H- (T- (C), 6) of the co bar complex.
XVIII.5 Cohomology of Coalgebras 433
PROOF. This results from the relations 6:,+rn (ww' ) = 6~ (w )w' if i ~ n, from
d-n( W ') 1'f Z. >
+ m ( WW ') -- wUrn
u>:in _ n + 1 ,an d f rom
xn)
6~O'n(XI Q9 ... Q9
(_1)n(n+I)/2 xn Q9 ... Q9 ,6,(x n + I -.J Q9 ... Q9 Xl
(_1)(n+l)(n+2)/2-(n+l) Xn Q9 .. Q9 ,6,0P(X n +l_ i ) Q9 ... Q9 Xl
(-1)"+IO'n+16~+I-i(XI Q9 ... Q9 XrJ
The second equality holds by the assumption on ,6,. D
in Sweedler's sigma notation. The examples we have in mind are the coal-
gebras U(g) and 8(g), on which 9 acts by the adjoint representation.
Equipping the tensor powers of C with the induced g-module structures,
we get the following result.
PROOF. It suffices to check that the maps /5~ of Section 5 are maps of
g-modules. Let c l , ... , cn be elements of C and x be in g. For /5~, we get
n
L 1 c l ... x . c k ... cn
k=l
X . /5~ (c l ... cn)
x . /5~ (c l ... cn )
by (6.1). o
Observe also that the subcomplexes (T(C), /5) are preserved by the g-
action where C is equipped with the identity involution.
We next restrict to the case when 9 is a finite-dimensional semisimple
Lie algebra acting on C such that C is a direct sum of finite-dimensional
XVIII.7 Computations for Symmetric Coalgebras 435
fl-modules. This is the case for C = 5(fl) and, hence, for the isomorphic
coalgebra U(fl). For any fl-module V, we define a fl-submodule V9 by
V9 = {v E V I x v = 0 \;fx E fl}.
Elements of V9 are called fl-invariant. The linear span flV of the elements
x . v where x runs over fl and v over V is also a fl-submodule of V.
V = V9 EElflV (6.2)
and
The rest of the section is devoted to the proof of Theorem 7.1. The idea
is to dualize the complex (T(S(V)), b) and to compute the homology of
the dual complex. For a definition of the exterior algebra, see II, Exercise 6.
We first need the concept of a graded dual vector space: if V = ffin>o Vn
is a vector space with a positive grading, we define the graded dual vector
space of V by
~~ = EB V;. (7.2)
n2:0
We can apply this to the vector spaces T(V), S(V), A(V) and T(S(V))
with their natural gradings. If V = ffin>o Vn and W = ffin>o Wn are
vector spaces with gradings, we may consider their tensor prodllct V W
graded by
(VW)n= VpWq' EB (7.3)
p+q=n
Lemma XVIII.7.2. Suppose V = ffin>o v" and W ffin>o Wn have
gmdings fOT which v" and Wn aTe finite-dimensional fOT all n.- Then theTe
is a canonical isomoTph'lsm
PROOF. This is straightforward. It uses the fact proved in II.2 that this
isomorphism holds for finite-dimensional vector spaces. 0
Let (C, ~, c, 1) be a graded coalgebra with unit, meaning that the under-
lying vector space C = ffin>o C n has a grading, that ~ and c are graded
maps (we equip k with the-trivial grading concentrated in degree zero),
XVIII. 7 Computations for Symmetric Coalgebras 437
(7.4)
The chain complex (Te(A), d) is called the bar complex of the augmented
algebra A.
PROOF. Apply both sides to Xl @ ... @ xn where Xl' ... ,xn belong to C.
D
PROOF. Let {VI' ... ,VN} be a basis of V. Then {vr1 ... V~N} <>1 +.+<>N=n
is a basis of sn(V). We define a basis {wr 1 ... w~N}<>1+.+<>N=n of sn(V)*
by
(7.5)
438 Chapter XVIII. Cohomology and Rigidity Theorems
< "'N)
( WI"'1 . "WN * (131
WI"
13 N ) 'Y1
'WN ,VI" 'VN
'Y N >
< WI"'1 "'N fQ.. 131
. "WN 'CYWI . "WN
f3N A( 'Y1
,L.l. VI
'YN)
.. 'VN
>
< WI"'1 "'N
... W N fQ..
'CY
f3N , ( VI
WI131 ... W N fQ..
'CY
1 + 1 'CY
fQ..
VI )'Y1 ...
(7.6)
which shows that the product on S(V);r is the product of the symmetric
algebra S(V*). The rest of the proof is left to the reader. D
then f-l is a chain map. We have (f-l oo:)(w) = n!w for all elements w
belonging to An(w).
PROOF. We proceed in six steps using the terminology and the results of
the Appendix.
XVIII. 7 Computations for Symmetric Coalgebras 439
(7.8)
where T~(A) = A Q9 A@n and the left A-linear differential d' is given by
n-1
d'(a o Q9 ... Q9 an) = L (-l)ia o Q9 .. Q9 a i - 1 Q9 a i a i +1 Q9 ai+2 Q9 .. Q9 an
i=O
(7.9)
for an, ... ,an EA. The reader may easily check that d' 0 d' = O.
2. We claim that the complex (T' (A), d') is a resolution of k by free left
A-modules. It suffices to prove that the complex
... ~T~(A)~T{(A)~TMA)~k -+ 0
s( ao Q9 .. Q9 an) = 1 Q9 a o Q9 .. Q9 an (7.10)
where a E S(W), w 1 , ... ,wn E Wand where the hat on wi again means
that we omit this element. Check that 808 = o.
We claim that (K.(W), 8) is a resolution of k. This again is due to the
existence of a homotopy: define a map h : Kn (W) -+ Kn+l (W) by
m
for all P in sm(w) and win An(w). Relation (7.14) shows the acyclicity
of the Koszul resolution in degree> O. As for degree 0, observe that the
440 Chapter XVIII. Cohomology and Rigidity Theorems
isomorphic to k.
Since c:(w) = for all W E W, (k S(W) K.(W),id k S(W) 8), the in-
duced complex, is isomorphic to the complex A (W) with zero differential:
where
ZI = L C:(O")Wu(l) ... Wu(n) ,
uESn
n-I
Z2 = L( _l)i L C:(o")WU(I) ... wu(i)wu(i+I) ... wu(n)
i=1 uESn
and
Z3 = (_l)n L C:(o")Wu(l) ... Wu(n-I)C:(Wu(n)'
uESn
Let us first deal with ZI' We have
n
ZI =L L C:(O")Wi wu(2) ... wu(n)'
i=l crESn.
u(I)=i
by (7.12). Relation (7.16) will be proved once we have checked the vanishing
of Z2 and of Z3. Concerning Z2' we have
L c:( 0" )WO"(I) IZi ... IZi wa(i) wO"(Hl) IZi ... IZi wO"(n)
aESn,
O"(i)<O"(Hl)
+
aESn.
O"(Hl)<O"(i)
(ids(w) IZi 0:)(3 = id + d'h l + hI d' and (3(id s (w) IZi 0:) = id + 8h 2 + h28.
(7.17)
Tensoring both resolutions on the left with k over S(W), we see that
< j1(Wil 00 wiJ, v jl A ... A Vjn >=< w il A ... A win' Vjl A ... A Vjn >
vanishes when (i l , ... , in) is not a permutation of (jl' ... ,jn). If it is, the
right-hand side is equal to the sign of this permutation. On the other hand,
(8.1)
Indeed, we have
R12<P312R13(<P132)-1 R23<P
<P321 R 23 (<P 231 ) -1 R 13 <P213 R 12
<P321R23(id 0 ~)(R)<p.
The first and last equalities follow from Proposition XV.2.2 while the mid-
dle one follows from Corollary XV.2.3. Next, apply the involution 7 13 to
A 0 A 0 A. Since ~ = ~op and R = R 21 , Relation (8.1) becomes
(8.2)
Proof of Theorem 8.1. We have to show that we can find a gauge transform-
ation taking <1'> to <1'>'. Suppose <1'> and <1'>' are equal modulo h n for some n 2': 1.
This always holds for n = 1 since <1'> and <1'>' are congruent to 1 1 1 modulo
h. Define 'P E U(I)3 by
of U (I) 3. The first step in the proof of Theorem 8.1 is the following lemma
with the same hypotheses.
Lemma XVIII.8.3. The element 'P is I-invariant and satisfies the rela-
tions 'P321 = -'P, Ant('P) = 0, and
(d) Using the pentagonal relations (XV.1.3) for <1> and <1>', we get
(1 <1>')(id ~ id)( <1>')( <1>' 1)(~ id id)( <1>') -1 (id id ~)( <1>')-1
= (l<1>)(id~id)(<1>)(<1>l)(~idid)(<1>)-l(idid~)(<1>)-l = 1.
Reducing these equalities modulo hn+l, we obtain the desired 5-term func-
tional equation for rp. 0
The first two relations in (8.7) mean that rp is a 3-cocycle in the co-
bar complex (T~(U(g)),b). We claim that rp is a coboundary. Using the
isomorphism TJ : 5(g) ----+ U(g) of co algebras , it suffices to check that
1jJ = (r)-l 7)-1 TJ-1)(rp) is a coboundary in (T~(5(g)),b). We also have
Ant(1jJ) = O. By Theorem 7.1 (c) we have H3(T~(5(g)),b) ~ A3 (g), the
isomorphism being induced by the map JL. It is therefore enough to check
that JL(1jJ) = O. Now Ant(1jJ) = 0 implies Ant((JLJLJL)(1jJ)) = O. An imme-
diate computation shows that a(JL( 1jJ)) = Ant ( (JL JL JL) (1jJ)). Therefore,
by Theorem 7.1 we have
1 1
JL(1jJ) = 6 JL(a(JL(1jJ))) = 6 JL(Ant((JL JL JL)(1jJ))) = 0,
We have so far proved Lemma 8.4 up to the fact that we can choose 1 to
be g-invariant.
This last fact is a consequence of Proposition 6.2 applied to the co algebra
U(g) on which 9 acts by the adjoint representation (this is where we use
the assumption that 9 is semisimple). Since (T~(U(g)),I5) splits into the
direct sum of (T~(U(g))g,l5) and of (gT~(U(g)))I5) and since rp belongs to
T~(U(g))g) then 1 necessarily belongs to T~(U(g))g. 0
446 Chapter XVIII. Cohomology and Rigidity Theorems
(8.9)
The g-invariance of f implies that [F, ~(x)] = 0 for all x E U(g). We also
have F2I = F. We already know that ~ and R remain unaffected by such
a gauge transformation. Let us compute <l>F modulo hn+l. From (8.8-8.9)
and from (XV.3.3) we get
and start the whole procedure all over again. By composing all the gauge
transformations obtained in this way, we obtain a gauge transformation
between the quasi-bialgebras A and A'. This completes the proof of Theo-
rem 8.1. 0
XVIII. 9 Exercises
1. Compute H2 (g, C) for all complex Lie algebras of dimension:::; 3.
XVIII. 10 Notes
The content of Sections 1-7 is classical. The cohomology of Lie algebras
was introduced by Chevalley and Eilenberg in [CE48] following ideas of E.
Cartan. See [Ger64] for a general deformation theory for algebras.
XVIII.ll Appendix. Complexes and Resolutions 447
(11.5)
for all n (by convention h_1 = 0). If there exists a homotopy between f
and 1', we say that the two chain maps are homotopic. Homotopy is an
equivalence relation. Homotopic chain maps f, f' induce the same map on
homology: f. = f~
One of the basic results in homological algebra is the following compar-
ison theorem for resolutions.
PROOF. Applying the above theorem to f-1 = id, we get chain maps f,g
such that 10 = 10' 0 fo and 10' = 10 0 go' Now go f is a chain map from (e., d)
to itself with 10 0 (go 0 fo) = E. SO is the identity on e .. By the second part
of the theorem, we see that g 0 f is necessarily homotopic to the identity.
A similar argument works for fog. D
Chapter XIX
Monodromy of the
Knizhnik-Zamolodchikov Equations
XIX.1 Connections
We assume some standard knowledge of differential geometry. Let us never-
theless recall a few facts. For more details, the reader may consult [KN63].
450 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
If V'1 and V' 2 are two connections on E, then the difference V'1 - V' 2 is
O(X)-linear, where O(X) is the ring of complex analytic functions on X.
Locally, we can write a section s in the form
(1.2)
where 11 , ... , Id are complex analytic functions on X and {e 1 , ... ,ed} is a
basis of the fibre. Any connection V' on E can be written locally as
V's = ds - rs (1.3)
where d is the de Rham differential and r is a differential I-form on X with
values in the endomorphism ring of E.
A section s of the bundle is horizontal for the connection V' if V's = 0,
i.e., if locally s is a solution of the system
ds = rs. (1.4)
Let "( : [0, 1]---t X be a smooth path in X from Xo = "((0) to Xl = "((1). We
may pull back the matrix r of differential forms on X along "( to a matrix
A(O)dO = "(*r of differential forms on the interval [0,1]. By the theory of
ordinary differential equations, there exists a unique smooth map A, from
[0, 1] into the group of linear automorphisms of the fibre bundle such that
A,(O) = id and w(O) = A,(O)w(O) is a solution of the differential equation
(1.6)
The holonomy group at Xo is defined as the subgroup of Aut(Fxo) gen-
erated by T, for all loops "( based at Xo at X. In general, the holonomy
depends on the local as well as on the global structure of X. In other words,
XIX.2 Braid Group Representations from Monodromy 451
(2.2)
dw = '"
L...J A
--'-J-(dzi - dzj)w. (2.3)
l~i<j~n'
z - zJ
r= I: A
--'J-(dz. - dz.)
z.-z. ' J
(2.4)
l~i<j~n' J
We have
r 1\ r = I: AijAk u ij 1\ uk' (2.5)
i<j, k<
Kp = I: AijAki' u ij 1\ uk'
i<j, k<
the set of indices {i < j, k < } C {I, ... ,n} running over all such subsets
of cardinality p. We now show that K 2 , K3 and K4 vanish separately due
to Relations (2.1-2.2). For K 2 , this results from Uij 1\ u ij = O. Let us deal
with K 4 : exchanging (i,j) and (k,), we get
K4 = I: AkAij uk 1\ uij'
i<j, k<
XIX.2 Braid Group Representations from Monodromy 453
_f
PROOF. Let i,j, k be distinct indices. Then
(dzi - dz j ) 1\ (dz j - dz k )
u ij 1\ ujk + ujk 1\ Uki + Uki 1\ Uij - (
Zi - Zj
)(
Zj - zk
)
where the symbol 1 means that we take the sum of the term under the
integral with the other two obtained by circular permutations of the indices.
We have
l.h.s
o
Let us resume the proof of Proposition 2.1. We still have to prove that
K3 = o. We break the sum K3 into three smaller pieces K3 = K5 + K6 + K 7
The first piece is
K5 = L AijAik u ij 1\ uik
i<jopk
K5 = L [A ij , Aikl u ij 1\ uik
i<j<k
Similarly,
L [A ij , Ajkl u ij 1\ ujk
i<j<k
454 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
Now, if the differential system (2.3) is invariant under the action of Sn'
then the connection 'V = d - r descends to a connection on E. If Rela-
tions (2.1-2.2) are satisfied, then it has a monodromy representation on the
fundamental group of Xn which, by Proposition X.6.14, is the full braid
group Bn'
dw =
h
L t
_'_J-(dzi - dzj)w (KZn)
2nA l:S;i<j:S;n zi - Zj
t
~
= "~ x(1)
r
0 0 x(n)
r
r
where XCi)
r = x r' X(j)
r = Yr and x(k)
r =1 otherwise .
Lemma XIX.3.2. The elements (tijh:S;i<j:S;n induce endomorphisms of
vn satisfying Relations (2.1-2.2).
PROOF. Relations (2.1) hold by definition oft ij . Relations (2.2) follow from
the g-invariance of t. We show this when i = 1, j = 2, and k = 3. We have
s r
by (3.1). D
456 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
Lemma 3.2 and Proposition 2.1 imply that the system (KZ r,) defines a
fiat connection on the trivial bundle over Yn with fibre V 18m and, conse-
quently, determines a monodromy representation
p~z : 'if l (Y,-,) --+ Aut(Vn).
dw = h "6 t
~(dZi - dzj)w. (3.3)
4'ifH ",-z
ls,i#.Js,n r J
It is clear from (3.3) that the system is invariant under the action of the
symmetric group. We thus obtain a monodromy representation
(3.4)
Here we take the point p = (1,2, ... , n) as the base-point of Xn and we
identify the fibre of the bundle E = (Yn x vn) / S n at the point p with
vn.
The main objective of this chapter is to compute the monodromy rep-
resentation p~z as explicitly as possible from the above data. This is a
difficult task. To begin with, let us consider the following special cases.
(a) Case when h = 0: the differential system reduces to dw = 0 which
has constant solutions over Y,-,. The corresponding monodromy is the rep-
resentation of Bn on Vn coming from the action (3.2) of the symmetric
group.
(b) Case when n = 2: the system (KZ 2 ) reduces to
h t
dw =
2'if
H-lzI -Z2 (dz l - dz 2 )w. (3.5)
dw = ht W(8) (3.6)
d8 2
XIX.3 The Knizhnik-Zamolodchikov Equations 457
(3.8)
for all VI' V2 E V. The flip appears in (3.8) as a consequence of the equality
[tij' t k ] = 0 (3.9)
holds for all i, j, k, . We claim that in this case the monodromy action of
En is given for each of the generators 0"1' ... , 0"n-l by
(3.10)
h
were T i ' Hl'd
= 1 V0(i-l) 161 TV V 1'dV0(n-i-l) . 'T'
f(A
.1.0 prove th e cI'
aIm, represent
the generator O"i of the braid group by the loop z( s) = (zl (s), ... , zn (s))
where
(3.11)
From Case (a) we see that p~z is congruent modulo h to the representation
coming from the symmetric group action (3.2).
We next observe that the monodromy is independent of the g-module V.
This again holds in full generality because the system has coefficients in the
tensor powers of U(g). One can prove this using Chen's theory of formal
connections and formal monodromy [Che73] [Che75] [Che77a].
The last remark is the following: in Case (c) above, the monodromy can
be derived from a topological braided bialgebra structure. Indeed, assume
that (U(g)[[h]],~, c) is the trivial topological bialgebra as above. Set R =
e ht / 2 . When t satisfies Relations (3.9), the element R of U(g)[[h]]0U(9)[[h]]
satisfies Relations (XVI. 4. 15-4. 17) with cp = 10101. The proof of this
claim is similar to the one used in XVI.5, Example 1. Therefore,
- - R = eht/2 )
Ag,t = (U(g) [[h]J, ~,c, cp = 1@1@1,
is a topological braided bialgebra whose universal R-matrix is symmetric:
R21 = R. The g-module V extends to the Ag,cmodule V[[h]] defined above.
The universal R-matrix R gives rise to a representation
Corollary XIX.4.4. The tensor functor (a*, id, rpf) is a braided tensor
equivalence from a braided tensor category of topologically free U(g)[[h]]-
modules equipped with associativity constraint induced by <P KZ and braiding
induced by R KZ , to the category U h (g)-Mod fr of XVII.3.
PROOF. Since ch is a counit for ~h' it follows that Ch is a counit for ~h'
Therefore
id = (Ch 129 id)~h = (Ch 129 id)(F'-l~F'),
which means that
(Ch 129 id)~(x) = xrl (5.4)
where = (ch 129 id)(F'). Using Sweedler's sigma notation, we get
Ch(x) ch'(2:x'c(x"))
(x)
c(2: Ch(X')X")
(x)
c(xrl)
c()c(x)c()-l
c(x)
by the counit axiom and (5.4). o
462 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
and
R,2 R;lR'
F" F'(O' 0 O')((Rhb)F';;} F";;} F"21 F~l (a 0 O')(Rh)F,-l F"-l
F" F'(O' 0 O')((Rhb Rh)F,-l F"-l.
Remark 5.4. The only property peculiar to Uh(g) used so far in this proof
is the one stated in Proposition XVII.3.2. Drinfeld [Dri90] actually proves
more: if A is any QUE for a complex semisimple Lie algebra g, then A
is necessarily isomorphic to the gauge transform of a topological braided
quasi-bialgebra of the form (U g[[h]],~, c, <I>, R) where R = R21 = ehO / 2
for some invariant symmetric 2-tensor e
on g. The trivial deformation
e
(U g[[h]],~, c, 10101, 101) corresponds to = 0. This concludes Step 2.
Step 3. Summing up Steps 1 and 2, we see that Drinfeld and Jimbo's
QUE Uh(g) is isomorphic as a topological braided bialgebra to the gauge
transform of a topological braided quasi-bialgebra of the form
Now the QUE Ag,t of Theorem 4.2 is of the same form, except that <I>' may
differ from the element <I>KZ of Ag,t. This discrepancy is taken care of by
Theorem XVIII.S.1 which implies the existence of a gauge transformation
F'" on U(g)0 2 [[h]] such that
G'(z) = h
A (A- + -B)
1 G(z) (6.1)
27r -1 z z-
where G(z) is a formal series in two non-commuting variables A and B
with coefficients which are analytic functions in the complex variable z. As
above, h is a formal parameter.
464 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
(6.4)
XIX.6 Drinfeld's Associator 465
The convergence of P(z) results from the general fact that a formal solution
of a regular singular equation is necessarily convergent. We refer to [Was87J,
11.5 for details.
Similarly, one proves that there exists an analytic function Q(z) defined
in a neighbourhood of 0 such that
(6.5)
when z is close to 1. D
Since Go and G 1 are both non-zero solutions of Equation (6.1), they have
to differ by an invertible element.
(6.6)
(6.7)
466 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
a~hBGo(1- a) a- hA P(a)-la hA
a-hB Q(a) a hB <P a- hA P( a) -la hA .
When a tends to 0, then Q(a) and P(a) tend to l. Consequently, the right-
hand side of the last equation tends to <P. D
(6.8)
n _ 1 ds and D _ 1 ~
-
21fA s -1'
~~o
21fA s 1 -
lim
a--+O
1a
1 a
- D(M)
(6.9)
where the complex numbers T(Pl' ql' ... ,Pk' qk) have been defined and com-
J:-
puted in the Appendix in terms of multiple zeta values. If the monomial M
begins with B or ends with A, then the integral a D(M) diverges as a
(6.10)
XIX.6 Drinfeld's Associator 467
G'(z) = h
21fH
(A -z a + B - ,8)G(z).
z-1
(6.12)
(6.13)
(6.14)
468 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
1+ L
k,?O
CkU k+ 1 v H1 = exp (f ((n) n (un
n=2 n(27rA)
+ v n- (u + v)n)).
(6.16)
From (6.16) we get CkC = cCk and ckO = cOk = -((k + 2)/(27rA)k+2 for
all k 2 O.
(7.1)
where h = h/(2JrH).
Lemma XIX.7.2. If W(Zl' ... ,zn) is a solution of (7.2)! then it also sat-
isfies the relations
and
l:C;i<j:C;n
(i = 0, 1) (7.5)
when IZ2 - zll IZ3 - zll, i.e., when IZ2 - zll/lz3 - zll tends to 0, and
W1 (z l' z 2' z)
3
rv (z 3 - Z2 )ht23(Z3 - z 1 )h(t 12 +t 13 ) (7.7)
when IZ2 - z31 IZ1 - z31 In view of Definitions 6.2 and 7.1, Wo and W 1
are related by
W O(zl,z2,z3) = W 1(Zl,Z2,Z3) <I>KZ' (7.8)
Let us determine the monodromy of (KZ3)' The change of variables (7.3)
has the following property: Zl is close to Z2 if and only if z is close to O.
Similarly, z3 is close to zl or to Z2 if and only if z is close to 00 or to 1
respectively. Now consider the generator (J1 of the braid group B3 with the
parametrization given by (3.11). An immediate computation shows that
(7.9)
XIX.8 Verification of the Axioms 471
dw
ds
and
for all x E g, where ~(2) = (~0 id)~ = (id 0 ~)~. By hypothesis, ~(x)
commutes with the 2-tensor t. By the following special case of Leibniz's
formula, we have
(8.7)
Relation (8.3). We shall be sketchy. For more details, see [Dri89b], Section
3. Recall the solutions Wo and WI of the system (KZ 3) described with their
asymptotic behaviour in (7.5-7.7). By permuting ZI' Z2, Z3 we get four other
solutions W 2 , W 3 , W 4 , W5 of (KZ3)' uniquely determined by the following
asymptotic behaviour:
W 2 (ZI' z2' z3) rv (z2 - Z3)ht23 (z2 - ZI)h(h2+ t ,3) when IZ3 - z21 IZ2 - zll,
W 3 (Zl' Z2, z3) rv (z3 - Zl) ht 13 (z2 - Zl)h(h2+ t 23) when IZ3 - zll IZ2 - zll,
W 4 (zl' Z2, z3) rv (Zl - Z3) ht 13(Z2 - Z3)h(h2+ t 23) when IZ3 - zll IZ2 - z31,
W5 (zl' Z2, z3) rv (Z2 - ZI)ht12 (Z2 - Z3)h(t 13 +t 23 ) when IZI - z21 IZ2 - z31
We observe that W 2 is obtained from WI by exchanging Z2 and z3' Letting
z3 pass in front of z2 following the loop 0"2 of the braid group B3 yields
W1= W 2 e ht23/2 =
W 2R 23' (8.9)
We next remark that W 2 and W3 are solutions of (KZ 3) where t12 and t 13
have been exchanged. Therefore, by definition of <I>, we have
(8.10)
(8.12)
Relation (8.4). One may proceed as for (8.3). An alternate proof consists
in first showing that <1.>321 = <1.>-\ which is done by replacing Z by 1 - Z
in Equation (6.1). Then, as in the proof of Lemma XVIII.8.2, apply the
involution T 13 to Relation (8.3) and use the fact that ~ = ~ op and R = R21
to derive Relation (8.4).
Relation (8.5). In order to prove the "pentagonal" relation we now consider
the system (KZ4)' The following lemma is due to Drinfeld [Dri90], Section
2, to which the reader is referred for a proof.
Lemma XIX.S.l. There exist solutions Xl' X 2, X 3, X 4 and X5 of (KZ 4)
uniquely determined by
X 1 (z l' Z2' Z3' z)
4
rv (z 2 - Z1 )iitl2 (z 3 - z 1 )ii(t l3 +t 23 )(z 4 - z 1 )ii(t,4+ t 24+ t 34) ,
For Xl the sign rv means that there exists an analytic function f (u, v)
such that f(O,O) = 1 and
Xl (Zl , Z2' Z3' Z4) = f (u, v) (Z2 - Zl )iit,2 (Z3 - Zl )ii(t,3+t23) (Z4 - Zl )ii( t'4 +t2d t 34)
where u = (Z2 - Zl)/(z4 - Zl) and v = (z3 - zl)/(z4 - Zl)' The reader will
be able to give a precise meaning to rv in the remaining cases.
The "pentagonal" relation (8.5) is an immediate consequence of the fol-
lowing lemma. This completes the proof of Part (i) of Theorem 4.2.
XIX.S Verification of the Axioms 475
X 4 =X5(id0id0.6.)(<I>-1), X5=Xl(.6.0id0id)(<I>-1).
PROOF. (a) We start with the proof of the first relation Xl = X 2 (<I> 0 1).
Set
V1 ( Zl,Z2'Z3,Z4 ) ~
~
X 1 ( Zl,Z2'Z3,Z4 )( Z4 - Zl )-h(h4+ t 24+ t 34)
and
av
-a =
- '"""' t lj
h L.....- ---V(Zl,z2,z3,z4)
-
+ hV(zl,z2,z3,z4) t14 + t24 + t34
,
Zl jil zl - Zj z4 - Zl
(8.15)
for i = 2,3, (8.16)
and
av
-a - '"""' t4j -
h L.....- ---V(zl,z2,z3,z4) - hV(Zl,Z2,z3,Z4)
=
t14 + t24 + t34
.
Z4 ji4 Z4 - Zj z4 - zl
(8.17)
We set z4 = 00 in Equations (8.15-8.16) (this is possible since the equations
are actually defined on the complex projective line). Then V l (Zl,Z2,z3'00)
and V2(Zl,Z2,z3'00) become solutions of the system (KZ3). Moreover, by
Lemma 8.1, V l (Zl,Z2,z3'00) and V2(Zl,Z2,Z3,00)(<I>-1 (1) have the same
asymptotic behaviour as the solutions Wo and WI of (KZ 3 ) respectively.
By uniqueness of these solutions, we get
and
476 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
(8.18)
for all ZI, z2, z3' Relation (8.18) and Equation (8.17) imply that VI and V2
coincide everywhere.
(b) To prove the second relation of Lemma 8.2, it is enough to check that
the functions U I and U2 coincide when we set
and
(8.20)
and
(8.21)
;:)aT -_ h- (t12 + t I3
+ ~) ( )
T ZI,Z2,Z4 , (8.22)
uZI ZI - Z2 ZI - Z4
(8.23)
XIX.S Verification of the Axioms 477
and
(8.24)
Now,
and
t24 + t34 = (id 6. id)(t23)'
Therefore Equations (8.22-8.24) imply that Tl and T2 are solutions of the
system (KZ 3 ) in which the coefficients tij have been replaced by new coef-
ficients (id 6. id)(t ij ). By the results of Section 7, there exist solutions
Ho and HI of this modified (KZ 3)-system such that
(8.25)
H
o(z l' z 2' z)
4
rv (z
2
- z )h(t12+h3)(Z
1 4
- z )h(t14+t24+t34)
1
when IZ2 - z41 IZI - z41 It follows from this, from Lemma 8.1, and
from the fact that t 23 commutes with t12 + t 13 , t14 + t24 + t 34 , t24 + t34
and with t12 + t 13 + tw that Tl and T2 (id 6. id)(<I-1 have the same
asymptotic behaviours as Ho and HI respectively. Consequently, Tl = Ho
and T2 (id6.id)(<I-1 = HI' Combining these relations with (8.25), we
conclude that Tl and T2 coincide. Therefore,
(8.26)
for all ZI' Z2' z4' Relation (8.26) and Equation (8.20) imply that the func-
tions U 1 and U2 coincide everywhere.
(c) The remaining relations of Lemma 8.2 are proved in a similar fashion:
in the case of the third relation, we send ZI to 00 whereas for the last two,
we have to identify z3 with z4' and ZI with Z2' respectively. The movements
of ZI' z2' z3 and z4 in this proof can be represented as a system of four
particles moving as in Figure 8.2. 0
478 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
Zl Z2 Z3 Z4
Xl
X2
X3
X4
X5
Zl Z2 Z3 Z4 Xl
(8.29)
Relation (8.27) is the translation of (8.5) while Relations (8.28-8.29) cor-
respond to (8.3-8.4) in view of Rij = e h~ij for 1 ~i <j ~ 3 and of
XIX.9 Exercises
1. Let 9 be a semisimple Lie algebra and t E 9 @ 9 be the 2-tensor
(XVII.1.6). Show that [tI2' td =f O.
2. (a) Compute Drinfeld's associator <I>(A, B) in the case that A and
B commute with the commutator [A, B].
(b) Show that <I>(A, B) is the exponential of a Lie series (Hint: prove
that Do <I> = <I> @ <I.
3. Prove that the projector f defined by (6.11) coincides with the con-
volution vB *id *v A where v A and vB are the algebra endomorphisms
of S determined by
5. Let VI' ... ,vn be analytic functions. Consider the differential equation
G'(z) = (t
i=1
Aiui)G(Z) (9.1)
XIX.I0 Notes
The material in Sections 1-2 is standard. For more on the configuration
space X n , see [Aom78] [Hai86] [Koh85].
The equations (KZn) were introduced by Knizhnik and Zamolodchikov
[KZ84] in connection with the Wess-Zumino-Witten model in conformal
field theory.
Theorem 4.1, which is the main result of this chapter, first appeared
in [Koh87] [Koh88]. For the proof we followed Drinfeld [Dri89b] [Dri90]
[Dri89c] closely. As a matter of fact, most results of Sections 4-8 are due
to Drinfeld. There is an exception in Section 6 where Proposition 6.4 is
480 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
(1l.1)
and
(1l.2)
(1l.3 )
(1l.5)
where a runs over all (n, m )-shuffies of the symmetric group Sn+m'
Iterated integrals occur in the solution of certain linear differential equa-
tions. Let us consider an equation of the form
dY
- = A(s)Y(s) (11.6)
ds
where Y (s) is a differentiable function defined on the real interval [a, b],
with values in the endomorphism ring of some complex vector space and
XIX.ll Appendix. Iterated Integrals 481
(11. 7)
by
Qp(s) = r
Jtl.p(a;s)
A(SI)A(S2) ... A(sp)ds I ds 2 ... ds p. (11.9)
dY = t Aj Y(s)
(11.10)
ds j=1 S - aj
where AI' ... , An are constant linear endomorphisms and aI' ... , an are dis-
tinct complex numbers lying outside the real interval [a, b]. By (11.1-11.2)
and (11.7-11.9) the unique formal solution Y(s) of (11.10) with Y(a) = id
is given by the formal series
where the complex functions La (all' ... , a jr Is) are defined as the following
iterated integrals
(11.12)
Functions of this kind already appeared in [Poi84], III and were investigated
at length by Lappo-Danilevsky in [Lap53], Memoire II under the name
"hyperlogarithms" .
We now concentrate on the hyperlogarithms built on the particular 1-
forms
no = 27ry11-11dsS- and ~fIIl _- 27ry11-11 ~
s- 1
. (11.13)
482 Chapter XIX. Monodromy of the Knizhnik-Zamolodchikov Equations
(11.14a)
I
and
b
Dk _ 1 (10 1- b) k (11.14b)
a 1 - (27rR)kk! g 1- a
(
T Pl,Ql"",Pk,qk
) = Jrl "p, "q,
HO HI
"Pk "qk
... HO HI (11.15)
o
where PI' Ql' ... , Pk' Qk are integers> O. We shall now compute the iterated
integrals (11.15) in terms of series reminiscent of Riemann's zeta function.
To this end, we introduce the convergent series
(11.16)
where i 1 , ... , i k are positive integers, x is a real number such that 0 <
x < 1, and m 1 , ... , mk run over the set of positive integers. The special
case L( n; x) is the n- th polylogarithm which appears in number theory,
geometry and algebraic K-theory. When n = 1, we have
L(I;x) = L -xm r
= -log(1- x) = -27rvCI In
1
D1 (11.17)
O<mm 0
(
Li 1, ... ,i k- 1,i k ;X)=Jo
r L(i ,,i
1 k _ 1 ,i k
s
-l;s)
ds. (11.18)
If i k = 1 we have
dL(i 1 ,,i k _ 1 ,I;x)
dx
XIX.ll Appendix. Iterated Integrals 483
dL(i j , i k- I, 1; x)
dx
L(il,,ik_I;X)
I-x
It follows that
l
x L(il,,ik_I;S)
L(i l ,,i k_ l ,l;x)= 0 -----'---ds.
l-s
(11.19)
(11.20)
The special case (( i l ) coincides with the value of Riemann's zeta function
at the positive integer i l . An easy induction using (11.15-11.20) expresses
the mixed iterated integrals T(PI' ql, ... , Pb qk) in terms of multiple zeta
values. To be precise, we get
( _l)Ql+'+Qk
T(PI' ql" .. ,Pk' qk) = (21TR)Pl +ql +"'+Pk+qk
(11.23)
with similar restrictions on the singularities. Singular knots and links are
represented by planar singular knot and singular link diagrams defined in
the same way as ordinary knot and link diagrams are (see X.3). One also
has an obvious notion of isotopy generalizing the one introduced in X.1.
Now any double point in a singular link diagram can be "desingularized"
by locally replacing the pattern X formed by the double point and the two
downwards oriented branches, by the patterns X+ and X_ described in
X.4. This observation allows us to extend any isotopy invariant of links to
any singular link. Indeed, let P be such an invariant with values in some
complex vector space V. Then the rule
(1.1)
defines the invariant P on the set of isotopy classes of all singular links
with one double point. Here L is a link diagram with one double point and
the ordinary link diagrams L+ and L_ are obtained from L by replacing
a neighbourhood X of the double point by X+ and X_, respectively. By
induction on the number of double points we may extend P to an isotopy
invariant for all singular links (with a finite number of double points).
Definition XX. I. I. Let m be a non-negative integer. An isotopy invari-
ant of oriented links is an invariant of degree ::; m if it vanishes on all
singular links with more than m double points.
There is a similar definition for framed links. Observe that an invari-
ant P is of degree 0 if and only if we have P(L+) = P(L_) on all link
diagrams, which means that the invariant P does not distinguish between
under crossings and overcrossings. Therefore it depends only on the number
of connected components of the link. In padicular, P is constant on the
space of all ordinary knots.
The first question we wish to address is the following: Are there any
non-trivial examples of finite-degree invariants of higher degree? Before
we answer this question, let us say that a (framed) link isotopy invariant
P(L) = L:m>O Pm(L) hm with values in V[[h]J, where V is a complex
vector space, IS of finite type if, for all m 2: 0, the V -valued invariant Pm is
of degree::; m. We now state the main source of invariants of finite type.
Proposition XX.I.2. Let P be a link isotopy invariant with values in
V[[h]] where V is a complex vector space. If, for any link L, we have
P(L+) == P(L_) modulo h at any crossing point of a link diagram of L,
then P is of finite type.
One similarly defines the vector space vt) of complex-valued framed knot
invariants of degree :S m. In this section we shall show that the spaces v(m)
and vjr;:) are all finite-dimensional and give a combinatorial description of
the quotients v(m)/v(m-l) and V(m)/V(m-1)
fT fT
To this end, we need the notion of a chord diagram on a circle: it is a
finite set of unordered pairs of distinct points on the circle considered up to
homeomorphisms preserving the orientation. To specify a pair one draws
a dashed line, called a chord, between the two points. Given 2m distinct
points there are
(2m - I)!! = 1 35 .. (2m - 1)
different ways to pair them. Indeed, given one point among 2m, we may
pair it with (2m - 1) points. Take another point among the remaining
(2m - 2); it may be paired to (2m - 3) other points, etc.
There is a relationship between invariants of finite degree and chord
diagrams on a circle which we explain now. Let D be a chord diagram
on the circle with m chords (Le., with 2m points paired two by two). By
an embedding of D into R3 we mean any singular knot f : Sl ....... R3 with
exactly m double points such that f(s) = f(s') if and only if s = s' or sand
s' are the two endpoints of a chord in D. There always exists an embedding
K D of D. If K'v is another embedding of D, then it can be obtained from
K D by a series of operations consisting in replacing an undercrossing by an
overcrossing and vice-versa. Suppose we are given a complex-valued knot
invariant P of degree :S m. Since P vanishes on singular knots with at least
m+ 1 double points, P remains constant by Rule (1.1) under the operations
transforming K D into K'v, which means that P( K D) is independent of the
embedding of D chosen to compute it.
Define Em as the complex vector space with a basis given by all chord
diagrams on the circle with m chords. The dimension of Em is finite and
:S (2m - I)!!. Then the evaluation of an invariant of degree :S m on an
embedding of a chord diagram with m chords gives rise to a pairing
< , > : v(m) Em ....... C. (2.2)
Suppose that < P, D > = 0 for all chord diagram with m chords. Since any
singular knot with m double points can be represented as an embedding
of a chord diagram, we see that P vanishes on all singular knots with m
double points, which means that P is an invariant of degree :S m - 1.
Consequently, the map P 1-+ < P, - > induces an injection
Y m : v(m) /v(m-l) ....... Hom(Em, C) (2.3)
of the quotient v(m) /v(m-1) into a finite-dimensional space. A similar
argument works for vjr;:) /Vjr;:-l). We get the following result.
PROOF. We have already noted that knot invariants of degree 0 are con-
stant. Therefore, V(O) = vj~ = C, which proves the assertion for m = O.
The rest follows by an easy induction on m using the injection (2.3). 0
Actually, the proof of Proposition 2.1 shows that the dimensions of v(m)
and vjr;) are bounded by 1 + 2::Z"=1 (2m - 1)!!.
What we next aim to, is to restrict the size of the image of the map
Ym . More precisely, we shall show that any linear form in the image of Ym
satisfies an important four-term relation. Let D be a chord diagram with
m - 2 chords. Consider the four pictures in Figure 2.1 involving each 2
chords.
v v v v v v v v v v v v
(2.5)
The argument in Part (b) of the proof above does not work for framed
knot invariants since K+ and K_ are not necessarily isotopic as framed
knots.
The relations in Proposition 2.2 appear as universal relations satisfied by
all invariants of finite degree. We may wonder whether there are more such
relations. The answer is negative. In order to make this more precise, we
define a vector space Am as the quotient of Em by the subspace generated
by all elements of the form
(2.6)
v(m)/v(m-l) ---+
]1']1'
Hom (A m' C) and v(m) /v(m-l) ---+ Hom(Am, C)
are isomorphisms.
(2.7)
m;:,D m;:,D
490 Chapter XX. Postlude. A Universal Knot Invariant
(2.8)
The invariant Z(K) and its image Z(K) in A are called the Kontse-
vich universal (fmmed) knot invariants. We now use Z(K) to prove Theo-
rem 2.3.
Proof of Theorem 2.3. We assume the existence of such an invariant Z. Let
w be a linear form on Am' By Property (i) above, Pw(K) = w(Zm(K)) is
an element of vjr;l.Define a map Xm from Hom(Am, C) to vjr;l/V)';-ll
by composing w 1-+ Pw with the projection onto vjr;l /vjr;-ll. By (2.10),
we have
for a unique family of linear maps (wi: Ai ----+ C)O<i<m' Consequently, for
any framed knot invariant P = Lm::::o Pmh m of finite type, with values in
XX.3 Algebra Structures on Chord Diagrams 491
the formal series algebra C[[h]], there exists a unique linear map w from A
to C[[h]] such that
P(K) = w(Z(K))
for all framed knots K. The bijection set up by Z(K) between framed
knot isotopy invariants of finite type and linear maps defined on A justifies
the qualifier "universal" for Z(K). We have a similar formulation for knot
invariants after replacing A by A: and Z(K) by Z(K). In particular, any
quantum group invariant can be obtained in this way (we shall give details
in Section 8).
Kontsevich's original definition ofthe universal knot invariant Z(K) used
complicated multiple integrals depending on the realization of the knot as
a smooth curve in the three-dimensional space. In Section 7 we shall give a
combinatorial construction of Z(K) using a planar diagram of the knot and
category theory in the spirit of what we did in Chapters X, XII, XIV.5.1
and XVII.3. The combinatorial construction is due to Cartier [Car93], Le-
Murakami [LM93c] and Piunikhin [Piu93].
where Em(T) is spanned by all chord diagrams with m chords. The sub-
space Eo(T) is the one-dimensional subspace spanned by the unique chord
diagram without chords. If f : T ---+ T' is a homeomorphism of tangles,
then f sends any chord diagram on T to a chord diagram on T', thus in-
ducing an isomorphism E(T) ~ E(T'). In particular, since any tangle is
homeomorphic to an "unknotted" tangle, the isomorphism class of E(T)
depends only on the number of circles and segments composing T.
Let T and T' be tangles such that s(T) = b(T') in the notation of X.5
and XII.2. Under this condition, the composition ToT' is defined. Placing
a chord diagram of T on top of a chord diagram of T', we get a chord
492 Chapter XX. Postlude. A Universal Knot Invariant
A(T) Q9 A(T') ----7 A(T 0 T') and A(T) Q9 A(T') ----7 A(T 0 T'), (3.5)
defined when s(T) = b(T'). The maps (3.5) preserve the gradings.
Next put a graded algebra structure on the vector space A = EBm>O Am'
Consider the braid In with n > 0 vertical segments oriented downwards
(defined in X.6). In the tangle category, In is the identity of the sequence
consisting of n +-signs. Since In 0 In = In the maps (3.2) and (3.5) yield
algebra structures on E(1n), A(I.,J and A(1rJ whose units are the chord
diagrams without chords.
We use these algebra structures to produce a family of elements of A(lrJ
satisfying the infinitesimal braid group relations (XIX.2.1-2.2). For integers
1 <:::: i =f j <:::: n, let tij be the unique chord diagram on In with a single
chord between the i-th and the j-th strands. We have t ji = tiJ by definition.
Using the algebra structure of E(lrJ we also have
(3.6)
in the quotient algebras A(1n) and A(ln) when i,j, k are distinct integers.
Consequently, the classes of the elements (tijh<;i<j<;n satisfy the infinites-
imal braid group relations in A(1n) and in A(1n).
XX.3 Algebra Structures on Chord Diagrams 493
The algebras A(1n) and A(ln) also have bialgebra structures. The co-
multiplication ~ is given by the formula
where D' runs over all sub diagrams of D including the chordless diagram
and D" is the subdiagram complementary to D' in D. The co unit is zero
on all chord diagrams with at least one chord and is 1 on the chordless
diagram 0. The reader may check that the comultiplication and the counit
are well-defined on A(IT/) and A(1n) and satisfy all the required axioms.
Observe that ~ is cocommutative. Actually, these bialgebras are Hopf al-
gebras as are all graded bialgebras whose zero-th part is equal to C. The
antipode S is defined inductively on the number of chords by S(0) = 0 and
We now consider the special case n = 1 and denote A(11) and A(1 1) by
AU) and AU) respectively. The following lemma holds in A(l).
(3.10)
The proof of the lemma now follows from (3.10) and from the equality
1 2 345 6 7
1 1 123 5 8
A final observation is in order: denote by C the image in A of the unique
chord diagram with one single chord and by (C) the two-sided ideal it
generates. We have A = Aj(C) and A ~ A[C].
and
tu,vw = tu,v 0 id w + (au,v 0id w )-1 0 (id v 0tu,w) 0 (au,v 0id w ) (4.3)
for all objects U, V, W in S.
A symmetric category as above equipped with an infinitesimal braiding is
called an infinitesimal symmetric category.
Observe that in view of (4.2), Relation (4.3) is equivalent to
(4.5)
PROOF. Part (a) follows by direct checking. To prove Part (b) we proceed as
in the proof of Proposition XIII. 1.4. The functoriality of the infinitesimal
braiding forces it to be of the form (4.7) with t = t H H (1 @ 1). The H-
linearity of the infinitesimal braiding implies that [,6. (~), t] = 0 for all a
in H. Conditions (4.2) and (4.3) yield t21 = t and (id@ ,6.)(t) = t12 + t 13
respectively. The fact that t belongs to the subspace generated by primitive
elements follows from an argument already used in the proof of Proposition
XVI.5.2. 0
U@V@W t u 09 V W U@V@W
1 1
)
tu.vidw tu,vidw
U @ V @ W tu 09 V.w) u@v@w
of the complex vector space AhoT(T) for some braid T where AhoT(T) has
the same definition as A(T) (see Section 3), except that we allow only hor-
izontal chords. The source [resp. the target] of such a chord diagram is the
sequence s(T) [resp. the sequence b(T)] defined in X.5. The composition of
morphisms is given by the map (3.5). The identity of an integer n is the
chord less diagram on the braid In (defined in X.6).
We put the same tensor product on AB as the one we put on the braid
category, namely, we have n @ m = n + m on objects while the tensor
product of morphisms is defined by placing chord diagrams side by side.
The tensor product is well-defined and strictly associative with unit I = O.
The braiding (XIII.2.1) of the braid category induces a braiding on the
category AB: it suffices to take the chord less diagrams on the correspond-
ing braids. Since we are considering braid chord diagrams up to homeo-
morphisms, we see that this braiding is symmetric in AB although it is not
in the braid category.
Given objects n, m of AB, define an endomorphism tn,m of n@m = n+m
as follows. If n or m = 0, set tn,m = O. Otherwise, set
n m
t n, Tn = ""'
L..-t ""'
L-t ti,n+j (5.1)
i=l j=l
where tij is the chord diagram (already defined in Section 3) with a unique
chord between the i-th and the j-th strands.
If0id~ (5.2)
tn =
n@m ~
and
498 Chapter XX. Postlude. A Universal Knot Invariant
where (Jij = id v0i (JV 0 (J-i-l) v id v0 (n-j). We have to check the relations
defining the morphisms of AB, including (3.6-3.7). Relation (3.6) is clear
while Relation (3.7) follows from (4.8). 0
Using Proposition 5.2, one may derive an equivalence between the cat-
egory S and a category of braided tensor functors preserving infinitesimal
braidings from AB to S.
Z : Bn --+ Bn x (1 + L A~r(1r,)hm)
7n:;,1
(6.4)
when 2 ::; i ::; n - 1. The associativity isomorphisms aV3(i-l),v,v have
to be computed from the Drinfeld series <P using (6.1) and (4.3-4.4). The
composition of Z with the projection onto Bn is the surjection sending each
braid to its permutation. In the next section, we shall extend the map Z
to all tangles.
Let 9 be a semisimple Lie algebra and t E g0g be the invariant symmetric
2-tensor given by (XVII. 1.6). Consider the infinitesimal symmetric category
U(g)-Mod f . We can reformulate precisely Drinfeld's Theorem XIX.4.3 and
Corollary XIX.4.4 as follows.
Corollary XX.6.2. In case <P = <P KZ , there is a braided tensor equiva-
lence between the braided tensor category U h (g)-Mod fT of XVII.3 and the
braided tensor category (U(g)-Modf)[[hl].
Theorem XX.7.1. Under these hypotheses, the strict braided tensor cat-
egory S[[hW tr is a ribbon category with twist Bv given by
(7.1)
and with left duality defined as follows: for any object V the dual object V*
is the same as in the category S; the structure maps bv and d v are defined
by
bv = b~ and d v = d~ 0 (AV: c>9 id v ) (7.2)
where AV* is the automorphism of V* defined by
vector spaces. This algebra is bigraded by the number of chords and the
number of connected components of the link. We have
preserving the duality and the twist, sending the object (+) of the category
R of ribbons to the object (+) of A[[h]]str. Consequently, the functor Z is
the identity on objects. The restriction of Z to braids is the morphism
defined by (6.3-6.4).
Let K be a framed link. It can be viewed as an endomorphism of the unit
object in the category R. Its image Z(K) is an isotopy invariant living in
PROOF. One proceeds as for Proposition 5.2. The main difference lies in the
existence of general chord diagrams in A. In order to show that Relation
(8.3) determines Fv on any chord diagram, we observe that any chord may
be arranged so as to be horizontal after possibly adding some maxima and
minima to the diagram. D
WD = Xj,XhxhYhXj4Yj,Yj4Xj5YhYj5
(here m = 5). Then the element CD is by definition
CD = (_l)m "'"
~
XJ, x J2 XJ3 y.J2 x J4 y.J' y.J4 X.15 y.J3 y..15' (8.4)
]1,,]5
XX.8 Recovering Quantum Group Invariants 503
i,j,k
(2: XjXkYjYk)V
j,k
CDv.
The third equality follows from S(YjXk) = S(Xk)S(Yj) = (-1)2 xkyj , which
holds because x k and Yj are primitive elements of H. 0
Theorem XX.8.3. Under the previous hypotheses, for all framed knots K
we have
Qg,v(K) = dimq(V) L
Wg,v(Zm(K))hm (8.8)
m:;,D
(8.10)
where the quantum trace and dimension are taken first in the ribbon cat-
egory (U(g)-Modj)[[hW tr , then in the equivalent category Uh(g)-Mod lr .
Combining the last set of equalities with (8.10) yields Theorem 8.3. D
XX.lO Notes 505
XX.9 Exercises
1. Find all primitive elements of degree :'S 4 in the Hopf algebra A of
Section 3.
2. Let O(N) be the framed trivial knot whose framing twists the knot by
27rN. Compute its Kontsevich invariant Z(O(N)) modulo h4. (Hint:
use Corollary XIX.6.5.)
XX.I0 Notes
The concept of a knot invariant of finite degree (also called "Vassiliev in-
variant" in the literature) was introduced by Gusarov [Gus91] and Vas-
siliev [Vas90] [Vas92] around 1989-90. Vassiliev's approch was based on
the theory of singularities. Soon after, a number of mathematicians made
substantial contributions to this new theory such as D. Bar-Natan, J. Bir-
man, P. Cartier, M. Kontsevich, Le T.Q.T., X.S. Lin, J. Murakami, S. Pi-
unikhin, T. Stanford (see [BN92] [Bir93] [BL93] [Car93] [Kon93] [LM93b]
[LM93a] [LM93c] [Lin91] [Piu92] [Piu93] [Sta92] [Sta93]). One will find a
review of their results in [Vog93]. A major step forward was undertaken by
Kontsevich who constructed the universal knot invariant Z(K) and proved
Theorem 2.3. Kontsevich's definition of Z(K) used complicated multiple
integrals. It was proved by Cartier [Car93], Le-Murakami [LM93c] and Pi-
unikhin [Piu93] that it could be defined in a simpler way using tangle
diagrams. Theorems 6.1 and 7.1 are due to Cartier [Car93].
The contents of Sections 5 and 8 seem to be new. For a generalization,
see [KT94].
References
[Ada56] J.F. Adams. On the cobar construction. Pmc. Nat. Acad. Sci.
USA 42 (1956), 409-412.
[Art25] E. Artin. Theorie der Zopfe. Abh. Math. Sem. Univ. Hamburg
4 (1925), 47-72.
[Che77b] K.T. Chen. Iterated path integrals. Bull. Amer. Math. Soc.
83 (1977),831-879.
[Coh7l] P.M. Cohn. Free Rings and their Relations, volume 2 of Lon-
don Math. Soc. Monographs. Academic Press, London, New
York, 1971.
[EM 53] S. Eilenberg and S. Mac Lane. On HJTr, n), 1. Ann. of Math.
(2) 58 (1953), 55-106.
[Enr92] B. Enriquez. Rational forms for twistings of enveloping al-
gebras of simple Lie algebras. Lett. Math. Phys. 25 (1992),
111-120.
[FY89] P.J. Freyd and D.N. Yetter. Braided compact closed categories
with applications to low-dimensional topology. Adv. Math. 77
(1989), 156-182. MR 91c:57019.
[Lin91] x.S. Lin. Vertex models, quantum groups and Vassiliev's knot
invariants. Columbia University, Preprint, 1991.
[Ser65] J.-P. Serre. Lie Algebras and Lie Groups. W.A. Benjamin,
Inc., New York, Amsterdam, 1965.
References 519
[Yan67] C.N. Yang. Some exact results for the many-body problem in
one dimension with repulsive delta-function interaction. Phys.
Rev. Lett. 19 (1967), 1312-1315.
[Yet88] D.N. Yetter. Markov algebras. In Braids (Santa Cruz, 1986),
volume 78 of Contemp. Math., pages 705-730. Amer. Math.
Soc., Providence, RI, 1988.
Mq(2), 78, 83, 89, 150, 194,418 X n , 267, 269, 454, 479
Mq(n), 91, 198 Y", 267, 270, 452, 454, 456
[n], 121 YB(C), 324
(n)q,74 Z(K), 490, 505
P n , 454 zn(fl, M), 421
Pn , 454 Zq' 133
Prim(C),48 Z(C),330
Prim(H), 495 Z[Xc(G)], 360
CJg,v, 411, 418, 486, 503
R, 313, 358, 411, 501
abelian Lie algebra, 94, 95, 117
Rh , 409, 412, 458, 459
acyclic chain complex, 447
R KZ , 471
adjoint corepresentation, 2:36
Sn' 454, 456
adjoint functors, 279, 291, 400
sl(2), 99, 405, 412
adjoint representation, 99, 107,208,
SL(2), 11,33,57,65, 114
404, 429
s[(n),94
of a Lie algebra, 208
SL p,q(2),91
affine
SLq(2), 83, 87, 90, 153, 196, 366,
line, 8
418
plane, 8, 65, 109
SLq(n), 91, 163
Alexander polynomial, 246, 486
T, 299, 305, 312, 360
T'(V),67' . algebra, 3, 39
of differential operators, 21
Tens(C, D), 288
of pseudo-differential opera-
Tensstr(C, D), 288
tors, 21
tr q' 354, 363
topologically generated by gen-
U~, 125, 139, 143
erators and relations, 407
U(s[(2)), 99, 109, 114, 118
antilinear, 86
Uh , 412
antipode, 50, 52, 55, 58, 84, 197
Uh(g), 407, 419, 430, 458
antisymmetric tensors, 37
Uh (s[(2)), 412, 419
antisymmetrization, 436
Uq , 122, 140, 412
associativity constraint, 281, 369,
Uq , 136, 224, 230, 238, 364 377, 460, 499
Uq (g), 138, 419 autonomous tensor category, 347,
Uq (s[(2)), 121, 122, 138, 140, 153, 353, 380
162, 168, 223, 412
Uq(s[(f + 1)), 139
Uq(s[(n)), 163 bar complex, 437, 447
V(A), 129, 138 basic hypergeometric series, 90
V(n), 102, 118, 417 basis of a free module, 26
Vn ,417 bialgebr~ 46, 58, 83, 285, 493
symmetrization trace, 21
map, 98 of an endomorphism, 31
of a universal R-matrix, 462 transpose, 343
symmetry, 320 transposition, 30
trefoil knot, 244, 256, 313, 361
tangle, 257, 299 triangle axiom, 282, 293, 369
category, 300, 306, 312, 360 trivial
diagram, 260, 302 comodule, 63
projection, 260 knot, 244, 312
target of a morphism, 276 link, 244
tensor module, 58, 59, 98, 102
algebra, 34, 47, 56 topological
category, 282, 293, 368 bialgebra, 457
coalgebra, 67 braided quasi-bialgebra, 394
equivalence, 288, 375 twist, 349, 359, 361, 366, 411, 500
equivalent categories, 288 two-variable Jones polynomial, 253
functor, 287, 375
product under crossing , 246, 260
algebra, 393 unit of a tensor category, 282, 284,
of algebras, 32 305, 316
of bialgebras, 207 universal
of coalgebras, 42, 45, 66 R-form, 185, 191, 195-197,365
of comodules, 63 R-matrix, 173, 175, 177, 179,
of linear maps, 26 198, 216, 220, 230, 371,
of modules, 57, 98 376, 393, 398, 409, 412,
of vector spaces, 23, 280 413, 419, 458, 459
theorem of Milnor and Moore, 494
topological Vassiliev invariant, 505
algebra, 392, 406, 424 Verma module, 130, 138
bialgebra, 393, 458 vertex, 242
braided quasi-bialgebra, 393,
395, 460, 469, 471 weight, 101, 127
crossed bimodule, 395, 414 Weyl algebra, 18
module, 394 Whitehead lemmas, 117, 429
quasi-bialgebra, 393 word,295
tensor product, 390, 410
Yang-Baxter
topologically free
equation, 167, 171, 178, 179,
algebra, 407
185, 191, 196, 198, 234,
module, 388, 390, 400, 409,
266, 306, 311, 317, 395,
418
418
torsion free module, 390
operator, 323
tortile tensor category, 366
Graduate Texts in Mathematics