Cellular algebra

From HandWiki
Short description: Term in abstract algebra

In abstract algebra, a cellular algebra is a finite-dimensional associative algebra A with a distinguished cellular basis which is particularly well-adapted to studying the representation theory of A.

History

The cellular algebras discussed in this article were introduced in a 1996 paper of Graham and Lehrer.[1] However, the terminology had previously been used by Weisfeiler and Lehman in the Soviet Union in the 1960s, to describe what are also known as coherent algebras. [2][3][4]

Definitions

Let [math]\displaystyle{ R }[/math] be a fixed commutative ring with unit. In most applications this is a field, but this is not needed for the definitions. Let also [math]\displaystyle{ A }[/math] be an [math]\displaystyle{ R }[/math]-algebra.

The concrete definition

A cell datum for [math]\displaystyle{ A }[/math] is a tuple [math]\displaystyle{ (\Lambda,i,M,C) }[/math] consisting of

  • A finite partially ordered set [math]\displaystyle{ \Lambda }[/math].
  • A [math]\displaystyle{ R }[/math]-linear anti-automorphism [math]\displaystyle{ i:A\to A }[/math] with [math]\displaystyle{ i^2 = \operatorname{id}_A }[/math].
  • For every [math]\displaystyle{ \lambda\in\Lambda }[/math] a non-empty finite set [math]\displaystyle{ M(\lambda) }[/math] of indices.
  • An injective map
[math]\displaystyle{ C: \dot{\bigcup}_{\lambda\in\Lambda} M(\lambda)\times M(\lambda) \to A }[/math]
The images under this map are notated with an upper index [math]\displaystyle{ \lambda\in\Lambda }[/math] and two lower indices [math]\displaystyle{ \mathfrak{s},\mathfrak{t}\in M(\lambda) }[/math] so that the typical element of the image is written as [math]\displaystyle{ C_\mathfrak{st}^\lambda }[/math].
and satisfying the following conditions:
  1. The image of [math]\displaystyle{ C }[/math] is a [math]\displaystyle{ R }[/math]-basis of [math]\displaystyle{ A }[/math].
  2. [math]\displaystyle{ i(C_\mathfrak{st}^\lambda)=C_\mathfrak{ts}^\lambda }[/math] for all elements of the basis.
  3. For every [math]\displaystyle{ \lambda\in\Lambda }[/math], [math]\displaystyle{ \mathfrak{s},\mathfrak{t}\in M(\lambda) }[/math] and every [math]\displaystyle{ a\in A }[/math] the equation
[math]\displaystyle{ aC_\mathfrak{st}^\lambda \equiv \sum_{\mathfrak{u}\in M(\lambda)} r_a(\mathfrak{u},\mathfrak{s}) C_\mathfrak{ut}^\lambda \mod A(\lt \lambda) }[/math]
with coefficients [math]\displaystyle{ r_a(\mathfrak{u},\mathfrak{s})\in R }[/math] depending only on [math]\displaystyle{ a }[/math], [math]\displaystyle{ \mathfrak{u} }[/math] and [math]\displaystyle{ \mathfrak{s} }[/math] but not on [math]\displaystyle{ \mathfrak{t} }[/math]. Here [math]\displaystyle{ A(\lt \lambda) }[/math] denotes the [math]\displaystyle{ R }[/math]-span of all basis elements with upper index strictly smaller than [math]\displaystyle{ \lambda }[/math].

This definition was originally given by Graham and Lehrer who invented cellular algebras.[1]

The more abstract definition

Let [math]\displaystyle{ i:A\to A }[/math] be an anti-automorphism of [math]\displaystyle{ R }[/math]-algebras with [math]\displaystyle{ i^2 = \operatorname{id} }[/math] (just called "involution" from now on).

A cell ideal of [math]\displaystyle{ A }[/math] w.r.t. [math]\displaystyle{ i }[/math] is a two-sided ideal [math]\displaystyle{ J\subseteq A }[/math] such that the following conditions hold:

  1. [math]\displaystyle{ i(J)=J }[/math].
  2. There is a left ideal [math]\displaystyle{ \Delta\subseteq J }[/math] that is free as a [math]\displaystyle{ R }[/math]-module and an isomorphism
[math]\displaystyle{ \alpha: \Delta\otimes_R i(\Delta) \to J }[/math]
of [math]\displaystyle{ A }[/math]-[math]\displaystyle{ A }[/math]-bimodules such that [math]\displaystyle{ \alpha }[/math] and [math]\displaystyle{ i }[/math] are compatible in the sense that
[math]\displaystyle{ \forall x,y\in\Delta: i(\alpha(x\otimes i(y))) = \alpha(y\otimes i(x)) }[/math]

A cell chain for [math]\displaystyle{ A }[/math] w.r.t. [math]\displaystyle{ i }[/math] is defined as a direct decomposition

[math]\displaystyle{ A=\bigoplus_{k=1}^m U_k }[/math]

into free [math]\displaystyle{ R }[/math]-submodules such that

  1. [math]\displaystyle{ i(U_k)=U_k }[/math]
  2. [math]\displaystyle{ J_k:=\bigoplus_{j=1}^k U_j }[/math] is a two-sided ideal of [math]\displaystyle{ A }[/math]
  3. [math]\displaystyle{ J_k/J_{k-1} }[/math] is a cell ideal of [math]\displaystyle{ A/J_{k-1} }[/math] w.r.t. to the induced involution.

Now [math]\displaystyle{ (A,i) }[/math] is called a cellular algebra if it has a cell chain. One can show that the two definitions are equivalent.[5] Every basis gives rise to cell chains (one for each topological ordering of [math]\displaystyle{ \Lambda }[/math]) and choosing a basis of every left ideal [math]\displaystyle{ \Delta/J_{k-1}\subseteq J_k/J_{k-1} }[/math] one can construct a corresponding cell basis for [math]\displaystyle{ A }[/math].

Examples

Polynomial examples

[math]\displaystyle{ R[x]/(x^n) }[/math] is cellular. A cell datum is given by [math]\displaystyle{ i = \operatorname{id} }[/math] and

  • [math]\displaystyle{ \Lambda := \lbrace 0,\ldots,n-1\rbrace }[/math] with the reverse of the natural ordering.
  • [math]\displaystyle{ M(\lambda) := \lbrace 1\rbrace }[/math]
  • [math]\displaystyle{ C_{11}^\lambda := x^\lambda }[/math]

A cell-chain in the sense of the second, abstract definition is given by

[math]\displaystyle{ 0 \subseteq (x^{n-1}) \subseteq (x^{n-2}) \subseteq \ldots \subseteq (x^1) \subseteq (x^0)=R[x]/(x^n) }[/math]

Matrix examples

[math]\displaystyle{ R^{\,d \times d} }[/math] is cellular. A cell datum is given by [math]\displaystyle{ i(A)=A^T }[/math] and

  • [math]\displaystyle{ \Lambda := \lbrace 1 \rbrace }[/math]
  • [math]\displaystyle{ M(1) := \lbrace 1,\dots,d\rbrace }[/math]
  • For the basis one chooses [math]\displaystyle{ C_{st}^1 := E_{st} }[/math] the standard matrix units, i.e. [math]\displaystyle{ C_{st}^1 }[/math] is the matrix with all entries equal to zero except the (s,t)-th entry which is equal to 1.

A cell-chain (and in fact the only cell chain) is given by

[math]\displaystyle{ 0 \subseteq R^{\!d \times d} }[/math]

In some sense all cellular algebras "interpolate" between these two extremes by arranging matrix-algebra-like pieces according to the poset [math]\displaystyle{ \Lambda }[/math].

Further examples

Modulo minor technicalities all Iwahori–Hecke algebras of finite type are cellular w.r.t. to the involution that maps the standard basis as [math]\displaystyle{ T_w\mapsto T_{w^{-1}} }[/math].[6] This includes for example the integral group algebra of the symmetric groups as well as all other finite Weyl groups.

A basic Brauer tree algebra over a field is cellular if and only if the Brauer tree is a straight line (with arbitrary number of exceptional vertices).[5]

Further examples include q-Schur algebras, the Brauer algebra, the Temperley–Lieb algebra, the Birman–Murakami–Wenzl algebra, the blocks of the Bernstein–Gelfand–Gelfand category [math]\displaystyle{ \mathcal{O} }[/math] of a semisimple Lie algebra.[5]

Representations

Cell modules and the invariant bilinear form

Assume [math]\displaystyle{ A }[/math] is cellular and [math]\displaystyle{ (\Lambda,i,M,C) }[/math] is a cell datum for [math]\displaystyle{ A }[/math]. Then one defines the cell module [math]\displaystyle{ W(\lambda) }[/math] as the free [math]\displaystyle{ R }[/math]-module with basis [math]\displaystyle{ \lbrace C_\mathfrak{s} \mid \mathfrak{s} \in M(\lambda)\rbrace }[/math] and multiplication

[math]\displaystyle{ aC_\mathfrak{s} := \sum_{\mathfrak{u}} r_a(\mathfrak{u},\mathfrak{s}) C_\mathfrak{u} }[/math]

where the coefficients [math]\displaystyle{ r_a(\mathfrak{u},\mathfrak{s}) }[/math] are the same as above. Then [math]\displaystyle{ W(\lambda) }[/math] becomes an [math]\displaystyle{ A }[/math]-left module.

These modules generalize the Specht modules for the symmetric group and the Hecke-algebras of type A.

There is a canonical bilinear form [math]\displaystyle{ \phi_\lambda: W(\lambda)\times W(\lambda)\to R }[/math] which satisfies

[math]\displaystyle{ C_\mathfrak{st}^\lambda C_\mathfrak{uv}^\lambda \equiv \phi_\lambda(C_\mathfrak{t},C_\mathfrak{u}) C_\mathfrak{sv}^\lambda \mod A(\lt \lambda) }[/math]

for all indices [math]\displaystyle{ s,t,u,v\in M(\lambda) }[/math].

One can check that [math]\displaystyle{ \phi_\lambda }[/math] is symmetric in the sense that

[math]\displaystyle{ \phi_\lambda(x,y) = \phi_\lambda(y,x) }[/math]

for all [math]\displaystyle{ x,y\in W(\lambda) }[/math] and also [math]\displaystyle{ A }[/math]-invariant in the sense that

[math]\displaystyle{ \phi_\lambda(i(a)x,y)=\phi_\lambda(x,ay) }[/math]

for all [math]\displaystyle{ a\in A }[/math],[math]\displaystyle{ x,y\in W(\lambda) }[/math].

Simple modules

Assume for the rest of this section that the ring [math]\displaystyle{ R }[/math] is a field. With the information contained in the invariant bilinear forms one can easily list all simple [math]\displaystyle{ A }[/math]-modules:

Let [math]\displaystyle{ \Lambda_0:=\lbrace \lambda\in\Lambda \mid \phi_\lambda\neq 0\rbrace }[/math] and define [math]\displaystyle{ L(\lambda):=W(\lambda)/\operatorname{rad}(\phi_\lambda) }[/math] for all [math]\displaystyle{ \lambda\in\Lambda_0 }[/math]. Then all [math]\displaystyle{ L(\lambda) }[/math] are absolute simple [math]\displaystyle{ A }[/math]-modules and every simple [math]\displaystyle{ A }[/math]-module is one of these.

These theorems appear already in the original paper by Graham and Lehrer.[1]

Properties of cellular algebras

Persistence properties

  • Tensor products of finitely many cellular [math]\displaystyle{ R }[/math]-algebras are cellular.
  • A [math]\displaystyle{ R }[/math]-algebra [math]\displaystyle{ A }[/math] is cellular if and only if its opposite algebra [math]\displaystyle{ A^{\text{op}} }[/math] is.
  • If [math]\displaystyle{ A }[/math] is cellular with cell-datum [math]\displaystyle{ (\Lambda,i,M,C) }[/math] and [math]\displaystyle{ \Phi\subseteq\Lambda }[/math] is an ideal (a downward closed subset) of the poset [math]\displaystyle{ \Lambda }[/math] then [math]\displaystyle{ A(\Phi):=\sum RC_\mathfrak{st}^\lambda }[/math] (where the sum runs over [math]\displaystyle{ \lambda\in\Lambda }[/math] and [math]\displaystyle{ s,t\in M(\lambda) }[/math]) is a two-sided, [math]\displaystyle{ i }[/math]-invariant ideal of [math]\displaystyle{ A }[/math] and the quotient [math]\displaystyle{ A/A(\Phi) }[/math] is cellular with cell datum [math]\displaystyle{ (\Lambda\setminus\Phi,i,M,C) }[/math] (where i denotes the induced involution and M, C denote the restricted mappings).
  • If [math]\displaystyle{ A }[/math] is a cellular [math]\displaystyle{ R }[/math]-algebra and [math]\displaystyle{ R\to S }[/math] is a unitary homomorphism of commutative rings, then the extension of scalars [math]\displaystyle{ S\otimes_R A }[/math] is a cellular [math]\displaystyle{ S }[/math]-algebra.
  • Direct products of finitely many cellular [math]\displaystyle{ R }[/math]-algebras are cellular.

If [math]\displaystyle{ R }[/math] is an integral domain then there is a converse to this last point:

  • If [math]\displaystyle{ (A,i) }[/math] is a finite-dimensional [math]\displaystyle{ R }[/math]-algebra with an involution and [math]\displaystyle{ A=A_1\oplus A_2 }[/math] a decomposition in two-sided, [math]\displaystyle{ i }[/math]-invariant ideals, then the following are equivalent:
  1. [math]\displaystyle{ (A,i) }[/math] is cellular.
  2. [math]\displaystyle{ (A_1,i) }[/math] and [math]\displaystyle{ (A_2,i) }[/math] are cellular.
  • Since in particular all blocks of [math]\displaystyle{ A }[/math] are [math]\displaystyle{ i }[/math]-invariant if [math]\displaystyle{ (A,i) }[/math] is cellular, an immediate corollary is that a finite-dimensional [math]\displaystyle{ R }[/math]-algebra is cellular w.r.t. [math]\displaystyle{ i }[/math] if and only if all blocks are [math]\displaystyle{ i }[/math]-invariant and cellular w.r.t. [math]\displaystyle{ i }[/math].
  • Tits' deformation theorem for cellular algebras: Let [math]\displaystyle{ A }[/math] be a cellular [math]\displaystyle{ R }[/math]-algebra. Also let [math]\displaystyle{ R\to k }[/math] be a unitary homomorphism into a field [math]\displaystyle{ k }[/math] and [math]\displaystyle{ K:=\operatorname{Quot}(R) }[/math] the quotient field of [math]\displaystyle{ R }[/math]. Then the following holds: If [math]\displaystyle{ kA }[/math] is semisimple, then [math]\displaystyle{ KA }[/math] is also semisimple.

If one further assumes [math]\displaystyle{ R }[/math] to be a local domain, then additionally the following holds:

  • If [math]\displaystyle{ A }[/math] is cellular w.r.t. [math]\displaystyle{ i }[/math] and [math]\displaystyle{ e\in A }[/math] is an idempotent such that [math]\displaystyle{ i(e)=e }[/math], then the algebra [math]\displaystyle{ eAe }[/math] is cellular.

Other properties

Assuming that [math]\displaystyle{ R }[/math] is a field (though a lot of this can be generalized to arbitrary rings, integral domains, local rings or at least discrete valuation rings) and [math]\displaystyle{ A }[/math] is cellular w.r.t. to the involution [math]\displaystyle{ i }[/math]. Then the following hold

  • [math]\displaystyle{ A }[/math] is split, i.e. all simple modules are absolutely irreducible.
  • The following are equivalent:[1]
  1. [math]\displaystyle{ A }[/math] is semisimple.
  2. [math]\displaystyle{ A }[/math] is split semisimple.
  3. [math]\displaystyle{ \forall\lambda\in\Lambda: W(\lambda) }[/math] is simple.
  4. [math]\displaystyle{ \forall\lambda\in\Lambda: \phi_\lambda }[/math] is nondegenerate.
  1. [math]\displaystyle{ A }[/math] is quasi-hereditary (i.e. its module category is a highest-weight category).
  2. [math]\displaystyle{ \Lambda=\Lambda_0 }[/math].
  3. All cell chains of [math]\displaystyle{ (A,i) }[/math] have the same length.
  4. All cell chains of [math]\displaystyle{ (A,j) }[/math] have the same length where [math]\displaystyle{ j:A\to A }[/math] is an arbitrary involution w.r.t. which [math]\displaystyle{ A }[/math] is cellular.
  5. [math]\displaystyle{ \det(C_A)=1 }[/math].
  • If [math]\displaystyle{ A }[/math] is Morita equivalent to [math]\displaystyle{ B }[/math] and the characteristic of [math]\displaystyle{ R }[/math] is not two, then [math]\displaystyle{ B }[/math] is also cellular w.r.t. a suitable involution. In particular [math]\displaystyle{ A }[/math] is cellular (to some involution) if and only if its basic algebra is.[8]
  • Every idempotent [math]\displaystyle{ e\in A }[/math] is equivalent to [math]\displaystyle{ i(e) }[/math], i.e. [math]\displaystyle{ Ae\cong Ai(e) }[/math]. If [math]\displaystyle{ \operatorname{char}(R) \neq 2 }[/math] then in fact every equivalence class contains an [math]\displaystyle{ i }[/math]-invariant idempotent.[5]

References

  1. 1.0 1.1 1.2 1.3 Graham, J.J; Lehrer, G.I. (1996), "Cellular algebras", Inventiones Mathematicae 123: 1–34, doi:10.1007/bf01232365, Bibcode1996InMat.123....1G 
  2. Weisfeiler, B. Yu.; A. A., Lehman (1968). "Reduction of a graph to a canonical form and an algebra which appears in this process" (in Russian). Scientific-Technological Investigations. 2 9: 12–16. 
  3. Higman, Donald G. (August 1987). "Coherent algebras". Linear Algebra and Its Applications 93: 209-239. doi:10.1016/S0024-3795(87)90326-0. 
  4. Cameron, Peter J. (1999). Permutation Groups. London Mathematical Society Student Texts (45). Cambridge University Press. ISBN 978-0-521-65378-7. https://archive.org/details/permutationgroup0000came. 
  5. 5.0 5.1 5.2 5.3 König, S.; Xi, C.C. (1996), "On the structure of cellular algebras", Algebras and Modules II. CMS Conference Proceedings: 365–386 
  6. Geck, Meinolf (2007), "Hecke algebras of finite type are cellular", Inventiones Mathematicae 169 (3): 501–517, doi:10.1007/s00222-007-0053-2, Bibcode2007InMat.169..501G 
  7. König, S.; Xi, C.C. (1999-06-24), "Cellular algebras and quasi-hereditary algebras: A comparison", Electronic Research Announcements of the American Mathematical Society 5 (10): 71–75, doi:10.1090/S1079-6762-99-00063-3 
  8. König, S.; Xi, C.C. (1999), "Cellular algebras: inflations and Morita equivalences", Journal of the London Mathematical Society 60 (3): 700–722, doi:10.1112/s0024610799008212