1: \documentclass[a4paper, 12pt, draft, reqno]{amsart}
2: \pagestyle{headings}
3: \usepackage[final]{graphicx}
4: \usepackage{amsmath}
5: \usepackage{amssymb}
6: \usepackage{latexsym}
7: \usepackage[cmtip,matrix,arrow]{xy}
8: \UseComputerModernTips
9: \CompileMatrices
10:
11: \setlength{\textheight}{8.9in}
12: \setlength{\textwidth}{6.5in}
13: \setlength{\topmargin}{0in}
14: \setlength{\oddsidemargin}{0in}
15: \setlength{\evensidemargin}{0in}
16:
17: \DeclareMathOperator{\sgn}{sgn}
18: \newcommand{\ZZ}{{\mathbb Z}}
19: \newcommand{\RR}{{\mathbb R}}
20: \newcommand{\NN}{{\mathbb N}}
21: \newcommand{\CC}{{\mathbb C}}
22: \newcommand{\DD}{{\mathcal D}}
23: \newcommand{\dd}{{\mathcal d}}
24: \DeclareMathOperator{\Hom}{Hom}
25: \DeclareMathOperator{\Ext}{Ext}
26: \DeclareMathOperator{\End}{End}
27: \DeclareMathOperator{\soc}{soc}
28: \DeclareMathOperator{\rad}{rad}
29: \DeclareMathOperator{\im}{im}
30: \DeclareMathOperator{\hd}{hd}
31: \DeclareMathOperator{\ch}{ch}
32: \DeclareMathOperator{\mn}{min}
33: \DeclareMathOperator{\res}{res}
34: \DeclareMathOperator{\add}{add}
35: \DeclareMathOperator{\remo}{rem}
36: \newcommand{\pow}{{\mathcal P}}
37: \newcommand{\lat}{{\mathcal L}}
38: \newcommand{\dual}{^{\ast}}
39: \newcommand{\frob}{^{\mbox{\rm\tiny F}}}
40: \newcommand{\tlam}{\tilde{\lambda}}
41: \newcommand{\bmu}{\bar{\mu}}
42: \newcommand{\otherwise}{\mbox{\rm otherwise}}
43: \newcommand{\wif}{\mbox{\rm if }}
44: \newcommand{\wand}{\mbox{\rm and }}
45: \DeclareMathOperator{\good}{{\mathcal F}(\nabla)}
46: \newcommand{\Mod}{\mathrm{mod}}
47: \newcommand{\MMod}{\mathrm{Mod}}
48: \newcommand{\even}{\ \mbox{\rm even}}
49: \newcommand{\odd}{\ \mbox{\rm odd}}
50: \newcommand{\bg}{\bar{G}}
51: \newcommand{\gi}{G_1}
52: \newcommand{\GL}{\mathrm{GL}}
53: \newcommand{\SL}{\mathrm{SL}}
54: \newcommand{\St}{\mathrm{St}}
55: \newcommand{\notequiv}{\not\equiv}
56: \newcommand{\wfor}{\mbox{\rm for }}
57: \DeclareMathOperator{\ind}{ind}
58: \DeclareMathOperator{\wt}{wt}
59: \DeclareMathOperator{\pr}{pr}
60: \DeclareMathOperator{\Stab}{Stab}
61: \DeclareMathOperator{\Grot}{Grot}
62: \newcommand{\fit}{\phantom{\Big(}}
63: \newcommand{\alphac}{\alpha\check{\ }\,}
64: \newcommand{\betac}{\beta\check{\ }\,}
65: \newcommand{\embeds}{\hookrightarrow}
66: \newcommand{\ses}{short exact sequence }
67: \newcommand{\Tt}{\mathtt{T}}
68: \newcommand{\Tu}{\mathtt{U}}
69: \newcommand{\Ts}{\mathtt{S}}
70: \newcommand{\rr}{\mathfrak{r}}
71: \newcommand{\Sym}{\mathfrak{S}}
72: \DeclareMathOperator{\partn}{\Lambda^+}
73: \DeclareMathOperator{\cpartn}{\Lambda^+_{\mbox{\scriptsize{\rm{col}}}}}
74: \DeclareMathOperator{\rpartn}{\Lambda^+_{\mbox{\scriptsize{\rm{row}}}}}
75: \newcommand{\half}{\frac{1}{2}}
76: \newcommand{\bnabla}{\overline{\nabla}}
77: \newcommand{\bi}{\bar{\imath}}
78: \newcommand{\bj}{\bar{\jmath}}
79: \newcommand{\too}{\longrightarrow}
80:
81:
82: \begin{document}
83: \theoremstyle{plain}
84: \newtheorem{thm}{Theorem}[section]
85: \newtheorem{prop}[thm]{Proposition}
86: \newtheorem{lem}[thm]{Lemma}
87: \newtheorem{cor}[thm]{Corollary}
88: \newtheorem{conj}[thm]{Conjecture}
89: \newtheorem{claim}[thm]{Claim}
90: \theoremstyle{definition}
91: \newtheorem{rem}[thm]{Remark}
92: \newtheorem{ass}[thm]{Assumption}
93: \newtheorem{defn}[thm]{Definition}
94: \newtheorem{example}[thm]{Example}
95:
96:
97: \setlength{\parskip}{1ex}
98:
99: \title[The blocks of the
100: Brauer algebra]{The blocks of the Brauer algebra\\ in characteristic zero}
101: \author{Anton Cox}
102: \email{A.G.Cox@city.ac.uk, M.Devisscher@city.ac.uk, P.P.Martin@city.ac.uk}
103: \author{Maud De Visscher}
104: \author{Paul Martin}
105: \address{Centre for Mathematical Science\\
106: City University\\
107: Northampton Square\\
108: London\\
109: EC1V 0HB\\
110: England.}
111: \subjclass[2000]{Primary 20G05}
112:
113: \begin{abstract}We determine the blocks of the Brauer algebra
114: in characteristic zero. We also give information on
115: the submodule structure of standard modules for this algebra.
116: \end{abstract}
117:
118: \maketitle
119: \medskip
120:
121: \section{Introduction}
122:
123: The Brauer algebra $B_n(\delta)$ was introduced in \cite{brauer}
124: in the study of the representation theory of orthogonal and sympletic
125: groups. Over $\CC$, and for integral values of $\delta$, its action on
126: tensor space $T = (\CC^{|\delta|})^{\otimes n}$ can be identified with
127: the centraliser algebra for the corresponding group action. This
128: generalises the Schur-Weyl duality between symmetric and general
129: linear groups \cite{weyl}.
130:
131:
132: If $n$ is fixed, then for all $\delta \geq n$ the centraliser algebra
133: $\End_{O(\delta)}(T)$ has multimatrix structure {\em independent} of
134: $\delta$, and Brauer's algebra $B_n(\delta)$ unifies these algebras,
135: having a basis independent of $\delta$, and a composition which makes
136: sense over any field $k$ and for any $\delta \in k$. The Brauer
137: algebra is well defined in particular for positive integral $\delta <
138: n$, but the action on $T$ is faithful for positive integral $\delta$
139: if and only if $\delta\geq n$.
140:
141: In classical invariant theory one is interested in the Brauer algebra
142: {\it per se} only in so far as it coincides with the centraliser of
143: the classical group action on $T$; i.e., in the case of $\delta$
144: integral with $|\delta|$ large compared to $n$. Here we take another
145: view, and consider the stable properties for fixed $\delta$ and
146: arbitrarily large $n$. In such cases $B_n(\delta)$ is not semisimple
147: for $\delta$ integral. However it belongs to a remarkable family of
148: algebras arising both in invariant theory and in statistical mechanics
149: for which this view is very natural. (For example when considered from
150: the point of view of transfer matrix algebras in statistical mechanics
151: \cite{marbook}.) Indeed much of the structure of $B_n(\delta)$ can be
152: recovered from a suitable global limit of $n$ by localisation (and in
153: this sense its structure does not depend on $n$).
154:
155: This family of algebras can be introduced as follows.
156: Consider the diagram of commuting actions on $T$, with $|\delta|=N$:
157: \[
158: \xymatrix@C=20pt@R=10pt{
159: GL(N) \ar[ddr] & & \ar[ddl] \CC\Sigma_n \\
160: \cup & & \cap \\
161: O(N) \ar[r] & T & \ar[l] B_n(N) \\
162: \cup & & \cap \\
163: \Sigma_N \ar[uur] & & \ar[uul] P_n(N)
164: }
165: \]
166: where the actions of the algebra on the right centralise the action of
167: the group on the left in the same row, and vice versa. The bottom row
168: consists of the diagonal action of $\Sigma_N$ permuting the standard
169: ordered basis of $\CC^N$ on the left, and the partition algebra
170: $P_n(N)$ on the right. The partition algebra $P_n(\delta)$ (for any
171: $\delta$) has a basis of partitions of two rows of $n$ vertices. The
172: Brauer algebra is the subalgebra with basis the subset of pair
173: partitions, and $\CC\Sigma_n$ is the subalgebra with basis the pair
174: partitions such that each pair contains a vertex from each row. The
175: Brauer algebra also has a subalgebra with basis the set of pair
176: partitions which can be represented by noncrossing lines drawn
177: vertex-to-vertex in an interval of the plane with the rows of vertices
178: on its boundary. This is the Temperley-Lieb algebra $T_n(\delta)$.
179:
180: All of these algebras are rather well understood over $\CC$, with the
181: exception of $B_n$. All their decomposition matrices are known, and
182: all of their blocks can be described by an appropriate geometric
183: linkage principle.
184: For $\Sigma_n$ both data are trivial, since it is semisimple.
185: For $T_n$ each standard module has either one or two composition
186: factors and its alcove geometry is affine $A_1$
187: (affine reflections on the real line).
188: For $P_n$ each standard module has either one or two composition
189: factors and its alcove geometry is affine $A_{\infty}$ (although
190: locally the block structure looks like affine $A_1$).
191:
192: Over $\CC$, the Brauer algebra is semisimple for $\delta$
193: sufficiently large, and is generically semisimple
194: \cite{brownbrauer}. Hanlon and Wales studied these algebras in a
195: series of papers \cite{hw1,hw2,hw3,hw4} and conjectured that
196: $B_n(\delta)$ is semi-simple for all non-integral choices of
197: $\delta$. This was proved by Wenzl \cite{wenzlbrauer}.
198:
199: In this paper we determine the blocks of $B_n$ for $\delta$ integral.
200: The simple modules of $B_n$ may be indexed by partitions of those
201: natural numbers congruent to $n$ modulo 2 and not exceeding $n$, and
202: hence by Young diagrams (if $\delta=0$ then the empty partition is
203: omitted). We will call these indexing objects {\it weights}. Given
204: $\delta \in R$ a ring we can associate a {\em charge}
205: $\ch(\epsilon)\in R$ to each box $\epsilon$ in a Young diagram, as
206: shown in Figure \ref{charge}. We will also refer later to the usual
207: {\em content} of boxes which, for the box $\epsilon$ in $i$-th row
208: and $j$-th column is $c(\epsilon)=j-i$. It is easy to see that
209: $ch(\epsilon)=\delta-1+2 c(\epsilon)$. For each pair of diagrams
210: $\lambda$ and $\mu$ we will also need to consider the
211: skew partitions $\lambda/(\lambda\cap\mu)$ and $\mu/(\lambda\cap\mu)$
212: consisting of those boxes occuring in $\lambda$ but not $\mu$ and in
213: $\mu$ but not $\lambda$.
214:
215: \begin{figure}[ht]
216: $$\begin{array}{|c|c|c|cc}
217: \hline \delta -1 & \delta +1 & \delta +3 &
218: \raisebox{0pt}[15pt][10pt]{$\cdots$} \\ \hline \delta -3 & \delta -1
219: & \delta +1 & \raisebox{0pt}[15pt][10pt]{$\cdots$} \\ \hline \delta
220: -5 & \delta -3 & \delta -1 & \raisebox{0pt}[15pt][10pt]{$\cdots$} \\
221: \hline \raisebox{0pt}[10pt][10pt]{$\vdots$} &
222: \raisebox{0pt}[15pt][10pt]{$\vdots$} &
223: \end{array}$$
224: \caption{\label{charge}The charges associated to boxes in a Young diagram}
225: \end{figure}
226:
227: %\]
228: \newcommand{\Yta}[1]{\begin{array}{|r|} \hline #1 \\
229: \hline \end{array}}
230: \newcommand{\Ytab}[2]{\begin{array}{|r|} \hline #1 \\ \hline #2 \\
231: \hline \end{array}}
232: %
233:
234: With these notations we can now state the two main results of the
235: paper (which are valid without restriction on $\delta$).
236:
237: \newtheorem*{thm1}{Corollary \ref{blocks}}
238: \newtheorem*{thm2}{Theorem \ref{lattice}}
239: \begin{thm1}\it
240: The simple modules $L(\lambda)$ and $L(\mu)$ are in the same block if
241: and only if
242:
243: (i) The boxes in $\lambda/(\lambda\cap \mu)$ (respectively
244: $\mu/(\lambda\cap\mu)$) can be put into pairs whose charges sum to
245: zero;
246:
247: (ii) if $\lambda/(\lambda\cap \mu)$ (respectively
248: $\mu/(\lambda\cap\mu)$ contains $\Ytab{1}{{\!\! -\! 1 }}$ with no
249: $\Yta{1}$ to the right of these boxes then it contains an even number
250: of 1/-1 pairs.
251: \end{thm1}
252:
253: Examples illustrating this result are given in Example \ref{exbet}.
254:
255: \begin{thm2}[Summary]\it
256: For any integral $\delta$ and natural number $l$ a standard module can be
257: constructed (for some $B_n(\delta)$) whose socle series length is
258: greater than $l$. This module also has a socle layer containing at
259: least $l$ simples.
260: \end{thm2}
261:
262: The second result shows that the structure of standard modules can
263: become arbitrarily complicated. This is in marked contrast to the
264: partition and Temperley-Lieb algebra, and symmetric group, cases.
265:
266:
267: To prove these results we use the theory of towers of recollement
268: developed in \cite{cmpx}. This approach is already closely modelled,
269: for $B_n$, in work of Doran, Wales, and Hanlon \cite{dhw} (since both
270: papers use the methods developed in \cite{mar1}). This key paper of Doran,
271: Wales and Hanlon will be the starting point for our work, and we will
272: generalise and refine several of their results.
273:
274: The `diagram' algebras $P_n \supset B_n \supset T_n$ are amenable to
275: many powerful representation theory techniques, and yet the
276: representation theory of the Brauer algebra is highly non-trivial in
277: comparison to the others. We shall see that, in terms of degree of
278: difficulty, the study of Brauer representation theory in
279: characteristic zero is an intermediate between the study of
280: `classical' objects in characteristic zero and the {\em grand theme}
281: of the representation theory of finite dimensional algebras, the study
282: of $\Sigma_n$ in characteristic $p$.
283:
284: Another such intermediate class of objects are the Hecke algebras of
285: type $A$ at roots of unity, which are Ringel dual to the generalised
286: Lie objects known as quantum groups. The Brauer algebra $B_n$ in
287: characteristic zero has, through its global limit, more
288: Lie-theory-like structure than $\Sigma_n$ in characteristic $p$ (for
289: which not even a good organisational scheme within which to address
290: the problem is known, for small primes $p$). This is reminiscent of
291: the virtual algebraic Lie theory discussed for the (generalised) blob
292: algebras in \cite{mryom, mwgen}. However in the Brauer algebra case,
293: any candidate for an alcove geometry formulation will be considerably
294: more complicated \cite{nazbrauer, orram}. For these reasons we
295: consider the further study of the Brauer algebra in characteristic
296: zero to be an important problem in representation theory.
297:
298:
299:
300: The paper begins with a section defining the various objects of
301: interest, and a review of their basic properties in the spirit of
302: \cite{cmpx}. This is followed by a brief section describing some basic
303: results about Littlewood-Richardson coefficients which will be needed
304: in what follows. In Section \ref{partial} we begin the analysis of
305: blocks by giving a necessary condition for two weights to be in the
306: same block. This is based on an analysis of the action of certain
307: central elements in the algebra on standard modules, and inductive
308: arguments using Frobenius reciprocity. Section \ref{detour} constructs
309: homomorphisms between standard modules in certain special cases,
310: generalising a result in \cite{dhw}. Although not necessary
311: for the main block result, this is of independent interest.
312:
313: The classification of blocks is completed in Section
314: \ref{blocksec}. The main idea is to show that every block contains a
315: unique minimal weight, and that there is a homomorphism from any
316: standard labelled by a non-minimal weight to one labelled by a
317: smaller weight. We also describe precisely which weights are minimal
318: in their blocks.
319:
320: In Section \ref{bigmod} we consider certain explicit choice of
321: weights, and show inductively, via Frobenius reciprocity arguments,
322: that the corresponding standards can have arbitrarily complicated
323: submodule structures. We conclude by outlining the modifications to
324: our arguments required in the case $\delta=0$.
325:
326: The structure of the Brauer algebra becomes mouch more complicated
327: when considered over an arbitrary field $k$. For general $k$ and
328: $\delta$ integral it is expected that this algebra still acts as a
329: centraliser algebra, and this has been shown in a recent series of
330: papers for the symplectic case \cite{dotpoly1,oe1,ddh}. A necessary
331: and sufficient condition for semisimplicity (which holds over
332: arbitrary fields) was given recently by Rui \cite{ruibrauer}. The study of
333: Young and permutation modules for these algebras has been started in
334: \cite{hpbrauer}.
335:
336:
337: \section{Preliminaries}
338:
339: In this section we will consider the Brauer algebra defined over a
340: general field $k$ of characteristic $p\geq 0$, although we will later
341: restrict attention to the case $k=\CC$. After reviewing the definition
342: of the Brauer algebra, we will show that families of such algebras
343: form towers of recollement in the sense of \cite{cmpx} (which we will
344: see follows from various results of Doran et.~al.~\cite{dhw}). This
345: will be the framework in which we base our analysis of these algebras.
346:
347: Given $n\in \NN$ and $\delta \in k$, the {\it Brauer algebra}
348: $B_n(\delta)$ is a finite dimensional associative $k$-algebra
349: generated by certain Brauer diagrams. A general {\it $(n,t)$-(Brauer)
350: diagram} consists of a rectangular box (or {\it frame}) with $n$
351: distinguished points on the northern boundary and $t$ distinguished
352: points on the southern boundary, which we call {\it nodes}. Each node
353: is joined to precisely one other by a line, and there may also be one
354: or more closed loops inside the frame. Those diagrams without closed
355: loops are called {\it reduced}. We will
356: label the northern nodes from left to right by $1,2,\ldots, n$ and the
357: southern nodes from left to right by
358: $\bar{1},\bar{2},\ldots,\bar{t}$. We identify diagrams if they
359: connect the same
360: pairs of labelled nodes, and have the same number of closed loops.
361: Lines which connect two nodes on the
362: northern (respectively southern) boundary will be called {\it
363: northern} (respectively {\it southern}) {\it arcs}; those connecting a
364: northern node to a southern node will be called {\it propagating
365: lines}.
366:
367: \begin{figure}[ht]
368: \includegraphics{prodex.eps}
369: \caption{Multiplication of two diagrams in $B_6(\delta)$}
370: \label{prodex}
371: \end{figure}
372:
373: Given an $(n,t)$-diagram $A$ and a $(t,u)$-diagram $B$, we define the
374: product $AB$ to be the $(n,u)$-diagram obtained by concatenation of
375: $A$ above $B$ (where we identify the southern nodes of $A$ with the
376: northern nodes of $B$ and then ignore the section of the frame common
377: to both diagrams). As a set, the Brauer algebra $B_n(\delta)$ consists
378: of linear combinations of $(n,n)$-diagrams. This has an obvious
379: additive structure, and multiplication is induced by concatenation. We
380: also impose the relation that any non-reduced diagram containing $m$
381: closed loops equals $\delta^m$ times the same diagram with all closed
382: loops removed. A basis is then given by the set of reduced
383: diagrams. An example of a product of two diagrams in given in Figure
384: \ref{prodex}. For convenience, we set $B_0(\delta)=k$. When
385: no confusion is likely to arise, we denote the algebra $B_n(\delta)$
386: simply by $B_n$.
387:
388: We will now apply as much as possible from the general setup of
389: \cite{cmpx} to the Brauer algebra. The labels (A1), (A2), etc., refer
390: to the axioms in that paper. Henceforth, we assume that $\delta\neq
391: 0$; for the case $\delta=0$ see Section \ref{last}.
392:
393: For $n\geq 2$ consider the idempotent $e_n$ in $B_n$ defined by $1/\delta$
394: times the Brauer diagram where $i$ is joined to $\bar{i}$ for
395: $i=1,\ldots n-2$, and $n-1$ is joined to $n$ and $\overline{n-1}$ is
396: joined to $\bar{n}$. This is illustrated in Figure
397: \ref{idem}.
398:
399: \begin{figure}[ht]
400: \includegraphics{idem.eps}
401: \caption{The idempotent $e_8$}
402: \label{idem}
403: \end{figure}
404:
405: \begin{lem}[A1]\label{A1} For each $n\geq 2$, we have an algebra isomorphism
406: $$\Phi_n \, : \, B_{n-2}\longrightarrow e_n B_n e_n$$ which takes a
407: diagram in $B_{n-2}$ to the diagram in $B_n$ obtained by adding an
408: extra northern and southern arc to the righthand end.
409: \end{lem}
410:
411: This allows us to define, following Green \cite{green}, an exact
412: localization functor
413: \begin{eqnarray*}
414: F_n \, :\, B_n\mbox{\rm -mod} &\longrightarrow& B_{n-2}\mbox{\rm -mod}\\
415: M &\longmapsto& e_n M
416: \end{eqnarray*}
417: and a right exact globalization functor
418: \begin{eqnarray*}
419: G_n\, : \, B_n\mbox{\rm -mod} &\longrightarrow& B_{n+2}\mbox{\rm -mod}\\
420: M &\longmapsto& B_{n+2}e_{n+2}\otimes_{B_n}M.
421: \end{eqnarray*}
422: Note that $F_{n+2}G_n(M)\cong M$ for all $M\in B_n\mbox{\rm -mod}$, and
423: hence $G_n$ is a full embedding.
424:
425: From this we can quickly deduce an indexing set for the isomorphism
426: classes of simple $B_n$-modules. It is easy to see that
427: \begin{equation}\label{symiso}
428: B_n/B_ne_nB_n\cong k\Sigma_n
429: \end{equation}
430: the group algebra of the symmetric
431: group on $n$ symbols. If the simple $k\Sigma_n$-modules are indexed by
432: the set $\Lambda^n$ then by \cite{green} and Lemma \ref{A1}, the
433: simple $B_n$-modules are indexed by the set
434: $$\Lambda_n = \Lambda^n \sqcup \Lambda_{n-2}= \Lambda^n \sqcup
435: \Lambda^{n-2} \sqcup \cdots \sqcup \Lambda^{\mn}$$ where $\min=0$ or
436: $1$ depending on the parity of $n$. If $p=0$ or $p>n$ then the set $\Lambda^n$
437: corresponds to the set of partitions of $n$; we write $\lambda \vdash n$ if
438: $\lambda$ is such a partition.
439:
440: For $m-n$ even we write $\Lambda_n^m$ for $\Lambda^m$ regarded as a
441: subset of $\Lambda_n$. (If $m>n$ then $\Lambda_n^m=\emptyset$.) We
442: also write $\Lambda$ for the disjoint union of all the $\Lambda^n$,
443: and call this the set of {\it weights} for the Brauer algebra. We
444: will henceforth abuse terminology and refer to weights as being in the
445: same block of $B_n$ if the corresponding simple modules are in the same block.
446:
447: For $n\geq 2$ and $0\leq t\leq n/2$, define the idempotent $e_{n,t}$
448: to be $1$ if $k=0$ or $1/\delta^t$ times the Brauer diagram with edges
449: between $i$ and $\bar{i}$ for all $1\leq i\leq n-2t$ and between $j$
450: and $j+1$, and $\bar{j}$ and $\overline{j+1}$ for $n-2t+1 \leq j\leq
451: n-1$. (This is the image of $e_t$ via the isomorphisms arising in
452: Lemma \ref{A1}.) Set $B_{n,t}=B_n / B_n e_{n,t} B_n$.
453:
454: \begin{lem}[A2]\label{A2} The natural multiplication map
455: $$B_{n,t}e_{n,t} \otimes_{e_{n,t}B_{n,t}e_{n,t}} e_{n,t}B_{n,t}
456: \longrightarrow B_{n,t}e_{n,t}B_{n,t}$$ is bijective. If $\delta\neq
457: 0$ and either $p=0$ or
458: $p>n$ then $B_n/B_ne_nB_n$ is semisimple.
459: \end{lem}
460: \begin{proof} The second part follows from (\ref{symiso}) and standard
461: symmetric group results. For the first part, the map is clearly
462: surjective so we only need to show that it is also injective. It is
463: easy to verify that: \\ (i) $B_{n,t}$ has a basis given by all reduced
464: diagrams having at least $n-2t$ propagating lines,\\ (ii)
465: $B_{n,t}e_{n,t}B_{n,t}$ has a basis given by all reduced diagrams
466: having exactly $n-2t$ vertical edges, and\\ (iii)
467: $e_{n,t}B_{n,t}e_{n,t} \cong k\Sigma_{n-2t}$.
468:
469: Now suppose that $X$ and $X'$ are diagrams in $B_{n,t}e_{n,t}$. Any
470: such diagram has a southern edge where the leftmost $n-2t$ nodes lie
471: on propagating lines, with the remaining southern nodes paired
472: consecutively. The northern edge has exactly $t$ northern arcs. We
473: will label such a diagram by $X_{v,1,\sigma}$, where $v$ represents
474: the configuration of northern arcs, $1$ represents the fixed southern
475: boundary, and $\sigma\in\Sigma_{n-2t}$ is the permutation obtained by
476: setting $\sigma(i)=j$ if the $i$th propagating northern node from the
477: left is connected to $\bar{j}$. (For later use we will denote the set
478: of elements $v$ arising thus by $V_{n,t}$, and call such elements {\it
479: partial one-row diagrams}.) Similarly a diagram $Y$ in
480: $e_{n,t}B_{n,t}$ will be labelled by $Y_{1,v,\sigma}$.
481:
482: It will be enough to show that the multiplication map is injective on
483: the set of tensor products of diagram elements. Given
484: $X=X_{v,1,\sigma}$ and $X'=X_{v',1,\sigma'}$ in $B_{n,t}e_{n,t}$ and
485: $Y=Y_{1,w,\tau}$, $Y'=Y_{1,w',\tau'}$ in $e_{n,t}B_{n,t}$, assume that
486: $XY=X'Y'$. Then we must have $v=v'$, $w=w'$ and $\sigma \circ \tau =
487: \sigma' \circ \tau'$. It now follows from the identifcation in (iii)
488: that $X\otimes Y = X'\otimes Y'$ in $B_{n,t}e_{n,t}
489: \otimes_{e_{n,t}B_{n,t}e_{n,t}} e_{n,t}B_{n,t}$.
490: \end{proof}
491:
492: We immediately obtain
493:
494: \begin{cor}[A2$'$]\label{A2'} If $\delta\neq 0$ and either
495: $p=0$ or $p>n$ then $B_n$ is a
496: quasi-hereditary algebra, with heredity chain given by
497: $$0\subset \cdots \subset B_n e_{n,k} B_n \subset \cdots \subset B_n e_{n,0}
498: B_n.$$ The partial ordering is given as follows: for $\lambda, \mu\in
499: \Lambda_n$ we have $\lambda \leq \mu$ if and only if either $\lambda =
500: \mu$ or $\lambda \in \Lambda_n^s$ and $\mu \in \Lambda_n^t$ with
501: $s>t$.
502: \end{cor}
503:
504: Henceforth we assume that $p$ satisfies the conditions in Corollary
505: \ref{A2'}. It follows from the quasi-hereditary structure that for
506: each $\lambda\in \Lambda_n$ we have a standard module
507: $\Delta_n(\lambda)$ having simple head $L_n(\lambda)$ and all other
508: composition factor $L_n(\mu)$ satisfying $\mu < \lambda$. Note that if
509: $\lambda \in \Lambda_n^n$ then
510: $$\Delta_n(\lambda)=L_n(\lambda) \cong S^\lambda$$ the lift to $B_n$
511: of the Specht module for $B_n/B_ne_n B_n \cong k\Sigma_n$.
512:
513:
514: Note also that by \cite[A1]{don2} and arguments as in
515: \cite[Proposition 3]{mryom}, the quasi-hereditary structure is
516: compatible with the globalization and localization functors. That is,
517: for all $\lambda\in \Lambda_n$ we have
518: \begin{eqnarray}\label{makestds}
519: &&G_n(\Delta_n(\lambda))=\Delta_{n+2}(\lambda)\\
520: &&F_n(\Delta_n(\lambda))=\left\{ \begin{array}{ll}
521: \Delta_{n-2}(\lambda) & \mbox{if $\lambda \in \Lambda_{n-2}$}\\ 0
522: & \mbox{otherwise} \end{array} \right.
523: \end{eqnarray}
524: As $F_n$ is exact we also have that
525: \begin{eqnarray}\label{Lres}
526: F_n(L_n(\lambda))=\left\{ \begin{array}{ll}
527: L_{n-2}(\lambda) & \mbox{if $\lambda \in \Lambda_{n-2}$}\\ 0
528: & \mbox{otherwise} \end{array} \right.
529: \end{eqnarray}
530:
531: For every partition $\mu$ of some $m=n-2t$ we can give an explicit
532: construction of the modules $\Delta_n(\mu)$. Let $e=e_{n,t}\in B_n$
533: be as above, so that $eB_ne \cong B_m$. If we denote by $S^{\mu}$ the
534: lift of the Specht module labelled by $\mu$ for $k\Sigma_m$ to $B_m$,
535: then by (\ref{makestds}) we have that
536: \begin{equation}\label{altdef}
537: \Delta_n(\mu)\cong B_n e\otimes_{eB_ne}S^{\mu}.
538: \end{equation}
539: Using this fact, it is
540: easy to give a basis for this module in terms of some basis ${\mathcal
541: B}(\mu)$ of $S^{\mu}$, using the notation introduced during the proof
542: of Lemma \ref{A2}.
543:
544: \begin{lem} \label{basisstandard} If $\mu$ is a partition of $n-2t$ then
545: the module $\Delta_n(\mu)$ has a basis given by
546: $$\{X_{v,1,id}\otimes x\,\, |\,\, v\in V_{n,t} ,\, x\in {\mathcal
547: B}(\mu)\}.$$
548: \end{lem}
549:
550: Via this Lemma we may identify our standard modules
551: $\Delta_n(\lambda)$ with the modules ${\mathcal S}_{\lambda}(n)$ in
552: \cite{dhw} (which in turn come from \cite{brownbrauer}).
553: Note that if we define $\Delta_n(\mu)$ as the tensor product in
554: (\ref{altdef}) then we have a definition that makes sense for all
555: values of $p$. In the non-quasi-hereditary cases these modules still
556: play an important role, as the algebras are cellular \cite{gl} with
557: the $\Delta_n(\mu)$ as cell modules.
558:
559:
560: We will frequently need a second way to relate different Brauer algebras.
561:
562: \begin{lem}[A3]\label{A3} For each $n\geq 1$, the algebra
563: $B_n$ can be identified as a subalgebra of
564: $B_{n+1}$ via the homomorphism which takes a Brauer diagram
565: $X$ in $B_n$ to the Brauer diagram in $B_{n+1}$
566: obtained by adding two vertices $n+1$ and
567: $\overline{n+1}$ with a line between them.
568: \end{lem}
569:
570: Lemma \ref{A3} implies that we can consider the usual restriction and
571: induction functors
572: \begin{eqnarray*}
573: \res_n \, :\, B_n\mbox{\rm -mod} &\longrightarrow& B_{n-1}\mbox{\rm -mod}\\
574: M &\longmapsto& M|_{B_{n-1}}
575: \end{eqnarray*}
576: and
577: \begin{eqnarray*}
578: \ind_n\, : \, B_n\mbox{\rm -mod} &\longrightarrow& B_{n+1}\mbox{\rm -mod}\\
579: M &\longmapsto& B_{n+1}\otimes_{B_n}M.
580: \end{eqnarray*}
581:
582: We can relate these functors to globalisation and localisation via
583:
584: \begin{lem}[A4]\label{A4} (i) For all $n\geq 2$ we have that
585: $$B_ne_n \cong B_{n-1}$$ as a left $B_{n-1}$, right
586: $B_{n-2}$-bimodule.\\
587: (ii) For all $B_n$-modules
588: $M$ we have $$\res_{n+2}(G_n(M))\cong \ind_n(M).$$
589: \end{lem}
590: \begin{proof}
591: (i) Every Brauer diagram in $B_ne_n$ has an edge between
592: $\overline{n-1}$ and $\bar{n}$. Define a map from $B_ne_n$ to
593: $B_{n-1}$ by sending a diagram $X$ to the diagram with $2(n-1)$
594: vertices obtained from $X$ by removing the line connecting
595: $\overline{n-1}$ and $\bar{n}$ and and the line from $n$, and pairing
596: the vertex $\overline{n-1}$ to the vertex originally paired with $n$
597: in $X$. It is easy to check that this gives an isomorphism.\\
598: (ii) Using (i) we have
599: \begin{eqnarray*}
600: \res_{n+2}(G_n(M)) &= (B_{n+2}e_{n+2}\otimes_{B_n}M)|_{B_{n+1}}\\
601: &\cong B_{n+1} \otimes_{B_n} M\cong \ind M.
602: \end{eqnarray*}
603: \end{proof}
604:
605:
606:
607: Let $\lambda$ be a partition of $n$ and $\mu$ be a partition of $n-1$.
608: We write $\lambda \rhd \mu$ and $\mu \lhd \lambda$ if $\mu$ is
609: obtained from $\lambda$ by removing a box from its Young diagram
610: (equivalently if $\lambda$ is obtained from $\mu$ by adding a box to
611: its Young diagram). The following result does {\it not} require any
612: restriction on the characteristic of our field.
613:
614: \begin{prop}[A5 and 6]\label{indres} For
615: $\lambda \in \Lambda_n$ we have short exact sequences
616: $$0\rightarrow \bigoplus_{\mu \lhd \lambda} \Delta_{n+1}(\mu)
617: \rightarrow \ind_n\, \Delta_n(\lambda) \rightarrow \bigoplus_{\mu
618: \rhd \lambda} \Delta_{n+1}(\mu)\rightarrow 0$$
619: and
620: $$0\rightarrow \bigoplus_{\mu \lhd \lambda} \Delta_{n-1}(\mu)
621: \rightarrow \res_n\, \Delta_n(\lambda) \rightarrow \bigoplus_{\mu
622: \rhd \lambda} \Delta_{n-1}(\mu)\rightarrow 0.$$
623: \end{prop}
624: \begin{proof} This was proved for $k=\CC$ in \cite[Theorem
625: 4.1 and Corollary 6.4]{dhw} (as the condition $\lambda\vdash n$ in
626: \cite[Corollary 6.4]{dhw} is not needed). However, their proof
627: is valid over any field.
628: \end{proof}
629:
630: Wenzl \cite{wenzlbrauer} has shown that $B_n$ is semisimple when
631: $k=\CC$ and $\delta\notin\ZZ$. (Over an arbitrary field, a necessary
632: and sufficient condition for semisimplicity has been given by Rui
633: \cite{ruibrauer}.) For this reason we do not consider the case of
634: non-integral $\delta$. As we will regularly need to appeal to the
635: representation theory of the symmetric group, which is not well
636: understood in positive characteristic, we will also only consider the
637: characteristic zero case. In summary:
638:
639: {\it Henceforth we will assume that $k=\CC$ and $\delta\in\ZZ\backslash\{0\}$,
640: unless otherwise stated.}
641:
642: \section{Some Littlewood-Richardson coefficients}
643:
644: One of the key results used by \cite{dhw} in their analysis of the
645: Brauer algebra is \cite[Theorem 4.1]{hw3} which decomposes standard
646: modules $\Delta_n(\lambda)$ with $\lambda\vdash n$ as symmetric group
647: modules. Recall that a partition is {\it even} if every part of the
648: partition is even, and that $c_{\mu\eta}^{\lambda}$ denotes a
649: Littlewood-Richardson coefficient. If $\lambda\vdash n$ and
650: $\mu\vdash m$ then \cite[Theorem 4.1]{hw3} states that either
651: $[\res_{\CC\Sigma_n}\Delta_n(\mu):S^{\lambda}]=0$ or
652: $m=n-2t$ for
653: some $t\geq 0$ and
654: \begin{equation}\label{ressymis}
655: [\res_{\CC\Sigma_n}\Delta_n(\mu):S^{\lambda}]=
656: \sum_{\begin{array}{c}\eta
657: \vdash 2t \\ \eta \,\,\mbox{even}\end{array}}c_{\mu \eta}^{\lambda}
658: \end{equation}
659: As this result is stated in terms of Littlewood-Richardson
660: coefficients, we will find it useful to calculate these in certain
661: special cases.
662:
663: \begin{lem}\label{LRstuff} If
664: $\mu \subset \lambda$ are partitions such that $\nu = \lambda / \mu$ is
665: also a partition
666: then
667: $$c_{\mu \eta}^\lambda = \left\{ \begin{array}{ll} 1 & \mbox{if $\eta
668: = \nu$}\\ 0 & \mbox{otherwise.} \end{array}\right.$$
669: \end{lem}
670: \begin{proof}
671: This follows immediately from the definition of
672: Littlewood-Richardson coefficients in terms of rectification of skew
673: tableaux (see \cite[Section 5.1, Corollary 2]{fultab})
674: \end{proof}
675:
676: For our second calculation we will need an alternative definition of
677: Littlewood-Richardson coefficients (which can be found in \cite[2.8.14
678: Corollary]{jk}). When considering a configuration of boxes labeled by
679: elements $b_{ij}$ we say that the configuration is {\it valid} if:\\
680: \phantom{ i} \quad\quad\quad (i) For all $i$, if $y<j$ then $b_{iy}$
681: is in a later column than $b_{ij}$.\\ \phantom{ } \quad\quad\quad(ii)
682: For all $j$, if $x<i$ then $b_{xj}$ is in an earlier row than
683: $b_{ij}$.\\ For each box $(i,j)$ of $\eta$ consider a symbol
684: $b_{ij}$. Then the Littlewood-Richardson coefficient $c_{\mu
685: \eta}^\lambda$ is the number of ways one can form $\lambda$ from $\mu$
686: by adding the boxes of $\eta$ to $\mu$ in the following manner. First
687: add $b_{11},b_{12},\ldots, b_{1\eta_1}$ to $\eta$ to form a new
688: partition $\eta^1$. Continue inductively by adding
689: $b_{i1},b_{i2},\ldots, b_{i\eta_i}$ to $\eta^{i-1}$ to form a new
690: partition $\eta^{i}$. We require that the final configuration of the
691: elements $b_{ij}$ is valid.
692:
693:
694: \begin{lem}\label{LRrec} If
695: $\mu \subset \lambda$ are partitions with $\lambda=(a^b)$ for some $a$
696: and $b$ then there is a unique partition $\eta=(\eta_1,\ldots,\eta_r)$
697: such that $c_{\mu\eta}^{\lambda}\neq 0$, and for this partition we
698: have $c_{\mu\eta}^{\lambda}=1$. Further,
699: $(\lambda/\mu)_i=\eta_{r-i}$.
700: \end{lem}
701: \begin{proof}
702: Consider valid extensions of $\mu$ by any $\eta$ to form $\lambda$.
703: As $\lambda$ is a rectangle, the final row of $\eta$ can only be
704: placed as illustrated in Figure \ref{etaplace}(a). Then the
705: penultimate row of $\eta$ must be placed as illustrated in Figure
706: \ref{etaplace}(b). Continuing in this way we see that the choice of
707: $\eta$ is unique, and the number of boxes in the final row of
708: $\lambda/\mu$ must equal $\eta_1$, in the penultimate row must equal
709: $\eta_2$, and so on.
710: \end{proof}
711:
712: \begin{figure}[ht]
713: $$
714: \begin{tabular}{c|c|c|c|c|c|c|c|c|c|}
715: \multicolumn{7}{l}{ }&
716: $\phantom{\raisebox{0pt}[15pt][10pt]{$\cdots$}}$\\
717: \cline{5-8}
718: \multicolumn{3}{l}{ } &
719: & $b_{(r-1)\eta_r}$ & $\raisebox{0pt}[15pt][10pt]{$\cdots$}$ & $b_{(r-1)2}$ &
720: $b_{(r-1)1}$\\
721: \cline{2-8}
722: $\phantom{\raisebox{0pt}[15pt][10pt]{$\cdots$}}$& $b_{(r-1)\eta_{r-1}}$&
723: $\raisebox{0pt}[15pt][10pt]{$\cdots$}$ & $b_{(r-1)(\eta_r+1)}$ &
724: $b_{r\eta_r}$ & $\raisebox{0pt}[15pt][10pt]{$\cdots$}$ & $b_{r2}$ &
725: $b_{r1}$ \\ \hline
726: \end{tabular}
727: $$\caption{\label{etaplace}The final two rows of $\eta$ in $\lambda$}
728: \end{figure}
729:
730:
731: \section{\label{partial}A partial block result}
732:
733: Doran, Wales, and Hanlon \cite{dhw} have given a necessary condition
734: for the existence of a non-zero homomorphism of $B_n$-modules from
735: $\Delta_n(\lambda)$ to $\Delta_n(\mu)$. We will first elevate this
736: condition to a partial block result, and then give a
737: stronger necessary condition that must also hold for two weights to be
738: in the same block. In section \ref{blocksec} we will see that this
739: stronger condition is also sufficient for two weights to be in the
740: same block.
741:
742: Let $\lambda$ be a partition. For a box $d$ in the corresponding Young
743: diagram $[\lambda]$, we denote by $c(d)$ the {\it content} of $d$.
744: Recall that if $d=(x,y)$ is in the $x$-th row (counting from top to
745: bottom) and in the $y$-th column (counting from left to right) of
746: $[\lambda]$, then $c(d)= y-x$. We denote by ${\bf c}(\lambda)$ the
747: {\it multiset} $\{ c(d)\,\, : \,\, d\in [\lambda]\}$. If $\mu$ is a
748: partition with $[\mu]\subseteq[\lambda]$ we write
749: $\mu\subseteq\lambda$, and denote the skew partition obtained by
750: removing $\mu$ from $\lambda$ by $\lambda/\mu$. We then denote by
751: ${\bf c}(\lambda/\mu)$ the multiset ${\bf c}(\lambda)\backslash{\bf
752: c}(\mu)$.
753:
754: Write $X_{i,j}$ for the Brauer diagram in $B_n$ with edges between $t$
755: and $\bar{t}$ for all $t\neq i,j$ and with edges between $i$ and $j$
756: and between $\bar{i}$ and $\bar{j}$. Note that $B_n$ is generated by
757: the elements $X_{i,j}$ together with the symmetric group $\Sigma_n$
758: (identified with the set of diagrams with $n$ propagating lines). We
759: denote by $T_n$ the element $\sum_{1\leq i <j \leq n} X_{i,j}$ in
760: $B_n$. Recall also the definition of partial one-row diagrams in the proof
761: of Lemma \ref{A2}.
762:
763: \begin{lem}
764: \label{Taction}
765: Let $\mu$ be a partition of $m$ with $m=n-2t$. For all $w\in V_{n,t}$ and
766: $x\in S^\mu$ we have that
767: $$ T_n(X_{w,1,id}\otimes x)=\Big( t(\delta -1) - \sum_{d\in [\mu]}
768: c(d) + \sum_ {1\leq i<j\leq n}(i,j)\Big) (X_{w,1,id}\otimes x)$$
769: where $(i,j)$ denotes the element of $\Sigma_n$ which transposes $i$ and $j$.
770: Hence for all $y\in \Delta_n(\mu)$ we have
771: $$T_ny=\Big( t(\delta -1) - \sum_{d\in [\mu]} c(d) + \sum_{1\leq
772: i<j\leq n}(i,j)\Big)y.$$
773: \end{lem}
774: \begin{proof} This is essentially \cite[Lemma 3.2]{dhw}, together with
775: observations in the proof of \cite[Theorem 3.3]{dhw}.
776: \end{proof}
777:
778: The next result is a slight strengthening of \cite[Theorem 3.3]{dhw}
779: (which in turn generalises \cite[formula before (2.13)]{nazbrauer},
780: which considers the case $\delta\in\NN$). The original results provide
781: a necessary condition for the existence of a homomorphism between two
782: standard modules, but can be refined to prove
783:
784: \begin{prop}\label{constar} Suppose that $[\Delta_n(\mu):L_n(\lambda)]\neq 0$.
785: Then either $\lambda=\mu$ or $\lambda\in \Lambda^r_n$ and $\mu \in
786: \Lambda_n^s$ for some $r-s=2t>0$. Further, we must have
787: $$\mu \subseteq \lambda\quad
788: \mbox{and}\quad t(\delta -1) + \sum_{d\in [\lambda/ \mu]} c(d)
789: =0.$$
790: \end{prop}
791: \begin{proof}
792: The first part of the proposition is clear from the quasi-hereditary
793: structure of $B_n$. For the second part, note that by using the
794: exactness of the localization functor we have
795: $$[\Delta_n(\mu):L_n(\lambda)]=[\Delta_r(\mu):L_r(\lambda)]$$ and
796: hence we may assume that $\lambda$ is a partition of $n$. In this
797: case, $L_n(\lambda)=\Delta_n(\lambda)=S^\lambda$, the lift of the
798: Specht module for $\CC\Sigma_n$ to $B_n(\delta)$, and so any Brauer
799: diagram having fewer than $n$ propagating lines must act as zero on
800: $L_n(\lambda)$. In particular, all the $X_{i,j}$'s act as zero and
801: hence so does $T_n$.
802:
803: The condition that $\mu \subseteq \lambda$ now follows by
804: regarding $\Delta_n(\mu)$ as a $\CC\Sigma_n$-module by restriction and
805: using (\ref{ressymis}) which describes the multiplicities of
806: composition factors of such a module.
807:
808: For the final condition, we know by assumption that there must exist a
809: $B_n$-submodule $M$ of $\Delta_n(\mu)$ and a $B_n$-homomorphism
810: $$\phi \, : \, L_n(\lambda) \longrightarrow \Delta_n(\mu) / M.$$
811: Let $N$ be the $B_n$-submodule of $\Delta_n(\mu)$ containing $M$ such that
812: $$\phi(L_n(\lambda))=N/M.$$ As $N|_{\CC\Sigma_n}$ is semisimple, we can
813: find a $\CC\Sigma_n$-submodule $W$ of $N$ such that $N=W\oplus M$ and
814: $W\cong S^{\lambda}$. By Lemma \ref{Taction} we have for all $y\in W$
815: that
816: $$T_ny=\Big( t(\delta -1) - \sum_{d\in [\mu]} c(d) + \sum_{1\leq
817: i<j\leq n}(i,j)\Big)y.$$
818:
819: But $W\cong S^\lambda$ is a simple $\CC\Sigma_n$-module and $\sum_{1\leq
820: i<j\leq n}(i ,j)$ is in the centre of $\CC\Sigma_n$, so it must act
821: as a scalar on $W$. It is well known \cite[Chapter 1]{Diaconis} that
822: this scalar is given by $\sum_{d\in [\lambda]}c(d)$. Hence we have
823: \begin{eqnarray*}
824: T_ny &=& \Big( t(\delta -1) - \sum_{d\in [\mu]} c(d) + \sum_{d\in
825: [\lambda]}c(d)\Big)y\\ &=&\Big( t(\delta -1) + \sum_{d\in
826: [\lambda / \mu]} c(d)\Big)y.
827: \end{eqnarray*}
828: But $T_n$ must act as zero on $N$
829: and hence $t(\delta-1) + \sum_{d\in
830: [\lambda/ \mu]}c(d)=0$.
831: \end{proof}
832:
833: By standard quasi-heredity arguments \cite[Appendix]{don2}
834: we deduce
835:
836: \begin{cor}\label{weak}
837: Suppose that $\lambda\in \Lambda_n^r$ and $\mu\in \Lambda_n^{s}$
838: with $s<r$. If $\lambda$ and $\mu$ are in the same block then
839: $s=r-2t$ for some $t\in\NN$ and
840: \begin{equation}\label{bal}
841: t(\delta -1) + \sum_{d\in [\lambda]}c (d) - \sum_{d\in [\mu]}c(d) =0.
842: \end{equation}
843: \end{cor}
844:
845: When $t=2$ \cite{dhw} gave a necessary and sufficient
846: condition for the existence of a standard module homomorphism. From
847: their results we obtain
848:
849: \begin{thm}\label{dhwhom}
850: Suppose that $\mu\subset\lambda$ with
851: $|\lambda/\mu|=2$. Then
852: $$\dim\Hom(\Delta_n(\lambda),\Delta_n(\mu))\leq 1$$
853: and is non-zero if and only if $\lambda$ and $\mu$ satisfy (\ref{bal}) with
854: $\lambda/\mu\neq(1^2)$. Indeed, if $\lambda/\mu=(1^2)$ then
855: $$[\Delta_n(\mu):L_n(\lambda)]=0.$$
856: \end{thm}
857: \begin{proof} It is enough to consider the case when
858: $\lambda\vdash n$, as the general case follows by globalisation. If
859: $\lambda$ and $\mu$ do not satisfy the required conditions then there
860: is no composition factor $L_n(\lambda)$ in $\Delta_n(\mu)$
861: (and hence no homomorphism) by Corollary \ref{weak} and the remarks after
862: \cite[Theorem 3.1]{dhw}. In the remaining cases the existence of such
863: a homomorphism was shown in \cite[Theorem 3.4]{dhw}. By the remarks
864: after \cite[Theorem 3.1]{dhw} the multiplicity of the simple module
865: $\Delta_n(\lambda)$ in $\Delta_n(\mu)$ is $1$, and the dimension
866: result is now immediate.
867: \end{proof}
868:
869: The next result is a strengthening of Proposition \ref{constar}.
870:
871: \begin{prop}\label{strongstar} Suppose that
872: $[\Delta_n(\mu):L_n(\lambda)]\neq 0$. Then
873: there is a pairing of the boxes in $\lambda / \mu$ such that the sum
874: of the content of the boxes in each pair is equal to $1-\delta$.
875: \end{prop}
876: \begin{proof}
877: We use induction on $n$; the case $n=2$ is covered by Proposition
878: \ref{constar}. Thus we assume that the result holds for $n-1$ and will
879: show that it holds for $n$.
880:
881: If $[\Delta_n(\mu):L_n(\lambda)]\neq 0$ then by Proposition
882: \ref{constar} we know that $\mu \subseteq \lambda$ and
883: \begin{equation}\label{tocompare}
884: t(\delta -
885: 1) + \sum_{d\in [\lambda / \mu]} c(d) =0
886: \end{equation}
887: where
888: $2t=|\lambda|-|\mu|$. Now suppose, for a contradiction, that there is
889: no pairing of the boxes of $[\lambda / \mu]$ satisfying the
890: condition of the proposition. By localising we may assume that
891: $\lambda$ is a partition of $n$, so that
892: $L_n(\lambda)=\Delta_n(\lambda)$. Thus $\Delta_n(\mu)$ has a submodule
893: $M$ such that $\Delta_n(\lambda)\hookrightarrow \Delta_n(\mu)/M $.
894:
895: The partition $\lambda$ has a removable box $\epsilon_i$ of
896: content $s$ say and by Proposition \ref{indres} we have a surjection
897: $\ind_{n-1}\Delta_{n-1}(\lambda - \epsilon_i) \rightarrow
898: \Delta_n(\lambda)$. Hence we have
899: $$\Hom(\ind_{n-1}\Delta_{n-1}(\lambda - \epsilon_i),
900: \Delta_n(\mu)/M)\neq 0$$
901: and so by Frobenius reciprocity we have
902: $${\rm Hom}(\Delta_{n-1}(\lambda - \epsilon_i),
903: {\res_n}(\Delta_n(\mu)/M))\neq 0.$$ This implies that
904: $\Delta_{n-1}(\lambda - \epsilon_i)=L_{n-1}(\lambda - \epsilon _i)$
905: is a composition factor of $\res_n(\Delta_n(\mu))$. Now using
906: Proposition \ref{indres} we see that either\\
907: (i) the weight $\mu$ has a removable box $\epsilon_j $ such that
908: $[\Delta_{ n-1}(\mu - \epsilon_j):L_{n-1}(\lambda - \epsilon_i)]\neq
909: 0$, or\\
910: (ii) the weight $\mu$ has an addable box $\epsilon_j$ such that
911: $[\Delta_{ n-1}(\mu + \epsilon_j):L_{n-1}(\lambda - \epsilon_i)]\neq
912: 0$.\\
913: We consider each case in turn.
914:
915: In case (i), Proposition \ref{constar} implies that $[\mu -
916: \epsilon_j]\subseteq [\lambda - \epsilon_i]$ and
917: $$t(\delta - 1) + \sum_{d\in [\lambda / \mu]}c(d) -
918: c(\epsilon_i) + c(\epsilon_j ) = 0.$$ Hence from (\ref{tocompare})
919: we must have
920: $$c(\epsilon_j)= c(\epsilon_i)=s$$ and by induction we can find a
921: pairing of the boxes in $(\lambda - \epsilon_i)/(\mu -\epsilon_j)$
922: such that the sum of the content of the boxes in each pair is equal
923: to $1-\delta$. But as multisets
924: $${\bf c}((\lambda - \epsilon_i) / (\mu - \epsilon_j)) = {\bf
925: c}(\lambda/ \mu) - c(\epsilon _i) + c(\epsilon_j) = {\bf
926: c}(\lambda / \mu)$$ and hence there is such a pairing for the
927: boxes of $\lambda/ \mu$. This gives the desired contradiction.
928:
929: Now consider case (ii). Here $\mu$ has an addable box $\epsilon_j$ such
930: that $[\mu + \epsilon_j] \subseteq [\lambda - \epsilon_i]$ and
931: $$(t-1)(\delta - 1) + \sum_{d\in [\lambda / \mu]}c(d) - c(\epsilon_i)
932: - c(\epsilon_j) = 0.$$ Comparing with (\ref{tocompare}) we deduce that
933: $$c(\epsilon_j) + c(\epsilon_i) = 1-\delta.$$ By induction there
934: is a pairing of the boxes of $(\lambda - \epsilon_i) / (\mu +
935: \epsilon_j)$ satisfying the condition of the Proposition. But as multisets
936: $${\bf c}((\lambda - \epsilon_i) / (\mu + \epsilon_j)) = {\bf
937: c}(\lambda / \mu) \, - c(\epsilon_i) - c(\epsilon_j)$$ and as
938: observed above the $c(\epsilon_i)$ and $c(\epsilon_j)$ can be paired
939: in the right way. Hence the boxes of $\lambda / \mu$ can be
940: paired appropriately, which again gives the desired contradiction.
941: \end{proof}
942:
943: When $\delta$ is even we will need a further refinement of Proposition
944: \ref{constar}. Given $\mu\subset\lambda$, consider the boxes with
945: content $-\frac{\delta}{2}$ and $\frac{2-\delta}{2}$ in
946: $\lambda/\mu$. If $[\Delta_n(\mu):L_n(\lambda)]\neq 0$ then these must
947: be paired by Proposition \ref{strongstar}, and so must be in one of
948: the two chain configurations illustrated in Figure \ref{diag} (for
949: some length of chain).
950:
951: \begin{figure}[ht]
952: \includegraphics{diag.eps}
953: \caption{The two possible configurations of paired boxes of
954: contents $-\frac{\delta}{2}$ and $\frac{2-\delta}{2}$}
955: \label{diag}
956: \end{figure}
957:
958: \begin{prop}\label{beststar} Suppose that
959: $[\Delta_n(\mu):L_n(\lambda)]\neq 0$ and $\delta$ is even. If the
960: boxes of content $-\frac{\delta}{2}$ and $\frac{2-\delta}{2}$ are
961: configured as in Figure \ref{diag}(b) then the number of columns in
962: this configuration must be even.
963: \end{prop}
964: \begin{proof}
965: We will show by induction on $n$ that in case
966: (b) the number of columns must be even. The case $n=2$ is covered by
967: Theorem \ref{dhwhom}.
968:
969: By repeated applications of $F$ we may assume that $\lambda\vdash
970: n$. Let $\epsilon_i$ be a removable box of $\lambda$. As in the proof
971: of Proposition \ref{strongstar} we have that if
972: $$[\Delta_n(\mu):L_n(\lambda)]\neq 0$$
973: then either
974: $$[\Delta_{n-1}(\mu-\epsilon_j):L_{n-1}(\lambda-\epsilon_i)]\neq 0$$ for
975: some removable box $\epsilon_j$ of $\mu$ with
976: $c(\epsilon_i)=c(\epsilon_j)$ and $\mu-\epsilon_j\subset\lambda-\epsilon_i$, or
977: $$[\Delta_{n-1}(\mu+\epsilon_j):L_{n-1}(\lambda-\epsilon_i)]\neq 0$$ for
978: some addable box $\epsilon_j$ of $\mu$ with
979: $c(\epsilon_i)+c(\epsilon_j)=1-\delta$ and
980: $\mu+\epsilon_j\subseteq\lambda-\epsilon_i$.
981:
982: If $c(\epsilon_i)$ is not equal to either $-\frac{\delta}{2}$ or
983: $\frac{2-\delta}{2}$ then the boxes of
984: $(\lambda-\epsilon_i)/(\mu-\epsilon_j)$ (respectively of
985: $(\lambda-\epsilon_i)/(\mu+\epsilon_j)$) of content
986: $-\frac{\delta}{2}$ and $\frac{2-\delta}{2}$ are the same as those
987: boxes in $\lambda/\mu$, and so the result follows by induction. Also,
988: by our assumption on the configuration of such boxes the partition
989: $\lambda$ does not have a removable box of content
990: $\frac{2-\delta}{2}$.
991: Thus we may assume that $\lambda$ has only one removable box
992: $\epsilon_i$ of content $-\frac{\delta}{2}$ (and hence that $\lambda$
993: is a rectangle).
994:
995: \begin{figure}[ht]
996: \includegraphics{rectlm.eps}
997: \caption{The partitions $\mu\subset\lambda$, with the
998: configuration as in Figure \ref{diag}(b) shaded}
999: \label{rectlm}
1000: \end{figure}
1001:
1002: We have that $\lambda$ and $\mu$ are of the form shown in Figure
1003: \ref{rectlm}, with $[\Delta_n(\mu):L_n(\lambda)]\neq 0$. So in particular
1004: $$[\res_{\CC\Sigma_n}\Delta_n(\mu):S^{\lambda}]\neq 0.$$
1005: By (\ref{ressymis}) we have
1006: $$[\res_{\CC\Sigma_n}\Delta_n(\mu):S^{\lambda}]= \sum_{\eta\ \mbox{\rm
1007: \tiny even}}c_{\mu\eta}^{\lambda}$$ and hence we must have
1008: $c_{\mu\eta}^{\lambda}\neq 0$ for some even partition
1009: $\eta=(\eta_1,\ldots,\eta_r)$. As $\lambda$ is a rectangle Lemma \ref{LRrec}
1010: implies there is only one possible $\eta$, and that each row of
1011: $\lambda/\mu$ has length $\eta_i$ for some $1\leq i\leq r$. But $\eta$
1012: was an even partition and hence these lengths are all even, which implies
1013: that the number of columns occupied by shaded boxes in Figure
1014: \ref{rectlm} is also even as required.
1015: \end{proof}
1016:
1017:
1018: \begin{defn} We say that $\lambda$ and $\mu$ are {\em $\delta$-balanced}
1019: (or just {\em balanced} when the context is clear) if:
1020: (i) there
1021: exists a pairing of the boxes in $\lambda/(\lambda\cap \mu)$
1022: (respectively in $\mu/(\lambda\cap\mu)$) such that
1023: the contents of each pair sum to $1-\delta$, and (ii) if $\delta$
1024: is even and the boxes with content $-\frac{\delta}{2}$ and
1025: $\frac{2-\delta}{2}$ in $\lambda/(\lambda\cap \mu)$
1026: (respectively in $\mu/(\lambda\cap\mu)$) are configured as
1027: in Figure \ref{diag}(b), then the number of columns in this
1028: configuration is even.
1029: \end{defn}
1030:
1031: Just as for Corollary \ref{weak} we can
1032: immediately deduce from Propositions \ref{strongstar} and
1033: \ref{beststar} the following block result.
1034:
1035: \begin{cor}\label{better}
1036: If $\lambda$ and $\mu$ are in the same block then they are balanced.
1037: \end{cor}
1038:
1039:
1040: \begin{figure}[ht]
1041: \includegraphics{whybetter.eps}
1042: \caption{The diagrams $[\lambda]$, $[\mu]$ and
1043: $[\tau]$ in Example \ref{exbet}(i)}
1044: \label{whybetter}
1045: \end{figure}
1046:
1047: \begin{example}\label{exbet} (i)
1048: Let $\lambda=(6,4^2,2,1)$, $\mu=(5,2^2)$, $\tau=\lambda/(\lambda\cap\mu)$,
1049: and $\delta=1$. The diagrams $[\lambda]$, $[\mu]$, and $[\tau]$ are
1050: illustrated (with their contents) in Figure \ref{whybetter}. Clearly
1051: $$\sum_{d\in [\lambda]}c (d) - \sum_{d\in [\mu]}c(d) =0$$ and hence
1052: $\lambda$ and $\mu$ satisfy the conditions in Corollary
1053: \ref{weak}. However, there is no pairing of the boxes in $[\tau]$ such
1054: that the content of each pair sums to zero, and hence $\lambda$ and
1055: $\mu$ cannot lie in the same block.\\ (ii) Let $\alpha=(5,4^4)$,
1056: $\beta=(5,1^4)$, $\gamma=\alpha/(\alpha\cap\beta)$, and $\delta=2$. The
1057: diagrams $[\alpha]$, $[\beta]$, and $[\gamma]$ are illustrated (with
1058: their contents) in Figure \ref{whybetter}. In this case the boxes in
1059: $[\gamma]$ can be put into pairs such that each pair sums to
1060: $1-\delta=-1$, but the boxes with contents $0$ and $-1$ are in
1061: configuration (b) from Figure \ref{diag}, and occupy an odd number of columns. Hence $\alpha$ and $\beta$ cannot lie in the same block.
1062: \end{example}
1063:
1064: \begin{figure}[ht]
1065: \includegraphics{why2better.eps}
1066: \caption{The diagrams $[\alpha]$, $[\beta]$ and
1067: $[\gamma]$ in Example \ref{exbet}(ii)}
1068: \label{why2better}
1069: \end{figure}
1070:
1071:
1072: By Corollary \ref{better} weights which are not balanced
1073: will lie in different blocks. Hence for a $B_n$-module $X$ we will
1074: denote by $\pr_{\lambda}X$ the direct summand of $X$ with composition
1075: factors $L_n(\mu)$ such that $\mu$ and $\lambda$ are balanced.
1076:
1077: \begin{lem} \label{nonsp}
1078: Suppose that $\lambda\vdash n$ and $\epsilon_i\in\remo(\lambda)$.\\
1079: (i) There exists a $B_n$-module $X$ and a short exact sequence
1080: $$0\too
1081: X\too\pr_{\lambda}\ind_{n-1}\Delta_{n-1}(\lambda-\epsilon_i)\too
1082: \Delta_n(\lambda)\too 0.$$ Here
1083: $X\cong\Delta_n(\lambda-\epsilon_i-\epsilon_j)$ if
1084: $(\lambda-\epsilon_i-\epsilon_j,\lambda)$ is a balanced pair
1085: or $X=0$ if no such $\epsilon_j$ exists. In the former case the
1086: sequence is non-split.
1087:
1088: \noindent (ii) If
1089: $$\Hom(\pr_{\lambda}\ind_{n-1}\Delta_{n-1}(\lambda-\epsilon_i),
1090: \Delta_n(\mu))\neq
1091: 0$$ then $[\Delta_n(\mu):L_n(\lambda)]\neq 0$.
1092: \end{lem}
1093: \begin{proof} (i)
1094: The existence of such a sequence, and the form of $X$, follows from
1095: Proposition \ref{indres} and Corollary \ref{better}.
1096: To see that the sequence
1097: is non split, we proceed by induction on $|\lambda|$, the case where
1098: $\lambda=\emptyset$ being clear. By Frobenius reciprocity we have
1099: \begin{equation}\label{lhrh}
1100: \Hom(\Delta_{n-1}(\lambda-\epsilon_i),
1101: \res\Delta_n(\lambda-\epsilon_i-\epsilon_j))\cong
1102: \Hom(\ind\Delta_{n-1}(\lambda-\epsilon_i),
1103: \Delta_n(\lambda-\epsilon_i-\epsilon_j))
1104: \end{equation}
1105: By (\ref{makestds}) and Lemma \ref{A4}(ii) the left-hand side equals
1106: $$\Hom(\Delta_{n-1}(\lambda-\epsilon_i),
1107: \ind\Delta_{n-2}(\lambda-\epsilon_i-\epsilon_j)).$$ As
1108: $\Delta_{n-1}(\lambda-\epsilon_i)$ is simple, we have by the induction
1109: hypothesis and Theorem \ref{dhwhom} that this Hom-space is one
1110: dimensional. Hence the right-hand side of (\ref{lhrh}) is also one
1111: dimensional, which by another application of Theorem \ref{dhwhom}
1112: implies that the desired sequence is non-split as required.
1113:
1114: Part (ii) is an immediate consequence of (i).
1115: \end{proof}
1116:
1117:
1118: \section{Computing some composition multiplicities\label{detour}}
1119:
1120: So far we have concentrated on conditions which imply that weights lie
1121: in different blocks of the algebra. In this section we will find
1122: certain pairs of weights which do lie in the same block, which we will
1123: demonstrate by determining certain composition factors of standard
1124: modules, and homomorphisms between such modules.
1125:
1126: We first consider the special case where the skew partition
1127: $\lambda/\mu$ is itself a partition. For such pairs we will be able to
1128: show precisely when $L_n(\lambda)$ is a composition factor of
1129: $\Delta_n(\mu)$. We first give a necessary condition, in Proposition
1130: \ref{rectanglenecess}, which is a generalisation of \cite[Corollary
1131: 9.1]{dhw} (the latter only considers the case $\mu=\emptyset$ and
1132: homomorphisms rather than composition factors).
1133:
1134: \begin{prop}\label{rectanglenecess}
1135: Let $\mu \subset \lambda$ are partitions such that $\nu = \lambda /
1136: \mu$ is also a partition. If
1137: $$[\Delta_n(\mu):L_n(\lambda)]\neq 0$$ then $\nu = (a^b)$ where $a$ is
1138: even and $b=\delta + a - 1 + 2c$, where $c$ is the content of the
1139: top lefthand box of $\nu$. Moreover, in this case we have
1140: $$[\Delta_n(\mu): L_n(\lambda)]=1.$$
1141: \end{prop}
1142: \begin{proof}
1143: As usual, by localisation we can assume that $\lambda$ is a partition of
1144: $n$. First suppose that $[\Delta_n(\mu) : L_n(\lambda)]\neq 0$. As
1145: $L_n(\lambda)$ is simply the lift of $S^{\lambda}$ for $\CC\Sigma_n$, we
1146: have that
1147: $$[\res_{\CC\Sigma_n}\Delta_n(\mu) \, : \, S^\lambda ]\neq 0.$$
1148: By (\ref{ressymis}) we have
1149: $$ [\res_{\CC\Sigma_n}\Delta_n(\mu):S^{\lambda}]=
1150: \sum_{\begin{array}{c}\eta
1151: \vdash 2k \\ \eta \,\,\mbox{even}\end{array}}c_{\mu \eta}^{\lambda}.
1152: $$ Hence we see that $\nu$ must be an even partition, and by
1153: Lemma \ref{LRstuff} that $[\Delta_n(\mu) : L_n(\lambda)]=1$.
1154:
1155: On the other hand, using Proposition
1156: \ref{strongstar} we know that there is a pairing of the boxes of
1157: $\nu$ such that the sum of the content of the boxes in each pair is
1158: equal to $1-\delta$. Clearly we have a submodule $M$ of
1159: $\Delta_n(\mu)$ and an embedding
1160: $$\Delta_n(\lambda) \hookrightarrow\Delta_n(\mu)/M.$$
1161: If $\epsilon_i$ is any removable box of
1162: $\lambda$ then we have a surjective homomorphism
1163: $${\ind_{n-1}}\, \Delta_{n-1}(\lambda - \epsilon_i) \rightarrow
1164: \Delta_n(\lambda).$$
1165: Composing these maps we see that
1166: $${\Hom}\, ({\ind_{n-1}}\, \Delta_{n-1}(\lambda - \epsilon_i),
1167: \Delta_n(\mu) / M ) \neq 0$$
1168: and so by Frobenius reciprocity we have
1169: $${\Hom}\, (\Delta_{n-1}(\lambda - \epsilon_i), {\res_n}\,
1170: (\Delta_n(\mu)/M) )\neq 0.$$
1171: Thus
1172: $$[{\res_n}\, \Delta_n(\mu): L_{n-1}(\lambda - \epsilon_i)]\neq 0$$
1173: and
1174: hence either $\mu$ must have a removable box $\epsilon_j$ such that
1175: $$[\Delta_{n-1}(\mu - \epsilon_j) : L_{n-1}(\lambda - \epsilon_i)]\neq 0$$
1176: or $\mu$ must have an addable box $\epsilon_j$ such that
1177: $$[\Delta_{n-1}(\mu + \epsilon_j) : L_{n-1}(\lambda - \epsilon_i)]\neq
1178: 0.$$
1179:
1180: In the first case we have $\mu - \epsilon_j \subset \lambda -
1181: \epsilon_i$ and hence $c (\epsilon_j) = c(\epsilon_i)$. However, as
1182: $\lambda/\mu$ is a partition this is impossible, as no removable box
1183: in $\mu$ can have the same content as some box in $\lambda/\mu$. Hence
1184: we must be in the second case with $\mu + \epsilon_j \subset \lambda -
1185: \epsilon_ i$, so in fact $\epsilon_j$ must be a box in $\nu = \lambda
1186: / \mu$. As $\nu$ is a partition, there is only one such addable box
1187: and its content is given by $c$. Thus we must have
1188: $$c(\epsilon_i) = 1-\delta - c.$$
1189:
1190: Now, if $\nu = \lambda / \mu$
1191: had another removable box then it would have to have the same
1192: content. But different removable boxes have different contents. Hence
1193: $\nu$ can only have one removable box, i.~e. it is a rectangle $\nu =
1194: (a^b)$, where $a$ is even as $\nu$ must be an even partition. The
1195: content of the only removable box of $\nu$ inside of $\lambda$ is
1196: given by $c + a-1 - (b-1) = c+a-b$ and this must be equal to $1-\delta
1197: -c$. Hence we get
1198: $$b=\delta - 1 +a +2c$$
1199: as required.
1200: \end{proof}
1201:
1202: We will show that the condition in Proposition
1203: \ref{rectanglenecess} is also sufficient. This generalises
1204: \cite[Theorem 9.2]{dhw}, which again only considers homomorphisms and
1205: the case $\mu=\emptyset$. Before doing this we will review some
1206: standard symmetric groups results which we will require. Details can be found in \cite[Chapter 7]{fultab}
1207:
1208: We will need to consider a set of idempotents $\{e_{\lambda}\, :\,
1209: \lambda\vdash n\}$ in $\CC\Sigma_n$, such that
1210: $\CC\Sigma_ne_{\lambda}\cong S^{\lambda}$. We will choose
1211: \begin{equation}\label{edef}
1212: e_{\lambda}=\frac{f^{\lambda}}{n!}\sum_{\sigma\in
1213: C_{\lambda}}\sum_{\tau\in R_{\lambda}}\sgn(\sigma)\sigma\tau
1214: \end{equation} where
1215: $f^{\lambda}=\dim S^{\lambda}$, $C_{\lambda}$ is the column stabiliser
1216: of $[\lambda]$ and $R_{\lambda}$ is the row stabiliser of
1217: $[\lambda]$. For example $e_{(2)}$ and $e_{(1,1)}$ (regarded as
1218: elements of $B_2$) are illustrated in Figure \ref{twoids}.
1219:
1220: \begin{figure}[ht]
1221: \includegraphics{twoids.eps}
1222: \caption{The elements $e_{(2)}$ and $e_{(1,1)}$}
1223: \label{twoids}
1224: \end{figure}
1225:
1226: We will also need the fact that
1227: $$\ind_{\CC(\Sigma_a\times\Sigma_b)}^{\CC\Sigma_{a+b}}\left(S^{\mu}\otimes
1228: S^{\nu}\right)\cong \bigoplus_{\lambda\vdash
1229: (n+m)}c_{\mu\nu}^{\lambda}S^\lambda.$$ As all these group algebras
1230: are semisimple, this implies by Frobenius reciprocity that
1231: \begin{equation}\label{resdec}
1232: \res^{\CC\Sigma_n}_{\CC(\Sigma_{a}\times \Sigma_b)}S^{\lambda} \cong
1233: \bigoplus_{\mu \vdash
1234: a,\ \nu\vdash b} c_{\mu\nu}^{\lambda}(S^\mu \otimes S^{\nu}).
1235: \end{equation}
1236: Particular values of $c_{\mu\nu}^{\lambda}$ which we will need are
1237: those where $\nu=(2)$, respectively $\nu=(1,1)$. In these cases
1238: $c_{\mu\nu}^{\lambda}$ is at most $1$, and is non-zero precisely when
1239: $\lambda/\mu$ consists of two boxes in different columns, respectively
1240: different rows.
1241:
1242: \begin{thm}
1243: \label{rectanglesuff} Supppose that
1244: $\mu \subset \lambda$ and $\lambda /\mu = \nu =(a^b)$. If $a$ is even
1245: and $b=\delta - 1 + a +2c$ where $c$ is the content of the top left
1246: box of $\nu$ then $$[\Delta_n(\mu)\, :\, L_n(\lambda)]=1.$$ Moreover,
1247: if $\lambda \vdash n$ then
1248: $${\rm Hom}_{B_n}(L_n(\lambda), \Delta_n(\mu))={\mathbb C}.$$
1249: \end{thm}
1250: \begin{proof}
1251: We can assume without loss of generality that $\lambda \vdash n$. We
1252: have seen in the proof of Proposition \ref{rectanglenecess} that
1253: $[\res_{\CC\Sigma_n}\Delta_n(\mu)\, :\, S^\lambda]=1$. Let $W=e_\lambda
1254: \Delta_n(\mu)$, which is isomorphic to $S^\lambda$ as a
1255: $\Sigma_n$-module. To show this is in fact a $B_n$-submodule of
1256: $\Delta_n(\mu)$, it will be enough to show that $X_{i,j}W=0$ for all
1257: $1\leq i < j \leq n$. Indeed, it is enough to show that this holds for a
1258: single choice of $i$ and $j$, as
1259: $$\sigma X_{i,j}\sigma^{-1}=X_{\sigma(i),\sigma(j)}$$
1260: for all $\sigma\in\Sigma_n$.
1261:
1262: So let us fix $i$ and $j$ with $1\leq i < j \leq n$ and use the embedding
1263: $$\Sigma_{n-2} \times \Sigma_{2} \subset \Sigma_{n}$$ where $\Sigma_2$
1264: is the symmetric group on $\{i,j\}$ and $\Sigma_{n-2}$ the symmetric
1265: group on $\{1,\ldots , n\} \setminus \{i,j\}$. By (\ref{resdec}) and
1266: the remarks following we have
1267: $$\res_{\CC(\Sigma_{n-2}\times \Sigma_2)}W \cong \bigoplus_{\alpha \vdash n-2}
1268: (S^\alpha \otimes S^{(1,1)}) \bigoplus_{\beta \vdash n-2} (S^\beta
1269: \otimes S^{(2)})$$ where we sum over all $\alpha$'s obtained from
1270: $\lambda$ by removing 2 boxes in different rows and over all $\beta$'s
1271: obtained from $\lambda$ by removing two boxes in different columns.
1272:
1273: The map $X_{i,j}\, : \, \Delta_n(\mu) \longrightarrow \Delta_n(\mu)$
1274: is a $\CC\Sigma_{n -2}\times \CC\Sigma_2$-homomorphism. Note that
1275: we have $X_{i,j}(\Delta_n(\mu)) \subset U$ where $U$ is the span of all
1276: elements of the form $X_{w,1,id} \otimes x$ where $w$ has an arc
1277: between $i$ and $j $ and $x\in S^\mu$. Regarding $U$ as a
1278: $B_{n-2}$-module acting on the strings excluding $i$ and $j$ it is
1279: easy to see that $U$ is isomorphic to $\Delta_{n-2}(\mu)$, and the
1280: restriction of this action to $\CC\Sigma_{n-2}$ is the same as
1281: restriction to the action of the first component of $\CC\Sigma_{n
1282: -2}\times \CC\Sigma_2$ regarded as a subalgebra of $B_n$. Also, it is
1283: clear that $X_{ij}$ kills the element $e_{(1,1)}$ in Figure
1284: \ref{twoids}, and hence kills the simple module $S^{(1,1)}$. Combining
1285: these observations with (\ref{ressymis})
1286: we deduce that, as a $\CC\Sigma_{n-2}\times
1287: \CC\Sigma_2$-module, $U$ decomposes as
1288: $$U=\bigoplus_\tau c_\tau (S^\tau \otimes S^{(2)})$$
1289: where
1290: $$c_\tau = \sum_{\begin{array}{c} \tau\vdash n-2 \\ \eta \,
1291: \mbox{even}\end{array}} c_{\mu \eta}^\tau.$$
1292:
1293: Consider the restriction $X_{i,j}\, : \,
1294: W\longrightarrow U$. We want to show that $X_{i,j}W=0$. Look at the
1295: simple summands of $W$. Every summand of the form $S^\alpha \otimes
1296: S^{(1,1)}$ is sent to zero as it does not appear in $U$. Moreover, if
1297: $\mu$ is not contained in $\beta$ then $S^\beta \otimes S^{(2)}$ is
1298: sent to zero as $U$ only contains simple modules $S^\eta \otimes
1299: S^{(2)}$ with $\mu \subset \eta$. So we only need to show that
1300: $$X_{i,j} (S^\beta \otimes S^{(2)})= 0$$ for any $\beta \vdash n-2$
1301: with $\mu \subset \beta$ and $\beta$ obtained from $\lambda$ by
1302: removing two boxes in different columns. But there is only one such $
1303: \beta$, namely the partition obtained from $\lambda$ by removing two
1304: boxes from the last row of $\nu$, i.e $\beta / \mu = (a^{b-1}, a-2)$,
1305: and by Lemma \ref{LRstuff} the coefficient of $S^{\beta}\otimes
1306: S^{(2)}$ in $U$ equals $1$.
1307:
1308: Write $W=V\oplus Y$ where $V=S^\beta \otimes S^{(2)}$. As $V$ is
1309: simple, either $X_{i,j}$ embeds $V$ into $U$ or $X_{i,j}V=0$. Label
1310: the boxes of the partition $\lambda$ with the numbers $1,2, \ldots ,
1311: n$ starting with the first row from left to right, then the second row
1312: from left to right, etc., until the last row. Say that the last box of
1313: the partition $\nu = (a^b)$ inside of $\lambda$ is labelled by $l$. Up
1314: until now $X_{i,j}$ was arbitrary; we now fix $i=l-1$ and $j= l$ and
1315: we want to show that $X_{l-1, l}V=0$.
1316:
1317: \begin{figure}[ht]
1318: \includegraphics{labels.eps}
1319: \caption{The labelling of $\lambda$, with $\nu$ shaded
1320: and $\mu$ unshaded}
1321: \label{labels}
1322: \end{figure}
1323:
1324: Fix a partial one-row diagram $w_0$ with $t$ arcs defined as follows:
1325: suppose the $u$-th row of $\nu$ inside of $\lambda$ is labelled by
1326: $x_u, x_u+1, \ldots, x_u + a -1$ for $1\leq u \leq b$, as illustrated
1327: in Figure \ref{labels}. Then $w_0$ is defined to have arcs $\{x_u, x_u
1328: +1\}, \{x_u+2, x_u +3\}, \ldots \{x_u+a-2, x_u+a-1\}$ for $1\leq u
1329: \leq b$. (Note that $x_b +a-1 = l$.) We will represent elements of
1330: $V_{n,t}$ by adding bars to the Young tableau joining each pair of
1331: nodes connected by an arc. Thus the element $w_0$ will be represented
1332: by the diagram in Figure \ref{bars}. Usually we will only represent
1333: the boxes of $\nu$ in such a diagram.
1334:
1335: \begin{figure}[ht]
1336: \includegraphics{bars.eps}
1337: \caption{A diagrammatic representation of the element
1338: $w_0$ }
1339: \label{bars}
1340: \end{figure}
1341:
1342: Now consider the element of
1343: $\Delta_n(\mu)$ given by $X_{w_0,1,id}\otimes x $ for some $x\in
1344: S^\mu$. Then $e_\lambda (X_{w_0,1,id}\otimes x)\in W$, so it decomposes
1345: as
1346: $$e_\lambda (X_{w_0,1,id}\otimes x) = v + y$$ where $v\in V$ and $y\in
1347: Y$. Note that this decomposition is independent of $\delta$. As
1348: observed above, we have $X_{l-1,l}e_\lambda(X_{w_0,1,id}\otimes
1349: x)=X_{l-1 ,l}v$. Consider the coefficient of $X_{w_0,1,id}\otimes x$
1350: in $X_{l-1,l}v$. We will show that it is a non-zero multiple of
1351: $$\delta -1 + a - b +2c.$$ Hence, as $v$ is independent of $\delta$ we
1352: see that $v\neq 0$, but when $\delta -1 + a - b +2c=0$, we have
1353: $X_{l-1,l}v=0$. Thus $X_{l-1,l}$ cannot embed $V$ into $U$ and so it
1354: must map $V$ to zero.
1355:
1356: Using the labelling of the boxes of $\lambda$ defined above, we will
1357: identify the row and column stabilisers $R_\lambda$ and $C_\lambda$ as
1358: subgroups of $\Sigma_n$, the symmetric group on
1359: $\{1,\ldots,n\}$. From (\ref{edef}) we have
1360: $$e_\lambda (X_{w_0,1,id}\otimes x) = \frac{f^\lambda}{n!}
1361: \sum_{\sigma\in C_\lambda} \sum_{\tau\in R_\lambda}{\rm
1362: sgn}(\sigma)\sigma \tau (X_{w_0,1,id}\otimes x),$$ and so
1363: $$X_{l-1,l}e_\lambda (X_{w_0,1,id}\otimes x) = \frac{f^\lambda}{n!}
1364: \sum_{\sigma \in C_\lambda} \sum_{\tau\in R_\lambda}{\rm
1365: sgn}(\sigma)X_{l-1,l}\sigma \tau (X_ {w_0,1,id}\otimes x).$$ We want
1366: to find the coefficient of $X_{w_0,1,id}\otimes x$ in this sum. We
1367: consider several cases.
1368:
1369:
1370: \noindent\textbf{Case 1:} Suppose that $\sigma\tau X_{w_0,v_k,id}$ has
1371: an arc $\{l-1,l\}$.
1372:
1373: In this case $X_{l-1,l}\sigma \tau (X_{w_0,1,id}\otimes
1374: x)= \delta \sigma \tau (X_{w_0,1 ,id}\otimes x)$. If we want
1375: $\sigma\tau (X_{w_0,1,id}\otimes x)$ to be in ${\rm span}\{X
1376: _{w_0,1,id}\otimes S^\mu\}$ then we must have
1377: \begin{eqnarray*}
1378: & \tau = \tau_1 \tau_2 \qquad &\mbox{with} \,\, \tau_1 \in
1379: R_{\mu\subset\lambda},
1380: \, \tau _2 \in
1381: R_{\lambda}^{0}\\ & \sigma = \sigma_1 \sigma_2 \qquad
1382: &\mbox{with} \,\, \sigma_1 \in C_{\mu\subset \lambda}, \, \sigma_2
1383: \in C_{\lambda}^{0}
1384: \end{eqnarray*}
1385: where $R_{\mu\subset\lambda}$ denotes the subgroup of $R_\lambda$
1386: (isomorphic to $R_{\mu}$) which preserves the rows of $\mu$ and fixes
1387: everything in $\nu$ and $R_{\lambda}^{0}$ denotes the subgroup of
1388: $R_\lambda$ which fixes $X_{w_0,1,id}$ as a diagram (i.e. fixes all
1389: but the $t$ northern arcs, which may be permuted amongst themselves
1390: and be reversed). In a similar way we define $C_{\mu\subset\lambda}$
1391: and $C_{\lambda}^{0}$.
1392:
1393: Set $r=|R_{\lambda}^0|$. As the $a$ columns of $\nu$ are paired by the
1394: bars in $w_0$, and each pair of such columns may be permuted freely by
1395: $C_{\lambda}^0$ we have $|C_{\lambda}^0|=(b!)^{a/2}$. Moreover ${\rm
1396: sgn}(\sigma_2)=1$ as $\sigma_2$ is an even permutation (as it is made
1397: up of pairs of identical permutations, corresponding to the paired
1398: ends of a bar) and so ${\rm sgn}(\sigma) = {\rm sgn}(\sigma _1)$.
1399: Hence in this case we get the contribution
1400: \begin{eqnarray*}
1401: && \frac{f^\lambda}{n!} \sum_{\sigma_2\in C_{\lambda}^0}
1402: \sum_{\sigma_1\in C_{\mu\subset\lambda}} \sum_{\tau _2\in
1403: R_{\lambda}^0} \sum_{\tau_1\in R_{\mu\subset\lambda}} {\rm
1404: sgn}(\sigma_1 \sigma_2)\sigma_1 \sigma_2 \tau_1 \tau_2
1405: (X_{w_0,1,id}\otimes x)\\ && = \frac{f^\lambda}{n!} \sum_{\sigma_2\in
1406: C_{\lambda}^0}\sum_{\tau_2\in R_{\lambda}^0} \sigma_2 \tau_2
1407: (X_{w_0,1,id}\otimes \sum_{\sigma_1\in C_{\mu\subset\lambda}}
1408: \sum_{\tau_1\in R_{\mu\subset\lambda}}{\rm sgn}(\sigma_1)\sigma_1
1409: \tau_1 (x))\\ && = \frac{f^\lambda}{n!} \frac{|\mu
1410: |}{f^\mu}\sum_{\sigma_2\in C_{\lambda}^0} \sum_{\tau_2\in
1411: R_{\lambda}^0} \sigma_2 \tau_2 (X_{w_0,1,id}\otimes e_\mu (x))\\ && =
1412: \frac{f^\lambda}{n!} \frac{|\mu |}{f^\mu} r
1413: (b!)^{a/2}(X_{w_0,1,id}\otimes x)
1414: \end{eqnarray*}
1415: using for the second equality the isomorphisms
1416: $C_{\mu\subset\lambda}\cong C_\mu$ and $R_{\mu\subset\lambda}\cong R_\mu$,
1417: and for the final equality the fact that $e_\mu (x)=x$ for all $x\in
1418: S^\mu$.
1419:
1420:
1421: \noindent
1422: \textbf{Case 2:} Suppose that neither $l-1$ nor $l$ is part of an
1423: arc in $\sigma \tau X_{w_0, 1,id}$.
1424:
1425: In this case $X_{l-1,l}\sigma\tau
1426: X_{w_0,1,id}$ has $t+1$ arcs in the top row and so $X
1427: _{l-1,l}(X_{w_0,1,id}\otimes x)=0$.
1428:
1429:
1430: \noindent
1431: \textbf{Case 3:} Suppose that in $\sigma \tau X_{w_0,v_k,id}$ there are
1432: arcs $\{l-1, i\}$ and $\{l,j\}$.
1433:
1434: In this case, $X_{l-1,l}\sigma\tau
1435: X_{w_0,1,id}$ is obtained from $\sigma \tau X _{w_0,1,id}$ by
1436: replacing the arcs $\{l-1, i\}$ and $\{ l,j\}$ by the arcs $\{
1437: i,j\}$ and $\{l-1, l\}$. Hence if we want to have $X_{l-1,l}\sigma\tau
1438: (X_{w_0,1,id}\otimes x)$ lying in ${\rm span}\, \{X_{w_0,1,id}\otimes
1439: S^\mu\}$ then $\{i,j\}$ must be an arc of $w$ and $i=j\pm 1$. Here
1440: we consider two subcases.
1441:
1442: \noindent
1443: \textbf{Subcase 3(a):} First assume that the pair $\{i, j\}$ is not in
1444: the last double column. Then $ \tau = \tau_2 \tau_1$ with $\tau_1\in
1445: R_{\mu\subset\lambda}$ and $\tau_2\in \tilde{\tau} R_{\lambda}^0$,
1446: where $\tilde{\tau}= (u-1,v)$ or $(u,v)$ such that $v$ is a box of
1447: $\nu$ in the same column as $l$ (possibly $l$ itself) and $u$ is the
1448: box of $\nu$ in the same row as $v$ and in the same column as
1449: $\max(i,j)$. An example of such a situation is illustrated in Figure
1450: \ref{3acase}.
1451:
1452: \begin{figure}[ht]
1453: \includegraphics{3acase.eps}
1454: \caption{An example of subcase 3(a)}
1455: \label{3acase}
1456: \end{figure}
1457:
1458:
1459: Thus we have $b$ choices for $v$ and $(\frac{a}{2}-1)$ choices for the
1460: position of $\{i,j\}$ (and hence of $u$), and so there are $2 b
1461: (\frac{a}{2}-1)$ choices for $\tilde{\tau}$. Hence there are $2 r b
1462: (\frac{a}{2}-1)$ choices for $\tau_2$. Now $\sigma = \sigma_2 \sigma
1463: _1$ where $\sigma_1 \in C_{\mu\subset\lambda}$, and $\sigma_2$
1464: permutes the pairs in all double columns, except the last and the
1465: double column containing $\{j-1,j\}$, arbitrarily. In the last double
1466: column it must send $v-1$ to $l-1$ and $v$ to $l$, and in the double
1467: column containing $\{j-1, j\}$, it can permute the pairs in any way
1468: (as $\{j-1,j\}$ can be any pair in this double column). So we get
1469: $(b!)^{\frac{a}{2}-2} (b-1)! \, b!$ possibilities for $\sigma_2$. Note
1470: also that $\sigma_ 2$ is always an even permutation and so ${\rm
1471: sgn}(\sigma) = {\rm sgn}(\sigma_1) $. Thus in this subcase, we get a
1472: contribution of
1473: \begin{eqnarray*}
1474: &&\frac{f^\lambda}{n!}\,
1475: 2\,r\,b\,(\frac{a}{2}-1)\,(b!)^{(\frac{a}{2}-2)}\,(b-1) !\, b!\,
1476: X_{w_0,1,id}\otimes \sum_{\sigma_1\in C_{\mu\subset\lambda}}
1477: \sum_{\tau_1\in R_{\mu\subset\lambda}}
1478: {\rm sgn}(\sigma_1) \sigma_1 \tau_1 (x)\\ && = \frac{f^\lambda}{n!}
1479: \, \frac{|\mu |!}{f^\mu}\, r \,(a-2)\, (b!)^{\frac{a}{ 2}}
1480: \,X_{w_0,1,id}\otimes x
1481: \end{eqnarray*}
1482: where the equality follows as in Subcase 1.
1483:
1484:
1485: \noindent{\bf Subcase 3(b):} Next assume that the pair $\{i,j\}$ is in
1486: the last column. We must have $\tau = \tau_2 \tau_1$ where $\tau_1 \in
1487: R_{\mu\subset\lambda}$ and $\tau_2 \in R_{\lambda}^0$. Also $\sigma =
1488: \sigma _2 \sigma_1$ where $\sigma_1 \in C_{\mu\subset\lambda}$ and
1489: $\sigma_2 \in (j,l)C_{\lambda}^0$. We have $b-1$ choices for $j$ being
1490: a box of $\nu$ in the same column as $l$. Note that in this case ${\rm
1491: sgn}(\sigma_2)=-1$ and so ${\rm sgn}(\sigma)=-{\rm sgn}(\sigma_1)$.
1492: Hence arguing as in Subcases 1 and 3(a) we get a
1493: contribution of
1494: $$-\, \frac{f^\lambda}{n!}\, \frac{|\mu |!}{f^\mu} \, r\, (b-1)\,
1495: (b!)^{\frac{a} {2}}\, X_{w_0,1,id}\otimes x.$$
1496:
1497: \noindent
1498: \textbf{Case 4:} Suppose that in $\sigma \tau X_{w_0,1,id}$ there is a
1499: link from $l-1$ to $i$, say, and $l$ is not part of an
1500: arc (or vice versa).
1501:
1502: In this case $X_{l-1,l} \sigma \tau X_{w_0,1,id}$ is obtained from
1503: $\sigma \tau X_{w_0,1 ,id}$ by replacing the arc $\{i,l-1\}$ (or
1504: $\{i,l\}$) with the arc $\{l-1, l\}$ and $i$ is not part of an arc
1505: any more. So, if we want to have $X_{l-1,l} \sigma \tau X_{w_0,1,id}$
1506: in ${\rm span}\{X_{w_0,1,id}\otimes S^\mu\}$ then $i$ cannot be one
1507: of the boxes of $\nu$. There are various potential subcases that can
1508: arise. After action by an element of $ R_{\mu\subset\lambda}$ the
1509: element $i$ may be in any box in the same row of $\mu$. There are
1510: three cases: (a) $i$ is now to the left of the first column of $\nu$;
1511: (b) $i$ is above $\nu$ but not above $l-1$ or $l$; (c) $i$ is above
1512: $l-1$ or $l$.
1513:
1514: \noindent{\bf Subcase 4(a):} First, assume that the box $i$ is in a
1515: column to the left of $\nu$ in $\lambda$. In this case, $\tau =
1516: \tau_2 \tau_1$ where $\tau_1 \in R_{\mu\subset\lambda}$ (as we have
1517: already acted by such an element to put $i$ in this case above) and
1518: $\tau_2 \in ( v-1, u)R_{\lambda}^0$, or $\tau_2 \in(v,u)R_{\lambda}^0$
1519: where $v$ is any box in $\nu$ in the same column as $l$ and $u$ is the
1520: box of $\mu$ in the same row as $v$ and in the same column as $i$. An
1521: example of such a situation is illustrated in Figure \ref{4acase}.
1522:
1523: \begin{figure}[ht]
1524: \includegraphics{4acase.eps}
1525: \caption{An example of subcase 4(a)}
1526: \label{4acase}
1527: \end{figure}
1528:
1529: Let $c_1$ be the number of columns of $\lambda$ to the left of $\nu$.
1530: Then there are $ 2 r \, b\, c_1$ possible choices of $\tau_2$. Now
1531: $\sigma=\sigma_2 \sigma_1$ where $\sigma_1 \in C_{\mu\subset\lambda}$
1532: (as $i$ is an arbitrary element in its column of $\mu$) and $\sigma_2$
1533: permutes the pairs in each of the first $(\frac{a}{2}-1)$ double
1534: columns of $\nu$ arbitrarily, and in the last double column sends
1535: $v-1$ to $l-1$ and $v$ to $l$ and then permutes the other pairs
1536: arbitrarily. Note that ${\rm sgn}(\sigma_1)={\rm sgn}(\sigma)$. Hence
1537: (arguing as in earlier cases) we get a contribution of
1538: \begin{eqnarray*}
1539: && \frac{f^\lambda}{n!}\, \frac{|\mu |!}{f^\mu} \, 2 \, r\, b\, c_1 \,
1540: (b!)^{(\frac{a}{2}-1)} \, (b-1)!\, X_{w_0,1,id}\otimes x\\ && =
1541: \frac{f^\lambda}{n!}\, \frac{|\mu |!}{f^\mu} \, r\, (b!)^{\frac{a}{2}}
1542: \, 2 c_1 \, X_{w_0,1,id}\otimes x.
1543: \end{eqnarray*}
1544:
1545:
1546: \noindent{\bf Subcase 4(b):} Suppose that $i$ is a box of $\mu$ which
1547: is above some column of $\nu$ but to the left of $l-1$. Then the only
1548: way to use row and column permutations not involving
1549: $R_{\mu\subset\lambda}$ (which we have already used to position $i$)
1550: to connect $i$ and $l$ (or $l-1$) is by some pair $\tau$ and $\sigma$
1551: similar to that shown in Figure \ref{4bcase}. But (as illustrated) any
1552: such pair does not preserve the remaining edges in $\nu$. Hence this
1553: subcase cannot arise.
1554:
1555: \begin{figure}[ht]
1556: \includegraphics{4bcase.eps}
1557: \caption{An example of the impossibility of subcase
1558: 4(b)}
1559: \label{4bcase}
1560: \end{figure}
1561:
1562:
1563: \noindent{\bf Subcase 4(c):} Finally we are left with the subcase
1564: where after action by $R_{\mu\subset\lambda}$ the element $i$ is in a
1565: box of $\mu$ which is either in the same column as $l-1$ or in the
1566: same column as $l$. In this case $\tau = \tau_2 \tau_1$ where $\tau_1
1567: \in R_{\mu\subset\lambda}$ and $\tau_2 \in R_{\lambda}^0$. Also,
1568: $\sigma = \sigma_2 \sigma_1$ where $\sigma_1 \in
1569: C_{\mu\subset\lambda}$ (as $i$ is an arbitrary element in its column
1570: of $\mu$) and either $\sigma_2 \in (i,l)C_{\lambda}^0$ or $\sigma_ 2
1571: \in (i, l-1)C_{\lambda}^0$. If $c_2$ is the number of columns above
1572: $\nu$ in $\lambda$ then there are $ 2 c_2$ choices for the position of
1573: $i$. Note that here ${\rm sgn}(\sigma_2) = -1 $ and so ${\rm
1574: sgn}(\sigma)= - {\rm sgn}(\sigma_1)$. Hence, in this case we get a
1575: contribution of
1576: $$- \frac{f^\lambda}{n!}\, \frac{|\mu |!}{f^\mu} \, r\,
1577: (b!)^{\frac{a}{2}} 2 c_2 \, \, X_{w_0,1,id}\otimes x.$$
1578:
1579: Note that the final sets of permutations obtained in Subcases 4(a) and
1580: 4(c) are disjoint, so there is no double counting in these
1581: contributions. Now on adding up all contributions from Cases 1--4 we
1582: see that the coefficient of $X_{w_0,1,id }\otimes x$ inside of
1583: $X_{l-1,l}e_\lambda (X_{w_0,1,id}\otimes x)$ is given by
1584: $$\frac{f^\lambda}{n!}\, \frac{|\mu |!}{f^\mu} \, r\,
1585: (b!)^{\frac{a}{2}}\, (\delta -1 +a -b +2 (c_1 - c_2)).$$
1586:
1587: The content of the top left box of the partition $\nu$ inside the
1588: partition $\lambda$ is given by $c=(c_1 +1) - (c_2 +1)=c_1 -
1589: c_2$. Thus we have proved that this coefficient is a non-zero
1590: multiple of $(\delta - 1) +a -b +2c$ as required.
1591: \end{proof}
1592:
1593: \section{\label{blocksec}The blocks of the Brauer algebra}
1594:
1595: In section \ref{partial} we saw that a necessary condition for two
1596: weights $\lambda$ and $\mu$ to be in the same block was that the pair
1597: was balanced. We will now show that this condition is also
1598: sufficient. The key idea will be to construct from any partition
1599: $\lambda$ in a balanced pair with some $\mu\subset\lambda$ a partition
1600: $\nu\subset \lambda$ and a homomorphism connecting $\Delta_n(\lambda)$
1601: and $\Delta_n(\nu)$. This will allow us to proceed by induction.
1602:
1603: Given a partition $\lambda$ we denote by $\add(\lambda)$ the set of
1604: {\it addable} boxes of $\lambda$ (i.e. the set of boxes which may be
1605: added to $\lambda$ such that the new shape is still a
1606: partition). Similarly we denote by $\remo(\lambda)$ the set of {\it
1607: removable} boxes of $\lambda$. If $\mu\subset\lambda$ then we denote
1608: the set of boxes in $\remo(\lambda)$ which are also boxes of
1609: $\lambda/\mu$ by $\remo(\lambda/\mu)$. Distinct boxes in
1610: $\add(\lambda)$ (respectively in $\remo(\lambda)$) have distinct
1611: contents, and we will identify such boxes by their contents. We will
1612: order the boxes in $\lambda$ with a given content by saying that box
1613: $\epsilon$ is {\it smaller} than box $\epsilon'$ if $\epsilon$ appears on an
1614: earlier row than $\epsilon'$.
1615:
1616: \begin{defn}\label{maxisub}
1617: Suppose that $\mu\subset\lambda$ is a balanced pair. For each
1618: $\epsilon_i\in\remo(\lambda/\mu)$ we wish to consider $\mu^i$, the
1619: {\em $i$-maximal balanced subpartition between $\mu$ and
1620: $\lambda$}. This is the maximal partition $\mu^i\subset\lambda$ such
1621: that $\mu^i$ does not contain $\epsilon_i$ and $\lambda$ and $\mu^i$
1622: form a balanced pair. We will construct $\mu^i$ by recursively
1623: defining a series of skew partitions $(\lambda/\mu^i)_j$ which will
1624: eventually equal the skew partition $\lambda/\mu^i$. There is by the
1625: pairing condition a maximal box (i.e. all others smaller) with
1626: content $c(\epsilon_i')$ such that
1627: $c(\epsilon_i)+c(\epsilon_i')=1-\delta$. Let
1628: $(\lambda/\mu^i)_0=\{\epsilon_i,\epsilon_i'\}$. Given
1629: $(\lambda/\mu^i)_m$, we set
1630: $$(\lambda/\mu^i)_{m+1}=(\lambda/\mu^i)_m\cup A_{m+1}\cup A_{m+1}'$$
1631: where $A_{m+1}$ is the set of boxes $\epsilon$ in $\lambda$ such that
1632: $\epsilon$ is to the right of or below a box in $(\lambda/\mu^i)_m$,
1633: and $A'_{m+1}$ is the set of boxes $\epsilon'$ in $(\lambda/\mu)$ such that
1634: $c(\epsilon)+c(\epsilon')=1-\delta$ for some $\epsilon\in A_{m+1}$ and
1635: $\epsilon'$ is maximal with such content among the boxes of
1636: $\lambda/\mu$ not already in $(\lambda/\mu^i)_m$.
1637:
1638: This iterative process eventually stabilises, and we obtain
1639: $(\lambda/\mu^i)_t$ which is a (possibly disconnected) subset of the
1640: edge of $\lambda/\mu$, having width one. (In particular it does not
1641: contain two boxes with the same content.) If $\delta$ is even and
1642: $(\lambda/\mu^i)_t$ does not contain a vertical pair of boxes with
1643: content $\frac{2-\delta}{2}$ and $-\frac{\delta}{2}$, or $\delta$ is
1644: odd and $(\lambda/\mu^i)_t$ does not contain a box of content
1645: $\frac{1-\delta}{2}$ then we set
1646: $\lambda/\mu^i=(\lambda/\mu^i)_t$. Otherwise if $\delta$ is even we
1647: set
1648: \begin{equation}\label{twocase}
1649: (\lambda/\mu^i)_{t+1}=(\lambda/\mu^i)_t\cup\{x,y\}
1650: \end{equation}
1651: where $x,y$ are
1652: the maximal boxes in $\lambda$ of content $\frac{2-\delta}{2}$ and
1653: $-\frac{\delta}{2}$ not in $(\lambda/\mu^i)_t$, and if $\delta$ is odd we set
1654: \begin{equation}\label{onecase}
1655: (\lambda/\mu^i)_{t+1}=(\lambda/\mu^i)_t\cup\{z\}
1656: \end{equation}
1657: where $z$ is the
1658: maximal box in $\lambda$ of content $\frac{1-\delta}{2}$ not in
1659: $(\lambda/\mu^i)_t$. This new skew partition is not necessarily stable
1660: under the addition of boxes $A$ and $A'$ as above, and we repeat that
1661: process again until the skew partition eventually stabilises at some
1662: step $s$. We then set $\lambda/\mu^i=(\lambda/\mu^i)_s$. Thus
1663: $\lambda/\mu^i$ is a removable subset of $\lambda/\mu$ having width at
1664: most two (so at most two boxes with any given content).
1665: \end{defn}
1666:
1667: \begin{figure}[ht]
1668: \includegraphics{3ex.eps}
1669: \caption{Two examples of the $\lambda/\mu^i$ construction}
1670: \label{3ex}
1671: \end{figure}
1672:
1673: \begin{example}
1674: We will now consider several examples of this construction. First let
1675: $\lambda=(6,5,5,2,1)$ and $\mu=(6,4,1)$; this is a balanced pair for
1676: $\delta=2$. If $\epsilon_i$ is any of the removable boxes in
1677: Figure \ref{3ex}(a), then $\lambda/\mu^i$ is the shaded region
1678: shown. For an example where the resulting skew
1679: partition is connected, consider $\lambda=(7,6,5,5,2,2)$ and
1680: $\mu=(7,4,4,1,1)$. This is a balanced pair for $\delta=2$. If $\epsilon_i$
1681: is any of the removable boxes in $\lambda/\mu$ then the skew
1682: partition $\lambda/\mu^i$ is the shaded region shown in Figure
1683: \ref{3ex}(b). In this case there is a pair of boxes in the skew
1684: partition with contents $\frac{2-\delta}{2}$ and $-\frac{\delta}{2}$
1685: (i.e. $0$ and $-1$), but we do not get a strip of width $2$ because
1686: these boxes are not vertically aligned.
1687:
1688: For an example of the full iterative process consider
1689: $\lambda=(7,6,4^4,1^2)$ and $\mu=(4,3^4)$. This is a balanced pair for
1690: $\delta=2$, and after the first part of the iterative process the skew
1691: partition stabilises into the lightly shaded region shown in Figure
1692: \ref{4ex}(a). However, we now have a vertical pair in the skew partition
1693: with contents $\frac{2-\delta}{2}$ and $-\frac{\delta}{2}$ (i.e. $0$
1694: and $-1$). Thus we have to apply (\ref{twocase}), and
1695: add the darkly shaded boxes with content $0$ and $-1$ to this skew
1696: partition. The complement of this is no longer a partition, so we
1697: remove the remaining darkly shaded region by one further application of
1698: the iterative procedure.
1699: \end{example}
1700:
1701: \begin{figure}[ht]
1702: \includegraphics{4ex.eps}
1703: \caption{More examples of the $\lambda/\mu^i$
1704: construction}
1705: \label{4ex}
1706: \end{figure}
1707:
1708: \begin{defn} \label{maxsub}
1709: We now wish to define a {\em maximal balanced subpartition
1710: between $\mu$ and $\lambda$}, which we will denote by $\lambda/\mu'$.
1711: Having constructed a skew partition $\lambda/\mu^i$ for each removable
1712: box $\epsilon_i$ of $\lambda$, we partially order this collection by
1713: inclusion. We then take $\lambda/\mu'$ to be some minimal element of
1714: this set.
1715: \end{defn}
1716:
1717: \begin{example}To see a non-trivial example of this choice, consider
1718: $\lambda=(7,6^2,5,4^2,2)$ and $\mu=(5,3,2^3,1)$. This is a balanced
1719: pair for $\delta=1$, but has several different associated skew
1720: partitions. If we take $\epsilon_i$ to be one of the removable boxes
1721: labelled by $6$ or $-5$ then $\lambda/\mu^i$ equals the entire shaded
1722: region in Figure \ref{4ex}(b). However, if we take $\epsilon_j$ to be
1723: any of the other removable boxes then $\lambda/\mu^j$ consists of the
1724: six darkly shaded boxes. As $\lambda/\mu^j\subset\lambda/\mu^i$, we
1725: take $\lambda/\mu'$ to equal $\lambda/\mu^j$ in this case, and hence
1726: $\mu'=(7,6,4^2,3,2^2)$. (Note that if this example had one additional
1727: box of content $0$ between the two darkly shaded regions, then we
1728: would have to apply (\ref{onecase}) and this box would have associated
1729: skew partition all of the darkly shaded region together with itself
1730: and the diagonally adjacent box with content $0$.)
1731: \end{example}
1732:
1733: The importance of this construction is given by
1734:
1735: \begin{thm}\label{balhom}
1736: If $\mu\subset\lambda$ is a balanced pair, then for any maximal
1737: balanced subpartition $\mu'$ between $\mu$ and $\lambda$ we have
1738: $$\Hom(\Delta_n(\lambda),\Delta_n(\mu'))\neq 0.$$
1739: \end{thm}
1740:
1741: \begin{proof} As usual, we may assume that $\lambda$ is a partition of $n$.
1742: Pick $\epsilon\in\remo(\lambda/\mu')$ with
1743: $|c(\epsilon)-\frac{1-\delta}{2}|$ maximal. (Note that there are at
1744: most two such boxes.) If $\delta$ is even and
1745: $c(\epsilon)=\frac{-\delta}{2}$ or $c(\epsilon)=\frac{2-\delta}{2}$
1746: then $\lambda/\mu'$ is one of the two cases in Figure \ref{triv}(a) or
1747: (b), while if $\delta$ is odd and $c(\epsilon)=\frac{1-\delta}{2}$
1748: then $\lambda/\mu'$ is as in Figure \ref{triv}(c). In each of these
1749: cases there is a non-zero homomorphism from $\Delta_n(\lambda)$ to
1750: $\Delta_n(\mu')$ by Theorem \ref{rectanglesuff} (or more directly by
1751: repeated applications of Frobenius reciprocity). Thus we henceforth
1752: assume we are not in any of these cases.
1753:
1754: \begin{figure}[ht]
1755: \includegraphics{triv.eps}
1756: \caption{Some small $\epsilon$ cases, with matched box denoted
1757: by $\epsilon'$}
1758: \label{triv}
1759: \end{figure}
1760:
1761: Suppose that $\epsilon$ is paired with a maximal $\epsilon'$ of
1762: content $1-\delta-c(\epsilon)$. We will assume that $\epsilon$ is
1763: above, or to the right of, $\epsilon'$, and leave the (obvious)
1764: modifications required for the other case to the reader.
1765:
1766: We will be able to proceed by induction using the following claim.
1767:
1768: \begin{claim}\label{claim}
1769: (i) There is no box of content $c(\epsilon)$ in $\remo(\mu')$.\\
1770: (ii)There is a unique box $\epsilon'$ of content
1771: $1-\delta-c(\epsilon)$ in $\add(\mu')$.\\ (iii) If $|\lambda/\mu'|>2$
1772: then the pair $\lambda-\epsilon$ and $\mu'+\epsilon'$ is balanced, and
1773: the associated skew partition is minimal in the set of those of the
1774: form $(\lambda-\epsilon)/(\mu+\epsilon')^k$, with $\epsilon_k$ in
1775: $\remo((\lambda-\epsilon)/(\mu'+\epsilon'))$. Equivalently, for every
1776: $\epsilon_k$ in $\remo((\lambda-\epsilon)/(\mu'+\epsilon'))$ we have
1777: $$(\lambda-\epsilon)/(\mu+\epsilon')^k=(\lambda-\epsilon)/(\mu+\epsilon').$$
1778: \end{claim}
1779:
1780: Before proving this claim, we show how it can be used to complete the
1781: proof of Theorem \ref{balhom}. Note that if $\lambda-\epsilon$ has
1782: a removable box $\tau$ with content $1-\delta-c(\epsilon)$ then by
1783: minimality $\lambda/\mu'=\{\epsilon,\tau\}$, and we are done
1784: by Theorem \ref{dhwhom} and our assumptions on $\lambda$ . Thus we
1785: assume that there is no such removable box. By Frobenius reciprocity,
1786: Corollary \ref{better}, and Lemma \ref{nonsp}, we have
1787: $$\begin{array}{ll}\Hom(\Delta_n(\lambda),\Delta_n(\mu'))
1788: &\cong\,\Hom(\pr_{\lambda}\ind_{n-1}\Delta_{n-1}(\lambda-\epsilon),
1789: \Delta_n(\mu'))\\
1790: &\cong\,\Hom(\Delta_{n-1}(\lambda-\epsilon),
1791: \pr_{\lambda-\epsilon}\res_{n}\Delta_n(\mu')).
1792: \end{array}$$
1793: By the first two parts of Claim \ref{claim} this latter Homspace is
1794: isomorphic to
1795: $$\Hom(\Delta_{n-1}(\lambda-\epsilon),\Delta_{n-1}(\mu'+\epsilon'))$$
1796: and by the final part of Claim \ref{claim} (and induction) this is
1797: non-zero as required.
1798:
1799: Thus it only remains to prove Claim \ref{claim}.
1800:
1801: \noindent{\bf Proof of Claim \ref{claim}:}
1802: (i) First suppose that there is only one box in $\lambda/\mu'$ with
1803: content $c(\epsilon)$. By construction, if there are any boxes above
1804: $\epsilon$ in $\lambda/\mu'$ then the one with largest content, or its
1805: matched pair, is removable. But this contradicts the choice of
1806: $\epsilon$. The other possibility is that there is a second box
1807: $\tau$ in
1808: $\lambda/\mu'$ with content $c(\epsilon)$, occuping the opposite corner
1809: of a two by two square. Arguing as in the previous case, if there are
1810: any boxes in $\lambda/\mu'$ above this square then this contradicts
1811: the choice of $\epsilon$. These two cases are illustrated in Figure
1812: \ref{corner}(a) and (b). In both these cases we deduce that $\mu'$
1813: cannot have a removable box of content $c(\epsilon)$, as there must be
1814: boxes to the right of any such box in $\mu'$.
1815:
1816: \begin{figure}[ht]
1817: \includegraphics{corner.eps}
1818: \caption{Two corner cases}
1819: \label{corner}
1820: \end{figure}
1821:
1822: (ii) Note that if $\lambda/\mu'$ consists of two boxes then the result
1823: is obvious, so we assume this is not the case. It is also clear that
1824: any addable box of a given content must be unique. Let $\epsilon'$ be
1825: the maximal box in $\lambda/\mu'$ with content
1826: $1-\delta-c(\epsilon)$.
1827:
1828: First suppose that $\lambda/\mu'$ has only one
1829: box with content $c(\epsilon)$, so that we are in the case shown in
1830: Figure \ref{corner}(a). The box $*'$ paired with $*$ in Figure
1831: \ref{corner}(a) must be to the right or above $\epsilon'$, and hence
1832: we are in one of the two configurations shown in Figure
1833: \ref{cor1}.
1834:
1835: \begin{figure}[ht]
1836: \includegraphics{cor1.eps}
1837: \caption{The first corner case}
1838: \label{cor1}
1839: \end{figure}
1840:
1841: The case in Figure \ref{cor1}(a) is impossible by our assumption on
1842: the size of $\lambda/\mu'$ (and minimality), as both $\epsilon$ and
1843: $\epsilon'$ are removable boxes. In the remaining case it is clear
1844: that $\mu'$ has addable box $\epsilon'$, as required.
1845:
1846: Next suppose that $\lambda/\mu'$ has two boxes with content
1847: $c(\epsilon)$, so that we are in the case shown in Figure
1848: \ref{corner}(b). As in the previous case, the box $\alpha'$ paired
1849: with $\alpha$ must be to the right or above $\epsilon'$. If it is
1850: above then we have a configuration similar to that in Figure
1851: \ref{cor1}(a), and hence $\epsilon'$ is a removable box. But this is
1852: impossible exactly as for the case in Figure \ref{cor1}(a). Hence
1853: $\alpha'$ must be to the right of $\epsilon'$, and we must have a
1854: configuration as in Figure \ref{cor2}. But this configuration clearly
1855: has an addable box, $\tau'$, of content $c(\epsilon')$.
1856:
1857: \begin{figure}[ht]
1858: \includegraphics{cor2.eps}
1859: \caption{The second corner case}
1860: \label{cor2}
1861: \end{figure}
1862:
1863: (iii) The two partitions $\lambda-\epsilon$ and $\mu'+\epsilon'$ are
1864: clearly balanced. For minimality we consider the various cases that
1865: can arise. If we are in the case shown in Figure \ref{corner}(a), then
1866: paired boxes are as shown in Figure \ref{cor1}(b). Suppose for a
1867: contradiction that $(\lambda-\epsilon)/(\mu'+\epsilon')$ is not
1868: minimal, and hence contains a smaller skew partition
1869: $\eta$. If $\eta$ does not involve $*$ and $*'$ then it is
1870: also contained in $\lambda/\mu'$, which contradicts the minimality of
1871: this original pair. If $\eta$ does involve $*$ and
1872: $*'$ then this contradicts $\lambda/\mu'$ being minimal, as
1873: $\lambda/\mu'$ contains $\eta\cup\{*,*'\}$, which is a smaller
1874: sub-skew partition of $\lambda/\mu'$.
1875:
1876: Now consider the case shown in Figure \ref{corner}(b), where the
1877: paired boxes are as in Figure \ref{cor2}. As before, suppose for a
1878: contradiction that $(\lambda-\epsilon)/(\mu'+\tau')$ is not minimal,
1879: and hence contains a smaller skew partition $\eta$. If $\eta$ does not
1880: involve $\alpha$ and $\alpha'$ then it is also contained in
1881: $\lambda/\mu'$. If $\eta$ does involve $\alpha$ and $\alpha'$ but not
1882: $\tau$ and $\epsilon'$, then $\eta\cup\{\epsilon,\epsilon'\}$ is a
1883: removable skew inside $\lambda/\mu'$. Finally, if $\eta$ involves all
1884: of $\alpha$, $\alpha'$, $\tau$, and $\epsilon'$, then $\eta$ must also
1885: involve $\beta$ and $\beta'$. Now the skew obtained from $\eta$ by
1886: replacing $\tau$ by $\epsilon$ can be removed from $\lambda/\mu'$. In
1887: each of these three cases we have found a proper removable skew inside
1888: $\lambda/\mu'$, which contradicts the minimality of
1889: $\lambda/\mu'$. Thus $(\lambda-\epsilon)/(\mu'+\tau')$ must be
1890: minimal, which completes the proof of Claim \ref{claim}, and hence
1891: also of Theorem \ref{balhom}.
1892: \end{proof}
1893:
1894: \begin{cor}\label{blocks}
1895: Two weights $\lambda$ and $\mu$ are in the same block of $B_n$ if and
1896: only if they are balanced. Each block contains a unique minimal weight.
1897: \end{cor}
1898: \begin{proof} In Corollary \ref{better} we proved that two weights in
1899: the same block must be balanced. For the reverse implication, we will
1900: proceed by induction. By Theorem \ref{balhom}, if $\lambda$ contains a
1901: smaller partition $\mu$ with which it is balanced, then there exists
1902: some $\mu'\subset\lambda$ with a non-zero homomorphism from
1903: $\Delta_n(\lambda)$ to $\Delta_n(\mu')$. In particular, $\lambda$ and
1904: $\mu'$ will lie in the same block of $B_n$. Thus it is enough to show
1905: that there is a unique minimal partition in the set of partitions
1906: which are balanced with $\lambda$.
1907:
1908: But if there are two such minimal partitions $\mu$ and $\nu$, then set
1909: $\eta=\mu\cap\nu$. Clearly $\eta$ is a partition, and it forms a balanced
1910: pair with both $\mu$ and $\nu$ (and hence with $\lambda$). This
1911: contradicts our assumption of minimality
1912: \end{proof}
1913:
1914: We conclude this section with a description of the minimal partitions
1915: in each block (and hence give a parametrisation of the blocks). We
1916: begin by constructing inductively a skew partition $\hat{\lambda}$
1917: related to $\lambda$. Let $\lambda(0)=\lambda$. Given $\lambda(i)$,
1918: consider $\epsilon\in\remo(\lambda(i))$ such that
1919: $|c(\epsilon)-\frac{1-\delta}{2}|$ is maximal. Suppose that
1920: there does not exist $\epsilon'\in[\lambda]$ with
1921: $c(\epsilon)+c(\epsilon')=1-\delta$ and $\epsilon'\neq
1922: \epsilon$. Hence either the set of rows $\lambda^t$ above and
1923: including the row containing $\epsilon$ (if
1924: $c(\epsilon)-\frac{1-\delta}{2}>0$) or the set of columns $\lambda^l$
1925: to the the left of and including the column containing $\epsilon$ (if
1926: $c(\epsilon)-\frac{1-\delta}{2}<0$) cannot be removed. In this case
1927: set $\lambda(i+1)=\lambda(i)/\lambda^t$, respectively
1928: $\lambda(i+1)=\lambda(i)/\lambda^l$. If there exists
1929: $\epsilon'\in[\lambda]$ with $c(\epsilon)+c(\epsilon')=1-\delta$ and
1930: $\epsilon'\neq \epsilon$ then $\hat{\lambda}=\lambda(i)$. This
1931: procedure will eventually terminate in the construction of
1932: $\hat{\lambda}$.
1933:
1934: \begin{figure}[ht]
1935: \includegraphics{hat.eps}
1936: \caption{An example of the construction of $\hat{\lambda}$}
1937: \label{hat}
1938: \end{figure}
1939:
1940: \begin{example}
1941: As an example of this construction, consider $\delta=1$ and
1942: $\lambda=(7^2,6,5,4,2,1^2)$, as illustrated in Figure \ref{hat}. At
1943: the first stage, we take $\epsilon$ to be the box labelled $-7$, and
1944: hence remove the first column. Next we take the box labelled $5$, and
1945: remove the first two rows. This is followed by the removal of the
1946: second column, then the third row, leaving the skew partition
1947: illustrated in the figure. As the two remaining removable nodes both
1948: have a paired partner (in this case each other) no more rows or
1949: columns need be removed, and we have constructed $\hat{\lambda}$.
1950: \end{example}
1951:
1952:
1953: \begin{prop}
1954: The minimal partitions in each block are precisely those for which
1955: either $\hat{\lambda}=\emptyset$ or a single row or column, or
1956: $\delta$ is even and $\hat{\lambda}$ consists of two rows, the second
1957: of which has final box of content $-\frac{\delta}{2}$.
1958: \end{prop}
1959: \begin{proof}
1960: Clearly if $\hat{\lambda}=\emptyset$ then $\lambda$ is minimal in its
1961: block. In the remaining cases, removal of any part of $\lambda$ can
1962: only involve boxes in $\hat{\lambda}$, and hence to be balanced must
1963: involve either a single unpaired box of content $\frac{1-\delta}{2}$
1964: or a single vertical pair in the configuration shown in Figure
1965: \ref{diag}(b). But this is impossible. Hence we assume that
1966: $\hat{\lambda}$ is not of the form given in the proposition, and will show
1967: that $\lambda$ is not minimal.
1968:
1969: First suppose that $\delta$ is odd. If $\hat{\lambda}$ contains two
1970: boxes of content $\frac{1-\delta}{2}$ then we can construct a maximal
1971: balanced subpartition of $\lambda$, mimicking the process in
1972: Definitions \ref{maxisub} and \ref{maxsub} by starting with
1973: $\epsilon$. Hence by Theorem \ref{balhom} $\lambda$ is non
1974: minimal. If $\hat{\lambda}$ only contains one box $\omega$ with
1975: content $\frac{1-\delta}{2}$ then, again by considering Definitions
1976: \ref{maxisub} and \ref{maxsub} and Theorem \ref{balhom}, any removable
1977: balanced skew-partition must involve $\omega$. The assumption also implies that
1978: $\epsilon$ is in the first row or column of $\hat{\lambda}$.
1979:
1980: Suppose that $\epsilon$ is in the first row of $\hat{\lambda}$ and
1981: there is more than one row (the case where $\epsilon$ is in the first
1982: column is similar). If $\lambda$ is minimal, then no final segment of
1983: this row has a removable paired segment in $\hat{\lambda}$; this can
1984: only arise if $\hat{\lambda}$ is of the form show in Figure
1985: \ref{oddcase} (where shaded areas indicate boxes definitely not in
1986: $\hat{\lambda}$), where $\tau$ is not paired with any box to the right
1987: of $\omega$. But this means that $\tau$ has content
1988: $\frac{1-\delta}{2}$ which is impossible, and hence $\lambda$ is not
1989: minimal.
1990:
1991: \begin{figure}[ht]
1992: \includegraphics{oddcase.eps}
1993: \caption{Possible configuration of $\hat{\lambda}$ when
1994: $\delta$ is odd}
1995: \label{oddcase}
1996: \end{figure}
1997:
1998: Now suppose that $\delta$ is even. If $\hat{\lambda}$ contains either
1999: of the configurations shown in Figure \ref{nogo}(a) and (b) then we
2000: can again construct a maximal balanced subpartition, and by Theorem
2001: \ref{balhom} $\lambda$ is not minimal.
2002:
2003: \begin{figure}[ht]
2004: \includegraphics{nogo.eps}
2005: \caption{Possible configurations in $\hat{\lambda}$ when
2006: $\delta$ is even}
2007: \label{nogo}
2008: \end{figure}
2009:
2010: If $\hat{\lambda}$ contains only one box with content either
2011: $-\frac{\delta}{2}$ or $1-\frac{\delta}{2}$ then this box is either at
2012: the end of the first row or bottom of the first column, which
2013: contradicts the definition of $\hat{\lambda}$. Thus we must have one of
2014: the configurations in Figure \ref{nogo}(c) or (d).
2015:
2016: In case (c) $\epsilon$ must lie at the end of the first column, and in
2017: case (d) at the end of the first row. Arguing as in the $\delta$ odd
2018: case, we see in case (c) that if $\lambda$ is minimal then
2019: $\hat{\lambda}$ must consist of a single column. However, in case (d),
2020: if $\lambda$ is minimal then we either have a single row or we are in
2021: a similar situation to that in Figure \ref{oddcase} and $\tau$ must
2022: have content $-\frac{\delta}{2}$. But this implies that
2023: $\hat{\lambda}$ consists of two rows with the final box having of the
2024: second having content $-\frac{\delta}{2}$, which contradicts our
2025: assumptions on $\lambda$.
2026:
2027: Thus the only cases where $\lambda$ is a minimal partition are those
2028: described in the theorem, and so we are done.
2029: \end{proof}
2030:
2031: \begin{example} To illustrate the last result, consider $\delta=1$
2032: with $\lambda=(7,6^2,5,2^2)$ as shown in Figure \ref{hat2}. The
2033: associated $\hat{\lambda}$ is also shown, and has only one row,
2034: and it is easy to see that $\lambda$ is indeed minimal inside its block.
2035: \end{example}
2036:
2037:
2038: \begin{figure}[ht]
2039: \includegraphics{hat2.eps}
2040: \caption{A minimal weight $\lambda$ and the associated
2041: $\hat{\lambda}$}
2042: \label{hat2}
2043: \end{figure}
2044:
2045:
2046: \section{On the submodule structure of certain standard modules\label{bigmod}}
2047:
2048: In this section we will show that the structure of standard modules
2049: can become arbitrarily complicated (as measured by their Loewy length
2050: and number of simples in each Loewy layer). For this it will be
2051: sufficient to consider certain special partitions which can be more
2052: easily analysed.
2053:
2054:
2055:
2056:
2057: \begin{lem}\label{spotl}
2058: If $\epsilon_i\in\remo(\lambda)$ then
2059: $$[\res_nL_n(\lambda):L_{n-1}(\lambda-\epsilon_i)]\neq 0.$$
2060: \end{lem}
2061: \begin{proof} By (\ref{makestds}) and (\ref{Lres}) we may assume
2062: that $\lambda\vdash n$; the result then follows from Proposition
2063: \ref{indres}.
2064: \end{proof}
2065:
2066: When considering a multi-skew-partition of
2067: differences these skew partitions will be listed in the order
2068: from top right to bottom left. We will extend the power
2069: notation for partitions to multi{\it partitions}, so $((2)^2,(21^3))$ will
2070: denote the triple of partitions $(2)$, $(2)$, and $(21^3)$.
2071:
2072: \begin{example}
2073: To illustrate these definitions we return to the partitions $\lambda$
2074: and $\mu$ considered in Figure \ref{whybetter}. In this case we have
2075: $\add(\lambda)=\{-5,-3,-1,3,6\}$ and
2076: $\remo(\lambda)=\{-4,-2,1,5\}$. Similarly $\add(\mu)=\{-3,1,5\}$ and
2077: $\remo(\mu)=\{-1,4\}$. The pair $(\lambda,\mu)$ is not
2078: $\delta$-balanced for any $\delta$, and $\lambda/(\lambda\cap\mu)$ has
2079: shape $((1),(2^2),(2,1))$.
2080: \end{example}
2081:
2082: We will be interested in $\delta$-balanced pairs $\mu\subset\lambda$
2083: such that the associated skew partition consists entirely of isolated
2084: boxes. If $\mu\subset\lambda$
2085: are balanced with $\lambda/\mu=((1)^{2m})$, denote the matched pairs
2086: of boxes in $\lambda/\mu$ by $\epsilon_1, \epsilon_1', \ldots,
2087: \epsilon_m, \epsilon_m'$ with respective contents $a_1, a_1',
2088: \ldots, a_m, a_m'$. Let ${\pow}(m)$ denote the power set of
2089: $\{1,2,\ldots,m\}$, and for $x\in{\pow}(m)$ set
2090: $$\lambda-x=\lambda-\sum_{i\in x}(\epsilon_i+\epsilon_i').$$
2091: For example, $\lambda-\{1,\ldots,m\}=\mu$.
2092:
2093: \begin{thm}\label{lattice}
2094: Let $\lambda\vdash n$ and $\mu\subset\lambda$ be a balanced pair with
2095: $\lambda/\mu=((1)^{2m})$. Then
2096: $$\dim\Hom(\Delta_n(\lambda),\Delta_n(\mu))=1$$ and
2097: $$[\Delta_n(\mu):L_n(\lambda-x)]=1$$ for all $x\in{\pow}(m)$.
2098:
2099: Further, denote by $\lat(\mu,\lambda)$ the induced lattice in the full
2100: submodule lattice of $\Delta_n(\mu)$ with vertices those simple
2101: modules of the form $L_n(\lambda-x)$ for some $x\in\pow(m)$. Then
2102: $\lat(\mu,\lambda)$ is isomorphic to the superset lattice on
2103: $\pow(m)$; i.e. every submodule of $\Delta_n(\mu)$ which contains
2104: $L_n(\lambda-x)$ contains $L_n(\lambda-y)$ for all $y\subset x$.
2105:
2106: In particular the length of the socle series of $\Delta_n(\mu)$ is at
2107: least $m+1$ and there is a socle series layer containing at least $m$ simple
2108: modules.
2109: \end{thm}
2110:
2111: \begin{rem}
2112: (i) Note that for the induced lattice we are only considering factors of the
2113: form $L_n(\lambda-x)$. In general the module $\Delta_n(\mu)$ will
2114: have many other composition factors. Thus an arrow $A\rightarrow B$ in
2115: our induced lattice structure is to be understood as representing some
2116: non-trivial extension in $\Delta_n(\mu)$ with $A$ in the head and $B$ in
2117: the socle.\\
2118: (ii) Clearly the final part of
2119: the theorem can be strengthened, but is already enough to show
2120: that standard modules can have arbitrarily large socle series lengths
2121: (and layers of arbitrary width).
2122: \end{rem}
2123:
2124: \begin{example} If $\lambda$ and $\mu$ are balanced with
2125: $\lambda/\mu=((1)^{6})=
2126: \{\epsilon_1,\epsilon_1',\epsilon_2,\epsilon_2',\epsilon_3,\epsilon_3'\}$
2127: then the lattice $\lat(\mu,\lambda)$ is illustrated in Figure \ref{latis}.
2128: \end{example}
2129:
2130: \begin{figure}
2131: $$\xymatrix{ & {L_n(\mu)} \ar[dl] \ar[d] \ar[dr]
2132: \\ {L_n(\mu+\epsilon_1+\epsilon_1')}
2133: \ar[d] \ar[dr] & {L_n(\mu+\epsilon_2+\epsilon_2')} \ar[dl]
2134: \ar[dr] & {L_n(\mu+\epsilon_3+\epsilon_3')} \ar[dl] \ar[d] \\
2135: {L_n(\lambda-\epsilon_3-\epsilon_3')}
2136: \ar[dr] & {L_n(\lambda-\epsilon_2-\epsilon_2')} \ar[d] &
2137: {L_n(\lambda-\epsilon_1-\epsilon_1')} \ar[dl] \\ & L_n(\lambda) }$$
2138: \caption{An example of $\lat(\mu,\lambda)$}
2139: \label{latis}
2140: \end{figure}
2141:
2142: \begin{proof}
2143: We proceed by induction on $m$, the result being obvious for $m=0$. By
2144: Frobenius reciprocity we have
2145: \begin{equation}\label{indstep}
2146: \Hom(\ind_{n-1}\Delta_{n-1}(\lambda-\epsilon_i),\Delta_n(\mu)) \cong
2147: \Hom(\Delta_{n-1}(\lambda-\epsilon_i),\res_{n}\Delta_n(\mu)).
2148: \end{equation}
2149: By Proposition \ref{indres} and Corollary \ref{better}, the only
2150: submodule of $\res_n\Delta_n(\mu)$ which can lie in the same block as
2151: $\Delta_{n-1}(\lambda-\epsilon_i)$ is isomorphic to
2152: $\Delta_{n-1}(\mu+\epsilon_i')$, and hence by the inductive hypothesis
2153: the right-hand side of (\ref{indstep}) is one dimensional. Lemma
2154: \ref{nonsp} now implies that $L_n(\lambda)$ is a composition factor of
2155: $\Delta_n(\mu)$. To show that
2156: $\dim\Hom(\Delta_n(\lambda),\Delta_n(\mu))=1$ it will be enough to
2157: show that there is precisely one copy of this composition factor in
2158: $\Delta_n(\mu)$ (which will necessarily lie in the socle).
2159:
2160: By assumption the pair
2161: $(\lambda,\mu)$ is balanced. We will define the {\it bias} of a pair
2162: $(\lambda,\tau)$ with $|\lambda\vartriangle\tau|=2t$ to be
2163: $$b(\lambda,\tau)= \left(\sum_{d\in
2164: \lambda\vartriangle\tau}c(d)\right)-t(1-\delta).$$ Thus a balanced
2165: pair has zero bias. Consider the restriction $\res_n\Delta_n(\mu)$. By
2166: Proposition \ref{indres} we have a short exact sequence
2167: \begin{equation}\label{resbias}
2168: 0\rightarrow \bigoplus_{\tau \lhd \mu} \Delta_{n-1}(\tau)
2169: \rightarrow \res_n\, \Delta_n(\mu) \rightarrow \bigoplus_{\tau \rhd
2170: \mu} \Delta_{n-1}(\tau)\rightarrow 0.
2171: \end{equation}
2172: Note that $\mu$ has no removable boxes with content $\pm a_i$ for
2173: $1\leq i\leq m$, as this would contradict the existence of an addable
2174: node with such a content. Thus the only modules $\Delta_{n-1}(\tau)$
2175: in the sequence (\ref{resbias}) with bias $\pm a_i$ are
2176: $\Delta_{n-1}(\mu+\epsilon_i)$ and $\Delta_{n-1}(\mu+\epsilon_i')$
2177:
2178:
2179: By Lemma \ref{spotl} we have that
2180: $$[\res_{n}L_n(\lambda-x):L_{n-1}(\lambda-x-\epsilon_i)]=1$$ provided
2181: that $i\notin x$. But (by the observations on bias above)
2182: $L_{n-1}(\lambda-x-\epsilon_i)$ can only occur in
2183: $\Delta_{n-1}(\mu+\epsilon_i')$, and by the inductive hypothesis it
2184: occurs there precisely once. By varying $i$ we deduce that there is at
2185: most one copy of each $L_n(\lambda-x)$ in $\Delta_n(\mu)$. But by
2186: induction we know that there is a homomorphism from
2187: $\Delta_{n'}(\lambda-x)$ to $\Delta_{n'}(\mu)$ where $n'=|\lambda-x|$, and
2188: hence by repeated applications of $G$ that there is a homomorphism
2189: from $\Delta_{n}(\lambda-x)$ to $\Delta_{n}(\mu)$. Hence we see that
2190: $L_n(\lambda-x)$ occurs exactly once in $\Delta_n(\mu)$.
2191:
2192: Now consider the summand $\Delta_{n-1}(\mu+\epsilon_i')$ in
2193: $\res_n\Delta_n(\mu)$. This is the only summand of the restriction in
2194: which $L_{n-1}(\lambda-x-\epsilon_i)$ (with $i\notin x$) can arise,
2195: and this simple appears in an extension below
2196: $L_{n-1}(\lambda-y-\epsilon_i)$ for all $y\supset x$ (with $i\notin
2197: y$), by the inductive hypothesis. In particular the copy of
2198: $L_{n-1}(\lambda-x-\epsilon_i)$ appearing in $\res_nL_n(\lambda-x)$
2199: appears below $L_{n-1}(\lambda-x-\epsilon_i-\epsilon_j-\epsilon_j')$
2200: in an extension, and this latter simple must come from
2201: $\res_nL_n(\lambda-x-\epsilon_j-\epsilon_j')$. It follows that
2202: $L_n(\lambda-x)$ must occur in some extension beneath
2203: $L_n(\lambda-x-\epsilon_j-\epsilon_j')$. This argument works for all
2204: $j$ and $x$, and hence verifies the claimed submodule structure except
2205: for the top two layers. However, these are forced by the structure of
2206: standard modules.
2207: \end{proof}
2208:
2209: \section{\label{last}The case $\delta=0$}
2210:
2211: In this section we will sketch the modifications to the preceding
2212: arguments which are required when $\delta=0$. The most obvious change
2213: is that the idempotents $e_n$ considered thus far no longer exist. This
2214: is easily remedied --- however a more serious complication is the
2215: failure of the algebras to be quasihereditary when $n$ is even.
2216:
2217: \begin{figure}[ht]
2218: \includegraphics{alte.eps}
2219: \caption{The element $\bar{e}_n$ in $B_n$}
2220: \label{alte}
2221: \end{figure}
2222:
2223: For $n\geq 3$ let $\bar{e}_n$ be the element illustrated in Figure
2224: \ref{alte}. This is an idempotent for every value of $\delta$, and
2225: satisfies (A1), i.e.
2226: $$\bar{e}_nB_n\bar{e}_n\cong B_{n-2}.$$ Unfortunately we can no longer
2227: prove an analogue of (A2) in general, as the algebras are not
2228: quasihereditary. If $n$ is odd then there are no problems, and the
2229: arguments in the $\delta\neq0$ case for (A1-6) go through unchanged.
2230: The results in Sections 4-7 also generalise, as the various results
2231: needed from \cite{dhw} include the case $\delta=0$, and we thus
2232: deduce the block result in this case.
2233:
2234: For $n$ even, we can no longer appeal directly to the general machinery in
2235: \cite{cmpx}. However, the algebras in this case are cellular, and the
2236: modules considered by \cite{dhw} are precisely the cell modules for
2237: these algebras. The necessary results coming from the general theory
2238: in \cite{cmpx} now have to be verified on an {\it ad hoc} basis, but
2239: this has been carried out in \cite{dhw}. Thus, again, the results in
2240: Sections 4-7 go through unchanged (noting that it is enough to analyse
2241: cell modules when determining blocks by \cite[(3.9.8)]{gl} (see
2242: \cite[2.22 Corollary]{mathas})).
2243:
2244:
2245: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
2246: \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
2247: % \MRhref is called by the amsart/book/proc definition of \MR.
2248: \providecommand{\MRhref}[2]{%
2249: \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
2250: }
2251: \providecommand{\href}[2]{#2}
2252: \begin{thebibliography}{DWH99}
2253:
2254: \bibitem[Bra37]{brauer}
2255: R.~Brauer, \emph{On algebras which are connected with the semisimple continuous
2256: groups}, Ann. of Math. \textbf{38} (1937), 857--872.
2257:
2258: \bibitem[Bro55]{brownbrauer}
2259: W.~Brown, \emph{An algebra related to the orthogonal group}, Michigan Math. J.
2260: \textbf{3} (1955), 1--22.
2261:
2262: \bibitem[CMPX]{cmpx}
2263: A.~G. Cox, P.~P. Martin, A.~E. Parker, and C.~Xi, \emph{Representation theory
2264: of towers of recollement: theory, notes, and examples}, J. Algebra, to
2265: appear.
2266:
2267: \bibitem[DDH]{ddh}
2268: R.~Dipper, S.~Doty, and J.~Hu, \emph{Brauer's centralizer algebras, symplectic
2269: {S}chur algebras and {Schur-Weyl} duality}, preprint.
2270:
2271: \bibitem[Dia]{Diaconis}
2272: P.~Diaconis, \emph{Group representations in probability and statistics},
2273: Institute of Mathematical Statistics Lecture Notes, Monograph Series 11.
2274:
2275: \bibitem[Don98]{don2}
2276: S.~Donkin, \emph{{The $q$-Schur} algebra}, LMS Lecture Notes Series, vol. 253,
2277: Cambridge University Press, 1998.
2278:
2279: \bibitem[Dot98]{dotpoly1}
2280: S.~Doty, \emph{Polynomial representations, algebraic monoids, and {S}chur
2281: algebras of classical type}, J. Pure Appl. Algebra \textbf{123} (1998),
2282: 165--199.
2283:
2284: \bibitem[DWH99]{dhw}
2285: W.~F. Doran, D.~B. Wales, and P.~J. Hanlon, \emph{On the semisimplicity of the
2286: {B}rauer centralizer algebras}, J. Algebra \textbf{211} (1999), 647--685.
2287:
2288: \bibitem[Ful97]{fultab}
2289: W.~Fulton, \emph{Young tableaux}, LMS Student Texts, vol.~35, Cambridge, 1997.
2290:
2291: \bibitem[GL96]{gl}
2292: J.~J. Graham and G.~I. Lehrer, \emph{Cellular algebras}, Invent. Math.
2293: \textbf{123} (1996), 1--34.
2294:
2295: \bibitem[Gre80]{green}
2296: J.~A. Green, \emph{Polynomial representations of {GL$_n$}}, Lecture Notes in
2297: Mathematics 830, Springer, 1980.
2298:
2299: \bibitem[HP]{hpbrauer}
2300: R.~Hartmann and R.~Paget, \emph{{Y}oung modules and filtration multiplicities
2301: for {B}rauer algebras}, preprint.
2302:
2303: \bibitem[HW89a]{hw2}
2304: P.~J. Hanlon and D.~B. Wales, \emph{Eigenvalues connected with {B}rauer's
2305: centralizer algebras}, J. Algebra \textbf{121} (1989), 446--476.
2306:
2307: \bibitem[HW89b]{hw1}
2308: \bysame, \emph{On the decomposition of {B}rauer's centralizer algebras}, J.
2309: Algebra \textbf{121} (1989), 409--445.
2310:
2311: \bibitem[HW90]{hw3}
2312: \bysame, \emph{Computing the discriminants of {B}rauer's centralizer algebras},
2313: Math. Comp. \textbf{54} (1990), 771--796.
2314:
2315: \bibitem[HW94]{hw4}
2316: \bysame, \emph{A tower construction for the radical in {B}rauer's centralizer
2317: algebras}, J. Algebra \textbf{164} (1994), 773--830.
2318:
2319: \bibitem[JK81]{jk}
2320: G.~D. James and A.~Kerber, \emph{The representation theory of the {S}ymmetric
2321: group}, Encyclopedia of Mathematics and its Applications, vol.~16,
2322: Addison-Wesley, 1981.
2323:
2324: \bibitem[Mar91]{marbook}
2325: P.~P. Martin, \emph{Potts models and related problems in statistical
2326: mechanics}, World Scientific, Singapore, 1991.
2327:
2328: \bibitem[Mar96]{mar1}
2329: \bysame, \emph{The structure of the {P}artition algebras}, J. Algebra
2330: \textbf{183} (1996), 319--358.
2331:
2332: \bibitem[Mat99]{mathas}
2333: A.~Mathas, \emph{Iwahori-{Hecke algebras and Schur algebras of the symmetric
2334: groups}}, University lecture series, vol.~15, American Mathematical Society,
2335: 1999.
2336:
2337: \bibitem[MRH04]{mryom}
2338: P.~P. Martin and S.~Ryom-Hansen, \emph{Virtual algebraic {L}ie theory: {Tilting
2339: modules and Ri}ngel duals for blob algebras}, Proc. LMS \textbf{89} (2004),
2340: 655--675.
2341:
2342: \bibitem[MW03]{mwgen}
2343: P.~P. Martin and D.~Woodcock, \emph{Generalized blob algebras and alcove
2344: geometry}, LMS J. of Comp. and Math. \textbf{6} (2003), 249--296.
2345:
2346: \bibitem[Naz96]{nazbrauer}
2347: M.~Nazarov, \emph{{Young's orthogonal form for Brauer's centralizer algebra}},
2348: J. Algebra \textbf{182} (1996), 664--693.
2349:
2350: \bibitem[Oeh01]{oe1}
2351: S.~Oehms, \emph{Centralizer coalgebras, {FRT}-construction and symplectic
2352: monoids}, J. Algebra \textbf{244} (2001), 19--44.
2353:
2354: \bibitem[OR01]{orram}
2355: R.~Orellana and A.~Ram, \emph{Affine braids, markov traces and the category
2356: {$\mathcal O$}}, preprint, 2001.
2357:
2358: \bibitem[Rui05]{ruibrauer}
2359: H.~Rui, \emph{A criterion on the semisimple {B}rauer algebras}, J. Comb. Theory
2360: Ser. A \textbf{111} (2005), 78--88.
2361:
2362: \bibitem[Wen88]{wenzlbrauer}
2363: H.~Wenzl, \emph{On the structure of {B}rauer's centralizer algebra}, Ann. Math.
2364: \textbf{128} (1988), 173--193.
2365:
2366: \bibitem[Wey46]{weyl}
2367: H.~Weyl, \emph{The classical groups, their invariants and representations},
2368: Priceton University Press, 1946.
2369:
2370: \end{thebibliography}
2371:
2372: \end{document}
2373:
2374: