math0108094/a.tex
1: 
2: \documentclass[dvips]{amsart}
3: \usepackage{amssymb,amsxtra,latexsym,graphics,pictex,epsfig}
4: 
5: %\input mac
6: \newcommand{\vanish}[1]{}
7: \newcommand{\F}{\mathcal{F}}
8: \newcommand{\G}{\mathcal{G}}
9: \newcommand{\A}{\mathcal{A}}
10: \newcommand{\B}{\mathcal{B}}
11: \renewcommand{\L}{\mathcal{L}}
12: \newcommand{\M}{\mathcal{M}}
13: \newcommand{\C}{\mathcal{C}}
14: %\newcommand{\D}{\mathcal{D}}
15: \newcommand{\D}{\Delta}
16: \newcommand{\R}{\mathbb{R}}
17: \newcommand{\CC}{\mathbb{C}}
18: \newcommand{\Z}{\mathbb{Z}}
19: \newcommand{\fq}{\mathbb{F}_q}
20: \newcommand{\abs}[1]{\lvert#1\rvert}
21: \newcommand{\norm}[1]{\lVert#1\rVert}
22: \newcommand{\zerohat}{\hat{0}}
23: \newcommand{\onehat}{\hat{1}}
24: \newcommand{\supp}{\operatorname{supp}}
25: \newcommand{\rank}{\operatorname{rank}}
26: \newcommand{\Span}{\operatorname{span}}
27: \newcommand{\dimm}{\operatorname{dim}}
28: \newcommand{\gl}{\operatorname{GL}}
29: \newcommand{\des}{\operatorname{des}}
30: \newcommand{\type}{\operatorname{type}}
31: \newcommand{\End}{\operatorname{End}}
32: \newcommand{\inv}{\operatorname{inv}}
33: \newcommand{\Stab}{\operatorname{Stab}}
34: \newcommand{\onto}{\twoheadrightarrow}
35: \newcommand{\into}{\hookrightarrow}
36: \newcommand{\iso}{\cong}
37: \newcommand{\isoto}{\stackrel{\cong}{\longrightarrow}}
38: \newcommand{\J}{\mathcal{J}}
39: \newcommand{\qbinom}[2]{\genfrac{[}{]}{0pt}{}{#1}{#2}}
40: %\newcommand{\lhor}{\rule[.7ex]{.2in}{.3pt}}%horiz. line for lattice
41: \newcommand{\lhor}{\rule[.7ex]{.2in}{.4pt}}%horiz. line for lattice
42: %\newsavebox{\lhorbox}
43: %\savebox{\lhorbox}{$\relbar\joinrel\relbar\joinrel\relbar$}
44: %\newcommand{\lhor}{\usebox{\lhorbox}}
45: 
46: %\newcommand{\kid}{\text{\textcircled{x}}}
47: %\newcommand{\nokid}{\text{\textcircled{}}}
48: 
49: \theoremstyle{plain}
50: \newtheorem*{corollary}{Corollary}
51: \newtheorem{lemma}{Lemma}
52: \newtheorem{theorem}{Theorem}
53: %\newtheorem*{theorem*}{Theorem}
54: \newtheorem*{main}{Main theorem}
55: \newtheorem{proposition}{Proposition}
56: 
57: \theoremstyle{definition}
58: \newtheorem{definition}{Definition}
59: 
60: \theoremstyle{remark}
61: \newtheorem*{remark}{Remark}
62: \newtheorem{example}{Example}
63: 
64: %%% Abbreviations for references%%%%%%%%%%%
65: \newcommand{\aigner}{aigner97:_combin}
66: \newcommand{\abra}{abramenko94:_walls_coxet}
67: \newcommand{\abels}{abels91}
68: \newcommand{\bjorner}{bjoerner84:_some_coxet_tits}
69: \newcommand{\bjor}{bjoerner80:_shell_cohen_macaul}
70: \newcommand{\bw}{bjoerner97:_shell}
71: \newcommand{\riffle}{bayerdiaconis92:_trail}
72: \newcommand{\bid}{bidigare97:_hyper}
73: \newcommand{\bergeron}{bergeron92:_descent}
74: \newcommand{\bhr}{bidigarehanlon98}
75: \newcommand{\bbd}{billera99:_random}
76: \newcommand{\bl}{billera98:_noncom}
77: \newcommand{\bir}{birkhoff79:_lattic}
78: \newcommand{\red}{bjoerner93:_om}
79: \newcommand{\brady}{brady01}
80: \newcommand{\brown}{brown89:_build}
81: \newcommand{\ken}{brown00:_semig_markov}
82: \newcommand{\bd}{browndiaconis:_random}
83: \newcommand{\borovik}{borovik95:_coxet}
84: \newcommand{\cp}{clifford61}
85: \newcommand{\desar}{desarmenien83:_une}
86: \newcommand{\desarw}{desarmenien93:_descen}
87: \newcommand{\drep}{diaconis88:_group}
88: \newcommand{\dicm}{diaconis98:icm}
89: \newcommand{\df}{diaconis90:_stron}
90: \newcommand{\dfp}{diaconisfill92:_top_to_random}
91: \newcommand{\diacfreed}{diaconis:_iterat}
92: \newcommand{\donnelly}{donnelly91}
93: \newcommand{\Fill}{fill96}
94: \newcommand{\greene}{greene73:_moebius}
95: \newcommand{\grillet}{grillet95:_semig}
96: \newcommand{\gruna}{gruenbaum72:_arran}
97: \newcommand{\gb}{grove85:_finit}
98: \newcommand{\mukherjea}{hoegnaes95:_probab}
99: \newcommand{\muhlherr}{muehlherr90}
100: \newcommand{\humphreys}{humphreys90:_reflec_coxet}
101: \newcommand{\kapoor}{kapoor91:_stoch}
102: \newcommand{\klein}{klein-barmen40:_uber_veral_verban}
103: \newcommand{\ot}{orlikterao92:_book}
104: \newcommand{\petrich}{petrich71}
105: \newcommand{\petrichbook}{petrich77:_lectur}
106: \newcommand{\phatarfod}{phatarfod91:_matrix}
107: \newcommand{\ronan}{ronan89:_lectur}
108: \newcommand{\dress}{dress87:_gated}
109: \newcommand{\scharlau}{scharlau85:_metric}
110: \newcommand{\schar}{scharlau95:_build}
111: \newcommand{\schmid}{muehlherr95:_stein}
112: \newcommand{\schutz}{schuetzenberger47:_sur}
113: \newcommand{\serre}{serre77:_linear}
114: \newcommand{\solomonb}{solomon67:_burns}
115: \newcommand{\solomon}{solomon76:_mackey}
116: \newcommand{\stanleyflag}{stanley96:_combin}
117: \newcommand{\stanley}{stanley97:_enumer1}
118: \newcommand{\stanleysuper}{stanley72:_super}
119: \newcommand{\tits}{tits74:_build_bn}
120: \newcommand{\wachs}{wachs89}
121: \newcommand{\wachswhite}{wachs91:_stirl}
122: \newcommand{\welsh}{welsh76:_matroid}
123: \newcommand{\whitney}{whitney35}
124: \newcommand{\zie}{ziegler95:_lectures}
125: \newcommand{\zas}{zaslavsky75:_facing}
126: \newcommand{\zasgen}{zaslavsky77}
127: 
128: \hyphenation{mul-ti-pli-ci-ties}
129: 
130: 
131: 
132: 
133: 
134: 
135: 
136: 
137: \begin{document}
138: \title{Shuffles on Coxeter groups}
139: \author{Swapneel Mahajan}
140: 
141: \address{Department of Mathematics\\
142: Cornell University\\
143: Ithaca, NY 14853}
144: \email{swapneel@math.cornell.edu}
145: %\date{\today}
146: %\date{September 20, 1999}
147: 
148: %\thanks{Preliminary version; comments and suggestions are welcome.}
149: 
150: \begin{abstract}
151: The random-to-top and the riffle shuffle are two well-studied methods
152: for shuffling a deck of cards. These correspond to the symmetric group
153: $S_n$, i.e., the Coxeter group of type $A_{n-1}$. In this paper, we
154: give analogous shuffles for the Coxeter groups of type $B_n$ and
155: $D_n$. These can be interpreted as shuffles on a ``signed'' deck 
156: of cards. With these examples as motivation, we abstract the notion of a
157: shuffle algebra which captures the connection between the algebraic
158: structure of the shuffles and the geometry of the Coxeter groups.
159: We also briefly discuss the generalisation to buildings which leads to 
160: $q$-analogues.
161: \end{abstract}
162: 
163: \maketitle
164: 
165: %\input s
166: \section{Introduction}
167: 
168: In a recent work, Ken Brown~\cite{\ken} used algebraic methods to
169: analyse random walks on a class of semigroups called ``left-regular
170: bands''. These walks include the hyperplane
171: chamber walks of Bidigare, Hanlon, and Rockmore~\cite{\bhr}. 
172: In this paper, we look at the special case of reflection arrangements
173: that arise in the study of Coxeter groups.
174: The random walks that we look at can be thought of as shuffles on the
175: Coxeter group.
176: The motivating examples are the riffle shuffle and the random-to-top
177: shuffle on a deck of $n$ cards.	
178: These correspond to the symmetric group
179: $S_n$, i.e., the Coxeter group of type $A_{n-1}$. 
180: 
181: The only Coxeter groups that we deal with are the ones of type $A_{n-1}$,
182: $B_n$ and $D_n$. 
183: To make this paper accessible to readers unfamiliar with Coxeter
184: groups we have included two appendices. 
185: Appendix~\ref{app:hyperplane} reviews the facts we need about hyperplane
186: arrangements. In Appendix~\ref{app:Coxeter}, we give a brief review of Coxeter
187: groups and then explain everything in
188: concrete terms for the three cases mentioned above.
189: For the general theory of Coxeter groups, we refer the reader to
190: \cite{\brown,\gb,\humphreys,\tits}. 
191: 
192: \subsection{The random walk and the method of analysis} \label{subs:rw}
193: Let $\Sigma$ be the simplicial (Coxeter) complex associated to a
194: Coxeter group $W$. Also let $\C$ be the set of chambers (or maximal
195: simplices) of $\Sigma$. Then $\Sigma$ is a semigroup containing $\C$
196: as an ideal. The product in $\Sigma$ is given by the projection maps
197: as explained in Appendix~\ref{app:hyperplane}.
198: We now describe the walk.
199: Let $\{w_x\}_{x\in \Sigma}$ be a probability distribution
200: on $\Sigma$. 
201: If the walk is in chamber $c \in \C$, then move to the chamber $xc$,
202: where $x \in \Sigma$ is chosen with probability $w_x$.
203: The product $xc$ is again a chamber because $\C$ is an ideal in $\Sigma$.
204: 
205: Next we describe the walk more algebraically.
206: Fix a commutative ring $k$ and consider the semigroup algebra
207: $k\Sigma$. The $k$-module $k\C$ spanned by the chambers
208: is an ideal in $k\Sigma$.  In particular, it is a module over 
209: $k\Sigma$; we therefore obtain a homomorphism
210: \[
211: k\Sigma\to \End_k(k\C),
212: \]
213: the latter being the ring of $k$-endomorphisms of $k\C$.  This map is
214: in fact an inclusion; so we can regard elements of $k\Sigma$ as
215: operators on $k\C$.  
216: Let $\sigma = \sum_{x\in \Sigma} w_x x$ be such that 
217: $\sum_{x \in \Sigma} w_x = 1$.
218: It is straightforward to check that the transition matrix of the random
219: walk determined by~$\{w_x\}$ is simply the matrix of the operator
220: ``left multiplication by~$\sigma$''.  
221: In other words, it is the image of $\sigma$ under the inclusion
222: $k\Sigma \into \End_k(k\C)$.
223: 
224: A way to analyse the random walk is to focus on $k\Sigma$ and
225: understand the structure
226: of the subalgebra $k[\sigma]$ that $\sigma$ generates in $k\Sigma$.
227: Because of the inclusion map above, we can also analyse the structure
228: of $k[\sigma]$ in $\End_k(k\C)$.
229: We will make use of both viewpoints in our examples.
230: The reader may be more familiar with this algebraic approach in the context of
231: groups~\cite{\drep} rather than semigroups~\cite{\ken}.
232: 
233: \subsection{Nature of our examples} \label{subs:ne}
234: In our examples, the element $\sigma$ will never be normalised. 
235: In other words, if we write $\sigma = \sum_{x\in \Sigma} w_x x$
236: then $\sum_{x \in \Sigma} w_x \not= 1$.
237: So, strictly speaking, we will not have a probability distribution on $\Sigma$.
238: The main reason for doing this is to simplify the algebra.
239: So everytime we describe the random walk associated to $\sigma$,
240: the description will be correct upto a normalisation factor.
241: This will also be true for all other objects that will show up
242: in the analysis.
243: To get rid of this anamoly,
244: we adopt the following convention.
245: If the random walk description says
246: ``Do xxx at random.''
247: then the normalisation factor is
248: ``the number of ways of doing xxx''.
249: In general, the random walk description will involve
250: a sequence of independent random acts in which case
251: the normalisation factor will be the product of the individual factors.
252: In most examples, 
253: we will prefer to first define all objects of interest 
254: as elements of $k\Sigma$
255: and later to provide motivation by interpreting the associated random walks.
256: Making this translation will always involve the normalisation factor
257: defined above.
258: 
259: Next we mention that in our examples, the element $\sigma$ will always
260: lie in the subalgebra of $W$-invariants of $k\Sigma$.
261: We now discuss this subalgebra.
262: The $W$-invariants of $k\Sigma$ under the natural $W$-action form
263: a $k$-algebra $(k\Sigma)^W$.  As a $k$-module, $(k\Sigma)^W$ is free with
264: one basis element for each $W$-orbit in~$\Sigma$, that basis element
265: being the sum of the simplices in the orbit.  Since orbits correspond
266: to types of simplices, we get a basis vector
267: \[
268: \sigma_J= \sum_{F\in\Sigma_J} F
269: \]
270: for each $J\subseteq I$, where
271: $\Sigma_J$ is the set of simplices of
272: type~$J$ and $I$ is
273: the set of all labels or types of vertices of $\Sigma$.
274: Also define $\Sigma_j$ to be the set of simplices of rank $j$.
275: 
276: As already mentioned, in our examples $\sigma \in (k\Sigma)^W$ and
277: hence $k[\sigma]$ is a certain commutative subalgebra 
278: of $(k\Sigma)^W$. Since the element $\sigma$ is always motivated by
279: some shuffle considerations, we will call $k[\sigma]$ a shuffle algebra.  
280: Bidigare \cite{\bid}
281: proved that $(k\Sigma)^W$ is anti-isomorphic to Solomon's descent
282: algebra~\cite{\solomon}, which is a certain subalgebra of the group
283: algebra $kW$.  So 
284: shuffle algebras are anti-subalgebras of Solomon's descent
285: algebra. 
286: 
287: \subsection{Organization of the paper}
288: In Section~\ref{s:mt} we define shuffle algebras and provide a 
289: semi-simplicity criterion for their analysis.
290: In the next three sections,
291: we give examples of shuffle algebras of type
292: $A_{n-1}, B_n$ and $D_n$ respectively.
293: These sections should be read in conjunction with
294: Appendices~\ref{subs:a}, \ref{subs:b} and~\ref{subs:d} respectively.
295: Our examples are motivated by shuffle considerations 
296: on the one hand and the geometry of the Coxeter complex on the other.
297: The striking similarity among the examples can be traced to maps among the 
298: three Coxeter complexes.
299: These maps which give a unified framework for our examples 
300: are explained in Section~\ref{s:maps}.
301: This section gives us a good overall picture
302: of how things fit together and 
303: can play an important role in the further development of the theory.
304: We see an example of this in Section~\ref{s:q}, 
305: where we briefly discuss the generalisation of the 
306: random walks to buildings.
307: 
308: 
309: \section{The main tool} \label{s:mt}
310: 
311: This section is somewhat technical and
312: it is better to read it lightly at present.
313: You may want to go over the details later 
314: with a concrete example in mind.
315: We first abstract the notion of a shuffle algebra
316: and then develop the main tool for the analysis of these algebras.
317: 
318: 
319: \subsection{Shuffle algebras} \label{subs:defn}
320: 
321: We first need a preliminary definition.
322: An additive (resp. multiplicative) \emph{shuffle
323: semigroup} is a subsemigroup of the semigroup of non-negative
324: (resp. positive)
325: integers under addition (resp. multiplication).
326: Note that there is only one additive shuffle semigroup upto isomorphism.
327: However in the multiplicative case, there could be many.
328: Now we are ready to define a shuffle algebra.
329: 
330: We say that a subalgebra $A\subseteq k\Sigma$ is an additive (resp. multiplicative)  \emph{shuffle algebra} if
331: it satisfies the following conditions:
332: 
333: 
334: \begin{enumerate}
335: \item[(1)]
336: $\dimm_kA \leq \dimm \Sigma + 2$. 
337: \item[(2)] 
338: $A$ has a basis of the form $\sigma_0 =
339: 1$,$\sigma_1$,$\sigma_2$,\ldots where $\sigma_j$ is an element of
340: $k\Sigma_j$, the span of simplices of rank $j$. 
341: \item[(3)] 
342: $A = k[\sigma_1]\subseteq k\Sigma$.
343: \item[(4)] 
344: $A$ contains an additive (resp. multiplicative) shuffle
345: semigroup $S$ as a spanning set.
346: \end{enumerate}
347: %
348: Note that the definition of a shuffle algebra $A$ depends on the
349: Coxeter complex $\Sigma$ under consideration. In fact, conditions
350: $(1),(2)$ and $(3)$ say that $A$ is in tune with the geometry of $\Sigma$.
351: Hence it is more correct to write a shuffle algebra as a pair
352: $(A,\Sigma)$.
353: But we will not bother with this since in our examples
354: the Coxeter complex $\Sigma$
355: will always be clear from context.
356: 
357: An alternate way to express condition $(4)$ is to say that
358: there exists an injective semigroup map $S \into A$ 
359: such that the image spans $A$.
360: In all our examples, the semigroup $S$ will have an interpretation 
361: in terms of card shuffles. This explains our terminology.
362: 
363: \begin{remark}
364: The above definition is mainly motivated by the infinite families
365: $A_{n-1}, B_n$ and $D_n$.
366: The significance of this definition for the sporadic Coxeter groups
367: of type $E_6$, $E_7$, etc. is not clear.
368: It is also natural to ask whether shuffle algebras can be classified.
369: We do not know the answer.
370: We would also like to point out that we have imposed the strictest
371: possible conditions that our examples led us to.
372: As a result, some of the algebras we consider are not shuffle
373: algebras;
374: for example, the shuffle double constructions in 
375: Sections~\ref{subs:brs} and~\ref{subs:drs}.
376: Hence a more flexible definition might work better,
377: say for the classification problem. 
378: 
379: \end{remark}
380: 
381: 
382: \subsection{A semisimplicity criterion}
383: 
384: Ken Brown~\cite{\ken} shows that if $k$ has characteristic 0 and $\sigma$ is a
385: non-negative integral linear combination in the canonical basis of
386: $k\Sigma$ consisting of all simplices
387: then $k[\sigma]$ is \emph{split-semisimple}, that is, $k[\sigma]
388: \iso k^n$ for $n=\dimm_kA$. To guarantee this result for an arbitrary
389: field he introduces an additional condition on $\sigma_1$.  In all our
390: examples, we assume that $k$ has characteristic 0. Also $\sigma_1$
391: will always be a non-negative integral linear combination in the
392: canonical basis of $k\Sigma$. Hence it follows by condition $(3)$ that
393: the shuffle algebras we consider are split-semisimple. We will not
394: rely on Brown's result however.  Instead we will use a semisimplicity
395: criterion which we now state.
396: 
397: \begin{proposition} \label{p:ss}
398: Let $A$ be an $n$-dim algebra containing a spanning semigroup $S$.
399: %Let $dim_kA = n$. 
400: Suppose that there exist characters $\chi_1$,\ldots,$\chi_n$ of $S$
401: and elements $e_1$,\ldots,$e_n$ of $A$ such that for every $s \in S$,
402: we have
403: \begin{equation} \tag{*}
404: s = \sum_{i=1}^n \chi_i(s) e_i.
405: \end{equation}
406: Then $A \isoto k^n$ with $e_i$ as the primitive idempotents.
407: Also the $\chi_i$'s extend to characters of $A$.
408: \end{proposition}
409: \begin{proof}
410: The elements $e_1$,\ldots,$e_n$ of $A$ span $S$ which in turn
411: spans $A$. 
412: Since $\dimm_kA = n$ this implies that $e_1$,\ldots,$e_n$ form a basis for $A$. 
413: This yields a vector space isomorphism $\Phi : A \isoto k^n$, where
414: $\sum_{i=1}^n a_i e_i \mapsto (a_1,\ldots,a_n)$.
415: Hence for $s \in S$,
416: $s \mapsto (\chi_1(s),\ldots,\chi_n(s))$.
417: Since $\chi_1,\ldots,\chi_n$ are characters of $S$, we have
418: $\Phi(s_1 s_2) = \Phi(s_1) \Phi(s_2)$ for $s_1,s_2 \in S$.
419: The fact that $\Phi$ respects the algebra structures on a spanning set
420: $S$ of $A$ implies that it is in fact an algebra homomorphism.
421: \end{proof}
422: 
423: As mentioned before, this proposition will apply to all our
424: examples. The algebra $A$ will be a shuffle algebra and $S$ the
425: shuffle semigroup contained in $A$. 
426: The characters $\chi$ of $S$ that we will use are
427: $\chi(a) = c^a$ if $S$ is additive and 
428: $\chi(a) = a^c$ if $S$ is multiplicative. 
429: Here $c$ is a fixed non-negative integer.
430: The specific values of $c$ that we need to choose 
431: depend on the example at hand.
432: The advantage of Proposition~\ref{p:ss}
433: is that it is tailor-made for shuffle algebras
434: and it gives an explicit isomorphism of $A$ with $k^n$.
435: 
436: \begin{remark}
437: The Coxeter complexes $\Sigma$ of type $A_{n-1}, B_n$ and $D_n$ can be
438: described using the language of ``ordered partitions''. 
439: Also, chambers of $\Sigma$ can be described as a ``deck of cards''. 
440: This is explained in Appendices~\ref{subs:a}, \ref{subs:b} and~\ref{subs:d} 
441: respectively.
442: We follow the convention that ``left to right'' for an ordered
443: partition is ``top to bottom'' for a deck of cards.
444: It is essential to understand this language before reading the examples.
445: The side shuffle and riffle shuffle of type $A_{n-1}$ considered in
446: Sections~\ref{subs:ass} and~\ref{subs:ars} are the motivating
447: examples of this paper. It is also a good idea to first understand these two
448: basic examples before proceeding to the rest.
449: \end{remark}
450: 
451: \section{Examples of type $A_{n-1}$}
452: 
453: We consider three examples; the side shuffle, the two sided shuffle
454: and the riffle shuffle.
455: The first two are additive in nature while the third is
456: multiplicative.
457: The analysis for the first and third example parallels that of
458: Diaconis, Fill and Pitman~\cite{\dfp}
459: and Bayer and Diaconis~\cite{\riffle} 
460: respectively.
461: The difference is that we prefer to work with $(k\Sigma)^W$,
462: which is the more geometric description of Solomon's descent algebra.
463: Elements of the label set $I$ will be written as $s_1, s_2, \ldots, s_{n-1}$;
464: see Figure 2.1 in Appendix~\ref{subs:cd}.
465: Recall that $\sigma_J$, the sum of simplices of type~$J$, for $J \subseteq I$
466: form a basis for $(k\Sigma)^W$. Hence these elements will play a key
467: role in the analysis.
468: 
469: 
470: \subsection{The side shuffle}\label{subs:ass}
471: This is more commonly known as the random-to-top shuffle or the
472: Tsetlin library.
473: The element of interest is $\sigma = \sigma_{s_1}$, that is, it is the
474: sum of all vertices of type $s_1$.
475: In terms of ordered partitions, $\sigma_{s_1}$ is the sum of all
476: ordered two block partitions of $[n]$ such that the first block is a
477: singleton.
478: As explained in Section~\ref{subs:rw}, 
479: there is a random walk on a deck of $n$ cards
480: associated to $\sigma$.
481: It consists of removing a
482: card at random and replacing it on top. 
483: When $n=4$, for example, 
484: $(\{2\}, \{1,3,4\})$ is a typical summand of $\sigma$. Its product with
485: the deck $(\{1\}$,$\{2\}$,$\{3\}$,$\{4\})$ gives
486: $(\{2\}$,$\{1\}$,$\{3\}$,$\{4\})$, that is, the overall effect is to
487: remove the card labelled $2$ and put it on top.
488: Note that the normalisation factor in this case is $n$,
489: consistent with the convention of Section~\ref{subs:ne}.
490: 
491: In addition to $\sigma = \sigma_1$,
492: we define $\sigma_j = \sigma_{J_j}$ where $J_j = \{s_1,\ldots,s_j\}
493: \subseteq I$ for $j = 1,\ldots,n-1$. 
494: To explain in words,
495: $\sigma_1$ is the sum of all the vertices of type $s_1$,
496: $\sigma_2$ is the sum of all the edges of type $s_1 s_2$ and so on till
497: $\sigma_{n-1}$ which is the sum of all chambers of $\Sigma$.
498: As ordered partitions, $\sigma_{j}$ is the sum of all
499: ordered $(j+1)$ block partitions of $[n]$ such that the first $j$
500: blocks are singletons.
501: Also for convenience, we define $\sigma_n = \sigma_{n-1}$. 
502: Just like $\sigma_1$, 
503: we can describe 
504: the random walk on a deck of $n$ cards
505: associated to $\sigma_j$.
506: 
507: $\sigma_j :$ Choose $j$ distinct cards at random and put them on top in a
508: random order. 
509: 
510: \noindent Note that $\sigma_n$ was previously defined
511: in an artificial way.
512: But the random walk description for $\sigma_n$ makes perfect sense
513: and we now see the motivation in setting $\sigma_n = \sigma_{n-1}$. 
514: The two are identical as random walks and by our convention
515: have the same normalisation factor of $n!$.
516: From now on, we will leave out the discussion on normalisation.
517: We now present two methods to
518: analyse $k[\sigma_1]$.
519: 
520: \medskip
521: \noindent
522: {\bf The shuffle method}.
523: 
524: Put $\sigma_0 = 1$.
525: Define $a$-shuffles $S_a$ for $a \geq 0$ by 
526: \begin{equation} \label{e:shuffle}
527: S_a = \sum_{j=0}^n S(a,j) \sigma_j,
528: \end{equation}
529: where $S(a,j)$ are the Stirling numbers of the second kind.
530: They count the number of ways in which $a$ elements can be divided
531: into $j$ non-empty subsets. 
532: We make the convention that 
533: $S(a,j) = 0$ for $a < j$. Also
534: $S(0,0) = 1$ and $S(a,0) = 0$ for $a > 0$.
535: Note that
536: $S_0 = 1$ and $S_1 = \sigma_1$.
537: The motivation behind this definition will soon become clear.
538: Unlike for $\sigma_j$,
539: the description of $S_a$ in terms of ordered partitions is not
540: clear. However, the associated random walk on a deck of $n$ cards 
541: can be readily described.
542: 
543: $S_a :$ Pick a card at random and mark it. Repeat this process $a$
544: times. There is no restriction on the number of times a given card may get
545: marked. Move all the marked cards to the top in a random order.
546: 
547: To see the equivalence of the two definitions, note
548: that $S_a$ can be expressed as a sum indexed by the number
549: of distinct cards that get marked. 
550: Suppose that $j$ distinct cards get marked. By our marking scheme for
551: $S_a$, there are exactly $S(a,j)$ ways in which the same set of $j$
552: cards gets marked. This leads to equation~\eqref{e:shuffle}.
553: With the probabilistic interpretation for $S_a$, one can also convince
554: oneself that $S_a S_b = S_{a+b}$.
555: This relation was the main motivation behind the definition of $S_a$.
556: Thus we get an additive shuffle semigroup $S = \{S_a : a \geq 0 \}$
557: contained in $A = k[\sigma_1] = k[S_1]$.
558: The additivity says that the semigroup $S$ spans $A$.
559: We can deduce from equation~\eqref{e:shuffle} that
560: $\sigma_0=1$,$\sigma_1$,\ldots,$\sigma_{n-1}$ lie in $A$. 
561: Also they span $S$ and hence $A$.
562: Since we know that they are linearly independent in $(k\Sigma)^W$, 
563: it now follows that they form a basis for $A$. 
564: Hence $A$ is a shuffle algebra.
565: 
566: The analysis so far shows that the element $\sigma_1$ determines what 
567: $\sigma_2,\ldots,\sigma_{n-1}$ should be, assuming that we want
568: $\sigma_j \in \Sigma_j$. We will see this more
569: directly in the second method. Now we show that $A \iso k^n$.
570: We do this by moulding equation~\eqref{e:shuffle} in the shape of
571: equation~(*) of Proposition~\ref{p:ss}. 
572: The key step is to use the following explicit formula for the Stirling
573: numbers $S(a,j)$; see~\cite[pg 34]{\stanley}.
574: \begin{equation*} %\label{e:tsetlinu}
575: S(a,j) = \sum_{i=0}^j
576: (-1)^{j-i}\binom{j}{i}\frac{i^a}{j!},
577: \end{equation*}
578: with the convention that $0^0=1$.
579: Substituting the above expression in the formula for $S_a$ and
580: rearranging terms, we get for $a \geq 0$
581: \begin{equation*} 
582: S_a=\sum_{i=0}^n i^a e_i, \quad \text{where} \quad
583: e_i=\sum_{j=i}^n(-1)^{j-i}\binom{j}{i}\frac{\sigma_j}{j!}\,. 
584: \end{equation*}
585: Observe that $e_{n-1} = 0$ because $\sigma_{n-1}=\sigma_{n}$.
586: The remaining $e_i$'s are clearly non-zero.
587: Now apply Proposition~\ref{p:ss} to $S$ and $A$ along with the $n$
588: characters of $S$ given by $\chi_i(S_a) = i^a$  for $i=0,1,2,\ldots,n-2,n$.
589: This gives
590: $A \isoto k^{n}$. 
591: The isomorphism maps $S_a$ to $(0^a,1^a,2^a,\ldots,(n-2)^a,n^a)$.
592: These give the possible set of eigenvalues of $S_a$ considered as an
593: operator on any $A$-module.  
594: The formulas for the $e_i$'s are identical to those obtained in~\cite{\ken}.
595: The minimal polynomial for $S_a$ is given by
596: $x(x-1)(x-2^a)\cdots(x-(n-2)^a)(x-n^a)$.
597: If we expand this polynomial, substitute $x=S_a$ and use
598: equation~\eqref{e:shuffle} then we get some identities involving
599: the Stirling numbers.
600: \vanish
601: {
602: Now set
603: $S' = \{S_a : a > 0 \}$ and
604: $A'$ to be the $k$-span $\sigma_1$,\ldots,$\sigma_{n-1}$.
605: The proposition applied to $S'$ and $A'$ along with the $n-1$
606: characters $\chi_i(S_a) = i^a$  for $i=1,2,\ldots,n-2,n$ of $S'$ gives
607: $A' \isoto k^{n-1}$. 
608: 
609: }
610: %\end{vanish}
611: 
612: \medskip
613: \noindent
614: {\bf The direct method}.
615: 
616: We first claim that
617: $\sigma_j \sigma_1 = j \sigma_j + \sigma_{j+1}$ if $j < n-1$ and 
618: $\sigma_{n-1} \sigma_1 = n \sigma_{n-1}$.
619: We see this from the description of the $\sigma$'s as ordered
620: partitions as follows. A typical summand of $\sigma_j$ is
621: $(\{1\}$,$\{2\}$,$\cdots$,$\{j\}$,$\{j+1 \cdots n\})$.
622: This term appears $j$ times in the product $\sigma_j \sigma_1$;
623: the $j$ summands of $\sigma_1$ that contribute
624: being the ones in which the element in the singleton block 
625: is one of $1,2, \ldots,j$.
626: This yields $j \sigma_j$. The remaining terms yield $\sigma_{j+1}$ and
627: the claim follows. For $j=1$, the claim says that $\sigma_1^2 = \sigma_1 +
628: \sigma_{2}$. Thus $\sigma_2 = \sigma_1^2 - \sigma_1$
629: is determined by $\sigma_1$ and the same is true for $\sigma_3$ and so on. 
630: More precisely, by induction we get
631: \begin{equation}\label{e:cob}
632: \sigma_{j+1} = \sigma_1(\sigma_1-1)\ldots(\sigma_1-j).
633: \end{equation}
634: This along with
635: $\sigma_{n-1} \sigma_1 = n \sigma_{n-1}$ implies that both
636: $\sigma_0=1$,$\sigma_1$,\ldots,$\sigma_{n-1}$ and
637: $\sigma_1^0$,$\sigma_1^1$,\ldots,$\sigma_1^{n-1}$
638: form a basis for $A$.
639: The basic relation satisfied by $\sigma_1$ is 
640: $\sigma_{n-1}(\sigma_1-n) = 0$, that is,
641: $\sigma_1(\sigma_1-1)\ldots(\sigma_1-n+2)(\sigma_1-n) = 0$.
642: Hence
643: $A \isoto \frac{k[x]}{x(x-1)\ldots(x-n+2)(x-n)}$,
644: where the map sends $\sigma_1$ to $x$.
645: This shows that $A$ is split semisimple.
646: Note that this time we did not use Proposition~\ref{p:ss}
647: to arrive at this conclusion.
648: 
649: We now give algebraic motivation for equation~\eqref{e:shuffle}
650: by rederiving it here.
651: Right now $\sigma_j$ is defined only for $0 \leq
652: j \leq n-1$.
653: But we may extend the definition of $\sigma_j$ to any $j$
654: using equation~\eqref{e:cob}. 
655: With this extension, one may check that 
656: $\sigma_n = \sigma_{n-1}$ and $\sigma_j = 0$ for
657: $j > n$.
658: We now find a formula for $\sigma_1^j$ by inverting
659: equation~\eqref{e:cob} formally.
660: Write 
661: \begin{equation*}
662: \sigma_1^a = \sum_{j=0}^a S(a,j) \sigma_j
663: \end{equation*}
664: for some constants $S(a,j)$.
665: Multiply both sides on the right by $\sigma_1$ and use the relation
666: $\sigma_j \sigma_1 = j \sigma_j + \sigma_{j+1}$ to obtain the
667: recursion:
668: $S(a,j) = j S(a-1,j) + S(a-1,j-1)$ with $S(a,a) = S(a,1) = 1$.
669: %Since $\sigma_n = \sigma_{n-1}$ and $\sigma_j = 0$ for $j > n$,
670: %the inversion above is not unique. But it is the most natural one to
671: %consider. 
672: The recursion and the initial conditions show that 
673: $S(a,j)$ are the Stirling numbers of the second kind.
674: We now define the $a$-shuffle $S_a$ to be $\sigma_1^a$.
675: This gives equation~\eqref{e:shuffle}.
676: By our definition, it follows directly that $S_a S_b = S_{a+b}$.
677: Note that this time we did not rely on any 
678: probabilistic interpretation of $S_a$
679: to derive this additive relation.
680: 
681: \begin{remark}
682: Consider the map $k[x] \onto A$ which sends $x$ to $\sigma_1$.
683: Along with $\{x^j\}_{j \geq 0}$,
684: the other sequence which played a prominent role in our analysis was
685: $\{x_{(j)}\}_{j \geq 0}$, where
686: $x_{(j+1)} = x(x-1)\ldots(x-j)$.
687: Both these sequences are polynomial sequences of binomial type~\cite{\aigner}. 
688: This kind of structure seems to be common to all the additive examples.
689: The first sequence is always the same but the second sequence varies.
690: And the relation between the two gives us various analogues of 
691: the Stirling numbers.
692: It would be worthwhile to understand this better.
693: 
694: 
695: It is possible to write down an explicit formula for $\sigma_i
696: \sigma_j$. We state it here for completeness.
697: $$\sigma_i\sigma_j=\sum_{k=0}^{min(i,j)}k!\binom{i}{k}\binom{j}{k}
698: \sigma_{i+j-k}.$$
699: There is also a $q$-analogue of the side shuffle which we will
700: explain in Section~\ref{s:q}.
701: \end{remark}
702: 
703: 
704: \subsection{The two sided shuffle} \label{subs:atss}
705: This example is similar to the side shuffle and it will be useful to
706: keep the analogy in mind.
707: The element of interest is $\sigma = \sigma_{s_1} + \sigma_{s_{n-1}}$. 
708: In terms of ordered partitions, $\sigma$ is the sum of all
709: ordered two block partitions of $[n]$ such that either the first block
710: or the second block is a singleton.  
711: The associated random walk on a deck of cards consists of removing a
712: card at random and replacing it either on top or at the bottom; whence
713: the name. 
714: 
715: In addition to $\sigma = \sigma_1$,
716: we define $\sigma_j = \sum_{k=0}^j \binom{j}{k}\sigma_{J_{j,k}}$, where
717: $J_{j,k} = \{s_1,\ldots,s_k\} \cup \{s_{n-(j-k)},\ldots,s_{n-1}\}
718: \subseteq I$ for $j = 1,\ldots,n-1$. 
719: Note that $J_{j,k}$ is the union of the first $k$ and last $j-k$
720: elements of the label set. Hence the cardinality of $J_{j,k}$ is always $j$.
721: Also as ordered partitions, $\sigma_{J_{j,k}}$ is the sum of all
722: ordered $(j+1)$ block partitions of $[n]$ such that the first $k$ and
723: the last $j-k$ blocks are singletons.
724: As is evident from the formula, expressing $\sigma_j$ in this language
725: is not pleasant. However the probability description will be simple as
726: we will see.
727: Also put $\sigma_0 = 1$ and
728: $\sigma_n = 2 \sigma_{n-1}$.
729: Now define $a$-shuffles $S_a$ for $a \geq 0$ by 
730: \begin{equation} \label{e:shuffle2}
731: S_a = \sum_{j=0}^n S(a,j) \sigma_j,
732: \end{equation}
733: where $S(a,j)$ counts the number of ways in which a set of $a$ elements can be divided
734: into $j$ non-empty subsets, where in each subset the elements are further
735: divided into two subsets. 
736: We will call the $S(a,j)$'s the \emph{signed} Stirling numbers.
737: They satisfy the recursion:
738: $S(a,j) = 2j S(a-1,j) + S(a-1,j-1)$ with $S(a,a) = 1$ and $S(a,1) =
739: 2^{a-1}$. An explicit formula is as under:
740: \begin{equation*} %\label{e:tsetlinu}
741: S(a,j) = \sum_{i=0}^j
742: (-1)^{j-i}\binom{j}{i}\frac{(2i)^a}{2^j j!}\,.
743: \end{equation*}
744: Note that
745: $S_0 = 1$ and $S_1 = \sigma_1$. 
746: These definitions are motivated by the following interpretations 
747: of the random walks associated to $\sigma_j$ and $S_a$.
748: 
749: $\sigma_j :$ Choose $j$ cards at random. Split these into two distinct
750: piles such that every split is equally likely. Move the cards in the
751: first pile to the top in a random order and those in the second
752: pile to the bottom in a random order. 
753: 
754: $S_a :$ Pick a card at random and mark it $T$ or $B$. Repeat this process $a$
755: times. If at any time we choose a card that has already been marked
756: then we overwrite that mark. Now move all cards marked $T$ to the top
757: and those marked $B$ to the bottom in a random order.
758: 
759: First check that $\sigma_n = 2 \sigma_{n-1}$.
760: Next note that $\sigma_j$ can be expressed as a sum depending on the
761: number of cards $k$ that are moved to the top.
762: There are $\binom{j}{k}$ ways for this to occur.
763: This leads us to the equation for $\sigma_j$ in terms of the
764: $\sigma_{J_{j,k}}$'s that we wrote earlier.
765: The interpretation of $S_a$ is slightly subtle. 
766: As for the side shuffle, express $S_a$ as a sum indexed by the number
767: of distinct cards that get marked. 
768: Suppose that $j$ distinct cards get marked. By our marking scheme for
769: $S_a$, there are exactly $2 S(a,j)$ ways in which the same set of $j$
770: cards gets marked. This is because we use distinct labels $T$
771: and $B$. However, this extra factor is compensated in the next step
772: where we use the last label on a marked card to decide whether it
773: goes to the top or the bottom. 
774: 
775: Now observe that $S_a S_b = S_{a+b}$.
776: This gives us an additive shuffle semigroup $S = \{S_a : a \geq 0 \}$
777: contained in $A = k[\sigma_1] = k[S_1]$.
778: We repeat the analysis for the side-shuffle to show that 
779: $A \isoto k^{n}$, where $S_a$ maps to $(0^a,2^a,4^a,\ldots,(2n-4)^a,(2n)^a)$. 
780: The formula for the primitive idempotents is given by
781: \begin{equation*} %\label{e:tsetlinu}
782: e_i = \sum_{j=i}^n(-1)^{j-i}\binom{j}{i}\frac{\sigma_j}{2^j j!}\,.
783: \end{equation*}
784: This time the relation $\sigma_{n}= 2 \sigma_{n-1}$ forces $e_{n-1} = 0$.
785: Also equation~\eqref{e:shuffle2} can be rewritten as
786: \[
787: S_a=\sum_{i=0}^n (2i)^a e_i.
788: \]
789: As for the side-shuffle, the direct method also works. Using the
790: description of the $\sigma$'s as ordered partitions, 
791: it is easy to see that
792: $\sigma_j \sigma_1 = 2j \sigma_j + \sigma_{j+1}$ if $j < n-1$ and 
793: $\sigma_{n-1} \sigma_1 = 2n \sigma_{n-1}$.
794: By induction we get
795: \begin{equation}\label{e:cob2}
796: \sigma_{j+1} = \sigma_1(\sigma_1-2)\ldots(\sigma_1-2j).
797: \end{equation}
798: This along with
799: $\sigma_{n-1} \sigma_1 = 2n \sigma_{n-1}$ implies that both
800: $\sigma_0=1$,$\sigma_1$,\ldots,$\sigma_{n-1}$ and
801: $\sigma_1^0=1$,$\sigma_1^1$,\ldots,$\sigma_1^{n-1}$
802: form a basis for $A$.
803: The basic relation satisfied by $\sigma_1$ is 
804: $\sigma_1(\sigma_1-2)\ldots(\sigma_1-2n+4)(\sigma_1-2n) = 0$.
805: Hence
806: $A \isoto \frac{k[x]}{x(x-2)\ldots(x-2n+4)(x-2n)}$.
807: This shows that $A$ is split semisimple.
808: Inverting equation~\eqref{e:cob2} formally and using the relation
809: $\sigma_j \sigma_1 = 2j \sigma_j + \sigma_{j+1}$ 
810: leads to the recursion of the signed Stirling numbers written earlier
811: and hence to equation~\eqref{e:shuffle2}.
812: 
813: \subsection{Riffle shuffle} \label{subs:ars}
814: A {\it riffle shuffle} is a common method people use for shuffling a
815: deck of cards. The Gilbert-Shannon-Reeds mathematical model for riffle
816: shuffling consists of first cutting the deck into two parts, making
817: the cut according to a binomial probability distribution, and then
818: interleaving the two parts in such a way that every interleaving is
819: equally likely. A natural generalisation is the $a$-shuffle in which
820: we cut the deck into $a$ parts (rather than just 2) and then interleave
821: them.
822: 
823: From our point of view, it is more natural to consider
824: the inverse riffle shuffle $S_2$ and more generally the inverse
825: $a$-shuffle $S_a$. It has the following description as a random walk on
826: a deck of $n$ cards. 
827: 
828: $S_a$ : Label each card randomly with an integer from $1$ to
829: $a$. Move all the cards labelled $1$ to the bottom of the deck,
830: preserving their relative order. Next move all the cards labelled $2$
831: above these again preserving their relative order and so on.
832: 
833: We can understand the inverse $a$-shuffle $S_a$ as an element of
834: $(k\Sigma)^W$. We begin with the element $\sigma_1 =
835: \sigma_{s_1}+\cdots+\sigma_{s_{n-1}}$; that is,
836: $\sigma_1$ is the sum of all the vertices in $\Sigma$.  In terms of
837: ordered partitions, it is the sum of all ordered two block partitions
838: of $[n]$.  This is closely related to the inverse riffle shuffle, in
839: fact, $S_2 = \sigma_1 + 2 \sigma_0$, where $\sigma_0 = 1$ is the one
840: block partition. We will show
841: this in more generality.
842: 
843: Let $\sigma_i$ be the sum of all the simplices of rank $i$ in $\Sigma$
844: for $i = 1,\ldots,n-1$. Then
845: define shuffles $S_a$ for $a \geq 1$ by
846: \begin{equation} \label{e:ars}
847: S_a = \sum_{j=1}^n \binom{a}{j} \sigma_{j-1}.
848: \end{equation}
849: In terms of ordered partitions, $\sigma_{j-1}$ is the sum of all
850: ordered $j$ block partitions
851: of $[n]$. To express $S_a$ in this language, we need the notion of a
852: weakly ordered 
853: partition of $[n]$. It is an ordered partition of $[n]$ in which the
854: blocks are allowed to be empty. Note that there are exactly 
855: $\binom{a}{j}$ weakly ordered $a$ block partitions that give the same
856: ordered $j$ block partition of $[n]$. 
857: It follows that $S_a$ is the sum of all weakly
858: ordered $a$ block partitions
859: of $[n]$. 
860: If we assign labels $1,2,\ldots,a$ to these weakly
861: ordered $a$ blocks such that the leftmost block is labelled $a$ and
862: the rightmost is labelled $1$, then
863: we get the random walk description for $S_a$ that we started with. 
864: It is clear from both viewpoints that 
865: $S_a S_b = S_{ab}$. 
866: This gives us a multiplicative shuffle semigroup $S = \{S_a : a \geq 1 \}$
867: contained in $A = k[\sigma_1] = k[S_2]$.
868: It now follows that 
869: $\sigma_0=1$,$\sigma_1$,\ldots,$\sigma_{n-1}$ is a basis for $A$ 
870: and that $A$ is a shuffle algebra.
871: 
872: Now we show that $A \iso k^n$.
873: Expand each $\binom{a}{j}$ as a polynomial in $a$ (with no constant
874: term):
875: \begin{equation*} %\label{e:tsetlinu}
876: \binom {a}{j} = \frac{a(a-1)\ldots(a-j+1)}{j!} =
877: \frac{1}{j!}\sum_{i=1}^j c_{ij} a^i\,
878: \end{equation*}
879: for some constants $c_{ij}$.
880: To be precise, $c_{ij}$ is the coefficient of $x^i$ in the polynomial
881: $x(x-1)\cdots(x-j+1)$.
882: Substituting the above expression in the formula for $S_a$ and
883: rearranging terms, we get for $a \geq 1$
884: \begin{equation*} 
885: S_a=\sum_{i=1}^n a^i e_i, \quad \text{where} \quad e_i=\sum_{j=i}^n
886: c_{ij}\frac{\sigma_{j-1}}{j!}\,. 
887: \end{equation*}
888: All the $e_i$'s are clearly non-zero.
889: Now apply Proposition~\ref{p:ss} to $S$ and $A$ along with the $n$
890: characters of $S$ given by $\chi_i(S_a) = a^i$  for $i=1,2,\ldots,n$.
891: This gives
892: $A \isoto k^{n}$. 
893: The isomorphism maps $S_a$ to $(a^1,a^2,\ldots,a^n)$.
894: These give the possible set of eigenvalues of $S_a$ considered as an
895: operator on any $A$-module.  
896: They are distinct for $a \neq 1$ and hence $S_a$ generates $A$ for all
897: $a \geq 2$. Also the minimal polynomial for $S_a$ on any faithful
898: $A$-module is given by $(x-a)(x-a^2)\cdots(x-a^n)$.
899: Write it as 
900: $\sum_{i=0}^n P_{n,i} x^i$.
901: Now if we substitute $x=S_a$ and use equation~\eqref{e:ars} then we
902: get identities involving binomial coefficients, namely
903: $\sum_{i=0}^n P_{n,i} \binom{a^i}{j} = 0$ for $j=1,2,\dots,n$.
904: It would be nice to have an explicit formula for the coefficients
905: $P_{n,i}$ also. 
906: 
907: The analysis that we gave for this example 
908: was the analogue of the shuffle method
909: of the previous two examples.
910: The direct computational method of those examples is not feasible here.
911: Writing a formula for $\sigma_j \sigma_1$ is not easy because
912: of the multiplicative nature of the example.
913: 
914: \section{Examples of type $B_n$}
915: 
916: We consider two examples, the side shuffle and the riffle shuffle.
917: They should not be regarded as mere generalisations of the ones
918: discussed so far.
919: In fact, we will see in Section~\ref{s:maps} that 
920: they are in a sense more fundamental and 
921: have a right to exist on their own.
922: In this section, the term partition always refers to a partition of
923: type $B_n$; see Appendix~\ref{subs:b}. 
924: Also elements of the label set $I$ will be written as $s_1, s_2, \ldots, s_{n}=t$;
925: see Figure 2.2 in Appendix~\ref{subs:cd}.
926: 
927: \subsection{The side shuffle}
928: The element of interest is $\sigma = \sigma_{s_1}$. 
929: In terms of partitions, $\sigma_{s_1}$ is the sum of all
930: three block partitions such that the first (and hence the last)
931: block is a 
932: singleton.  
933: The associated random walk on a deck of $n$ signed cards 
934: consists of removing a
935: card at random and replacing it on top with either the same sign or
936: its reverse both with equal probability. 
937: Note that the normalisation factor is $2n$.
938: 
939: In addition to $\sigma = \sigma_1$,
940: we define $\sigma_j = \sigma_{J_j}$ where $J_j = \{s_1,\ldots,s_j\}
941: \subseteq I$ for $j = 1,\ldots,n$. 
942: To explain in words,
943: $\sigma_1$ is the sum of all the vertices of type $s_1$,
944: $\sigma_2$ is the sum of all the edges of type $s_1 s_2$ and so on till
945: $\sigma_{n}$ which is the sum of all chambers of $\Sigma$.
946: As partitions, $\sigma_{j}$ is the sum of all
947: $(2j+1)$ block partitions such that the first (and hence the last) $j$
948: blocks are singletons.
949: Define $a$-shuffles for $a \geq 0$ by 
950: \begin{equation*} 
951: S_a = \sum_{j=0}^n S(a,j) \sigma_j,
952: \end{equation*}
953: where $S(a,j)$ are the signed Stirling numbers defined in
954: Section~\ref{subs:atss}.
955: The associated random walks on a deck of $n$ signed cards 
956: are as follows:
957: 
958: $\sigma_j :$ Choose $j$ cards at random. Move them to the top in a
959: random order. Now with equal probability either flip or do not flip
960: the sign of the chosen cards.  
961: 
962: $S_a :$ Pick a card at random and with equal probability either flip
963: or do not flip its sign. Repeat this process $a$ times. There is no
964: restriction on the number of times the same card may get picked. Move
965: all the picked cards to the top in a random order.
966: 
967: The interpretation of $S_a$ involves the same kind of subtlety as the
968: two sided shuffle of type $A_{n-1}$ discussed in Section~\ref{subs:atss}.
969: We observe that $S_a S_b = S_{a+b}$. After this,
970: the analysis works exactly like the two sided shuffle for type
971: $A_{n-1}$. The only difference is that $e_{n-1} \neq 0$ since we do
972: not have any relation like $\sigma_{n} = 2 \sigma_{n-1}$.
973: This gives $A \isoto k^{n+1}$, where $S_a$ maps to $(0^a,2^a,4^a,\ldots,(2n-2)^a,(2n)^a)$. 
974: 
975: Like the additive examples of type $A_{n-1}$, 
976: we can also use the direct method here.
977: Using the partition description of the $\sigma$'s,
978: it is easy to see that
979: $\sigma_j \sigma_1 = 2j \sigma_j + \sigma_{j+1}$ if $j \leq n-1$ and 
980: $\sigma_{n} \sigma_1 = 2n \sigma_{n}$ and the rest is similar.
981: 
982: \subsection{The riffle shuffle}\label{subs:brs}
983: This example is motivated by the riffle shuffle of type $A_{n-1}$.  In
984: this case, the shuffles $S_a$ split into two cases depending on the
985: parity of $a$. They can be described using the inverse $a$-shuffles
986: of Section~\ref{subs:ars} and a special shuffle $S_2$.
987: We will call $S_2$ the inverse {\it
988: signed} riffle shuffle or the inverse riffle shuffle of type $B_n$.
989: 
990: $S_2$: For every card, we either flip or do not flip its sign with equal
991: probability. The cards with unchanged signs move to the top 
992: in the same relative order and the
993: rest move to the bottom in the reverse relative order. 
994: 
995: $S_{2a}$: We do a usual inverse $a$-shuffle with labels
996: $1,2,\ldots,a$. Then within each of the $a$ blocks with a fixed label 
997: we do an inverse signed riffle shuffle $S_2$.
998: 
999: $S_{2a+1}$: We do a usual inverse $(a+1)$-shuffle with labels
1000: $0,1,2,\ldots,a$. Then we do an inverse signed riffle shuffle $S_2$ 
1001: on each block except the one labelled 0.
1002: 
1003: These shuffles are multiplicative. 
1004: One can check directly that $S_{a} S_{b} = S_{ab}$ for any $a,b \geq
1005: 1$ irrespective of parity. However, we will derive this relation by
1006: considering the shuffles $S_a$ as elements of
1007: $(k\Sigma)^W$. It is natural to first analyse the even and odd shuffles
1008: separately and then to put them together later.
1009: 
1010: \medskip
1011: \noindent
1012: {\bf The even part}.
1013: 
1014: The element of interest is $\sigma_1 = \sigma_{t}$. 
1015: In terms of partitions, $\sigma_{t}$ is the sum of all
1016: three block partitions such that the second (zero)
1017: block is  empty.
1018: 
1019: If we try to analyse $\sigma_1^2$, $\sigma_1^3$ and so on then we
1020: observe that they involve only those partitions that have an
1021: empty zero block. In geometric language, they involve only those faces
1022: whose type contains the letter $t$. This motivates the definition of
1023: $\sigma_j$ which we now give. Put $\sigma_j = \sum_{\abs{J}=j,t \in J}
1024: \sigma_J$ for $j=1,2,\ldots,n$.
1025: In terms of partitions, $\sigma_{j}$ is the sum of all
1026: $(2j+1)$ block partitions such that the zero
1027: block is  empty.
1028: In the spirit of the riffle shuffle of type $A_{n-1}$, we define shuffles
1029: $S_{2a}$ for $a \geq 1$ by
1030: \begin{equation}\label{e:ershuffle}
1031: S_{2a} = \sum_{j=1}^n \binom{a}{j} \sigma_{j}.
1032: \end{equation}
1033: Note that $S_2 = \sigma_1$.
1034: Also put $S_1 = \sigma_0 = 1$.
1035: To express $S_{2a}$ in words, we use the notion of a weak 
1036: partition that we defined in Appendix~\ref{subs:b}. It is a partition
1037: in which the 
1038: signed blocks are also allowed to be empty. 
1039: Since the partitions are anti-symmetric, there are exactly 
1040: $\binom{a}{j}$ weak $(2a+1)$ block partitions that give the same
1041: $(2j+1)$ block partition. 
1042: It follows that $S_{2a}$ is the sum of all weak
1043: $(2a+1)$ block partitions
1044: such that the zero block is always empty; so we have
1045: upto $2a$ non-empty blocks.
1046: This explains the term ``$2a$-shuffle''. 
1047: If we assign labels $1,2,\ldots,a$ in decreasing order to the first 
1048: $a$ blocks with the leftmost (signed) block labelled $a$, then
1049: we get the random walk description for $S_{2a}$ that we started with. 
1050: 
1051: With the partition description it is immediate that 
1052: $S_{2a} S_{2b} = S_{4ab}$. 
1053: This gives us a multiplicative shuffle semigroup $\{S_{2a} : a > 0
1054: \} \cup S_1$
1055: contained in $k[\sigma_1] = k[S_2]$.
1056: It now follows that 
1057: $\sigma_0=1$,$\sigma_1$,\ldots,$\sigma_{n}$ is a basis for
1058: $k[\sigma_1]$ and that it is a shuffle algebra.
1059: 
1060: \medskip
1061: \noindent
1062: {\bf The odd part}.
1063: 
1064: This case is completely analogous to the even case,
1065: the only difference being that now we put no
1066: restriction on the zero block.
1067: We begin with the element $\sigma^{\prime}_1 =
1068: \sigma_{s_1}+\cdots+\sigma_{s_{n}}$; that is,
1069: $\sigma^{\prime}_1$ is the sum of all the vertices in $\Sigma$.  
1070: We also
1071: define $\sigma^{\prime}_j$ as the sum of all the simplices of rank $j$ in $\Sigma$
1072: for $j = 1,\ldots,n$. 
1073: In terms of partitions, $\sigma^{\prime}_{j}$ is the sum of all
1074: $(2j+1)$ block partitions.
1075: Note that $\sigma^{\prime}_n = \sigma_n$.
1076: Also put
1077: $\sigma^{\prime}_0 = \sigma_0 = 1$.
1078: Next 
1079: define shuffles $S_{2a+1}$ for $a \geq 0$ by
1080: \begin{equation}\label{e:orshuffle}
1081: S_{2a+1} = \sum_{j=0}^n \binom{a}{j} \sigma^{\prime}_{j}.
1082: \end{equation}
1083: To express $S_{2a+1}$ in words, we again use the notion of a
1084: weak partition. 
1085: Exactly as in the even case, it follows that $S_{2a+1}$ is the sum
1086: of all weak $(2a+1)$ block partitions.
1087: The random walk description is obtained by assigning 
1088: labels $0,1,\ldots,a$ in decreasing order to the first 
1089: $(a+1)$ blocks with the leftmost (signed) block labelled $a$.
1090: 
1091: 
1092: With the partition description it is immediate that 
1093: $S_{2a+1} S_{2b+1} = S_{(2a+1)(2b+1)}$. 
1094: This gives us a multiplicative shuffle semigroup $\{S_{2a+1} : a \geq 0\}$
1095: contained in $k[\sigma^{\prime}_1] = k[S_3]$.
1096: It now follows that 
1097: $\sigma^{\prime}_0=1$,$\sigma^{\prime}_1$,\ldots,$\sigma^{\prime}_{n}$ is a
1098: basis for $k[\sigma^{\prime}_1]$ and that it is a shuffle algebra.
1099: 
1100: \medskip
1101: \noindent
1102: {\bf The shuffle double.}
1103: 
1104: We now put the even and odd parts together.
1105: Define $S = \{S_{2a} : a \geq 1\} \cup \{S_{2a+1} : a \geq 0\}$.
1106: It is immediate from the descriptions of $S_{2a}$ and $S_{2a+1}$ in
1107: terms of partitions that 
1108: $S_{a} S_{b} = S_{ab}$ for all $a,b \geq 1$ irrespective of parity. 
1109: Thus we get a multiplicative shuffle semigroup $S$
1110: as a spanning set in $A = k[\sigma_1,\sigma^{\prime}_1]$.
1111: It follows that $A$ is a $2n$-dimensional algebra with basis  
1112: $\sigma_0=1,\sigma_1,\sigma^{\prime}_1,\ldots,\sigma_{n-1},\sigma^{\prime}_{n-1},\sigma_{n}=\sigma^{\prime}_{n}$.
1113: We call it a shuffle ``double'' algebra. 
1114: It is not a shuffle algebra because the first three conditions are each
1115: violated by a factor of two.
1116: For example, in condition $(3)$, instead of a single generator $\sigma_1$,
1117: we have two generators $\sigma_1$ and $\sigma^{\prime}_1$.
1118: 
1119: We now show that $A \iso k^{2n}$.
1120: Following the example of the riffle shuffle of type $A_{n-1}$, we write
1121: equation~\eqref{e:ershuffle} as 
1122: \begin{equation}\label{e:1} 
1123: S_{2a}=\sum_{i=1}^n (2a)^i e_i, \quad \text{where} \quad e_i = \sum_{j=i}^n
1124: c_{ij}\frac{\sigma_{j}}{2^j j!}\,. 
1125: \end{equation}
1126: Here $c_{ij}$ is the coefficient of $x^i$ in the polynomial
1127: $x(x-2)(x-4)\cdots(x-2(j-1))$.
1128: It is $2^{j-i}$ times the $c_{ij}$ that occurred in the riffle shuffle
1129: of type $A_{n-1}$. For notational convenience, we set $e_0=0$.
1130: Similarly, we rewrite equation~\eqref{e:orshuffle} as
1131: \begin{equation}\label{e:2} 
1132: S_{2a+1}=\sum_{i=0}^n (2a+1)^i e^{\prime}_i, \quad \text{where} \quad e^{\prime}_i = \sum_{j=i}^n
1133: c^{\prime}_{ij}\frac{\sigma^{\prime}_{j}}{2^j j!}\,. 
1134: \end{equation}
1135: Here $c^{\prime}_{ij}$ is the coefficient of $x^i$ in the polynomial
1136: $(x-1)(x-3)\cdots(x-(2j-1))$.
1137: Note that $e^{\prime}_n = e_n = \frac{\sigma_n}{2^n n!}$.
1138: We next define two different kinds of characters of $S$. 
1139: For any non-negative integer $i$, define
1140: $\chi_i(S_{a}) = a^i$ for $a\geq1$.
1141: Also define 
1142: $\chi^{\prime}_i(S_{2a}) = 0$ for $a\geq1$ and $\chi^{\prime}_i(S_{2a+1}) = (2a+1)^i$
1143: for $a\geq0$.
1144: Now observe that for any $s \in S$ 
1145: \begin{equation*} 
1146: s = \sum_{i=1}^n \chi_i(s) e_i + \sum_{i=0}^n \chi^{\prime}_i(s) (e^{\prime}_i - e_i).
1147: \end{equation*}
1148: This can be checked separately for $s=S_{2a},S_{2a+1}$ using
1149: equations~\eqref{e:1} and~\eqref{e:2}. 
1150: The second summation actually goes only till $n-1$, since $e^{\prime}_n = e_n$.
1151: Now apply Proposition~\ref{p:ss} to $S$ and $A$ along with 
1152: the $2n$ characters $\chi_i$ for $i=1,\ldots,n$
1153: and $\chi^{\prime}_i$ for $i=0,1,\ldots,n-1$
1154: to conclude that $A \iso k^{2n}$. 
1155: 
1156: It is clear that $k[\sigma_1] \isoto k^{n+1}$, 
1157: where the $n+1$ factors correspond to the idempotents 
1158: $e^{\prime}_0,e_1,e_2,\ldots,e_n$. 
1159: The isomorphism maps $S_{2a}$ to $(0,(2a)^1,(2a)^2,\ldots,(2a)^n)$.
1160: To see that $k[\sigma^{\prime}_1] \iso k^{n+1}$, 
1161: first note that $\chi^{\prime}_i(s) = \chi_i(s)$ for $s \in \{S_{2a+1} : a \geq
1162: 0\}$. 
1163: Hence we need to lump together the factors corresponding to $e_i$ and
1164: $e^{\prime}_i - e_i$. Thus we get $k[\sigma^{\prime}_1] \isoto k^{n+1}$, with
1165: $e^{\prime}_0,e^{\prime}_1,e^{\prime}_2,\ldots,e^{\prime}_n$ as the primitive idempotents.
1166: The isomorphism maps $S_{2a+1}$ to $((2a+1)^0,(2a+1)^1,(2a+1)^2,\ldots,(2a+1)^n)$.
1167: 
1168: 
1169: 
1170: \section{Examples of type $D_n$}
1171: 
1172: We consider two examples, the side shuffle and the riffle shuffle.
1173: In this section, the term partition always refers to a partition of
1174: type $D_n$; see Appendix~\ref{subs:d}. 
1175: Elements of the label set $I$ will be written as $s_1, s_2, \ldots, s_{n-2},u,v$;
1176: see Figure 2.3 in Appendix~\ref{subs:cd}.
1177: We also recall that the random walk operates on 
1178: an almost signed deck of cards or a deck of type $D_n$. 
1179: It is a deck in
1180: which every card, except the bottomost, is signed.
1181: Also, one of the two basic algebras of Section~\ref{subs:drs} 
1182: does not strictly satisfy
1183: condition $(2)$ in the definition of a shuffle algebra. 
1184: This is relevant to the remark in Section~\ref{subs:defn}.
1185: 
1186: \subsection{The side shuffle}
1187: The element of interest is $\sigma = \sigma_{s_1}$.  In terms of
1188: partitions, $\sigma_{s_1}$ is the sum of all three block partitions
1189: such that the first (and hence the last) block is a singleton.  
1190: The associated random walk on an almost signed deck of $n$ cards
1191: consists of removing a card at random and replacing it on top with
1192: either the same sign or its reverse both with equal probability.
1193: If the bottomost unsigned card is chosen then we give it a sign and
1194: then erase the sign on the new bottomost card.
1195: 
1196: In addition to $\sigma = \sigma_1$,
1197: we define $\sigma_j = \sigma_{J_j}$ where $J_j = \{s_1,\ldots,s_j\}
1198: \subseteq I$ for $j = 1,\ldots,n-2$. 
1199: Note that the choice for $\sigma_{n-1}$ is not immediately obvious.
1200: We set 
1201: $\sigma_{n-1} = \sigma_{\{s_1,\ldots,s_{n-1},u,v\}}$.
1202: The motivation becomes clearer from the following.
1203: As partitions, $\sigma_{j}$ is the sum of all
1204: $(2j+1)$ block partitions such that the first (and hence the last) $j$
1205: blocks are singletons.
1206: Observe that for $j=n-1$, this does give us the above definition of
1207: $\sigma_{n-1}$.
1208: We also set $\sigma_n = 2 \sigma_{n-1}$.
1209: Define $a$-shuffles for $a \geq 0$ by 
1210: \begin{equation*} 
1211: S_a = \sum_{j=0}^n S(a,j) \sigma_j,
1212: \end{equation*}
1213: where $S(a,j)$ are the signed Stirling numbers defined in
1214: Section~\ref{subs:atss}.
1215: The associated random walks on a deck of type $D_n$ are as follows.
1216: 
1217: $\sigma_j :$ Choose $j$ cards at random. Move them to the top in a
1218: random order. Now with equal probability either flip or do not flip
1219: the sign of the chosen cards.  
1220: If the bottomost unsigned card is one of the chosen cards then we give
1221: it a sign and 
1222: then erase the sign on the new bottomost card.
1223: 
1224: $S_a :$ Assign the sign $+$ to the bottomost card. Now do the side
1225: shuffle $S_a$ of type $B_n$. Drop the sign of the bottomost card.
1226: 
1227: Observe that $\sigma_n = 2 \sigma_{n-1}$. We did not have any such
1228: relation for the side shuffle of type $B_n$. It is interesting how
1229: this relation emerges by just making the bottomost card unsigned.
1230: Also observe that $S_a S_b = S_{a+b}$. At this point, it is clear that
1231: the rest is identical to the two sided shuffle for type
1232: $A_{n-1}$ and we omit it. 
1233: 
1234: \subsection{The riffle shuffle} \label{subs:drs}
1235: 
1236: 
1237: The shuffles $S_a$ in this example can be described in the same way as
1238: the inverse $a$-shuffles of type $B_{n}$ with a minor modification.  
1239: This is because we are now operating on a deck of type $D_n$
1240: rather than of type $B_n$.
1241: The difference between the two is that in the former the bottomost
1242: card is unsigned.
1243: We now describe $S_a$, the inverse $a$-shuffle of type $D_n$.
1244: 
1245: $S_a :$ Assign the sign $+$ to the bottomost card. Now do an inverse
1246: $a$-shuffle of type $B_n$. Drop the sign of the bottomost card.
1247: 
1248: It is useful to describe these shuffles in another equivalent way
1249: using the analogy with the previous situation 
1250: rather than the end result. For that,
1251: we first define the inverse riffle shuffle of type $D_n$.
1252: 
1253: For every signed card, we either flip or do not flip its sign with equal
1254: probability. The cards whose signs were flipped move below the
1255: unsigned card (in the reverse relative order). The rest stay on top.
1256: 
1257: Note that after this shuffle, we may not have a deck of type $D_n$, 
1258: since the unsigned card is not necessarily at the bottom. Hence we
1259: define a \emph{correction} operation which assigns a sign to the unsigned
1260: card and then erases the sign of the bottomost card.
1261: We now give a description of our shuffles as operators on a deck of
1262: type $D_n$.
1263: 
1264: $S_{2a}$: We do a usual inverse $a$-shuffle with labels
1265: $1,2,\ldots,a$. Then within each block with a fixed label we do an
1266: inverse riffle shuffle of type $D_n$ or $B_n$ depending on whether the
1267: block contains the unsigned card or not. Lastly, we do the correction.
1268: 
1269: $S_{2a+1}$: We do a usual inverse $(a+1)$-shuffle with labels
1270: $0,1,2,\ldots,a$. Then we repeat the above on each
1271: block except the one labelled 0 and then do the correction.
1272: 
1273: Note that $S_2$ is just an inverse riffle shuffle of type $D_n$
1274: followed by the correction.
1275: 
1276: \medskip
1277: \noindent
1278: {\bf The even part}.
1279: 
1280: The element of interest is $\sigma_1 = \sigma_{u} + \sigma_{v}$. 
1281: In terms of partitions, $\sigma_{1}$ is the sum of all
1282: three block partitions such that the central
1283: block is  empty.
1284: 
1285: 
1286: Put $\sigma_j = \sum_{\abs{J}=j,u \in J}\sigma_J + \sum_{\abs{J}=j,v \in J}\sigma_J$ for $j=1,2,\ldots,n$.
1287: Explicitly for $n=3$, we have
1288: $\sigma_1 = \sigma_{u} + \sigma_{v},
1289: \sigma_2 = \sigma_{\{s,u\}} + \sigma_{\{s,v\}} + 2 \sigma_{\{u,v\}},
1290: \sigma_3 = 2 \sigma_{\{s,u,v\}}$.
1291: Note that we consider only those faces whose type contains either $u$
1292: or $v$. If it contains both $u$ and $v$ then we put a coefficient of
1293: $2$ in front.
1294: In terms of partitions, $\sigma_{j}$ is the sum of all
1295: $(2j+1)$ block partitions such that the central
1296: block is  empty.
1297: In this description we follow the convention that the partitions 
1298: $(\{1,\overline{2}\},\{3\},\{\},\{\overline{3}\},\{2,\overline{1}\})$ 
1299: and
1300: $(\{1,\overline{2}\},\{\overline{3}\},\{\},\{3\},\{2,\overline{1}\})$
1301: are counted separately, even though they are both equal to 
1302: $(\{1,\overline{2}\},\{3,\overline{3}\},\{2,\overline{1}\})$.
1303: This explains the coefficient of $2$ in front of faces whose type
1304: contains both $u$ and $v$.
1305: 
1306: \medskip
1307: \noindent
1308: {\bf The odd part}.
1309: 
1310: The element of interest is $\sigma^{\prime}_1 = (\sigma_{u} + \sigma_{v}) +
1311: (\sigma_{s_1} + \ldots + \sigma_{s_{n-2}}) + \sigma_{\{u,v\}}$. 
1312: Note that in contrast to all our examples so far,
1313: $\sigma^{\prime}_1$ is a combination of elements of $k\Sigma_1$
1314: and $k\Sigma_2$. Next we define
1315: $\sigma^{\prime}_j = \sigma_j + \sum_{\abs{J}=j,u \notin J, v \notin J}\sigma_J +
1316: \sum_{\abs{J}=j+1,\{u,v\} \subset J}\sigma_J$ for $j=1,2,\ldots,n$.
1317: Here $\sigma_j$ is as defined in the even case.
1318: Thus we see that $\sigma^{\prime}_j$ is a combination of elements of $k\Sigma_j$
1319: and $k\Sigma_{j+1}$.
1320: This is the only example where the geometric description of 
1321: objects of interest is somewhat complicated.
1322: For better motivation we now consider partitions. 
1323: We have already interpreted $\sigma_j$ in terms of partitions.
1324: Note that the two summation terms that we added give the sum 
1325: of all
1326: $(2j+1)$ block partitions with a non-empty central
1327: block.
1328: Hence we see that $\sigma^{\prime}_j$ is the sum 
1329: of all
1330: $(2j+1)$ block partitions.
1331: While counting partitions, we adopt the same convention as in the even case.
1332: 
1333: \medskip
1334: \noindent
1335: {\bf The shuffle double.}
1336: 
1337: We define shuffles $S_{2a}$ and $S_{2a+1}$ by
1338: equations~\eqref{e:ershuffle} and~\eqref{e:orshuffle} and the analysis is
1339: identical word for word to the case of $B_n$. We only point out that
1340: partition now means partition of type $D_n$ and $\sigma_j$ and
1341: $\sigma^{\prime}_j$ refer to the definitions that we made above. 
1342: The shuffles can be described using weak partitions 
1343: and this leads to the two random walk interpretations
1344: that we gave earlier.
1345: They are slightly complicated now
1346: because of the more involved nature
1347: of the partitions of type $D_n$.
1348: A clear conceptual explanation of the close connection between the
1349: riffle shuffles of type $B_n$ and $D_n$ is given in the next section.
1350: 
1351: \begin{remark}
1352: The riffle shuffle examples in Sections~\ref{subs:ars}
1353: and~\ref{subs:brs}(odd part) show that for types $A_{n-1}$ and $B_n$,
1354: $\sigma_i \sigma_j = \sigma_j \sigma_i$, 
1355: where
1356: $\sigma_i$ is the sum of all the simplices of rank $i$.
1357: This is because $\sigma_i$ and $\sigma_j$
1358: are elements of a commutative (shuffle) algebra.
1359: The fact that
1360: $\sigma_i \sigma_j = \sigma_j \sigma_i$
1361: can also be proven by a direct geometric argument. The key property is
1362: that the simplicial 
1363: complex induced on the support of any face of the Coxeter complex of type
1364: $A_{n-1}$ or $B_n$ is again of the same type.
1365: This property fails for $D_n$.
1366: It is incomplete in this geometric sense.
1367: Note that the $\sigma_i$'s as defined above did not play any role in the
1368: riffle shuffle for $D_n$.
1369: In fact, their role for type $D_n$ in this theory is far from being clear.
1370: \end{remark}
1371: 
1372: 
1373: 
1374: \section{Maps between Coxeter complexes} \label{s:maps}
1375: 
1376: Recall that the reflection arrangements of type $B_n$, $D_n$
1377: and $A_{n-1}$ are given by the hyperplanes
1378: $x_i= \pm x_j, x_i = 0$; $x_i= \pm x_j$ and $x_i=x_j$ 
1379: $(1\le i<j\le n)$ respectively.
1380: Let $\Sigma(B_n)$, $\Sigma(D_n)$
1381: and $\Sigma(A_{n-1})$ be the corresponding Coxeter complexes.
1382: Observe that the arrangement for $D_n$ (resp. $A_{n-1}$) is
1383: obtained from the one for $B_n$ (resp. $D_n$) by deleting some
1384: hyperplanes. 
1385: Following the discussion in Appendix~\ref{subs:map}, we get maps
1386: $\Sigma(B_n) \rightarrow \Sigma(D_n) \rightarrow \Sigma(A_{n-1})$.
1387: These are semigroup homomorphisms and
1388: they induce algebra homomorphisms
1389: $k\Sigma(B_n) \rightarrow k\Sigma(D_n) \rightarrow k\Sigma(A_{n-1})$.
1390: We will use the language of
1391: partitions to make these maps more explicit.
1392: 
1393: \begin{figure}[hbt]
1394: \centering
1395: \begin{tabular}{c@{\qquad}c}
1396: \mbox{\epsfig{file=maps.eps,height=2.7cm,width = 12cm}}
1397: \end{tabular}
1398: \caption{The case $n=3$.}
1399: \label{maps}
1400: \end{figure}
1401: 
1402: \noindent
1403: Figure~\ref{maps} shows the intersection of the hyperplane arrangements
1404: for $B_3$, $D_3$ and $A_2$
1405: with the boundary of a cube centred at the origin.
1406: For $B_3$ and $D_3$, 
1407: this gives us the corresponding Coxeter complex.
1408: For $A_2$, we do not quite get $\Sigma(A_2)$ 
1409: because the braid arrangement is not essential.
1410: So we cut the braid arrangement 
1411: by the hyperplane $x_1+x_2+x_3=0$.
1412: The complex $\Sigma(A_2)$ can be seen in Figure~\ref{maps}
1413: as the dotted hexagon 
1414: with three vertices each of type $s_1$ and $s_2$.
1415: Also in Figure~\ref{maps},
1416: we have labelled only some of the vertices
1417: since they uniquely determine the remaining labels.
1418: 
1419: \subsection{The map $\Sigma(B_n) \rightarrow \Sigma(D_n)$} \label{subs:map1}
1420: 
1421: For the case $B_n$ (resp. $D_n$) we have split the partition
1422: description for a face $F$ into 4 (resp. 3) cases; see
1423: Appendix~\ref{subs:b} and~\ref{subs:d}. From those descriptions, it is
1424: straightforward to describe the map. 
1425: Let $K \subset \{s_1,\ldots,s_{n-2}\}$. 
1426: 
1427: \begin{itemize}
1428: \item[(i)]
1429: A face of type $K$ in $\Sigma(B_n)$ maps to a face of type $K$ in
1430: $\Sigma(D_n)$. 
1431: 
1432: \item[(ii)]
1433: A face of type $K \cup t$ in $\Sigma(B_n)$  maps to a face either of type 
1434: $K \cup u$ or $K \cup v$ depending on a parity condition. 
1435: 
1436: \item[(iii)]
1437: A face of type $K \cup s_{n-1}$ in $\Sigma(B_n)$ maps to a face of one
1438: higher dimension of type  
1439: $K \cup \{u,v\}$. 
1440: For $n=3$, this is also clear from Figure~\ref{maps},
1441: where we see that 
1442: $\{s_2\} \mapsto \{u,v\}$ and $\{s_1,s_2\}\mapsto\{s_1,u,v\}$. 
1443: 
1444: \item[(iv)]
1445: A face of type $K \cup \{s_{n-1},t\}$ in $\Sigma(B_n)$ maps to a face of type 
1446: $K \cup \{u,v\}$. 
1447: \end{itemize}
1448: %
1449: The maps in the first three cases are bijective.
1450: More precisely, the faces of type $K$ in $\Sigma(B_n)$
1451: are in bijection with the faces of type $K$ in $\Sigma(D_n)$
1452: and so on. However in the last case, 
1453: the map is two to one. This is again
1454: illustrated in Figure~\ref{maps},
1455: where we see that there are two faces of type $\{s_1,s_2,t\}$
1456: which map to a face of type $\{s_1,u,v\}$.
1457: 
1458: Recall that $\sigma_J$ is the sum of all faces of type $J$.
1459: From the above discussion it follows that 
1460: for the map $k\Sigma(B_n) \rightarrow k\Sigma(D_n)$,
1461: we have
1462: $\sigma_{K} \mapsto \sigma_{K}$,
1463: $\sigma_{K \cup s_{n-1}} \mapsto \sigma_{K \cup \{u,v\}}$,
1464: $\sigma_{K \cup t} \mapsto \sigma_{K \cup u} + \sigma_{K \cup v}$ and
1465: $\sigma_{K \cup \{s_{n-1},t\}} \mapsto 2 \sigma_{K \cup \{u,v\}}$. 
1466: 
1467: 
1468: 
1469: \subsection{The map $\Sigma(B_n) \rightarrow \Sigma(A_{n-1})$} \label{subs:map2}
1470: It is easier to describe this composite map than the map from
1471: $\Sigma(D_n)$ to $\Sigma(A_{n-1})$. In contrast to the previous map,
1472: this composite map is hard to describe using face types but easy to
1473: describe using partitions. 
1474: To get an ordered partition of $[n]$, 
1475: starting with an anti-symmetric
1476: ordered partition of $[n,\overline{n}]$, we simply forget the set $[\overline{n}]$. 
1477: For example,
1478: $(\{2\},\{\overline{3}\},\{1,\overline{1}\},\{3\},\{\overline{2}\})$
1479: maps to $(\{2\},\{1\},\{3\})$.
1480: We explain some special cases.
1481: 
1482: Observe that a face of type $s_1$ maps to a face either of type $s_1$
1483: or $s_{n-1}$ depending on whether the element in the singleton first
1484: block has a plus sign or minus. It follows that
1485: for the map $k\Sigma(B_n) \rightarrow k\Sigma(A_{n-1})$,
1486: $\sigma_{s_1}$ maps to $\sigma_{s_1}+\sigma_{s_{n-1}}$.
1487: This is the primary reason why the side shuffle of type $B_n$
1488: is more closely related to the two-sided shuffle 
1489: (rather than the side shuffle) of type $A_{n-1}$.
1490: 
1491: Next we show that inverse $a$-shuffles map to inverse $a$-shuffles.
1492: For that
1493: observe that weak $(2a+1)$ block partitions of type $B_n$
1494: correspond to weak $(2a+1)$ block partitions of type $A_{n-1}$.
1495: For example,
1496: $(\{\overline{2}\},\{\},\{1,\overline{1}\},\{\},\{{2}\})
1497: \leftrightarrow (\{\},\{\},\{1\},\{\},\{{2}\})$.
1498: Similarly, weak $(2a+1)$ block partitions of type $B_n$ with an empty
1499: zero block correspond to weak $2a$ block partitions of type $A_{n-1}$.
1500: For example, when $a=1$, we get that 
1501: $\sigma_{t}$ maps to $\sigma_{s_1}+\ldots+\sigma_{s_{n-1}}+2\sigma_0$.
1502: In other words, 
1503: the inverse riffle shuffle of type $B_n$ maps to the inverse
1504: riffle shuffle of type $A_{n-1}$.
1505: 
1506: 
1507: \subsection{Maps between side shuffles}
1508: 
1509: We recall the definitions of $\sigma_j$ for the side
1510: shuffles of type $B_n$ and $D_n$ and the two-sided shuffle of type
1511: $A_{n-1}$.
1512: 
1513: \begin{itemize}
1514: \item[$B_n$] $: 
1515: \sigma_j = \sigma_{J_j}$ where $J_j = \{s_1,\ldots,s_j\}
1516: \subseteq I$ for $j = 1,\ldots,n$. 
1517: 
1518: \item[$D_n$] $: 
1519: \sigma_j = \sigma_{J_j}$ where $J_j = \{s_1,\ldots,s_j\}
1520: \subseteq I$ for $j = 1,\ldots,n-2$ and $J_{n-1} = \{s_1,\ldots,s_{n-1},u,v\}$. 
1521: Also $\sigma_n = 2 \sigma_{n-1}$.
1522: 
1523: \item[$A_{n-1}$] $: 
1524: \sigma_j = \sum_{k=0}^j \binom{j}{k}\sigma_{J_{j,k}}$ where
1525: $J_{j,k} = \{s_1,\ldots,s_k\} \cup \{s_{n-(j-k)},\ldots,s_{n-1}\}
1526: \subseteq I$ for $j = 1,\ldots,n-1$ and $\sigma_n = 2 \sigma_{n-1}$.
1527: 
1528: \end{itemize}
1529: %
1530: We claim that 
1531: $\sigma_j \mapsto \sigma_j \mapsto \sigma_j$
1532: under the maps 
1533: $k\Sigma(B_n) \rightarrow k\Sigma(D_n) \rightarrow k\Sigma(A_{n-1})$.
1534: From the discussion in Section~\ref{subs:map1},
1535: it is easy to see the claim for the first map.
1536: A point worth noting is that $\sigma_n \mapsto 2 \sigma_{n-1}$
1537: and the claim holds for $j=n$
1538: because of the relation $\sigma_n = 2 \sigma_{n-1}$ 
1539: for $D_n$.
1540: Similarly, using the discussion in Section~\ref{subs:map2},
1541: we can show the claim for the composite map.
1542: In earlier sections,
1543: the introduction of $\sigma_n$ and the relation 
1544: $\sigma_n = 2 \sigma_{n-1}$
1545: for types $D_n$ and $A_{n-1}$ was somewhat artificial and justified 
1546: only from the random walk perspective.
1547: However we now also have a clear geometric perspective.
1548: 
1549: Recall that the shuffles are given by the same formula 
1550: $S_a = \sum_{j=0}^n S(a,j) \sigma_j$
1551: in all three cases with $S(a,j)$ being the signed Stirling numbers. 
1552: Hence we get
1553: $S_a \mapsto S_a \mapsto S_a$.
1554: To summarise, we have:
1555: 
1556: \begin{displaymath}
1557: \begin{array}{ccccc}
1558: \left\{
1559: \begin{array}{c}
1560: \text {side shuffle}\\
1561: \text {of type $B_n$}
1562: \end{array}
1563: \right\}&\longrightarrow&\left\{
1564: \begin{array}{c}
1565: \text {side shuffle}\\
1566: \text {of type $D_n$}\\
1567: \end{array}
1568: \right\}&\isoto&\left\{
1569: \begin{array}{c}
1570: \text {two sided shuffle}\\
1571: \text {of type $A_{n-1}$}\\
1572: \end{array}
1573: \right\}\\
1574: \sigma_j & \mapsto & \sigma_j & \mapsto & \sigma_j \\
1575: S_a & \mapsto & S_a & \mapsto & S_a 
1576: \end{array}
1577: \end{displaymath}
1578: 
1579: \subsection{Maps between riffle  shuffles}
1580: 
1581: Recall the definitions of $\sigma_j$ and $\sigma^{\prime}_j$ for the riffle
1582: shuffles of type $B_n$ and $D_n$ from 
1583: Sections~\ref{subs:brs} and \ref{subs:drs}. 
1584: Following the discussion in
1585: Section~\ref{subs:map1}, it is clear that 
1586: $\Sigma(B_n) \rightarrow \Sigma(D_n)$
1587: maps $\sigma_j$ to $\sigma_j$ and $\sigma^{\prime}_j$ to $\sigma^{\prime}_j$.
1588: So the map restricted to the algebras $A = k[\sigma_1,\sigma^{\prime}_1]$ 
1589: is an algebra map that maps a basis to a basis.
1590: Hence it is an isomorphism. 
1591: It also clearly maps $S_a$ to $S_a$.
1592: Again we see that the somewhat unmotivated definitions of
1593: $\sigma_j$ and $\sigma^{\prime}_j$ for $D_n$
1594: have a more natural meaning as the images of the corresponding 
1595: elements in $\Sigma(B_n)$.
1596: 
1597: For the map $\Sigma(B_n) \rightarrow \Sigma(A_{n-1})$, 
1598: the discussion at the end of Section~\ref{subs:map2} 
1599: and the interpretation of the inverse $a$-shuffles $S_a$ as weak partitions shows
1600: that $S_a \mapsto S_a$. This gives us a surjective map
1601: $A = k[\sigma_1,\sigma^{\prime}_1] = k[S_2,S_3] \rightarrow A = k[S_2]$.  
1602: However we do not know a good way to describe 
1603: the images of $\sigma_j$ and $\sigma^{\prime}_j$ under this map. 
1604: To summarise, we have:
1605: 
1606: \begin{displaymath}
1607: \begin{array}{ccccc}
1608: \left\{
1609: \begin{array}{c}
1610: \text {riffle shuffle}\\
1611: \text {of type $B_n$}
1612: \end{array}
1613: \right\}&\isoto&\left\{
1614: \begin{array}{c}
1615: \text {riffle shuffle}\\
1616: \text {of type $D_n$}\\
1617: \end{array}
1618: \right\}&\longrightarrow&\left\{
1619: \begin{array}{c}
1620: \text {riffle shuffle}\\
1621: \text {of type $A_{n-1}$}\\
1622: \end{array}
1623: \right\}\\
1624: \sigma_j & \mapsto & \sigma_j & \mapsto & ? \\
1625: \sigma^{\prime}_j & \mapsto & \sigma^{\prime}_j & \mapsto & ? \\
1626: S_a & \mapsto & S_a & \mapsto & S_a 
1627: \end{array}
1628: \end{displaymath}
1629: 
1630: \vanish{
1631: \matrix
1632: {\left\{
1633: \begin{tabular}{c}
1634: side shuffle\\
1635: of type $B_n$\\
1636: \end{tabular}
1637: \right\}&\rightarrow&\left\{
1638: \begin{tabular}
1639: side shuffle\\
1640: of type $B_n$\\
1641: \end{tabular}
1642: \right\}&\rightarrow&\left\{
1643: \begin{tabular}
1644: side shuffle\\
1645: of type $B_n$\\
1646: \end{tabular}
1647: \right\}\cr}
1648: }
1649: 
1650: \begin{remark}
1651: The Coxeter complex of type $B_n$ is probably the correct place
1652: to look for shuffle algebras.
1653: The images of these shuffle algebras under our maps
1654: would then yield shuffle algebras (in a suitably generalised sense)
1655: of type $D_n$ and $A_{n-1}$.
1656: Among the examples in this paper,
1657: the only one that we failed to describe in this manner
1658: was the side shuffle of type $A_{n-1}$.
1659: \end{remark}
1660: 
1661: 
1662: 
1663: \section{Generalisation to buildings}\label{s:q}
1664: 
1665: We first give a brief review of buildings.
1666: For the general theory
1667: see~\cite{\brown,\schar,\tits}. 
1668: Let $W$ be a Coxeter group and $\Sigma(W)$ its Coxeter complex.
1669: Roughly,
1670: a building $\Delta$ of type $W$ is a union of subcomplexes $\Sigma$ (called
1671: apartments) which fit together nicely.
1672: Each apartment $\Sigma$ is isomorphic to $\Sigma(W)$.
1673: As a simplicial complex, $\Delta$ is pure and labelled.
1674: The term \emph{pure} means that all maximal simplices (chambers)
1675: have the same dimension. 
1676: For any two simplices in $\Delta$, there is an apartment $\Sigma$
1677: containing both of them.
1678: Using this fact, we can define a product on $\Delta$ as follows.
1679: 
1680: For $x,y \in \Delta$, we choose an apartment $\Sigma$ containing $x$ and $y$
1681: and define $xy$ to be their product in $\Sigma$.
1682: Since $\Sigma$ is a Coxeter complex, we know how to do this.
1683: Furthermore, it can be shown that the product does not depend on the
1684: choice of $\Sigma$.
1685: So this defines a product on $\Delta$ 
1686: with the set of chambers $\C$ contained as an ideal.
1687: As in Section~\ref{subs:rw},
1688: we can now define a random walk on $\C$ 
1689: starting with a probability distribution on $\Delta$.
1690: Unfortunately, the product on $\Delta$ is not associative,
1691: that is,
1692: $\Delta$ is no longer a semigroup.
1693: As a result $k\C$ is not a module over $k\Delta$
1694: and so our algebraic methods break down.
1695: However, we know three examples where they do work.
1696: This is because in these cases, the subalgebra of interest,
1697: namely $k[\sigma]$, turns out to be associative.
1698: 
1699: 
1700: A theorem of Tits \cite[Theorem 11.4]{\tits} says that
1701: if $\Delta$ is a finite, irreducible Moufang building
1702: then it is the building of an
1703: absolutely simple algebraic group $G$ over a finite field $\fq$.
1704: The buildings we consider satisfy this hypothesis;
1705: hence in each example we will also specify the algebraic group $G$.
1706: The group $G$ acts by simplicial type-preserving automorphisms on
1707: $\Delta$. To avoid getting into details, 
1708: we just mention that the action of $G$ is very closely related to
1709: the geometry of the building.
1710: 
1711: The three examples we consider can be thought of as 
1712: $q$-analogues of the side shuffles of 
1713: type $A_{n-1}$, $B_n$ and $D_n$.
1714: We use the same definitions for the $\sigma_j$'s as before,
1715: except that we now apply them to the building $\Delta$
1716: rather than the Coxeter complex $\Sigma$.
1717: We will use the notation $[j] = 1 + q + \ldots + q^{j-1}$
1718: to denote the $q$-numbers. They will show up a lot in our analysis.
1719: 
1720: 
1721: 
1722: \subsection{The q-side shuffle for $A_{n-1}$} \label{subs:qass}
1723: 
1724: We first briefly describe the building of type $A_{n-1}$
1725: associated to the algebraic group $GL_n(\fq)$ (also $SL_n(\fq)$).
1726: Let $V$ be the $n$-dimensional vector space over $\fq$
1727: and $\mathcal{L}_{n}$ be the lattice of subspaces
1728: of $V$. 
1729: The building of type $A_{n-1}$ associated to $GL_n(\fq)$ 
1730: is simply the flag (order) complex
1731: $\Delta(\mathcal{L}_n)$. 
1732: It is a labelled simplicial complex.
1733: A vertex of $\Delta(\mathcal{L}_n)$ is by definition a proper subspace of $V$
1734: and we label it by its dimension.
1735: To be consistent with earlier notation,
1736: we label the vertices $s_1, \ldots, s_{n-1}$ 
1737: instead of just $1, \ldots, n-1$.
1738: In particular, vertices of type $s_1$ are the one dimensional
1739: subspaces of $V$.
1740: 
1741: Let $\mathcal{B}_{n}$ be
1742: the Boolean lattice of rank $n$ consisting
1743: of all subsets of an $n$-set ordered under inclusion.
1744: Note that the order complex $\Delta(\mathcal{B}_n)$ is the Coxeter complex of type
1745: $A_{n-1}$, 
1746: which we earlier denoted by $\Sigma(A_{n-1})$.
1747: Also note that a
1748: choice of a basis for $V$ gives an 
1749: embedding of $\mathcal{B}_n$ into $\mathcal{L}_n$.
1750: The subcomplexes $\Delta(\mathcal{B}_n)$, for various embeddings $\mathcal{B}_n \into \mathcal{L}_n$,
1751: play the role of apartments.
1752: Alternatively, it is also useful to think of $\mathcal{B}_n$ as a specialisation of $\mathcal{L}_n$
1753: for the degenerate case $q=1$.
1754: In this sense $\Delta(\mathcal{L}_n)$
1755: may be regarded as a $q$-analogue of $\Delta(\mathcal{B}_n) = \Sigma(A_{n-1})$.
1756: For the rest of the section
1757: we will denote the building $\Delta(\mathcal{L}_n)$ by $\Delta$.
1758: 
1759: The product in $\Delta$ can be made more explicit. 
1760: A face of $\Delta$ is a chain in $\mathcal{L}_n$
1761: and a chamber is a maximal chain in $\mathcal{L}_n$.
1762: For $x, y \in \Delta$, the face $xy$ is the chain in $\mathcal{L}_n$ obtained by
1763: refining the chain $x$ by the chain $y$ (using meets and joins as in a
1764: Jordan--H\"{o}lder product).
1765: This is a generalisation of the product in $\Sigma(A_{n-1})$
1766: which we defined
1767: in terms of refinement of one ordered partition by another.
1768: 
1769: The side shuffle of type $A_{n-1}$ (Section~\ref{subs:ass})
1770: generalises in a straightforward way to this setting. Let
1771: $\sigma_1$ be the sum of all the vertices of type $s_1$,
1772: $\sigma_2$ be the sum of all the edges of type $s_1 s_2$ and so on till
1773: $\sigma_{n-1}$ which is the sum of all chambers of $\Delta$.
1774: The number of summands in $\sigma_{j}$, 
1775: which is same as the number of faces of type
1776: $s_1 s_2 \ldots s_j$, is
1777: $[n][n-1] \ldots [n-j+1]$.
1778: It follows directly from the definition that 
1779: $\sigma_j \sigma_1 = \sigma_1 \sigma_j = 
1780: [j] \sigma_j + q^j \sigma_{j+1}$ if $j < n-1$ and 
1781: $\sigma_{n-1} \sigma_1 = [n] \sigma_{n-1}$.
1782: As a consistency check, 
1783: we can also verify that the number of terms 
1784: on each side is the same.
1785: These facts imply that $\sigma_1$ is power associative.
1786: Hence we may write the following without ambiguity. 
1787: \begin{equation}\label{e:qcob}
1788: q^{\binom{j+1}{2}}\sigma_{j+1} = \sigma_1(\sigma_1-[1])\ldots(\sigma_1-[j]).
1789: \end{equation}
1790: This along with
1791: $\sigma_{n-1} \sigma_1 = [n] \sigma_{n-1}$ implies that both
1792: $\sigma_0=1$,$\sigma_1$,\ldots,$\sigma_{n-1}$ and
1793: $\sigma_1^0=1$,$\sigma_1^1$,\ldots,$\sigma_1^{n-1}$
1794: form a basis for the associative algebra $A = k[\sigma_1]$.
1795: The basic relation satisfied by $\sigma_1$ is 
1796: $\sigma_1(\sigma_1-[1])\ldots(\sigma_1-[n-2])(\sigma_1-[n]) = 0$.
1797: Hence
1798: $A \isoto \frac{k[x]}{x(x-[1])\ldots(x-[n-2])(x-[n])}$,
1799: where $\sigma_1$ maps to $x$.
1800: Since the roots are distinct
1801: we conclude that $A$ is split semisimple.
1802: 
1803: As before,
1804: we extend the definition of $\sigma_j$ to any $j$
1805: using~\eqref{e:qcob} to obtain 
1806: $\sigma_n = \sigma_{n-1}$ and $\sigma_j = 0$ for
1807: $j > n$.
1808: Inverting equation~\eqref{e:qcob} formally, we get
1809: \begin{equation*}
1810: \sigma_1^a = \sum_{j=0}^a S(a,j) \sigma_j
1811: \end{equation*}
1812: where $S(a,j)$ satisfies the recursion:
1813: $S(a,j) = [j] S(a-1,j) + q^{j-1} S(a-1,j-1)$ with $S(a,1) = 1$ and 
1814: $S(a,a) = q^{\binom{a}{2}}$. 
1815: %Since $\sigma_n = \sigma_{n-1}$ and $\sigma_j = 0$ for $j > n$,
1816: %the inversion above is not unique. But it is the most natural one to
1817: %consider. 
1818: The numbers $S(a,j)$ give a $q$-analogue to the Stirling numbers of
1819: the second kind. For more information on these numbers, see~\cite{\wachswhite} and the references therein.
1820: 
1821: 
1822: 
1823: \begin{remark}
1824: The analysis we gave for this example was the analogue of the
1825: direct method of Section~\ref{subs:ass}.
1826: The difficulty with the shuffle method is that
1827: we do not know of a good random walk interpretation of $\sigma_1^a$.
1828: However as the direct method shows,
1829: the algebra $A$ satisfies all the properties of 
1830: an additive shuffle algebra 
1831: (except that now we are in the more general context of buildings).
1832: There are also natural ways to define the q-analogue of the two-sided
1833: shuffle and the riffle shuffle. However the analysis breaks down
1834: because the basic elements of interest are not
1835: power associative.
1836: \end{remark}
1837: 
1838: \subsection{The q-side shuffle for $B_{n}$} \label{subs:qbss}
1839: 
1840: There are two possibilities for the building depending 
1841: on whether the associated algebraic group is
1842: the symplectic group $Sp_{2n}(\fq)$ or
1843: the orthogonal group $O_{2n}(\fq)$.
1844: We consider them separately.
1845: They may both be regarded as $q$-analogues of $\Sigma(B_n)$.
1846: The side shuffle for $B_{n}$ generalises to both cases.
1847: As expected, the computations have the same spirit as the previous section. 
1848: 
1849: \medskip
1850: \noindent
1851: {\bf The symplectic case.}
1852: Let $V$ be a (even dimensional) vector space over $\fq$ 
1853: with a skew-symmetric non-degenerate bilinear form $Q$.
1854: More explicitly, $V$ has a basis
1855: $e_1, e_2, \ldots, e_n, f_1, f_2, \ldots, f_n$ 
1856: such that $Q(e_i,f_i) = 1$, $Q(f_i,e_i) = -1$
1857: and all other pairings between basis vectors are zero.
1858: The symplectic group $Sp_{2n}(\fq)$ 
1859: is the group of automorphisms of $V$ 
1860: that preserve the bilinear form $Q$.
1861: 
1862: The building of type $B_{n}$
1863: associated to the symplectic group
1864: is the flag (order) complex of all isotropic subspaces of $V$.
1865: We will denote it by $\Delta$.
1866: The definition of the product in $\Delta$ is similar to that in
1867: Section~\ref{subs:qass}.
1868: The difference occurs due to the fact that 
1869: the sum of two isotropic subspaces may not be isotropic.
1870: 
1871: Let
1872: $\sigma_1$ be the sum of all the vertices of type $s_1$,
1873: $\sigma_2$ be the sum of all the edges of type $s_1 s_2$ and so on till
1874: $\sigma_{n}$ which is the sum of all chambers of $\Delta$.
1875: It follows from the definition that 
1876: $\sigma_j \sigma_1 = \sigma_1 \sigma_j = 
1877: (1+q^{2n-j})[j] 
1878: \sigma_j + q^j \sigma_{j+1}$ if $j \leq n-1$ and 
1879: $\sigma_{n} \sigma_1 = \sigma_{1} \sigma_n = [2n] \sigma_{n}$.
1880: An easy way to check this is to first figure out the coefficient of 
1881: $\sigma_{j+1}$. It is $q^j$, the same answer that we got for the
1882: $A_{n-1}$ case.
1883: There is no difference in the two situations
1884: with respect to this coefficient.
1885: With this information, we can find the coefficient of
1886: $\sigma_{j}$ by simply counting the number of terms involved 
1887: on both sides. 
1888: The number of summands in $\sigma_{j}$, 
1889: which is same as the number of faces of type
1890: $s_1 s_2 \ldots s_j$, is
1891: $[2n][2n-2] \ldots [2n-2j+2]$.
1892: The identity is then a consequence of the simple relation
1893: $[2n] = (1+q^{2n-j})[j] + q^j [2n-2j]$.
1894: It is a good exercise to compute the coefficient of $\sigma_j$ directly.
1895: 
1896: These facts imply that $\sigma_1$ is power associative.
1897: Hence we may write the following without ambiguity. 
1898: \begin{equation}\label{e:qcob2}
1899: q^{\binom{j+1}{2}}\sigma_{j+1} = \sigma_1(\sigma_1-(1+q^{2n-1})[1])\ldots(\sigma_1-(1+q^{2n-j})[j]).
1900: \end{equation}
1901: This along with
1902: $\sigma_{n} \sigma_1 = [2n] \sigma_{n}$ implies that both
1903: $\sigma_0=1$,$\sigma_1$,\ldots,$\sigma_{n}$ and
1904: $\sigma_1^0=1$,$\sigma_1^1$,\ldots,$\sigma_1^{n}$
1905: form a basis for the associative algebra $A = k[\sigma_1]$.
1906: The basic relation satisfied by $\sigma_1$ is 
1907: $\sigma_1(\sigma_1- (1+q^{2n-1})[1])\ldots(\sigma_1-(1+q^{n+1})[n-1])(\sigma_1-[2n]) = 0$.
1908: Hence
1909: $A \isoto \frac{k[x]}{x(x-(1+q^{2n-1})[1])\ldots(x-(1+q^{n+1})[n-1])(x-[2n])}$.
1910: This shows that $A$ is split semisimple.
1911: 
1912: As before,
1913: we extend the definition of $\sigma_j$ to any $j$
1914: using~\eqref{e:qcob2} to obtain 
1915: $\sigma_j = 0$ for
1916: $j > n$.
1917: Inverting equation~\eqref{e:qcob2} formally, we get
1918: \begin{equation*}
1919: \sigma_1^a = \sum_{j=0}^a S(a,j) \sigma_j
1920: \end{equation*}
1921: where $S(a,j)$ satisfies the recursion:
1922: $S(a,j) = (1+q^{2n-j})[j] S(a-1,j) + q^{j-1} S(a-1,j-1)$ 
1923: with $S(a,1) = (1+q^{2n-1})^{a-1}$ and 
1924: $S(a,a) = q^{\binom{a}{2}}$. 
1925: %Since $\sigma_n = \sigma_{n-1}$ and $\sigma_j = 0$ for $j > n$,
1926: %the inversion above is not unique. But it is the most natural one to
1927: %consider. 
1928: The numbers $S(a,j)$ give a $q$-analogue to the signed
1929: Stirling numbers.
1930: Note that they now depend on $n$.
1931: We do not know whether they have been considered before or
1932: whether they have an explicit formula.
1933: 
1934: \medskip
1935: \noindent
1936: {\bf The orthogonal case.}
1937: Let $V$ be a (even dimensional) vector space over $\fq$ 
1938: with a symmetric non-degenerate bilinear form $Q$.
1939: We also assume that $V$ has an isotropic subspace of dimension $n$.
1940: More explicitly, $V$ has a basis
1941: $e_1, e_2, \ldots, e_n, f_1, f_2, \ldots, f_n$ 
1942: such that $Q(e_i,f_i) = Q(f_i,e_i) = 1$
1943: and all other pairings between basis vectors are zero.
1944: The orthogonal group $O_{2n}(\fq)$ 
1945: is the group of automorphisms of $V$ 
1946: that preserve the bilinear form $Q$.
1947: 
1948: The building of type $B_{n}$
1949: associated to the orthogonal group
1950: is the flag (order) complex of all isotropic subspaces of $V$.
1951: We will denote it by $\Delta(B_n)$.
1952: The analysis parallels the symplectic case.
1953: With the same definitions of the $\sigma_j$'s,
1954: we get
1955: $\sigma_j \sigma_1 = \sigma_1 \sigma_j = 
1956: (1+q^{2n-j-1})[j] 
1957: \sigma_j + q^j \sigma_{j+1}$ if $j \leq n-1$ and 
1958: $\sigma_{n} \sigma_1 = \sigma_{1} \sigma_{n} = (1+q^{n-1})[n] \sigma_{n}$.
1959: The reason for the difference is that now
1960: the number of summands in $\sigma_{j}$ is
1961: $(1+q^{n-1})[n](1+q^{n-2})[n-1] \ldots (1+q^{n-j})[n-j+1]$.
1962: The rest is similar to the symplectic case.
1963: We only point out that the Stirling numbers now satisfy the recursion:
1964: $S(a,j) = (1+q^{2n-j-1})[j] S(a-1,j) + q^{j-1} S(a-1,j-1)$ 
1965: with $S(a,1) = (1+q^{2n-2})^{a-1}$ and 
1966: $S(a,a) = q^{\binom{a}{2}}$. 
1967: 
1968: \begin{remark}
1969: The building $\Delta(B_n)$ that we considered here is not thick.
1970: Classically, the group $O_{2n}(\fq)$ is considered 
1971: to be of type $D_n$ since there is a thick building $\Delta(D_n)$
1972: associated to it. We will discuss this next.
1973: For more information on this, see~\cite[pgs 123-127]{\brown}.
1974: \end{remark}
1975: 
1976: 
1977: 
1978: \subsection{The $q$-side shuffle for $D_n$} \label{subs:qdss}
1979: 
1980: Let $V$ and $Q$ be as in the orthogonal case of Section~\ref{subs:qbss}.
1981: The building of type $D_n$,
1982: which we denote by $\Delta(D_n)$,
1983: is the flag complex of the following so-called ``oriflamme geometry''.
1984: The vertices in $\Delta(D_n)$ are the non-zero isotropic subspaces of $V$
1985: of dimension $\not=n-1$.
1986: Two such subspaces are called incident 
1987: if one is contained in the other or
1988: if both have dimension $n$ and their intersection has dimension $n-1$.
1989: A simplex in $\Delta(D_n)$ is a set of pairwise incident vertices.
1990: 
1991: There is a natural product preserving map 
1992: $\Delta(B_n) \rightarrow \Delta(D_n)$,
1993: which generalises the map $\Sigma(B_n) \rightarrow \Sigma(D_n)$
1994: in Section~\ref{subs:map1}.
1995: To describe this map,
1996: we identify the vector space $V$ and the bilinear form $Q$
1997: in the two cases.
1998: A vertex in $\Delta(B_n)$ is a non-zero isotropic subspace of $V$.
1999: If it has dimension $\not = n-1$
2000: then we map it to itself.
2001: Otherwise there are exactly two maximal isotropic subspaces 
2002: of dimension $n$ in $V$ that contain it.
2003: And we map it to the edge joining the two vertices in $\Delta(D_n)$
2004: representing these two subspaces.
2005: This again explains how a vertex of type $s_{n-1}$ 
2006: maps to an edge of type $uv$. 
2007: 
2008: In the orthogonal case of Section~\ref{subs:qbss},
2009: we have the shuffle algebra $k[\sigma_1]$ in $\Delta(B_n)$.
2010: The image of this shuffle algebra under the above map
2011: yields a shuffle algebra in $\Delta(D_n)$.
2012: As a $q$-analogue of the map between side shuffles of 
2013: type $B_n$ and $D_n$,
2014: we get a map
2015: \begin{displaymath}
2016: \begin{array}{ccc}
2017: \left\{
2018: \begin{array}{c}
2019: \text {$q$-side shuffle}\\
2020: \text {of type $B_n$}
2021: \end{array}
2022: \right\}&\longrightarrow&\left\{
2023: \begin{array}{c}
2024: \text {$q$-side shuffle}\\
2025: \text {of type $D_n$}\\
2026: \end{array}
2027: \right\}\\
2028: \sigma_j & \mapsto & \sigma_j \\
2029: \end{array}
2030: \end{displaymath}
2031: %
2032: The $\sigma_j$'s on the right are defined 
2033: as for the side shuffle of type $D_n$ and 
2034: we again get the relation $\sigma_n = 2 \sigma_{n-1}$.
2035: The basic relation satisfied by $\sigma_1$ is 
2036: $\sigma_1(\sigma_1- (1+q^{2n-2})[1])\ldots(\sigma_1-(1+q^{n+1})[n-2])(\sigma_1-(1+q^{n-1})[n]) = 0$.
2037: Note that the term corresponding to $n-1$ is absent.
2038: The rest of the analysis is routine and we omit it.
2039: 
2040: \begin{remark}
2041: In light of the discussion in Section~\ref{s:maps},
2042: we may say the following.
2043: The failure to get a $q$-two sided shuffle of type $A_{n-1}$
2044: is the result of our failure to define a product preserving map
2045: $\Delta(D_n) \rightarrow \Delta(A_{n-1})$.
2046: Also we failed to get a $q$-riffle shuffle of type $B_n$
2047: and as a result also failed on the other two fronts.
2048: \end{remark}
2049: 
2050: 
2051: 
2052: 
2053: \vanish{
2054: The element of interest is $\sigma = \sigma_{s_1}$. 
2055: In addition to $\sigma = \sigma_1$,
2056: we define $\sigma_j = \sigma_{J_j}$ where $J_j = \{s_1,\ldots,s_j\}
2057: \subseteq I$ for $j = 1,\ldots,n-2$. 
2058: Next we set 
2059: $\sigma_{n-1} = \sigma_{\{s_1,\ldots,s_{n-1},u,v\}}$, that is,
2060: $\sigma_{n-1}$ is the sum of all chambers of $\Delta$.
2061: It turns out that
2062: Hence we may write the following without ambiguity. 
2063: \begin{equation}\label{e:qcob3}
2064: q^{\binom{j+1}{2}}\sigma_{j+1} = \sigma_1(\sigma_1-(1+q^{2n-2})[1])\ldots(\sigma_1-(1+q^{2n-j-1})[j]).
2065: \end{equation}
2066: This along with
2067: $\sigma_{n-1} \sigma_1 = (1+q^{n-1})[n] \sigma_{n-1}$,
2068: implies that both
2069: $\sigma_0=1$,$\sigma_1$,\ldots,$\sigma_{n-1}$ and
2070: $\sigma_1^0=1$,$\sigma_1^1$,\ldots,$\sigma_1^{n-1}$
2071: form a basis for the associative algebra $A = k[\sigma_1]$.
2072: We conclude that
2073: $A \isoto \frac{k[x]}{x(x-(1+q^{2n-2})[1])\ldots(x-(1+q^{n+1})[n-2])(x-(1+q^{n-1})[n])}$.
2074: This shows that $A$ is split semisimple.
2075: 
2076: As before,
2077: we extend the definition of $\sigma_j$ to any $j$
2078: using~\eqref{e:qcob} to obtain 
2079: 
2080: $\sigma_j = 0$ for
2081: $j > n$.
2082: Inverting equation~\eqref{e:qcob3} formally, we get
2083: \begin{equation*}
2084: \sigma_1^a = \sum_{j=0}^a S(a,j) \sigma_j
2085: \end{equation*}
2086: where $S(a,j)$ satisfies 
2087: %Since $\sigma_n = \sigma_{n-1}$ and $\sigma_j = 0$ for $j > n$,
2088: %the inversion above is not unique. But it is the most natural one to
2089: %consider. 
2090: The numbers $S(a,j)$ give another $q$-analogue to the signed
2091: Stirling numbers different from the one in the previous section.
2092: }
2093: 
2094: 
2095: \section{Future prospects}
2096: 
2097: We conclude by suggesting some problems for future consideration.
2098: 
2099: \subsection*{Shuffle algebras}
2100: One problem is to classify them 
2101: with minor modifications of the definition if necessary
2102: (see remark in Section~\ref{subs:defn}). 
2103: Also it is natural to ask whether there is a sensible
2104: generalisation to the case of infinite Coxeter groups.
2105: 
2106: \subsection*{Buildings}
2107: There is no good theory for random walks on buildings 
2108: the major difficulty being the non-associativity of the product. 
2109: Our methods in Section~\ref{s:q} suggest that it would be worthwhile
2110: to systematically understand 
2111: the power associative elements of a building
2112: and also to classify product preserving maps between buildings. 
2113: 
2114: 
2115: \subsection*{Complex reflection groups}
2116: In this paper, we considered only real reflection groups.
2117: The difficulty in passing to the complex case 
2118: is the absence of an analogue of the Coxeter complex.
2119: As an example, consider the complex reflection group
2120: $S_n \ltimes \Z_r^n$.
2121: For $r=2$, it specialises to the Coxeter group of type $B_n$. 
2122: The first step would be to generalise the semigroup
2123: of ordered partitions of type $B_n$. 
2124: However, it is not clear how to do this.
2125: 
2126: \subsection*{Multiplicities and derangement numbers}
2127: In \cite{\ken}, Brown defines a random walk associated to matroids.
2128: And he relates the eigenvalue multiplicities to
2129: invariants of the lattice of flats.
2130: He calls these invariants the generalised derangement numbers.
2131: The motivating examples are the side and $q$-side shuffle of type $A_{n-1}$
2132: and the multiplicities then are the usual and $q$-derangement numbers.
2133: In this sense, we may think of ordinary matroids as related to
2134: the Coxeter group of type $A_{n-1}$.
2135: 
2136: There is a notion of a $W$-matroid~\cite{\borovik} for any 
2137: Coxeter group $W$. We ask whether it is possible to generalise the above 
2138: to this setting.
2139: As a positive result in this direction,
2140: Bidigare \cite[pgs 147-148]{\bid} shows that 
2141: for the side shuffle of type $B_n$,
2142: the multiplicities are the signed derangement numbers.
2143: 
2144: 
2145: 
2146: 
2147: \vanish{
2148: 
2149: Expect Tsetlin, riffle shuffle, etc.
2150: }
2151: 
2152: 
2153: 
2154: 
2155: 
2156: %\input e
2157: 
2158: %\newpage
2159: 
2160: %\input app
2161: \appendix
2162: 
2163: \section{The hyperplane face semigroup} \label{app:hyperplane}
2164: 
2165: Most of this section and a part of the next is taken directly from~\cite{\ken}.
2166: More details concerning the material reviewed here can be found in
2167: \cite{\bhr,\bbd,\red,\brown,\bd,\ot,\zie}.  Throughout this section
2168: $\A=\{H_i\}_{i\in I}$ denotes a finite set of affine hyperplanes in
2169: $V=\R^n$.  Let $H_i^+$ and $H_i^-$ be the two open halfspaces
2170: determined by $H_i$; the choice of which one to call $H_i^+$ is
2171: arbitrary but fixed.
2172: 
2173: \subsection{Faces and chambers}
2174: 
2175: The hyperplanes $H_i$ induce a partition of $V$ into convex sets
2176: called \emph{faces} (or \emph{relatively open faces}).  These are the
2177: nonempty sets $F \subseteq V$ of the form
2178: \[
2179: F = \bigcap_{i\in I} H_i^{\varepsilon_i},
2180: \]
2181: where $\varepsilon_i \in \{+,-,0\}$ and $H^0_i = H_i$.  Equivalently, if we
2182: choose for each $i$ an affine function $f_i\colon V \rightarrow \R$
2183: such that $H_i$ is defined by $f_i = 0$, then a face is a nonempty set
2184: defined by equalities and inequalities of the form $f_i > 0$, $f_i <
2185: 0$, or $f_i = 0$, one for each $i \in I$.  The sequence $\varepsilon =
2186: (\varepsilon_i)_{i\in I}$ that encodes the definition of $F$ is called the
2187: \emph{sign sequence} of $F$ and is denoted $\varepsilon(F)$.
2188: 
2189: The faces such that $\varepsilon_i \ne 0$ for all $i$ are called
2190: \emph{chambers}.  They are convex open sets that partition the
2191: complement $V - \bigcup_{i\in I} H_i$.  In general, a face $F$ is
2192: open relative to its \emph{support}, which is defined to be the affine
2193: subspace
2194: \[
2195: \supp F = \bigcap_{\varepsilon_i (F) = 0} H_i.
2196: \]
2197: Since $F$ is open in $\supp F$, we can also describe $\supp F$ as the
2198: affine span of~$F$.
2199: 
2200: \subsection{The face relation} 
2201: 
2202: The \emph{face poset} of $\A$ is the set $\F$ of faces, ordered as
2203: follows:  $F \le G$ if for each $i \in I$ either $\varepsilon_i(F) = 0$ or
2204: $\varepsilon_i(F) = \varepsilon_i(G)$.  In other words, the description of $F$
2205: by linear equalities and inequalities is obtained from that of $G$ by
2206: changing zero or more inequalities to equalities.
2207: We say that $F$ is a face of $G$.
2208: Note that the chambers are precisely the maximal elements 
2209: of the face poset.
2210: We denote the set of chambers by $\C$.
2211: 
2212: \subsection{Product}
2213: The set $\F$ of faces is also a semigroup.  Given $F,G \in \F$, their
2214: \emph{product} $FG$ is the face with sign sequence
2215: \[
2216: \varepsilon_i(FG) = 
2217: \begin{cases}
2218:   \varepsilon_i(F)  &\text{if $\varepsilon_i(F) \ne 0$} \\
2219:          \varepsilon_i(G) &\text{if $\varepsilon_i(F) = 0$.}
2220: \end{cases}
2221: \]
2222: This has a geometric interpretation:  If we move on a straight line
2223: from a point of~$F$ toward a point of $G$, then $FG$ is the face we
2224: are in after moving a small positive distance.  
2225: Hence, we may think of $FG$ as the projection of $G$ on $F$.
2226: Notice that the face
2227: relation can be described in terms of the product:  One has
2228: \begin{equation} \label{e:poset}
2229: F\le G \iff FG=G.
2230: \end{equation}
2231: Also note that the set of chambers $\C$ is an ideal of $\F$.
2232: 
2233: \subsection{The semilattice of flats}
2234: 
2235: A second poset associated with the arrangement $\A$ is the
2236: \emph{semilattice of flats}, also called the \emph{intersection
2237: semilattice}, which we denote by~$\L$.  It consists of all nonempty
2238: affine subspaces $X \subseteq V$ of the form $X = \bigcap_{H\in\A'}
2239: H$, where $\A' \subseteq \A$ is an arbitrary subset (possibly empty).
2240: We order $\L$ by inclusion.  [Warning: Many authors order $\L$ by
2241: reverse inclusion.]  Notice that any two elements $X,Y$ have a
2242: least upper bound $X\vee Y$ in~$\L$, which is the intersection of
2243: all hyperplanes $H\in\A$ containing both $X$ and~$Y$; hence $\L$
2244: is an \emph{upper semilattice} (poset with least upper bounds).  It is
2245: a lattice if the arrangement $\A$ is \emph{central}, i.e., if
2246: $\bigcap_{H\in\A} H \ne\emptyset$.  Indeed, this intersection is then
2247: the smallest element of~$\L$, and a finite upper semilattice with a
2248: smallest element is a lattice \cite[Section 3.3]{\stanley}.  The
2249: support map gives a surjection
2250: \[
2251: \supp\colon\F\onto\L,
2252: \]
2253: which preserves order and also behaves nicely with respect to the
2254: semigroup structure.  Namely, we have
2255: \begin{equation}
2256: \supp(FG) = \supp F \vee \supp G
2257: \end{equation}
2258: and
2259: \begin{equation}
2260: FG = F \iff \supp G \le \supp F.
2261: \end{equation}
2262: 
2263: \subsection{The forgetful map} \label{subs:map}
2264: Let $\A=\{H_i\}_{i\in I}$ and $\A^{\prime}=\{H_i\}_{i\in I^{\prime}}$ be two hyperplane
2265: arrangements such that $I^{\prime} \subset I$, that is, the arrangement $\A^{\prime}$ is
2266: obtained from $\A$ by deleting some of the hyperplanes.
2267: Let $\F(A)$ and $\F(A^{\prime})$ be the respective hyperplane face
2268: semigroups. Then there is a natural map $\F(A) \rightarrow \F(A^{\prime})$.
2269: An element of $\F(A)$ is a face $F$ with a sign sequence, say 
2270: $\varepsilon = (\varepsilon_i)_{i\in I}$. We map it to the face of
2271: $\A^{\prime}$ with the sign sequence $\varepsilon^{\prime} = (\varepsilon_i)_{i\in I^{\prime}}$.
2272: In other words, we forget the signs of the elements of $I$ that are not in
2273: $I^{\prime}$.
2274: It is clear that this map is a semigroup homomorphism and preserves
2275: the face relation. The second fact is implied by the first in view of
2276: equation~\eqref{e:poset}. 
2277: 
2278: \subsection{Spherical representation} \label{sub:cell}
2279: 
2280: Suppose now that $\A$ is a \emph{central} arrangement, i.e., that the
2281: hyperplanes have a nonempty intersection.  We may assume that this
2282: intersection contains the origin.  Suppose further that $\bigcap_{i\in
2283: I} H_i = \{0\}$, in which case $\A$ is said to be \emph{essential}.
2284: (There is no loss of generality in making this assumption; for if it
2285: fails, then we can replace $V$ by the quotient space $V/\bigcap_i H_i$.)
2286: The hyperplanes then induce a cell-decomposition of the unit sphere,
2287: the cells being the intersections with the sphere of the faces
2288: $F\in\F$.  Thus $\F$, as a poset, can be identified with the poset of
2289: cells of a regular cell-complex $\Sigma$, homeomorphic to a sphere.
2290: Note that the face $F=\{0\}$, which is the identity of the
2291: semigroup~$\F$, is not visible in the spherical picture; it
2292: corresponds to the empty cell.  The cell-complex $\Sigma$ plays a
2293: crucial role in~\cite{\bd}, to which we refer for more details.
2294: 
2295: Thus the cell-complex $\Sigma$ is a semigroup 
2296: which we call the hyperplane face semigroup.
2297: The maximal cells (which we again call chambers and denote $\C$)
2298: is an ideal in $\Sigma$.
2299: 
2300: 
2301: 
2302: \section{Reflection arrangements} \label{app:Coxeter}
2303: 
2304: We work with an
2305: arbitrary finite Coxeter group~$W$ and its associated hyperplane face
2306: semigroup $\Sigma$ (the Coxeter complex of~$W$).  
2307: But we will 
2308: explain everything in concrete terms for the cases $W=A_{n-1},B_n,D_n$.
2309: This should make the discussion accessible to readers unfamiliar with
2310: Coxeter groups.
2311: 
2312: 
2313: 
2314: \subsection{Finite reflection groups} \label{sub:reflection}
2315: 
2316: We begin with a very quick review of the basic facts that we need
2317: about finite Coxeter groups and their associated simplicial complexes
2318: $\Sigma$.  Details can be found in many places, such as
2319: \cite{\brown,\gb,\humphreys,\tits}.  A \emph{finite reflection group}
2320: on a real inner-product space $V$ is a finite group of orthogonal
2321: transformations of $V$ generated by reflections $s_H$ with respect to
2322: hyperplanes $H$ through the origin.  The set of hyperplanes $H$ such
2323: that $s_H\in W$ is the \emph{reflection arrangement} associated
2324: with~$W$.  Its hyperplane face semigroup $\Sigma$ is called the
2325: \emph{Coxeter complex} of~$W$.  Geometrically, this complex is obtained
2326: by cutting the unit sphere in~$V$ by the hyperplanes~$H$, as in
2327: Section~\ref{sub:cell}.  (As explained there, one might have to first
2328: pass to a quotient of~$V$.)  
2329: It turns out that the Coxeter complex $\Sigma$
2330: is always a simplicial complex.
2331: Furthermore, the action of~$W$ on~$V$ induces an action of
2332: $W$ on~$\Sigma$, and this action is simply-transitive on the chambers.
2333: Thus the set $\C$ of chambers can be identified with~$W$, once a
2334: ``fundamental chamber'' $C$ is chosen.
2335: 
2336: 
2337: \subsection{Types of simplices}
2338: 
2339: The number $r$ of vertices of a chamber of $\Sigma$ is called the
2340: \emph{rank} of $\Sigma$ (and of $W$); thus the dimension of $\Sigma$
2341: as a simplicial complex is $r-1$.  It is known that one can color the
2342: vertices of $\Sigma$ with $r$ colors in such a way that vertices
2343: connected by an edge have distinct colors.  The color of a vertex is
2344: also called its \emph{label}, or its \emph{type}, and we denote by $I$
2345: the set of all types.  We can also define $\type(F)$ for any
2346: $F\in\Sigma$; it is the subset of~$I$ consisting of the types of the
2347: vertices of~$F$.  For example, every chamber has type $I$, while the
2348: empty simplex has type $\emptyset$.  The action of $W$ is
2349: type-preserving; moreover, two simplices are in the same $W$-orbit if
2350: and only if they have the same type.  
2351: 
2352: \subsection{The Coxeter diagram}\label{subs:cd}
2353: Choose a fundamental chamber $C$. It is known that the
2354: reflections $s_i$ in the facets of $C$ generate $W$.
2355: In fact, $W$ has a presentation of the form
2356: $<s_1,\ldots,s_r \mid (s_i s_j)^{m_{ij}}>$
2357: with $m_{ii}=1$ and ${m_{ij}}={m_{ji}}\geq2$.
2358: This data is conveniently encoded in a picture called the Coxeter
2359: diagram of $W$. This diagram is a graph, with vertices and edges,
2360: defined as follows: There are $r$ vertices, one for each generator
2361: $i=1,2,\ldots,r$, and the vertices corresponding to $i$ and $j$ are
2362: connected by an edge if and only if $m_{ij}\geq3$. If $m_{ij}\geq4$
2363: then we simply label the edge with the number $m_{ij}$. 
2364: The figures show the Coxeter diagrams which are of
2365: interest to us, namely the ones of type $A_{n-1}$, $B_n$ and $D_n$.
2366: It is customary to use the generators of $W$, or the vertices of the
2367: Coxeter diagram to label the vertices of
2368: its Coxeter complex $\Sigma$. 
2369: A vertex of the fundamental chamber $C$ is labelled $s_i$ if
2370: it is fixed by all the fundamental reflections except $s_i$.
2371: Since $W$ acts transitively on $\C$ and the action is type-preserving,
2372: this determines the type of all the vertices of $\Sigma$.
2373: 
2374: 
2375: %% uses pictex macros:  prepictex, pictex, postpictex
2376: %% for Latex 2.09
2377: %%%Dynkin diagram of type A_l %%%
2378: \beginpicture
2379: \setcoordinatesystem units <1mm, 1mm>
2380: \setplotarea x from 0 to 80, y from -10 to 20
2381: %%%  caption  %%
2382: \put {\small Figure 2.1: Coxeter Diagram of Type $A_{n-1}$} 
2383:   at  40 -5
2384: %\put { $\scriptstyle {1}$}  at 10 15
2385: \circulararc 360 degrees from 11 10 center at 10 10
2386: \put  {$\scriptstyle {s_1}$} at 10 5
2387: \setlinear
2388: \plot 11 10  29 10 /
2389: %\put { $\scriptstyle {1}$}  at 30 15
2390: \circulararc 360 degrees from 31 10 center at 30 10
2391: \put  {$\scriptstyle{s_2}$} at 30 5
2392: \plot 31 10  45 10 /
2393: \put {${\bf \ldots}$} at 50 10
2394: \plot 55 10 69 10 /
2395: %\put { $\scriptstyle {1}$}  at 70 15
2396: \circulararc 360 degrees from 71 10 center at 70 10
2397: \put {$\scriptstyle{s_{n-1}}$} 
2398:  at 70 5
2399: \endpicture
2400: 
2401: 
2402: 
2403: 
2404: 
2405: %% uses pictex macros:  prepictex, pictex, postpictex
2406: %% for Latex 2.09
2407: %%%Dynkin diagram of type B_l %%%
2408: \beginpicture
2409: \setcoordinatesystem units <1mm, 1mm>
2410: \setplotarea x from 0 to 80, y from -10 to 20
2411: %%%  caption  %%
2412: \put {\small Figure 2.2: Coxeter Diagram of Type $B_n$} 
2413:   at  50 -5
2414: %\put { $\scriptstyle{1}$}  at 10 15
2415: \circulararc 360 degrees from 11 10 center at 10 10
2416: \put  { $\scriptstyle{s_1}$} at 10 5
2417: \setlinear
2418: \plot 11 10  29 10 /
2419: %\put { $\scriptstyle{2}$}  at 30 15
2420: \circulararc 360 degrees from 31 10 center at 30 10
2421: \put  { $\scriptstyle{s_2}$} at 30 5
2422: \plot 31 10  45 10 /
2423: \put {${\bf \ldots}$} at 50 10
2424: \plot 55 10 69 10 /
2425: %\put { $\scriptstyle{2}$}  at 70 15
2426: \circulararc 360 degrees from 71 10 center at 70 10
2427: \put {$\scriptstyle{s_{n-1}}$} at 70 5
2428: \vanish{
2429: %%%% double bond %%%%
2430: \plot 71 10.5 88 10.5 /
2431: \plot 71 9.5  88 9.5  /
2432: %%% arrow for root size  %%%
2433: \plot 89 10  87 12 /
2434: \plot 89 10  87  8 /
2435: }
2436: \setlinear
2437: \plot 71 10 89 10 /
2438: %\put { $\scriptstyle{2}$}  at 90 15
2439: \put { $\scriptstyle{4}$}  at 80 15
2440: \circulararc 360 degrees from 91 10 center at 90 10
2441: \put {$\scriptstyle{s_n=t}$} 
2442:  at 90 5
2443: \endpicture
2444: 
2445: 
2446: %% uses pictex macros:  prepictex, pictex, postpictex
2447: %% for Latex 2.09
2448: %%%Dynkin diagram of type C_l %%%
2449: \beginpicture
2450: \setcoordinatesystem units <1mm, 1mm>
2451: \setplotarea x from 0 to 80, y from -10 to 25
2452: %%%  caption  %%
2453: \put {\small Figure 2.3: Coxeter Diagram of Type $D_n$} 
2454:   at  50 -10
2455: %\put { $\scriptstyle{2}$}  at 10 15
2456: \circulararc 360 degrees from 11 10 center at 10 10
2457: %\put  { $\scriptstyle{\varepsilon_1-\varepsilon_2}$} at 10 5
2458: \put  { $\scriptstyle{s_1}$} at 10 5
2459: \setlinear
2460: \plot 11 10  29 10 /
2461: %\put { $\scriptstyle{2}$}  at 30 15
2462: \circulararc 360 degrees from 31 10 center at 30 10
2463: %\put  { $\scriptstyle{\varepsilon_2-\varepsilon_3}$} at 30 5
2464: \put  { $\scriptstyle{s_2}$} at 30 5
2465: \plot 31 10  45 10 /
2466: \put {${\bf \ldots}$} at 50 10
2467: \plot 55 10 69 10 /
2468: %\put { $\scriptstyle{2}$}  at 70 15
2469: \circulararc 360 degrees from 71 10 center at 70 10
2470: %\put { $\scriptstyle{\varepsilon_{l-1}-\varepsilon_l}$} 
2471: % at 70 5
2472: \put { $\scriptstyle{s_{n-2}}$} 
2473:  at 70 5
2474: %%%% double bond %%%%
2475: %\plot 72 10.5  89 10.5 /
2476: %\plot 72  9.5  89  9.5 /
2477: %%% arrow for root size  %%%
2478: %\plot 71 10  73 12 /
2479: %\plot 71 10  73  8 /
2480: %\put { $\scriptstyle{1}$}  at 90 15
2481: %\circulararc 360 degrees from 91 10 center at 90 10
2482: %\put { $\scriptstyle{2\varepsilon_l}$} 
2483: % at 90 5
2484: \circulararc 360 degrees from 86 20 center at 85 20
2485: \circulararc 360 degrees from 86 0 center at 85 0
2486: \plot 71 10.5  84 19.5 /
2487: \plot 71 9.5  84 0.5 /
2488: \put { $\scriptstyle{u}$}  at 87 16
2489: \put { $\scriptstyle{v}$}  at 87 -4
2490: \endpicture
2491: 
2492: 
2493: 
2494: \subsection{The Coxeter group of type $A_{n-1}$} \label{subs:a}
2495: 
2496: The Coxeter group $W=S_n$ acts on $\R^n$ by permuting the
2497: coordinates. 
2498: The arrangement in this case is the \emph{braid
2499: arrangement} in $\R^n$. It is discussed in detail in
2500: \cite{\bid,\bhr,\bbd,\bd}. 
2501: It consists of the $\binom{n}{2}$
2502: hyperplanes $H_{ij}$ defined by $x_i=x_j$, where $1\le i<j\le n$.
2503: Each chamber is determined by an ordering of the coordinates, so it
2504: corresponds to a permutation.  
2505: The faces of a chamber are obtained by
2506: changing to equalities some of the inequalities defining that chamber.
2507: 
2508: We fix  
2509: $x_1<x_2<\ldots<x_n$ 
2510: to be the fundamental chamber $C$.
2511: The supports of the facets of $C$ are hyperplanes of the form
2512: $x_i=x_{i+1}$, where $1\le i \le n-1$.
2513: The reflection in the hyperplane $x_i=x_{i+1}$
2514: correponds to the generator $s_i$ that
2515: interchanges the coordinates $x_i$
2516: and $x_{i+1}$.
2517: The chamber $C$
2518: has $n-1$ vertices, namely
2519: 
2520: $s_1 : x_1<x_2=\ldots=x_n$, 
2521: 
2522: $s_2 : x_1=x_2<x_3=\ldots=x_n$,$\ldots$, 
2523: 
2524: $s_{n-1} : x_1=\ldots=x_{n-1}<x_n$.
2525: 
2526: \noindent The labels $s_1$,$s_2$,$\ldots$,$s_{n-1}$ are assigned by the rule
2527: mentioned in subsection~\ref{subs:cd}. Applying the action of $W$ we see,
2528: for example, that $x_{\pi(1)}<x_{\pi(2)}=\ldots=x_{\pi(n)}$ gives all
2529: vertices of type $s_1$ as $\pi$ varies over all permutations of~$[n]=\{1,\dots,n\}$. 
2530: \vanish{
2531: When $n=4$, for example, one of the 24
2532: chambers is the region defined by $x_2>x_3>x_1>x_4$, corresponding to
2533: the permutation 2314.  
2534: For
2535: example, the chamber $x_2>x_3>x_1>x_4$ has a face given by
2536: $x_2>x_3>x_1=x_4$, which is also a face of the chamber
2537: $x_2>x_3>x_4>x_1$.
2538: }
2539: 
2540: We encode the system of equalities and inequalities
2541: defining a face~$F$ by an ordered partition $(B_1,\dots,B_k)$ of
2542: $[n]$.  Here $B_1,\dots,B_k$ are disjoint nonempty sets
2543: whose union is $[n]$, and their order counts.  
2544: \vanish{
2545: For example, the face
2546: $x_2>x_3>x_1=x_4$ corresponds to the 3-block ordered partition
2547: $(\{2\},\{3\},\{1,4\})$, and the face $x_2>x_1=x_3=x_4$ corresponds to
2548: the 2-block ordered partition $(\{2\}, \{1,3,4\})$.
2549: }
2550: Thus the simplices of $\Sigma$ are ordered partitions $B=(B_1,\dots,B_l)$ 
2551: of the set $[n]$. 
2552: For example, the vertices of type $s_1$ are ordered two block partitions
2553: such that the first block is a singleton.
2554: 
2555: The product in $\Sigma$ is also easy to
2556: describe. We multiply two ordered
2557: partitions by taking intersections and ordering them
2558: lexicographically; more precisely, if $B=(B_1,\dots,B_l)$ and
2559: $C=(C_1,\dots,C_m)$, then
2560: \[
2561: BC=(B_1\cap C_1,\dots,B_1\cap C_m,\dots, B_l\cap C_1,\dots,B_l\cap
2562: C_m)\sphat\,,
2563: \]    
2564: where the hat means ``delete empty intersections''. The 1-block
2565: ordered partition is the identity.  
2566: 
2567: 
2568: Note that $B$ is a face of $C$ if and only if $C$ consists of
2569: an ordered partition of $B_1$ followed by an ordered partition
2570: of~$B_2$, and so on, that is, if and only if $C$ is a refinement of
2571: $B$. The chambers are the ordered partitions into
2572: singletons, so they correspond to the permutations of~$[n]$ or a deck
2573: of $n$ cards.
2574: 
2575: \vanish{
2576: It is useful to have a second description of $\B$.  Ordered partitions
2577: $(B_1,\dots,B_l)$ of~$[n]$ are in 1--1 correspondence with chains of
2578: subsets $\emptyset=E_0<E_1<\cdots<E_l=[n]$, the correspondence being
2579: given by $B_i=E_i-E_{i-1}$.  So we may identify $\B$ with the set of
2580: such chains.  
2581: The
2582: vertices are the proper nonempty subsets $X\subset[n]=\{1,\dots,n\}$,
2583: and the simplices are the chains of such subsets.  The $S_n$-action is
2584: induced by the action of $S_n$ on~$[n]$. The
2585: chambers of~$\Sigma$ correspond to permutations $w$ of $[n]$, with $w$
2586: corresponding to the chamber
2587: \[
2588: \{w(1)\}<\{w(1),w(2)\}<\cdots<\{w(1),w(2),\dots,w(n-1)\}.
2589: \]
2590: This is the same as the identification of $\C$ with $S_n$ that results
2591: from choosing
2592: \[
2593: \{1\}<\{1,2\}<\cdots<\{1,2,\dots,n-1\}
2594: \]
2595: as fundamental chamber.
2596: The product is then described as follows:  Given a chain
2597: $E$ as above and a second chain $F\colon\emptyset=F_0<F_1<\cdots<
2598: F_m=[n]$, their product $EF$ is obtained by using $F$ to refine~$E$.
2599: More precisely, consider the sets $G_{ij}=(E_{i-1}\cup F_j)\cap E_i =
2600: E_{i-1}\cup (F_j\cap E_i)$.  For each $i=1,2,\dots,l$ we have
2601: \[
2602: E_{i-1} = G_{i0}\subseteq G_{i1}\subseteq\cdots\subseteq G_{im}=E_i.
2603: \]
2604: Deleting repetitions gives a chain from $E_{i-1}$ to~$E_i$, and
2605: combining these for all~$i$ gives the desired refinement $EF$ of~$E$.
2606: 
2607: This construction is used in one of the standard proofs of the
2608: Jordan--H\"{o}lder theorem.
2609: }
2610: 
2611: \subsection{The Coxeter group of type $B_n$} \label{subs:b}
2612:  
2613: The group of signed permutations $W=S_n \ltimes \Z_2^n$ acts on $\R^n$
2614: with the subgroup $S_n$ permuting the
2615: coordinates and the subgroup $\Z_2^n$ flipping the signs of the
2616: coordinates. 
2617: The reflection arrangement in this case consists of the hyperplanes
2618: defined by $x_i= \pm x_j$ and $x_i = 0$, where $1\le i<j\le n$.
2619: A chamber is given by an ordering of the coordinates, their negatives
2620: and zero. For example, for $n = 3$, 
2621: $x_2<-x_3<x_1<0<-x_1<x_3<-x_2$ specifies a chamber.
2622: Note that the inequalities that appear on the left of $0$ completely
2623: determine a chamber (and the same is true for any face since it is obtained by changing to equalities some of the inequalities defining a chamber).
2624: Thus we see that a chamber corresponds to a signed permutation.  
2625: 
2626: 
2627: We fix
2628: $x_1<x_2<\ldots<x_n<0$ 
2629: to be the fundamental chamber $C$. 
2630: The supports of the facets of $C$ are hyperplanes of the form
2631: $x_i=x_{i+1}$, where $1\le i \le n-1$ and $x_n = 0$.
2632: The generators $s_i$ interchange the coordinates $x_i$
2633: and $x_{i+1}$ and the generator $t$ flips the sign of $x_n$.
2634: The $n$ vertices of $C$ along with their labels are as follows.
2635: 
2636: $s_1:x_1<x_2=\ldots=x_n=0$, 
2637: 
2638: $s_2:x_1=x_2<x_3=\ldots=x_n=0$,$\ldots$, 
2639: 
2640: $s_{n-1}:x_1=x_2=\ldots=x_{n-1}<x_n=0$,
2641: 
2642: $t:x_1=x_2\ldots=x_n<0$.
2643: 
2644: \noindent Applying the group action, we see, for example, that the
2645: vertices of type $t$ have the form
2646: $\varepsilon_1 x_1=\varepsilon_2 x_2\ldots=\varepsilon_n
2647: x_n<0$
2648: where $\varepsilon_i \in \{\pm 1 \}$.
2649: 
2650: It is convenient to describe a face~$F$ by an anti-symmetric ordered partition
2651: $(B_1,\dots,B_k,Z,\overline{B}_k,\ldots,\overline{B}_1)$ of
2652: $[n,\overline{n}]=\{1,\dots,n,\overline{1},\ldots,\overline{n}\}$. 
2653: We will call this a {\it partition of type $B_n$}.
2654: For example, for $n = 3$, the face
2655: $x_2<-x_3<x_1=0=-x_1<x_3<-x_2$ is written
2656: $(\{2\},\{\overline{3}\},\{1,\overline{1}\},\{3\},\{\overline{2}\})$.
2657: The set $Z$ satisfies $Z = \overline{Z}$ and is allowed to be
2658: empty. We call it the zero block. It is the only set that is allowed
2659: to contain both a number and its negative. 
2660: Also, the sets $B_1,\dots,B_k$ are necessarily non-empty.
2661: We also define a \emph{weak partition of type $B_n$} to be a partition as
2662: above but where
2663: the sets $B_1,\dots,B_k$ are allowed to be empty.
2664: Note that for a face $F$, the zero block of its partition is empty if
2665: and only if the type of $F$ contains the letter 
2666: $t$. 
2667: We split the description for a face $F$
2668: into four cases, depending on whether the type of $F$ contains 
2669: \begin{itemize}
2670: \item[(i)] neither $s_{n-1}$ nor $t$: 
2671: 
2672: An ordered partition
2673: $(B_1,\dots,B_k,Z,\overline{B}_k,\ldots,\overline{B}_1)$ of
2674: $[n,\overline{n}]$ of type $B_n$ with the added
2675: restriction that the zero block $Z$ has at least $4$ elements.
2676: 
2677: \item[(ii)] $t$ but not $s_{n-1}$:
2678: 
2679: An ordered partition
2680: $(B_1,\dots,B_k,Z,\overline{B}_k,\ldots,\overline{B}_1)$ of
2681: $[n,\overline{n}]$, with the restriction that the zero block is empty
2682: and $B_k$ has at least $2$ elements. 
2683: 
2684: \item[(iii)] $s_{n-1}$ but not $t$:
2685: 
2686: An ordered partition
2687: $(B_1,\dots,B_k,Z,\overline{B}_k,\ldots,\overline{B}_1)$ of
2688: $[n,\overline{n}]$, with the restriction that the zero block has exactly
2689: $2$ elements. 
2690: 
2691: \item[(iv)] both $s_{n-1}$ and $t$:
2692: 
2693: An ordered partition
2694: $(B_1,\dots,B_k,Z,\overline{B}_k,\ldots,\overline{B}_1)$ of
2695: $[n,\overline{n}]$, with the restriction that the zero block is empty
2696: and $B_k$ has at exactly $1$ element. 
2697: 
2698: \end{itemize}
2699: %
2700: The product in $\Sigma$ is defined exactly as for the $A_{n-1}$ case,
2701: where we refine the first partition by the second. 
2702: Note that $B$ is a face of $C$ if and only if $C$ is a refinement of
2703: $B$.
2704: 
2705: 
2706: As mentioned earlier, the second half of the partition following $Z$
2707: contains no new information. Hence we may describe a $k-1$ dimensional
2708: face by a $(k+1)$ block partition 
2709: $(B_1,\dots,B_k,Z)$ of $[n]$. We think of $B_1,\dots,B_k$ as signed sets
2710: and $Z$ (possibly empty) as an unsigned set.
2711: The chambers are then $(n+1)$ block
2712: partitions into $n$ singletons and an empty zero block, so they correspond
2713: to signed permutations of~$[n]$ or a deck of $n$ signed cards.
2714: We will call such a deck as a \emph{deck of type $B_n$}.
2715: However, the product and the face relation is not so natural now. So we
2716: will mainly use the first description.
2717: 
2718: \subsection{The Coxeter group of type $D_n$} \label{subs:d}
2719: 
2720: The group of even signed permutations $W=S_n \ltimes G$ is an index 2
2721: subgroup of the group of signed permutations $S_n \ltimes \Z_2^n$.
2722: Here $G$ is the index 2 subgroup of $\Z_2^n$ consisting of $n$-tuples
2723: that have an even number of negative signs.
2724: The group $W$ acts on $\R^n$
2725: with the subgroup $S_n$ permuting the
2726: coordinates and the subgroup $G$ flipping the signs of the
2727: coordinates. The
2728: reflection arrangement in this case consists of the hyperplanes
2729: defined by $x_i= \pm x_j$, where $1\le i<j\le n$.
2730: It is obtained from the reflection arrangement of type $B_n$ by
2731: deleting the coordinate hyperplanes $x_i=0$ for $1\le i \le n$.
2732: On the level of chambers, this has the effect of merging together pairs
2733: of adjacent chambers of the Coxeter complex of type $B_n$; 
2734: see Figure~\ref{maps}.
2735: For example, the chamber of the Coxeter complex of type $D_n$,
2736: $x_1<x_2<\ldots<x_{n-1}<\pm x_n<-x_{n-1}<\ldots<-x_1$
2737: (which we fix as our fundamental chamber $C$) is the union of the two
2738: chambers of the Coxeter complex of type $B_n$, namely, 
2739: $x_1<x_2<\ldots<x_n<0$ and $x_1<x_2<\ldots<-x_n<0$.
2740: The supports of the facets of $C$ are hyperplanes of the form
2741: $x_i=x_{i+1}$, where $1\le i \le n-1$ and $x_{n-1} = - x_n$.
2742: The generators $s_i$ interchange the coordinates $x_i$
2743: and $x_{i+1}$, the generator $u$ interchanges the coordinates $x_{n-1}$
2744: and $x_{n}$ and the generator $v$ interchanges $x_{n-1}$ and $x_n$ and
2745: flips the signs of both.
2746: The $n$ vertices of $C$ along with their labels are as follows.
2747: 
2748: $s_1:x_1<x_2=\ldots=x_n=0$, 
2749: 
2750: $s_2:x_1=x_2<x_3=\ldots=x_n=0$,$\ldots$, 
2751: 
2752: $s_{n-2}:x_1=x_2=\ldots=x_{n-2}<x_{n-1}=x_n=0$,
2753: 
2754: $u:x_1=x_2\ldots=x_{n-1}=-x_n<x_n=-x_{n-1}=\ldots=-x_{2}=-x_1$,
2755: 
2756: $v:x_1=x_2\ldots=x_{n-1}=x_n<-x_n=-x_{n-1}=\ldots=-x_{2}=-x_1$.
2757: 
2758: \noindent Applying the group action, we see, for example, that the
2759: vertices of type $u$ and $v$ both have the form
2760: $\varepsilon_1 x_1=\varepsilon_2 x_2\ldots=\varepsilon_n
2761: x_n<-\varepsilon_n x_n=\ldots=-\varepsilon_2
2762: x_{2}=-\varepsilon_1 x_1$, 
2763: where $\varepsilon_i \in \{\pm 1 \}$
2764: and they are distinguished by the parity of the product 
2765: $\varepsilon_1 \varepsilon_2 \ldots \varepsilon_n$.
2766: Also observe that the edge of the fundamental chamber $C$ of type
2767: $\{u,v\}$ is given by 
2768: $x_1=x_2\ldots=x_{n-1}<\pm x_n<-x_{n-1}=\ldots=-x_{2}=-x_1$.
2769: 
2770: As in the $B_n$ case, we 
2771: describe a face~$F$ by an anti-symmetric ordered partition
2772: $(B_1,\dots,B_k,C,\overline{B}_k,\ldots,\overline{B}_1)$ of
2773: $[n,\overline{n}]$.
2774: We repeat that the sets $B_1,\dots,B_k$ are
2775: non-empty and cannot contain both a number and its negative. 
2776: We call the block $C=\overline{C}$ in the middle, the central
2777: block rather than the zero block. 
2778: It always has an even number of elements.
2779: Furthermore, we impose the relation
2780: $(B_1,\dots,B_{k-1},C,\overline{B}_{k-1},\ldots,\overline{B}_1) =
2781: (B_1,\dots,B_{k-1},B_{k},C^{\prime},\overline{B}_{k},\overline{B}_{k-1},\ldots,\overline{B}_1)$ 
2782: where $C^{\prime}$ is empty, $C$ has exactly two elements, namely, a number and
2783: its negative, and $C=B_{k} \cup C^{\prime} \cup \overline{B}_{k}$. 
2784: We split the description for a face $F$
2785: into three cases, depending on whether the type of $F$ contains 
2786: \begin{itemize}
2787: \item[(i)] neither $u$ nor $v$: 
2788: 
2789: An ordered partition
2790: $(B_1,\dots,B_k,C,\overline{B}_k,\ldots,\overline{B}_1)$ of
2791: $[n,\overline{n}]$, with the
2792: restriction that the central block $C$ (in this case, we may think of
2793: it as the zero block) has at least $4$ elements.
2794: 
2795: \item[(ii)] exactly one of $u$ and $v$:
2796: 
2797: An ordered partition
2798: $(B_1,\dots,B_k,C,\overline{B}_k,\ldots,\overline{B}_1)$ of
2799: $[n,\overline{n}]$, with the restriction that $C$ is empty and $B_k$
2800: has at least $2$ elements. 
2801: 
2802: \item[(iii)] both $u$ and $v$:
2803: 
2804: An ordered partition
2805: $(B_1,\dots,B_{k-1},C,\overline{B}_{k-1},\ldots,\overline{B}_1)$ of
2806: $[n,\overline{n}]$, with the restriction that $C$ has exactly $2$
2807: elements, namely, a number and its negative.  
2808: 
2809: An ordered partition
2810: $(B_1,\dots,B_{k-1},B_{k},C^{\prime},\overline{B}_{k},\overline{B}_{k-1},\ldots,\overline{B}_1)$ 
2811: where $C^{\prime}$ is empty and $B_{k}$ has exactly $1$ element.
2812: 
2813: \end{itemize}
2814: %
2815: Here the number $k$ is always the rank of the
2816: face. 
2817: The first three descriptions are obtained directly from the equalities
2818: and inequalities that define a face. The reason why we call $C$ the
2819: central block rather than the zero block should be clear now.
2820: The motivation for the second description in case $(iii)$ is as follows. 
2821: To give an example, the product of
2822: $(\{2,\overline{3}\},\{1,\overline{5},\overline{4}\},\{4,5,\overline{1}\},\{3,\overline{2}\})$
2823: with
2824: $(\{4,1,\overline{5}\},\{2,\overline{3},3,\overline{2}\},\{5,\overline{1},\overline{4}\})$,
2825: must be
2826: $(\{2,\overline{3}\},\{1,\overline{5}\},\{\overline{4},4\},\{5,\overline{1}\},\{3,\overline{2}\})$.
2827: But if we refine the first partition by the second then
2828: we get
2829: $(\{2,\overline{3}\},\{1,\overline{5}\},\{\overline{4}\},\{4\},\{5,\overline{1}\},\{3,\overline{2}\})$.
2830: This partition has two singletons in the middle and does not fit any
2831: of the first three descriptions given above. So we allow ourselves to merge the
2832: two singletons in the middle.
2833: This identification allows us to multiply two partitions $B$ and $C$
2834: in the same way as before; that 
2835: is, we refine $B$ using $C$. 
2836: 
2837: %We will call this a {\it partition of type $D_n$}.
2838: As in the $B_n$ case, if we throw off the (redundant) second half of
2839: the partition then we see that chambers correspond to an \emph{almost}
2840: signed deck of cards or a \emph{deck of type $D_n$}. It is a deck in
2841: which every card, except the bottomost, is signed.
2842: We also define a \emph{weak partition of type $D_n$} to be a partition 
2843: of type $D_n$ where
2844: the sets $B_1,\dots,B_k$ are allowed to be empty.
2845: 
2846: 
2847: 
2848: \subsection*{Acknowledgments} It is a pleasure to thank my advisor
2849: Ken Brown for introducing me to this area of mathematics and sharing
2850: his insights with me. I would also like to thank Marcelo Aguiar for
2851: giving valuable suggestions.
2852: 
2853: 
2854: 
2855: 
2856: 
2857: 
2858: 
2859: %\input bib
2860: 
2861: %\bibliographystyle{amsplain}
2862: %\bibliography{hyperplane}
2863: 
2864: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
2865: \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
2866: % \MRhref is called by the amsart/book/proc definition of \MR.
2867: \providecommand{\MRhref}[2]{%
2868:   \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
2869: }
2870: \providecommand{\href}[2]{#2}
2871: \begin{thebibliography}{10}
2872: 
2873: \bibitem{aigner97:_combin}
2874: Martin Aigner, \emph{Combinatorial theory}, Springer-Verlag, Berlin, 1997,
2875:   Reprint of the 1979 original. \MR{1 434 477}
2876: 
2877: \bibitem{bayerdiaconis92:_trail}
2878: Dave Bayer and Persi Diaconis, \emph{Trailing the dovetail shuffle to its
2879:   lair}, Ann. Appl. Probab. \textbf{2} (1992), no.~2, 294--313.
2880: 
2881: \bibitem{bidigare97:_hyper}
2882: T.~Patrick Bidigare, \emph{Hyperplane arrangement face algebras and their
2883:   associated {M}arkov chains}, Ph.D. thesis, University of Michigan, 1997.
2884: 
2885: \bibitem{bidigarehanlon98}
2886: T.~Patrick Bidigare, Phil Hanlon, and Daniel~N. Rockmore, \emph{A combinatorial
2887:   description of the spectrum for the {T}setlin library and its generalization
2888:   to hyperplane arrangements}, Duke Math. J. \textbf{99} (1999), no.~1,
2889:   135--174.
2890: 
2891: \bibitem{billera99:_random}
2892: Louis~J. Billera, Kenneth~S. Brown, and Persi Diaconis, \emph{Random walks and
2893:   plane arrangements in three dimensions}, Amer. Math. Monthly \textbf{106}
2894:   (1999), no.~6, 502--524. \MR{1 699 244}
2895: 
2896: \bibitem{bjoerner93:_om}
2897: Anders Bj{\"o}rner, Michel Las~Vergnas, Bernd Sturmfels, Neil White, and
2898:   G{\"u}nter~M. Ziegler, \emph{Oriented matroids}, Encyclopedia of Mathematics
2899:   and its Applications, vol.~46, Cambridge University Press, Cambridge, 1993.
2900: 
2901: \bibitem{borovik95:_coxet}
2902: Alexandre~V. Borovik and K.~Sian Roberts, \emph{Coxeter groups and matroids},
2903:   Groups of Lie type and their geometries (Como, 1993), Cambridge Univ. Press,
2904:   Cambridge, 1995, pp.~13--34. \MR{96e:20062}
2905: 
2906: \bibitem{brown89:_build}
2907: Kenneth~S. Brown, \emph{Buildings}, Springer-Verlag, New York, 1989.
2908: 
2909: \bibitem{brown00:_semig_markov}
2910: \bysame, \emph{Semigroups, rings, and {M}arkov chains}, J. Theoret. Probab.
2911:   \textbf{13} (2000), no.~3, 871--938. \MR{1 785 534}
2912: 
2913: \bibitem{browndiaconis:_random}
2914: Kenneth~S. Brown and Persi Diaconis, \emph{Random walks and hyperplane
2915:   arrangements}, Ann. Probab. \textbf{26} (1998), no.~4, 1813--1854. \MR{1 675
2916:   083}
2917: 
2918: \bibitem{diaconis88:_group}
2919: Persi Diaconis, \emph{Group representations in probability and statistics},
2920:   Institute of Mathematical Statistics Lecture Notes---Monograph Series,
2921:   vol.~11, Institute of Mathematical Statistics, Hayward, CA, 1988.
2922: 
2923: \bibitem{diaconisfill92:_top_to_random}
2924: Persi Diaconis, James~Allen Fill, and Jim Pitman, \emph{Analysis of top to
2925:   random shuffles}, Combin. Probab. Comput. \textbf{1} (1992), no.~2, 135--155.
2926: 
2927: \bibitem{grove85:_finit}
2928: Larry~C. Grove and Clark~T. Benson, \emph{Finite reflection groups}, second
2929:   ed., Graduate Texts in Mathematics, vol.~99, Springer-Verlag, New York, 1985.
2930: 
2931: \bibitem{humphreys90:_reflec_coxet}
2932: James~E. Humphreys, \emph{Reflection groups and {C}oxeter groups}, Cambridge
2933:   Studies in Advanced Mathematics, vol.~29, Cambridge University Press,
2934:   Cambridge, 1990.
2935: 
2936: \bibitem{orlikterao92:_book}
2937: Peter Orlik and Hiroaki Terao, \emph{Arrangements of hyperplanes}, Grundlehren
2938:   der Mathematischen Wissenschaften, vol. 300, Springer-Verlag, Berlin, 1992.
2939: 
2940: \bibitem{scharlau95:_build}
2941: Rudolf Scharlau, \emph{Buildings}, Handbook of incidence geometry,
2942:   North-Holland, Amsterdam, 1995, pp.~477--645. \MR{97h:51018}
2943: 
2944: \bibitem{solomon76:_mackey}
2945: Louis Solomon, \emph{A {M}ackey formula in the group ring of a {C}oxeter
2946:   group}, J. Algebra \textbf{41} (1976), no.~2, 255--264.
2947: 
2948: \bibitem{stanley97:_enumer1}
2949: Richard~P. Stanley, \emph{Enumerative combinatorics. {V}ol. 1}, Cambridge
2950:   Studies in Advanced Mathematics, vol.~49, Cambridge University Press,
2951:   Cambridge, 1997, with a foreword by Gian-Carlo Rota, corrected reprint of the
2952:   1986 original.
2953: 
2954: \bibitem{tits74:_build_bn}
2955: Jacques Tits, \emph{Buildings of spherical type and finite {B}{N}-pairs},
2956:   Springer-Verlag, Berlin, 1974, Lecture Notes in Mathematics, Vol. 386.
2957: 
2958: \bibitem{wachs91:_stirl}
2959: Michelle Wachs and Dennis White, \emph{$p,q$-{S}tirling numbers and set
2960:   partition statistics}, J. Combin. Theory Ser. A \textbf{56} (1991), no.~1,
2961:   27--46. \MR{92b:05004}
2962: 
2963: \bibitem{ziegler95:_lectures}
2964: G{\"u}nter~M. Ziegler, \emph{Lectures on polytopes}, Graduate Texts in
2965:   Mathematics, vol. 152, Springer-Verlag, New York, 1995.
2966: 
2967: \end{thebibliography}
2968: 
2969: 
2970: 
2971: \end{document}
2972: 
2973: 
2974: 
2975: 
2976: 
2977: 
2978: