• No results found

Higher Spin Algebras and Universal Enveloping Algebras

N/A
N/A
Protected

Academic year: 2020

Share "Higher Spin Algebras and Universal Enveloping Algebras"

Copied!
77
0
0

Loading.... (view fulltext now)

Full text

(1)

Higher Spin Algebras

and Universal Enveloping Algebras

Bolin Han

Supervised by: Peter Bouwknegt

May 2019

(2)
(3)

Declaration

The work in this thesis is my own except where otherwise stated.

(4)
(5)

Acknowledgements

First of all, I must acknowledge my supervisor Peter Bouwknegt for his sup-port and encouragement throughout this year. He is always passionate about our project and patient to answer every single one of my questions. I have been impressed by his talented ideas many times throughout the work. Without his guidance, I would never have survived this tough year.

I would also like to thank all of my lectures and professors who lead me to the beautiful multiverse of mathematics, especially Joan Licata, who initialled my interests in algebra; Jim Borger and Vigleik Angeltveit, who broadened my eyes on the algebra universe; Griff Ware who instructed me on academic writing; Mark Bugden, who infected me with his enthusiasm on this subject and inspired me in the discussions; Amnon Neeman, who not only taught me to be a responsible writer, but also spent his time listening to my troubles with tears and generously gave me warmth.

To all my fellows and friends, you have all contributed to my year in some way or another. Especially, I want to thank for Christopher Hone and Yuzheng Yan for their generous help on some parts of this thesis. And I also need to thank for Kenny Wiratama who always stimulates me when I am pressured or depressed. I would also like to thank for Shiqiu Qiu for his accompany and inspiring discussions with me. I must also thank for Yiming Xu, Marcus Cai, Xilin Lu and Fredrick Yuan who created so many unforgettable memories in my undergraduate life.

Last but not least, I owe a huge amount of gratitude to my parents for loving me and supporting me even remotely.

(6)
(7)

Abstract

Higher spin algebras, arising from the study of the underlying global symme-tries of massless higher-spin particles in physics, have become an interesting area in mathematics since people realised these algebras are deeply related to the the-ory of minimal representations. A well-studied special one-parameter family hs[λ] is shown to be equivalent to a quotient of the universal enveloping algebra (UEA) of sl2. In this thesis, we review the results on hs[λ] with some modifications and then construct new higher spin algebras from the UEA of the semi-direct product sl2 nV2. In addition, we also study the centralisers in the UEA of sl2 nVm for other values of m in preparation to construct more higher spin algebras.

(8)
(9)

Contents

Acknowledgements v

Abstract vii

Introduction 1

1 One-parameter Family: hs[λ] 3

2 Algebraic Structure of U(sl2) 9

3 Construction of hs[λ] from U(sl2) 19

4 An Analogue of hs[λ] Constructed from U(sl2nV2) 27

5 Centralisers in U(sl2nVm) 39

5.1 Centre of U(sl2 nVm) . . . 39 5.2 Centralisers of {F, H, E}in U(sl2nVm) . . . 40 5.3 Centralisers of the elementH inU(sl2nVm) . . . 47

Appendix A Background Knowledge 51

Appendix B Bilinear Form of hs[λ] 55

Appendix C Some Calculation Details 57

C.1 Proof of Equation (3.8) . . . 57 C.2 Identification of even-dimensional sl2-modules in U(sl2nV2)[λ] . . 59 C.3 Another basis of hs[gλ] . . . 60

Appendix D Further Conjectures 63

D.1 Roots of φrs

2(r−1)(λ) . . . 63 D.2 Explicit inverse map of the symmetrisation . . . 63

(10)

D.3 More possible higher spin algebras . . . 64

(11)

Introduction

The study of massless higher spin particles in theoretical physics gives rise to higher spin algebras which describe the underlying global symmetries. The higher spin algebra that is known to be fully consistent in four-dimensional space-time was first considered in [10] from the construction of higher spin cubic interaction. Then the generalization of this algebra to any dimension has been studied based on oscillators.[28]

The unusual feature of interacting higher spin gauge theories is that they require non-flat vacuum solutions. The most symmetric vacuum they admit is known as Anti-de Sitter (AdS) space, frequently used in quantum gravity theories.[24] This brings higher spin algebras to the context of the famous AdS/CFT correspondence which proposes a duality between the AdS space formulated in terms of string theory, or M-theory, and conformal field theories (CFT) which are quantum field theories.[21]

There is a special one-parameter family of higher spin algebras known as hs[λ] which corresponds to a family of three-dimensional interacting equations.[23] It was realised that hs[λ] can also be constructed as a quotient of the universal enveloping algebra of the isometry algebrasl2. And in fact, the coset construction of hs[λ] is deeply related to the minimal representations of sl2 whose kernels are known as Joseph ideals.[17]

Having found the connection between higher spin algebras and the minimal representation problem, mathematicians started to generalise the notion of higher spin algebras to Lie algebras beyond the isometry algebras, while the term “spin” is often translated to the term “weight” in representation theory. The construc-tion of higher spin algebras from the universal enveloping algebra of any semi-simple Lie algebra has been mostly studied.[18] The higher spin superalgebras are also of interests and in fact they play important roles in the area of AdS/CFT.[29] Part of this thesis will be devoted to a fairly less studied family of higher spin alge-bras that comes from the universal enveloping algebra of sl2nV2 which is defined

(12)

to be the semi-direct product between sl2 and the two-dimensional irreducible sl2-module V2.

(13)

Chapter 1

One-parameter Family: hs

[

λ

]

In this chapter, we introduce the one-parameter family higher spin algebras hs[λ] and make some observations on the analytic formula of structure constants. The underlying field of algebras mentioned in this thesis will always be C unless elsewhere stated. To proceed the discussion, let us set up some notations first.

Notation 1.1(Pochhammer symbols). In the following, [a]ndenotes the descend-ing Pochhammer symbol and (a)n denotes the ascending Pochhammer symbol. That is,

[a]n=

Γ(a+ 1)

Γ(a+ 1−n) =a(a−1). . .(a−n+ 1), (a)n = Γ(a+n)

Γ(a) =a(a+ 1). . .(a+n−1).

Notation 1.2 (Generalised hypergeometric function). In the following, the gen-eralised hypergeometric function is denoted as

pFq "

a1, . . . , ap ; z

b1, . . . , bq #

=

X

n=0

(a1)n· · ·(ap)n (b1)n· · ·(bq)n

zn

n!.

Definition 1.3. The one-parameter family of higher spin Lie algebras, denoted by hs[λ], has generators {Vmr :r ≥2,|m|< r}, with commutation relations

[Vmr, Vns] =

r+s−1 X

t=2,t even

gtrs(m, n, λ)Vm+nr+s−t (1.1)

where the structure constants are given by

gtrs(m, n, λ) = 2q t−1 (t−1)!φ

rs t (λ)N

rs

t (m, n) (1.2)

(14)

where qis a scaling constant (which is chosen to be 1

4 from now on), λ∈Cis the parameter, and φrs

t (λ), Ntrs(m, n) are given by

φrst (λ) = 4F3 "

1 2 +λ,

1 2 −λ,

2−t 2 ,

1−t

2 ; 1 3

2 −r, 3 2 −s,

1

2 +r+s−t #

, (1.3)

Ntrs(m, n) = t−1 X

k=0 (−1)k

t−1

k

[r−1 +m]t−1−k[r−1−m]k[s−1−n]t−1−k[s−1 +n]k. (1.4) The commutation relations of hs[λ] were stated in [22]. Here we adapt the notation given in [13].

One can show that hs[λ] is indeed a Lie algebra by directly verifying the skew-symmetry and Jacobi identity on the structure constant grst (m, n, λ). The skew-symmetry is fairly easy to see with noticing φrst (λ) = φtsr(λ) and Ntrs(m, n) = (−1)t−1Ntsr(n, m). The latter equality is obtained by changing the dummy index in the expression of Ntrs(m, n) from k tot−1−k. However, showing the Jacobi identity is more technical. It is subject to relating grst (m, n, λ) to the so-called 6j-symbol that is well-studied in the area of quantum physics and then take advantages of some known properties of the 6j-symbol.[22]

Although not necessary, one can expect realising the abstract Lie bracket of hs[λ] as the usual commutator. In fact there exists an associative product, denoted by ?: hs[λ]×hs[λ]→hs[λ], such that

[Vmr, Vns] =Vmr ? Vns−Vns? Vmr.

The product map ? has been found explicitly in [22] and is given by

Vmr ? Vns = 1 2

r+s−1 X

t=1

gtrs(m, n, λ)Vm+nr+s−t. (1.5)

According to [22], the algebra hs[λ] has an invariant symmetric bilinear form given by

hVmr, Vnsi= 3·4

r−3√πq2r−4Γ(r)(1)r−m−1 (λ21)Γ(r+ 1

2)(2r−2)!

(1−λ)r−1(1 +λ)r−1Γ(r+m)Γ(r−m)δrsδmn

= 3

4q(λ21)g rs

r+s−1(m, n, λ).

(15)

5

Proposition 1.4. hs[λ] is not simple if λ ∈Z\{0,±1}.

Proof. Suppose λ = N ∈ Z\{0,±1}. From Equation (1.6) we note that when

r ≥ |N|+ 1, (1−N)r−1 has a factor (1−N +|N| −1) and (1 +N)r−1 has a factor (1 +N +|N| −1), so we have either (1−N)r−1 = 0 or (1 +N)r−1 = 0. This implies

Span{Vmr :r ≥ |N|+ 1} ⊂radN, where radN denotes the radical of the bilinear form on hs[N].

When r≤N, it is easy to seehVr

m, Vmri 6= 0 for any |m|< r. Hence we have Span{Vmr :r≥ |N|+ 1}= radN.

Now because the bilinear form is invariant, that is,h[Vmr, Vlt], Vnsi=hVmr,[Vlt, Vns]i, we know the radical of the bilinear form is a Lie ideal of hs[N]. And since|N| ≥2, Span{Vmr :r ≥ |N|+ 1} is a proper ideal, then the result follows.

We should mention that an alternative proof can be obtained from a more algebraic point of view after we relate hs[λ] to U(sl2) in Chapter 3. And in fact the converse of Proposition 1.4 is also true (see Proposition 1.10).[9]

Next, let us get a closer look at the complicated expression of the structure constants by computing some commutators explicitly.

Example 1.5. By straightforward calculation we have [V12, V02] =g222(1,0, λ)V12 =V12,

[V02, V21] =g222(0,−1, λ)V21 =V21,

[V12, V21] =g222(1,−1, λ)V02 = 2V02.

which indicates that {V2

−1, V02, V12}forms a basis of the special linear Lie algebra sl2 and hence we know hs[λ] contains sl2 as a subalgebra.

Lemma 1.6. Nrs

t (m, n) = 0 for all t ≥2 min{r, s}.

Proof. Without loss of generality, assume r ≤s and m≥0. Recall

Ntrs(m, n) = t−1 X

k=0 (−1)k

t−1

k

[r−1 +m]t−1−k[r−1−m]k[s−1−n]t−1−k[s−1 +n]k,

(16)

Suppose t ≥ 2r, then for each 0 ≤ k ≤ t−1, we have either k ≥ r−m

or t −1−k ≥ t −r+m ≥ r +m. If k ≥ r, then [r−1−m]k has a factor (r−1−(m+ (r−m)−1)) = 0, so Nrs

t (m, n) = 0. If t−1−k ≥ r+m, then [r −1 + m]t−1−k has a factor (r−1 +m −(r +m) + 1) = 0, so we still have

Nrs

t (m, n) = 0.

Example 1.7. Using Lemma 1.6, we have

[Vm2, Vns] =g2st (m, n, λ)Vm+ns = (m(s−1)−n)Vm+ns

which, together with Example 1.5, indicates that for each fixed s≥ 2, the space Span{Vs

n :|n|< s} is a sl2-module.

It is well-known that the universal enveloping algebra of sl2, herein denoted as U(sl2), contains all irreducible sl2-modules, so this observation also suggests that hs[λ] could be closely related to U(sl2). This turns out to be true and will be discussed in Chapter 3.

Remark 1.8. It is worth to mention that although the generalised hypergeomet-ric function is defined as an infinite series, φrs

t (λ) is always a finite sum because of the presence of 2−2t and 1−2t. When t is a positive integer, one of 2−2t and 1−2t must be a non-positive integer, say −k, and then it is easy to see that only the first k terms of φrs

t (λ) are non-zero.

Example 1.9. Using Lemma 1.6 and Remark 1.8, we have [Vm3, Vns] =g3s2 (m, n, λ)Vm+ns+1 +g3s4 (m, n, λ)Vm+ns−1

= (m(s−1)−2n)Vm+ns+1 + 1 192φ

3s 4 (λ)N

3s

4 (m, n)V s−1 m+n

where

φ3s4 (λ) = 1− (

1 2 −λ)(

1 2 +λ) (32 −s)(s− 1

2) and

N43s(m, n) =24m−24m3+ 24n−72m2n−72mn2−24n3−76ms+ 52m3s

−48ns+ 84m2ns+ 36mn2s+ 72ms2−36m3s2+ 24ns2−24m2ns2

−20ms3+ 8m3s3.

It is easy to solve that φ3s

(17)

7

Proposition 1.10. hs[λ] is simple if λ /∈Z\{0,±1}.

Proof. Suppose λ /∈ Z\{0,±1} and let I be a non-zero ideal of hs[λ]. Consider the set S ={r :Vr

n ∈I for some n}. ThenS =∅ and S has a least element s. We shall do a proof by contradiction. Suppose I is proper, then s≥3. Note thatVs

n ∈I for somenactually impliesVns ∈Ifor all|n|< sas observed in Example 1.7. In particular, for s≥3 we know Vs

−1, V−s2 ∈I. Then we have

I 3[V13, Vs1] =g3s2 (1,−1, λ)V0s+1+g43s(1,−1, λ)V0s−1

= (1 +s)V0s+1− 1

16φ 3s

4 (λ)s(s−2)(s−1)Vs

−1 0 ,

I 3[V23, Vs2] =g3s2 (2,−2, λ)V0s+1+g43s(2,−2, λ)V0s−1

= 2(1 +s)V0s+1+ 1 8φ

3s

4 (λ)s(s−1)(s+ 1)V s−1 0 .

Now since the roots of φ3s4 (λ) = 0 areλ =±(s−1) ands is an integer that is no less than 3, we see φ3s4 (λ)6= 0 unless λ∈Z\{0,±1}. And then we have

I 3[V23, Vs2]−2[V13, Vs1] = 1 8φ

3s

4 (λ)s(s−1)(2s−1)V s−1 0 6= 0

(18)
(19)

Chapter 2

Algebraic Structure of

U

(

sl

2

)

In this chapter, we review some essential definitions and results on the alge-braic structure of the universal enveloping algebraU(sl2). Hereaftergwill always denote a Lie algebra and any additional hypothesis will be stated explicitly.

Definition 2.1 (Universal Enveloping Algebra). Given a Lie algebra g, consider the n-fold tensor product

Tng=g⊗ · · · ⊗g | {z }

ntimes

and denote T0g =

C. Then consider the vector space Tg =

L n=0

Tng and the two-sided idealIgofTggenerated by{x⊗y−y⊗x−[x, y] :x, y ∈g}. We define the universal enveloping algebra to be the quotient algebra

U(g) := Tg

Ig.

Remark 2.2. With respect to the operation of tensor product, U(g) is an as-sociative algebra, and later on we shall omit ⊗ in the discussion. On the other hand, U(g) is also a Lie algebra with the Lie bracket defined to be the usual (matrix) commutator, that is,

[A, B]U(g) =AB−BA, forA, B ∈U(g).

Since U(g) has two multiplicative operations, when we use the word ”ideal” for U(g), we should be careful on the terminology and notation. To be clear, we will use (S)g to denote the Lie ideal ofU(g) generated by S ⊂U(g) and use hSi

to denote the two-sided ideal ofU(g) as an associative algebra.

(20)

Remark 2.3. U(g) can be considered as a g-module via either the left (or right) multiplication or the Lie bracket operation. In this thesis, we are more concerned with the latter case.

Theorem 2.4 (Poincar´e-Birkhoff-Witt). Given a finite-dimensional Lie algebra g, let {x1, x2, . . . , xn}be an ordered basis ofg, then{xi11x

i2

2 · · ·xinn :i1, . . . , in∈N} is a basis of U(g), called Poincar´e-Birkhoff-Witt (PBW) basis.

This is a well-known result with a non-trivial proof that can be found in any standard textbook of Lie algebra and representation theory such as [7].

Let us explore some useful properties and results of the commutators in the universal enveloping algebra, using U(sl2) as an example.

Lemma 2.5. For any A, B ∈U(g), [An, B] =

n X

i=1

Ai−1[A, B]An−i.

This identity could be easily proved by induction. Using this identity we could compute a family of commutators in U(sl2).

Throughout this thesis, we choose an ordered basis ofsl2 to be{F, H, E}with commutation relations

[H, E] = 2E, [H, F] =−2F, [E, F] =H. (2.1) By straightforward calculation we have

[H, Fn] =−2nFn, (2.2)

[E, Fn] =nFn−1H+n(n−1)Fn−1. (2.3) With some effort, we can also get

[F, Hj] = j X

i=1

(−2)i−1

j i

F Hj−i. (2.4)

In order to derive more useful identities forU(sl2), we can bring in the adjoint representation to clean up the notations.

Notation 2.6. Denote the adjoint representation by ad : g→gl(g), that is, (adA) (B) = [A, B] for all A, B ∈g.

And we shall denote (adA)n:= (adA)◦ · · · ◦(adA) | {z }

(21)

11

Bearing in mind that the representation map preserves Lie structure and tak-ing the advantage of the adjoint representation, we can neatly derive a recursive relation that turns out to be very useful later. Note

[E,(adF)r−1−m(Er−1)] = (adE)◦(adF)r−1−m(Er−1)

= (adF)◦(adE)◦(adF)r−1−(m+1)(Er) + (adH)◦(adF)r−1−(m+1)(Er) where

(adH)◦(adF)r−1−(m+1)(Er)

=(r−1−(m+ 1)) (ad[H, F]) (Er) + (adF)r−1−(m+1)◦(adH) (Er−1) =2(m+ 1) (adF)r−1−(m+1)(Er−1).

Hence

[E,(adF)r−1−m(Er−1)] =2 r−1−m

X

k=1

(m+k) (adF)r−1−(m+1)(Er−1)

=(r+m)(r−1−m) (adF)r−1−(m+1)(Er−1).

(2.5)

Lemma 2.7. For all A, B, C ∈U(g), we have the identity (adC)n(AB) =

n X

k=0

n k

(adC)n−k(A)(adC)k(B) (2.6)

The identity is again easy to prove by induction. For further use we shall do an example for U(sl2).

Example 2.8. Applying the identity above we have (adF)n(Er) =(adF)n(Er−1E)

=(adF)n(Er−1)E−n(adF)n−1(Er−1)H−n(n−1)(adF)n−2(Er−1)F

The next thing one would care about U(sl2) is usually the ideals it contains. One of the most important Lie ideals of a universal enveloping algebra is the centre, so we shall describe the centre of U(sl2) explicitly.

Notation 2.9. Denote the centre of U(g) by Z(U(g)), that is,

(22)

Notation 2.10. Denote

C = 4F E+H2+ 2H

which is known as the quadratic Casimir of U(sl2).

Lemma 2.11 (Centre of U(sl2)). The centre of U(sl2) is the polynomial ring generated by the quadratic Casimir C, that is, Z(U(sl2)) = C[C].

In fact this statement can be proven as a special case of Harish-Chandra’s the-orem which describes the centre Z(U(g)) for any semi-simple Lie algebra g.[26] However the general statement and proof of Harish-Chandra’s theorem are in-volved, so we shall approach to an alternative proof following the idea in [4]. Proof. It is easy to check that C ∈ Z(U(sl2)), so C[C] ⊂ Z(U(sl2)). For the converse inclusion, consider the U(sl2)-algebra

e

U(sl2) :=

U(sl2)[F−1]

hF F−11, F−1F 1i

which is also known as the localisation of U(sl2) at F. We shall formally denote

F−n := (F−1)n. And one can check e

U(sl2) is a Lie algebra with extra commuta-tion relacommuta-tions extended from

[H, F−1] = 2F−1, [E, F−1] =−F−2H+ 2F−2.

which agree with Equations (2.2) and (2.3). It is also known that Ue(sl2) has a basis {FiHjEk:i∈Z, j, k ∈N}. And inUe(sl2) we have

E = 1 4F

−1 CH2+ 2H

.

It follows that we can choose a more convenient basis of Ue(sl2), which is

{FiHjCk:i∈Z, j, k ∈N}.

We claim Z(U(sl2)) =U(sl2)∩Z(Ue(sl2)).

It is clear thatU(sl2)∩Z(Ue(sl2))⊂Z(U(sl2)). To show the converse inclusion, take any element z ∈ Z(U(sl2)), then it suffices to check [z, F−1] = 0. To see this, note [z, E] = 0 and

[z, E] =[z,1

4F

−1 CH2+ 2H ] =1

4F

(23)

13

Now since C−H2+ 2H 6= 0, we must have [z, F−1] = 0.

Therefore to showZ(U(sl2)) =C[C], it suffices to determine Z(Ue(sl2)). Con-sider a generic element

A=X

j∈N

Hjaj(F, C)∈Z(Ue(sl2)) where aj(F, C) =

P i∈Z

P

k∈NaijkFiCk and aijk ∈ C are some coefficients. Note only finitely many aijk are non-zero, so there exists a largest value ofj such that

aijk is non-zero for some i, k. Say the value is J. Then since A, C ∈ Z(Ue(sl2)), we have

0 = [F, A] = J X

j=0

[F, Hjaj(F, C)]

= J X

j=0

[F, Hj]aj(F, C)

= J X j=0 j X n=1

(−2)n−1

j n

F Hj−naj(F, C)

= J X n=1 J X j=n

(−2)j−n

j j−n+ 1

F Hn−1aj(F, C)

where we used Equation (2.4). This implies aj(F, C) = 0 whenever j 6= 0. Therefore A can be written as

A=X

i∈Z

Fibi(C)

where bi(C) = P

k∈NbikC

k, with only finitely many of b

ik non-zero. And similar to the above we have

0 = [H, A] =X i∈Z

[H, Fi]bi(C) =X i∈Z

(−2i)Fibi(C)

where we used Equation (2.2). This implies bi(C) = 0 whenever i6= 0. Hence we have

A=b0(C) = X k∈N

ckCk

where only finitely many ck are non-zero. Now the result follows.

(24)

Now we introduce a quotient algebra ofU(sl2) which will be our main interest in the rest of this chapter and also the next chapter.

Notation 2.12. Consider the two-sided ideal ofU(sl2) generated by C−(λ21) as an associative algebra. Denote

U(sl2)[λ] :=

U(sl2)

hC−(λ21)i

where λ∈C. The choice of this parametrisation will be explained in Chapter 3.

Remark 2.13. The idealshC−µiwithµ∈Cturn out to be the maximal ideals of U(sl2).[5]

It is claimed in [6] that there is a well-known result saying

U(sl2) = [U(sl2), U(sl2)]⊕Z(U(sl2)).

However, contrary to one’s expectation, we found this result not easy to prove. For the purpose of this thesis, it suffices to prove the weaker statement stated below.

Lemma 2.14. U(sl2)[λ] = [U(sl2)[λ], U(sl2)[λ]]⊕C

To prove this, we need to introduce the Harish-Chandra character ofU(g).

Definition 2.15 (Harish-Chandra character). The Harish-Chandra character of

U(g) is a linear functionχ:U(g)→C satisfying the following conditions: (1) χ(AB) = χ(BA) for all A, B ∈U(g);

(2) χ(1) = 1;

(3) χ(z1z2) = χ(z1)χ(z2) for all z1, z2 ∈Z(U(g)).

It is proven in [14] that a Harish-Chandra character is uniquely determined by its value on the centre. Precisely speaking, we have

Lemma 2.16. Given a semi-simple Lie algebra g, let χ1, χ2 be Harish-Chandra characters of U(g). If χ1(z) = χ2(z) for all z ∈Z(U(g)), then χ1 =χ2.

Proof of Lemma 2.14 using Lemma 2.16. It follows from Lemma 2.11 that

Z(U(sl2)[λ]) = C.

(25)

15

By Definition 2.15 we observe thatχλ([A, B]) = 0 for allA, B ∈U(sl2)[λ] and

χλ(c) =cfor all c∈C. This implies [U(sl2), U(sl2)]∩C={0}.

It remains to show U(sl2)[λ] = [U(sl2)[λ], U(sl2)[λ]] +C which we shall ap-proach by contradiction.

Denote V := [U(sl2)[λ], U(sl2)[λ]] +Cand supposeV 6=U(sl2)[λ]. Then there exists a subspace S ∈ U(sl2)[λ] such that U(sl2)[λ] = V ⊕S. Choose a basis of

S, say {ei : 1≤i < N}, whereN is either a positive integer or infinity. Define a function χ0λ :U(sl2)[λ]→C by declaring

χ0λ(v) =χλ(v) for all v ∈V , χ0λ(ei) =χλ(ei) + 1 for all 1≤i < N and extending the map linearly. Thenχ0λ satisfies the conditions in Definition 2.15 because χλ does and χ0λ|V = χλ|V. Therefore χ0λ is a Harish-Chandra character of U(sl2)[λ]. But by construction χ0λ 6= χλ, which contradicts to Lemma 2.16. Hence we must have U(sl2)[λ] =V = [U(sl2)[λ], U(sl2)[λ]] +C.

We may devote the last part of this chapter to a simple but yet interesting result that reveals the structure of U(sl2) (and consequently U(sl2)[λ]) via its representations.

Proposition 2.17. Let Vn be ann-dimensional irreducible module ofU(sl2) with the representation homomorphism π:U(sl2)→gln satisfying π(C) = (n2−1)1, where 1 denotes the identity operator in gln. Then

U(sl2)/ker(π)∼=gln.

Bearing the first isomorphism theorem in mind, this follows immediately from the famous classical theorem stated below.

Theorem 2.18 (Burnside). Let V be a finite-dimensional vector space over an algebraically closed field. Let gl(V)be the operator algebra consisting of all linear operators on V. If a non-trivial subalgebra A ⊂ gl(V) is irreducible, then A = gl(V).

Burnside’s theorem of matrix algebras was first proved in 1905. Interestingly, people have made many efforts on simplifying its proof. Here we adapt the proof given in [25], which is claimed to be the simplest proof.

(26)

Proof. We assumed A 6= {0}. Note the algebra U = {kI : k ∈ C}, where I

denotes the identity operator, is not irreducible unless dim(V) = 1. If dim(V) = 1, then the theorem is trivially true, so we may assume dim(V) > 1. Now by irreducibility of A, we have A 6= U. Thus there exists a transformation T ∈ A

such that T /∈ U.

IfT is not invertible, then we are done with S =T. Suppose T is invertible, then T has an eigenvalue λ such that T −λI is not invertible and not the zero-operator. If I ∈ A, then we are done with S = T −λI. Suppose I /∈ A, let

S =T(T −λI) = T2−λT, which belongs to A, and is not invertible (because if it is, then (S−1T) (T −λI) = I, which contradicts with T −λI not invertible), and is not the zero-operator. This proves the lemma.

Lemma 2.20. Any irreducible algebra A contains a rank-one operator.

Proof. We shall prove it by induction on the dimension of V. For the base case dimV = 1, the lemma is trivial. Now suppose the lemma holds when dimV ∈ {1,2,· · ·, n −1}, then for dimV = n > 1, consider the subspace V0 = S(V) and the family A0 = {(SA)|V0 : A ∈ A}, where S is a non-zero non-invertible

operator in A. SinceS is non-invertible, dimV0 < n.

Note A0 is a subalgebra of gl(V0). We claim A0 is irreducible. It’s well-known that transitivity is equivalent to irreducibility, so it suffices to show A0 is transitive on V0. Take 06= v ∈V0, then Av =V because A is irreducible hence transitive. Then A0v =S(Av) = S(V) =V0, so A0 is transitive on V0.

Now by the induction hypothesis, A0 contains a rank-one operator, which means there exists a operator T ∈ A such that (ST)|V0 has rank one. Thus

dimST S(V) = 1, which means ST S ∈ A has rank one.

Lemma 2.21. If a subalgebra A is irreducible, then its adjoint-algebra A∗

=

{A∗ :A ∈ A} is also irreducible.

Proof. We shall prove the contrapositive statement. SupposeA∗ has a non-trivial

A∗-invariant subspace M V, that is, v M,A∈ A, Av V. We claim

the orthogonal complement M⊥ is a non-trivial A-invariant subspace. To see this, take any u∈M⊥and anyA ∈ A, thenhAu, vi=hu, A∗vi= 0 for all v ∈M

since A∗v ∈ M for all v ∈ M. This means Au ∈ M⊥. And M⊥ is non-trivial because M is proper.

(27)

17

denote this as S = v ⊕ w. It can be checked that for any T ∈ A, we have

T(v⊕w)T = (T v)⊕(T∗w), which implies (T v)⊕(T∗w)∈ A.

Note A is transitive, so Av =V, which meansx⊕(T∗w)∈ A for any x∈V. Now noteA∗ is also transitive, soAw=V and hencexy∈ Afor anyx, y V,

that is, A contains all rank-one operators which span gl(V), so A=gl(V). This proves the theorem.

Proof of Proposition 2.17 using Burnside’s theorem. By the first isomorphism the-orem, we have U(sl2)/ker(π) = im(∼ π)⊂ gln. Note that Vn being an irreducible module is equivalent to im(π) being an irreducible subalgebra. And the con-straint that π(C) = (n21)I ensures im(π) is non-trivial. Hence im(π) =gl

n by Burnside’s theorem of matrix algebra.

Besides of taking the advantage of the theorem, here we also provide a more constructive proof of Proposition 2.17 in the case of sl2. And in Chapter 4 we will be able to use the same approach to prove a new result.

Constructive proof of 2.17. Let Vn be the set of polynomials of single variable, say x, with degree less than n, denoted by C[x]n. Choose an ordered basis of C[x]n to be {1, x, . . . , xn−1}. Then we have a realization of U(sl2) as differential operator given by

π(E) = ∂

∂x, π(H) = −2x ∂

∂x + (n−1), π(F) =−x

2 ∂

∂x + (n−1)x.

We shall denote A·v :=π(A)v for A∈U(sl2), v ∈C[x]n.

One can routinely check this is indeed a well-defined U(sl2)-representation. Here we only comment that, by construction we have

F ·xn−1 =−x2∂(x

n−1)

∂x + (n−1)x

n

=−(n−1)xn+ (n−1)xn= 0 which shows that F acting on xn−1 does not cause any trouble. We may check

C=4F E+HH + 2H= (n−1)2+ 2(n−1) =n2−1 as desired. Also note that for k∈ {0,1, . . . , n−1}

E·xk =kxk−1 (2.7)

H·xk= (−2k+n−1)xk (2.8)

(28)

Let Mij denote the basis operator of gln such that

Mijxi =xj, Mijxk = 0, for all k6=i. Now we can construct an element Aij ∈U(sl2) such that π(Aij) = Mij.

For the special case i=j, note that

Aii:= Y

k6=i,k<n 1

2(k−i)(H+ 2k+ 1−n). satisfies

Aiixi =xi, Aiixk = 0, for all k 6=i.

according to Equation (2.8). Now to constructAij forj 6=i, we only need to care about how to send xi toxj. An obvious choice of A

ij fori < j is

Aij :=

(n−1−j)! (n−1−i)!F

j−iA ii

and if i > j we may take

Aij :=

j!

i!E i−jA

ii.

LetBij denote the equivalence class ofAij in the quotient algebraU(sl2)/ker(π). Then {Bij :i, j ∈ {0,1, . . . , n−1}} forms a basis of gln.

Corollary 2.22. With the notation used in Proposition 2.17, for n≥2, consider

Vn as a [U(sl2)[n], U(sl2)[n]]-module with induced representation homomorphism ˜

π: [U(sl2)[n], U(sl2)[n]]→gln. Then

[U(sl2)[n], U(sl2)[n]]/ker(˜π)∼=sln

Proof. The unique Harish-Chandra character χn of U(sl2)[n] allows us to define an invariant symmetric bilinear form on U(sl2)[n], which we shall call the trace, given by

T r(AB) =χn(AB)

(29)

Chapter 3

Construction of hs

[

λ

]

from

U

(

sl

2

)

As mentioned in Example 1.7, it is known that hs[λ] is indeed isomorphic to the derived algebra of the quotient algebra U(sl2)[λ] constructed fromU(sl2). Although this result is stated in many articles on higher spin algebras, to the best of our knowledge, there is no paper that gives a complete proof. The main content of this chapter is to organize a comprehensive proof. For clarity, we will not present all details when they can be found in the literature or can be done by straightforward computation.

Notation 3.1. Let

Vmr = (r+m−1)!

(2r−2)! (adF) r−1−m

(Er−1)∈U(sl2)

for r ≥ 2, |m| < r, and denote the identity element of U(sl2) by V01, that is,

V1

0 := 1 ∈U(sl2).

Proposition 3.2. The set

{CkVmr :r≥1,|m|< r, k ≥0}

forms a basis of U(sl2).

Proof. Recall a PBW basis of U(sl2) is {FiHjEk :i, j, k≥0}. Define the degree of a monomial FiHjEk to be deg(FiHjEk) :=i+j+k, then as a vector space,

U(sl2) can be decomposed into the direct sums of graded pieces via the degree, that is,

U(sl2) =

M

t=0

Grt(sl2)

where Grt(sl2) := Span{FiHjEk : deg(FiHjEk) = t}. (This decomposition can be formalised by using the machinery tool called filtration [7] which takes the Lie

(30)

structure of U(sl2) into account, but for the purpose of this proof it is not harm to restrict ourselves to the decomposition of U(sl2) as a vector space.) We may also denote

U(sl2)n= n M

t=0

Grt(sl2).

And we shall then perform the proof by inducting on n, that is, we are going to prove that for each n, {CkVr

m : k ≥0, r ≥ 1,2k +r−1 = n,|m| < r} is a basis of U(sl2)n.

For n = 0, this holds trivially. For n = 1, we see k could only be 0 and it is easy to compute V2

1 = E, V02 = − 1

2H and V 2

−1 = −F, so it is clear that

{V2

−1, V02, V−21}is a basis of U(sl2)1 ∼=sl2. Forn≥2, suppose{CkVr

m :k ≥0, r ≥1,2k+r−1 =s,|m|< r}is a basis of

U(sl2)sfor alls < n. Then to show{CkVmr :k ≥0, r≥1,2k+r−1 =n,|m|< r} is a basis of U(sl2)n, it suffices to check the vectors {CkVmr : 2k +r−1 = n} are independent, because then the set is automatically a basis of U(sl2)n by counting the dimension. To check the independence we need to look at the explicit expression of Vn+1

m .

By recursively using Example 2.8, we see

Vmn+1 = 1 (2n)!

X

i+j+k=n

(−1)i+j[n+m]2k+j[n−m]2i+jP FiHjEk

(3.1)

where [n+m]2k+j is the Pochhammer symbol andP FiHjEk

denotes the sum of all possible configurations of F, H, E’s such that F appears exactly itimes, H

appears exactly j times and E appears exactlyk times. For example, P(F HE) is F HE+F EH +EF H +EHF +HEF +HF E.

For fixed i, j, k with i +j +k = n, the total number of all such possible configurations is ni n−ji. Now writeVn+1

m in terms of the PBW basis, we have

Vmn+1 = n X

i+j=0

(−1)i+j (2n)!

n i

n−i j

[n+m]2n−2i−j[n−m]j+2iFiHjEn−i−j+U(sl2)n−1 where “+ U(sl2)n−1” means “+ some element in U(sl2)n−1”. Since we are only concerned with the highest degree part of Vmn+1, hereafter we secretly use Vmn+1

to denote Vmn+1 modulo U(sl2)n−1.

Now note that for any fixed m, the term FiHjEn−i−j is non-vanished only when 2i+j =n−m, so we have

Vmn+1 =

b(n−m)/2c

X

i=max{−m,0}

(n+m)!(n−m)!(−1)n−m−i (2n)!

n i

n−i m+i

(31)

21

For each max{−m,0}< i≤ b(n−m)/2c, we see

FiHn−m−2iEm+i =F E(Fi−1Hn−m−2iEm+i−1) +U(sl2)n−1

Fi−1Hn−m−2i+2Em+i−1 =H2(Fi−1Hn−m−2iEm+i−1) +U(sl2)n−1.

Therefore with the cost of adding extra elements inU(sl2)n−1, we can factor out the quadratic Casimir C = 4F E+H2+ 2H, that is,

Vmn+1 = fnm X

k=dm

(−1)n−m−dm 4k−dm cknm

!

FdmHn−m−2dmEm+dm

+C

fnm X

i=dm+1 fnm X

k=i

(−1)n−m−i 4k−i cknm

!

Fi−1Hn−m−2iEm+i−1

(3.2)

wheredm = max{−m,0}, fnm=b(n−m)/2c, cknm =

(n+m)!(n−m)! (2n)!

n k

n−k m+k

. And it can be computed with the help of MathematicaTM that the coefficient of the first term in Equation (3.2) is either (−1)n2−m−n or (−2)n−m, which are non-zero for all |m|< n+ 1.

By the inductive hypothesis, the second term in Equation (3.2) can be written as a linear combination of the vectors in{CkVmr :k, r≥1,2k+r−1 =n,|m|< r}. Then because of the presence of the term in (3.2), we seeVmn+1 is independent on the vectors in {CkVmr :k, r≥1,2k+r−1 =n,|m|< r}.

Also by the inductive hypothesis we see the vectors in{Vr

m :r−1< n,|m|< r} are independent and hence the vectors {CkVr

m :k, r≥1,2k+r−1 =n,|m|< r} are also independent. Now by the PBW Theorem 2.4 we can see the vectors in

{Vn+1

m : |m| < n+ 1} are independent. Therefore we can conclude that all the vectors in {CkVr

m : k ≥ 0, r ≥ 1,2k+r−1 = n,|m| < r} are independent, as desired.

Notation 3.3. In U(sl2)[λ], Vmr denotes the corresponding equivalence class.

Corollary 3.4.

U(sl2)[λ] = Span{Vmr :r≥2,|m|< r} ⊕C Proof. This is an immediate consequence of Proposition 3.2.

Corollary 3.5.

(32)

Proof. This follows from Corollary 3.4 and Lemma 2.14.

Proposition 3.6. The set of generators {Vr

m : r ≥ 2,|m| < r} ⊂ U(sl2)[λ] satisfies the commutation relation (1.1) given in the definition ofhs[λ], which we shall recall here.

[Vmr, Vns] =

r+s−1 X

t=2,t even

gtrs(m, n, λ)Vm+nr+s−t

The key point to prove proposition 3.6 is to figure out the composition law of

U(sl2)[λ] under the basis {Vmr :r≥1,|m|< r}.

From Proposition 2.17 we see whenλ is an integer,U(sl2)[λ] is closely related to the matrix algebra, so it is not surprising that we could get some power from the theory of matrix algebras. Indeed, under a new basis introduced by Racah, a composition law of the matrix algebra obtained as an application of the Wigner-Eckart theorem [3] is helpful. It is proved in [11] that one can derive a manifest expression of the composition law between{Vmr}from the Racah composition law for integer N, and one can then obtain the law for any λ ∈ C by the analytic continuation. To state the manifest expression, we first introduce a few notations.

Notation 3.7 (Clebsch-Gordan coefficient). There are many different notations for Clebsch-Gordan coefficients, here we adapt the convention in [3] and set

Cj1j2j

m1m2m :=δ(m1+m2, m)

×

(2j+ 1)(j1+j2−j)!(j1−m1)!(j2−m2)!(j−m)!(j+m)! (j1+j2+j+ 1)!(j +j1 −j2)!(j +j2 −j1)!(j1+m1)!(j2+m2)!

12

×X

t

(−1)j1−m1+t(j

1+m1+t)!(j+j2−m1−t)!

t!(j−m−t)!(j1−m1−t)!(j2−j+m1+t)!

where t goes from |min{j2−j+m1,0}| to min{j −m, j1−m1, j +j2 −m1}.

Notation 3.8 (Triangle coefficient). Let (a, b, c) be a triple obeying the triangle conditions. Define the triangle coefficient

∆(a, b, c) := s

(a+b−c)!(a−b+c)!(−a+b+c)! (a+b+c+ 1)! .

Notation 3.9 (Modified 6j-symbol). Let "

s s0 s00

j j j

#

(33)

-23

symbol defined in [11] which is explicitly given by "

s s0 s00

j j j

#

:=s!s0!s00!∆(s, s0, s00)X t

(−1)t(2j+ 1 +s00+t)!(2j+ 1−s00−1)! (2j−s−s0+t)!(2j+ 1 +s00)!

× 1

t!(t+s00s)!(t+s00s0)!(s+s0s00t)!(st)!(s0t)!

where ∆(s, s0, s00) is the triangle coefficient.

Now according to [11], the composition law ofU(sl2)[λ] under our chosen basis is given by

VmrVns = X s0,m0

(−1)r+s+s0(2s0)!(s1n)!(r1m)!

(2s−2)!(2r−2)!(s0+m0)!

×f(r−1, s−1, s0|λ)C((rm)(1)(sn)m1)s00V

s0+1 −m0

(3.3)

where C((rm)(1)(sn)m1)s00 are the Clebsh-Gordan coefficients and

f(r−1, s−1, s0|λ) = p(2r−1)(2s−1) "

r s s0

λ−1 2

λ−1 2

λ−1 2

#

. (3.4)

It is possible to directly relate Equation (3.3) to Equation (1.5) to prove Proposition 3.6, since the hypergeometric functions have occurred in the study of Clebsh-Gordan coefficients [20] and the relation between hypergeometric func-tions and the 6j-symbol has also been partly revealed in [22]. However, ap-proaching the proof in this way requires a large amount of knowledge in special functions, so we decided to outline a more practical proof here.

Proof Outline of Proposition 3.6. We shall prove it by a double induction, where the primary induction is on r and the secondary induction in on m.

The Equation (3.3) with r = 2, m=−1 reads

V21Vns = X s0,m0

(−1)s+s0(2s0)!(s−1−n)!

(2s−2)!(s0 +m0)! f(1, s−1, s 0|

λ)C1(1(sn)m1)s00V

s0+1

−m0 (3.5)

Note the delta-function δ(1−n, m0) in the Clebsh-Gordan coefficient forcesm0 = 1−n. And the triangle condition that is secretly required inf(1, s−1, s0|λ) implies

s0 could only be s, s−1 or s−2. Then it remains to calculate the coefficients of

Vns+11, Vs

n−1 and V s−1

(34)

omitted. Up to a normalisation factor, the result should be

V21Vns =Vns+11 − (s−1 +n)

2 V

s n−1

+ (s−2 +n)(s−1 +n)(s−1−λ)(s−1 +λ) 4(2s−3)(2s−1) V

s−1 n−1 =g12s(−1, n, λ)Vns+11 +g22s(−1, n, λ)Vns1+g2s3 (−1, n, λ)Vns11

(3.6)

which agrees with Equation (1.5).

Now suppose for all 2≤k≤r, it is true that

Vkk+1Vns= k+s−1

X

t=1

gtks(−k+ 1, n, λ)Vnk+sk+1−t.

Then to show the formula also holds for k=r+ 1, we note thatV−r+1r =V−21V−rr+1 by Equation (3.6), so it amounts to show

2gt(r+1)s(−r, n, λ) = 2gtrs(1−r, n, λ)−(s+n−t+ 1)grst1(1−r, n, λ)

+(s+n−t+ 1)(s+n−t+ 2)(r+s−t−λ+ 1)(r+s−t+λ+ 1) 2(2r+ 2s−2t+ 1)(2r+ 2s−2t+ 3)

×grst2(1−r, n, λ) for all t. Recallgtrs(m, n, λ) = 4t−1(t21)!φ

rs

t (λ)Ntrs(m, n) and note that

Nt(r+1)s(−r, n) =− 1

(2r−t)(s+n−t+ 1)N rs

t (1−r, n)

Ntrs−2(1−r, n) =

1

(2r−t)(2r−t+ 1)(s+n−t+ 1)(s+n−t+ 2)N rs

t (1−r, n), we see it suffices to show

0 =

r− t

2+ 1

2 r−

t

2 r+s−t+ 1

2 r+s−t+ 3 2

φrst (λ)

+2

t

2 − 1

2 r−

t

2+ 1

2 r+s−t+ 1

2 r+s−t+ 3 2

φrst1(λ)

−r

r− 1

2 r+s−t+ 1

2 r+s−t+ 3 2

φ(r+1)st (λ)

+ t 2 − 1 2 t

2 −1

(r+s−t−λ+ 1) (r+s−t+λ+ 1)φrst2(λ).

(3.7)

(35)

hypergeo-25

metric functions.[1]

0 =b4F3

"

a, b+ 1, a3, a4

; 1

b1, b2, b3

#

−a4F3

"

a+ 1, b, a3, a4

; 1

b1, b2, b3

#

+ (a−b)4F3

"

a, b, a3, a4

; 1

b1, b2, b3

#

0 =c4F3

"

a, a2, a3, a4

; 1

c, b2, b3

#

−a4F3

"

a+ 1, a2, a3, a4

; 1

c+ 1, b2, b3

#

+ (a−c)4F3

"

a, a2, a3, a4

; 1

c+ 1, b2, b3

#

0 =d4F3

"

a1, a2, a3, a4

; 1

c+ 1, d, b3

#

−c4F3

"

a1, a2, a3, a4

; 1

c, d+ 1, b3

#

+ (a−c)4F3

"

a1, a2, a3, a4

; 1

c+ 1, d+ 1, b3

#

Hence Vr

−r+1Vns = r+s−1

X

t=1

gtrs(−r+ 1, n, λ)Vnr+sk+1−t is true for all r ≥ 2. Now as observed in Chapter 1 that φrs

t = φsrt and Ntrs(m, n) = (−1)t

−1Nsr

t (n, m) for all

t, we know grs

t (−r+ 1, n, λ) = (−1)t

−1gsr

t (n,−r+ 1, λ) and therefore [Vrr+1, Vns] =

r+s−1 X

t=2,t even

grst (−r+ 1, n, λ)Vr+sr+1+n−t .

Suppose for each fixed r, [Vr

k, Vns] =

r+s−1 X

t=2,t even

gtrs(k, n, λ)Vk+nr+s−t holds for all

−r+ 1≤k < m. Then to show it also holds for k =m, we first note that using the recursive relation [E, Vr

m−1] = (r−m)Vmr read from Equation (2.5) together with the Jacobi identity, we have

(r−m)[Vmr, Vns] = [E,[Vmr1, Vns]]−[Vmr1,[E, Vns]],

so it suffices to show 0 = r−m+s−t−n

r−m N

rs

t (m−1, n)−

s−1−n r−m N

rs

t (m−1, n+ 1)−N rs t (m, n)

(3.8) for all t even, with noting that other factors in grs(m, n, λ) are all independent on m and n. This may be done by rewriting the function Nrs

t (m, n) in terms of hypergeometric functions in the type of3F2and then again using some contiguous relations. Alternatively, we are also able to prove it by fairly straightforward calculation (see Appendix C.1). Therefore the commutation relation

[Vmr, Vns] =

r+s−1 X

t=2,t even

gtrs(m, n, λ)Vm+nr+s−t

holds for all required values of r, m, s and n.

(36)

Proof. This follows from Corollary 3.5 and Proposition 3.6.

With Theorem 3.10 in hands, now as promised in Chapter 1, we are able to prove that hs[λ] is not simple when λ ∈ Z\{0,±1} (Proposition 1.4) using the algebraic property of U(sl2)[λ]. To some extent, this proof suggests that we can actually get benefits from the algebraic aspect of hs[λ].

(37)

Chapter 4

An Analogue of hs

[

λ

]

Constructed

from

U

(

sl

2

n

V

2

)

So far we have seen that U(sl2)[λ] = hs[λ] ⊕ C as an sl2-module can be decomposed into the direct sum of all irreducible sl2-modules of odd dimension (see Remark 1.7, Remark 2.3 and Proposition 3.2), that is,

hs[λ]⊕C∼=

M

n=1

V2n−1

where V2n−1 is the (2n−1)-dimensional irreducible sl2-module. Then a natural question would be: Can we construct an analogue of hs[λ] that can be decomposed into the direct sum of all irreducible sl2-modules, including the ones of even dimension? That is, we attempt to construct an algebra, hereafter denoted as ]

hs[λ], such that

]

hs[λ]⊕C∼=

M

n=1

Vn.

To this goal, we shall first introduce a family of Lie algebras affiliated to sl2.

Notation 4.1 (sl2nVm). LetVm be them-dimensional irreducible sl2-module, viewed as an abelian Lie algebra. Then denote the semi-direct product of sl2 and Vm by sl2 n Vm. Choose a basis of Vm, say {X0, X1, . . . , Xm−1}, then we have an ordered basis {F, H, E, X0, X1, . . . , Xm−1} for sl2 nVm. Explicitly, the commutation relations of sl2nVm are given by

[E, F] =H, [H, E] = 2E, [H, F] =−2F, [H, Xn] = (m−2n−1)Xn, [E, Xn] =n(m−n)Xn−1, [F, Xn] =Xn+1.

Example 4.2. In this chapter, we are particularly interested in sl2nV2, which has an ordered basis {F, H, E, X0, X1} for sl2nV2 and commutation relations

(38)

[E, F] =H, [H, E] = 2E, [H, F] =−2F, [E, X0] = 0, [H, X0] =X0, [F, X0] =X1, [E, X1] =X0, [H, X1] =−X1, [F, X1] = 0.

Let us first investigate the structure of U(sl2 nV2) as we did for U(sl2) in Chapter 2. As promised before, some methods developed in Chapter 2 and 3 can be inherited here.

Example 4.3. A set of commutation relations can be obtained with the help of Lemma 2.5. We shall list some useful ones here for later reference.

[X1, Hj] = j X n=1 j n

Hj−nX1, [X0, Hj] = j X

n=1 (−1)n

j n

Hj−nX0

[X1, Ek] =−kEk−1X0, [X0, Fi] =−iFi−1X1, [F, X0i] =iX0i−1X1, [E, X1j] =jX0Xj

−1 1 , [H, X0i] =iX0i, [H, X1j] =−jX1j.

Lemma 4.4 (Centre ofU(sl2nV2)). The centre of U(sl2nV2) is the polynomial ring generated by Z :=F X2

0 −HX0X1−EX12, that is,

Z(U(sl2nV2)) =C[Z].

Proof. It is easy to checkC[Z]⊂Z(U(sl2nV2)). To show the converse inclusion, similar to the proof of Lemma 2.11, consider the localisation of U(sl2nV2) atX1 defined as

e

U(sl2nV2) :=

U(sl2nV2)[X1−1]

X1X1−1 −1, X

−1

1 X1−1

and denote X1−n:= X1−1n. Note inUe(sl2nV2) we have

E = F X02−HX0X1−Z

X1−2.

With the similar reasoning as in the proof of Lemma 2.11, again it suffices to determine Z(Ue(sl2nV2)).

Expand a generic element z ∈Z(Ue(sl2nV2)) as

z =X

j∈N

Hjaj(F, Z, X0, X1)

where aj(F, Z, X0, X1) = P

i,k,m∈N

P

n∈ZaijkmnF

iZkXm

0 X1n with finitely many non-zero aijkmn∈C. Then

0 = [X1, z] = X

j∈N

j X n=1 j n

(39)

29

implies aj(F, Z, X0, X1) = 0 unless j = 0, soz can be written as

z = X

m∈N

X0mbm(F, Z, X1)

where bm(F, Z, X1) = Pi,kN

P

n∈ZbikmnFiZkX1n with finitely many non-zero

bikmn∈C. Now

0 = [F, z] = X m∈N

mX0m−1X1bm(F, Z, X1)

implies bm(F, Z, X1) = 0 unless m= 0. Thereforez is in the form of

z =X

i∈N

Fici(Z, X1)

where ci(Z, X1) = Pk∈N

P

n∈ZciknZ

kXn

1 with finitely many non-zero cikn ∈C. 0 = [X0, z] =−

X

i∈N

iFi−1X1ci(Z, X1)

then implies ci(Z, X1) = 0 unless i= 0, so

z =X

n∈Z

X1ndn(Z)

where dn(Z) = P

k∈NdnkZ

k. Finally

0 = [E, z] =X n∈Z

nX0X1n−1dn(Z)

implies dn(Z) = 0 unless n = 0. Hence

z =X

k∈N

pkZk

with finitely many non-zero pk ∈C.

One would also like to have a result that is analogous to Proposition 2.17. Unfortunately, the classification of U(sl2nV2)-modules is not as simple as that of U(sl2), so it is almost impossible to rely on a general theorem such as the Burnside’s theorem.[2] This is the place that our constructive proof of 2.17 could be motivating.

(40)

check that Wm is a (sl2 nV2)-module with the representation homomorphism

τ :sl2nV2 →Wm determined by

τ(H) = x0

∂ ∂x0

−x1

∂ ∂x1

, τ(E) = x0

∂ ∂x1

, τ(F) =x1

∂ ∂x0

,

τ(X0) =

∂ ∂x1

, τ(X1) =−

∂ ∂x0

.

And clearly (Wm, τ) can be extended to aU(sl2nV2)-module. Fortunately, even though these modules are not irreducible, it turns out that we indeed have a result parallel to Proposition 2.17.

Proposition 4.5. Consider the U(sl2nV2)-module Wm defined above. Then

U(sl2nV2)/ker(τ)∼=p1,2,...,m

where pd1,...,dn denotes the parabolic subalgebra of gld1+···+dn with blocks of dimen-sion d1, . . . , dn.

Before proving the proposition, we first need to fix an ordering of the monomi-als in Wm. For convenience, we choose the graded reverse lexicographic ordering here.[19]

Definition 4.6 (Graded reverse lexicographic ordering). Given a polynomial ring R[x0, x1, . . . , xn], whereR is a ring. Let t=xe00· · ·xenn andt

0 =xe00

0 · · ·x e0

n n be monomials in R[x0, x1, . . . , xn]. And let deg(t) := Pni=0ei. The graded reverse lexicographic ordering, also known as grevlex, is given by saying that t ≤ t0 if

t = t0 or deg(t) < deg(t0), or deg(t) = deg(t0) and ei > e0i for the largest index i with ei 6=e0i.

Example 4.7. ForW3, the grevlex ordering gives 1≤x1 ≤x0 ≤x21 ≤x0x1 ≤x20, which suggests we may choose {1, x1, x0, x12, x0x1, x20} to be an ordered basis of

W3.

Proof of Proposition 4.5. Choose an ordered basis ofWmto be the set of monomi-als with graded reverse lexicographic ordering. By the first isomorphism theorem,

U(sl2nV2)/ker(τ)∼= im(τ), so it suffices to show im(τ) =pd1,...,dn (with respect to the chosen basis of Wm).

(41)

31

By straightforward calculation we have

C·xi0xj1 = (i+j)(i+j + 2)xi0xj1, H·xi0xj1 = (i−j)xi0xj1.

Let A(i,j)(p,q)∈U(sl2nV2) with p+q ≤i+j be an element such that τ

A(i,j)(p,q)

satisfies

τA(i,j)(p,q)xi0xj1 =xp0xq1,

τA(i,j)(p,q)xk0xl1 =0, for all (k, l)∈ {(k, l) :k 6=iorl 6=j}.

For the special case p=i, q=j, up to a normalisation factor, we can take

A(i,j)(i,j):= Y k+l=i+j k−l6=i−j

(H−k+l) Y t<m t6=i+j

(C−t(t+ 2)).

When p≤i, q≤j, up to a normalisation factor, we can take

A(i,j)(p,q) :=X0j−qX1i−pA(i,j)(i,j).

When p < i, q > j, up to a normalisation factor, we can take

A(p,q)(i,j) :=Fq−jX1i+j−p−qA(i,j)(i,j).

When p > i, q < j, up to a normalisation factor, we can take

A(p,q)(i,j) :=Ep−iX0i+j−p−qA(i,j)(i,j).

Now the corresponding equivalence classes of A(i,j)(p,q) in U(sl2 nV2)/ker(τ) forms a basis of p1,2...,m.

In Chapter 2 and 3, one of the reason we were interested in the idealshC−µi

with µ∈C(where C = 4F E+HH+ 2H is the quadratic Casimir) is that they are the maximal ideals of U(sl2). In fact, it is shown in [2] that they are also maximal in U(sl2nV2). This suggests that we might be able to construct hs[]λ] by considering the quotient algebra

U(sl2nV2)[λ] := U(sl2nV2)

hC−(λ21)i.

(42)

a central element, now the Lie ideal (C)sl

2nV2 considerably enriches the structure

of hC−(λ21)i. It is not completely obvious now what kind of elements are contained in hC−(λ21)i. In the following discussion, we will find a fairly illustrating generating set of hC−(λ21)i to help us understand the structure of U(sl2nV2)[λ].

Lemma 4.8. The set

{CkVmrX0iX1j :i, j, k ≥0, r≥1,|m|< r}

is a basis of U(sl2nV2).

Proof. This follows from Proposition 3.2 and PBW Theorem 2.4.

Proposition 4.9. With the notation above, the set consists of all elements in the following forms:

Ck+1VmrX0iX1j, VmrX0i+2X1j, VmrX0X1j+1, V r mX

j+2 1 , V

r

m[C, X0], Vmr[C, X1], where k ≥0, r ≥1, m <|r|, i, j ≥0, is a generating set of hCi.

Proof. Denote the set given above by SC from now on. We first show all the elements in SC are contained in hCi.

It is clear that Ck+1Vr mX0iX

j

1 ∈ hCi for all required k, r, m, i, j. Also note

Vr

m[C, X0] = [C, VmrX0] and Vmr[C, X1] = [C, VmrX1] are in (C)sl2nV2 ⊂ hCi for all required r, m. Now note that

(C)sl

2nV2 3[[C, X0], X0] =X

2 0,

which implies X0X1, X12 are also in (C)sl2nV2 by commuting withF and therefore

Vr mX

i+2 0 X

j

1, VmrX0X1j+1, VmrX j+2

1 are contained in hCifor all required r, m, i, j. To showhCi ⊂Span(SC), note that by Lemma 4.8 we know

{Ck0Vmr00Xi 0

0X j0

1 Ck+1VmrX0iX j 1 :k, k

0

, i, i0, j, j0 ≥0, r, r0 ≥1,|m|< r,|m0|< r0}.

is a generating set of hCi. We now try to rewrite each generator above in terms of the elements in SC.

Ifi0 =j0 = 0, then by Proposition 3.2 we know

Ck0Vmr00Ck+1VmrX0iX1j =Ck 0+k+1

(43)

33

If i0+j0 ≥2, then since Span{CkVr mX

i+2 0 X

j

1, CkVmrX0X1j+1, CkVmrX j+2 1 }is a Lie ideal of U(sl2nV2) as one can check, we see

Ck0Vmr00Xi 0

0X j0 1 C

k+1Vr mX

i 0X

j 1 =Ck0+k+1Vmr00VmrXi

0+i

0 X j0+j 1 +C

k0

Vmr00[Xi 0

0X j0 1 , C

k+1Vr m]X

i 0X

j 1

∈Span{Ck+1VmrX0iX1j, VmrX0i+2X1j, VmrX0X1j+1, V r mX

j+2

1 } ⊂Span(SC)

(4.1)

Ifi0+j0 = 1, k0 ≥1, then from Equation (4.1) and Lemma 4.8 we can conclude

Ck0Vmr00Xi 0

0X j0

1 C k+1

VmrX0iX1j ∈Span{Ck+1VmrX0iX1j} ⊂Span(SC)

If i0 +j0 = 1 and k0 = 0, we need to look more closer at [Xn, Ck+1Vmr] where

n ∈ {0,1}. Without loss of generality we may assume k = 0, then [Xn, CVmr] =−V

r

m[C, Xn]−[[Vmr, Xn], C] +C[Xn, Vmr] where we used the Jacobi identity.

One might be able to easily observe that [Vmr, Xn] could be written in the form of aVmr−01X0 +bVmr−0+11 X1 for some a, b ∈ C,|m0| < r−1. But here we decide to

present some careful calculation for later reference.

By recursively using the [F, Vmr] = (m+r−1)Vmr1, we have [Vmr, X1] =

1 (m+r)r−1−m

(adF)r−1−m([Vrr1, X1])

= r−1

(m+r)r−1−m

(adF)r−2−(m−1)(Er−2X0)

where (m+r)r−1−m is the Pochhammer symbol. And by Lemma 2.8 we have (adF)r−2−m(Er−2X0)

=(adF)r−2−m(Er−2)X0 + (r−2−m)(adF)r−3−m(Er−2)[F, X0] = (2r−4)!

(m+r−1)! (m+r−1)V r−1

m X0+ (r−2−m)Vm+1r−1X1

,

(4.2)

and therefore we see

[Vmr, X1] =

(m+r−1)

2(2r−3) (m+r−2)V r−1

m−1X0+ (r−1−m)Vmr−1X1

(4.3)

[Vmr, X0] =

1 (m+r)r−1−m

(adF)r−1−m([Vrr1, X0])−

(r−1−m) 2(2r−3) [V

r

m+1, X1] =− (r−1−m)

2(2r−3) (m+r−1)V r−1

m X0+ (r−2−m)Vm+1r−1X1

(44)

And using the formula given in Equation (1.5) one could easily compute [C, X0] =−4V02X0+ 4V12X1−3X0

[C, X1] =−4V−21X0+ 4V02X1−3X1.

(4.5)

Now combining all the results given in Equation (4.4), (4.3), (4.5) and using Proposition 3.2, we see if i+j ≥1, then

Vmr00Xi 0

0X j0 1 C

k+1Vr mX

i 0X

j 1

∈Span{Ck+1VmrX0iX1j, VmrX0i+2X1j, VmrX0X1j+1, VmrX j+2

1 } ⊂Span(SC), and if i=j = 0, then

Vmr00Xi 0

0X j0 1 C

k+1Vr

m ∈Span{C k+1Vr

mX i 0X

j 1, V

r

m[C, X0], Vmr[C, X1]} ⊂Span(SC). From Proposition 4.9 analysed above, it should be recognised that

U(sl2nV2)−

C−(λ2−1)⊂Span{Vmr, VmrX0, VmrX1 :r≥1,|m|< r}. As remarked at the start of this chapter, we know for eachr, Span{Vr

m :|m|< r} is a (2r−1)-dimensional irreducible sl2-module, which we denote byV2r−1. Now we also notice that

Span{Vmr, VmrX0, VmrX1 :r≥1,|m|< r}=

M

r=1

(V2r−1⊕(V2r−1⊗V2)). It is known that V2r−1⊗V2 can be decomposed asV2r−2⊕V2r. This can be shown via some machinery tools of the representation theory of sl2 such as the Young tableau.[12] We shall include a constructive proof here with giving the highest weight vectors for later use.

Lemma 4.10. Let V2r−1 :=Span{Vmr :|m|< r} and V2 :=Span{X0, X1}. Then

V2r−1⊗V2 =V2r−2 ⊕V2r.

Moreover, up to a normalisation factor, the highest weight vector ofV2r isVrr−1X0 and for r ≥2 the highest weight vector of V2r−2 is Vrr−2X0−Vrr−1X1.

Proof. Let aVmrX0 +bVm+1r X1 be an element in V2r−1 ⊗V2, where a, b ∈ C. For this element to be a highest weight vector, we must have

0 = [E, aVmrX0+bVm+1r X1] =a[E, Vmr]X0+b[E, Vm+1r ]X1+bVm+1r [E, X1]

(45)

35

which requiresa(r−1−m) +b = 0 and either b= 0 or m=r−2. Ifb = 0, then

m=r−1, soVr

r−1X0 is the highest weight vector ofV2r. Ifb 6= 0 thenm =r−2 and a+b = 0, soVr

r−2X0−Vrr−1X1 is the highest weight vector of V2r−2.

To show the independence, let us first explore all basis vectors of V2r and

V2r−2. Recall Equation (4.2), we see the basis vectors ofV2r are

(adF)r−1−m(Vrr1X0) =

(2r−2)!

(m+r)! (m+r)V r

mX0 + (r−1−m)Vm+1r X1

(4.6)

where −r ≤m ≤r−1. And using [F, Vr

m] = (m+r−1)Vmr−1, it is not hard to find the basis vectors of V2r−2 which are given by

(adF)r−1−m Vrr2X0 −Vrr−1X1

= [2r−3]r−1−m Vmr−1X0−VmrX1

(4.7) where −r+ 2≤ m ≤r−1 and [2r−3]r−1−m is the Pochhammer symbol. Now by matching the weights, it suffices to check (adF)r−1−m Vr

r−2X0−Vrr−1X1

and (adF)r−m(Vr

r−1X0) are independent for −r+ 2 ≤m≤r−1. Suppose 0 = c (m+r−1)Vmr1X0+ (r−m)VmrX1

+d Vmr1X0−VmrX1

,

then c(m+r−1) +d = c(r−m)−d = 0, which implies c(2r−1) = 0. This requires c = 0 since r is an integer. But then d = c(r −m) = 0. This shows the sum of V2r−2 +V2r is direct. And then the result follows by counting the dimension of both sides.

Now we have found that Span{Vmr, VmrX0, VmrX1 : r ≥ 1,|m| < r} contains each odd-dimensional irreducible sl2-module once and each non-trivial even di-mensional irreducible sl2-module twice. From the explicit analysis of the gener-ating set SC of hC−(λ2−1)i, we know that for any r ≥ 1 neither Vrr−−21X0 nor

Vrr2X0−Vrr−1X1 is contained inhC−(λ2−1)i, so the 2r-dimensionalsl2-modules are not trivial in U(sl2nV2)[λ].

Remember our intention is to construct hs[]λ] that contains each non-trivial irreducible sl2-module once, so it would be good if we can get rid of one copy of

V2r for each r ≥ 1. Fortunately, U(sl2nV2)[λ] indeed accomplishes this goal in an elegant manner.

Notice that the element Vrr1[C, X0]∈ hC−(λ2−1)iexpands to

Vrr21[C, X0] =−4Vrr−2X0+ 4Vrr−1X1−(2r−1)Vrr−−21X0

according to Equation (4.5) and (1.5). Together with Lemma 4.10, we see for each

(46)

highest weight vector Vr

r−2X0−Vrr−1X1 of V2r ⊂V2r+1⊗V2 up to a normalisation factor. By the nature of the sl2-modules, we can then conclude that the two 2r -dimensionalsl2-modules are identified as the same inU(sl2nV2)[λ] (see Appendix C.2 for calculations). Hereafter we shall denote the highest weight vector of this

V2r ⊂ U(sl2 nV2)[λ] by Urr−1 and define Umr :=

(r+m−2)!

(2r−3)! (adF) r−1−m

(Ur r−1) for −r≤m ≤r−1.

It remains to check that inU(sl2nV2)[λ] the sums between even-dimensional sl2-modules are still direct, that is, we want to show

min{m+r, r−1−m}

X

i=0

ciUmr−i = 0 only if ci = 0 for alli.

Suppose Pmini=0{m+r, r−1−m}ciUmr−i = 0 and there exists ci 6= 0. Consider the set

I ={i:ci 6= 0} and letj be the largest element in I, then 0 = (adE)r−j−1−m

min{m+r, r−1−m}

X

i=0

ciUmr−i 

=c U r−j r−j−1

where the coefficient c ∈ C is non-zero. But this implies Urrjj1 = 0, which is impossible.

Although we have not yet found a manifest expression of the structure con-stants for U(sl2 nV2)[λ] analogous to the one given in Equation (1.1) (see Ap-pendix C.3 for some progress), from Equation (4.3) and (4.4) we can at least deduce that [Ur

m, Uns] = 0 for any r, s, m, n and

[Vmr, Uns]∈Span{Umr :r≥1,−r≤m≤r−1}.

This implies

[U(sl2nV2)[λ], U(sl2nV2)[λ]]∩C={0}.

Hence we may let hs[]λ] := [U(sl2 nV2)[λ], U(sl2 nV2)[λ]]. And finally we are ready to state the new result.

Theorem 4.11. The Lie algebra hs][λ] with a basis

{Vmr :r≥2,|m|< r} ∪ {Uns:, s≥1,−s≤n ≤s−1}

where

Vmr := (r+m−1)! (2r−2)! (adF)

(47)

37

and

Uns:= (s+n−2)! (2s−3)! (adF)

s−1−n(Vs

s−2X0−Vss−1X1)

decomposes, as a sl2-module, into the direct sum of irreducible sl2-modules of dimension larger than 1. That is,

] hs[λ] =

M

k=2

Vk

where fork odd, Vk is thek-dimensionalsl2-module generated byV(k(k+1)/21)/2 and for

(48)
(49)

Chapter 5

Centralisers in

U

(

sl

2

n

V

m

)

The success on constructing an interesting algebra fromU(sl2nV2) motivates us to explore more on the structure of U(sl2 nVm) for generic m. We decided to start with investigating the centre of U(sl2 nVm), which is one of the most natural Lie ideals.

5.1

Centre of

U

(

sl

2

n

V

m

)

The centres of U(sl2 nV1) ∼= U(sl2) and U(sl2 n V2) have been found in Lemmas 2.11 and 4.4 respectively. Another studied case is for m = 3. And in fact the proof given in [4] is the prototype of our proofs for m = 1,2, so here we only present the result and briefly sketch the proof.

Lemma 5.1. Let Z1 := 2X0X2−X12, Z2 :=EX2+HX1−2F X0. Then

Z(U(sl2nV3)) =C[Z1, Z2].

Proof outline. Consider a localisation ofU(sl2nV3) atX2, which we hereafter de-note byUe(sl2nV3), where we can make sense ofX2−1so thatX0 = 12(Z1+X12)X2−1 and E = (Z2+ 2F X0−HX1)X2−1. Then a basis of Ue(sl2nV3) is

{FiHjX1kX2lZ1mZ2n:i, j, k, m, n∈N, l ∈Z}.

Now for a generic central element z ∈ Ue(sl2nV3), using [X2, z] = 0 and X2 commutes with every basis element except for those with j ≥ 1, we can deduce

z could only be a linear combination of {FiXk

1X2lZ1mZ2n : i, k, m, n ∈ N, l ∈ Z}. And then [F, z] = 0 implies z ∈ Span{FiXl

2Z1mZ2n : i, m, n ∈ N, l ∈ Z}. Now [X1, z] = 0 gives z ∈ Span{X2lZ1mZ2n : m, n ∈ N, l ∈ Z} and finally [H, z] = 0

References

Related documents

Table 20: Result - Neighborhood Change Score by Household Size. To further examine the relationship between household size and

Abstract—This paper presents a model for developing long- term human-robot interactions and social relationships based on the principle of 'human' cognitive

I think, your point is a good one, but as Sarah says, social research takes a bit of time, and writing it up also takes time, which is why any journal can’t be as quick as

These results suggest that the use of informal job search methods (direct approach and friends/relatives) gener- ates substantially higher hourly wages for the unemployed than

of the study that private consumption in Romania is an AR(1) process (the current value of private consumption is based on the immediately preceding value) was

- Investigation the performance of CRM binders used in Isfahan Province shows that the existing neat binder samples covered only four out of seven required per- formance

In order to discover new bioactive molecules and pesticides acting on tobacco bacterial wilt, we sought to combine the active structure of hydrazone and pyridine together to design

Retirees and doctors who had ‘retired and returned’ were asked to indicate which, if any, of the following factors had influenced their decision to retire when they did: pressure