118
ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS By Justin M. Allman A Thesis Submitted to the Graduate Faculty of WAKE FOREST UNIVERSITY in Partial Fulfillment of the Requirements for the Degree of MASTER OF ARTS in the Department of Mathematics May 2009 Winston-Salem, North Carolina Approved By: James Kuzmanovich, Ph.D., Advisor Examining Committee: Ellen Kirkman, Ph.D., Chairperson Hugh Howards, Ph.D. Stephen Robinson, Ph.D

ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

  • Upload
    others

  • View
    10

  • Download
    0

Embed Size (px)

Citation preview

Page 1: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE,NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

By

Justin M. Allman

A Thesis Submitted to the Graduate Faculty of

WAKE FOREST UNIVERSITY

in Partial Fulfillment of the Requirements

for the Degree of

MASTER OF ARTS

in the Department of Mathematics

May 2009

Winston-Salem, North Carolina

Approved By:

James Kuzmanovich, Ph.D., Advisor

Examining Committee:

Ellen Kirkman, Ph.D., Chairperson

Hugh Howards, Ph.D.

Stephen Robinson, Ph.D

Page 2: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Table of Contents

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

Chapter 1 Algebras and coalgebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1 Definitions and examples . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Computation in coalgebras, sigma notation . . . . . . . . . . . . . . . 6

1.3 Morphisms, coideals, quotient structures . . . . . . . . . . . . . . . . 11

1.4 Duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

1.5 Group-like and primitive elements . . . . . . . . . . . . . . . . . . . . 25

Chapter 2 Bialgebras and Hopf algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2.1 Bialgebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2.2 Convolution, antipodes, and Hopf algebras . . . . . . . . . . . . . . . 33

2.3 Some examples of Hopf algebras . . . . . . . . . . . . . . . . . . . . . 42

2.4 kG, (kG)∗, and (co)commutativity . . . . . . . . . . . . . . . . . . . 46

2.5 Calculation in the dual . . . . . . . . . . . . . . . . . . . . . . . . . . 52

Chapter 3 Hopf algebra actions on rings . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

3.1 Definitions and Examples . . . . . . . . . . . . . . . . . . . . . . . . 61

3.2 Actions of the Taft algebras . . . . . . . . . . . . . . . . . . . . . . . 70

3.2.1 Actions on k[u, v] . . . . . . . . . . . . . . . . . . . . . . . . . 70

3.2.2 Actions on k[u1, . . . , ur] for r ≥ 3 . . . . . . . . . . . . . . . . 73

Chapter 4 Reflection groups and Hopf algebras . . . . . . . . . . . . . . . . . . . . 79

4.1 Reflection groups, integrals, and trace functions . . . . . . . . . . . . 79

4.2 Semisimple Hopf algebras of dimension 8 . . . . . . . . . . . . . . . . 83

4.2.1 Actions on k[u1, . . . , un] . . . . . . . . . . . . . . . . . . . . . 84

4.2.2 Actions on kq[u, v] . . . . . . . . . . . . . . . . . . . . . . . . 86

4.3 Semisimple Hopf algebras of dimension 12 . . . . . . . . . . . . . . . 92

4.3.1 Actions of A± on k[x1, . . . , xn] . . . . . . . . . . . . . . . . . . 96

4.3.2 Actions of A+ on kq[x, y] . . . . . . . . . . . . . . . . . . . . . 99

Appendix A Maple Worksheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

i

Page 3: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

ii

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

Vita . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

Page 4: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Acknowledgments

I would like to thank Kate for her love and support, as well as her evolving admi-ration for mathematics. I owe much gratitude to Jim Kuzmanovich, who introducedme to the subject of algebra and instilled me with many of his mathematical “bi-ases.” To the members of my thesis committee, I appreciate your careful reading ofthis document and confess you may not have been adequately warned of its length. Iwould also like to thank some of my teachers: Stephen Robinson and Edward Allen,whose willingness to challenge future high school teachers exposed me to rich andrigorous mathematics; and Leah McCoy, who guided me to deeper understanding ofthe discipline by teaching me to reflect on how others learn mathematics. Finally, Iwould like to thank my family, especially my mother, who taught me how to constructa Mobius band and add polynomial expressions.

iii

Page 5: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Abstract

Justin M. Allman

ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE,

NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Thesis under the direction of James Kuzmanovich, Ph.D.

In 1954, Shephard and Todd [20] showed that if A is a polynomial ring and G isa finite group acting as automorphisms on A, then the ring of invariants AG = {a ∈A | g · a = a, ∀g ∈ G} is again a polynomial ring exactly when G is generated byreflections. The major goal of this thesis is the computation of several examples enroute to a conjecture for an analogous result regarding the ring of invariants for someclass of “nice” algebras under finite dimensional Hopf algebra actions.

We begin with an introduction to the general study of Hopf algebras and theirbasic properties, then explain why they are a natural choice to generalize the actionof finite groups on rings. We then show that in order to generalize existing theories,we must consider actions of “nontrivial” Hopf algebras, in particular, those that arenot isomorphic to group rings or their duals. We compute several examples of suchactions, and in particular, we prove that there are no actions of nontrivial semisimpleHopf algebras with dimension less than or equal to 15 on polynomial algebras.

iv

Page 6: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Chapter 1: Algebras and coalgebras

Much of the material in the introductory chapters, 1 and 2, covering basic con-

structions, results, and examples of algebras, coalgebras, bialgebras and Hopf alge-

bras, is an amalgam of material found in Dascalescu, et. al. [4], Montgomery [18],

and Sweedler [21]. In many cases, we expand on these texts, either providing simpler

proofs which emphasize only what is necessary for the purpose of this thesis, or give

explicit calculation and verification for much of the “folklore” of the subject.

1.1 Definitions and examples

Let k be a field. All vector spaces and tensor products are assumed to be over k.

Throughout we will assume the reader to be familiar with concepts of the tensor prod-

uct of two vector spaces and many properties of algebras over the field k. At certain

points, we will restrict our discussion to algebraically closed fields of characteristic

zero, and in fact, simply to C. We begin by presenting the traditional definition of a

k-algebra.

Definition 1.1.1. A unitary ring A becomes a k-algebra with the structure map φ :

k → A, a unitary ring morphism, such that the image of φ is contained in the center

of A.

In order to dualize the notion of an algebra, we give an equivalent definition using

commutative diagrams.

Definition 1.1.2. A k-algebra is a triple (A,M, u) where A is a k-vector space, and

M : A ⊗ A → A, u : k → A are k-linear maps such that the following diagrams are

1

Page 7: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

2

commutative:

A⊗ A⊗ AM ⊗ I

- A⊗ A

A⊗ A

I ⊗M

?

M- A

M

?

(1.1)

A⊗ A

k ⊗ A -

u⊗I-

A

M

?� A⊗ k

I ⊗u (1.2)

We have used I to denote the identity map on A, while the unadorned arrows are the

canonical isomorphisms given by scalar multiplication. The maps M and u are called

the multiplication and unit, respectively, of the algebra A.

Remark 1.1.3. Note that diagram (1.1) ensures that the multiplication of A is

associative. Also, using diagram (1.2), if λ ∈ k such that u(λ) = 0, then for any

a ∈ A, we have

M ◦ (u⊗ I)(λ⊗ a) = M(0⊗ a) = 0.

Thus, by the commutativity of (1.2), λa = 0 for all a ∈ A and therefore u is injective.

Also, since

M ◦ (u⊗ I)(1k ⊗ a) = u(1k)a = a, and

M ◦ (I ⊗ u)(a⊗ 1k) = au(1k) = a,

we see that u(1k) = 1A. Note that we have shown the statement u(1k) = 1A is

equivalent to the commutativity of (1.2) and hence u is an embedding of the field k

into A. Thus, u(k) ⊆ Z(A).

Page 8: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

3

Therefore, Definitions 1.1.1 and 1.1.2 are equivalent. The map M endows A with

the structure of a unitary ring with 1A = u(1k). The role of the map φ is simply

played by u. Conversely, if we have a unitary ring with structure map φ, we define

the map M by x⊗ y 7→ xy and let u = φ. �

Now, simply by reversing the arrows in Definition 1.1.2, we obtain the dual structure

of a k-coalgebra.

Definition 1.1.4. A k-coalgebra is a triple (C,∆, ε) where C is a k-vector space,

and ∆ : C → C ⊗ C, ε : C → k are morphisms of k-vector spaces such that the

following diagrams are commutative:

C∆

- C ⊗ C

C ⊗ C

?

∆⊗ I- C ⊗ C ⊗ C

I ⊗∆

?

(1.3)

k ⊗ C �1⊗

C⊗1- C ⊗ k

C ⊗ C

?I⊗ε

-�

ε⊗I

(1.4)

The maps ∆ and ε are called the comultiplication and counit, respectively, of the

coalgebra C. Alternatively, we sometimes refer to the counit as an augmentation

map. We call the commutativity of diagram (1.3) coassociativity.

Now we give several examples of algebras and coalgebras, the second of which, the

group algebra, will be our prototypical example throughout this text.

Page 9: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

4

Example 1.1.5 (The Group-like Coalgebra). Let S be a nonempty set. Denote by

kS the vector space formed using S as a basis. Then kS becomes a coalgebra via

∆(s) = s⊗ s, and ε(s) = 1

for all s ∈ S. First, we check the coassociativity:

(I ⊗∆) ◦∆(s) = s⊗∆(s) = s⊗ (s⊗ s) = (s⊗ s)⊗ s = ∆(s)⊗ s = (∆⊗ I) ◦∆(s).

The counit property is easily verified as well. �

This shows that any k-vector space can be endowed with a coalgebra structure. Now,

we introduce a very important example.

Example 1.1.6 (The Group Algebra). Let G be a multiplicative group. Denote by

kG the group algebra which has G as a basis for its vector space. The multiplication

of kG is the obvious one, induced by the multiplication of G and extended linearly.

The unit map is given by u(α) = α1G, for all α ∈ k. Additionally, we can endow

kG with the structure of a coalgebra by using the preceding example, letting S = G.

Later, we will be very interested in objects, such as kG, which can be simultaneously

endowed with the structure of an algebra and coalgebra. �

Example 1.1.7 (The Trigonometric Coalgebra). Let C be the k-vector space with

basis {s, c}. Then define comultiplication ∆ and counit ε by

∆(s) = s⊗ c+ c⊗ s, ε(s) = 0∆(c) = c⊗ c− s⊗ s, ε(c) = 1.

This forms a coalgebra, with a resemblance to the angle sum formulas. Since we

may extend our results linearly by the basis elements, we must only check that the

comultiplication is coassociative on s and c. Further, we must check that diagram

Page 10: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

5

(1.4), the counit property, holds for both s and c. Below are the details for c:

(I ⊗∆) ◦∆(c)

=(I ⊗∆) ◦ (c⊗ c− s⊗ s)

=c⊗ (c⊗ c− s⊗ s)− s⊗ (s⊗ c+ c⊗ s)

=c⊗ c⊗ c− s⊗ s⊗ c− s⊗ c⊗ s− c⊗ s⊗ s

=(c⊗ c− s⊗ s)⊗ c− (s⊗ c+ c⊗ s)⊗ s

=∆(c)⊗ c−∆(s)⊗ s

=(∆⊗ I) ◦∆(c).

Now we check the counit property; that is,

(ε⊗ I) ◦∆(c) = ε(c)⊗ c− ε(s)⊗ s = 1⊗ c, and

(I ⊗ ε) ◦∆(c) = c⊗ ε(c)− s⊗ ε(s) = c⊗ 1.

The computations for s are similar. �

Example 1.1.8. The field k is a coalgebra with comultiplication defined by ∆(λ) =

λ⊗ 1, the canonical isomorphism, and counit ε = I. Note that this is a special case

of Example 1.1.5 when S is a set of only one element. �

Before introducing the next examples, we need to define a linear map which will be

utilized throughout.

Definition 1.1.9. Let V and W be k-vector spaces. The k-linear map T : V ⊗W →

W ⊗ V defined by v ⊗ w 7→ w ⊗ v, is called the twist map.

Example 1.1.10 (The Opposite Algebra). Let (A,M, u) be an algebra. The op-

posite algebra (Aop,M op, u) has A as its underlying k-vector space but the twisted

multiplication:

M op(x⊗ y) = (M ◦ T )(x⊗ y).

Page 11: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

6

It is not difficult to check Aop is an algebra. �

Example 1.1.11 (The Coopposite Coalgebra). Let (C,∆, ε) be a coalgebra. The

coopposite coalgebra (Ccop,∆cop, ε) has C as its underlying k-vector space but the

twisted comultiplication:

∆cop(c) = (T ◦∆)(c).

Again, the verification is straightforward. �

Example 1.1.12 (The Tensor Product of Coalgebras). Let (C,∆C , εC) and (D,∆D, εD)

be coalgebras. We claim that the tensor product, C ⊗D becomes a coalgebra by the

following maps:

∆(c⊗ d) = (I ⊗ T ⊗ I) ◦ (∆C ⊗∆D)(c⊗ d), and ε(c⊗ d) = εC(c)εD(d).

We postpone the verification of this example for the next section since it motivates a

method of simplification for the unwieldy calculations in a coalgebra. �

1.2 Computation in coalgebras, sigma notation

It is clear that computations in a coalgebra will be more difficult than those in an

algebra. When the comultiplication is applied to a single element, it expands into a

finite sum of pairs of elements. In the case of algebras, we may expand by induction the

notion of associativity to as many multiplicands as possible, so that the multiplication

produces unique results. It is natural to inquire whether or not an analogous property

holds for coalgebras. To investigate this, let (C,∆, ε) be a coalgebra and recursively

define the sequence of maps, {∆n}n≥1, by:

∆1 = ∆, ∆n : C → C ⊗ · · · ⊗ C (with C appearing n+ 1 times),

∆n = (∆⊗ In−1) ◦∆n−1, for n ≥ 2.

Page 12: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

7

The following proposition dualizes the desired notion to coalgebras. We will call this

property generalized coassociativity.

Proposition 1.2.1. Let (C,∆, ε) be a coalgebra. Then for any n ≥ 2, j ∈ {0, 1, . . . , n−

1}, the following equality holds:

∆n = (Ij ⊗∆⊗ In−1−j) ◦∆n−1.

Proof. The proof is by induction on n. Notice that for n = 2 we must show (I ⊗

∆) ◦ ∆ = (∆ ⊗ I) ◦ ∆, but this is just the commutativity of diagram (1.3), that is,

coassociativity. Now, we assume the equation holds for n and let j ∈ {1, . . . , n}.

Then,

(Ij ⊗∆⊗ In−j) ◦∆n

=(Ij ⊗∆⊗ In−j) ◦ (Ij−1 ⊗∆⊗ In−j) ◦∆n−1

=(Ij−1 ⊗ ((I ⊗∆) ◦∆)⊗ In−j) ◦∆n−1

=(Ij−1 ⊗ ((∆⊗ I) ◦∆)⊗ In−j) ◦∆n−1

=(Ij−1 ⊗∆⊗ In−j+1) ◦ (Ij−1 ⊗∆⊗ In−j) ◦∆n−1

=(Ij−1 ⊗∆⊗ In−j+1) ◦∆n.

Where we have again made use of the coassociativity in the fourth line above. So

we have “shifted” the comultiplication by one place. Notice that if j = 0 then

∆n = (Ij ⊗ ∆ ⊗ In−1−j) ◦ ∆n−1, by definition. The result follows by induction on

j. �

In practice, this tells us that if we are to apply ∆ to a ci in the string ∆n−1(c) =

c1 ⊗ c2 ⊗ · · · ⊗ cn in order to obtain ∆n, it is irrelevant which term we choose to

expand, our answer is always unique. This fact allows us to make use of a very helpful

notation, introduced by Heyneman and Sweedler [7], when performing calculations in

a coalgebra.

Page 13: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

8

Notation 1.2.2. Let (C,∆, ε) be a coalgebra. In general, we could write the comul-

tiplication applied to the element c ∈ C as

∆(c) =∑i

ci1 ⊗ ci2.

Because we wish to emphasize mostly the form of the element ∆(c) ∈ C ⊗ C, we

simply suppress the subscript ‘i’ and denote for any c ∈ C

∆(c) =∑

c1 ⊗ c2.

We call this convention the sigma notation. Furthermore, generalized coassociativ-

ity allows us to write unambiguously for any c ∈ C,

∆n(c) =∑

c1 ⊗ · · · ⊗ cn+1.

We now present some formulas making use of the sigma notation to aid in our calcu-

lations.

Facts 1.2.3. Let (C,∆, ε) be a coalgebra and c ∈ C. The following are reformulations

of the coassociativity and counit properties.

(i) (Coassociativity) ∆2(c) =∑

∆(c1)⊗ c2 =∑

c1 ⊗∆(c2) =∑

c1 ⊗ c2 ⊗ c3

(ii) (Counit)∑

ε(c1)c2 =∑

c1ε(c2) = c.

Proof. (i) The first assertion is clear by chasing through diagram (1.3) in the definition

of a coalgebra. We have already agreed to write the unique element in C ⊗ C ⊗ C

given by the first two expressions in (i) as∑c1 ⊗ c2 ⊗ c3. Alternatively, we could

have written:

∆2(c) =∑

(c1)1 ⊗ (c1)2 ⊗ c2 =∑

c1 ⊗ (c2)1 ⊗ (c2)2 =∑

c1 ⊗ c2 ⊗ c3.

Page 14: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

9

(ii) We may justify this formulation for the counit property by first writing the

canonical isomorphisms, φL : k ⊗ C → C, and φR : C ⊗ k → C, given by scalar

multiplication. Then, diagram (1.4) may be written as:

I = φL ◦ (ε⊗ I) ◦∆ = φR ◦ (I ⊗ ε) ◦∆,

and so, c =∑ε(c1)c2 =

∑c1ε(c2). �

We now return to the concept of the tensor product of two coalgebras to illustrate

the power of this notation.

Proposition 1.2.4. Let (C,∆C , εC) and (D,∆D, εD) be coalgebras. Then (C ⊗

D,∆, ε) is a coalgebra with the following comultiplication and counit maps:

∆(c⊗ d) = (I ⊗ T ⊗ I) ◦ (∆C ⊗∆D)(c⊗ d)

ε(c⊗ d) = εC(c)εD(d).

Proof. First, we note that we may rewrite the comultiplication ∆ in terms of the

sigma notation as ∆(c⊗ d) =∑

c1 ⊗ d1 ⊗ c2 ⊗ d2. Then we have that,

(∆⊗ IC⊗D) ◦∆(c⊗ d)

=(∆⊗ IC⊗D)(∑

c1 ⊗ d1 ⊗ c2 ⊗ d2)

=∑

(c1)1 ⊗ (d1)1 ⊗ (c1)2 ⊗ (d1)2 ⊗ c2 ⊗ d2

=∑

c1 ⊗ d1 ⊗ c2 ⊗ d2 ⊗ c3 ⊗ d3,

and also,

(IC⊗D ⊗∆) ◦∆(c⊗ d)

=(IC⊗D ⊗∆)(∑

c1 ⊗ d1 ⊗ c2 ⊗ d2)

=∑

c1 ⊗ d1 ⊗ (c2)1 ⊗ (d2)1 ⊗ (c2)2 ⊗ (d2)2

=∑

c1 ⊗ d1 ⊗ c2 ⊗ d2 ⊗ c3 ⊗ d3,

Page 15: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

10

showing that the comultiplication is coassociative. We use the reformulation of the

counit property above and notice that

∑ε(c1 ⊗ d1)(c2 ⊗ d2)

=∑

εC(c1)εD(d1)(c2 ⊗ d2)

=(∑

εC(c1)c2

)⊗(∑

εD(d1)d2

)=c⊗ d.

Similarly,∑

(c1 ⊗ d1)ε(c2 ⊗ d2) = c⊗ d, and hence, (C ⊗D,∆, ε) is a coalgebra. �

We conclude this section with two important categorical definitions. The first is

equivalent to the usual definition for commutativity; the second is its dual.

Definition 1.2.5. (i) An algebra (A,M, u) is said to be commutative if the follow-

ing diagram commutes:

A⊗ AT

- A⊗ A

A�

MM

-

(ii) A coalgebra (C,∆, ε) is said to be cocommutative if the following diagram

commutes:C

C ⊗ CT

-�

C ⊗ C

-

In sigma notation, we may write this as∑c1 ⊗ c2 =

∑c2 ⊗ c1 for any c ∈ C.

Page 16: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

11

1.3 Morphisms, coideals, quotient structures

In this section we verify that all of the usual constructions for algebraic structures are

valid in the case of coalgebras. We define the maps between algebras the usual way,

but use commutative diagrams for the purpose of dualizing the definitions. We then

proceed to construct the “ideals” for coalgebras; that is, the objects that will lead to

natural quotient structures.

Definition 1.3.1. Let (A,M, u) and (A′,M ′, u′) be k-algebras. A k-linear map f :

A→ A′ is called an algebra morphism if the following diagrams are commutative:

A⊗ Af ⊗ f

- A′ ⊗ A′

A

M

?

f- A′

M ′

?

Af

- A′

k

u′

-�

u

Alternatively, we could write for any a, b ∈ A, λ ∈ k, that f(ab) = f(a)f(b) and

f(u(λ)) = u′(λ), or equivalently, f(1A) = 1A′.

Dualizing this definition to the case of coalgebras we have:

Definition 1.3.2. Let (C,∆, ε) and (C ′,∆′, ε′) be k-coalgebras. A k-linear map g :

C → C ′ is called a coalgebra morphism if the following diagrams are commutative:

Cg

- C ′

C ⊗ C

?

g ⊗ g- C ′ ⊗ C ′

∆′

?

k

Cg

-

ε

-

C ′

ε ′

In sigma notation, we could write for any c ∈ C,

∆′(g(c)) =∑

g(c)1 ⊗ g(c)2 =∑

g(c1)⊗ g(c2), and ε(c) = ε′(g(c)).

Page 17: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

12

Definition 1.3.3. Let (C,∆, ε) be a coalgebra. A subspace J ⊆ C is called a

(i) co-subalgebra if ∆(J) ⊆ J ⊗ J ,

(ii) coideal if ∆(J) ⊆ J ⊗ C + C ⊗ J and ε(J) = 0.

Here we extend some important results for algebras to the case of coalgebras.

Proposition 1.3.4. Let (C,∆C , εC), (D,∆D, εD) be coalgebras and f : C → D a

coalgebra morphism. Then,

(i) f(C) is a co-subalgebra of D,

(ii) Ker(f) is a coideal of C.

Proof. (i) f(C) is a subspace of D since f is k-linear. Using the fact that f is a

coalgebra morphism, we have that

∆D(f(C)) =(f ⊗ f)∆C(C)

⊆(f ⊗ f)(C ⊗ C)

=f(C)⊗ f(C),

showing f(C) is a co-subalgebra.

(ii) Again, we know that Ker(f) is certainly a subspace of C. Let {vα}α∈A1 be a basis

for Ker(f) which we complete with the linearly independent set, {vα}α∈A2 so that

{vα}α∈A1∪A2 is a basis for all of C. Then {f(vα)}α∈A2 is a linearly independent set in

D. Note that since f is a coalgebra morphism, 0 = ∆D(f(Ker f)) = (f⊗f)∆C(Ker f).

Thus,

∆C(Ker f) ⊆ Ker(f ⊗ f). (1.5)

Now, take x ∈ Ker(f ⊗ f). That is,

x =∑

α,β∈A1∪A2

λαβvα ⊗ vβ,

Page 18: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

13

so that after applying f ⊗ f , we have

∑α,β∈A1∪A2

λαβf(vα)⊗ f(vβ) = 0.

Now, for each α ∈ A1 we have that f(vα) = 0. So we may rewrite the above sum

(less redundantly) as ∑α,β∈A2

λαβf(vα)⊗ f(vβ) = 0.

Then by the linearly independence of the set {f(vα)⊗ f(vβ)}α,β∈A2 , (indeed, it forms

a basis for f(C)⊗f(C)), we have that 0 = λαβ for all α, β ∈ A2. Thus, we can rewrite

x =∑

β∈A1∪A2

∑α∈A1

λαβvα ⊗ vβ +∑

α∈A1∪A2

∑β∈A1

λαβvα ⊗ vβ,

which implies that

x ∈ Ker(f)⊗ C + C ⊗Ker(f). (1.6)

Combining (1.5) and (1.6) gives ∆C(Ker(f)) ⊆ Ker(f⊗f) ⊆ Ker(f)⊗C+C⊗Ker(f).

Moreover, if x ∈ Ker(f), then 0 = εD(f(x)) = εC(x), so εC(Ker(f)) = 0. Therefore,

Ker(f) is a coideal. �

Now we can introduce quotient structures for coalgebras. Let (C,∆, ε) be a coal-

gebra and J be a coideal. Define the map π : C → C/J to be the canonical vector

space projection. We will write π(c) = c as the coset of c modulo J . We must first

recall a property of quotient vector spaces.

Lemma 1.3.5 (Universal Property of Quotient Vector Spaces). Let V be a k-vector

space, and W a subspace of V . Whenever W ′ is a k-vector space and ψ : V → W ′ a

linear map such that W ⊆ Ker(ψ) there exists a unique linear map φ : V/W → W ′

such that ψ = φ ◦ π.

Proof. Define the map φ : V/W → W ′ by φ(v) = ψ(v), that is, ψ = φ ◦ π. The map,

φ, is well defined since whenever v1, v2 ∈ V are representatives of the coset v ∈ V/W ,

Page 19: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

14

then v1 = v2 + w for some w ∈ W . Hence,

ψ(v1) = ψ(v2 + w) = ψ(v2) + ψ(w) = ψ(v2).

Moreover, φ is linear, since

φ(v1 + v2) = φ(v1 + v2) = ψ(v1 + v2) = ψ(v1) + ψ(v2) = φ(v1) + φ(v2).

Finally, φ is unique. For if φ′ is such that φ′ ◦ π = ψ = φ ◦ π, then φ′(v) = φ(v) for

all v ∈ V . Since π is surjective, φ′ = φ. �

Let (C,∆, ε) be a coalgebra, and J ⊆ C be a coideal. We can now prove the following:

Proposition 1.3.6. There exist linear maps ∆ : C/J → C/J⊗C/J and ε : C/J → k

such that (C/J,∆, ε) has a unique coalgebra structure.

Proof. Consider the linear map (π ⊗ π)∆ : C → C/J ⊗ C/J , and note that (π ⊗

π)∆(J) ⊆ (π⊗ π)(J ⊗C +C ⊗ J) = 0. Hence J ⊆ Ker((π⊗ π)∆). So, there exists a

unique linear map, ∆ : C/J → C/J ⊗ C/J such that (π ⊗ π)∆ = ∆ ◦ π, that is, the

following diagram commutes:

- C/J

C ⊗ C

? π ⊗ π- C/J ⊗ C/J

?

(1.7)

Page 20: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

15

This map is explicitly given by ∆(c) =∑c1 ⊗ c2 for c ∈ C. Then,

(∆⊗ IC/J)∆(c)

=(∆⊗ IC/J)∑

c1 ⊗ c2

=(∆⊗ IC/J)∑

π(c1)⊗ π(c2)

=∑

∆(π(c1))⊗ π(c2)

=∑

(π ⊗ π)∆(c1)⊗ π(c2)

=∑

π(c1)⊗ π(c2)⊗ π(c3), (by the coassociativity of C)

=∑

c1 ⊗ c2 ⊗ c3.

Similarly, (IC/J ⊗ ∆)∆(c) =∑c1 ⊗ c2 ⊗ c3. So ∆ is coassociative. Furthermore,

J ⊆ Ker(ε) by definition. Then, there exists a unique linear map ε : C/J → k such

that ε = ε ◦ π. That is, the following diagram commutes.

- k

C/J

ε

-

π

-

(1.8)

Then for any c ∈ C, we have

∑ε(c1)c2 =

∑ε(π(c1))π(c2)

=∑

ε(c1)π(c2)

=π(∑

ε(c1)c2

), (linearity of π)

=π(c), (counit property in C)

=c.

Similarly,∑

(c1)ε(c2) = c, and hence (C/J,∆, ε) is a coalgebra. The uniqueness of

the structure follows directly from the uniqueness of the linear maps ∆ and ε. �

Page 21: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

16

Corollary 1.3.7. π is a coalgebra morphism.

Proof. Follows from the commutativity of diagrams (1.7) and (1.8). �

Proposition 1.3.8 (The Fundamental Isomorphism Theorem for Coalgebras). Let

f : C → D be a surjective morphism of coalgebras and J = Ker(f). Then D ∼= C/J .

Proof. Consider the diagram,

Cf

- D

C/J

f

-

π

-

where f : C/J → D via c 7→ f(c). This map is well defined for the same reason as φ in

the proof of Lemma 1.3.5. Now, suppose c ∈ Ker(f). Then 0 = f(c) = f(c). Hence,

c ∈ Ker(f) = J and c = 0, showing f is injective. Moreover, since f is surjective, for

any d ∈ D, there exists c ∈ C such that d = f(c) = f(c). Therefore, f is a bijection.

Finally, using that f is a coalgebra morphism we have,

(f ⊗ f)∆(c) =∑

f(c1)⊗ f(c2)

=∑

f(c1)⊗ f(c2)

=(f ⊗ f)∆C(c)

=∆Df(c)

=∆Df(c),

and, εD◦f(c) = εD(f(c)) = εC(c) = ε(c), showing that f is a coalgebra morphism. �

1.4 Duality

We now emphasize a relationship between algebras and coalgebras, by examining

their dual spaces. Explicitly, the dual of a coalgebra is an algebra, and the dual of a

Page 22: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

17

finite dimensional algebra is a coalgebra, once defining appropriate, and quite natural,

structure maps.

For any k-vector space V we let V ∗ = Homk(V, k), the linear dual of V . Usually,

when there is no ambiguity, we will simply write, Hom(V, k). Furthermore, if V and

W are two k-vector spaces and φ : V → W is a morphism of k-vector spaces, then we

denote the transpose of the map φ by φ∗ : W ∗ → V ∗, given by

φ∗(f)(v) = f(φ(v))

for all f ∈ W ∗, v ∈ V . Also, we have the following fact.

Proposition 1.4.1. Let dim(V ) = n < ∞ and {ei}ni=1 be a basis for V , then there

exists a basis {e∗i }ni=1 for V ∗ given by

e∗i (ej) = δij,

where δij is the Kronecker delta. The linearly independent set {e∗i } is called a dual

basis to {ei} for V ∗. �

The following Lemma will also be necessary.

Lemma 1.4.2. Let V and W be k-vector spaces. Define the linear map ρ : V ∗⊗W ∗ →

(V ⊗W )∗, given by

ρ(f ⊗ g)(v ⊗ w) = f(v)g(w),

for f ∈ V ∗, g ∈ W ∗, v ∈ V,w ∈ W . Then

(i) ρ is injective.

(ii) if V and W are finite dimensional, then ρ is an isomorphism.

Proof. (i) Let x =∑

i fi ⊗ gi, (a finite sum) with the fi, gi linearly independent

elements of V ∗,W ∗, respectively, such that ρ(x) = 0. Note that this is possible since

Page 23: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

18

we can write bases {fα}α∈A and {gβ}β∈B for V ∗ and W ∗, respectively. Then the family

{fα ⊗ gβ}α∈A,β∈B forms a basis for V ∗ ⊗W ∗. We can then choose the finite family

{fi ⊗ gi}i ⊆ {fα ⊗ gβ}α∈A,β∈B. (Note: we may need to take scalar multiples of the fα

to form the fi, but they are still linearly independent.)

Now suppose there exists v ∈ V such that fi(v) 6= 0 for some i. Then the sum∑i fi(v)gi 6= 0 by the linear independence of the family, {gi}i. Hence, there must

exist w ∈ W such that

ρ(x)(v ⊗ w) = ρ

(∑i

fi ⊗ gi

)(v ⊗ w) =

∑i

fi(v)gi(w) 6= 0,

a contradiction. Thus, fi(v) = 0 for all v ∈ V , and thus x =∑

i fi ⊗ gi = 0.

(ii). Let {vi}ni=1, {wj}mj=1 be bases for V and W , respectively. Form the dual bases

{v∗i }ni=1 and {w∗j}mj=1 for V ∗ and W ∗, respectively. Then, {v∗i ⊗ w∗j}i,j is a basis for

V ∗⊗W ∗. Thus, dim(V ∗⊗W ∗) = nm. On the other hand, the family {vi⊗wj}i,j forms

a basis for V ⊗W , and we can form the dual basis {(vi⊗wj)∗}i,j for (V ⊗W )∗. Hence

dim(V ⊗W )∗ = nm and the two vector spaces must be isomorphic. Since ρ provides

an embedding V ∗ ⊗W ∗ ↪→ (V ⊗W )∗, it must be the desired isomorphism. �

In order to define a coalgebra structure on the dual of a finite dimensional algebra,

we will also need the following corollary.

Corollary 1.4.3. For any k-vector spaces V1, . . . , Vn the map θn : V ∗1 ⊗ · · · ⊗ V ∗n →

(V1 ⊗ · · · ⊗ Vn)∗ given by

θn(f1 ⊗ · · · ⊗ fn)(v1 ⊗ · · · ⊗ vn) = f1(v1) · · · fn(vn)

is injective. Furthermore, if each Vi is finite dimensional, then θn is an isomorphism.

Proof. Follows from induction on the above lemma. �

Page 24: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

19

Remark 1.4.4. Notice that we have already made use of the map ρ when defining the

tensor product of coalgebras. Indeed, for (C,∆C , εC) and (D,∆D, εD) two coalgebras,

recall that the counit ε for C ⊗D is given by

ε(c⊗ d) = ρ(εC ⊗ εD)(c⊗ d) = εC(c)εD(d),

which is exactly the map given in Example 1.1.12. �

Now, we return to the task at hand. Let (C,∆, ε) be a coalgebra. We wish to

define an algebra structure on C∗. The preceding Lemma ensures that since we can

embed C∗⊗C∗ into (C ⊗C)∗, we can restrict the map ∆∗ : (C ⊗C)∗ → C∗ to a map

m : C∗ ⊗ C∗ → C∗ given by

m(f ⊗ g)(c) = ∆∗ρ(f ⊗ g)(c) = ρ(f ⊗ g)∆(c) =∑

f(c1)g(c2). (1.9)

That is, m = ∆∗ρ. Moreover, if we take the map ε∗ : k∗(∼= k) → C∗, we can form

the unit map, U = ε∗i−1 where i : k∗ → k is the canonical isomorphism given by

i(f) = f(1), (and hence, i−1(λ) = λI). Then, we can write for any λ ∈ k that

U : k → C∗ is given by

U(λ)(c) = λε(c). (1.10)

Proposition 1.4.5. (C∗,m, U) is an algebra.

Proof. For simplicity, we will denote m(f ⊗ g) as f ∗ g. Now, for any f, g, h ∈ C∗ and

any c ∈ C, making use of the coassociativity in C, we have

((f ∗ g) ∗ h)(c)

=∑

(f ∗ g)(c1)h(c2)

=∑

f(c1)g(c2)h(c3)

=∑

f(c1)(g ∗ h)(c2)

=(f ∗ (g ∗ h))(c),

Page 25: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

20

and therefore, m is associative. Now we need to check the unit property. To do this,

we verify that U(1) is the identity element of C∗, i.e. that U(1) ∗ f = f ∗ U(1) = f

for any f ∈ C∗.

Now, by the counit property of C, we have that∑ε(c1)c2 = c for all c ∈ C. Then

we see,

(U(1) ∗ f)(c)

=∑

U(1)(c1)f(c2)

=∑

ε(c1)f(c2)

=f(∑

ε(c1)c2

), (by linearity of f)

=f(c).

Similarly, using∑c1ε(c2) = c, we obtain f ∗ U(1) = f . �

Now, if we start with an algebra (A,M, u), and try to endow A∗ with a coalgebra

structure, problems arise. Attempting to proceed as above, we consider the multipli-

cation M : A⊗ A→ A, and try to define a comultiplication on A∗ via the transpose

map M∗ : A∗ → (A⊗A)∗. But, since A∗⊗A∗ ⊆ (A⊗A)∗, there is no guarantee that

our result will lie in A∗⊗A∗, as desired. If, however, we restrict to finite dimensional

algebras, then the map ρ : A∗ ⊗ A∗ → (A⊗ A)∗ defined in Lemma 1.4.2 is bijective.

So we may define a comultiplication δ : A∗ → A∗ ⊗ A∗ by δ = ρ−1M∗ and a counit

map E : A∗ → k by E = iu∗, where i is the canonical isomorphism k∗ → k, defined

as before.

Facts 1.4.6. If we write δ(f) =∑

i gi ⊗ hi, a finite sum, then

(i) f(ab) =∑

i gi(a)hi(b) for all a, b ∈ A.

(ii) Moreover, if {xj, yj}j is any finite family of elements in A∗ such that∑

j xj(a)yj(b) =

f(ab), then∑

j xj ⊗ yj =∑

i gi ⊗ hi.

Page 26: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

21

Proof. (i) First, we note that M∗(f)(a ⊗ b) = f(ab) for all a, b ∈ A. Then, the

definition of ρ gives that if

δ(f)(a⊗ b) = ρ−1M∗(f)(a⊗ b) =

(∑i

gi ⊗ hi

)(a⊗ b),

then

f(ab) = M∗(f)(a⊗ b) = ρ

(∑i

gi ⊗ hi

)(a⊗ b) =

∑gi(a)hi(b)

for all a, b ∈ A. Assertion (ii) follows immediately from the injectivity of ρ. �

Now we are ready to prove the following proposition:

Proposition 1.4.7. Let (A,M, u) be a finite dimensional algebra. Then (A∗, δ, E) is

a coalgebra.

Proof. Take f ∈ A∗ and write δ(f) =∑

i gi⊗hi. Also, we will write δ(gi) =∑

j g′i,j⊗

g′′i,j, and δ(hi) =∑

j h′i,j ⊗ h′′i,j. Then,

(δ ⊗ I) ◦ δ(f) =∑i,j

g′i,j ⊗ g′′i,j ⊗ hi, and

(I ⊗ δ) ◦ δ(f) =∑i,j

gi ⊗ h′i,j ⊗ h′′i,j.

Now using the associativity of A, the fact that f(ab) =∑gi(a)hi(b) for all a, b ∈ A,

and the map, θ3 : A∗ ⊗ A∗ ⊗ A∗ → (A⊗ A⊗ A)∗, (as defined in Corollary 1.4.3), we

have that,

θ3((δ ⊗ I) ◦ δ(f))(a⊗ b⊗ c)

=θ3(∑i,j

g′i,j ⊗ g′′i,j ⊗ hi)(a⊗ b⊗ c)

=∑i,j

g′i,j(a)g′′i,j(b)hi(c)

=∑i

gi(ab)hi(c)

=f(abc)

Page 27: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

22

Moreover,

θ3((I ⊗ δ) ◦ δ(f))(a⊗ b⊗ c)

=θ3(∑i,j

gi ⊗ h′i,j ⊗ h′′i,j)(a⊗ b⊗ c)

=∑i,j

gi(a)h′i,j(b)h′′i,j(c)

=∑i

gi(a)hi(bc)

=f(abc).

Then, by the injectivity of θ3, (δ ⊗ I) ◦ δ(f) = (I ⊗ δ) ◦ δ(f), for any f ∈ A∗. Thus,

the coassociativity is checked. By the unit property of A, we note that E(f) = f(1)

for f ∈ A∗, and hence:(∑i

E(gi)hi

)(a) =

∑i

gi(1)hi(a) = f(1 · a) = f(a).

Similarly,∑

i giE(hi) = f . Thus (A∗, δ, E) is a coalgebra. �

At this point, it is obvious to ask what happens to algebra morphisms and coalgebra

morphisms under dual constructions.

Proposition 1.4.8. Let (C,∆C , εC) and (D,∆D, εD) be k-coalgebras and f : C →

D a coalgebra morphism. Then (C∗,mC∗ , UC∗) and (D∗,mD∗ , UD∗) are algebras (as

defined above) and f ∗ : D∗ → C∗ is an algebra morphism.

Proof. We need to show that the following diagrams commute.

D∗ ⊗D∗f ∗ ⊗ f ∗

- C∗ ⊗ C∗

D∗

mD∗

?

f ∗- C∗

mC∗

?

D∗f ∗

- C∗

k

UC∗

-�

UD ∗

Page 28: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

23

Let x, y ∈ D∗ and c ∈ C. Then

(f ∗ ◦mD∗)(x⊗ y)(c)

=ρ(x⊗ y)∆D(f(c))

=ρ(x⊗ y)(∑

f(c)1 ⊗ f(c)2

)=ρ(x⊗ y)

(∑f(c1)⊗ f(c2)

)=∑

(x(f)(c1))(y(f)(c2))

=ρ(x(f)⊗ y(f))∆C(c)

=mC∗(f∗(x)⊗ f ∗(y))(c)

=mC∗(f∗ ⊗ f ∗)(x⊗ y)(c),

where we have used the fact that f is a coalgebra morphism in the third line. This

shows the commutativity of the first diagram above. Now, we use that εD ◦ f = εC

(f is a coalgebra morphism), and obtain for λ ∈ k, c ∈ C,

(f ∗ ◦ UD)(λ)(c) = UD(λ)(f(c))

= λεD(f(c))

= λεC(c)

= UC(λ)(c).

Thus, f ∗ is an algebra morphism. �

Proposition 1.4.9. Let (A,MA, uA) and (B,MB, uB) be finite dimensional k-algebras

and g : A → B an algebra morphism. Then (A∗, δA∗ , EA∗) and (B∗, δB∗ , EB∗) are

coalgebras (as defined above) and g∗ : B∗ → A∗ is a coalgebra morphism.

Page 29: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

24

Proof. We need to show the following diagrams commute.

B∗g∗

- A∗

B∗ ⊗B∗

δB∗

?

g∗ ⊗ g∗- A∗ ⊗ A∗

δA∗

?

k

B∗g∗

-

EB∗

-

A∗

EA ∗

Let x ∈ B∗ and a′ ⊗ a′′ ∈ A ⊗ A. First, note that if δB∗(x) =∑

i x′i ⊗ x′′i , where

x′i, x′′i ∈ B∗, then

δB∗(x)(g ⊗ g) =∑

i x′i(g)⊗ x′′i (g)

=∑

i g∗(x′i)⊗ g∗(x′′i )

= (g∗ ⊗ g∗) (∑

i x′i ⊗ x′′i )

= (g∗ ⊗ g∗)δB∗(x).

(1.11)

So,

(δA∗ ◦ g∗)(x)(a′ ⊗ a′′)

=δA∗(x(g))(a′ ⊗ a′′)

=x(g(a′a′′))

=x(g(a′)g(a′′))

=δB∗(x)(g(a′)⊗ g(a′′))

=δB∗(x)(g ⊗ g)(a′ ⊗ a′′)

=(g∗ ⊗ g∗)δB∗(x)(a′ ⊗ a′′)

where we have used that g is an algebra morphism in the third line, and (1.11). Hence,

the commutativity of the first diagram is checked. Now, for any x ∈ B∗, b ∈ B, we

Page 30: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

25

have

(EA∗ ◦ g∗)(x)(b) = (EA∗ ◦ x(g))(b)

= x(g ◦ uA)(b)

= x(uB)(b)

= EB∗(x)(b),

where we have used the property that g ◦ uA = uB since g is an algebra morphism.

This checks the commutativity of the second diagram. �

We end this subsection with an important result that we will use later, but will not

prove. A proof can be found in Dascalescu, et. al. [4].

Proposition 1.4.10. Let A be a finite dimensional algebra and C a finite dimensional

coalgebra. Then

(i) A and (A∗)∗ are isomorphic as algebras,

(ii) C and (C∗)∗ are isomorphic as coalgebras. �

1.5 Group-like and primitive elements

In this section we will define and give some important results for two special classes

of elements in a coalgebra. Throughout, let (C,∆, ε) be a coalgebra.

Definition 1.5.1 (Group-like elements). Let g ∈ C such that g 6= 0 and ∆(g) = g⊗g.

Then g is called a group-like element of C. The set of all group-like elements is

denoted by G(C).

Facts 1.5.2. We have the following facts regarding G(C):

(i) If g ∈ G(C), then ε(g) = 1.

Page 31: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

26

(ii) G(C) is a linearly independent set.

Proof. Assertion (i) follows immediately from the counit property since g =∑ε(g1)g2 =

ε(g)g. To prove (ii), we assume the set G(C) is linearly dependent and seek a con-

tradiction. For, let n be the least natural number so that there exists g, g1, . . . , gn ∈

G(C), distinct elements such that

g =n∑i=1

αigi (1.12)

for some scalars αi ∈ k. Notice that if n = 1, then g = α1g1 and hence ε(g) = α1ε(g1).

Then α1 = 1, and by (i) we have g = g1, a contradiction.

So we must have that n ≥ 2. Moreover, we may assume that all the αi are non-

zero, or else we would obtain equation (1.12) for a natural number smaller than n.

Apply ∆ to (1.12) to get

g ⊗ g =n∑i=1

αigi ⊗ gi.

Then using (1.12) to replace g on the left hand side we have,

n∑i,j=1

αiαjgi ⊗ gj =n∑i=1

αigi ⊗ gi.

and therefore

0 =∑i 6=j

αiαjgi ⊗ gj +∑i

(α2i − αi)gi ⊗ gi.

Since g1, . . . , gn are linearly independent (again, otherwise we would have equation

(1.12) for a smaller n), we have that {gi ⊗ gj}i,j is linearly independent in C ⊗ C.

Thus, we must have for i 6= j, that αiαj = 0. This is a contradiction since n ≥ 2 and

all αi 6= 0. �

Let A and B be algebras. Then denote the set of algebra maps from A to B by

Alg(A,B). That is,

Alg(A,B) = {f ∈ Hom(A,B) | f(ab) = f(a)f(b), ∀a, b ∈ A}.

Page 32: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

27

Now, if A is a finite dimensional algebra, we can describe the group-like elements of

A∗ exactly.

Proposition 1.5.3. G(A∗) = Alg(A, k).

Proof. If f ∈ G(A∗), then δ(f) = f ⊗ f , and by Fact 1.4.6(i), f(ab) = f(a)f(b) for

all a, b ∈ A. Moreover, f(1) = E(f) = 1, so f ∈ Alg(A, k). The reverse inclusion is

similar. �

Definition 1.5.4 (Primitive elements). Let (C,∆, ε) be a coalgebra and c 6= 0 in C

such that ∆(c) = 1⊗ c+ c⊗ 1. Then c is called a primitive element of C. The set

of all primitive elements is denoted by P (C). We can similarly define, for x, y ∈ C

the set of x, y-primitive elements, denoted by Px,y(C), where if c ∈ Px,y(C) then

∆(c) = x⊗ c+ c⊗ y.

Facts 1.5.5. We have the following facts about primitive elements:

(i) Let x ∈ G(C), 0 6= y ∈ C. If c ∈ Px,y(C), then ε(c) = 0. (Similarly, if

y ∈ G(C), 0 6= x ∈ C.)

(ii) For any x ∈ G(C) and 0 6= y ∈ C, Px,y(C) is a coideal of C.

Proof. (i) The counit property gives that c =∑ε(c1)c2. So c = ε(x)c + ε(c)y =

c+ε(c)y. Thus ε(c) = 0. To prove (ii), we note that Px,y(C) is a subspace of C, since

for any c, d ∈ Px,y(C),

∆(c+ d) = ∆(c) + ∆(d)

= x⊗ c+ c⊗ y + x⊗ d+ d⊗ y

= x⊗ (c+ d) + (c+ d)⊗ y.

Then, for any c ∈ Px,y(C), ∆(c) = x ⊗ c + c ⊗ y ∈ C ⊗ Px,y(C) + Px,y(C) ⊗ C and

part (i) gives Px,y(C) ⊆ Ker(ε). �

Page 33: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Chapter 2: Bialgebras and Hopf algebras

We will now examine k-vector spaces that are simultaneously endowed with both

an algebra and coalgebra structure. In particular, we will want to investigate spaces

in which the two structures relate in a “nice” way. Let (H,M, u) be an algebra which

is also a coalgebra, (H,∆, ε). Recall that we have a natural algebra structure induced

on H ⊗H, as well as a coalgebra structure described in Proposition 1.2.4. We must

also recall the canonical coalgebra structure on the field k as described in Example

1.1.8.

2.1 Bialgebras

Proposition 2.1.1. Let (H,M, u) be an algebra which is simultaneously a coalgebra,

(H,∆, ε). Then the following are equivalent:

(i) M and u are coalgebra morphisms.

(ii) ∆ and ε are algebra morphisms.

Proof. (i) ⇒ (ii) M is a coalgebra morphism exactly when the following diagrams

commute:

H ⊗HM

- H

H ⊗H ⊗H ⊗H

∆⊗∆?

H ⊗H ⊗H ⊗H

I ⊗ T ⊗ I?

M ⊗M- H ⊗H

?

(2.1)

28

Page 34: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

29

H ⊗HM

- H

k ⊗ k

ε⊗ ε?

k

φ?

I- k

ε

?

(2.2)

where φ : k ⊗ k → k is the canonical isomorphism φ(α ⊗ β) = αβ. Moreover, u is a

coalgebra morphism exactly when the following diagrams commute:

ku

- H

k ⊗ k

φ−1

?

u⊗ u- H ⊗H

?

(2.3)

ku

- H

k�

εI

-

(2.4)

Where φ−1 is the comultiplication map for k defined in Example 1.1.8, and given by

α 7→ α⊗1, for α ∈ k. We note that ∆ is an algebra morphism if and only if diagrams

(2.1) and (2.3) commute and ε is an algebra morphism if and only if diagrams (2.2)

and (2.4) commute. Then, using the same diagrams, the reverse implication is also

clear. �

Remark 2.1.2. For easier calculation in the future, we note that in the sigma nota-

tion, if ∆ and ε are to be algebra morphisms, we must have

∆(xy) =∑

x1y1 ⊗ x2y2, ∆(1) = 1⊗ 1,

ε(xy) = ε(x)ε(y), ε(1) = 1,

Page 35: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

30

for all x, y ∈ H. �

We are now in a position to define a bialgebra.

Definition 2.1.3. Let H be a k-vector space with algebra structure (H,M, u) and

coalgebra structure (H,∆, ε) such that M and u are coalgebra morphisms (and hence

∆, ε are algebra morphisms). Then (H,M, u,∆, ε) is called a bialgebra.

We return to our prototypical example.

Example 2.1.4. Let G be a group and kG the associated group algebra. Let kG

have a coalgebra structure as in Example 1.1.6. Then kG is a bialgebra. To verify

this, we check the conditions of Remark 2.1.2. Let g, h ∈ G. Then, gh ∈ G, so

∆(gh) = gh⊗ gh

= (g ⊗ g)(h⊗ h)

= ∆(g)∆(h),

and ε(g)ε(h) = 1 · 1 = 1 = ε(gh). Furthermore, ∆(1) = 1 ⊗ 1 and ε(1) = 1, by

definition. �

Example 2.1.5. Let (B,M, u,∆, ε) be a bialgebra. Then we can form bialgebras

Bop and Bcop, where Bop (resp. Bcop) is the vector space B, but with algebra (resp.

coalgebra) structure given by the opposite algebra (resp. coopposite coalgebra) de-

fined in Example 1.1.10 (resp. 1.1.11). Bcopop has both the opposite and coopposite

structures. It is not difficult to check that each of these are bialgebras. �

We have shown that the structures of (finite dimensional) algebras and coalgebras

are, in fact, dual constructions. This leads to another method of obtaining new

bialgebras from existing ones.

Page 36: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

31

Proposition 2.1.6. Let (H,M, u,∆, ε) be a finite dimensional bialgebra. Then H∗

is a bialgebra with algebra structure dual to the coalgebra structure on H (as defined

by Proposition 1.4.5), and coalgebra structure dual to the algebra structure on H (as

defined by Proposition 1.4.7).

Proof. We have already shown that (H∗,m, U) is an algebra when we define the

multiplication m = ∆∗ρ and unit U = ε∗i−1. We have also shown that (H∗, δ, E) is a

coalgebra when we define the comultiplication δ = ρ−1M∗ and counit E = iu∗.

We need only to show that m and U are coalgebra morphisms (or equivalently,

δ and E are algebra morphisms), but this follows readily from Proposition 1.4.9.

Indeed, since H is a bialgebra, ∆ and ε are algebra morphisms, and thus ∆∗ and ε∗

are coalgebra morphisms. Since m can be thought of simply as a restriction of ∆∗,

m is also a coalgebra morphism. Similarly, U is a coalgebra morphism since it is ε∗

following the canonical isomorphism i−1. �

For the remainder of the section, let H and B be bialgebras, with respective structure

maps (MH , uH ,∆H , εH) and (MB, uB,∆B, εB).

Definition 2.1.7. The k-linear map f : H → B is called a bialgebra morphism

if f is both an algebra morphism and coalgebra morphism.

Definition 2.1.8. A subspace J ⊆ H is called

(i) a bi-subalgebra if it is a subalgebra and co-subalgebra.

(ii) a biideal if it is an ideal and coideal.

The following propositions are immediate consequences of the preceding definitions:

Proposition 2.1.9. If H and B are finite dimensional, and f : H → B is a bialgebra

morphism, then f ∗ : B∗ → H∗ is a bialgebra morphism.

Page 37: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

32

Proof. Since H and B are finite dimensional, we can be sure that H∗ and B∗ are

indeed bialgebras. Then, f ∗ must be an algebra morphism since f is a coalgebra

morphism. Similarly, f ∗ must also be a coalgebra morphism since f is an algebra

morphism. �

Proposition 2.1.10. Let J ⊆ H be a biideal. Then (H/J,M, u,∆, ε) is a bialgebra.

Proof. Since J is an ideal, it defines the unique algebra structure (H/J,M, u), and

similarly, since J is a coideal we obtain the unique coalgebra structure (H/J,∆, ε).

Let π : H → H/J be the canonical vector space projection and write π(h) = h as

the coset of h modulo J . Recall that π is both an algebra and coalgebra morphism.

Now we verify the conditions of Remark 2.1.2,

∆((x)(y))

=∆(π(x)π(y))

=∆(π(xy)), (π is an algebra morphism)

=(π ⊗ π)∆(xy), (π is a coalgebra morphism)

=(π ⊗ π)(∑

x1y1 ⊗ x2y2

), (∆ is an algebra morphism)

=∑

π(x1y1)⊗ π(x2y2)

=∑

π(x1)π(y1)⊗ π(x2)π(y2), (π is an algebra morphism)

=∑

(x1)(y1)⊗ (x2)(y2).

Page 38: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

33

Also,

ε((x)(y))

=ε(π(x)π(y))

=ε(π(xy)), (π is an algebra morphism)

=ε(xy), (π is a coalgebra morphism)

=ε(x)ε(y), (ε is an algebra morphism)

=ε(π(x))ε(π(y)), (π is a coalgebra morphism)

=ε(x)ε(y).

It is easy to check that ∆(1) = 1⊗ 1 and ε(1) = 1. �

Proposition 2.1.11. Let f : H → B be a bialgebra morphism. Then,

(i) f(H) is a bi-subalgebra of B.

(ii) Ker(f) is a biideal of H.

Proof. Combine the analogous results for algebras and coalgebras. �

Proposition 2.1.12 (The Fundamental Isomorphism Theorem for Bialgebras). Let

f : H → B be a surjective bialgebra morphism and J = Ker(f). Then B ∼= H/J .

Proof. Combine the analogous results for algebras and coalgebras, plus Proposition

2.1.10. �

2.2 Convolution, antipodes, and Hopf algebras

We begin by letting (C,∆, ε) be a coalgebra and (A,M, u) be an algebra. Now,

consider Hom(C,A), the family of k-linear maps from C to A. We can define an

Page 39: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

34

algebra structure on Hom(C,A) via the so-called convolution product. We will

denote the product of f, g ∈ Hom(C,A) by f ∗ g. The product is explicitly given by

(f ∗ g)(c) = M(f ⊗ g)∆(c) =∑

f(c1)g(c2).

Note that it was necessary to have C a coalgebra and A an algebra in order to define

the convolution product. Now, we should justify the title “algebra” for Hom(C,A)

under convolution by checking that the product is associative. Indeed, for f, g, h ∈

Hom(C,A) and c ∈ C, we have

((f ∗ g) ∗ h)(c) =∑

(f ∗ g)(c1)h(c2)

=∑

f(c1)g(c2)h(c3)

=∑

f(c1)(g ∗ h)(c2)

= (f ∗ (g ∗ h))(c),

where we have made use of the coassociativity of C. Additionally, we note that uε is

the identity element of Hom(C,A), since for any f ∈ Hom(C,A), c ∈ C,

(f ∗ uε)(c) =∑

f(c1)u(ε(c2))

=∑

f(c1)ε(c2)u(1)

=∑

f(c1)ε(c2)

= f(∑

c1ε(c2))

= f(c),

where we have used that u is linear in the second line, and the counit property in

the last line. Similarly, using∑ε(c1)c2 = c we obtain uε ∗ f = f . We have already

considered a special case of the convolution product in defining an algebra structure

on C∗, that is, Hom(C, k). Now, consider the special case in which (H,M, u,∆, ε) is

a bialgebra. Then it makes sense to talk about the algebra Hom(Hc, Ha), where we

think of Hc as the underlying coalgebra and Ha as the underlying algebra.

Page 40: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

35

Definition 2.2.1. Let H be a bialgebra. A map S ∈ Hom(Hc, Ha) is called an

antipode if it is the inverse to the identity map under convolution. H, together with

an antipode, is called a Hopf algebra.

We have the following facts.

Facts 2.2.2. Let (H,M, u,∆, ε, S) be a Hopf algebra and h, g ∈ H. Then,

(i) (Antipode Property)∑S(h1)h2 =

∑h1S(h2) = ε(h)1,

(ii) S(hg) = S(g)S(h) and S(1) = 1. That is, S is an anti-algebra morphism,

(iii) ∆(S(h)) =∑S(h)1 ⊗ S(h)2 =

∑S(h2) ⊗ S(h1) and ε(S(h)) = ε(h). That is,

S is an anti-coalgebra morphism,

Proof. (i) This follows directly from the definition of S. For, since S ∗ I = I ∗S = uε,

then for any h ∈ H,

(S ∗ I)(h) =∑

S(h1)h2

(I ∗ S)(h) =∑

h1S(h2)

u(ε(h)) = ε(h)u(1) = ε(h)1.

(ii) Let H ⊗ H have the tensor product of coalgebras structure and H an algebra

structure. Then, it makes sense to form the algebra Hom(H⊗H,H). Define the maps

X, Y ∈ Hom(H ⊗H,H) by X(h⊗ g) = S(hg), Y (h⊗ g) = S(g)S(h). Let h, g ∈ H,

Page 41: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

36

and consider,

(X ∗M)(h⊗ g)

=∑

X((h⊗ g)1)M((h⊗ g)2)

=∑

X(h1 ⊗ g1)M(h2 ⊗ g2), (comultiplication in H ⊗H)

=∑

S(h1g1)h2g2

=∑

S((hg)1)(hg)2, (M is a coalgebra morphism)

=εH(hg)1, (antipode property)

=εH(h)εH(g)1

=(uHεH⊗H)(h⊗ g).

So X is a left inverse of M . Now, we consider

(M ∗ Y )(h⊗ g)

=∑

M((h⊗ g)1)Y ((h⊗ g)2)

=∑

M(h1 ⊗ g1)Y (h2 ⊗ g2), (comultiplication in H ⊗H)

=∑

h1g1S(g2)S(h2)

=∑

h1(εH(g)1)S(h2), (antipode property)

=∑

h1S(h2)εH(g)1

=εH(h)εH(g)1, (antipode property)

=(uHεH⊗H)(h⊗ g)

So Y is a right inverse for M . Since Hom(H ⊗H,H) forms an algebra, and therefore

has unique inverses, X = Y , and we conclude that S(hg) = S(g)S(h). Consider the

element 1 ∈ H. Applying (i), we get that S(1)1 = ε(1)1, that is, S(1) = 1.

(iii) Similar to (ii), we think of H with a coalgebra structure and H ⊗ H with

an algebra structure to define Hom(H,H ⊗ H). Let F,G ∈ Hom(H,H ⊗ H) with

Page 42: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

37

F (h) =∑S(h)1⊗S(h)2 = ∆(S(h)) and G(h) =

∑S(h2)S(h1). Then for any h ∈ H,

(∆ ∗ F )(h)

=∑

∆(h1)F (h2)

=∑

∆(h1)∆(S(h2))

=∑

∆(h1S(h2)), (∆ is an algebra morphism)

=∆(∑

h1S(h2)), (linearity of ∆)

=∆(ε(h)1), (antipode property)

=ε(h)1⊗ 1

=(uH⊗HεH)(h).

Thus F is a right inverse for ∆. Moreover,

(G ∗∆)(h)

=∑

G(h1)∆(h2)

=∑

(S((h1)2)⊗ S((h1)1))((h2)1 ⊗ (h2)2)

=∑

(S(h2)⊗ S(h1))(h3 ⊗ h4), (generalized coassociativity)

=∑

S(h2)h3 ⊗ S(h1)h4

=∑

S((h2)1)(h2)2 ⊗ S(h1)h3, (generalized coassociativity)

=∑

ε(h2)1⊗ S(h1)h3, (antipode property)

=∑

1⊗ S(h1)ε((h2)1)(h2)2, (generalized coassociativity)

=∑

1⊗ S(h1)h2, (counit property)

=1⊗ ε(h)1, (antipode property)

=ε(h)1⊗ 1

=(uH⊗HεH)(h).

Hence G is a left inverse for ∆, and G = F . Finally, for any h ∈ H we have that

Page 43: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

38

ε(h)1 =∑h1S(h2). Now, applying ε to both sides we obtain

ε(h) =∑

ε(h1)ε(S(h2)), (ε is an algebra morphism)

=∑

ε(ε(h1)S(h2)), (linearity of counit)

=∑

ε(S(ε(h1)h2)), (linearity of antipode)

= ε(S(∑

ε(h1)h2

)), (again using linearity)

= ε(S(h)),

by the counit property. �

We now define the ideals in Hopf algebras.

Definition 2.2.3. Let H be a Hopf algebra. A subspace J ⊆ H is called

(i) a Hopf-subalgebra if it is a bi-subalgebra and S(J) ⊆ J .

(ii) a Hopf ideal if it is a biideal and S(J) ⊆ J .

It seems reasonable for ideals and maps to have some preserving effect on an-

tipodes. With this in mind, we turn our attention to maps between Hopf algebras.

Let H and B be Hopf algebras with antipodes SH and SB, respectively. We will say

that a linear map f : H → B preserves antipodes whenever SBf = fSH . Let f be

a bialgebra morphism. We can define the algebra Hom(H,B) with identity element

uBεH . Of course, f ∈ Hom(H,B), so consider

((SBf) ∗ f)(h)

=∑

(SBf)(h1)f(h2)

=∑

SB(f(h)1)f(h)2, (f is a coalgebra morphism)

=εB(f(h))1, (antipode property for B)

=εH(h)1, (f is a coalgebra morphism)

=(uBεH)(h),

Page 44: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

39

and

(f ∗ (fSH))(h)

=∑

f(h1)(fSH(h2))

=∑

f(h1SH(h2)), (f is an algebra morphism)

=f(εH(h)1), (antipode property for H)

=εH(h)1, (f is an algebra morphism)

=(uBεH)(h),

showing that SBf is a left inverse of f and fSH is a right inverse for f . Thus

SBf = fSH , and we have proven:

Proposition 2.2.4. If H and B are Hopf algebras and f : H → B is a bialgebra

morphism, then f preserves antipodes. �

Therefore, the following definition is justified:

Definition 2.2.5. Let H and B be Hopf algebras and f : H → B. f is called a Hopf

morphism if f is a bialgebra morphism.

The following proposition is immediate.

Proposition 2.2.6. If f : H → B is a Hopf morphism, then

(i) f(H) is a Hopf-subalgebra of B.

(ii) Ker(f) is a Hopf ideal of H.

Proof. (i) f is a bialgebra morphism, so we have that f(H) is a bi-subalgebra. More-

over, f preserves antipodes, so SB(f(H)) = f(SH(H)) ⊆ f(H).

(ii) We need only to show that S(Ker(f)) ⊆ Ker(f). Indeed, since f is a Hopf

morphism, it preserves antipodes. Then for h ∈ Ker(f), f(SH(h)) = SB(f(h)) =

SB(0) = 0. Hence, SH(h) ∈ Ker(f). �

Page 45: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

40

Also, we have the analogous results for quotient structures.

Proposition 2.2.7. Let (H,M, u,∆, ε, S) be a Hopf algebra and J a Hopf ideal. Then

there is a unique Hopf algebra structure on H/J .

Proof. We already know from Proposition 2.1.10 that H/J has a unique bialgebra

structure. It becomes a Hopf algebra with induced map S : H/J → H/J given by

S(h) = S(h). This map makes sense because J is a Hopf ideal, that is, S(J) ⊆ J .

Then we can check, for h ∈ H,∑S(h1)h2 =

∑S(h1)(h2)

=∑

π(S(h1))π(h2)

= π(∑

S(h1)h2

)= π(ε(h)1), (antipode property of H)

= ε(h)1

= ε(h)1.

Similarly,∑

(h1)S(h2) = ε(h)1. �

Proposition 2.2.8 (The Fundamental Isomorphism Theorem for Hopf Algebras).

Let H,B be Hopf algebras with antipodes SH , SB, respectively. Let f : H → B be a

surjective Hopf morphism and J = Ker(f). Then B ∼= H/J .

Proof. Use the map f from the proof of Proposition 1.3.8. We know that f is a

bialgebra morphism and therefore a Hopf morphism. �

Just as with bialgebras, we may create new Hopf algebras from existing ones, using

the dual space.

Proposition 2.2.9. If H is a finite dimensional Hopf algebra with antipode S, then

H∗ is a Hopf algebra with antipode S∗.

Page 46: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

41

Proof. It remains only to show that S∗ is an antipode. As usual, we will denote the

multiplication and unit in H∗ by m and U , respectively, and the comultiplication and

counit by δ and E, respectively. We write for any h∗ ∈ H∗, that δ(h∗) =∑h∗1 ⊗ h∗2

by the sigma notation. Also, recall from Fact 1.4.6(i) that for any x, y ∈ H, h∗(xy) =∑h∗1(x)h∗2(y). Also, E(h∗) = h∗(1), so we have for any h ∈ H,(∑

S∗(h∗1)h∗2

)(h) =

∑S∗(h∗1)(h1)h

∗2(h2)

=∑

h∗1(S(h1))h∗2(h2)

=∑

h∗(S(h1)h2)

= h∗(ε(h)1), (antipode property of H)

= ε(h)h∗(1)

= E(1)ε(h)

Similarly,∑h∗1S

∗(h∗2) = E(1)ε, so S is an antipode of H∗. �

Remark 2.2.10. Let H be a Hopf algebra. We will call H commutative if its underly-

ing algebra is commutative. We will call H cocommutative if its underlying coalgebra

is cocommutative. Additionally, we can talk about Hopf algebras with neither or both

properties. �

We have the following proposition.

Proposition 2.2.11. Let H be a finite dimensional Hopf algebra. Then,

(i) H is commutative if and only if H∗ is cocommutative.

(ii) H is cocommutative if and only if H∗ is commutative.

Proof. We will show the forward implications for both (i) and (ii). The reverse

implications then follow from Proposition 1.4.10. For (i), let f ∈ H∗, and x, y ∈ H.

Using Fact 1.4.6(i), we see∑f1(x)f2(y) = f(xy) = f(yx)

∑f2(x)f1(y). Hence

Page 47: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

42

∑f1⊗f2 =

∑f2⊗f1, that is, H∗ is cocommutative. For (ii), let f, g ∈ H∗ and x ∈ H.

Then, m(f ⊗ g)(x) =∑f(x1)g(x2) =

∑f(x2)g(x1) =

∑g(x1)f(x2) = m(g ⊗ f)(x).

Hence, H∗ is commutative. �

2.3 Some examples of Hopf algebras

In this section we give various examples and small results for Hopf algebras, beginning

with our prototype.

Example 2.3.1 (The Group Algebra). Let G be a multiplicative group. Then kG

becomes a Hopf algebra with bialgebra structure from Example 2.1.4 and antipode

S : kG → kG given by S(g) = g−1 for all g ∈ G. S is indeed an antipode. For

consider g ∈ G,

∑S(g1)g2 = S(g)g = g−1g = 1kG = ε(g)1kG.

Similarly,∑g1S(g2) = ε(g)1kG. �

Example 2.3.2 (The coopposite Hopf algebra). As in 2.1.5, if H is a Hopf algebra

with antipode S, then certainly Hcop is a bialgebra. Suppose that Hcop is a Hopf

algebra with antipode S. This means that∑

(Sh2)h1 =∑h2(Sh1) = ε(h)1 for all

h ∈ H, and since S is an antipode for Hcop it is an anti-algebra and anti-coalgebra

morphism. S is called a twisted antipode for H. Moreover, for any h ∈ H we have

SSh =∑SS(ε(h2)h1) =

∑ε(h2)SSh1 =

∑h3(Sh2)SSh1

=∑h3S((Sh1)h2) =

∑h2S(ε(h1)) =

∑ε(h1)h2 = h.

Similarly, SSh = h for any h ∈ H. Therefore, when this situation occurs, we must

have that S and S are composition inverses. This gives a necessary condition for Hcop

to be a Hopf algebra. To complete the example, we have the following proposition. �

Proposition 2.3.3. Let H be a bialgebra. The following are equivalent:

Page 48: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

43

(i) H is a Hopf algebra with composition invertible antipode S.

(ii) Hcop is a Hopf algebra with composition invertible antipode S.

That is, Hcop is a Hopf algebra if and only if H has a composition invertible antipode.

Proof. We show that (i) ⇒ (ii); the reverse direction is similar. Let S ′ be the

composition inverse of S. It is not difficult to check that S ′ must be an anti-algebra

and anti-coalgebra morphism. Now take any h ∈ H and consider

∑(S ′h2)h1 =

∑(S ′h2)(S

′Sh1) = S ′(∑

(Sh1)h2

)= S ′(ε(h)1) = ε(h)1.

Similarly,∑h2(S

′h1) = ε(h)1, so S ′ is an antipode for Hcop. By the analysis of the

previous example, S ′ = S. �

The following example is due to Sweedler and is the smallest non-commutative, non-

cocommutative Hopf algebra.

Example 2.3.4 (The Sweedler algebra). Assume that ch(k) 6= 2. Let H be the

algebra given by the two generators g and x with the following relations:

g2 = 1, x2 = 0, xg = −gx

Then H has dimension four as a vector space, with basis {1, g, x, gx}. Define a

coalgebra structure by

∆(g) = g ⊗ g, ∆(x) = g ⊗ x+ x⊗ 1.

That is, g ∈ G(H) and x ∈ Pg,1(H) so by Facts 1.5.2(i) and 1.5.5(i) we must have

ε(g) = 1 and ε(x) = 0. Finally, define S : H → H to be the antipode given by

S(g) = g, S(x) = −gx.

One can check that these relations define a Hopf algebra. �

Page 49: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

44

The Sweedler algebra was later generalized by Taft [22] to any dimension that is a

perfect square.

Example 2.3.5 (The Taft algebras). Let n ≥ 2 and λ a primitive n-th root of unity

in k (in particular this means that ch(k) does not divide n). Again, let Hn(λ) be the

algebra generated by g and x, with relations

gn = 1, xn = 0, xg = λgx.

Then Hn(λ) is an algebra with vector space basis {gixj}n−1i,j=0. Thus dim(Hn(λ)) = n2.

We introduce a coalgebra structure just as in Example 2.3.4:

∆(g) = g ⊗ g, ∆(x) = g ⊗ x+ x⊗ 1,

so that again we must have, ε(g) = 1 and ε(x) = 0. The antipode is given by

S(g) = gn−1 and S(x) = −gn−1x. Notice that when n = 2 we obtain the Sweedler

algebra. �

Remark 2.3.6. (Antipodes of Arbitrary Order.) More than the fact that Hn(λ) is a

generalization of the Sweedler algebra, it has historical importance. Taft constructed

the algebras Hn(λ) to give explicit examples of Hopf algebras having antipodes of

arbitrarily large orders. We say that the antipode has order m, whenever m is the

least positive integer such that Sm = I, under composition. In particular, one can

check that the order of the antipode in Hn(λ) is 2n. To see this, we first check the

effect of composition of the antipode on the group-like generator g. We can show by

induction on m that

Sm(g) = (S ◦ · · · ◦ S︸ ︷︷ ︸m

)(g) = g(n−1)m

.

For m = 1, S(g) = gn−1 by definition. Then if the assertion is true for all natural

numbers < m, since S is an anti-algebra morphism,

Sm(g) = S(Sm−1(g)) = S(g(n−1)m−1

) = (S(g))(n−1)m−1

= (gn−1)(n−1)m−1

= g(n−1)m

.

Page 50: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

45

Now, since

(n− 1)m =m∑k=0

(mk

)nm−k(−1)k,

we see that (n− 1)m ≡ (−1)m mod n. Therefore,

Sm(g) =

{g , if m even

gn−1 , if m odd.

So if m is to be the order of the antipode, then m must be even. We also have the

following recursive relation for Sm(x):

Sm(x) =

−Sm−1(x)g , if m even

−gn−1Sm−1(x) , if m odd.

To check this, we note that since S is an anti-algebra morphism, Sr is also an anti-

algebra morphism for all odd r. On the other hand, if r is even, then Sr is an algebra

morphism. Then for m an even integer,

Sm(x) = Sm−1(−gn−1x)

= −Sm−1(x)Sm−1(gn−1)

= −Sm−1(x)(Sm−1(g))n−1

= −Sm−1(x)(gn−1)n−1

= −Sm−1(x)gn2−2n+1

= −Sm−1(x)g.

The calculation for m an odd integer is similar. Then we can see recursively that

S2(x) = −S(x)g = −(−gn−1x)g = λx,

S3(x) = −gn−1S2(x) = −gn−1(λx) = −λgn−1x

S4(x) = −S3(x)g = −(−λgn−1x)g = λ2x

...

S2j(x) = −S2j−1(x)g = −(−λj−1gn−1x)g = λjx.

Page 51: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

46

Therefore, the least positive integer m such that Sm(x) = x is m = 2n. Of course,

this is even, and therefore S2n(g) = g as well. Hence, the order of S is 2n. �

2.4 kG, (kG)∗, and (co)commutativity

We will look more closely later at Hopf algebras, such as the Taft algebras, that are

neither commutative nor cocommutative.

Remark 2.4.1. Notice that the group algebra kG is always cocommutative. Thus,

by Proposition 2.2.11 the dual of a group algebra is always commutative. Conversely,

it is not difficult to check that kG is a commutative algebra exactly when G is an

abelian group. �

Next, we justify the name “group-like elements” with the following proposition.

Proposition 2.4.2. Let H be a Hopf algebra. The group-like elements G(H) form a

multiplicative group under the multiplication of H.

Proof. Let M denote the multiplication of H. Certainly the operation is associative

since M is associative. Furthermore, G(H) is closed under M , for if g, h ∈ G(H),

then since ∆ is an algebra morphism,

∆(gh) = ∆(g)∆(h)

= (g ⊗ g)(h⊗ h)

= (gh)⊗ (gh),

showing gh ∈ G(H). We already know that for ∆ to be an algebra morphism we

must have ∆(1) = 1 ⊗ 1, so 1 ∈ G(H). Finally, the elements of G(H) are invertible

by the antipode:

gS(g) =∑

g1S(g2) = ε(g)1 = 1.

Similarly, S(g)g = 1. �

Page 52: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

47

The concept of the group algebra has proved very important in the task of classifying

finite dimensional Hopf algebras. In fact, kG is the measure of triviality for Hopf

algebras. The following result is a small example.

Proposition 2.4.3. Let ch(k) 6= 2 and (H,M, u,∆, ε, S) be a Hopf algebra of dimen-

sion two. Then H is isomorphic to the group algebra kC2.

Proof. First, we note that we can write H = k1 ⊕ Ker(ε). Choose x ∈ Ker(ε) such

that {1, x} is a basis for H. Then {1⊗ 1, 1⊗ x, x⊗ 1, x⊗ x} is a basis for H ⊗H, so

write

∆(x) = α1⊗ 1 + β1⊗ x+ γx⊗ 1 + δx⊗ x

for some field elements α, β, γ, δ. Now ε is an algebra morphism, so Ker(ε) is a two-

sided ideal of the algebra H. Hence x2 = ax for some a ∈ x. Now, we apply the

counit property to x and obtain

x = αε(1)1 + βε(1)x+ γε(x)1 + δε(x)x

= α1 + βx,

and

x = α1ε(1) + β1ε(x) + γxε(1) + δxε(x)

= α1 + γx.

Therefore, α = 0 and β = γ = 1. Now since ∆ is an algebra morphism,

∆(x2) = ∆(x)∆(x)

= (1⊗ x+ x⊗ 1 + δx⊗ x)2

= a1⊗ x+ ax⊗ 1 + (2 + 4aδ + a2δ2)x⊗ x,

Page 53: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

48

where we have used that x2 = ax. On the other hand,

∆(x2) = ∆(ax)

= a∆(x)

= a1⊗ x+ ax⊗ 1 + aδx⊗ x.

So we obtain the quadratic equation in aδ, 2+4aδ+a2δ2 = aδ, with solutions aδ = −1

and aδ = −2. Now, we apply the antipode property to x, that is,∑S(x1)x2 =

ε(x)1 = 0. Recalling that S(1) = 1, we have,

S(1)x+ S(x) + δS(x)x = x+ S(x)(1 + δx) = 0.

If we write, S(x) = c1 + dx for some c, d ∈ k, then we have

0 = x+ (c1 + dx)(1 + δx)

= x+ c1 + cδx+ dx+ adδx

= c1 + (1 + cδ + d+ daδ)x.

Immediately, we see that c = 0, so we obtain the equation 1 + d+ daδ = 0. Thus we

cannot have aδ = −1, and we must have aδ = −2. Since ch(k) 6= 2, then a 6= 0, and

δ = −2a

. Finally, we can check that the element g = 1 + δx ∈ H is group-like, since

∆(g) = ∆(1 + δx)

= ∆(1) + δ∆(x)

= 1⊗ 1 + δ1⊗ x+ δx⊗ 1 + δ2x⊗ x

= 1⊗ 1 + δx⊗ 1 + 1⊗ δx+ δx⊗ δx

= (1 + δx)⊗ 1 + (1 + δx)⊗ δx

= (1 + δx)⊗ (1 + δx)

= g ⊗ g.

Page 54: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

49

Hence, there are two distinct group-like elements, and since group-likes must be lin-

early independent, a dimension argument yields H ∼= kC2. �

The preceding exercise is a special case of a conjecture made in 1975 by Kaplansky

[9] which was proven in 1993 by Zhu [23].

Theorem 2.4.4 (Kaplansky’s Conjecture). Let k be an algebraically closed field of

characteristic zero and H a Hopf algebra over k with prime dimension p. Then

H ∼= kCp. �

Remark 2.4.5. Recall that in §2.3 we claimed that Sweedler’s 4-dimensional algebra

was the smallest non-commutative, non-cocommutative Hopf algebra. The above

proposition justifies this claim since if H is a Hopf algebra with dimension two or

three, then H ∼= kC2 or H ∼= kC3, both commutative and cocommutative Hopf

algebras. �

We now give a condition under which a finite dimensional group algebra is self-dual.

That is, kG ∼= (kG)∗. We will need the following lemma:

Lemma 2.4.6. Let G and H be multiplicative groups. Then k{G×H} ∼= kG⊗ kH.

Proof. Define the map φ : k{G×H} → kG⊗ kH given by φ(g, h) = g⊗ h. It is easy

to check that φ is a bijection. We now check that φ is a morphism of bialgebras, and

therefore a Hopf morphism. Indeed, for g, g′ ∈ G and h, h′ ∈ H,

φ((g, h)(g′, h′)) = φ(gg′, hh′)

= gg′ ⊗ hh′

= (g ⊗ h)(g′ ⊗ h′)

= φ(g, h)φ(g′, h′),

Page 55: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

50

showing φ is an algebra morphism. Moreover,

(φ⊗ φ)∆k{G×H}(g, h) = (φ⊗ φ)((g, h)⊗ (g, h))

= g ⊗ h⊗ g ⊗ h

= (I ⊗ T ⊗ I) ◦ (∆G ⊗∆H)(g ⊗ h)

= ∆kG⊗kH(φ(g, h)),

as desired. �

Proposition 2.4.7. Let n ≥ 2 and k a field containing a primitive n-th root of unity.

Let Cn be the cyclic group of order n. Then kCn is self-dual.

Proof. Let λ be a primitive n-th root of unity. Denote kCn by H and write Cn = 〈c〉.

Now, {1, c, . . . , cn−1} forms a basis for H, and let {h∗1, h∗c , . . . , h∗cn−1} be the dual basis

for H∗. Hence dim(H) = dim(H∗) = n.

We already know from Proposition 1.5.3 that the group-like elements in H∗ are

the algebra morphisms. Thus if f ∈ G(H∗), then 1 = f(1) = f(cn) = (f(c))n.

Hence, f(c) = λi for some 0 < i < n − 1. On the other hand, if we choose any such

i, then there exists a unique algebra morphism, fi : H → k, given by fi(c) = λi.

In particular, fi(cj) = λij, and extended linearly to all of H. Now, the group-like

elements, G(H∗) form a multiplicative group by Proposition 2.4.2 and Fact 1.5.2(iii)

ensures that they are linearly independent. Thus, after counting dimensions, we see

that H∗ = kG(H∗).

We have only to show that G(H∗) ∼= Cn as a group, that is, fi = f i1. The proof

is by induction on i. Recall that the multiplication for H∗ is given by m(f ⊗ g)(c) =∑f(c1)g(c2) = f(c)g(c) since H is a group algebra. Then,

f i+11 (c) = m(f i1 ⊗ f1)(c) = f i1(c)f1(c) = (λi)λ = λi+1 = fi+1(c)

Then G(H∗) ∼= Cn, and therefore, (kCn)∗ = H∗ = kG(H∗) ∼= kCn. �

Page 56: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

51

Corollary 2.4.8. Let G be a finite abelian group. Then kG is self-dual.

Proof. First we note that G ∼= Cn1 × · · · ×Cnm for some natural numbers n1, . . . , nm.

We may induct on Lemma 2.4.6 to obtain that kG ∼= kCn1 ⊗ · · · ⊗ kCnm . Then we

may use the isomorphism θm defined in Corollary 1.4.3 to get

(kG)∗ ∼= (kCn1 ⊗ · · · ⊗ kCnm)∗ ∼= (kCn1)∗ ⊗ · · · ⊗ (kCnm)∗.

Using the preceding proposition, we obtain

(kG)∗ ∼= (kCn1)∗ ⊗ · · · ⊗ (kCnm)∗ ∼= kCn1 ⊗ · · · ⊗ kCnm

∼= kG,

showing that kG is self-dual. �

We end this section with an important result in the classification of Hopf alge-

bras. We have already stated that kG is always cocommutative and hence its dual

is always commutative. However, the following result from the collective work done

by Milnor and Moore [17], Cartier [2], and Kostant [11], gives the converse in the

finite dimensional case. That is, it ensures that all finite dimensional cocommutative

Hopf algebras over an algebraically closed field with characteristic 0 are isomorphic

to group algebras.

Theorem 2.4.9 (The Cartier-Kostant-Milnor-Moore Theorem). A cocommutative

Hopf algebra over an algebraically closed field of characteristic zero is a semidirect

product of a group algebra and the enveloping algebra of a Lie algebra. In particular,

a finite dimensional cocommutative Hopf algebra is a group algebra. �

Corollary 2.4.10. A finite dimensional commutative Hopf algebra over an alge-

braically closed field of characteristic 0 is the dual of a group algebra. �

Corollary 2.4.11. If H is a finite dimensional commutative, cocommutative Hopf

algebra over an algebraically closed field of characteristic 0, then H ∼= (kG)∗ for an

abelian group G. �

Page 57: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

52

2.5 Calculation in the dual

The goal of this section is to give some examples of calculations in a dual Hopf algebra,

in particular, the dual of a group algebra (kG)∗, and the dual of a Taft algebra

Hn(λ)∗. If H is the Hopf algebra in question, we will write, for both H and H∗, the

comultiplication as ∆ and the counit as ε, since the intention should be clear from

the context. In contrast, we will write the antipode of H (resp. H∗) as S (resp. S∗

the transpose map of S). We will also need to recall several important results of §1.4,

including Fact 1.4.6(i); that is, for any f ∈ H∗ and a, b ∈ H, f(ab) =∑f1(a)f2(b).

First we establish a useful lemma.

Lemma 2.5.1. Let H be a finite dimensional Hopf algebra with basis {ei}i. If the

dual basis is written {e∗i }i, then for any f ∈ H∗,

∆(f) =∑i,j

f(eiej)e∗i ⊗ e∗j .

Proof. Since {e∗i ⊗ e∗j}i,j is a basis for H∗ ⊗H∗, we can write

∆(f) =∑i,j

ai,je∗i ⊗ e∗j

for some scalars ai,j. By Fact 1.4.6 we have

f(eser) =∑i,j

ai,je∗i (es)e

∗j(er) = as,r.

Therefore, ∆(f) =∑

i,j f(eiej)e∗i ⊗ e∗j , as desired. �

Proposition 2.5.2 (Calculation of (kG)∗). Let G be a finite group and kG the asso-

ciated Hopf algebra. Let {pg}g∈G denote the basis for (kG)∗ which is dual to the basis

{g}g∈G for kG. Then the Hopf algebra structure of (kG)∗ can be expressed in terms

of {pg}g∈G by:

pgph = 0 if g 6= h, p2g = pg,

∑g∈G

pg = 1(kG)∗.

Page 58: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

53

∆(pg) =∑h∈G

pgh−1 ⊗ ph =∑h∈G

ph ⊗ ph−1g, ε(pg) = δ1,g.

S∗(pg) = pg−1.

Proof. Given the definition for multiplication in the dual of a coalgebra (§1.4), we

have for x ∈ G,

(pg ∗ ph)(x) =∑

pg(x1)ph(x2) = pg(x)ph(x) =

{1 , g = h = x0 , otherwise

.

Thus, pgph = 0 for g 6= h and p2g = pg. In other words, {pg}g∈G is a collection of

orthogonal idempotents. Next, we write 1(kG)∗ =∑

g∈G agpg, for some scalars ag. We

again use the definition for multiplication in the dual and consider, for any f ∈ (kG)∗

and g ∈ G,

(1(kG)∗ ∗ f)(g) =∑

1(kG)∗(g1)f(g2) = 1(kG)∗(g)f(g) = agf(g).

Now, if 1(kG)∗ is to be the identity element of (kG)∗ then we must have that ag = 1

for all g ∈ G. Thus, 1(kG)∗ =∑

g∈G pg.

For the comultiplication, we can begin by using Lemma 2.5.1 to obtain that

∆(pg) =∑x,y∈G

pg(xy)(px ⊗ py).

Now, the only nonzero terms above are those for which xy = g; that is, x = gy−1 or

y = x−1g, and in this case, pg(xy) = 1. Hence, the comultiplication in (kG)∗ is given

in terms of the dual basis by

∆(pg) =∑x∈G

px ⊗ px−1g =∑y∈G

pgy−1 ⊗ py.

Alternatively, we could write,

∆(pg) =∑xy=g

px ⊗ py.

The counit comes directly from the definition given in §1.4, that for any f ∈ H∗,

ε(f) = f(1). Thus ε(pg) = δ1,g. The antipode follows from the computation,

S∗(pg)(h) = pg(S(h)) = pg(h−1) for all g, h ∈ G. �

Page 59: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

54

Now we turn our attention to the Taft algebras Hn(λ). Recall that these algebras

are generated by the pair 〈g, x〉 with vector space basis given by {gixj} for 0 ≤ i, j ≤

n − 1. We will denote the dual basis in Hn(λ)∗ by {pi,j} where pi,j(gkx`) = 1 if and

only if i = k, j = `, otherwise, pi,j(gkx`) = 0.

Lemma 2.5.3. For any f ∈ Hn(λ)∗,

∆(f) =n−1∑

i,j,k,`=0

λjkf(gi+kxj+`)pi,j ⊗ pk,`

Proof. We can immediately write, using Lemma 2.5.1,

∆(f) =n−1∑

i,j,k,`=0

f(gixjgkx`)pi,j ⊗ pk,`.

Using the relations in Hn(λ), specifically xg = λgx, we have,

gixjgkx` = gixj−1(xg)gk−1x`

= gixj−1(λgx)gk−1x`

= λgixj−1g(xg)gk−2x`

...= (λk)gixj−1gkx1+`

...= (λjk)gi+kxj+`.

Therefore,

∆(f) =n−1∑

i,j,k,`=0

f(gixjgkx`)pi,j ⊗ pk,`

=n−1∑

i,j,k,`=0

f(λjkgi+kxj+`)pi,j ⊗ pk,`

=n−1∑

i,j,k,`=0

λjkf(gi+kxj+`)pi,j ⊗ pk,`,

as desired. �

Theorem 2.5.4. The Sweedler algebra H2(−1) is self-dual.

Page 60: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

55

Proof. It is possible to establish this result by showing that the Sweedler algebra is

the unique non-commutative, non-cocommutative Hopf algebra of dimension four and

hence must be self-dual, but here we calculate the dual explicitly for emphasis. We

will take as a basis that which is dual to {1, g, x, gx}, the standard basis in H2(−1).

We will denote this basis by {p1, pg, px, pgx}. Notice that as a special case of the

Taft algebra (n = 2), this set corresponds to the notation above for the dual basis

when written as {p0,0, p1,0, p0,1, p1,1}, respectively. We begin by using Lemma 2.5.3 to

calculate the comultiplication for each of these elements.

Comultiplication of p1: By 2.5.3, we can write

∆(p1) =1∑

i,j,k,`=0

(−1)jkp1(gi+kxj+`)pi,j ⊗ pk,`.

Using the lemmas to aid calculation, one can check that

∆(p1) = p1 ⊗ p1 + pg ⊗ pg. (2.5)

Comultiplication of pg: Similarly, some computation gives

∆(pg) = p1 ⊗ pg + pg ⊗ p1 (2.6)

Comultiplication of px: Using Lemma 2.5.3, the only terms which are nonzero have

i + k = 0 mod 2 and j + ` = 1. Notice we require the sum j + ` to be exactly one

since it represents an exponent on the element x ∈ H2(−1). If this exceeds one, then

the resulting product gi+kxj+` = 0 for all i, k. This results in the comultiplication,

∆(px) = p1 ⊗ px + pg ⊗ pgx + px ⊗ p1 − pgx ⊗ pg. (2.7)

Comultiplication of pgx: Similarly, we have that

∆(pgx) = p1 ⊗ pgx + pg ⊗ px − px ⊗ pg + pgx ⊗ p1. (2.8)

Using the definition of the counit in a dual Hopf algebra, it is easy to see that ε(p1) =

1, while ε(pg) = ε(px) = ε(pgx) = 0.

Page 61: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

56

Next, we seek a multiplication table in H2(−1)∗ in terms of {p1, pg, px, pgx}. We

use the fact that for any α, β ∈ H∗ and h ∈ H, the relation (α∗β)(h) =∑α(h1)β(h2)

always holds. We know that in H2(−1)∗, we can write any such product as

αβ = c(αβ)1 p1 + c(αβ)

g pg + c(αβ)x px + c(αβ)

gx pgx,

for some scalars c(αβ)1 , c

(αβ)g , c

(αβ)x , c

(αβ)gx . Consider the following example calculation:

(pgpx)(x) = pg(g)px(x) + pg(x)px(1) = 1, ⇒ c(pgpx)x = 1.

Similar computations yield the following multiplication table, with products written

as “element from left column” times “element from top row”:

∗ p1 pg px pgxp1 p1 pgxpg pg pxpx pxpgx pgx

where blank cells are understood to be zero. We are now in a position to see that

H2(−1)∗ ∼= H2(−1). Indeed, if we set

1∗ = p1 + pg,

g∗ = p1 − pg,

x∗ = px − pgx,

we can check that these elements mimic the roles of 1, g, x ∈ H2(−1), respectively.

Using the multiplication table above, we can immediately see that g∗x∗, the element

that should mimic gx ∈ H2(−1), can be written as g∗x∗ = −(px + pgx). It is easy

to check that the set {1∗, g∗, x∗, g∗x∗} is a basis for H2(−1)∗. Next, we verify each

of the algebraic relations. Let f ∈ H2(−1)∗, then for scalars α, β, γ, δ, let h =

Page 62: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

57

α1 + βg + γx+ δgx. Consider

(1∗ ∗ f)(h) =∑

1∗(h1)f(h2)

= 1∗(1)f(α1) + 1∗(g)f(βg) + 1∗(g)f(γx)

+ 1∗(x)f(γ1) + 1∗(1)f(δgx) + 1∗(gx)f(δg)

= f(α1) + f(βg) + f(γx) + 0 + f(δgx) + 0

= f(α1 + βg + γx+ δgx)

= f(h).

Similarly, (f ∗ 1∗)(h) = f(h). Hence, 1∗ is the identity of H2(−1)∗. Now, using the

multiplication table, consider

(g∗)2 = (p1 − pg)2 = p21 − p1pg − pgp1 + p2

g = p1 + pg = 1∗, and

(x∗)2 = (px − pgx)2 = p2x − pxpgx − pgxpx + p2

gx = 0.

So g∗ squares to the identity and x∗ squares to 0, as desired. Moreover,

x∗g∗ = (px − pgx)(p1 − pg)

= pxp1 − pxpg − pgxp1 + pgxpg

= px + 0 + 0 + pgx

= −g∗x∗.

So H2(−1)∗ ∼= H2(−1) as an algebra. To check the coalgebra structure, we use

equations (2.5) and (2.6) to see

∆(g∗) = ∆(p1)−∆(pg)

= p1 ⊗ p1 + pg ⊗ pg − (p1 ⊗ pg + pg ⊗ p1)

= p1 ⊗ (p1 − pg)− pg ⊗ (p1 − pg)

= (p1 − pg)⊗ (p1 − pg)

= g∗ ⊗ g∗,

Page 63: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

58

and of course, ε(g∗) = g∗(1) = 1. From equations (2.7) and (2.8) we observe that

∆(x∗) = ∆(px)−∆(pgx)

= p1 ⊗ px + pg ⊗ pgx + px ⊗ p1 − pgx ⊗ pg

− (p1 ⊗ pgx + pg ⊗ px − px ⊗ pg + pgx ⊗ p1)

= p1 ⊗ (px − pgx)− pg ⊗ (px − pgx)

+ px ⊗ (p1 + pg)− pgx ⊗ (p1 + pg)

= (p1 − pg)⊗ (px − pgx) + (px − pgx)⊗ (p1 + pg)

= g∗ ⊗ x∗ + x∗ ⊗ 1∗,

and furthermore, ε(x∗) = x∗(1) = 0, as desired. Finally, we check the antipodes.

Again, consider the general element h = α1 + βg + γx+ δgx. Then

S∗(g∗)(h) = g∗(S(h)) = g∗(α1 + βg − γgx+ δx) = α− β.

Hence S∗(g∗) = p1 − pg = g∗. For x∗, we have that

S∗(x∗)(h) = x∗(S(h)) = x∗(α1 + βg − γgx+ δx) = γ + δ.

Thus, S∗(x∗) = px + pgx = −g∗x∗, and the verification is complete. �

What, if any, of the computations above can be generalized to the duals of the

larger Taft algebras, Hn(λ)∗ (n > 2)? As before, we will attempt to compute elements

g∗, x∗ ∈ Hn(λ)∗ which correspond exactly to the generators 〈g, x〉 of Hn(λ), giving

rise to the same Hopf algebra structure. Recall that we will write the basis dual to

{gixj}, 0 ≤ i, j ≤ n − 1, as {pi,j}. We have already calculated a formula for the

comultiplication in terms of this basis in Lemma 2.5.3. We begin much as we did

for the Sweedler algebra, and explicitly write the comultiplication for each pi,j with

j = 0. That is, the comultiplication for dual basis elements corresponding to the

group-like elements {1, g, g2, . . . , gn−1} in Hn(λ).

Page 64: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

59

The formula in Lemma 2.5.3 now says that the nonzero terms in ∆(p0,0) are those

in which i+ k = 0 mod n, and j + ` = 0. There are n possible cases, one for each of

the values 0 ≤ i ≤ n− 1, and hence we have

∆(p0,0) = p0,0 ⊗ p0,0 + p1,0 ⊗ pn−1,0 + p2,0 ⊗ pn−2,0 + · · ·+ pn−1,0 ⊗ p1,0. (2.9)

In general, to determine ∆(pr,0) (0 ≤ r ≤ n − 1), the process is the same, and we

have that

∆(pr,0) =∑

ps,0 ⊗ pt,0, (2.10)

where the sum is taken over all pairs (s, t) such that 0 ≤ s, t ≤ n − 1 and s + t = r

mod n. Just as in the Sweedler algebra, each coefficient is known to be 1 since

j = 0⇒ λjk = 1. In terms of multiplication, we can check, using similar methods to

those used for the Sweedler algebra, that the following relations hold for the elements

{pi,0}:

pi,0 ∗ pk,0 = δi,k,n−1∑i=0

pi,0 = 1Hn(λ)∗ . (2.11)

We are now in a position to prove an interesting proposition regarding the group-

like elements in Hn(λ)∗. This is a first step toward comparing Hn(λ) and Hn(λ)∗,

implying that, at least their group-like structures are isomorphic.

Proposition 2.5.5. G(Hn(λ)∗) ∼= Cn.

Proof. Recall that if f ∈ G(Hn(λ)∗) then f is an algebra morphism. Therefore,

writing f =∑

i,j ai,jpi,j, for some scalars ai,j, we can check that 1 = f(1) = a0,0, and

moreover 1 = f(gn) = f(g)n = (a1,0)n. Hence, a1,0 = ξ, some n-th root of unity. In

addition, we have the relations:

a2,0 = f(g2) = f(g)2 = (a1,0)2,

a3,0 = f(g3) = f(g)3 = (a1,0)3,

...an−1,0 = f(gn−1) = f(g)n−1 = (a1,0)

n−1.

Page 65: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

60

We also have 0 = f(0) = f(xn) = f(x)n = (a0,1)n, so a0,1 = 0, and the relations

a0,2 = f(x2) = f(x)2 = (a0,1)2 = 0,

a0,3 = f(x3) = f(x)3 = (a0,1)3 = 0,

...a0,n−1 = f(xn−1) = f(x)n−1 = (a0,1)

n−1 = 0.

Hence for all j 6= 0, ai,j = f(gixj) = f(gi)f(xj) = ai,0a0,j = 0. Therefore,

f = p0,0 + ξp1,0 + ξ2p2,0 + · · ·+ ξn−1pn−1,0.

Now λ is a primitive n-th root of unity, so the element g∗ = p0,0 +λp1,0 +λ2p2,0 + · · ·+

λn−1pn−1,0 can generate all such f ∈ G(Hn(λ)∗). Indeed, since λ is primitive, there

must exist m < n such that λm = ξ. Using the multiplicative relations in (2.11), we

observe that

(g∗)m = (p0,0 + λp1,0 + λ2p2,0 + · · ·+ λn−1pn−1,0)m

= p0,0 + (λm)p1,0 + (λm)2p2,0 + · · ·+ (λm)n−1pn−1,0

= p0,0 + ξp1,0 + ξ2p2,0 + · · ·+ ξn−1pn−1,0

= f.

Again we use (2.11):

(g∗)n = (p0,0 + λp1,0 + λ2p2,0 + · · ·+ λn−1pn−1,0)n

= p0,0 + (λn)p1,0 + (λn)2p2,0 + · · ·+ (λn)n−1pn−1,0

= p0,0 + p1,0 + p2,0 + · · ·+ pn−1,0

= 1Hn(λ)∗ .

Therefore, G(Hn(λ)∗) = 〈g∗〉 ∼= Cn. �

Page 66: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Chapter 3: Hopf algebra actions on rings

In practice, we have demonstrated in §2.4 that when working over an algebraically

closed field of characteristic zero, to study finite dimensional cocommutative (resp.

commutative) Hopf algebras is to study group algebras kG (resp. duals of group

algebras (kG)∗). As a consequence, we will see that to generate the most general ac-

tions on algebras, we will have to study Hopf algebras that are both non-commutative

and non-cocommutative. In fact, we will refer to any finite dimensional Hopf algebra

which is either commutative or cocommutative as a trivial Hopf algebra. We begin

with some definitions.

3.1 Definitions and Examples

Throughout this chapter, H is a Hopf algebra over the field k with comultiplication

∆, counit ε, and antipode S. A is a k-algebra. First we recall the definition of a

module.

Definition 3.1.1. Let (A,m, u) be a k-algebra. Then a left A-module is a k-space

M with a k-linear map γ : A⊗M →M such that the following diagrams commute:

A⊗ A⊗Mm⊗ I

- A⊗M

A⊗M

I ⊗ γ

?

γ- M

γ

?

k ⊗Mu⊗ I

- A⊗M

M

γ

?-

where I denotes the identity map and the unadorned diagonal arrow indicates scalar

multiplication.

61

Page 67: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

62

Just as algebras and coalgebras are dual constructions, there is an analogous dualized

structure for modules.

Definition 3.1.2. Let (C,∆, ε) be a k-coalgebra. Then a right C-comodule is a

k-space M with a k-linear map ρ : M → M ⊗ C such that the following diagrams

commute:

Mρ- M ⊗ C

M ⊗ C

ρ

?

ρ⊗ I- M ⊗ C ⊗ C

I ⊗∆

?

Mρ- M ⊗ C

M ⊗ k

I ⊗ ε

?

⊗1-

where I denotes the identity map.

The maps in a comodule are defined in the obvious way:

Definition 3.1.3. Let (M,ρM) and (N, ρN) be right C-comodules and f : M → N .

f is called a comodule map if ρN ◦ f = (f ⊗ I) ◦ ρM .

Now we can define a Hopf algebra action.

Definition 3.1.4. A k-algebra A is a left H-module algebra if A is a left H-module

and the following hold:

h · (ab) =∑

(h1 · a)(h2 · b) (3.1)

h · 1A = ε(h)1A (3.2)

for all a, b ∈ A and h ∈ H, where we have denoted the action of the element h ∈ H

on the element a ∈ A by h · a. In this case, we say that H acts on A.

Right H-module algebras can be defined analogously, but we will primarily be

concerned with left actions, so henceforth, the phrase “H acts on A” should be inter-

preted as the left action as given above, unless explicitly stated otherwise.

Page 68: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

63

Definition 3.1.5. An algebra (A,M, u) is called a right H-comodule algebra if

(i) A is a right H-comodule

(ii) M and u are right H-comodule maps.

Proposition 3.1.6. Let H be a finite-dimensional Hopf algebra. Then A is a left

H-module algebra if and only if A is a right H∗-comodule algebra.

Proof. See Montgomery [18]. �

Armed with a definition for Hopf actions, we turn our attention to the major

goal of this thesis. We will demonstrate that Hopf algebras are natural candidates

to replace finite groups as actors on rings. Moreover, we wish to ask what remains

invariant under such an action and will compute several specific examples. Before

this however, we will see that we need an updated idea of invariance.

Traditionally, if an element a is invariant under an action, we require that h ·a = a

for all h ∈ H. However, in the case of Hopf algebras, problems arise. For a simple

example, suppose a, b ∈ A both have the property that h · a = a and h · b = b for all

h ∈ H. Further, suppose that H contains, for instance, a primitive element x. That

is, ∆(x) = 1⊗ x+ x⊗ 1. Then consider,

x · (ab) = (1 · a)(x · b) + (x · a)(1 · b) = (a)(b) + (a)(b) = 2ab.

Although a and b are both “fixed” under the action of H, the same is not necessarily

true of their product. Hence, the set {a ∈ A |h · a = a ∀h ∈ H} may not be a

subalgebra of A. We wish to preserve this property, after all it is true for group

actions on rings, which motivates the following alternative definition of the invariant

algebra.

Page 69: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

64

Definition 3.1.7. Let A be an H-module algebra. Then the algebra of invariants

under the action of H is the set

AH = {a ∈ A | h · a = ε(h)a, ∀h ∈ H}.

We may alternatively refer to this set as the fixed algebra.

Notice that this definition is consistent with the traditional one whenever H = kG

since ε(g) = 1 for all g ∈ G. Moreover, this will solve the dilemma in the example

above since x ∈ P (H) guarantees that ε(x) = 0. Thus, this definition has the

surprising consequence that a primitive element (or any other element in Ker ε) has

the property that it “fixes” only those elements which it sends to zero. Though this

may at first seem bizarre, Definition 3.1.7 is the correct generalization, since it indeed

produces a subalgebra.

Proposition 3.1.8. Let A be an H-module algebra. Then AH is a subalgebra of A.

Proof. The subspace condition is easily checked. The property 3.2 insures that 1A ∈

AH . Certainly the multiplication is associative. We must show that the multiplication

is closed. Take any a, b ∈ AH . Then

h · (ab) =∑

(h1 · a)(h2 · b)

=∑

(ε(h1)a)(ε(h2)b)

=∑

ε(h1)ε(h2)(ab)

= ε(∑

ε(h1)h2

)(ab), (linearity of ε)

= ε(h)(ab), (counit property)

showing that ab ∈ AH . �

Sometimes we will want to talk about the elements of A invariant under only a

single element h ∈ H. We will denote this set by Ah = {a ∈ A |h ·a = ε(h)a}. By the

Page 70: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

65

same process as above, one can check that Ah is a subalgebra of A. Now we consider

actions of our prototypical examples, kG and (kG)∗.

Example 3.1.9. Let H = kG and suppose that H acts on A. Since ∆(g) = g ⊗ g,

g ·(ab) = (g ·a)(g ·b) for all a, b ∈ A, g ∈ G. Moreover, if a ∈ AH , then g ·a = ε(g)a = a

for all g ∈ H. Thus the action of a group algebra mimics the action of groups acting

as automorphisms. More generally, we can say that for any H-module algebra A, the

set of group-like elements G(H) acts like a group of automorphisms on A. �

Example 3.1.10. Let G be a finite group and A an algebra graded by G. Then, write

A =⊕

g∈GAg, such that AgAh ⊆ Agh, ∀g, h ∈ G and in particular 1A ∈ A1G. Now,

any a ∈ A can be written as a =∑

g∈G ag, where we let ag denote the g-component

of a. Let H = (kG)∗ and denote the dual basis by {pg}g∈G. It is not difficult to check

that the action pg · a = ag for all g ∈ G makes A an H-module. Using the grading

and the comultiplication in (kG)∗ calculated in Proposition 2.5.2, we see that for any

g ∈ G,

pg · (ab) =∑xy=g

px(a)py(b) =∑xy=g

(ax)(by).

On the other hand,

(ab)g =

((∑x∈G

ax

)(∑y∈G

by

))g

=∑xy=g

(ax)(by),

which establishes property (3.1) in Definition 3.1.4. Furthermore, for any grading of

A, we must have that (1A)g = δ1,g, satisfying property (3.2).

Conversely, suppose that A is a left (kG)∗-module algebra. Alternatively, this

means that A is a right (kG)-comodule algebra with some structure map ρ. We will

begin by writing ρ(a) =∑

g ag ⊗ g and attempt to calculate the “coefficients” ag.

Using the fact that A is a comodule we have that (ρ ⊗ I) ◦ ρ = (I ⊗ ∆) ◦ ρ and

thus we have (ag)h = δg,hag. Therefore, ρ(ag) = ag ⊗ g. Then for each g, define

Page 71: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

66

the subspace Ag = {ag | a ∈ A} and by our computation we have that Ag ∩ Ah = 0

for g 6= h. Moreover, the comodule structure requires that for all a ∈ A we have

a ⊗ 1 = (I ⊗ ε)ρ(a). Thus,∑

g ag = a for all a ∈ A. Therefore we may write

A =⊕

g∈GAg so that A is a G-graded vector space. Furthermore, this decomposition

provides an algebra grading. Indeed, let ag ∈ Ag and bh ∈ Ah and recall that the

multiplication of A must be a comodule map to see that ρ(agbh) = ρ(ag)ρ(bh) =

agbh ⊗ gh. That is, AgAh ⊆ Agh for all g, h ∈ G. Similarly, since u is a comodule

map, 1A = u(1) ∈ A1G. �

Remark 3.1.11. The importance of the two previous examples cannot be overstated.

Recall that over algebraically closed fields of characteristic 0, if H is cocommutative,

H ∼= kG. On the other hand, when H is commutative, H ∼= (kG)∗. Hence, when

working over algebraically closed fields of characteristic 0, to study actions of co-

commutative Hopf algebras is to study group actions. Similarly, to study actions of

commutative Hopf algebras is to study G-graded algebras for finite groups G. �

In the sections that follow, we will introduce some specific examples of actions of

Hopf algebras which are both non-commutative and non-cocommutative on various

rings. In terms of the algebras being acted upon, we will be most interested in two

specific examples: polynomial rings k[x1, . . . , xn] and the so-called quantum plane,

kq[x, y] = k〈x, y | yx− qxy〉. In particular, both of these algebras share the property

that they are graded by degree. That is, they can be written as A =⊕

nAn where each

An is the vector space of all monomials of degree n and AiAj ⊂ Ai+j. Additionally,

all of the actions we consider will preserve degree. That is, if h ∈ H and ai ∈ Ai,

then h · ai ∈ Ai. The following propositions deal with matrix representations for

automorphisms on these algebras, and will prove useful as we progress.

Proposition 3.1.12. Let A = k[x1, . . . , xn], a polynomial ring and suppose H acts

on A. Conjugation in the representation of H induces an automorphism of A.

Page 72: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

67

Proof. The result follows since conjugation is accomplished by multiplication of an

invertible matrix. Any such matrix represents an automorphism of A. �

Proposition 3.1.13. Let q = −1 and A = kq[x, y]. Suppose σ acts as a degree pre-

serving automorphism of A. Then the 2×2 matrix representing σ as a transformation

on A1 must be either a diagonal or skew diagonal matrix.

Proof. First write

σ 7→[a bc d

].

Now σ is an automorphism, so it must be invertible, thus ad 6= bc. Moreover

σ · xy = (σ · x)(σ · y) = (ax+ cy)(bx+ dy) = abx2 + (ad− bc)xy + cdy2,

and

σ · yx = (σ · y)(σ · x) = (bx+ dy)(ax+ cy) = abx2 + (bc− ad)xy + cdy2.

From the relation σ(yx) = −σ(xy), we obtain the equations ab = 0 and cd = 0.

Moreover, from ad 6= bc, we note that we cannot have both a = c = 0. Similarly, we

cannot have both b = d = 0. It is not difficult to check that if b = 0 we obtain a

diagonal representation and when b 6= 0 we obtain a skew diagonal representation. �

Proposition 3.1.14. Let A = kq[x, y] and q2 6= 1. If σ is a degree preserving

automorphism of A, then σ must be represented in 2 × 2 matrices by a diagonal

matrix.

Proof. As before, begin by letting

σ 7→[a bc d

],

such that the above matrix has full rank, that is ad 6= bc. Consider,

σ · xy = (σ · x)(σ · y) = (ax+ cy)(bx+ dy) = abx2 + (ad+ qbc)xy + cdy2,

Page 73: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

68

and

σ · yx = (σ · y)(σ · x) = (bx+ dy)(ax+ cy) = abx2 + (bc+ qad)xy + cdy2.

We use the relation σ · yx = q(σ · xy) and equate coefficients to obtain the following

system of equations:

0 = ab(1− q)

0 = cd(1− q)

0 = bc(1− q2).

Some computation gives that the only solution is b = c = 0, implying σ is represented

by a diagonal matrix. �

Lemma 3.1.15. Any matrix over C with finite order is diagonalizable.

Proof. Let M be such a matrix and suppose that Mn = I. Then the minimal poly-

nomial for M divides zn − 1, a polynomial with distinct roots over C. �

We conclude this section with another important idea that will be utilized throughout.

Definition 3.1.16. Let A be an H-module algebra and I ⊆ H a Hopf ideal. The

action of H on A is said to be factorable through I if it induces an action of H/I

via h · a = h · a such that A is an H/I-module algebra.

Proposition 3.1.17. Let A be an H-module algebra and I ⊆ H a Hopf ideal. Then

the action of H on A is factorable through I if and only if for every a ∈ A, h ∈ I,

h · a = 0.

Proof. Suppose the action is factorable through I. Then if h ∈ I, 0 = h ∈ H/I. The

module condition implies that for any a ∈ A, 0 = 0 · a = h · a = h · a. Conversely,

suppose that h · a = 0 for all h ∈ I and a ∈ A. We wish to show that the action

Page 74: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

69

h · a = h · a endows A with the structure of an H/I-module algebra. The module

condition is not difficult to check and when h /∈ I, the other conditions are easily

established as well. We must verify that equations (3.1) and (3.2) hold in the case

that h ∈ I. As above, since h ∈ I, h = 0 so that h · a = 0 for all a ∈ A. For any

a, b ∈ A we must have,∑(h1 · a)(h2 · b) =

∑(h1 · a)(h2 · b) = 0,

by definition of the comultiplication in H/I and the fact that either h1 ∈ I or h2 ∈ I

for each term in the sum above. (recall ∆(h) ∈ I ⊗H +H ⊗ I for a Hopf ideal). This

is consistent with 0 = h · ab and thus establishes equation (3.1). Moreover,

ε(h)1 = ε(h)1 = 0,

again by the definition of ε and the requirement that ε(I) = 0 for a Hopf ideal. This

coincides with the condition h · 1 = 0 and therefore equation (3.2) is established. �

Remark 3.1.18. Suppose that an element h ∈ H generates a Hopf ideal. Now

imagine that we find a representation ρ : H → M(k) into an appropriate matrix

ring so that ρ gives rise to an action on the degree graded algebra A =⊕

nAn and

furthermore ρ(h) acts as the zero map on all degrees. That is, h · An = 0 for all

n. The preceding proposition says that the action factors through I = 〈h〉 and we

may as well regard the action as simply that of H/I. As we progress, we will want

to consider actions of Hopf algebras with a specified dimension. Hence, we will not

consider representations giving rise to actions that factor since H/I has dimension

strictly less than that of H. �

In Chapter 4 we will prove that there are no actions by nontrivial semisimple Hopf

algebras with dimension less than or equal to 15 on polynomials. However, if we relax

the condition that the Hopf algebra be semisimple we can construct an example of a

polynomial H-module algebra. The rest of this chapter is dedicated to this example.

Page 75: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

70

3.2 Actions of the Taft algebras

Motivated by the comments of Remark 3.1.11 we turn our attention to the actions

of nontrivial Hopf algebras, that is, those that are both non-commutative and non-

cocommutative. The Taft algebras Hn(λ) provided our first examples of such Hopf

algebras and we will exhaust the study of their actions on polynomial rings here.

Notice that Hn(λ) is not semisimple since the ideal generated by x is nilpotent.

From this point forward, we will assume that our field k is algebraically closed of

characteristic zero. In fact, unless explicitly stated otherwise, we assume k = C.

3.2.1 Actions on k[u, v]

We begin with Hopf algebra actions for H = Hn(λ) on polynomials in two variables;

that is, A = k[u, v]. We seek a representation of Hn(λ) as an algebra in M2(C). Recall

from Example 2.3.5 that we must satisfy the following relations:

gn = 1, xn = 0, xg = λgx, (3.3)

where λ is a primitive n-th root of 1. We begin with a lemma.

Lemma 3.2.1. 〈x〉 is a Hopf ideal.

Proof. 〈x〉 is a coideal since ε(x) = 0 and ∆(x) = g⊗x+x⊗ 1 ∈ Hn(λ)⊗〈x〉+ 〈x〉⊗

Hn(λ). Moreover, S(〈x〉) ⊆ 〈x〉 since S(x) = −gn−1x ∈ 〈x〉. �

It is not difficult to show that if x acts as the zero map on degree one monomials,

i.e. it is represented by the zero matrix, then by induction it acts as the zero map on

all of A. This will cause the action of H to factor, and by the rationale of Remark

3.1.18 we will not consider such representations. Recall that since A is a polynomial

algebra, by Proposition 3.1.12, we may make a change of variables in A so that x is

Page 76: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

71

in its Jordan form. Now x is nilpotent, but is not the zero matrix, so it must have

Jordan form

x 7→(

0 10 0

).

Then, writing g 7→ (gij) in this basis and using the relations in (3.3), we obtain the

following equations:

g22 = λg11

g21 = 0.

Thus, we can write

g 7→(α β0 λα

)for some α, β ∈ C. We must now ensure that Hn(λ) can act as a Hopf algebra on

A = k[u, v]. It is sufficient to check the generators for both Hn(λ) and A. That

is, it is sufficient to guarantee that x · (uv) = x · (vu) and g · (uv) = g · (vu). The

matrix for g is invertible provided α 6= 0 so that g, a group-like element, is in fact an

automorphism. From this, it is clear the relation for g holds. Consider the relation

for x. Given the coalgebra structure in Example 2.3.5, we require that

x · (uv) = (g · u)(x · v) + (x · u)v = αu2,

and

x · (vu) = (g · v)(x · u) + (x · v)u = u2

be equal, from which we obtain that α = 1. Thus the following representation for

Hn(λ) gives a Hopf algebra action on A = k[u, v]:

x 7→(

0 10 0

), g 7→

(1 β0 λ

),

for some β ∈ C. Suppose that β 6= 0. The eigenvalues of the matrix representation

for g are 1 and λ, with corresponding eigenvectors 〈1, 0〉 and 〈 βλ−1

, 1〉, respectively.

Hence, we can make another change of variables and write

A = k[u, v] ∼= k[u,β

λ− 1u+ v],

Page 77: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

72

in the new eigenbasis. Of course, with these generators, g is represented by a diagonal

matrix, (1

λ

).

Moreover, the representation for x in this basis is the same,

x 7→(

0 10 0

)since x · u = 0 and x · ((β/(λ − 1))u + v) = u. Therefore, we need only to consider

the case when β = 0.

We now calculate the fixed algebra. If we denote the set of elements fixed by g

as Ag, then it is clear that AHn(λ) ⊆ Ag. We see immediately that u ∈ Ag since

g · u = u = ε(g)u. Moreover, we notice that for i ∈ N that g · (vi) = g · (vi−1)(g · v)

and by induction on i, we obtain that g · (vi) = (g · v)i. Therefore, g · (vi) = (λv)i, so

that if i = n, g · (vn) = ε(g)vn. Hence, Ag = k[u, vn].

Now, we need to find all elements of k[u, vn] which are also fixed by x, that is,

such that x · a = ε(x)a = 0. It is immediately clear that x · u = 0, but again we must

consider powers of v. Note that x ·(v2) = (g ·v)(x ·v)+(x ·v)v = λuv+uv = (λ+1)uv.

By induction on `, we can show:

Lemma 3.2.2. The following relation holds for all ` ∈ N.

x · (v`) = (λ`−1 + λ`−2 + · · ·+ λ+ 1)uv`−1.

Proof. If the assertion is true for ` > 2, then

x · (v`+1) = (g · (v`))(x · v) + (x · (v`))v

= (λv)`u+ (λ`−1 + λ`−2 + · · ·+ 1)uv`

= (λ` + λ`−1 + λ`−2 + · · ·+ 1)uv`,

as desired. �

Page 78: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

73

So we have, x · (vn) = (λn−1 + λn−2 + · · ·+ 1)uvn−1 = 0, and therefore

AHn(λ) = Ag = k[u, vn].

3.2.2 Actions on k[u1, . . . , ur] for r ≥ 3

In the preceding subsection, we determined all possible Hopf algebra actions of Taft

algebras on polynomials in two variables. In this section, we show that all Hopf

algebra actions on polynomials in three or more variables reduce to actions on two

variables in the following sense:

Theorem 3.2.3.

(a) In terms of the representation of H (as an algebra), there is only one possible

Jordan form for x, and in this representation, x2 = 0 (and therefore, xn = 0).

(b) Given this matrix representation, we may write the matrix representing g as:(G 00 I

),

where I and 0 represent the identity and zero matrices of the appropriate sizes,

and G is the 2 × 2 matrix that represents g in the preceding subsection, §3.2.1.

That is, G represents g in defining a Hopf algebra action of Hn(λ) on polynomials

in two variables.

(c) AHn(λ) = k[u1, un2 , u3, . . . , ur] ∼=

(k[u1, u2]

Hn(λ))

[u3, . . . , ur]. That is, the fixed

algebra for A = k[u1, . . . , ur] is just the algebra formed by appending r − 2

independent variables to the fixed algebra calculated for k[u, v] in the preceding

subsection.

Proof of this theorem will be the goal of this subsection. As before, x cannot be

represented by the zero matrix, but does have zero as its only eigenvalue. To begin,

we restrict the possible representations with the following lemma.

Page 79: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

74

Lemma 3.2.4. Suppose we define a Hopf algebra action for Hn(λ) on A. The Jordan

form of the matrix representing x can have blocks no larger than 2× 2.

Proof. Without loss of generality we may assume that the non-zero Jordan blocks

occur in the uppermost leftmost corner, in order of decreasing size. Suppose that x

contains a block with dimension greater than 2× 2. We write,

x 7→

0 1 0 . . .0 0 1 . . .0 0 0 . . ....

......

. . .

*

.

Let g = (gij) for i, j = 1, . . . , r. We have the products:

xg 7→

g21 g22 g23

g31 g32 g33

0 0 0*

* *

and gx 7→

0 g11 g12

0 g21 g22

0 g31 g32

*

* *

.

Then from the relation, xg = λgx, we obtain that 0 = g21 = g31 = g32, g33 = λg22 =

λ2g11, and g23 = λg12. Thus for some α, β, γ ∈ C, g is represented byα β γ0 λα λβ0 0 λ2α

*

* *

.

Now consider the action of x on the product, u1u2.

x · (u1u2) = (g · u1)(x · u2) + (x · u1)u2

= αu21,

and

x · (u2u1) = (g · u2)(x · u1) + (x · u2)u1

= u21.

Page 80: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

75

Thus, as was the case for actions on k[u, v], we have that α = 1. Now, we let x act

on the product, u2u3.

x · (u2u3) = (g · u2)(x · u3) + (x · u2)u3

= (βu1 + λu2)u2 + u1u3

= λu22 + βu1u2 + u1u3,

but

x · (u3u2) = (g · u3)(x · u2) + (x · u3)u2

= (γu1 + λβu2 + λ2u3)u1 + u22,

by which we see that no matter what values we choose for β and γ, we must have

that λ = 1, a contradiction. �

Proof of Theorem 3.2.3(a). We need to show that x is represented by

x 7→

0 10 0

0

.

Conversely, suppose that x has the form0 10 0

0 10 0

*

.

As before, writing g = (gij) where i, j = 1, . . . , r, and using the relation xg = λgx,

we obtain the equations g22 = λg11, g24 = λg13, g42 = λg31, and g44 = λg33. Now,

consider the action of x on the product u1u2.

x · (u1u2) = (g · u1)(x · u2) + (x · u1)u2

= (r∑`=1

g`1u`)u1,

Page 81: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

76

and

x · (u2u1) = (g · u2)(x · u1) + (x · u2)u1

= u21.

Hence, g11 = 1 and g`1 = 0 for all ` 6= 1. Similarly, applying x to the product u3u4

yields that g33 = 1 and g`3 = 0 for all ` 6= 3. Then, applying x to u2u4, we see that

x · (u2u4) = (g · u2)(x · u4) + (x · u2)u4

= (g12u1 + λu2 + · · ·+ gr2ur)u3 + u1u4,

but

x · (u4u2) = (g · u4)(x · u2) + (x · u4)u2

= (g14u1 + g34u3 + λu4 + · · ·+ gr4ur)u1 + u3u2.

Comparing coefficients in either of the terms containing u2u3 or u1u4, implies that

λ = 1, again a contradiction. �

Proof of Theorem 3.2.3(b). Now that we know the Jordan form of x, we can ex-

plicitly determine g. As before, the relation xg = λgx yields the following:

g 7→

α β0 λα

*

* *

.

Then, if we assume that Hn(λ) acts on A and consider x · (u1u2) and x · (u2u1), we

observe that we must have 0 = gi1 for all i ≥ 2, and g11 = 1 (thus g22 = λ). Now, we

consider the general product, u2uj, when j ≥ 3. In this case,

x · (u2uj) = (g · u2)(x · uj) + (x · u2)uj

= u1uj,

Page 82: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

77

and

x · (uju2) = (g · uj)(x · u2) + (x · uj)u2

= (r∑i=1

gijui)u1.

Therefore, gij = δij for all j ≥ 3, where δij is the Kronecker delta. Hence, g has the

form: 1 a1

λa3 1...

. . .

ar 1

.

This will give rise to an action on A since x · (uiuj) = 0 for all choices of i, j 6= 2

and we have computed explicitly the action on products when either i = 2 or j = 2.

Notice that the vectors, 100...0

,

a1

λ−1

1a3

λ−1...ar

λ−1

,

001...0

, . . . ,

000...1

are linearly independent eigenvectors of g corresponding to the eigenvalues 1, λ, 1, . . . , 1

respectively. One can check,1 a1

λa3 1...

. . .

ar 1

a1

λ−1

1a3

λ−1...ar

λ−1

=

a1 + a1

λ−1

λa3 + a3

λ−1...

ar + ar

λ−1

= λ

a1

λ−1

1a3

λ−1...ar

λ−1

,and the others are easy. Thus if we consider the algebra A = k[u1,

a1

λ−1u1 + u2 +

a3

λ−1u3 + · · ·+ ar

λ−1um, u3, . . . , um] ∼= k[u1, . . . , ur], we see that the representation for g

can be written as

g 7→

1

λ1

. . .

1

,

Page 83: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

78

the desired diagonalized matrix. �

Proof of Theorem 3.2.3(c). As was the case in the previous subsection, checking

the actions of x on the new generators for A gives the same representation,

x =

0 10 0

0

.

First, we calculate Ag. It is easy to see that if i 6= 2, then g · ui = ui = ε(g)ui.

However, we can show by induction that g · (uj2) = (g · u2)j = (λu2)

j and hence

Ag = k[u1, un2 , u3, . . . , ur].

In order to compute AHn(λ) we simply must find which generators of Ag are fixed

by x. That is, for which a ∈ Ag do we have x · a = ε(x)a = 0. Clearly, for any i 6= 2,

we have x · ui = 0 so that all such ui ∈ AHn(λ). Analogously to Lemma 3.2.2 we can

also show that

x · (u`2) = (λ`−1 + λ`−2 + · · ·+ λ+ 1)u1u`−12 .

Thus, x · (u`2) 6= 0 for any ` < n, but x · (un2 ) = 0. Therefore,

AHn(λ) = Ag = k[u1, un2 , u3, . . . , ur].

Hence, AHn(λ) ∼=(k[u1, u2]

Hn(λ))

[u3, . . . , ur]. �

Page 84: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Chapter 4: Reflection groups and Hopf algebras

Our major goal is to describe a generalization of group actions on rings. Through-

out the rest of this thesis, our principal motivation will be a generalization of the

Shephard-Todd-Chevalley Theorem, a classical result describing the invariant subal-

gebras of polynomial rings under group actions.

4.1 Reflection groups, integrals, and trace func-

tions

Let V be a finite dimensional complex vector space. A transformation T : V → V

of a finite subgroup of GL(V ) is called a pseudoreflection if it fixes a codimension

one subspace of V (and is not the identity transformation). In practice, this means

that T has all but one of its eigenvalues equal to one, and the last equal to some

other λ ∈ C. Let G be a finite group of transformations on V . If additionally, G is

generated by pseudoreflections, then G is called a complex reflection group.

Theorem 4.1.1 (Shephard-Todd-Chevalley Theorem). Let G be a group of automor-

phisms acting on a polynomial algebra A = k[x1, . . . , xn]. Then the ring of invariants

AG is isomorphic to a polynomial algebra if and only if G is a complex reflection

group. �

One method of generalization first replaces the polynomial ring with a non-

commutative analog. A finite group of automorphisms acts on an Artin-Schelter

regular algebra A and one considers when the fixed algebra AG is Artin-Schelter

regular. This is a generalization to non-commutative algebras since the only commu-

tative Artin-Schelter regular algebras are polynomials. En route to a full generaliza-

79

Page 85: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

80

tion to Hopf algebra actions, Kirkman, Kuzmanovich, and Zhang investigate when

AH is Artin-Schelter Gorenstein (a slightly weaker condition than Artin-Schelter

regular) provided that A is Artin-Schelter regular. For precise definitions of Artin-

Schelter regular and Artin-Schelter Gorenstein algebras as well as their formulation in

this context see [1] and [10]. Henceforth we will simply say “regular” when we mean

“Artin-Schelter regular.” Suppose that A is regular and H is a finite-dimensional

semisimple Hopf algebra acting on A, a complete generalization of Shephard-Todd-

Chevalley to Hopf algebra actions must answer:

Question 4.1.2. What are the properties of the action of H on A such that the

invariant algebra AH is regular? By analogy to Theorem 4.1.1, such a Hopf algebra

H will be called a quantum reflection group.

Before continuing, we give a few definitions and results for H-module algebras

which are analogous to useful ideas in the case of group actions. From this point

forward, we will be mostly interested with semi-simple Hopf algebras (that is, their

underlying algebra structure has no nilpotent ideals). Many of the following results

should be viewed as a justification for this specialization.

Definition 4.1.3. Let H be a Hopf algebra. A left integral in H is an element

t ∈ H such that ht = ε(h)t, for all h ∈ H. Right integrals are defined analogously.∫ lH

denotes the space of left integrals and∫ rH

denotes the space of right integrals. If∫ lH

=∫ rH

, then H is called unimodular and we simply write∫H

for the space of

integrals.

Example 4.1.4. When H = kG, then t =∑

g∈G g is a left and right integral for

H. This is not difficult to check using that ε(g) = 1 for all g ∈ G and that gh = gk

implies h = k for all g, h, k ∈ G since the elements of G are invertible. �

Theorem 4.1.5. Let H be a finite dimensional Hopf algebra. Then the following are

equivalent.

Page 86: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

81

(i) H is semisimple.

(ii) H is unimodular.

(iii) ε(∫ lH

) 6= 0.

(iv) ε(∫ rH

) 6= 0.

Proof. It turns out that (i) and (ii) are equivalent even in the infinite dimensional

case. For a full proof, see Montgomery [18]. �

Recall that when a finite group G acts on an algebra A we can define a trace

function tr : A→ AG by a 7→∑

g∈G g·a. We have already commented that t =∑

g∈G g

is a left (right integral) in H = kG. This suggests the following generalization of trace

to Hopf algebra actions.

Lemma 4.1.6. Let H be a finite dimensional Hopf algebra, A an H-module algebra,

and 0 6= t ∈∫ lH

. Then the map t : A → A defined by a 7→ t · a is an AH-bimodule

map with values in AH (Note: here “M is an R-bimodule” means M is both a left

and right R-module.)

Proof. Checking that t is a bimodule map is straightforward. We verify that the

values are indeed in AH . Let a ∈ A, h ∈ H. Then

h · (t(a)) = h · (t · a) = (ht) · a = (ε(h)t) · a = ε(h)(t · a) = ε(h)(t(a)),

showing t(a) ∈ AH . �

Definition 4.1.7. A map t : A → AH as above is called a left trace function for

H on A.

Remark 4.1.8. Now suppose that H is semisimple. Then by Theorem 4.1.5 we may

choose t ∈∫H

such that ε(t) = 1. Then for any a ∈ AH we have that t(a) = t · a =

ε(t)a = a. Then, it follows that the image t(A) = AH . That is, in the case of finite

dimensional semisimple Hopf algebras, the trace function is always surjective. �

Page 87: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

82

Now we make more precise some of the properties that are important about the

algebras we consider. Let A be a locally finite N-graded algebra, that is

A = A0 ⊕ A1 ⊕ A2 ⊕ · · ·

where AiAj ⊆ Ai+j and 1 ∈ A0 with dim(Ai) < ∞ for all i, j ∈ N. Additionally,

when A0 = k the algebra is said to be connected. Notice that polynomial algebras

and the quantum plane are both connected locally finite N-graded algebras.

Definition 4.1.9. The Hilbert series of A is

HA(t) =∞∑n=0

dim(An)tn

As a formal power series, the Hilbert series measures how quickly the algebra is

growing as the degree increases. Stated another way, HA(t) tells how many indepen-

dent generators there are in each degree. The computation of this series can uniquely

determine the algebra in question. In particular, consider the following example.

Example 4.1.10. Let A = k[x, y] with deg(x) = deg(y) = 1, that is, a polynomial

algebra in 2 variables. Notice that in degree zero there is one generator since A0 = k,

in degree one we have A1 = 〈x, y〉, in degree two there are three generators A2 =

〈x2, y2, xy〉, and so on. Thus we have

HA(t) = 1 + 2t+ 3t2 + 4t3 + · · · = 1

(1− t)2.

In fact, A is the only commutative algebra which is 2-generated in degree one with

this Hilbert series. Now suppose instead that deg(y) = 2. Then it is not difficult to

show that

HA(t) = 1 + t+ 2t2 + 2t3 + 3t4 + 3t5 + · · · = 1

(1− t)(1− t2).

We comment that there are two terms in the denominators above, each corresponding

to one of the generators x and y. The quantity (1−t2) in the second series corresponds

to the fact that deg(y) = 2. �

Page 88: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

83

We end this section with another generating function that will be useful in the

calculation of our examples. There is another notion of trace in the realm of formal

power series. Suppose that H acts on a graded algebra A, then for h ∈ H one can

define the trace of h as the series

Tr(h, t) =∑n

tr(h|An)tn,

where tr(h|An) denotes the trace of the linear transformation induced by h by restrict-

ing to the degree n component subspace An. When H has finite dimension and is

semisimple, we know that for τ ∈∫H

such that ε(τ) = 1, the associated trace function

τ : A→ AH is surjective. Using this fact, Kirkman, Kuzmanovich, and Zhang extend

a well-known theorem of Molien (regarding invariant algebras under group actions)

to the case of Hopf algebras.

Theorem 4.1.11 (Molien’s Theorem). Let H be a finite-dimensional semisimple

Hopf algebra acting on the connected locally finite N-graded algebra A such that A is

an H-module algebra. Then HAH (t) = Tr(τ, t) where τ ∈∫H

with ε(τ) = 1.

Proof. See [10], and for more context see [8]. �

4.2 Semisimple Hopf algebras of dimension 8

Larson and Radford [13] proved that a semisimple Hopf algebra H of odd dimension

≤ 19 over any field k is both commutative and cocommutative. Therefore, as we

have mentioned earlier, if k is algebraically closed, H is isomorphic to kG ∼= (kG)∗,

for an abelian group G. Further, we have demonstrated (Proposition 2.4.3) that

the same is true for H with dimension = 2, since H ∼= kC2∼= (kC2)

∗, and the same

conclusion holds for dimension 4. Masuoka [15] has shown that there are no nontrivial

semisimple Hopf algebras of dimension = 6, but there is one such isomorphism class

of Hopf algebras having dimension 8. The unique (and hence self-dual) semisimple

Page 89: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

84

Hopf algebra of dimension 8, henceforth denoted H8, that is neither commutative nor

cocommutative, is generated as an algebra by x, y, z with relations,

x2 = y2 = 1, z2 =1

2(1 + x+ y − xy), yx = xy, zx = yz, zy = xz.

The coalgebra structure and antipodes are,

∆(x) = x⊗ x, ∆(y) = y ⊗ y, ε(x) = ε(y) = 1,

∆(z) =1

2(1⊗ 1 + 1⊗ x+ y ⊗ 1− y ⊗ x)(z ⊗ z), ε(z) = 1,

S(x) = x, S(y) = y, S(z) = z.

4.2.1 Actions on k[u1, . . . , un]

We wish to find examples for actions of H8 on algebras A, and we will begin by

considering the case when A is isomorphic to a polynomial ring, k[u1, . . . , un], of n

independent variables. In this subsection we will show that there are no such actions.

We will assume that k = C, so our first task is to find a representation of the algebra

in the appropriate matrix ring, Mn(C). Notice that since x and y commute and

are automorphisms of A, we may assume that they are simultaneously diagonalized

since conjugation is an automorphism of A. Furthermore, they must square to the

identity, implying that each has eigenvalues consisting of only ±1. Therefore, up to

automorphism in A, we can represent x as,

x 7→(−In1

In2

)where Ini

is an identity matrix of dimension ni × ni and n1 is not necessarily equal

to n2. To narrow our choices further, we have the following result.

Lemma 4.2.1. The ideals 〈1− x〉, 〈1− y〉, and 〈x− y〉 are Hopf ideals.

Proof. We will prove the result for the latter; the others are similar. Certainly ε(〈x−

Page 90: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

85

y〉) = 0, and moreover,

∆(x− y) = x⊗ x− y ⊗ y

= x⊗ x− y ⊗ y + x⊗ y − x⊗ y

= x⊗ (x− y) + (x− y)⊗ y

∈ H8 ⊗ 〈x− y〉+ 〈x− y〉 ⊗H8,

showing that 〈x−y〉 is a coideal. The antipode requirement holds since S(x−y) = x−y

so S(〈x− y〉) ⊆ 〈x− y〉. �

Remark 4.2.2. It is not difficult to check that since x (respectively y) is an auto-

morphism, if x (resp. y) is represented by the identity matrix, then x (resp. y) acts

as the identity map on monomials of all degrees in A. Thus 1 − x (resp. 1 − y)

acts as the zero map in all degrees and any action of H on A will factor through the

ideal I = 〈1 − x〉 (resp. 〈1 − y〉). Similarly, if x and y act identically on degree one

monomials, then since both are automorphisms, an inductive argument ensures they

will act identically on monomials of all degrees. Thus x− y will act as the zero map

on all of A and any action of H on A will factor through the ideal I = 〈x− y〉. �

Lemma 4.2.3. If ρ : H8 → Mn(C) is an algebra map such that ρ(x) 6= ρ(y), then ρ

is not a representation of H8.

Proof. Suppose that ρ(x) 6= ρ(y). By the results above, we may represent x and y

without loss of generality as

x 7→

−In1

−In2

In3

, y 7→

−In1

In2

In3

,

where Inirepresents an identity matrix of dimension ni×ni and n1, n2, n3 are natural

numbers such that n1 + n2 + n3 = n. We will write,

z 7→

A B CD E FG H J

,

Page 91: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

86

where each entry is a block matrix of corresponding size to those in x and y, for

example A has dimension n1 × n1. Using the relations zx = yz we observe that

C,D,E,G,H = 0. Additionally, from xz = zy we obtain B,F = 0. Then,

z2 7→

A2

0B2

.

On the other hand,

z2 =1

2(1 + x+ y − xy) 7→

−I II

,

yielding the matrix equation 0 = I, a contradiction. �

Theorem 4.2.4. The polynomial ring k[u1, . . . , un] cannot have the structure of a

non-factorable H8-module algebra.

Proof. By Lemma 4.2.1 we cannot have x and y represented by the same matrix,

while Lemma 4.2.3 says we must have x and y represented by the same matrix. �

4.2.2 Actions on kq[u, v]

Though we have shown that H8 cannot be made to act on polynomial algebras, in

this subsection we compute all of its actions on the algebra k−1[u, v]. Throughout, let

A = k−1[u, v] and k = C. The following example is due to Kirkman, Kuzmanovich,

and Zhang [10].

Example 4.2.5. One can check that the following is a representation for H8 into

2× 2 matrices:

x 7→[

0 11 0

], y 7→

[0 −1−1 0

], z 7→

[1 00 −1

].

Further, it is easily verified that the above representation gives A the structure of

an H8-module algebra. Let G = 〈x, y〉 denote the group generated by the group-

like elements and AG be the algebra of invariants under the action of kG. Some

Page 92: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

87

computation yields that AG = k[a, b, c] where

a↔ u3v − uv3, b↔ u2v2, c↔ u2 + v2, (4.1)

and the generators satisfy the relation 4b2−bc2−a2 = 0. Using the generalized version

of Molien’s Theorem and the concept of surjective trace, it can then be demonstrated

that the full invariant algebra is given by

AH8 = k[a, c].

Note that this algebra is abstractly isomorphic to a polynomial ring in two variables

since ac = ca. Thus, we should regard H8 as a “quantum reflection group” in this

case, even though the representation given above does not generate a reflection group

in the classical sense. In fact, even the group G(H8) = 〈x, y〉 is not generated by

classical pseudoreflections. �

The following example is similar, only using a different representation for the Hopf

algebra. Later, we will verify that the representation below and the one used in the

above example are the only representations which make A an H8-module algebra.

Example 4.2.6. It is not difficult to check that the following is a representation of

H8 in 2× 2 matrices.

x 7→[

0 11 0

], y 7→

[0 −1−1 0

], z 7→

[0 i−i 0

].

With the above representation, we claim that A has the structure of an H8-module

algebra. Indeed, since ∆(x) = x⊗ x and ∆(y) = y⊗ y, one can check that x · (vu) =

−x ·(uv), and y ·(vu) = −y ·(uv). To check that the action of z preserves the relation,

Page 93: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

88

recall that ∆(z) = 12(1⊗ 1 + 1⊗ x+ y ⊗ 1− y ⊗ x)(z ⊗ z). Then,

z · (vu) =1

2((z · v)(z · u) + (z · v)(xz · u) + (yz · v)(z · u)− (yz · v)(xz · u))

=1

2((iu)(−iv) + (iu)(−iu) + (−iv)(−iv)− (−iv)(−iu))

=1

2

(uv + u2 − v2 + vu)

)=

1

2

(u2 − v2

).

On the other hand,

z · (uv) =1

2((z · u)(z · v) + (z · u)(xz · v) + (yz · u)(z · v)− (yz · u)(xz · v))

=1

2((−iv)(iu) + (−iv)(iv) + (iu)(iu)− (iu)(iv))

=1

2

(vu+ v2 − u2 + uv

)=

1

2

(v2 − u2

),

showing that, indeed, z · (vu) = −z · (uv). Now, we compute the algebra of invariants,

AH8 . We begin by noting which elements of A are fixed by the group-like elements

x, y ∈ H8. Notice that G = 〈x, y〉 is represented exactly as in Example 4.2.5. Thus

AG = k[a, b, c] where a, b, c are given as in (4.1). Checking the action of z on these

elements gives

z · a = −a, z · b = b− 1

4a2, z · c = c.

Hence, we have that c ∈ AH8 and z · b ∈ AG. Furthermore, we note that k[a, b, c] is

a commutative ring and let B = k[a2, c]. We claim that AH8 = B. First, note that

indeed a2 ∈ AH8 since,

z · a2 =1

2((z · a)(z · a) + (z · a)(xz · a) + (yz · a)(z · a)− (yz · a)(xz · a))

=1

2(a2 + a2 + a2 − a2)

= a2.

Page 94: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

89

Now suppose that α ∈ AH8 and β ∈ A. In this case, we can show that z · (αβ) =

α(z · β). For since α ∈ AH8 ,

z · (αβ) =1

2(α(z · β) + α(xz · β) + α(z · β)− α(xz · β)) = α(z · β).

A similar calculation gives that z·ab = −a(z·b). Moreover, we have thatB = k[a2, c] ⊆

AH8 ⊆ AG = k[a, b, c] = B[a][b]. Since a, b, c satisfy the relation b2 − 14c2b − 1

4a2 = 0

we can write any element d ∈ AG in the form

d = (f1(a2, c) + g1(a

2, c)a) + (f2(a2, c) + g2(a

2, c)a)b,

for some polynomials fi, gj ∈ B. Now suppose that d ∈ AH8 and consider the result

of z · d. Using the results from our previous calculations, we observe that

z · d = z · (f1 + g1a+ f2b+ g2ab)

= f1 + g1(z · a) + f2(z · b) + g2(z · ab)

= f1 − g1a+ f2(z · b)− g2a(z · b)

= f1 − g1a+ (f2 − g2a)(z · b)

= f1 − g1a+ (f2 − g2a)(b− 1

4a2).

Now if d is to be invariant under z, collecting terms with like degree gives the system

of equations

f1 = f1 −1

4f2a

2

g1 = −g1 +1

4g2a

2

g2 = −g2,

from which we immediately obtain g1, g2 and f2 are identically zero. Hence B = AH8 .

Again, notice that as in Example 4.2.5, the invariant algebra is regular. �

Proposition 4.2.7. The only representations which endow A with the structure of a

non-factorable H8-module algebra are given in Examples 4.2.5 and 4.2.6.

Page 95: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

90

Proof. The proof is by brute force calculation. We begin by noting that since x is

group-like (and hence an automorphism) it must be represented by either a diagonal

or skew diagonal matrix, and moreover, x must square to the identity. Then without

loss of generality, in the diagonal case we are free to assume

x 7→(−1 00 1

),

while in the skew case

x 7→(

0 ζ1ζ

0

)for some ζ ∈ C. When x is diagonal, this is because we cannot have x 7→ I or x 7→ −I.

In the latter scenario, we would then have y 7→ I after some calculation. The only

other possibility is

x 7→(

1 00 −1

),

which is equivalent after the automorphism exchanging the generators inA = k−1[u, v].

In the skew case, we may improve our situation further by the automorphism given

by u 7→ u, v 7→ ζv. With the new generators in A, x is represented by

x 7→(

0 11 0

).

Next we compute all possible representations in these two cases.

Case I: x is represented diagonally

Setting up a system of equations using the algebra relations and computing the

solutions in Maple gives the following possible representations ofH8 into 2×2 matrices:

(i) y 7→(−1 00 1

)and z 7→

((±)i 0

0 (±)1

),

(ii) y 7→(

1 00 −1

)and z 7→

(0 α1α

0

).

Notice that in (i) we have x and y represented by the same matrix. Hence by Lemma

4.2.1 even if this gives rise to an action, it will factor through 〈x− y〉. Now, suppose

Page 96: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

91

that (ii) gives A the structure of an H8-module algebra. Some computation yields

that z · (uv) = uv while we also have z · (vu) = uv, a contradiction since vu = −uv.

Thus, there are no actions on A when x is represented diagonally.

Case II: x is represented by a skew diagonal matrix

Assuming x is represented with 1 along the off diagonal, another system of equa-

tions in Maple yields the following solutions:

(i) y 7→(

0 11 0

)and z 7→ ±

(α 1− α

1− α α

)where 2α2 − 2α + 1 = 0,

(ii) y 7→(

0 −1−1 0

)and z 7→

(−γ β−β γ

)such that β2 = γ2 − 1.

In (i), y is represented by the same matrix as x, so any induced action is factorable.

We claim that (ii) gives rise to both actions we have computed as examples in this

subsection. In order to check that (ii) endows A with the structure of an H8-module

algebra, it is sufficient to check the action of z on uv and vu (the rest of the relations

follow from the module requirement and the fact that x and y act as automorphisms).

After some computation, we obtain the relations

z · (uv) =1

2[u2(−γ2 + 2γβ + β2) + v2(γ2 + 2γβ − β2)]

z · (vu) =1

2[u2(γ2 + 2γβ − β2) + v2(−γ2 + 2γβ + β2)].

Now, imposing the requirement z · (vu) = −z · (uv), we obtain γβ = 0. When β = 0

we have that γ2 = 1 giving rise to the action in Example 4.2.5. On the other hand,

when γ = 0, we have β2 = −1 giving rise to the action in Example 4.2.6. Finally, for

completeness we note that we have not considered the negatives of the representations

for z in Examples 4.2.5 and 4.2.6, but it is not difficult to check that these give rise

to the same H8-module structures. �

Page 97: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

92

4.3 Semisimple Hopf algebras of dimension 12

As alluded to in the previous section, the classification of finite dimensional semisimple

Hopf algebras is still a topic of ongoing investigation. In addition to results of the

previous section, Masuoka [16] has proven the following theorem.

Theorem 4.3.1. Let H be a semisimple Hopf algebra over an algebraically closed field

k, with ch(k) = 0. Let dim(H) = 2p with p an odd prime. Then H, as Hopf algebra,

is isomorphic to one of kC2p, kD2p or (kD2p)∗. Where D2n denotes the dihedral group

of order 2n. �

In particular, this implies that there are no nontrivial Hopf algebras of dimensions

10 or 14. Recall we have already considered odd orders ≤ 19, as well as orders 2, 4, 6

and 8 in the previous sections.

For Hopf algebras of dimension less than 20, this leaves only dimension 12, 16,

and 18. Suppose that k is algebraically closed of characteristic ch(k) 6= 2, 3 (in

particular, what follows is true in characteristic zero). In this case Fukuda [6] has

shown that there exist exactly two distinct isomorphism classes of semisimple Hopf

algebras of dimension 12 which are neither commutative nor cocommutative. Per

Fukuda’s notation, we will denote these Hopf algebras as A+ and A−. As algebras,

A+∼= A−, and both are generated by appending a central idempotent element to the

group algebra kS3. That is, the algebras A± are generated by σ, τ, υ with relations

σ3 = 1, τ 2 = υ2 = 1, στ = τσ2, aυ = υa (a ∈ A±).

Page 98: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

93

The coalgebra structures and antipodes of A+ (respectively A−) are given by:

∆(σ) = συ ⊗ σ + σ(1− υ)⊗ σ2, ε(σ) = 1,

∆(τ) = τ ⊗ τ (resp. τυ ⊗ τ + τ(1− υ)⊗ τ(2υ − 1)), ε(τ) = 1,

∆(υ) = υ ⊗ υ + (1− υ)⊗ (1− υ), ε(υ) = 1,

S(σ) = σ(1− υ) + σ2υ,

S(τ) = τ (resp. τ(2υ − 1)), S(υ) = υ.

Given this structure, it is not difficult to prove that

Lemma 4.3.2. The element µ ∈ A± given by

µ =1

6υ(1 + σ + σ2 + τ + τσ + τσ2)

is a left integral. Since A± is semisimple and ε(µ) = 1, this implies the associated

trace function µ is surjective.

Proof. Consider left multiplication by the generators:

σµ = σ(1

6υ(1 + σ + σ2 + τ + τσ + τσ2))

=1

6υ(σ + σ2 + 1 + στ + στσ + στσ2)

=1

6υ(σ + σ2 + 1 + τσ2 + τ + τσ)

= µ = ε(σ)µ.

The calculations for υ and τ are similar. �

We also note that Fukuda has shown that both A± are self dual. Furthermore, if

H is a semisimple Hopf algebra of dimension 12, it must be the case that |G(H)| = 4,

and in particular, G(A+) ∼= C2 × C2 while G(A−) ∼= C4. This fact is not difficult to

Page 99: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

94

check. In fact, for both A± it is easy to see that 2υ − 1 is group-like:

∆(2υ − 1) = 2∆(υ)−∆(1)

= 2υ ⊗ υ + 2(1− υ)⊗ (1− υ)− 1⊗ 1

= 2υ ⊗ υ + 2(1⊗ 1)− 2(1⊗ υ)− 2(υ ⊗ 1) + 2(υ ⊗ υ)− 1⊗ 1

= 4(υ ⊗ υ)− 2(1⊗ υ)− 2(υ ⊗ 1) + 1⊗ 1

= 2υ ⊗ (2υ − 1)− 1⊗ (2υ − 1)

= (2υ − 1)⊗ (2υ − 1).

Then G(A+) = {1, 2υ − 1, τ, τ(2υ − 1)}. Moreover, (2υ − 1)2 = 1 and τ 2 = 1 so

G(A+) ∼= C2 × C2. In A− one can check that the element g = iτ + (1 − i)τυ is

group-like and that g2 = 2υ − 1. Thus, G(A−) = 〈iτ + (1 + i)τυ〉 ∼= C4.

Shortly, we will want to compute representations for A± which allow us to endow

algebras with the structure of an A±-module algebra.

Lemma 4.3.3. Let A =⊕

nAn be an A±-module algebra. If σ is represented by the

identity matrix; that is, σ acts as the identity map on A0, then σ = idA. The same

result holds for υ.

Proof. Suppose that a ∈ A is a monomial with deg(a) = d ≥ 2. Then we may write

a = a1a2 where each ai is a monomial of degree strictly less than d. We induct on d

to obtain,

σ · a1a2 = (συ · a1)(σ · a2) + (σ(1− υ) · a1)(σ2 · a2)

= (υσ · a1)a2 + ((1− υ)σ · a1)a2 (υ is central)

= (υ · a1)a2 + ((1− υ) · a1)a2

= ((υ + 1− υ) · a1)a2 (A is a Hopf module)

= a1a2.

Page 100: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

95

To show the analogous result for υ, we simply note that when υ is represented by the

identity matrix, (1− υ) 7→ 0. Thus

υ · a1a2 = (υ · a1)(υ · a2) + ((1− υ) · a1)((1− υ) · a2) = a1a2 + 0.

Again, the result follows by induction on the degree of a. �

Lemma 4.3.4. I = 〈1− σ〉 and J = 〈1− υ〉 are Hopf ideals in both algebras, A±.

Proof. First we check that I is a coideal. I satisfies the counit requirement, since

ε(1− σ) = 0. Moreover,

∆(1− σ) = 1⊗ 1− σv ⊗ σ − σ(1− v)⊗ σ2

= 1⊗ 1− σv ⊗ σ − σ ⊗ σ2 + σv ⊗ σ2

= 1⊗ 1− σv ⊗ (σ − σ2)− σ ⊗ σ2

= 1⊗ 1− 1⊗ σ + 1⊗ σ − σ ⊗ σ2 − σv ⊗ (1− σ)σ

= 1⊗ (1− σ) + 1⊗ σ − 1⊗ σ2 + 1⊗ σ2 − σ ⊗ σ2 − σv ⊗ (1− σ)σ

= 1⊗ (1− σ) + 1⊗ (1− σ)σ + (1− σ)⊗ σ2 − σv ⊗ (1− σ)σ,

where the first, second and last terms in the final line above are elements of A± ⊗ I

and the third term is an element of I ⊗A±. Therefore, ∆(I) ⊂ I ⊗A±+A±⊗ I. We

now need only to show that S(I) ⊆ I. Consider

S(1− σ) = S(1)− S(σ)

= 1− (σ(1− υ) + σ2υ)

= (1− σ) + (1− σ)συ

= (1− σ)(1− συ) ∈ I.

Page 101: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

96

Therefore, 〈1− σ〉 is a Hopf ideal. For J , we check

∆(1− υ) = 1⊗ 1− (υ ⊗ υ + (1− υ)⊗ (1− υ))

= −υ ⊗ υ + 1⊗ υ + υ ⊗ 1− υ ⊗ υ

= (1− υ)⊗ υ + υ ⊗ (1− υ)

showing ∆(J) ⊆ J⊗A±+A±⊗J . The counit and antipode conditions are clear. �

Thus, if we define an action of A± on A in which either σ or υ act like the identity

matrix on degree one monomials, then the action is factorable. In A+ we can restrict

even further.

Lemma 4.3.5. I = 〈1− τ〉 is a Hopf ideal of A+.

Proof. In A+, τ is group-like and the verification is straightforward. �

4.3.1 Actions of A± on k[x1, . . . , xn]

In this subsection we prove the following result:

Theorem 4.3.6. The polynomial algebra A = k[x1, . . . , xn] cannot be endowed with

the structure of a non-factorable A±-module algebra for any n.

Lemma 4.3.7. Let A = k[x1, . . . , xn] be a polynomial ring and suppose that A± acts

on A. If υ is represented by the zero matrix, then σ is represented by the identity

matrix.

Proof. Begin by choosing generators x1, . . . , xn such that υ and τ are diagonal. This is

possible since υτ = τυ and τ has finite order, hence is diagonalizable. The conjugation

achieving this diagonal form must be an automorphism of A. Consider a generator

xi. Then υ · xi = 0 and therefore,

σ · x2i = (συ · xi)(σ · xi) + (σ(1− υ) · xi)(σ2 · xi) = (σ · xi)(σ2 · xi).

Page 102: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

97

Similarly, σ2 · x2i = (σ2 · xi)(σ · xi) so that σ · x2

i = σ2 · x2i since A is a commutative

ring. Acting on the left and right sides of this equation with σ2 gives that σ ·x2i = x2

i .

Similarly, σ2 · x2i = x2

i . Now, write

σ · xi =∑j

ajxj, and σ2 · xi =∑`

b`x`.

Then we must have

x2i = σ · x2

i = (σ · xi)(σ2 · xi) =

(∑j

ajxj

)(∑`

a`x`

).

We immediately obtain the equation aibi = 1 by considering the coefficient on the

term x2i on the right. In particular, both ai and bi are nonzero. Next we compute

the coefficient of the term xixj for j 6= i and obtain the equation 0 = aibj + ajbi.

Multiplying through by bi gives the relation bj = −ajb2i . Then consider the coefficient

on x2j for j 6= i. We must have 0 = ajbj = aj(−ajb2i ) after substitution, and we see

that a2j = 0 for all j 6= i. Thus aj = 0 for all such j and moreover, this implies bj = 0.

Hence we have that σ · xi = aixi 6= 0 and σ2 · xi = bixi 6= 0. This must be true for all

xi, and therefore both σ and σ2 are represented by diagonal matrices.

Denote the matrix representing σ by S and the matrix representing τ by T . Since

both are diagonal, they must commute and we have the matrix equation ST = TS =

S2T where we have also made use the relation σ2τ = τσ. Since T is invertible, we

must have that S = I. �

Proof of Theorem 4.3.6. Suppose that A is an A±-module algebra. After choosing

generators so that υ and τ are simultaneously diagonalized (as above) Lemmas 4.3.4

and 4.3.7 allow us to write, without loss of generality, the following block matrices

representing τ and υ:

τ 7→[D1 00 D2

], and υ 7→

[I 00 0

],

Page 103: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

98

where D1 and D2 are both diagonal matrices such that each entry on the diagonal

is equal to ±1. Write {x1, . . . , xm} and {xm+1, . . . , xn} for the set of generators such

that υ ·x` = x` for 1 ≤ ` ≤ m and υ ·x` = 0 for m < ` ≤ n. Choose any i ∈ {1, . . . ,m}

and j ∈ {m+ 1, . . . , n} and consider

σ · xixj = (συ · xi)(σ · xj) + (σ(1− υ) · xi)(σ2 · xj) = (σ · xi)(σ · xj).

On the other hand, taking advantage of commutativity we have

σ · xixj = σ · xjxi = (συ · xj)(σ · xi) + (σ(1− υ) · xj)(σ2 · xi) = (σ · xj)(σ2 · xi).

Therefore 0 = (σ · xj)((σ · xi)− (σ2 · xi)). Suppose that σ · xj = 0. Then the matrix

representing σ must have the zero vector as its j-th column. However, this matrix

must cube to I by the relation σ3 = 1, a contradiction. Thus we conclude that

σ · xi = σ2 · xi and acting on both sides with σ2 gives that σ · xi = xi. This must be

true for all i ∈ {1, . . . ,m} and we have the following representation:

σ 7→[I S1

0 S2

], τ 7→

[D1 00 D2

], and υ 7→

[I 00 0

].

where corresponding blocks have identical dimensions. Using the relation συ = υσ,

we see that S1 = 0. Now, since τ ·xi = ±xi for all i, the action of A± on k[x1, . . . , xn]

induces an action on k[xm+1, . . . , xn] with the following representation:

σ 7→ S2, τ 7→ D2, and υ 7→ 0.

Then, by Lemma 4.3.7 we must have that S2 = I. Therefore S = I, giving rise to a

factorable action. �

Combining the results of this section with those in §4.2, we have proven the

following theorem:

Theorem 4.3.8. Let H be a nontrivial semisimple Hopf algebra over an algebraically

closed field of characteristic zero with dimH ≤ 15. Then for any n ≥ 2, there are no

Page 104: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

99

non-factorable actions of H on the polynomial algebra A = k[x1, . . . , xn] which make

A an H-module algebra. �

4.3.2 Actions of A+ on kq[x, y]

Let A = kq[x, y]. In this subsection, we will consider non-factorable actions of A+,

and we begin by considering representations into 2× 2 matrices over C. Let

υ 7→[a bc d

]so that 1− υ 7→

[1− a −b−c 1− d

].

Apply the action of υ to obtain

υ · yx = (2ab− b)x2 + (qad+ 2bc+ q(1− a)(1− d))xy + (2cd− c)y2

υ · xy = (2ab− b)x2 + (ad+ 2qbc+ (1− a)(1− d))xy + (2cd− c)y2.

Requiring that υ · yx = q(υ · xy) and equating coefficients gives the conditions,

x2 : q = 1 or b(2a− 1) = 0y2 : q = 1 or c(2d− 1) = 0xy : q2 = 1 or bc = 0.

(4.2)

We have already considered the case when q = 1 in the previous subsection. Now

suppose q2 6= 1. Then (4.2) and the algebra relation υ2 = υ together provide a

system of equations in a, b, c and d. Computing the solutions in Maple gives only the

following as possible representations for υ:[1 00 1

],

[1 00 0

],

[0 00 1

],

[0 00 0

].

Per our previous results (Lemma 4.3.4), the first representation (in which υ is the

identity matrix) will give rise to a factorable action. Also up to interchanging the

generators of A, the second and third matrices above are the same. Moreover, we

note that τ is an automorphism and must be represented diagonally when q2 6= 1,

reducing without loss of generality to the following two possible representations,

σ 7→[r st u

], τ 7→

[−1 00 1

],

υ 7→[

1 00 0

]or υ 7→ 0

Page 105: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

100

Consider the first representation. From the relations συ = υσ, σ3 = 1, and στ = τσ2

we obtain a system of equations in r, s, t, u. Some computation gives the only solution

as r = u = 1, s = t = 0. That is, σ 7→ I. By Lemma 4.3.4, this representation factors

the action of A+ on A.

When υ is represented by the zero matrix, there is one additional solution. The

corresponding representation for σ is

σ 7→[−1

2− 3

ξ −12

], σ2 7→

[−1

234ξ

−ξ −12

]. (4.3)

Notice that if σ acts on the product yx = qxy we obtain

σ · yx =3

8ξx2 +

(3

4+

1

4q

)xy +

ξ

2y2,

σ · xy = − 3

8ξx2 +

(3

4q +

1

4

)xy − ξ

2y2,

where in particular we must have 38ξ

= −q 38ξ

after equating coefficients of x2. Thus

we must have q = −1 showing that (4.3) cannot give rise to an action on kq[x, y],

when q2 6= 1.

The only remaining case is q = −1. By Proposition 3.1.13 we must have that

the automorphism τ is represented by either a diagonal or skew diagonal matrix. As

before, we begin by considering representations for υ. Using that υ is idempotent and

the equations (4.2), Maple gives five solutions to the resulting system of equations.

They are given below:

υ 7→[

1 00 1

],

[1 00 0

],

[0 00 1

],

[0 00 0

], or

[12

14ξ

ξ12

], (ξ 6= 0).

Again we will not consider the first representation υ 7→ I and the second and third

representations are equivalent up to exchanging the variables. Next, we use Maple to

check for representations of the full algebra when υ is represented by either[1 00 0

], or

[12

14ξ

ξ12

], (ξ 6= 0).

Page 106: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

101

For υ given by either matrix above, Maple gives that the only representations for

the full algebra are those in which σ 7→ I. Recall that we do not wish to consider

such representations (Lemma 4.3.4) since they will give rise to factorable actions.

Therefore, we must have that that υ is represented by the zero matrix.

Lemma 4.3.9. When υ is represented by the zero matrix, υ acts as the identity map

on monomials of even degree and the zero map on monomials of odd degree.

Proof. If a ∈ A is a monomial of degree one, then υ · a = 0. If a has degree two, then

it can be written as the product of two degree one monomials, a = bc. Then

υ · a = υ · bc = (υ · b)(υ · c) + ((1− υ) · b)((1− υ) · c) = 0 + bc = a.

The result follows by induction. �

Suppose that τ is skew diagonal. Some computation gives that only the following

representation for A+ (up to an automorphism scaling the generators of A) gives rise

to a non-factorable action:

σ 7→

[12−√

32√

32−1

2

], σ2 7→

[−1

2

√3

2

−√

32

12

], τ 7→

[0 11 0

], υ 7→ 0 (4.4)

For brevity, we will not include the calculations for this example, but note the theo-

retical result is similar to what follows.

Finally, suppose instead that τ is represented by a diagonal matrix. A Maple

computation yields that the following is the only representation for A+ into 2 × 2

matrices (up to automorphism in A):

σ 7→[−1

2− 3

ξ −12

], σ2 7→

[−1

234ξ

−ξ −12

], τ 7→

[−1 00 1

], υ 7→ 0 (4.5)

where ξ is any nonzero complex number. Note that exchanging ξ with its negation

simply transposes the representation for σ with that of σ2 in (4.5). These two elements

are completely symmetric in terms of their relations in the Hopf algebra, so henceforth

Page 107: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

102

we will assume that ξ is in the right half-plane, {z ∈ C : =(z) ≥ 0}. In particular, if

ξ is real, then ξ is positive. It is not difficult to check that this representation gives

rise to an action on A. To aid in computation later, we have the following lemma.

Lemma 4.3.10. Let a, b be monomials in A with even degree at least 2. When A+

is represented as in (4.5), we have the following relation: σ · ab = (σ · a)(σ · b).

Proof. Making use of the Lemma 4.3.9, we have

σ · ab = (συ · a)(σ · b) + (σ(1− υ) · a)(σ2 · b)

= (σ · a)(σ · b) + 0

= (σ · a)(σ · b),

as desired. �

Example 4.3.11. Let A+ be represented as in (4.5). The action on degree one

monomials is given by:

σ · x = −12x+ ξy, τ · x = −x, υ · x = 0

σ · y = − 34ξx− 1

2y, τ · y = y, υ · y = 0

Notice that both x2 and y2 are fixed under the actions of τ and υ. Hence, we have

(Aυ)τ = k[x2, y2] = T,

a commutative polynomial ring. The actions of x2 and y2 under σ are

σ · x2 = 0 + (σ(1− υ) · x)(σ2x)

=1

4x2 + ξxy − ξ2y2.

and

σ · y2 = 0 + (σ(1− υ) · y)(σ2y)

= − 9

16ξ2x2 +

3

4ξxy +

1

4y2.

Page 108: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

103

We now have enough information to begin calculation of the fixed algebra under this

action. Suppose a generic element of degree two, αx2 + βy2 ∈ T , for some scalars

α, β, is also invariant under the action of σ. Then we obtain a system of equations

from the relation αx2 +βy2 = σ ·(αx2 +βy2). Maple gives the solution, β = −4ξ2α/3.

Thus we write a = 3x2 − 4ξ2y2, as a generator of the invariant algebra. A similar

calculation on degree four elements emits no new solutions; in particular, the only

degree four invariant element is the square of a.

We now seek generators of degree six. If such an element is to be an independent

generator, that is, not a power of a, we can without loss of generality assume that

any y6 term can be subtracted off with an appropriate multiple of a3. Thus we

write b = αx6 + βx4y2 + γx2y4 for some scalars α, β, γ. Again, we solve the system

obtained from b = σ · b. Using Lemma 4.3.10 it is not difficult to compute the action

of σ on b, and using Maple to solve the resulting equations, obtain the generator

b = x6 + 8ξ2x4y2 + 16ξ4x2y4. Then we have,

B = k[a, b] ⊆ AA+

At this point, it is natural to inquire whether or not the two generators a and b are

algebraically independent. Let R = k[a] denote the algebra of all polynomials over k

in a. Suppose that b satisfies a polynomial, f ∈ R[t]. We will show that f must be

the zero polynomial by inducting on the degree of f . We cannot have f be a nonzero

linear polynomial in b, by our construction of the element b, found above.

When deg f ≥ 2, we begin by showing that f can have no constant term in

R. Suppose 0 = f(b) = r0 + r1b + · · · + rnbn with ri ∈ R and rn 6= 0. Write

r0 = α0 + α1a + · · · + αmam for scalars αj ∈ k. Without loss of generality, we

assume that αm 6= 0. Since a = 3x2 − 4ξ2y2, after expanding r0 in terms of x and

y, there is a term (obtained from αmam) containing y2m, but no other powers of x

or y. This term cannot be canceled by any other term of f(b) since no power of

Page 109: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

104

b = x6 + 8ξ2x4y2 + 16ξ4x2y4 contains terms consisting solely of powers in y. Hence,

αm = 0. Similarly, we must have αj = 0, for all other j. Thus, r0 = 0. Then we have

0 = f(b) = r1b+ · · ·+ rnbn.

We may divide both sides through by b to obtain 0 = r1 + r2b + · · · + rnbn−1, a

polynomial over R, satisfied by b, with degree strictly less than f . By our inductive

hypothesis, r1 = r2 = · · · = rn = 0, and f must also be the zero polynomial. Hence,

there are no relations between the elements a and b.

We conjecture that the algebra B = k[a, b] obtained above is equal to the invariant

subalgebra AA+ . Now deg(a) = 2 and deg(b) = 6, and standard calculation in formal

power series yields that the Hilbert series of B is given by:

HB(t) = 1 + t2 + t4 + 2t6 + 2t8 + 2t10 + 3t12 + · · · = 1

(1− t2)(1− t6). (4.6)

We will use Molien’s Theorem (4.1.11) to calculate the Hilbert series of the fixed

algebra. If they are the same, then B = AA+ since B is the unique commutative

algebra having the Hilbert series in (4.6). Recall (Lemma 4.3.2) that the element

µ = 16υ(1 + σ + σ2 + τ + τσ + τσ2) is a left integral for A+ with ε(µ) = 1. Thus we

have

HAA+ (t) = Tr(µ, t) =∞∑n=0

tr(µ|An)tn.

Now µ will act as the zero map on An for n odd since υ acts the zero map on monomials

of odd degree. Thus, tr(µ|Am) = 0 for m odd and we need only compute tr(µ|An) for

even integers n. Thus, we may restrict to the subalgebra T = k[x2, y2] = (Aυ)τ ⊆ A

for the purposes of calculation. We can simplify our calculation further. As an

Page 110: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

105

example, consider

µ · x2 =1

6(1 + σ + σ2 + τ + στ + σ2τ)υ · x2

=1

6(1 + σ + σ2 + τ + στ + σ2τ) · x2

=1

6((1 + σ + σ2) · x2 + (1 + σ + σ2)τ · x2

=1

6(2(1 + σ + σ2) · x2)

=1

3(1 + σ + σ2) · x2.

Similarly, we may inductively argue that for any d ∈ T that µ · d = 13(1 + σ + σ2) · d.

Therefore, for any positive number n,

tr(µ|T2n) =1

3

(tr(1|T2n) + tr(σ|T2n) + tr(σ2|T2n)

)=

1

3

((n+ 1) + tr(σ|T2n) + tr(σ2|T2n)

).

However, we can do even better.

Lemma 4.3.12. For all n ≥ 0, we have tr(σ|T2n) = tr(σ2|T2n).

Proof. The proof is combinatorial and utilizes the anti-symmetry in the actions of σ

and σ2 on x2 and y2. �

After this simplification we have that tr(µ|T2n) = 13(1 + n+ 2 tr(σ|T2n)). If we are

to reproduce the series in (4.6), we must show that as n takes on the values 0, 1, 2, . . .

that tr(σ|T2n) follows the repeating sequence

1,1

2, 0, 1,

1

2, 0, . . .

so that for example, we want to show tr(µ|T0) = 13(1 + 0 + 2(1)) = 1, tr(µ|T2) =

13(1 + 1 + 2(1

2)) = 1, and tr(µ|T4) = 1

3(1 + 2 + 2(0)) = 1.

Page 111: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

106

Consider T0. It has a single generator 1k and it easy to see that tr(σ|T0) = 1. In

T2 there are two generators x2 and y2. Recall the action of σ on these elements:

σ · x2 =1

4x2 + ξxy − ξ2y2

σ · y2 = − 9

16ξ2x2 +

3

4ξxy +

1

4y2.

Thus, tr(σ|T2) = 14

+ 14

= 12. In T4 we have

σ · x4 = (σ · x2)2 = (1

4x2 + ξxy − ξ2y2)2

σ · x2y2 = (σ · x2)(σ · y2) = (1

4x2 + ξxy − ξ2y2)(− 9

16ξ2x2 +

3

4ξxy +

1

4y2)

σ · y4 = (σ · y2)2 = (− 9

16ξ2x2 +

3

4ξxy +

1

4y2)2,

and some careful multiplication yields tr(σ|T4) = (14)2 + ((1

4)2− 3

4+ 9

16) + (1

4)2 = 0, as

desired. Some Maple worksheets are appended which verify the calculation through

dimension 14. �

Remark 4.3.13. In the example above, it is not difficult to show that the invariant

algebra under the action of the group-like elements G(A+) = 〈τ, 2υ − 1〉 is AG =

(Aυ)τ = k[x2, y2] = T . Now, if B = k[a, b] is indeed the invariant subalgebra AA+ we

make the following observation. The invariant algebra is regular (it is a polynomial

ring) makingA+ a “quantum reflection group” even though σ is clearly not a reflection

of T , a polynomial ring. In fact σ · x2, σ · y2 /∈ T . Neither is σ a reflection of the

original algebra A−1[x, y]. �

Page 112: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Appendix A: Maple Worksheets

The following calculations were done in Maple. One can check that these confirm

the desired sequence of values for tr(σ|T2n). That is, these calculations verify that the

sequence

1,1

2, 0, 1,

1

2, 0, . . .

is correct up through dimension 14, that is n = 7. We have already shown these values

are correct for n = 0, 1, 2 so these calculations begin with n = 3, that is, degree 6.

107

Page 113: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Denote x^2, y^2, and xy by new variables X, Y, and Z. Note that X and Y commute with everything, while Z^2= -XY. We will begin by defining the variables 'sx' and 'sy', the respective actions of sigma on X and Y.

> sx:=1/4*X+a*Z-a^2*Y;sy:=-9/(16*a^2)*X+3/(4*a)*Z+1/4*Y;

:= sx + − X4

a Z a2 Y

:= sy − + + 9 X

16 a2

3 Z4 a

Y4

We begin in degree 6. Consider the actions of sigma on x^6, x^4y^2, x^2y^4, y^6. By symmetry, we need only compute the action on x^4y^2. This is because the coefficient of the x^4y^2 term in s(x^4y^2) will be the same as the x^2y^4 term in s(x^2y^4). The calculations of the x^(2n) term in s(x^2n) are always easy, and are given by (1/4)^n. Similarly, those for y^(2n).> f24:=collect(collect(collect(expand(sx^2*sy^1),X),Y),Z);

:= f24 + + + − − + 3 a Z3

4

⎝⎜⎜

⎠⎟⎟− −

5 a2 Y4

3 X16

Z2⎛

⎝⎜⎜

⎠⎟⎟ + −

7 X a Y8

a3 Y2

415 X2

64 aZ

a4 Y3

411 X a2 Y2

169 X3

256 a2

19 X2 Y64

Now we collect terms which give the coefficient of x^4y^2=X^2Y. Recall, when we use a term above containing Z^2, we must insert a negative sign.> trace6:=2*(1/4)^3+2*(3/16+19/64);

:= trace6 1Next we check degree 8. We need Maple to tell us the actions on x^6y^2 and x^4y^4. The rest of the information can be obtained by symmetry.> f62:=collect(collect(collect(expand(sx^3*sy^1),X),Y),Z);

f623 a2 Z4

42 a3 Z3 Y

⎝⎜⎜

⎠⎟⎟ + −

32

a4 Y2 34

X a2 Y9

32X2 Z2

⎝⎜⎜

⎠⎟⎟− + −

3 a3 X Y2

23 X2 a Y

43 X3

32 aZ

a6 Y4

43 X a4 Y3

415 X2 a2 Y2

32 − + + − + − :=

7 X3 Y64

9 X4

1024 a2 + −

> f44:=collect(collect(collect(expand(sx^2*sy^2),X),Y),Z);

f449 Z4

16⎛

⎝⎜⎜

⎠⎟⎟− −

9 X16 a

3 a Y4

Z3⎛

⎝⎜⎜

⎠⎟⎟− + −

a2 Y2

821 X Y

169 X2

128 a2 Z2⎛

⎝⎜⎜

⎠⎟⎟− − + +

7 X a Y2

1621 X2 Y

64 aa3 Y3

427 X3

256 a3 Za4 Y4

165 X a2 Y3

16 + + + + − :=

45 X3 Y

256 a2

59 X2 Y2

12881 X4

4096 a4 − + +

> trace8:=2*(1/4)^4+2*(7/64-(-9/32))+(59/128-(21/16)+(9/16));

:= trace812

Degree 10:> f82:=collect(collect(collect(expand(sx^4*sy^1),X),Y),Z);

f823 a3 Z5

4⎛

⎝⎜⎜

⎠⎟⎟− +

114

a4 Y316

a2 X Z4 ⎛

⎝⎜⎜

⎠⎟⎟ + −

72

a5 Y2 14

a3 X Y932

X2 a Z3 ⎛

⎝⎜⎜

⎠⎟⎟− − − +

32

a6 Y3 158

X a4 Y2 21128

X3 3932

X2 a2 Y Z2 + + + :=

⎝⎜⎜

⎠⎟⎟ − − + −

9 a5 X Y3

451 a3 X2 Y2

3233 X4

1024 a25 X3 a Y

64a7 Y4

4Z

a8 Y5

413 a6 Y4 X

169 X5

4096 a2

21 X2 a4 Y3

3229 X3 a2 Y2

12837 X4 Y

1024 + + − − + − +

> f64:=collect(collect(collect(expand(sx^3*sy^2),X),Y),Z);

f649 a Z5

16

⎝⎜⎜

⎠⎟⎟− −

21 a2 Y16

27 X64

Z4⎛

⎝⎜⎜

⎠⎟⎟ + −

5 a3 Y2

827 X a Y

1627 X2

128 aZ3

⎝⎜⎜

⎠⎟⎟ − + +

3 a4 Y3

857 X a2 Y2

3245 X3

512 a2

9 X2 Y128

Z2 + + + :=

⎝⎜⎜

⎠⎟⎟ + + − −

3 a3 X Y3

1687 X2 a Y2

128189 X4

4096 a3

93 X3 Y256 a

3 a5 Y4

16Z

a6 Y5

1621 X a4 Y4

6481 X5

16384 a4

69 X2 a2 Y3

128149 X3 Y2

512261 X4 Y

4096 a2 + − + + − + −

> trace10:=2*(1/4)^5+2*(37/1024-(-21/128))+2*(149/512-(9/128)+(-27/64));

:= trace10 0Degree 12:> f102:=collect(collect(collect(expand(sx^5*sy^1),X),Y),Z);

f1023 a4 Z6

4⎛

⎝⎜⎜

⎠⎟⎟− +

72

a5 Y38

a3 X Z5 ⎛

⎝⎜⎜

⎠⎟⎟− + −

58

a4 X Y254

a6 Y2 1564

a2 X2 Z4 ⎛

⎝⎜⎜

⎠⎟⎟ − − −

2516

a3 X2 Y 5 a7 Y3 54

a5 X Y2 1564

X3 a Z3 + + + :=

⎝⎜⎜

⎠⎟⎟ + − + −

54

a8 Y4 5564

X3 a2 Y10532

X2 a4 Y2 154

X a6 Y3 751024

X4 Z2 +

Page 114: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

⎝⎜⎜

⎠⎟⎟− + − − + +

21 X5

2048 a45 a5 X2 Y3

1625 a7 X Y4

865 a3 X3 Y2

6485 X4 a Y

512a9 Y5

2Z

a10 Y6

47 a8 Y5 X

823 X5 Y

204855 X2 a6 Y4

649 X6

16384 a2 + − + + − −

25 X3 a4 Y3

6495 X4 a2 Y2

1024 + −

> f84:=collect(collect(collect(expand(sx^4*sy^2),X),Y),Z);

f849 a2 Z6

16⎛

⎝⎜⎜

⎠⎟⎟− −

158

a3 Y932

X a Z5 ⎛

⎝⎜⎜

⎠⎟⎟ − +

5732

X a2 Y81256

X2 3116

a4 Y2 Z4⎛

⎝⎜⎜

⎠⎟⎟ − − +

45 X2 a Y64

a5 Y3

453 a3 X Y2

169 X3

256 aZ3 + + + :=

⎝⎜⎜

⎠⎟⎟− − + + +

9 a6 Y4

16111 X3 Y

25621 X2 a2 Y2

12833 X a4 Y3

16279 X4

4096 a2 Z2 +

⎝⎜⎜

⎠⎟⎟ − + + − +

135 X5

8192 a3

75 a3 X2 Y3

643 a5 X Y4

32211 X3 a Y2

256411 X4 Y2048 a

a7 Y5

8Z

a8 Y6

1611 a6 Y5 X

32171 X5 Y

8192 a2

159 X2 a4 Y4

25681 X6

65536 a4 + + − − + +

109 X3 a2 Y3

256559 X4 Y2

4096 − +

> f66:=collect(collect(collect(expand(sx^3*sy^3),X),Y),Z);

f6627 Z6

64⎛

⎝⎜⎜

⎠⎟⎟− −

27 a Y32

81 X128 a

Z5⎛

⎝⎜⎜

⎠⎟⎟ + +

243 X Y128

81 X2

1024 a2

9 a2 Y2

64Z4

⎝⎜⎜

⎠⎟⎟− + − +

243 X2 Y256 a

7 a3 Y3

1681 X a Y2

64189 X3

1024 a3 Z3 + + + :=

⎝⎜⎜

⎠⎟⎟− − + − −

3 a4 Y4

64297 X3 Y

1024 a2

783 X2 Y2

51233 X a2 Y3

64243 X4

16384 a4 Z2 +

⎝⎜⎜

⎠⎟⎟− − + − + −

729 X5

32768 a5

87 X2 a Y3

25651 a3 X Y4

128261 X3 Y2

1024 a1377 X4 Y

8192 a3

3 a5 Y5

32Z

a6 Y6

6415 X a4 Y5

1281215 X5 Y

32768 a4

327 X2 a2 Y4

1024 + − + + −

729 X6

262144 a6

385 X3 Y3

10242943 X4 Y2

16384 a2 − + −

> trace12:=2*(1/4)^6+2*(23/2048-(-75/1024))+2*(559/4096-(-111/256)+(-81/256))+(385/1024-(783/512)+(243/128)-(27/64));

:= trace12 1Degree 14:> f122:=collect(collect(collect(expand(sx^6*sy^1),X),Y),Z);

f1223 a5 Z7

4⎛

⎝⎜⎜

⎠⎟⎟ −

916

a4 X174

a6 Y Z6 ⎛

⎝⎜⎜

⎠⎟⎟− − +

158

a5 X Y9

64a3 X2 39

4a7 Y2 Z5 ⎛

⎝⎜⎜

⎠⎟⎟− + + −

454

a8 Y3 1516

a6 X Y2 10564

a4 X2 Y75256

a2 X3 Z4 + + + :=

⎝⎜⎜

⎠⎟⎟ − − + +

254

a9 Y4 16532

a5 X2 Y2 1351024

X4 a154

a7 X Y3 9564

a3 X3 Y Z3 +

⎝⎜⎜

⎠⎟⎟− − − + + −

34

a10 Y5 1174096

X5 345128

X3 a4 Y2 4651024

X4 a2 Y22532

X2 a6 Y3 10516

X a8 Y4 Z2 +

⎝⎜⎜

⎠⎟⎟ + − − − + −

129 X5 a Y2048

33 a9 X Y5

851 X6

16384 a285 a7 X2 Y4

64525 a3 X4 Y2

1024135 a5 X3 Y3

643 a11 Y6

4Z

a12 Y7

415 a10 Y6 X

16141 X5 a2 Y2

4096 + + − −

69 X2 a8 Y5

6455 X6 Y16384

155 X3 a6 Y4

2569 X7

65536 a2

195 X4 a4 Y3

1024 + + − − +

> f104:=collect(collect(collect(expand(sx^5*sy^2),X),Y),Z);

f1049 a3 Z7

16⎛

⎝⎜⎜

⎠⎟⎟− −

964

a2 X3916

a4 Y Z6 ⎛

⎝⎜⎜

⎠⎟⎟ + −

6116

a5 Y2 5132

a3 X Y99256

X2 a Z5 ⎛

⎝⎜⎜

⎠⎟⎟− − + −

3516

a6 Y3 29564

X a4 Y2 375256

X2 a2 Y45

1024X3 Z4 + + + :=

⎝⎜⎜

⎠⎟⎟− + − − +

5 a7 Y4

16315 X4

4096 a175 a3 X2 Y2

12875 X3 a Y

25685 a5 X Y3

16Z3 +

⎝⎜⎜

⎠⎟⎟ + + − − −

11 a8 Y5

16549 X5

16384 a2

665 X3 a2 Y2

5121545 X4 Y

4096105 X2 a4 Y3

128135 a6 Y4 X

64Z2 +

⎝⎜⎜

⎠⎟⎟− − + + + − −

717 X5 Y8192 a

13 a7 X Y5

32351 X6

65536 a3

465 a5 X2 Y4

2562225 X4 a Y2

4096395 a3 X3 Y3

256a9 Y6

16Z

a10 Y7

1623 a8 Y6 X

64901 X5 Y2

16384 + − + +

181 a6 Y5 X2

256423 X6 Y

65536 a2

595 X3 a4 Y4

102481 X7

262144 a4

995 X4 a2 Y3

4096 − − + + −

> f86:=collect(collect(collect(expand(sx^4*sy^3),X),Y),Z);

Page 115: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

f8627 a Z7

64

⎝⎜⎜

⎠⎟⎟− −

135 X256

81 a2 Y64

Z6⎛

⎝⎜⎜

⎠⎟⎟ + −

63 a3 Y2

64297 X a Y

12881 X2

1024 aZ5

⎝⎜⎜

⎠⎟⎟ − − +

19 a4 Y3

64801 X a2 Y2

256567 X2 Y

1024837 X3

4096 a2 Z4 + + + :=

⎝⎜⎜

⎠⎟⎟− + + − +

31 a5 Y4

64513 X4

16384 a3

1107 X2 a Y2

512729 X3 Y1024 a

55 a3 X Y3

64Z3 +

⎝⎜⎜

⎠⎟⎟− − + + − +

3 a6 Y5

641701 X5

65536 a4

855 X3 Y2

20481809 X4 Y

16384 a2

1023 X2 a2 Y3

512231 X a4 Y4

256Z2 +

⎝⎜⎜

⎠⎟⎟ − − + − + +

3321 X5 Y

32768 a3

39 a5 X Y5

1282187 X6

262144 a5

123 a3 X2 Y4

10246741 X4 Y2

16384 a559 X3 a Y3

10245 a7 Y6

64Z

a8 Y7

6431 a6 Y6 X

2565373 X5 Y2

65536 a2 + + − −

357 X2 a4 Y5

10243159 X6 Y

262144 a4

1867 X3 a2 Y4

4096729 X7

1048576 a6

4483 X4 Y3

16384 + + − − +

> trace14:=2*(1/4)^7+2*(55/16384-(-117/4096))+2*(901/16384-(-1545/4096)+(-45/1024))+2*(4483/16384-(855/2048)+(-567/1024)-(-135/256));

:= trace1412

At this point, to continue, we will have to start calculating at least four terms at a time. A more general proof that the pattern continues is required.>

Page 116: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Bibliography

[1] M. Artin and W.F. Schelter, Graded algebras of global dimension 3, Adv. in

Math. 66 (1987), 171-216.

[2] P. Cartier, Groupes algebriques et groupes formels, in “Colloq. Theorie des

Groupes Algebriques (Bruxelles, 1962)”, 87-111, Librairie Universitaire, Louvain;

Gauthier villars, Paris.

[3] C. Chevalley, Invariants of finite groups generated by reflections, Amer. J. Math.

67 (1955), 778-782.

[4] S. Dascalescu, C. Nastasescu and S. Rainu, Hopf Algebras, Marcel Dekker New

York, 2001.

[5] V.G. Drinfeld, Quantum groups, Proc. Int. Cong. Math, Berkeley, 1 (1986), 789-

820.

[6] N. Fukuda, Semisimple Hopf algebras of dimension 12, Tsukuba J. Math. 21

(1997), 43-54.

[7] R.G. Heyneman and M.E. Sweedler, Affine Hopf algebras I, J. Algebra 13 (1969),

192-241.

[8] N. Jing and J.J. Zhang, On the trace of graded automorphisms, J. Algebra 100

(1986), 363-379.

[9] I. Kaplansky, Bialgebras, University of Chicago Lecture Notes, 1975.

[10] E. Kirkman, J.Kuzmanovich and J.J. Zhang, Gorenstein subrings of invariants

under Hopf algebra actions, to appear.

[11] B. Kostant, Graded manifolds, graded Lie theory, and prequantization, Lect.

Notes Math. 570 (1975), 177-306.

[12] R.G. Larson and D.E. Radford, Semisimple cosemisimple Hopf algebras, Amer.

J. Math. 109 (1987), 187-195.

[13] R.G. Larson and D.E. Radford, Semisimple Hopf algebras, J. Algebra 117 (1995),

5-35.

[14] S. Majid, Physics for algebraists: non-commutative and non-cocommutative Hopf

algebras by a bicrossproduct construction, J. Algebra 130 (1990), 17-64.

111

Page 117: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

112

[15] A. Masuoka, Semisimple Hopf algebras of dimesion 6,8, Israel J. Math. 92 (1995),

361-373.

[16] A. Masuoka, Semisimple Hopf algebras of dimension 2p, Comm. Alg. 23 (1995),

1931-1940.

[17] J. Milnor and J.C. Moore, On the structure of Hopf algebras, Ann. of Math. 81

(1965), 211-264.

[18] S. Montgomery, Hopf algebras and their actions on rings, CBMS Reg. Conf.

Series 82, Providence, R.I., 1993.

[19] D.E. Radford, Minimal quasi-triangular Hopf algebras, J. Algebra 157 (1993),

285-315.

[20] G.C. Shephard and J.A. Todd, Finite unitary reflection groups, Canadien J.

Math. 6 (1954), 274-304.

[21] M.E. Sweedler, Hopf Algebras, Benjamin New York, 1969.

[22] E.J. Taft, The order of the antipode of finite-dimensional Hopf algebra, Proc.

Natl. Acad. Sci. USA 68 (1971), 2631-2633.

[23] Y. Zhu, Hopf algebras of prime dimension, Internat. Math. Res. Notices (1994),

53-59.

Page 118: ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON ...jma/allman_ma.pdf · Justin M. Allman ACTIONS OF FINITE DIMENSIONAL NON-COMMUTATIVE, NON-COCOMMUTATIVE HOPF ALGEBRAS ON RINGS

Vita

Justin Allman was born May 17, 1982 to Dr. Richard and Connie Allman of Mor-

gantown, WV. He graduated from Vestavia Hills High School, Birmingham, AL in

May 2001 and completed a Bachelor of Science in May 2005 at the Georgia Institute of

Technology, with a major in Physics and minor in Mathematics. In December of that

year, he was married to Sara Katherine Robb Allman of Memphis, TN. He received

a Master of Arts in Education from Wake Forest University along with secondary

teaching certification in August 2006, after which he taught high school algebra and

trigonometry at West Forsyth High School, in Clemmons, NC, until returning to

Wake Forest in August 2007. He is a proud member of the mathematics faculty at

the North Carolina Governor’s School - West, housed each summer at Salem College

in Winston-Salem. In May 2009, he will be awarded the degree of Master of Arts in

Mathematics by Wake Forest University, and begin work towards a Doctor of Phi-

losophy in Mathematics in August 2009 at the University of North Carolina, Chapel

Hill.

113