%%================================================================ %% a plain LaTeX file for a 42 page document %% %% This LaTeX file is completely self-contained, it contains %% all of its own graphics, and macros needed for LaTeX. %%================================================================ \documentstyle[12pt]{article} %%=============================================================== %% If you want to change size of the page and margins %% this is the place to do it %%=============================================================== \setlength{\topmargin}{-.45in} \setlength{\textheight}{9in} \setlength{\textwidth}{6.4in} %\setlength{\evensidemargin}{.5in} \setlength{\oddsidemargin}{.4in} %%============================================================= %% Inputing standard AMS definitions %%============================================================= \input amssym.def \input amssym %%================================================================ %% AUTHORs MACROs %%================================================================ \def\AA {{\cal A}} \def\BB {{\cal B}} \def\CC {{\cal C}} \def\DD {{\cal D}} \def\G{{\cal G}} \def\LLL{\cal L} \def\LL {{\Lambda}} \def\SS {\ifmmode {\cal S\!\it h} \else ${\cal S\!\it h}$ \fi} \def\abs#1{{| #1 |}} \def\aalpha{{\underline\alpha}} \def\bbeta{{\underline\beta}} \def\n{\ifmmode n \else $n$\fi} \def\dddelta#1#2{\delta_{x_#1,y_#2}} \def\dx#1#2{\delta_{x_#1,x_#2}} \def\dy#1#2{\delta_{y_#1,y_#2}} \def\interval#1#2{z_{{#1},{#2}}} \def\xi#1{x_{i_{#1}}} \def\yi#1{y_{i_{#1}}} \def\d{\partial} \def\dt{\partial^t} \def\zz#1{ z_{x_{#1},y_{#1}}} \def\ww{\wedge} \def\wdw{\wedge\cdots\wedge} \def\wdd{\wedge \dots} \def\ddw{ \dots \wedge} \def\odo{\otimes \cdots \otimes} \def\zzeta#1{\zz1\ww\zz2\wdw\zz#1} \def\PPP{{\Pi}} \def\CS#1{{\Bbb {C}S_{#1}}} \def\DONE{{\rule {5pt} {5pt}}} \def\CrP{{\cal C}_r(P)} \def\binom#1#2{\mbox{\small $\left(\!\!\begin{array}{c}{#1}\\{#2}\end{array} \!\!\right)$}} \def\smbinom#1#2{\mbox{\tiny $\left(\!\!\!\!\begin{array}{c}{#1}\\{#2}\end{array} \!\!\!\!\right)$}} \def\zex#1#2#3#4{z_{1,#1}\ww z_{2,#2}\ww z_{3,#3}\ww z_{4,#4}} \def\ith{\ifmmode {i^{\mbox{\small\,th}}\/ } \else {$i^{\mbox{\small\,th}}\/ $} \fi} \def\nth{\ifmmode {n^{\mbox{\small\,th}}\/ } \else {$n^{\mbox{\small\,th}}\/ $} \fi} \def\kth{\ifmmode {k^{\mbox{\small\,th}}\/ } \else {$k^{\mbox{\small\,th}}\/ $} \fi} \def\<{\langle} \def\>{\rangle} \def\u#1{{\underline #1}} \def\ch{{\cal C\!\it h}} \def\taken{\framebox[7pt]{$\cdot$}} \def\one{\framebox[7pt]{\phantom{$\cdot$}}} \def\em{\bf} %%============================================================== %% END of AUTHORs MACROs %%============================================================== \begin{document} %%============================================================= %% RUNNING HEADS -- ELECTRONIC JOURNAL OF COMBINATORICS %%============================================================= \baselineskip 15pt \pagestyle{myheadings} \markright{\sc the electronic journal of combinatorics 2 (1995), \#R14\hfill} \thispagestyle{empty} %%============================================================== \newtheorem{theorem}{Theorem} \newtheorem{conjecture}[theorem]{Conjecture} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{cor}[theorem]{Corollary} \newtheorem{claim}[theorem]{Claim} \newtheorem{prop}[theorem]{Proposition} \newtheorem{defi}{Definition}[section] \newtheorem{fact}[theorem]{Fact} \begin{center} {\Large The eigenvalues of the Laplacian for the homology of the Lie algebra corresponding to a poset} \vspace{1.0cm} {Iztok Hozo} \vspace{0.5cm} Department of Mathematics Indiana University Northwest Gary, In 46408 {email: ihozo@iunhaw1.iun.indiana.edu} \bigskip Submitted: April 6, 1995; Accepted: July 21, 1995. \end{center} \bigskip \begin{abstract} In this paper we study the spectral resolution of the Laplacian ${\cal L}$ of the Koszul complex of the Lie algebras corresponding to a certain class of posets. Given a poset $P$ on the set $\{1,2,\dots,n\}$, we define the nilpotent Lie algebra $L_P$ to be the span of all elementary matrices $z_{x,y}$, such that $x$ is less than $y$ in $P$. In this paper, we make a decisive step toward calculating the Lie algebra homology of $L_P$ in the case that the Hasse diagram of $P$ is a rooted tree. We show that the Laplacian $\LLL$ simplifies significantly when the Lie algebra corresponds to a poset whose Hasse diagram is a tree. The main result of this paper determines the spectral resolutions of three commuting linear operators whose sum is the Laplacian $\LLL$ of the Koszul complex of $L_P$ in the case that the Hasse diagram is a rooted tree. We show that these eigenvalues are integers, give a combinatorial indexing of these eigenvalues and describe the corresponding eigenspaces in representation-theoretic terms. The homology of $L_P$ is represented by the nullspace of $\LLL$, so in future work, these results should allow for the homology to be effectively computed. \medskip\noindent AMS Classification Number: 17B56 (primary) 05E25 (secondary) \end{abstract} %================================================== % preliminaries: Definitions and known % results needed for later work %------------------------------------------------------------------------------------------------------------- % % %------------------------------------------------------------------------------------------------------------- \section{Preliminaries} \subsection{Definitions} A {\em partially ordered set} $P$ (or {\em poset}, for short) is a set (which by abuse of notation we also call $P$), together with a binary relation denoted $\le$ (or $\le_P$ when there is a possibility of confusion), satisfying the following three axioms: \begin{enumerate} \item For all $x \in P$\/, $ x\le x$. {\em (reflexivity)} \item If $x\le y$ and $y \le x$, then $x = y$. {\em (antisymmetry)} \item If $x\le y$ and $y\le z$, then $x\le z$. {\em (transitivity)} \end{enumerate} A {\em chain} (or {\em totally ordered set} or {\em linearly ordered set}) is a poset in which any two elements are comparable. A subset $C$ of a poset $P$ is called a {\em chain} if $C$ is a chain when regarded as a subposet of $P$. \begin{defi} A poset $P$ is {\em linear} if for any two comparable elements $x,~y~\in~P$, the interval $[x,y]$ is a chain, i.e., if every interval has the structure of a chain. \end{defi} The {\em length} $l(C)$ of a finite chain $C$ is defined by $l(C)=\abs C -1$. %======================================================================= \subsection{The homology of a poset} The combinatorial approach to a homology theory for posets was developed by Rota \cite{rota}, Farmer \cite{farmer}, Lakser \cite{lakser}, Mather \cite{mather}, Crapo \cite{crapo} and others (more references can be found in \cite{walker}). A systematic development of the relationship between the combinatorial and topological properties of posets was begun by K.~Baclawski \cite{baclawski} and A.~Bj\"orner \cite{bjorner} and continued by J.~Walker \cite{walker}. Define the set $\CrP$ to be the set of 0-1 chains of length $r$ in the poset $P$. By abuse of notation we will use the same name for the complex vector space $C_r$ or $\CrP$, with basis the set of $r$-chains. The $C_r$'s are called {\em chain spaces}. The map $\partial_r : C_r \rightarrow C_{r-1}$, called the {\em boundary map}, is defined by: $$ \partial_r (\hat 0 < x_1 < \ldots $ on the product $\oplus \Gamma_r$, such that $\< \Gamma_r,\Gamma_s\>~=~0$ whenever $r\not = s$. We will restrict our attention to the subspaces of the nilpotent Lie algebra $T_n({\Bbb C})$ of all strictly upper triangular matrices over the complex numbers, with standard basis $\{ z_{i,j} ~:~1\le i < j \le n\}$, so we can define this product naturally: \begin{defi} Let $L$ be a Lie algebra, $L\subset T_n({\Bbb C})$. Define an inner product for standard basis elements $v,w \in L$ by: $$ \< v, w\> = \left \{ \begin{array} {lr} 1 & \quad \mbox{ if } v=w \\ 0 & \mbox{otherwise}\\ 0 & \mbox{ if $v$ and $w$ have different exterior degrees}\\ \end{array} \right. $$ \end{defi} Extend this to the exterior algebra, i.e., to the complexes mentioned above. \begin{defi} Suppose that $ v = v_1 \wdw v_k$ and $w = w_1 \wdw w_k$. Then define the inner product: $$ \< v,w\> = det ( \< v_i, w_j\> )_{1\le i,j \le k} $$ \end{defi} Note that this can be written also as $$ \< v,w\> = \sum_{\sigma \in S_n} sgn(\sigma) \prod_i \< v_i, w_{\sigma(i)}\> = \left \{ \begin{array} {lr} sgn(\sigma) & \mbox{ iff } v_i = w_{\sigma(i)} \mbox{ for all $i$}\\ 0 & \mbox{ otherwise }\\ \end{array} \right. $$ In other words, the product of two pure wedges of basis elements is nonzero if and only if two pure wedges differ only in the order of the elements, and in that case, the product is just the sign of the permutation that changes one into another. Define $\delta_r$ mapping $\Gamma_r$ into $\Gamma_{r+1}$ by $$ \<\delta_r v, w\> = \< v, \partial_{r+1} w\> $$ over all $v \in \Gamma_r$, and all $w\in \Gamma_{r+1}$. It is enough to calculate $\delta$ on pure wedges (as in our definitions), since the inner product and $\delta$ are both linear functions. \begin{lemma} The map $\delta$ is given by \begin{eqnarray*} \lefteqn{\delta_r(\zz 1 \wedge \zz 2 \wedge \dots \wedge \zz r)} \\ &=&\sum_{s=1}^r (-1)^{s-1} { \sum_{x_s < l < y_s} \zz 1 \wedge\dots\wedge z_{x_s,l}\wedge z_{l,y_s}\wedge \dots \wedge \zz r } \end{eqnarray*} \end{lemma} {\em Note:} It is easy to check that $\delta_{r+1} \delta_r = 0$, thus $\delta_*$ defines a coboundary operator, and so we can define the cohomology to be $$ H^r(L) = Ker( \delta_r)/Im(\delta_{r-1})$$ Proof: But to prove that, it is enough to show that the coefficient of the pure wedge $z_{x_1,y_1}\wedge z_{x_2,y_2} \wedge \dots \wedge z_{x_r,y_r}$ in $\d (\zz 1 \wdw z_{x_s,l} \ww z_{l,y_s} \wdw \zz r )$ is $(-1)^{s-1}$ for any $l \in (x_s,y_s)$, i.e., \begin{eqnarray*} \lefteqn{\d(z_{x_1,y_1}\wedge\dots\wedge z_{x_s,l}\wedge z_{l,y_s}\wedge \dots \wedge z_{x_r,y_r})} \\ &=&\dots + (-1)^{s-1} (z_{x_1,y_1}\wedge z_{x_2,y_2} \wedge \dots \wedge z_{x_r,y_r}) + \dots \end{eqnarray*} and this is not difficult by the definition of $\d$. Note that we can change the order of the elements in the pure wedges, and obtain a slightly different form for $\delta$: \begin{eqnarray*} \lefteqn{\delta_r(z_{x_1,y_1} \wedge z_{x_2,y_2} \wedge \dots \wedge z_{x_r,y_r})}\\ &=& \sum_{s=1}^r (-1)^{s-1} { \sum_{x_s < l < y_s} z_{x_1,y_1}\wedge\dots\wedge z_{x_s,l}\wedge z_{l,y_s}\wedge \dots \wedge z_{x_r,y_r} }\\ &=&\sum_m \sum_{x_m < l < y_m} (z_{x_m,l} \ww \zz 1 \wdw z_{l,y_m} \wdw \zz k) \end{eqnarray*} This is the form for the $\delta=\dt$ we will use. \begin{defi} Define the {\em Laplacian operator} $L_r: \Gamma_r \rightarrow \Gamma_r$ by $$ L_r = \delta_{r-1} \partial_r + \partial_{r+1} \delta_r $$ \end{defi} \begin{theorem}[Kostant, \cite{kostant} \label{kost1}] Let $B = \{\beta_1,\ldots,\beta_d\}$ be a basis for $Ker(L_r)$. Then $B$ is simultaneously a complete set of representatives of $H^r(L)$ and $H_r(L)$. In particular $\dim(H^r(L)) = \dim(H_r(L)) = \dim( Ker(L_r))$. \end{theorem} Sometimes, the Laplacian $L_r$ will turn out to be very simple. In these cases, Theorem~\ref{kost1} is a very efficient method for evaluating the homology and cohomology of a Lie algebra. One famous result obtained in this way is given by Kostant~\cite{kostant}. \subsection{Kostant's Theorem} We need some preliminary definitions. Suppose $\cal G$ is a semisimple Lie algebra, with the root system $R$, whose basis is $\Delta$. Thus ${\cal G} = H \oplus ( \oplus_{\alpha \in R} \< z_\alpha\>)$, where $H$ is the torus. Suppose that $S \subset \Delta$, and let $R_S$ be the set of roots in the $\Bbb Z$ (integer) module spanned by elements of $S$. Define $\cal G_S$ to be ${\cal G_S} = H \oplus \< z_\alpha : \alpha \in R_S \>$. Define a ${\cal G}_S$ module $N_S$ to be $N_S = \< z_\alpha : \alpha \in R^+ \setminus R_S^+ \>$. We will state a couple of facts without proof: \begin{itemize} \item $N_S$ is a nilpotent subalgebra of $\cal G$. \item Let $W$ be a $\cal G$-module. Then $W$ is also a $N_S$-module and a $\cal G_S$-module. \item Thus we can compute $H(N_S ;W^\mu)$ as $\cal G_S$-module, where $W^\mu$ is an irreducible $\cal G$-module. Kostant used the Laplacian operator to prove the following theorem: \end{itemize} \begin{theorem} [Kostant, Theorem 5.7,\cite{kostant}] Let $\lambda$ be a dominant weight for $\cal G$, and let $\mu$ be a minimal weight for $\cal G_S$. Let $V$ be a $\cal G_S$-invariant subspace of $W^\lambda \otimes \bigwedge^r N_S$ isomorphic to the $\cal G_S$-irreducible (indexed by $\mu$) with minimal weight $\mu$. \begin{itemize} \item The Laplacian $L=\delta \partial + \partial \delta$ preserves $V$. \item Then, $L\vert_V$ is a scalar, and the scalar is given by $$ \frac{1}{2} (\vert \rho + \lambda \vert^2 - \vert \rho - \mu \vert^2) $$ where $\rho$ is half of the sum of the positive roots of $\cal G$. \end{itemize} \end{theorem} \subsection{The Lie Algebra corresponding to a Poset} \begin{defi} A {\em standard labeling} of the poset $P$ is a total ordering of the elements of $P$ such that whenever $x <_P y$, $x$ precedes $y$ in that total ordering. \end{defi} Since $P$ is a partial order, i.e. transitive , there always is such labeling. Fix a standard labeling of the poset $P$. We can define a Lie algebra $L_P$ corresponding to the poset $P$ in the following way. First, for every relation $x<_P y$ in the poset $P$, i.e., for every two elements $x,y \in P$ such that $x<_P y$ we can define the matrix $z_{x,y}$ having all entries equal to zero, except for exactly one entry equal to 1, namely the entry at the position $x,y$ in the standard labeling of the poset $P$. All matrices $z_{x,y}$ are strictly upper triangular because of our labeling. So $L_P$ is a subalgebra of $T_n$. The Lie algebras $L_P$ obtained from distinct labellings are isomorphic -- the labeling only specifies embedding of $L_P$ in the $n\times n$ matrices. %========================================================== % the main part %------------------------------------------------------------------------------------------------------------------- % %------------------------------------------------------------------------------------------------------------- \section{The Formula for Laplacian of a Linear Poset \label{ch4}} In this section we will present a significant simplification of the Lie algebra Laplacian in the case of linear posets. That will allow us to prove our main result on the eigenvalues of those Laplacians. \subsection{Simplification } Recall the Lie algebra boundary map: \begin{eqnarray*} \lefteqn{\d(\zz 1\wedge \dots \wedge \zz k)} \\ &=&\sum_{i y_3$. All together, the relations are: \begin{eqnarray*} \begin{array}{c} y_2 \\ y_1 \\ \end{array} > x_2 > y_3 > \begin{array}{c} x_3 \\ x_1 \\ \end{array} \end{eqnarray*} So in this case we have $y_1 > y_3$, and $x_2 > x_3$. Let \begin{eqnarray*} \CC &=&(x_2,x_3)\cdot(y_1,y_3)\cdot (\zz 1 \ww \zz 2 \ww \zz 3) \\ &=&(x_2,x_3)\cdot(z_{x_1,y_3} \ww z_{x_2,y_2}\ww z_{x_3,y_1})\\ &=& - ( z_{x_1,y_3} \ww z_{x_2,y_1} \ww z_{x_3,y_2}) \end{eqnarray*} and \begin{eqnarray*} \DD&=&(y_1,y_3) \cdot (x_2,x_3) \cdot (\zz 1 \ww \zz 2 \ww \zz 3) \\ &=& - (y_1,y_3) \cdot (z_{x_1,y_1} \ww z_{x_2,y_3} \ww z_{x_3,y_2} )\\ &=& 0 \end{eqnarray*} since $x_2 > y_3$. The expressions $\AA$ and $\CC$ are two summands of the product $L_X L_Y$, while $\BB$ and $\DD$ are two summands of the product $L_Y L_X$. As we can see, $\AA + \CC = \BB + \DD$. Thus $L_X \cdot L_Y = L_Y \cdot L_X$. \DONE \bigskip In view of Lemma~\ref{lemma5.1}, $L_X$, $L_Y$ and $L_D$ are commuting linear transformations. So, to analyze the spectrum of their sum, we can compute the eigenvalues and eigenspaces of each separately. We will begin with $L_Y$. \subsection{ A poset tableau of type $(X,Y)_P$} \begin{defi} The {\em diagram of the L-block}, $P[X,Y]$, spanned by the sets $(X,Y)_P$, is the Hasse diagram of the subposet $X \cup Y$ with order inherited from the poset $P$. Furthermore every vertex of $P$, which is in the intersection $X \cap Y$ is split into two nodes, with the $x$-node above the $y$-node. \end{defi} \begin{defi} Given a node $v$ in $P[X,Y]$ define the {\em repetition number of $v$ }, $k(v)$, to be the number of times that $v$ appears in the multiset $X$ if $v$ is an $x$-node of $P[X,Y]$, or the multiset $Y$ if $v$ is a $y$-node of $P[X,Y]$. \end{defi} Let $C(v)$ be the set of covers of node $v$ in $P[X,Y]$. If $v$ is a maximal node, than $C(v) = \emptyset$. \begin{defi}\label{def:poset} A {\em poset tableau of type $(X,Y)_P$ } (or just of type $(X,Y)$) is any labeling $\LL$ of the diagram, $P[X,Y]$, of the L-block $V$ spanned by $(X,Y)$, where the labels are partitions $\LL(v)$, such that $\LL(v)$ is a partition of the number $\sum_{w\ge v} \epsilon(w) k(w)$, where $$\epsilon(w) = \left\{ \begin{array}{ll} +1 & \mbox{ if $w$ is a $y$-node} \\ -1 & \mbox{ if $w$ is an $x$-node.} \end{array} \right. $$ \end{defi} Given a poset tableau $\LL$ we will define the {\em multiplicity of $\LL$, $m(\LL)$}, and the {\em eigenvalues of $\LL$, $e(\LL)$.} \begin{defi} \begin{itemize} \item Let $v$ be a $y$-node of the diagram $P[X,Y]$, labeled with the partition $\LL(v)$ and with repetition number $k(v)$. Let $C(v)= \{ v_1, v_2, \dots , v_l\}$ be the set of covers of $v$. Let $\lambda_i$ denote $\LL(v_i)$, and let $k_i$ denote the repetition numbers, $k(v_i)$. The {\em multiplicity of $\LL$ at $v$} is defined to be $$ m_v(\LL) = c_{\lambda_1, \dots, \lambda_l, k(v)}^{\LL(v)} $$ \item Let $v$ be an $x$-node of the diagram $P[X,Y]$, labeled with the partition $\LL(v)$ and with repetition number $k(v)$. Let $C(v)= \{ v_1, v_2, \dots , v_l\}$ be the set of covers of $v$. Let $\lambda_i$ denote $\LL(v_i)$, and let $k_i$ denote the repetition numbers, $k(v_i)$. The {\em multiplicity of $\LL$ at $v$} is defined to be $$ m_v(\LL) = \sum_{\mu} c_{\lambda_1, \dots, \lambda_l}^{\mu} c^{\mu}_{\LL(v), 1^{k(v)}}. $$ \end{itemize} \end{defi} If the multiplicity $m_v(\LL) = 0$ then we know that that particular labeling is not valid. Now, we will define the $y$-eigenvalues for each $y$-node $v$ of the diagram $P[X,Y]$. We want to have as many $y$-eigenvalues as the value of multiplicity. From the representation theory of the symmetric group, we know that \begin{equation}\label{eig1} c_{\lambda_1, \dots, \lambda_l, k(v)}^{\LL(v)} = \sum_{ \mbox{\tiny $\LL(v)/\mu=k(v)${--horizontal strip}} } c_{\lambda_1, \dots, \lambda_l}^{\mu}. \end{equation} The {\em node--eigenvalue, $e_v(\LL)$, } for each node $v$, is the set of the sums of the content over all squares in $\LL(v)/\mu$ for all possible $\mu$ for which $\Lambda(v)/\mu$ is a $k(v)$--horizontal strip minus the binomial coefficient $\binom{k(v)}{2}$. Recall that the content of a square is given by $c(i,k) = k-i$ if the square is at position $(i,k)$ in a partition (\ith row and \kth column). This gives $m_v(\LL)$ eigenvalues at each $y$-node $v$. We now define {\em $y$-eigenvalue of $\LL$}, $e_y(\LL)$, to be the set of numbers obtained by taking a sum of one element of $e_v(\LL)$ for each $y$-node $v$. So $\abs{e_y(\LL)} = \prod_{\mbox{$y$-nodes $v$}} m_v(\LL)$. \subsection{Example \label{exa44}} Let the poset $P = \{ 1,2,3,4,5 \}$ with the relations $1<_P 2$, $2<_P 3$, $3<_P 4$ and $4<_P 5$. The Hasse diagram of this poset is given in figure~\ref{V--9}. %=================================================================== %\begin{figure} %\centerline{\picture{1}{1.208}{2.083}{pic51.1.ps}} %\caption{Example: The poset $P$ from Example~\ref{exa44}}\label{V--9} %\end{figure} \begin{figure}[t] \begin{center} \begin{picture}(60,140)(30,0) %---------------------------------------- \put(0,0){\line(0,1){120}} \put(3,-1){1} \put(3,29){2} \put(3,59){3} \put(3,89){4} \put(3,119){5} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(0,120){\circle*{3}} %---------------------------------------------- \end{picture} \caption{ Example: poset $P$}\label{V--9} \end{center} \end{figure} Let $X$ and $Y$ be the sets $X=\{1,2,3\}$ and $Y=\{ 4,4,5\}$. So the node $4$, is a node with non-trivial repetition number $k(4) = 2$. The L-block $V$ is spanned by the following pure wedges: \begin{eqnarray*} \zeta &=& z_{1,4}\ww z_{2,4} \ww z_{3,5} \\ \tau &=& z_{1,4}\ww z_{2,5} \ww z_{3,4} \\ \eta &=& z_{1,5}\ww z_{2,4} \ww z_{3,4} . \end{eqnarray*} Thus the L-block $V$ is 3-dimensional. We calculate the Laplacian $L_Y$ on these three elements. Note that the Laplacian $L_Y$ is in fact $L_Y = (4,5)$, since those are the only two comparable $y$'s. \begin{eqnarray*} L(\zeta) &=& \tau + \eta \\ L(\tau) &=& \zeta + \eta \\ L(\eta) &=& \zeta + \tau \end{eqnarray*} The matrix representation of $L_Y$ with respect to the basis $<\zeta, \tau, \eta >$ is thus \begin{eqnarray*} L_Y = \left( \begin{array}{ccc} 0 & 1 & 1 \\ 1 & 0 & 1 \\ 1 & 1 & 0 \\ \end{array} \right) \end{eqnarray*} So the eigenvalues of the Laplacian are $-1, -1, +2$. Now we will evaluate the $y$--eigenvalue for each of the poset tableaux for this L-block. The only nontrivial node is node $4$. Thus, the $y$--eigenvalue, $e_y(\Lambda)$, is the node-eigenvalue, $e_4(\Lambda) = c(1,2) + c(1,3) - \binom{2}{2}$. The result is given in figure~\ref{51.9}. %========================================================== %\begin{figure} %\centerline{\picture{1}{3.250}{2.444}{pic51.9.ps}} %\caption{Example: the $y$-eigenvalues}\label{51.9} %\end{figure} \begin{figure}[t] \begin{center} \begin{picture}(300,200)(0,0) %---------------------------------------- \put(50,50){\begin{picture}(60,140)(30,0) \put(0,0){\line(0,1){120}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one\one}} \put(8,89){\makebox(0,0)[bl]{\one\one\one}} \put(8,119){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(0,120){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = 2$}} \end{picture}} %---------------------------------------------- %---------------------------------------- \put(150,50){\begin{picture}(60,140)(30,0) \put(0,0){\line(0,1){120}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one\one}} \put(8,89){\makebox(0,0)[bl]{\one\one}} \put(8,78){\makebox(0,0)[bl]{\one}} \put(8,119){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(0,120){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = -1$}} \end{picture}} %---------------------------------------------- %---------------------------------------- \put(250,50){\begin{picture}(60,140)(30,0) \put(0,0){\line(0,1){120}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one}} \put(8,48){\makebox(0,0)[bl]{\one}} \put(8,89){\makebox(0,0)[bl]{\one\one}} \put(8,78){\makebox(0,0)[bl]{\one}} \put(8,119){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(0,120){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = -1$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{Example: the $y$-eigenvalues}\label{51.9} \end{center} \end{figure} Note that the $y$--eigenvalues of this labeling give exactly the same numbers as the eigenvalues of the Laplacian $L_Y$. In the next section we will show that this is not coincidental. %-------------------------------------------------------------------------------------------- %-------------------------------------------------------------------------------------------- % %-------------------------------------------------------------------------------------------- \section{Centerpiece Theorem for $L_Y$} \begin{theorem}[$L_Y$-Centerpiece] Let $P$ be a linear poset with a minimum element, $\hat 0$. Let $X$ and $Y$ be two (multi-)sets, subsets of $P$. For every labeling $\LL$ of positive multiplicity, each element in $e_y(\LL)$ is an eigenvalue of $L_Y$ with multiplicity $\prod_{\mbox{$x$-nodes $v$}} m_v(\LL)$. \end{theorem} Proof: The proof of this theorem will be by induction on the sizes of the (multi)-sets $X$ and $Y$. So let $n=\abs{X} = \abs{Y}$ (counting multiplicities). If $n=1$ --- there is nothing to prove as the Laplacian $L_Y$ has no pairs to switch, and the only $y$-node is the maximal element for the diagram of the L-block. The Laplacian $L_Y$ is the one-by-one zero matrix and the eigenvalue of this unique pair is zero. Suppose $n=2$. There are several different possible combinations of relations between sets $X=\{x_1,x_2\}$ and $Y=\{y_1, y_2\}$. \begin{itemize} \item The most obvious one is $x_1 < x_2 < y_1 < y_2$. In that case the Laplacian $L_Y = (y_1, y_2)$, and the two possible elements are $\zeta_1 = \zz 1 \ww \zz 2$, $\zeta_2 = z_{x_1,y_2} \ww z_{x_2,y_1}$. The Laplacian $L_Y$ has the following matrix representation with respect to the basis $\< \zeta_1 , \zeta_2 \>$: \begin{eqnarray*} L_Y = \left( \begin{array}{cc} 0 & 1\\ 1 & 0 \end{array} \right) \end{eqnarray*} The eigenvalues of $L_Y$ are $+1$ and $-1$. The eigenvalue of the poset tableau of type $(X,Y)_P$ is given in figure~\ref{sl51.11}. It also gives values $+1$ and $-1$, so the claim of the theorem holds. \item The second case is when \begin{eqnarray*} x_1 < x_2 < \begin{array}{c}{y_1}\\{y_2}\end{array} \end{eqnarray*} %\begin{figure} %\centerline{\picture{1}{1.930}{1.944}{pic51.11.ps}} %\caption{poset tableaux}\label{sl51.11} %\end{figure} %========================================================== \begin{figure}[t] \begin{center} \begin{picture}(200,120)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){90}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one\one}} \put(8,89){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = +1$}} \end{picture}} %---------------------------------------------- %---------------------------------------- \put(150,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){90}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one}} \put(8,48){\makebox(0,0)[bl]{\one}} \put(8,89){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = -1$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{ poset tableaux}\label{sl51.11} \end{center} \end{figure} In that case the Laplacian $L_Y$ does nothing (since $y_1$ and $y_2$ are not comparable), thus the eigenvalues of $L_Y$ are $0$. The dimension of the L-block spanned by $(X,Y)$ is two. The eigenvalues of the poset tableaux give the same values (figure~\ref{sl51.12}), where the ''$\cdot$'' in a box denotes which square was deleted in that step. %\begin{figure} %\centerline{\picture{1}{2.875}{1.750}{pic51.12.ps}} %\caption{poset tableaux}\label{sl51.12} %\end{figure} %========================================================== \begin{figure}[t] \begin{center} \begin{picture}(200,120)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){30}} \put(0,30){\line(1,1){30}} \put(0,30){\line(-1,1){30}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one\taken}} \put(38,59){\makebox(0,0)[bl]{\one}} \put(-38,59){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(30,60){\circle*{3}} \put(-30,60){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = 0$}} \end{picture}} %---------------------------------------------- %---------------------------------------- \put(150,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){30}} \put(0,30){\line(1,1){30}} \put(0,30){\line(-1,1){30}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,18){\makebox(0,0)[bl]{\taken}} \put(38,59){\makebox(0,0)[bl]{\one}} \put(-38,59){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(30,60){\circle*{3}} \put(-30,60){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = 0$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{ poset tableaux}\label{sl51.12} \end{center} \end{figure} Thus in this case the theorem checks too. \item $x_1 < y_1 < x_2 < y_2$ or equivalently (for our purpose) $x_1 < y_1 = x_2 < y_2$.\\ There is only one poset tableau spanned by these sets $X$ and $Y$, namely the one shown on the figure~\ref{sl504a}. %\begin{figure} %\centerline{\picture{0.7}{0.680}{1.944}{pic504.ps}} %\caption{\label{sl504a} poset tableau} %\end{figure} %========================================================== \begin{figure} \begin{center} \begin{picture}(100,120)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){90}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{$\emptyset$}} \put(8,89){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = 0$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{\label{sl504a} poset tableau} \end{center} \end{figure} The $y$-eigenvalue for the poset tableau is zero in both cases. The Laplacian $L_Y$ can not switch the $y$'s, since that would produce the element $z_{x_2,y_1}$ which doesn't exist. So the Laplacian $L_Y$ also acts as zero. \item $x_1 < x_2 < y_1 = y_2$.\\ There is only one poset tableau spanned by these sets $X$ and $Y$, namely the one shown on the figure~\ref{sl505a}. %\begin{figure} %\centerline{\picture{0.7}{0.680}{1.444}{pic505.ps}} %\caption{\label{sl505a} poset tableau} %\end{figure} %========================================================== \begin{figure} \begin{center} \begin{picture}(100,90)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,80)(30,0) \put(0,0){\line(0,1){60}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = 0$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{\label{sl505a} poset tableau} \end{center} \end{figure} The $y$-eigenvalue is again zero (contents of the partition $(2)$ minus the binomial coefficient $\binom{2}{2}$). The Laplacian $L_Y$ doesn't have two distinct $y$'s to switch, thus, it is zero. \item $x$'s are the same. $$x_1 = x_2 < \begin{array}{c} y_1 \\ y_2 \end{array}$$ There is only one poset tableau spanned by these sets $X$ and $Y$, namely the one shown on the figure~\ref{sl506a}. %\begin{figure} %\centerline{\picture{0.7}{1.125}{1.250}{pic506.ps}} %\caption{\label{sl506a} poset tableau} %\end{figure} %========================================================== \begin{figure} \begin{center} \begin{picture}(100,50)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,60)(30,0) \put(0,0){\line(1,1){30}} \put(0,0){\line(-1,1){30}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(36,29){\makebox(0,0)[bl]{\one}} \put(-40,29){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(30,30){\circle*{3}} \put(-30,30){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = 0$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{\label{sl506a} poset tableau} \end{center} \end{figure} The $y$-eigenvalue is zero. The Laplacian $L_Y$ has no comparable $y$'s to switch - thus $L_Y=0$. \item $x_1 = x_2 < y_1 < y_2$. There is only one poset tableau spanned by these sets $X$ and $Y$, namely the one shown on the figure~\ref{sl507a}. %\begin{figure} %\centerline{\picture{0.7}{0.680}{1.444}{pic507.ps}} %\caption{\label{sl507a} poset tableau} %\end{figure} %========================================================== \begin{figure} \begin{center} \begin{picture}(100,90)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,80)(30,0) \put(0,0){\line(0,1){60}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one\one}} \put(8,59){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e_y(\Lambda) = 1$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{\label{sl507a} poset tableau} \end{center} \end{figure} The $y$-eigenvalue is equal to 1. The Laplacian $L_Y$ can switch $y_1$ and $y_2$ but the result would be the same element, since the $x$'s are indistinguishable. Thus the Laplacian $L_Y = Id$, with the eigenvalue 1. \item In the trivial case when the $x$'s are not comparable the $y$'s are not comparable because of linearity and the existence of a minimal element. So we have $$ \begin{array}{ccc} x_1 & < & y_1 \\x_2 & < & y_2 \end{array}. $$ The poset tableau again gives zero as the $y$-eigenvalue, and since the $y$'s are not comparable, the Laplacian $L_Y$ is also zero. So the theorem holds for the case $n=2$. \end{itemize} Now, we will treat the general case $n>2$. \begin{itemize} \item Label the $y$-nodes of the diagram of the L-block using the depth-first algorithm: \begin{enumerate} \item Start with a leftmost minimal $y$-element $v$. \item{\label{step0}} If $v$ is not the maximal unlabeled $y$-node go to the leftmost unlabeled cover of $v$, and repeat this step. Otherwise label $v$ with next available number from the set $\{1,2,\dots, \abs{Y}\}$. \end{enumerate} From this labeling we see that, $y_i > y_j \Rightarrow i x_{a+1} > \cdots > x_n $, where all $>$ are covering relations. \item Either \begin{itemize} \item[Case 1: ] There is more than one element covering $x_a$. \item[Case 2: ] $x_a$ has unique cover in $P[X,Y]$ but it is a $y$-element. \end{itemize} \end{enumerate} Let $B = \{ x_a, \dots , x_n\}$, and let $G = $Sym($B$). \begin{lemma}\label{first.one.lemma} Let $\sigma \in G$, and let $\zeta = z_{x_{i_1},y_1} \ww z_{x_{i_2},y_2} \wdw z_{x_{i_n},y_n}$ be non-zero. Then $$ \zeta^{\sigma} = z_{\sigma(x_{i_1}),y_1} \ww z_{\sigma(x_{i_2}),y_2} \wdw z_{\sigma(x_{i_n}),y_n} $$ is also non-zero. \end{lemma} Proof: It is sufficient to prove the lemma for the transposition $(x_{i_k},x_{i_l}) \in G$. $$ \zeta^{(x_{i_k},x_{i_l})} = z_{x_{i_1},y_1} \wdw z_{x_{i_l},y_l} \wdw z_{x_{i_k},y_k} \wdw z_{x_{i_n},y_n}. $$ Now, since $(x_{i_k},x_{i_l}) \in B$, i.e., $x_{i_k},x_{i_l}$ are both less or equal to $x_a$, which is below all of the $y$'s -- the lemma is clear.\DONE \begin{lemma}\label{second.one.lemma} This action of $G$ commutes with $L_Y$. \end{lemma} Proof: Let $\zeta = z_{x_{i_1},y_1} \ww z_{x_{i_2},y_2} \wdw z_{x_{i_n},y_n}$ be in our L-block, $V$, let $y_k <_P y_l$ and let $\sigma \in G$. Then $$ \sigma \cdot (y_k,y_l) \cdot \zeta = z_{\sigma(x_{i_1}),y_1} \wdw z_{\sigma(x_{i_k}),y_l} \wdw z_{\sigma(x_{i_l}),y_k} \wdw z_{\sigma(x_{i_n}),y_n} $$ and $$ (y_k,y_l)\cdot \sigma \cdot \zeta = z_{\sigma(x_{i_1}),y_1} \wdw z_{\sigma(x_{i_k}),y_l} \wdw z_{\sigma(x_{i_l}),y_k} \wdw z_{\sigma(x_{i_n}),y_n} . $$ So they are equal unless one of the expressions above is zero, and the other is not. The only way for that to happen is in the middle step, i.e., either $\sigma \cdot \zeta =0$ or $(y_k,y_l)\cdot \zeta = 0$. But we assumed the $\zeta \not = 0$, and by our lemma above $\sigma \cdot \zeta \not = 0$. So the only possible conflict is $(y_k,y_l) \cdot \zeta =0$, and $(y_k,y_l) \cdot \sigma \cdot \zeta \not = 0$. But since $\sigma \in G$, it only moves elements of $B$ which are allowed to be paired with any $y_k$.\DONE Let $C(x_a)=\{v_1, v_2, \dots, v_l\}$ be the set of covers of $x_a$. We will prove the following generalization of the Centerpiece Theorem: \begin{theorem} Let $\LL$ be a poset tableau of positive multiplicity, and let $\lambda_i$ be the label of $v_i$ in $\LL$. Let $\lambda$ be a partition such that $$ c_{\lambda_1, \dots, \lambda_l}^\lambda c_{\LL(x_a), 1^{k(x_a)}}^{\lambda} \not = 0. $$ Then the occurrences of $G$-irreducibles $S^\lambda$ in $V$ can be indexed by such poset tableaux $\LL$, and the Laplacian $L_Y$ acts on $S^\lambda$ as one of the scalars in $e_Y(\LL)$. \end{theorem} % % %====================== C A S E 1 =================== % % \subsubsection{Case 1} %\begin{figure}[t] %\centerline{\picture{0.66}{2.750}{3.972}{pic811.ps}} %\caption{Case 1}\label{slucaj1} %\end{figure} \begin{figure} \begin{center} \begin{picture}(100,160)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,150)(30,0) \put(0,0){\line(0,1){20}} \put(0,60){\line(0,1){60}} \put(0,100){\line(2,1){40}} \put(0,100){\line(-1,1){20}} \put(0,0){\circle*{3}} \put(0,20){\circle*{3}}\put(0,30){\circle*{3}}\put(0,40){\circle*{3}} \put(0,50){\circle*{3}}\put(0,60){\circle*{3}}\put(0,80){\circle*{3}} \put(0,100){\circle*{3}}\put(0,120){\circle*{3}} \put(-20,120){\circle*{3}} \put(40,120){\circle*{3}} \put(10,120){\circle*{3}}\put(20,120){\circle*{3}}\put(30,120){\circle*{3}} \put(8,-3){\makebox(0,0)[bl]{$x_n$}} \put(8,77){\makebox(0,0)[bl]{$x_{a+1}$}} \put(8,97){\makebox(0,0)[bl]{$x_{a}$}} \put(-22,112){\makebox(0,0)[bl]{$\mu_1$}} \put(4,112){\makebox(0,0)[bl]{$\mu_2$}} \put(44,112){\makebox(0,0)[bl]{$\mu_r$}} \put(-20,120){\line(0,1){10}}\put(-20,130){\makebox(0,0)[bl]{$T_1$}} \put(0,120){\line(0,1){10}}\put(0,130){\makebox(0,0)[bl]{$T_2$}} \put(40,120){\line(0,1){10}}\put(40,130){\makebox(0,0)[bl]{$T_r$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{Case 1}\label{slucaj1} \end{center} \end{figure} Suppose there are two or more subtrees above the node $x_a$, in our poset $P$ (as in figure~\ref{slucaj1}). Label the subtrees above the $x_a$ by $T_1, T_2, \dots , T_r$. Let $Y = Y_1 \cup Y_2 \cup \cdots \cup Y_r$ ($Y_i \subset T_i$), where $Y_i \cap Y_j = \emptyset$. Let $k_i = \abs{Y_i}$, and let $$ Y_i = \{ y_{k_1+k_2+\cdots +k_{i-1}+1} , \dots , y_{k_1+\cdots +k_{i-1}+k_i} \}. $$ Because of our labeling, we know that all relations between $y$'s are contained within the sets $Y_i$, i.e., $y_i <_P y_j$ implies that both $y_i, y_j$ are in the same $Y_k$. Let $b_i$ be the number of $y_j$'s in $T_i$ minus the number of $x_j$'s in $T_i$ (note that in general $T_i$ will have more $y_j$'s than $x_j$'s). In other words, $b_i = \abs{Y_i} - \abs{X \cap T_i}$. Consider element $\zeta = z_{x_{i_1},y_1} \ww z_{x_{i_2},y_2} \wdw z_{x_{i_n},y_n}$. If we want $\zeta$ to be non-zero, there will be exactly $b_i$ $x_j$'s from $B$ in $\zeta$ paired up with the $y_j$'s of $T_i$. Split the L-block \begin{equation} V = \oplus_{ ( S_1,S_2,\dots,S_r) } V[S_1,S_2,\dots,S_r] , \label{spliting:eq} \end{equation} where $S_1 \cup S_2 \cup \cdots \cup S_r = B$, $\abs{S_i}=b_i$, and $V[S_1,\dots,S_r]$ is the span of all $\zeta$ with exactly the elements of $S_i$ paired with the $y$'s from $T_i$. \begin{lemma}\label{LLLEEEMMMAAA} \begin{enumerate} \item[1.] As a vector space $$ V[S_1,\dots, S_r] \cong V(X_1\cup S_1, Y_1) \otimes V(X_2\cup S_2, Y_2) \odo V(X_r\cup S_r, Y_r), $$ where $X_i$ and $Y_i$ are the multisets of the $x$ and $y$ elements of the subtree $T_i$. \item[2.] With respect to the decomposition in 1., the Laplacian $L_Y$ acts as: $$ L_Y(v_1 \odo v_r) = \sum_i v_1 \odo (L_{Y_i}v_i) \odo v_r . $$ \end{enumerate} \end{lemma} Proof: Statement 1. is clear by the definition of $V[S_1,\dots, S_r]$. To prove statement 2., we only need to recall that the Laplacian $L_Y$ switches comparable $y$'s, and that $y$'s in different subtrees can not be comparable because of linearity of the poset. $L_Y$ can switch only $y$'s in the same subtree $T_i$.\DONE Let $G_i = \mbox{Sym($S_i$)}$. Note that $G_1 \times G_2 \times \dots \times G_r$ acts on $V[S_1,S_2, \dots ,S_r]$. Let $s_i$ be the minimal node of subtree $T_i$. Now apply the induction hypothesis to L-blocks, $V_i = V(X_i\cup S_i, Y_i)$. According to our theorem this gives the decomposition of the L-block $V_i$ as $G_i$-module, and the eigenvalues of $L_{Y_i}$ are indexed by poset tableau of shape $\mu_i \vdash \abs{S_i}$. Moreover each poset tableau of shape $\mu_i$ with eigenvalue $e_i$ represents a copy of the irreducible $S^{\mu_i}$ in the $e_i$-eigenspace. Now, as a $G_1 \times \dots \times G_r$ module we know the eigenspaces of $L_Y$ are given by our labeling up to the points $s_i$ where the partitions $\mu_i$ come together at $x_a$. \begin{lemma} As a $G = $Sym($B$)-module, the space $V$ is $$ V \cong \mbox{\rm ind}_{(\mbox{\tiny Sym($S_1^0$) $\times \cdots \times $ Sym($S_r^0$) )}}^{\mbox{\tiny Sym($B$)}} (V[S_1^0, \dots, S_r^0]) , $$ where $(S_1^0, \dots , S_r^0)$ is any fixed ordered partition of $B$. \end{lemma} Proof: Choose $S_i^0 = \{ x_{a+b_1+\cdots+b_{i-1}}, x_{a+b_1+\cdots+b_{i-1}+1}, \dots , x_{a+b_1+\cdots+b_{i}-1} \}$. Let \SS denote the set of permutations $\sigma \in \mbox{Sym}(B)$ such that $\sigma(u) < \sigma(v)$ whenever $u,v$ are in the same set $S_i^0$ for some $i$. There is 1--1 correspondence between the $\sigma \in \SS$ and the sequences indexing the summands in the~\ref{spliting:eq}, namely $$ \sigma \leftrightarrow (S_1^\sigma, \dots , S_r^\sigma) $$ where $S_i^\sigma = \{ \sigma(u) : u \in S_i^0 \} $. Also \SS is a collection of coset representatives for $\mbox{Sym}(S_1^0) \times \cdots \times \mbox{Sym}(S_r^0)$ in Sym($B$). Thus we have a natural vector space isomorphism between $V$ and $$ V[S_1^0, \dots, S_r^0] \otimes_{(\mbox{\tiny Sym($S_1^0$) $\times \cdots \times $ Sym($S_r^0$)})} \mbox{Sym}(B). $$ It is straightforward to check that this isomorphism commutes with the action of Sym($B$).\DONE Let $\LL_i$ be a poset tableau of type $(X_i\cup S_i, Y_i)$ , where $\mu_i \vdash b_i$ is the label of the vertex $s_i$. By our inductive hypothesis the Laplacian $L_{Y}$ acts as a scalar on the irreducible $S^{\mu_i}$, i.e., $L_{Y}\vert_{S^{\mu_i}} = e_Y(\LL_i)$. Applying Lemma~\ref{LLLEEEMMMAAA} part 2., we have \begin{eqnarray*} L_Y(v_1 \odo v_r) &=& \sum_i v_1 \odo (L_{Y_i}v_i) \odo v_r \\ &=& \sum_i v_1 \odo e_Y(\LL_i)v_i \odo v_r \\ &=& (\sum_i e_Y(\LL_i)) (v_1 \odo v_r). \end{eqnarray*} Now, we will use the fact (\cite{james,james-kerber,macdonald}) that $$ (S^{\mu_1}\odo S^{\mu_r})\uparrow_{G_1\times \cdots \times G_r}^G =\oplus_{\lambda \vdash \abs{B}} c_{\mu_1, \mu_2, \dots , \mu_r}^{\lambda} S^\lambda. $$ Thus we have $c_{\mu_1, \mu_2, \dots , \mu_r}^{\lambda}$ copies of the $G$-module $S^\lambda$ which explains why this is the multiplicity of the label $\lambda$ on node $x_a$ in our labeling. \item Now we have to decide what is the dimension of each eigenspace. But that is something we will have to do in the second case too - so we will do it for both cases at the end. % % %=============== C A S E 2 ================================= % % \subsubsection{Case 2} %\begin{figure}[h] %\centerline{\picture{0.7}{1.708}{5.138}{pic55.1.eps}} %\caption{Case 2}\label{prikaz.55.1} %\end{figure} %============================================================ \begin{figure} \begin{center} \begin{picture}(100,160)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,150)(30,0) \put(0,0){\line(0,1){120}} \put(0,120){\line(1,1){10}} \put(0,120){\line(-1,1){10}} \put(0,0){\circle*{3}} \put(0,10){\circle*{3}}\put(0,20){\circle*{3}}\put(0,30){\circle*{3}} \put(0,40){\circle*{3}} \put(0,50){\circle*{3}}\put(0,60){\circle*{3}} \put(0,70){\circle*{3}}\put(0,85){\circle*{3}} \put(0,100){\circle*{3}}\put(0,115){\circle*{3}} \put(0,120){\circle*{3}} \put(8,-3){\makebox(0,0)[bl]{$x_n$}} \put(8,27){\makebox(0,0)[bl]{$x_{a}$}} \put(8,37){\makebox(0,0)[bl]{$y_n$}} \put(8,67){\makebox(0,0)[bl]{$y_{n-k+1}$}} \put(8,82){\makebox(0,0)[bl]{$t$}} \put(20,83){\makebox(0,0)[bl]{$ \rangle NB$}} \put(8,97){\makebox(0,0)[bl]{$s$}} \put(20,98){\makebox(0,0)[bl]{$\rangle \hat{A}$}} \put(8,112){\makebox(0,0)[bl]{$v_0$}} \put(20,30){\line(2,-3){10}} \put(20,0){\line(2,3){10}}\put(31,12){\makebox(0,0)[bl]{$B$}} \put(20,70){\line(2,-3){10}} \put(20,40){\line(2,3){10}}\put(31,52){\makebox(0,0)[bl]{$A$}} \put(-2,123){\makebox(0,0)[bl]{$T$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{Case 2}\label{prikaz.55.1} \end{center} \end{figure} Let $A = \{y_{n-k+1}, \dots , y_n \}$ be the largest possible set so that $$x_a \le_P y_n \le_P y_{n-1} \le_P \cdots \le_P y_{n-k+1}$$ and there are no $x_i$'s with $y_n \le_P x_i _{S_{B'}} &=& \< S^{\lambda'}, S^{\mu'}\otimes \mbox{sgn}_{\gamma_1} \odo \mbox{sgn}_{\gamma_n} \>_{S_{B'}\times S_{\gamma_1} \times \dots \times S_{ \gamma_n}} \\ &=& \mbox{ \begin{minipage}[t]{3.82in} (\# of ways to get $\mu'$ from $\lambda'$ by successively removing vertical strips of lengths $c_1, c_2, \dots$) \end{minipage} } \\ \end{eqnarray*} Now consider the next step of going up from $\mu'$ to $\lambda$. At this point we have a piece of the $w$-eigenspace on which $B'$ acts like $S^{\mu'}$. The multiplicity coming from this $S^{\mu'}$ is $\mbox{dim}(\PPP_A \mbox{ind}_{G'}^G (S^{\mu'}) \PPP_B)$. We need to check how $\PPP_A$ acts on the induction $\mbox{ind}_{G'}^G(S^{\mu'}) = S^{\mu'} \otimes_{{\Bbb C}G'} {\Bbb C}G$. The $\sigma \in \PPP_A$ permutes the $y$'s. We identified the induction by identifying the sequence of the $x_i$'s that are paired with the set $A$. So switching $y_i$'s has the effect in Sym($B$) of switching the positions corresponding to $B\backslash B'$. In other words, $\PPP_A - \PPP_B$ has the effect of projecting onto the (trivial~$\otimes$~sgn) characters of $(S_{\alpha_1} \times \dots \times S_{\alpha_l}) \times (S_{\beta_1}\times \dots \times S_{\beta_m}) \subset S_{B\backslash B'} \times S_B$ where $S_{B\backslash B'} \times S_B$ is acting on $S^{\mu'} \otimes_{{\Bbb C}G'} {\Bbb C}G$ via left multiplication on ${\Bbb C}G$ by $ S_{B\backslash B'} $ and right multiplication on ${\Bbb C}G$ by $S_B$. So to determine $\mbox{dim}(\PPP_A \mbox{ind}_{G'}^G (S^{\mu'}) \PPP_B)$ it will be helpful to know the decomposition of the induction $\mbox{ind}_{G'}^G(S^{\mu'})$ as a $S_{B\backslash B'} \times S_B$-module. \begin{lemma}[\cite{hanlon}] Let $\mu' \vdash m$, $G=S_r$, $G'=S_m$ and $H=S_{r-m}$ (acting on $\{m+1, \dots, r\}$). Then as a $H\times G$ module, the induced representation $S^{\mu'}\otimes_{{\Bbb C}G'} {\Bbb C}G$ decomposes as $$ \mbox{ind}_{G'}^G(S^{\mu'})= \oplus_{\lambda \vdash r, \mu\subset\lambda} S^{\lambda /\mu} \otimes S^\lambda. $$ \end{lemma} Now armed with that lemma, let us return to the dimension count. \end{itemize} \begin{eqnarray*} \lefteqn{ \mbox{dim}(\PPP_A \mbox{ind}_{G'}^G (S^{\mu'}) \PPP_B) =}\\ &=& \< (\mbox{ind}_{G'}^G(S^{\mu'}))\downarrow , (\epsilon_{\alpha_1} \odo \epsilon_{\alpha_l})\otimes (\mbox{sgn}_{\beta_1} \odo \mbox{sgn}_{\beta_m})\> \\ &=& \sum_{\lambda \vdash n, \mu' \subset \lambda} \< (S^{\lambda /{\mu'}})\downarrow , (\epsilon_{\alpha_1} \odo \epsilon_{\alpha_l})\> \< S^\lambda\downarrow ,\mbox{sgn}_{\beta_1}\odo \mbox{sgn}_{\beta_m}\> \\ &=& \sum_{\mu'\subset\lambda \vdash n} \< (S^{\lambda /{\mu'}} \otimes S^\lambda)\downarrow , (\epsilon_{\alpha_1}\odo\epsilon_{\alpha_l})\otimes (\mbox{sgn}_{\beta_1}\odo\mbox{sgn}_{\beta_m})\>\\ &=& \sum_{\lambda\vdash n} \mbox{ \begin{minipage}[t]{4.75in} (\# of ways to get $\lambda$ from $\mu'$ by adding horizontal strips of lengths $\alpha_1,\alpha_2, \dots$ ) $\cdot$ (\# of ways to get $\lambda$ from $\emptyset$ by removing a vertical strips of lengths $\beta_1, \beta_2, \dots $ ) \bigskip \end{minipage}}\\ &=& \mbox{\begin{minipage}[t]{5.045in} (\# of poset tableaux labellings from $\mu'$ up to $\lambda$ then down to $\emptyset$, which add a horizontal strip of length $k$ for every $y$-vertex of repetition number $k$ and subtract a vertical strip of length $k$ for every $x$-vertex of repetition number $k$.) \end{minipage}} \\ \end{eqnarray*} This completes the proof of the theorem.\DONE %===========KRAJ TEOREMA====================================== %-------------------------------------------------------------------------------------------- %-------------------------------------------------------------------------------------------- %-------------------------------------------------------------------------------------------- %-------------------------------------------------------------------------------------------- %============================================================================ % A D I N G T H E Lx %===================================================================== \subsection{Adding the $L_X$} Consider the Laplacian $L_X$. Since we have identified the L-block $V$ with a subspace of the symmetric group algebra ${\Bbb C}S_n$, by fixing the order on the $x$'s, every time the Laplacian $L_X$ switches a pair of $x$'s, it is actually putting a minus sign in front of the corresponding basis element, with the $x$'s ordered. Since $L_X$ acts as a sum of transpositions, every eigenvalue we obtain from the $L_X$, will have a minus sign. Recall the multiplicity of the $x$-node. Let $v$ be an $x$-node of the diagram $P[X,Y]$, labeled with the partition $\LL(v)$ and with repetition number $k(v)$. Let $C(v)= \{ v_1, v_2, \dots , v_l\}$ be the set of covers of $v$. Let $\lambda_i$ denote $\LL(v_i)$, and let $k_i$ denote the repetition numbers, $k(v_i)$. The {\em multiplicity of $\LL$ at $v$} is defined to be $$ m_v(\LL) = \sum_{\mu} c_{\lambda_1, \dots, \lambda_l}^{\mu} c^{\mu}_{\LL(v), 1^{k(v)}}. $$ The {\em node--eigenvalue, $e_v(\LL)$, } for each node $v$, is the set of sums of the content over all squares in $\mu /\LL(v)$ above for a given $\mu$ minus the binomial coefficient $\binom{k(v)}{2}$. This gives $m_v(\LL)$ eigenvalues at each $x$-node $v$. We now define the {\em $x$-eigenvalue of $\LL$}, $e_x(\LL)$, to be the set of numbers obtained by taking a sum of one element of $e_v(\LL)$ for each $x$-node $v$. So $\abs{e_x(\LL)} = \prod_{\mbox{$x$-nodes $v$}} m_v(\LL)$. \begin{theorem}[$L_X$-Centerpiece] Let $P$ be a linear poset with a minimum element, $\hat 0$. Let $X$ and $Y$ be two (multi-)sets, subsets of $P$. For every labeling $\LL$ of positive multiplicity, each element in $e_x(\LL)$ is an eigenvalue of $L_X$. \end{theorem} Proof: The proof of this theorem is similar to the proof of the $L_Y$-Centerpiece Theorem. We will omit the details here.\DONE Since $L_X$ and $L_Y$ commute (as established in Lemma~\ref{lemma5.1} ), the eigenvalues of $L_X + L_Y$ will be the sum of the eigenvalues on the corresponding irreducibles of the eigenspaces. Recall that the complete Laplacian $L$ is the sum of three things (from the beginning of this section): $$ L=L_D + L_X + L_Y $$ The $L_D$ component is the diagonal matrix, which on the L-block $V$ spanned by the sets $(X,Y)$, has value: \begin{eqnarray*} e_D (X,Y) &=& w(X,Y) + \Delta(X,Y) \\ w(X,Y) &=& \sum_{m=1}^k \abs{(x_m,y_m)} \\ \Delta(X,Y) &=& \sum_{i,j} \delta_{x_i,y_j} \end{eqnarray*} The computer evidence strongly supports the following conjecture: \section{Complete Centerpiece Conjecture} \begin{conjecture} Let $P$ be a linear poset with a minimum element, $\hat 0$. Let $X$ and $Y$ be two multisets of vertices of $P$. Let $e_D(X,Y)$ be defined as above. For every poset tableau $\LL$ of positive multiplicity, let $\tau_Y(\LL) \in e_Y(\LL)$ and let $\tau_X(\LL) \in e_X(\LL)$. Then the scalars $e(\LL) = \tau_Y({\LL}) - \tau_X({\LL}) + e_D(X,Y)$ are the complete set of eigenvalues of the Laplacian $L$. \end{conjecture} This conjecture claims that the same poset tableau will work simultaneously for both Laplacians ($L_X$ and $L_Y$),i.e., that the eigenvalues of the Laplacian $L$ are the sum of the eigenvalues of $L_Y$ and the eigenvalues of $L_X$ evaluated simultaneously with the same poset tableau. %========================================================================= % H O M O L O G Y %========================================================================= \section{Homology} The object of the paper is to get a step closer to evaluating the homology of any Lie algebra corresponding to a linear poset, using only combinatorial properties of the poset. In these two small cases (n=1 and n=2) we had no difficulty. For larger n, we need some extra results. \subsection{$H_1$} For example, if we want to evaluate the homology $H_1(L_P)$ of a Lie algebra $L_P$ corresponding to a linear poset $P$, with $\hat 0$, our construction gives an immediate answer. An L-block $V$ of size 1, is determined by the sets $(X,Y)$, $X=\{x\}$, and $Y=\{y\}$. Obviously, if we want $V$ to be non-zero, $x<_P y$. So the corresponding diagram of this L-block is given in figure~\ref{sl501}. %\begin{figure}[h] %\centerline{\picture{1}{0.333}{0.625}{pic501.ps}} %\caption{\label{sl501} A diagram of the L-block} %\end{figure} \begin{figure} \begin{center} \begin{picture}(100,50)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,60)(30,0) \put(0,0){\line(0,1){30}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \end{picture}} %---------------------------------------------- \end{picture} \caption{\label{sl501} A diagram of the L-block} \end{center} \end{figure} Both indicators $e_Y$ and $e_X$ are zero, so the eigenvalues are given by $e_D$. But $\Delta(X,Y) = 0$ too, since $X\cap Y = \emptyset$. Thus $L(z_{x,y}) = w(X,Y) z_{x,y} = \abs{(x,y)} z_{x,y}$. In other words, the eigenvectors of the Laplacian $L_1$ are the basis vectors $z_{x,y}$, and the corresponding eigenvalues are $\abs{(x,y)}$, i.e., the number of the vertices in the poset $P$, between $x$ and $y$. The dimension of the homology is the number of zero eigenvalues, i.e., the number of the intervals $z_{x,y}$, such that $y$ covers $x$. Thus $$ \dim (H_1(L_P)) = \mbox{ (\# of covering relations in $P$)}. $$ \subsection{$H_2$} In this case the L-block $V$ in question is spanned by the (multi--)sets $(X,Y)$, each of size 2, i.e., $X=\{ x_1, x_2 \}$ and $Y=\{ y_1, y_2\}$. There are several possibilities for the L-block. \begin{enumerate} \item All four elements are comparable, and $x$'s are below the $y$'s. $$x_1 < x_2 < y_1 < y_2 .$$ All possible poset tableaux are shown in figure~\ref{sl502}. %\begin{figure} %\centerline{\picture{0.7}{1.930}{1.944}{pic502.ps}} %\caption{\label{sl502} poset tableaux} %\end{figure} \begin{figure}[t] \begin{center} \begin{picture}(200,120)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){90}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one\one}} \put(8,89){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e(\Lambda) = 0$}} \end{picture}} %---------------------------------------------- %---------------------------------------- \put(150,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){90}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one}} \put(8,48){\makebox(0,0)[bl]{\one}} \put(8,89){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e(\Lambda) = 0$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{\label{sl502} poset tableaux} \end{center} \end{figure} As we can see, both $e_Y$ and $e_X$ eigenvalues are zero. So we don't have to worry how to add them up - we will always get zero. $\Delta$ is also zero, since the sets $X$ and $Y$ are disjoint. Thus again, the Laplacian is $L(\zz 1 \ww \zz 2) = w(X,Y) (\zz 1 \ww \zz 2)$. But in this case, both intervals contain at least one element, so $w(X,Y)>0$. Thus in this case we never get a zero eigenvalue, which might contribute to the homology $H_2$. \item $y$'s are not comparable. $$ x_1 < x_2 < \begin{array}{c} y_1 \\ y_2 \end{array}. $$ All possible poset tableaux are shown in figure~\ref{sl503}. % %\begin{figure} %\centerline{\picture{0.7}{2.875}{1.750}{pic503.ps}} %\caption{\label{sl503} poset tableaux} %\end{figure} \begin{figure}[t] \begin{center} \begin{picture}(200,120)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){90}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one\one}} \put(8,89){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e(\Lambda) = -1$}} \end{picture}} %---------------------------------------------- %---------------------------------------- \put(150,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){90}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{\one}} \put(8,48){\makebox(0,0)[bl]{\one}} \put(8,89){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e(\Lambda) = +1$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{\label{sl503} poset tableaux} \end{center} \end{figure} The value of $e_Y$ is zero, while the values of $e_X$ are +1 and -1. Since $e_D$ is always non-negative, the value +1 cannot contribute to homology $H_2$. The other value can, but only if $e_D$ is +1. That means that all of the relations indicated: $x_1 < x_2$, $x_2 < y_1$ and $x_2< y_2$ are covering relations in the poset $P$. Whenever we have a four-element subset of the poset $P$, with covering relations $x_1 < x_2$, $x_2 < y_1$ and $x_2< y_2$, we will call that a ''Y``-configuration. Thus in this case every occurrence of ''Y``-configuration (described as above) in the Hasse diagram of the poset contributes to the homology $H_2$. \item $x_1 < y_1 < x_2 < y_2$ or equivalently (for our purpose) $x_1 < y_1 = x_2 < y_2$.\\ There is only one poset tableau spanned by these sets $X$ and $Y$, namely the one shown in figure~\ref{sl504}. %\begin{figure} %\centerline{\picture{0.7}{0.680}{1.944}{pic504.ps}} %\caption{\label{sl504} poset tableau} %\end{figure} \begin{figure} \begin{center} \begin{picture}(100,120)(0,0) %---------------------------------------- \put(50,30){\begin{picture}(60,110)(30,0) \put(0,0){\line(0,1){90}} \put(8,-1){\makebox(0,0)[bl]{$\emptyset$}} \put(8,29){\makebox(0,0)[bl]{\one}} \put(8,59){\makebox(0,0)[bl]{$\emptyset$}} \put(8,89){\makebox(0,0)[bl]{\one}} \put(0,0){\circle*{3}} \put(0,30){\circle*{3}} \put(0,60){\circle*{3}} \put(0,90){\circle*{3}} \put(-15,-15){\line(1,0){30}} \put(-20,-30){\makebox(0,0)[bl]{$e(\Lambda) = 0$}} \end{picture}} %---------------------------------------------- \end{picture} \caption{\label{sl504} poset tableau} \end{center} \end{figure} Since the space is one dimensional, we will add the eigenvalues in the only possible way. In the first case, the eigenvalue will be zero, if and only if both relations, $x_1> %% %% ACKNOWLEDGMENTS %% %% <<====================================================>> \vspace{1.0cm} \noindent{\bf Acknowledgement.} I would like to thank my advisor, Professor Phil Hanlon, for the guidance, help, and for his invaluable suggestions. \vspace{1.0cm} %======================================================= % BIBLIOGRAPHY FILE %-------------------------------------------------------------------------------------------- %======================================================= % B I B L I O G R A P H Y %======================================================== \begin{thebibliography}{1234} \bibliographystyle{plain} \def\em{\it} \bibitem{baclawski} K.~Baclawski, {\em Cohen-Macaulay ordered sets}, J. Algebra, [63] (1980), 226-258. \bibitem{bjorner} A.~Bj\"orner, {\em On the homology of geometric lattices}, Algebra Universalis, [14] (1982), 107-128. \bibitem{garsia} A.~Bj\"orner, A.~M.~Garsia, R.~P.~Stanley, {\em An introduction to Cohen-Macaulay partially ordered sets}, Ordered Sets (I. Rival ed.), 563-615, Boston: D. Reidel Publishing Co., 1981. \bibitem{cartan} H.~Cartan, S.~Eilenberg, {\em Homological algebra}, Princeton University Press, Princeton, 1956. \bibitem{crapo} H.~H.~Crapo, {\em The M\"obius function of a lattice}, Journal of Comb. Theory, [1] (1966), 126--131. \bibitem{eilenberg} C.~Chevalley, S.~Eilenberg, {\em Cohomology theory of Lie groups and Lie algebras}, Trans. of the AMS, [63] (1948), 85--124. \bibitem{chow} Y.~Chow, {\em General theory of Lie algebras}, Gordon and Breach, New York, 1978. \bibitem{farmer} F.~D. Farmer, {\em Cellular homology for posets}, Math. Japonica, [23] (1979), 607-613. \bibitem{fuks} D.~B. Fuks, {\em Cohomology of infinite-dimensional Lie algebras}, Consultants Bureau, New York, 1986. \bibitem{shack} M.~Gerstenhaber, S.D.~Schack, {\em The cohomology of presheaves of algebras I.}, Trans. of the AMS, [310] (1988), no.1 , 135-165. \bibitem{serre} G.~P.~Hochschild, J.~P.~Serre, {\em Cohomology of Lie algebras}. Ann. of Math., [57] (1953), 591--603. \bibitem{hanlon} P.~Hanlon, {\em Some twisted random walks indexed by sequences}, preprint, (1993). \bibitem{hochschild} G.~P.~Hochschild, {\em Basic theory of algebraic groups and Lie algebras}, Sprin\-ger-Verlag, New York, 1981. \bibitem{humphreys} J.~E.~Humphreys, {\em Introduction to Lie algebras and representation theory}, Springer-Verlag, New York, 1972. \bibitem{jacobson} N.~Jacobson, {\em Lie algebras}, Dover Publications Inc., New York, 1962. \bibitem{james} G. D. James, {\em The representation theory of the symmetric group}, Springer--Verlag, Berlin~Heidelberg~New York, 1978. \bibitem{james-kerber} G.~James, A.~Kerber, {\em The representation theory of the symmetric group}, Addison--Wesley Publishing Company, Reading, MA, 1981. \bibitem{knapp} A.~W. Knapp, {\em Lie groups, Lie algebras, and cohomology}, Princeton University Press, Princeton, 1988. \bibitem{kostant} B.~Kostant, {\em Lie algebra cohomology and the generalized Borel-Weil theorem}, Ann. of Math., [74] (1961), 329--387. \bibitem{kostant1} B.~Kostant, {\em Lie algebra cohomology and generalized Schubert cells}, Ann. of Math., [77] (1963), 72--144. \bibitem{koszul} J.~L. Koszul. {\em Homologie et cohomologie des alg\`ebres de Lie}, Bull. Soc. Math. France, [78] (1950), 65--127. \bibitem{lakser} H.~Lakser, {\em The homology of lattice}, Discrete Math., [1] (1971), 187--192. \bibitem{leger} G. Leger, E. Luks, {\em Cohomology and weight systems for nilpotent Lie algebras}, Bull. Amer. Math. Soc., [80] (1974), 77-80. \bibitem{macdonald} I.~G.~Macdonald, {\em Symmetric functions and Hall polynomials}, Oxford University Press, New York, 1979. \bibitem{mather} J.~Mather, {\em Invariance of the homology of a lattice}, Proc. Amer. Math. Soc., [17] (1966), 1120--1124. \bibitem{murphy} G.~E. Murphy, {\em A new construction of Young's seminormal representation of the symmetric group}, Journal of Algebra, [69] (1981), 287--297. \bibitem{peel} M. H.~Peel, {\em Specht modules and symmetric groups}, Journal of Algebra, [36] (1975), 88-97. \bibitem{robinson} G.~de~B. Robinson, {\em Representation theory of the symmetric group}, University of Toronto Press, 1961. \bibitem{rota} G.-C.~Rota, {\em On the foundations of combinatorial theory. I: Theory of M\"obius functions}, Z. Wahrsch. Verw. Gebiete, [2] (1964), 340--368. \bibitem{sagan} B. E.~Sagan, {\em The symmetric group}, Wadsworth \& Brooks/Cole, Advanced Books \& Software, Pacific Grove, California, 1991. \bibitem{stanley} R.~P. Stanley, {\em Enumerative combinatorics}, v.~1, Wadsworth and Brooks/Cole Mathematics Series, Monterey, California, 1986. \bibitem{thomas} G. P.~Thomas, {\em On Schensted's construction and the multiplication of Schur functions}, Advances in Math., [30] (1978), 8--32. \bibitem{walker} J.~Walker, {\em Homotopy type and Euler characteristic of partially ordered sets}, European J. Combinatorics, [2] (1981), 373-384. \bibitem{white} D. E.~White, {\em Some connections between the Littlewood--Richardson rule and the construction of Schensted}, Journal of Comb. Theory, Series A, [30] (1981), 237-247. \end{thebibliography} \end{document} .